Você está na página 1de 578

Eyal Buks

Quantum Mechanics - Lecture


Notes
March 1, 2015

Technion
Preface

The dynamics of a quantum system is governed by the celebrated Schrödinger


equation
d
i |ψ = H |ψ , (0.1)
dt

where i = −1 and  = 1.05457266 × 10−34 J s is Planck’s h-bar constant.
However, what is the meaning of the symbols |ψ and H? The answers will
be given in the first part of the course (chapters 1-4), which reviews several
physical and mathematical concepts that are needed to formulate the theory
of quantum mechanics. We will learn that |ψ in Eq. (0.1) represents the
ket-vector state of the system and H represents the Hamiltonian operator.
The operator H is directly related to the Hamiltonian function in classical
physics, which will be defined in the first chapter. The ket-vector state and
its physical meaning will be introduced in the second chapter. Chapter 3
reviews the position and momentum operators, whereas chapter 4 discusses
dynamics of quantum systems. The second part of the course (chapters 5-7)
is devoted to some relatively simple quantum systems including a harmonic
oscillator, spin, hydrogen atom and more. In chapter 8 we will study quantum
systems in thermal equilibrium. The third part of the course (chapters 9-13)
is devoted to approximation methods such as perturbation theory, semiclas-
sical and adiabatic approximations. Light and its interaction with matter are
the subjects of chapter 14-15. Finally, systems of identical particles will be
discussed in chapter 16 and open quantum systems in chapter 17. Most of
the material in these lecture notes is based on the textbooks [1, 2, 3, 4, 6, 7].
Contents

1. Hamilton’s Formalism of Classical Physics . . . . . . . . . . . . . . . . 1


1.1 Action and Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Principle of Least Action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Poisson’s Brackets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2. State Vectors and Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15


2.1 Linear Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3 Dirac’s notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4 Dual Correspondence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5 Matrix Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.6 Observables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.6.1 Hermitian Adjoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.6.2 Eigenvalues and Eigenvectors . . . . . . . . . . . . . . . . . . . . . . 23
2.7 Quantum Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.8 Example - Spin 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.9 Unitary Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.10 Trace . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.11 Commutation Relation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.12 Simultaneous Diagonalization of Commuting Operators . . . . . 35
2.13 Uncertainty Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.15 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3. The Position and Momentum Observables . . . . . . . . . . . . . . . . 49


3.1 The One Dimensional Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.1.1 Position Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.1.2 Momentum Representation . . . . . . . . . . . . . . . . . . . . . . . . 54
3.2 Transformation Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.3 Generalization for 3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Contents

3.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4. Quantum Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.1 Time Evolution Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.2 Time Independent Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.3 Example - Spin 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.4 Connection to Classical Dynamics . . . . . . . . . . . . . . . . . . . . . . . . 73
4.5 Symmetric Ordering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
4.7 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

5. The Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101


5.1 Eigenstates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.2 Coherent States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.3 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.4 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

6. Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149


6.1 Angular Momentum and Rotation . . . . . . . . . . . . . . . . . . . . . . . . 150
6.2 General Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
6.3 Simultaneous Diagonalization of J2 and Jz . . . . . . . . . . . . . . . . 152
6.4 Example - Spin 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
6.5 Orbital Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
6.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
6.7 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

7. Central Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203


7.1 Simultaneous Diagonalization of the Operators H, L2 and Lz 204
7.2 The Radial Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
7.3 Hydrogen Atom . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
7.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
7.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215

8. Density Operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223


8.1 Time Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
8.2 Quantum Statistical Mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
8.3 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
8.4 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239

9. Time Independent Perturbation Theory . . . . . . . . . . . . . . . . . . 293


9.1 The Level En . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
9.1.1 Nondegenerate Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
9.1.2 Degenerate Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
9.2 Example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
9.3 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300

Eyal Buks Quantum Mechanics - Lecture Notes 6


Contents

9.4 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306

10. Time-Dependent Perturbation Theory . . . . . . . . . . . . . . . . . . . . 337


10.1 Time Evolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
10.2 Perturbation Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
10.3 Transition Probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
10.3.1 The Stationary Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
10.3.2 The Near-Resonance Case . . . . . . . . . . . . . . . . . . . . . . . . . 343
10.3.3 H1 is Separable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
10.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
10.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346

11. WKB Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351


11.1 WKB Wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
11.2 Turning Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 354
11.3 Bohr-Sommerfeld Quantization Rule . . . . . . . . . . . . . . . . . . . . . . 358
11.4 Tunneling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
11.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
11.6 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361

12. Path Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 369


12.1 Charged Particle in Electromagnetic Field . . . . . . . . . . . . . . . . . 369
12.2 Classical Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
12.3 Aharonov-Bohm Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
12.3.1 Two-slit Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376
12.3.2 Gauge Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
12.4 One Dimensional Path Integrals . . . . . . . . . . . . . . . . . . . . . . . . . . 379
12.4.1 One Dimensional Free Particle . . . . . . . . . . . . . . . . . . . . . 380
12.4.2 Expansion Around the Classical Path . . . . . . . . . . . . . . . 381
12.4.3 One Dimensional Harmonic Oscillator . . . . . . . . . . . . . . . 383
12.5 Semiclassical Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
12.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
12.7 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389

13. Adiabatic Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397


13.1 Momentary Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 397
13.2 Gauge Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
13.3 Adiabatic Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
13.4 The Case of Two Dimensional Hilbert Space . . . . . . . . . . . . . . . 400
13.5 Transition Probability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
13.5.1 The Case of Two Dimensional Hilbert Space . . . . . . . . . 403
13.6 Slow and Fast Coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
13.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
13.8 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410

Eyal Buks Quantum Mechanics - Lecture Notes 7


Contents

14. The Quantized Electromagnetic Field . . . . . . . . . . . . . . . . . . . . . 415


14.1 Classical Electromagnetic Cavity . . . . . . . . . . . . . . . . . . . . . . . . . 415
14.2 Quantum Electromagnetic Cavity . . . . . . . . . . . . . . . . . . . . . . . . 420
14.3 Periodic Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 422
14.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
14.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 424

15. Light Matter Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431


15.1 Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
15.2 Transition Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
15.2.1 Spontaneous Emission . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432
15.2.2 Stimulated Emission and Absorption . . . . . . . . . . . . . . . . 433
15.2.3 Selection Rules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 434
15.3 Semiclassical Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
15.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
15.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438

16. Identical Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445


16.1 Basis for the Hilbert Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 445
16.2 Bosons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
16.3 Fermions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 449
16.4 Changing the Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 451
16.5 Many Particle Observables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453
16.5.1 One-Particle Observables . . . . . . . . . . . . . . . . . . . . . . . . . . 453
16.5.2 Two-Particle Observables . . . . . . . . . . . . . . . . . . . . . . . . . . 454
16.6 Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 455
16.7 Momentum Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458
16.8 Spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
16.9 The Electron Gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
16.10Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 463
16.11Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 465

17. Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483


17.1 Macroscopic Wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483
17.1.1 Single Particle in Electromagnetic Field . . . . . . . . . . . . . 483
17.1.2 Drude Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
17.1.3 The Macroscopic Quantum Model . . . . . . . . . . . . . . . . . . 489
17.1.4 London Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 489
17.2 The Josephson Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493
17.2.1 The First Josephson Relation . . . . . . . . . . . . . . . . . . . . . . 493
17.2.2 The Second Josephson Relation . . . . . . . . . . . . . . . . . . . . 494
17.2.3 The Energy of a Josephson Junction . . . . . . . . . . . . . . . . 495
17.3 RF SQUID . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 496
17.3.1 Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
17.3.2 Flux Quantum Bit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499

Eyal Buks Quantum Mechanics - Lecture Notes 8


Contents

17.3.3 Qubit Readout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500


17.4 BCS Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
17.4.1 Phonon Mediated Electron-Electron Interaction . . . . . . 509
17.4.2 The Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
17.4.3 Bogoliubov Transformation . . . . . . . . . . . . . . . . . . . . . . . . 513
17.4.4 The Energy Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 516
17.4.5 The Ground State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 518
17.4.6 Pairing Wavefunction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 520
17.5 The Josephson Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 521
17.5.1 The Second Josephson Relation . . . . . . . . . . . . . . . . . . . . 522
17.5.2 The Energy of a Josephson Junction . . . . . . . . . . . . . . . . 524
17.5.3 The First Josephson Relation . . . . . . . . . . . . . . . . . . . . . . 527
17.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 528
17.7 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 530

18. Open Quantum Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 541


18.1 Classical Resonator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 541
18.2 Quantum Resonator Coupled to Thermal Bath . . . . . . . . . . . . . 542
18.2.1 The closed System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 542
18.2.2 Coupling to Thermal Bath . . . . . . . . . . . . . . . . . . . . . . . . 543
18.2.3 Thermal Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 546
18.3 Two Level System Coupled to Thermal Bath . . . . . . . . . . . . . . . 549
18.3.1 The Closed System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 549
18.3.2 Coupling to Thermal Baths . . . . . . . . . . . . . . . . . . . . . . . . 550
18.3.3 Thermal Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554
18.3.4 Correlation Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 555
18.3.5 The Bloch Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
18.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
18.5 Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567

Eyal Buks Quantum Mechanics - Lecture Notes 9


1. Hamilton’s Formalism of Classical Physics

In this chapter the Hamilton’s formalism of classical physics is introduced,


with a special emphasis on the concepts that are needed for quantum me-
chanics.

1.1 Action and Lagrangian


Consider a classical physical system having N degrees of freedom. The clas-
sical state of the system can be described by N independent coordinates qn ,
where n = 1, 2, · · · , N . The vector of coordinates is denoted by

Q = (q1 , q2 , · · · , qN ) . (1.1)

Consider the case where the vector of coordinates takes the value Q1 at time
t1 and the value Q2 at a later time t2 > t1 , namely
Q (t1 ) = Q1 , (1.2)
Q (t2 ) = Q2 . (1.3)
The action S associated with the evolution of the system from time t1 to
time t2 is defined by
t2
S= dt L , (1.4)
t1

where L is the Lagrangian function of the system. In general, the Lagrangian


is a function of the coordinates Q, the velocities Q̇ and time t, namely
 
L = L Q, Q̇; t , (1.5)

where

Q̇ = (q̇1 , q̇2 , · · · , q̇N ) , (1.6)

and where overdot denotes time derivative. The time evolution of Q, in turn,
depends on the trajectory taken by the system from point Q1 at time t1
Chapter 1. Hamilton’s Formalism of Classical Physics

Q
Q2

Q1
t
t1 t2

Fig. 1.1. A trajectory taken by the system from point Q1 at time t1 to point Q2
at time t2 .

to point Q2 at time t2 (see Fig. 1.1). For a given trajectory Γ the time
dependency is denoted as
Q (t) = QΓ (t) . (1.7)

1.2 Principle of Least Action


For any given trajectory Q (t) the action can be evaluated using Eq. (1.4).
Consider a classical system evolving in time from point Q1 at time t1 to point
Q2 at time t2 along the trajectory QΓ (t). The trajectory QΓ (t), which is
obtained from the laws of classical physics, has the following unique property
known as the principle of least action:
Proposition 1.2.1 (principle of least action). Among all possible trajec-
tories from point Q1 at time t1 to point Q2 at time t2 the action obtains its
minimal value by the classical trajectory QΓ (t).
In a weaker version of this principle, the action obtains a local minimum
for the trajectory QΓ (t). As the following theorem shows, the principle of
least action leads to a set of equations of motion, known as Euler-Lagrange
equations.
Theorem 1.2.1. The classical trajectory QΓ (t), for which the action obtains
its minimum value, obeys the Euler-Lagrange equations of motion, which are
given by

Eyal Buks Quantum Mechanics - Lecture Notes 2


1.2. Principle of Least Action

d ∂L ∂L
= , (1.8)
dt ∂ q̇n ∂qn
where n = 1, 2, · · · , N.

Proof. Consider another trajectory QΓ ′ (t) from point Q1 at time t1 to point


Q2 at time t2 (see Fig. 1.2). The difference

δQ = QΓ ′ (t) − QΓ (t) = (δq1 , δq2 , · · · , δqN ) (1.9)

is assumed to be infinitesimally small. To lowest order in δQ the change in


the action δS is given by
t2
δS = dt δL
t1
t2  N N

 ∂L  ∂L
= dt δqn + δ q̇n
n=1
∂qn n=1
∂ q̇n
t1
t2  N N

 ∂L  ∂L d
= dt δqn + δqn .
n=1
∂qn n=1
∂ q̇n dt
t1
(1.10)
Integrating the second term by parts leads to
t2 N 
 
∂L d ∂L
δS = dt − δqn
n=1
∂qn dt ∂ q̇n
t1
N

t2
∂L

+ δqn

.
n=1
∂ q̇n t1
(1.11)
The last term vanishes since

δQ (t1 ) = δQ (t2 ) = 0 . (1.12)

The principle of least action implies that

δS = 0 . (1.13)

This has to be satisfied for any δQ, therefore the following must hold
d ∂L ∂L
= . (1.14)
dt ∂ q̇n ∂qn

Eyal Buks Quantum Mechanics - Lecture Notes 3


Chapter 1. Hamilton’s Formalism of Classical Physics

Q
Q2

Γ’
Q1
t
t1 t2

Fig. 1.2. The classical trajectory QΓ (t) and the trajectory QΓ ′ (t).

In what follows we will assume for simplicity that the kinetic energy T of
the system can be expressed as a function of the velocities Q̇ only (namely,
it does not explicitly depend on the coordinates Q). The components of the
generalized force Fn , where n = 1, 2, · · · , N , are derived from the potential
energy U of the system as follows
∂U d ∂U
Fn = − + . (1.15)
∂qn dt ∂ q̇n
When the potential energy can be expressed as a function of the coordinates
Q only (namely, when it is independent on the velocities Q̇), the system is
said to be conservative. For that case, the Lagrangian can be expressed in
terms of T and U as
L=T −U . (1.16)
Example 1.2.1. Consider a point particle having mass m moving in a one-
dimensional potential U (x). The Lagrangian is given by
mẋ2
L=T −U = − U (x) . (1.17)
2
From the Euler-Lagrange equation
d ∂L ∂L
= , (1.18)
dt ∂ ẋ ∂x
one finds that
∂U
mẍ = − . (1.19)
∂x

Eyal Buks Quantum Mechanics - Lecture Notes 4


1.3. Hamiltonian

1.3 Hamiltonian
The set of Euler-Lagrange equations contains N second order differential
equations. In this section we derive an alternative and equivalent set of equa-
tions of motion, known as Hamilton-Jacobi equations, that contains twice the
number of equations, namely 2N, however, of first, instead of second, order.
Definition 1.3.1. The variable canonically conjugate to qn is defined by
∂L
pn = . (1.20)
∂ q̇n
Definition 1.3.2. The Hamiltonian of a physical system is a function of
the vector of coordinates Q, the vector of canonical conjugate variables P =
(p1 , p2 , · · · , pN ) and time, namely

H = H (Q, P ; t) , (1.21)

is defined by
N

H= pn q̇n − L , (1.22)
n=1

where L is the Lagrangian.


Theorem 1.3.1. The classical trajectory satisfies the Hamilton-Jacobi equa-
tions of motion, which are given by
∂H
q̇n = , (1.23)
∂pn
∂H
ṗn = − , (1.24)
∂qn
where n = 1, 2, · · · , N.
Proof. The differential of H is given by
N

dH = d pn q̇n − dL
n=1
 

N  
 ∂L ∂L  ∂L
 
= q̇n dpn + pn dq̇n − ∂qn dqn − ∂ q̇n dq̇n  − ∂t dt
n=1    
d ∂L pn
dt ∂ q̇n

N
 ∂L
= (q̇n dpn − ṗn dqn ) − dt .
n=1
∂t
(1.25)

Eyal Buks Quantum Mechanics - Lecture Notes 5


Chapter 1. Hamilton’s Formalism of Classical Physics

Thus the following holds


∂H
q̇n = , (1.26)
∂pn
∂H
ṗn = − , (1.27)
∂qn
∂L ∂H
− = . (1.28)
∂t ∂t
Corollary 1.3.1. The following holds
dH ∂H
= . (1.29)
dt ∂t
Proof. Using Eqs. (1.23) and (1.24) one finds that
N  
dH  ∂H ∂H ∂H ∂H
= q̇n + ṗn + = . (1.30)
dt n=1
∂qn ∂pn ∂t ∂t
  
=0

The last corollary implies that H is time independent provided that H


does not depend on time explicitly, namely, provided that ∂H/∂t = 0. This
property is referred to as the law of energy conservation. The theorem below
further emphasizes the relation between the Hamiltonian and the total energy
of the system.

Theorem 1.3.2. Assume that the kinetic energy of a conservative system is


given by

T = αnm q̇n q̇m , (1.31)
n,m

where αnm are constants. Then,the Hamiltonian of the system is given by

H =T +U , (1.32)

where T is the kinetic energy of the system and where U is the potential
energy.

Proof. For a conservative system the potential energy is independent on ve-


locities, thus
∂L ∂T
pl = = , (1.33)
∂ q̇l ∂ q̇l
where L = T − U is the Lagrangian. The Hamiltonian is thus given by

Eyal Buks Quantum Mechanics - Lecture Notes 6


1.4. Poisson’s Brackets

N

H= pl q̇l − L
l=1
 ∂T
= q̇l − (T − U )
∂ q̇l
l
 
  ∂ q̇ ∂ q̇m 
 n 
= αnm q̇m + q̇n  q̇l − T + U
 ∂ q̇l ∂ q̇l 
l,n,m    
δnl δ ml

=2 αnm q̇n q̇m − T + U
n,m
  
T
= T +U .
(1.34)

1.4 Poisson’s Brackets

Consider two physical quantities F and G that can be expressed as a function


of the vector of coordinates Q, the vector of canonical conjugate variables P
and time t, namely
F = F (Q, P ; t) , (1.35)
G = G (Q, P ; t) , (1.36)
The Poisson’s brackets are defined by

N  
∂F ∂G ∂F ∂G
{F, G} = − , (1.37)
n=1
∂qn ∂pn ∂pn ∂qn

The Poisson’s brackets are employed for writing an equation of motion for a
general physical quantity of interest, as the following theorem shows.

Theorem 1.4.1. Let F be a physical quantity that can be expressed as a


function of the vector of coordinates Q, the vector of canonical conjugate
variables P and time t, and let H be the Hamiltonian. Then, the following
holds
dF ∂F
= {F, H} + . (1.38)
dt ∂t
Proof. Using Eqs. (1.23) and (1.24) one finds that the time derivative of F
is given by

Eyal Buks Quantum Mechanics - Lecture Notes 7


Chapter 1. Hamilton’s Formalism of Classical Physics

N  
dF ∂F ∂F ∂F
= q̇n + ṗn +
dt n=1
∂qn ∂pn ∂t
N  
∂F ∂H ∂F ∂H ∂F
= − +
n=1
∂qn ∂p n ∂p n ∂q n ∂t
∂F
= {F, H} + .
∂t
(1.39)
Corollary 1.4.1. If F does not explicitly depend on time, namely if ∂F/∂t =
0, and if {F, H} = 0, then F is a constant of the motion, namely
dF
=0. (1.40)
dt

1.5 Problems
1. Consider a particle having charge q and mass m in electromagnetic field
characterized by the scalar potential ϕ and the vector potential A. The
electric field E and the magnetic field B are given by
1 ∂A
E = −∇ϕ − , (1.41)
c ∂t
and
B=∇×A. (1.42)
Let r = (x, y, z) be the Cartesian coordinates of the particle.
a) Verify that the Lagrangian of the system can be chosen to be given
by
1 2 q
L= mṙ − qϕ + A · ṙ , (1.43)
2 c
by showing that the corresponding Euler-Lagrange equations are
equivalent to Newton’s 2nd law (i.e., F = mr̈).
b) Show that the Hamilton-Jacobi equations are equivalent to Newton’s
2nd law.
c) Gauge transformation — The electromagnetic field is invariant un-
der the gauge transformation of the scalar and vector potentials
A → A + ∇χ , (1.44)
1 ∂χ
ϕ → ϕ− (1.45)
c ∂t
where χ = χ (r, t) is an arbitrary smooth and continuous function
of r and t. What effect does this gauge transformation have on the
Lagrangian and Hamiltonian? Is the motion affected?

Eyal Buks Quantum Mechanics - Lecture Notes 8


1.6. Solutions

L C

Fig. 1.3. LC resonator.

2. Consider an LC resonator made of a capacitor having capacitance C in


parallel with an inductor having inductance L (see Fig. 1.3). The state
of the system is characterized by the coordinate q , which is the charge
stored by the capacitor.
a) Find the Euler-Lagrange equation of the system.
b) Find the Hamilton-Jacobi equations of the system.
c) Show that {q, p} = 1 .
3. Show that Poisson brackets satisfy the following relations
{qj , qk } = 0 , (1.46)
{pj , pk } = 0 , (1.47)
{qj , pk } = δ jk , (1.48)
{F, G} = − {G, F } , (1.49)
{F, F } = 0 , (1.50)
{F, K} = 0 if K constant or F depends only on t , (1.51)
{E + F, G} = {E, G} + {F, G} , (1.52)
{E, F G} = {E, F } G + F {E, G} . (1.53)
4. Show that the Lagrange equations are coordinate invariant.
5. Consider a point particle having mass m moving in a 3D central po-
tential,
 namely a potential V (r) that depends only on the distance
r = x2 + y 2 + z 2 from the origin. Show that the angular momentum
L = r × p is a constant of the motion.

1.6 Solutions

1. The Lagrangian of the system (in Gaussian units) is taken to be given


by
1 q
L = mṙ2 − qϕ + A · ṙ . (1.54)
2 c

Eyal Buks Quantum Mechanics - Lecture Notes 9


Chapter 1. Hamilton’s Formalism of Classical Physics

a) The Euler-Lagrange equation for the coordinate x is given by

d ∂L ∂L
= , (1.55)
dt ∂ ẋ ∂x
where
 
d ∂L q ∂Ax ∂Ax ∂Ax ∂Ax
= mẍ + + ẋ + ẏ + ż , (1.56)
dt ∂ ẋ c ∂t ∂x ∂y ∂z

and
 
∂L ∂ϕ q ∂Ax ∂Ay ∂Az
= −q + ẋ + ẏ + ż , (1.57)
∂x ∂x c ∂x ∂x ∂x

thus
∂ϕ q ∂Ax
mẍ = −q −
 ∂x  c ∂t 
qEx
 
 
 
 
    
q  ∂Ay ∂Ax ∂Ax ∂Az 
+  ẏ − −ż −  ,
c ∂x ∂y ∂z ∂x 
      

 (∇ ×A)z (∇ ×A)y 
  
(ṙ×(∇×A))x

(1.58)
or
q
mẍ = qEx + (ṙ × B)x . (1.59)
c
Similar equations are obtained for ÿ and z̈ in the same way. These 3
equations can be written in a vector form as
 
1
mr̈ = q E + ṙ × B . (1.60)
c

b) The variable vector canonically conjugate to the coordinates vector


r is given by
∂L q
p= = mṙ+ A . (1.61)
∂ ṙ c
The Hamiltonian is thus given by

Eyal Buks Quantum Mechanics - Lecture Notes 10


1.6. Solutions

H = p · ṙ − L
 
1 q
= ṙ · p − mṙ − A + qϕ
2 c
1 2
= mṙ + qϕ
2
 2
p− qc A
= + qϕ .
2m
(1.62)
The Hamilton-Jacobi equation for the coordinate x is given by
∂H
ẋ = , (1.63)
∂px
thus
px − qc Ax
ẋ = , (1.64)
m
or
q
px = mẋ+ Ax . (1.65)
c
The Hamilton-Jacobi equation for the canonically conjugate variable
px is given by
∂H
ṗx = − , (1.66)
∂x
where
 
q ∂Ax ∂Ax ∂Ax q ∂Ax
ṗx = mẍ+ ẋ + ẏ + ż + , (1.67)
c ∂x ∂y ∂z c ∂t

and  
∂H q px − qc Ax ∂Ax py − qc Ay ∂Ay pz − qc Az ∂Az ∂ϕ
− = + + −q
∂x c m ∂x m ∂x m ∂x ∂x
 
q ∂Ax ∂Ay ∂Az ∂ϕ
= ẋ + ẏ + ż −q ,
c ∂x ∂x ∂x ∂x
(1.68)
thus
   
∂ϕ q ∂Ax q ∂Ay ∂Ax ∂Ax ∂Az
mẍ = −q − + ẏ − − ż − .
∂x c ∂t c ∂x ∂y ∂z ∂x
(1.69)

The last result is identical to Eq. (1.59).

Eyal Buks Quantum Mechanics - Lecture Notes 11


Chapter 1. Hamilton’s Formalism of Classical Physics

c) Clearly, the fields E and B, which are given by Eqs. (1.41) and (1.42)
respectively, are unchanged since
 
∂χ ∂ (∇χ)
∇ − =0, (1.70)
∂t ∂t
and
∇ × (∇χ) = 0 . (1.71)
Thus, even though both L and H are modified, the motion, which
depends on E and B only, is unaffected.
2. The kinetic energy in this case T = Lq̇ 2 /2 is the energy stored in the
inductor, and the potential energy U = q 2 /2C is the energy stored in the
capacitor.
a) The Lagrangian is given by
Lq̇ 2 q2
L=T −U = − . (1.72)
2 2C
The Euler-Lagrange equation for the coordinate q is given by
d ∂L ∂L
= , (1.73)
dt ∂ q̇ ∂q
thus
q
Lq̈ + =0. (1.74)
C
This equation expresses the requirement that the voltage across the
capacitor is the same as the one across the inductor.
b) The canonical conjugate momentum is given by
∂L
p= = Lq̇ , (1.75)
∂ q̇
and the Hamiltonian is given by
p2 q2
H = pq̇ − L = + . (1.76)
2L 2C
Hamilton-Jacobi equations read
p
q̇ = (1.77)
L
q
ṗ = − , (1.78)
C
thus
q
Lq̈ + =0. (1.79)
C

Eyal Buks Quantum Mechanics - Lecture Notes 12


1.6. Solutions

c) Using the definition (1.37) one has


∂q ∂p ∂q ∂p
{q, p} = − =1. (1.80)
∂q ∂p ∂p ∂q
3. All these relations
 are easily proven using the definition (1.37).
4. Let L = L Q, Q̇; t be a Lagrangian of a system, where Q = (q1 , q2 , · · · )
is the vector of coordinates, Q̇ = (q̇1 , q̇2 , · · · ) is the vector of veloci-
ties, and where overdot denotes time derivative. Consider the coordinates
transformation

xa = xa (q1 , q2 , ..., t) , (1.81)

where a = 1, 2, · · · . The following holds


∂xa ∂xa
ẋa = q̇b + , (1.82)
∂qb ∂t
where the summation convention is being used, namely, repeated indices
are summed over. Moreover
∂L ∂L ∂xb ∂L ∂ ẋb
= + , (1.83)
∂qa ∂xb ∂qa ∂ ẋb ∂qa
and
   
d ∂L d ∂L ∂ ẋb
= . (1.84)
dt ∂ q̇a dt ∂ ẋb ∂ q̇a

As can be seen from Eq. (1.82), one has


∂ ẋb ∂xb
= . (1.85)
∂ q̇a ∂qa
Thus, using Eqs. (1.83) and (1.84) one finds
   
d ∂L ∂L d ∂L ∂xb
− =
dt ∂ q̇a ∂qa dt ∂ ẋb ∂qa
∂L ∂xb ∂L ∂ ẋb
− −
∂x ∂q ∂ ẋ ∂q
b  a  b a
d ∂L ∂L ∂xb
= −
dt ∂ ẋb ∂xb ∂qa
  
d ∂xb ∂ ẋb ∂L
+ − .
dt ∂qa ∂qa ∂ ẋb
(1.86)

As can be seen from Eq. (1.82), the second term vanishes since

Eyal Buks Quantum Mechanics - Lecture Notes 13


Chapter 1. Hamilton’s Formalism of Classical Physics
 
∂ ẋb ∂ 2 xb ∂ 2 xb d ∂xb
= q̇c + = ,
∂qa ∂qa ∂qc ∂t∂qa dt ∂qa

thus
    
d ∂L ∂L d ∂L ∂L ∂xb
− = − . (1.87)
dt ∂ q̇a ∂qa dt ∂ ẋb ∂xb ∂qa

The last result shows that if the coordinate transformation is reversible,


namely if det (∂xb /∂qa ) = 0 then Lagrange equations are coordinate
invariant.
5. The angular momentum L is given by
 
x̂ ŷ ẑ
L = r × p = det  x y z  , (1.88)
px py pz

where r = (x, y, z) is the position vector and where p = (px , py , pz ) is the


momentum vector. The Hamiltonian is given by

p2
H= + V (r) . (1.89)
2m
Using
{xi , pj } = δ ij , (1.90)
Lz = xpy − ypx , (1.91)
one finds that
 2   
p , Lz = p2x , Lz + p2y , Lz + p2z , Lz
 
= p2x , xpy − p2y , ypx
= −2px py + 2py px
=0,
(1.92)
and
 2   
r , Lz = x2 , Lz + y 2 , Lz + z 2 , Lz
 
= −y x2 , px + y 2 , py x
=0.
(1.93)
  2  2
Thus f r , Lz = 0 for arbitrary smooth function f r , and con-
sequently {H, Lz } = 0. In a similar way one can show that {H, Lx } =
{H, Ly } = 0, and therefore H, L2 = 0.

Eyal Buks Quantum Mechanics - Lecture Notes 14


2. State Vectors and Operators

In quantum mechanics the state of a physical system is described by a state


vector |α, which is a vector in a vector space F, namely
|α ∈ F . (2.1)
Here, we have employed the Dirac’s ket-vector notation |α for the state vec-
tor, which contains all information about the state of the physical system
under study. The dimensionality of F is finite in some specific cases (no-
tably, spin systems), however, it can also be infinite in many other cases
of interest. The basic mathematical theory dealing with vector spaces hav-
ing infinite dimensionality was mainly developed by David Hilbert. Under
some conditions, vector spaces having infinite dimensionality have properties
similar to those of their finite dimensionality counterparts. A mathematically
rigorous treatment of such vector spaces having infinite dimensionality, which
are commonly called Hilbert spaces, can be found in textbooks that are de-
voted to this subject. In this chapter, however, we will only review the main
properties that are useful for quantum mechanics. In some cases, when the
generalization from the case of finite dimensionality to the case of arbitrary
dimensionality is nontrivial, results will be presented without providing a
rigorous proof and even without accurately specifying what are the validity
conditions for these results.

2.1 Linear Vector Space


A linear vector space F is a set {|α} of mathematical objects called vectors.
The space is assumed to be closed under vector addition and scalar multipli-
cation. Both, operations (i.e., vector addition and scalar multiplication) are
commutative. That is:
1. |α + |β = |β + |α ∈ F for every |α ∈ F and |β ∈ F
2. c |α = |α c ∈ F for every |α ∈ F and c ∈ C (where C is the set of
complex numbers)
A vector space with an inner product is called an inner product space.
An inner product of the ordered pair |α , |β ∈ F is denoted as β |α. The
inner product is a function F 2 → C that satisfies the following properties:
Chapter 2. State Vectors and Operators

β |α ∈ C , (2.2)

β |α = α |β , (2.3)
α (c1 |β 1  + c2 |β 2 ) = c1 α |β 1  + c2 α |β 2  , where c1 , c2 ∈ C , (2.4)

α |α ∈ R and α |α ≥ 0. Equality holds iff |α = 0 .


(2.5)
Note that the asterisk in Eq. (2.3) denotes complex conjugate. Below we list
some important definitions and comments regarding inner product:

• The real number α |α is called the norm of the vector |α ∈ F.
• A normalized vector has a unity norm, namely α |α = 1.
• Every nonzero vector 0 = |α ∈ F can be normalized using the transfor-
mation
|α
|α →  . (2.6)
α |α

• The vectors |α ∈ F and |β ∈ F are said to be orthogonal if β |α = 0.


• A set of vectors {|φn }n , where |φn  ∈ F is called a complete orthonormal
basis if
— The vectors are all normalized and orthogonal to each other, namely

φm |φn  = δ nm . (2.7)

— Every |α ∈ F can be written as a superposition of the basis vectors,


namely

|α = cn |φn  , (2.8)
n

where cn ∈ C.
• By evaluating the inner product φm |α, where |α is given by Eq. (2.8)
one finds with the help of Eq. (2.7) and property (2.4) of inner products
that
! "
 
φm |α = φm cn |φn  = cn φm |φn  = cm . (2.9)
n n
  
=δnm

• The last result allows rewriting Eq. (2.8) as


  
|α = cn |φn  = |φn  cn = |φn  φn |α . (2.10)
n n n

Eyal Buks Quantum Mechanics - Lecture Notes 16


2.3. Dirac’s notation

2.2 Operators
Operators, as the definition below states, are function from F to F:
Definition 2.2.1. An operator A : F → F on a vector space maps vectors
onto vectors, namely A |α ∈ F for every |α ∈ F.

Some important definitions and comments are listed below:


• The operators X : F → F and Y : F → F are said to be equal, namely
X = Y , if for every |α ∈ F the following holds

X |α = Y |α . (2.11)

• Operators can be added, and the addition is both, commutative and asso-
ciative, namely
X +Y = Y +X , (2.12)
X + (Y + Z) = (X + Y ) + Z . (2.13)
• An operator A : F → F is said to be linear if

A (c1 |γ 1  + c2 |γ 2 ) = c1 A |γ 1  + c2 A |γ 2  (2.14)

for every |γ 1  , |γ 2  ∈ F and c1 , c2 ∈ C.


• The operators X : F → F and Y : F → F can be multiplied, where

XY |α = X (Y |α) (2.15)

for any |α ∈ F.


• Operator multiplication is associative

X (Y Z) = (XY ) Z = XY Z . (2.16)

• However, in general operator multiplication needs not be commutative

XY = Y X . (2.17)

2.3 Dirac’s notation


In Dirac’s notation the inner product is considered as a multiplication of two
mathematical objects called ’bra’ and ’ket’

β |α = β| |α . (2.18)



bra ket

While the ket-vector |α is a vector in F, the bra-vector β| represents a


functional that maps any ket-vector |α ∈ F to the complex number β |α.

Eyal Buks Quantum Mechanics - Lecture Notes 17


Chapter 2. State Vectors and Operators

While the multiplication of a bra-vector on the left and a ket-vector on the


right represents inner product, the outer product is obtained by reversing the
order

Aαβ = |α β| . (2.19)

The outer product Aαβ is clearly an operator since for any |γ ∈ F the object
Aαβ |γ is a ket-vector

Aαβ |γ = (|β α|) |γ = |β α |γ ∈ F . (2.20)


  
∈C

Moreover, according to property (2.4), Aαβ is linear since for every |γ 1  , |γ 2  ∈


F and c1 , c2 ∈ C the following holds
Aαβ (c1 |γ 1  + c2 |γ 2 ) = |α β| (c1 |γ 1  + c2 |γ 2 )
= |α (c1 β |γ 1  + c2 β |γ 2 )
= c1 Aαβ |γ 1  + c2 Aαβ |γ 2  .
(2.21)
With Dirac’s notation Eq. (2.10) can be rewritten as
! "

|α = |φn  φn | |α . (2.22)
n

Since the above identity holds for any |α ∈ F one concludes that the quantity
in brackets is the identity operator, which is denoted as 1, namely

1= |φn  φn | . (2.23)
n

This result, which is called the closure relation, implies that any complete
orthonormal basis can be used to express the identity operator.

2.4 Dual Correspondence

As we have mentioned above, the bra-vector β| represents a functional map-


ping any ket-vector |α ∈ F to the complex number β |α. Moreover, since
the inner product is linear [see property (2.4) above], such a mapping is linear,
namely for every |γ 1  , |γ 2  ∈ F and c1 , c2 ∈ C the following holds

β| (c1 |γ 1  + c2 |γ 2 ) = c1 β |γ 1  + c2 β |γ 2  . (2.24)

The set of linear functionals from F to C, namely, the set of functionals F : F


→ C that satisfy

Eyal Buks Quantum Mechanics - Lecture Notes 18


2.4. Dual Correspondence

F (c1 |γ 1  + c2 |γ 2 ) = c1 F (|γ 1 ) + c2 F (|γ 2 ) (2.25)

for every |γ 1  , |γ 2  ∈ F and c1 , c2 ∈ C, is called the dual space F ∗ . As


the name suggests, there is a dual correspondence (DC) between F and F ∗ ,
namely a one to one mapping between these two sets, which are both linear
vector spaces. The duality relation is presented using the notation

α| ⇔ |α , (2.26)

where |α ∈ F and α| ∈ F ∗ . What is the dual of the ket-vector |γ =
c1 |γ 1  + c2 |γ 2 , where |γ 1  , |γ 2  ∈ F and c1 , c2 ∈ C? To answer this question
we employ the above mentioned general properties (2.3) and (2.4) of inner
products and consider the quantity γ |α for an arbitrary ket-vector |α ∈ F
γ |α = α |γ∗
= (c1 α |γ 1  + c2 α |γ 2 )∗
= c∗1 γ 1 |α + c2 γ 2 |α
= (c∗1 γ 1 | + c∗2 γ 2 |) |α .
(2.27)
From this result we conclude that the duality relation takes the form

c∗1 γ 1 | + c∗2 γ 2 | ⇔ c1 |γ 1  + c2 |γ 2  . (2.28)

The last relation describes how to map any given ket-vector |β ∈ F
to its dual F = β| : F → C, where F ∈ F ∗ is a linear functional that
maps any ket-vector |α ∈ F to the complex number β |α. What is the
inverse mapping? The answer can take a relatively simple form provided that
a complete orthonormal basis exists, and consequently the identity operator
can be expressed as in Eq. (2.23). In that case the dual of a given linear
functional F : F → C is the ket-vector |FD  ∈ F, which is given by

|FD  = (F (|φn ))∗ |φn  . (2.29)
n

The duality is demonstrated by proving the two claims below:

Claim. |β DD  = |β for any |β ∈ F, where |β DD  is the dual of the dual of
|β.

Proof. The dual of |β is the bra-vector β|, whereas the dual of β| is found
using Eqs. (2.29) and (2.23), thus

Eyal Buks Quantum Mechanics - Lecture Notes 19


Chapter 2. State Vectors and Operators

|β DD  = β |φn ∗ |φn 
n
  
=φn |β

= |φn  φn |β
n

= |φn  φn | |β
n
  
=1
= |β .
(2.30)

Claim. FDD = F for any F ∈ F ∗ , where FDD is the dual of the dual of F .

Proof. The dual |FD  ∈ F of the functional F ∈ F ∗ is given by Eq. (2.29).


Thus with the help of the duality relation (2.28) one finds that dual FDD ∈ F ∗
of |FD  is given by

FDD = F (|φn ) φn | . (2.31)
n

Consider an arbitrary ket-vector |α ∈ F that is written as a superposition


of the complete orthonormal basis vectors, namely

|α = cm |φm  . (2.32)
m

Using the above expression for FDD and the linearity property one finds that

FDD |α = cm F (|φn ) φn |φm 
n,m
  
δmn

= cn F (|φn )
n
! "

=F cn |φn 
n
= F (|α) ,
(2.33)
therefore, FDD = F .

2.5 Matrix Representation


Given a complete orthonormal basis, ket-vectors, bra-vectors and linear op-
erators can be represented using matrices. Such representations are easily
obtained using the closure relation (2.23).

Eyal Buks Quantum Mechanics - Lecture Notes 20


2.5. Matrix Representation

• The inner product between the bra-vector β| and the ket-vector |α can
be written as
β |α = β| 1 |α

= β |φn  φn | α
n
 
φ |α
  φ1 |α
= β |φ1  β |φ2  · · ·  2  .
..
.
(2.34)
Thus, the inner product can be viewed as a product between the row vector
 
β| =˙ β |φ1  β |φ2  · · · , (2.35)
which is the matrix representation of the bra-vector β|, and the column
vector
 
φ1 |α
 
˙ φ2 |α ,
|α = (2.36)
..
.
which is the matrix representation of the ket-vector |α. Obviously, both
representations are basis dependent.
• Multiplying the relation |γ = X |α from the right by the basis bra-vector
φm | and employing again the closure relation (2.23) yields

φm |γ = φm | X |α = φm | X1 |α = φm | X |φn  φn |α , (2.37)
n

or in matrix form
    
φ1 |γ φ1 | X |φ1  φ1 | X |φ2  · · · φ1 |α
φ2 |γ φ2 | X |φ1  φ2 | X |φ2  · · · φ2 |α
 =   . (2.38)
.. .. .. ..
. . . .
In view of this expression, the matrix representation of the linear operator
X is given by
 
φ1 | X |φ1  φ1 | X |φ2  · · ·
 
X= ˙ φ2 | X |φ1  φ2 | X |φ2  · · · . (2.39)
.. ..
. .
Alternatively, the last result can be written as
Xnm = φn | X |φm  , (2.40)
where Xnm is the element in row n and column m of the matrix represen-
tation of the operator X.

Eyal Buks Quantum Mechanics - Lecture Notes 21


Chapter 2. State Vectors and Operators

• Such matrix representation of linear operators can be useful also for mul-
tiplying linear operators. The matrix elements of the product Z = XY are
given by

φm | Z |φn  = φm | XY |φn  = φm | X1Y |φn  = φm | X |φl  φl | Y |φn  .
l
(2.41)

• Similarly, the matrix representation of the outer product |β α| is given
by
 
φ1 |β
  
|β α| =˙ φ2 |β α |φ1  α |φ2  · · ·
..
.
 
φ1 |β α |φ1  φ1 |β α |φ2  · · ·
 
= φ2 |β α |φ1  φ2 |β α |φ2  · · · .
.. ..
. .
(2.42)

2.6 Observables
Measurable physical variables are represented in quantum mechanics by Her-
mitian operators.

2.6.1 Hermitian Adjoint

Definition 2.6.1. The Hermitian adjoint of an operator X is denoted as X †


and is defined by the following duality relation

α| X † ⇔ X |α . (2.43)

Namely, for any ket-vector |α ∈ F, the dual to the ket-vector X |α is the
bra-vector α| X † .
Definition 2.6.2. An operator is said to be Hermitian if X = X † .
Below we prove some simple relations:
Claim. β| X |α = α| X † |β∗
Proof. Using the general property (2.3) of inner products one has
  ∗
β| X |α = β| (X |α) = α| X † |β = α| X † |β∗ . (2.44)

Note that this result implies that if X = X † then β| X |α = α| X |β∗ .

Eyal Buks Quantum Mechanics - Lecture Notes 22


2.6. Observables
 †
Claim. X † = X

Proof. For any |α , |β ∈ F the following holds


 ∗  †
β| X |α = β| X |α∗ = α| X † |β∗ = β| X † |α , (2.45)
 †
thus X † = X.

Claim. (XY ) = Y † X †

Proof. Applying XY on an arbitrary ket-vector |α ∈ F and employing the


duality correspondence yield
 
XY |α = X (Y |α) ⇔ α| Y † X † = α| Y † X † , (2.46)

thus

(XY )† = Y † X † . (2.47)

Claim. If X = |β α| then X † = |α β|

Proof. By applying X on an arbitrary ket-vector |γ ∈ F and employing the


duality correspondence one finds that

X |γ = (|β α|) |γ = |β (α |γ) ⇔ (α |γ)∗ β| = γ |α β| = γ| X † ,
(2.48)

where X † = |α β|.

2.6.2 Eigenvalues and Eigenvectors

Each operator is characterized by its set of eigenvalues, which is defined


below:

Definition 2.6.3. A number an ∈ C is said to be an eigenvalue of an op-


erator A : F → F if for some nonzero ket-vector |an  ∈ F the following
holds

A |an  = an |an  . (2.49)

The ket-vector |an  is then said to be an eigenvector of the operator A with


an eigenvalue an .

The set of eigenvectors associated with a given eigenvalue of an operator


A is called eigensubspace and is denoted as

Fn = {|an  ∈ F such that A |an  = an |an } . (2.50)

Eyal Buks Quantum Mechanics - Lecture Notes 23


Chapter 2. State Vectors and Operators

Clearly, Fn is closed under vector addition and scalar multiplication, namely


c1 |γ 1  + c2 |γ 2  ∈ Fn for every |γ 1 , |γ 2  ∈ Fn and for every c1 , c2 ∈ C. Thus,
the set Fn is a subspace of F. The dimensionality of Fn (i.e., the minimum
number of vectors that are needed to span Fn ) is called the level of degeneracy
gn of the eigenvalue an , namely

gn = dim Fn . (2.51)

As the theorem below shows, the eigenvalues and eigenvectors of a Her-


mitian operator have some unique properties.

Theorem 2.6.1. The eigenvalues of a Hermitian operator A are real. The


eigenvectors of A corresponding to different eigenvalues are orthogonal.

Proof. Let a1 and a2 be two eigenvalues of A with corresponding eigen vectors


|a1  and |a2 
A |a1  = a1 |a1  , (2.52)
A |a2  = a2 |a2  . (2.53)
Multiplying Eq. (2.52) from the left by the bra-vector a2 |, and multiplying
the dual of Eq. (2.53), which since A = A† is given by

a2 | A = a∗2 a2 | , (2.54)

from the right by the ket-vector |a1  yield


a2 | A |a1  = a1 a2 |a1  , (2.55)
a2 | A |a1  = a∗2 a2 |a1  . (2.56)
Thus, we have found that

(a1 − a∗2 ) a2 |a1  = 0 . (2.57)

The first part of the theorem is proven by employing the last result (2.57)
for the case where |a1  = |a2 . Since |a1  is assumed to be a nonzero ket-
vector one concludes that a1 = a∗1 , namely a1 is real. Since this is true for
any eigenvalue of A, one can rewrite Eq. (2.57) as

(a1 − a2 ) a2 |a1  = 0 . (2.58)

The second part of the theorem is proven by considering the case where
a1 = a2 , for which the above result (2.58) can hold only if a2 |a1  = 0.
Namely eigenvectors corresponding to different eigenvalues are orthogonal.

Consider a Hermitian operator A having a set of eigenvalues {an }n . Let


gn be the degree of degeneracy of eigenvalue an , namely gn is the dimension
of the corresponding eigensubspace, which is denoted by Fn . For simplic-
ity, assume that gn is finite for every n. Let {|an,1  , |an,2  , · · · , |an,gn } be

Eyal Buks Quantum Mechanics - Lecture Notes 24


2.6. Observables

an orthonormal basis of the eigensubspace Fn , namely an,i′ |an,i  = δ ii′ .


Constructing such an orthonormal basis for Fn can be done by the so-called
Gram-Schmidt process. Moreover, since eigenvectors of A corresponding to
different eigenvalues are orthogonal, the following holds

an′ ,i′ |an,i  = δ nn′ δ ii′ , (2.59)

In addition, all the ket-vectors |an,i  are eigenvectors of A

A |an,i  = an |an,i  . (2.60)

Projectors. Projector operators are useful for expressing the properties of


an observable.
Definition 2.6.4. An Hermitian operator P is called a projector if P 2 = P .
Claim. The only possible eigenvalues of a projector are 0 and 1.
Proof. Assume that |p is an eigenvector of P with an eigenvalue p, namely
P |p = p |p. Applying the operator P on both sides and using the fact that
P 2 = P yield P |p = p2 |p, thus p (1 − p) |p = 0, therefore since |p is
assumed to be nonzero, either p = 0 or p = 1.

A projector is said to project any given vector onto the eigensubspace


corresponding to the eigenvalue p = 1.
Let {|an,1  , |an,2  , · · · , |an,gn } be an orthonormal basis of an eigensub-
space Fn corresponding to an eigenvalue of an observable A. Such an ortho-
normal basis can be used to construct a projection Pn onto Fn , which is given
by
gn

Pn = |an,i  an,i | . (2.61)
i=1

Clearly, Pn is a projector since Pn† = Pn and since


gn
 gn

Pn2 = |an,i  an,i |an,i′  an,i′ | = |an,i  an,i | = Pn . (2.62)
i,i′ =1
   i=1
δii′

Moreover, it is easy to show using the orthonormality relation (2.59) that the
following holds

Pn Pm = Pm Pn = Pn δ nm . (2.63)

For linear vector spaces of finite dimensionality, it can be shown that the
set {|an,i }n,i forms a complete orthonormal basis of eigenvectors of a given
Hermitian operator A. The generalization of this result for the case of ar-
bitrary dimensionality is nontrivial, since generally such a set needs not be

Eyal Buks Quantum Mechanics - Lecture Notes 25


Chapter 2. State Vectors and Operators

complete. On the other hand, it can be shown that if a given Hermitian oper-
ator A satisfies some conditions (e.g., A needs to be completely continuous)
then completeness is guarantied. For all Hermitian operators of interest for
this course we will assume that all such conditions are satisfied. That is, for
any such Hermitian operator A there exists a set of ket vectors {|an,i }, such
that:
1. The set is orthonormal, namely

an′ ,i′ |an,i  = δ nn′ δ ii′ , (2.64)

2. The ket-vectors |an,i  are eigenvectors, namely

A |an,i  = an |an,i  , (2.65)

where an ∈ R.
3. The set is complete, namely closure relation [see also Eq. (2.23)] is satis-
fied
gn
 
1= |an,i  an,i | = Pn , (2.66)
n i=1 n

where
gn

Pn = |an,i  an,i | (2.67)
i=1

is the projector onto eigen subspace Fn with the corresponding eigenvalue


an .
The closure relation (2.66) can be used to express the operator A in terms
of the projectors Pn
gn
  gn

A = A1 = A |an,i  an,i | = an |an,i  an,i | , (2.68)
n i=1 n i=1

that is

A= an Pn . (2.69)
n

The last result is very useful when dealing with a function f (A) of the
operator A. The meaning of a function of an operator can be understood in
terms of the Taylor expansion of the function

f (x) = fm xm , (2.70)
m

Eyal Buks Quantum Mechanics - Lecture Notes 26


2.6. Observables

where
1 dm f
fm = . (2.71)
m! dxm
With the help of Eqs. (2.63) and (2.69) one finds that

f (A) = fm Am
m
! "m
 
= fm an Pn
m n
 
= fm am
n Pn
m n

= fm am
n Pn ,
n m
  
f (an )

(2.72)
thus

f (A) = f (an ) Pn . (2.73)
n

Exercise 2.6.1. Express the projector Pn in terms of the operator A and


its set of eigenvalues.
Solution 2.6.1. We seek a function f such that f (A) = Pn . Multiplying
from the right by a basis ket-vector |am,i  yields

f (A) |am,i  = δ mn |am,i  . (2.74)

On the other hand

f (A) |am,i  = f (am ) |am,i  . (2.75)

Thus we seek a function that satisfy

f (am ) = δ mn . (2.76)

The polynomial function


#
f (a) = K (a − am ) , (2.77)
m
=n

where K is a constant, satisfies the requirement that f (am ) = 0 for every


m = n. The constant K is chosen such that f (an ) = 1, that is
# a − am
f (a) = , (2.78)
an − am
m
=n

Eyal Buks Quantum Mechanics - Lecture Notes 27


Chapter 2. State Vectors and Operators

Thus, the desired expression is given by


# A − am
Pn = . (2.79)
an − am
m
=n

2.7 Quantum Measurement

Consider a measurement of a physical variable denoted as A(c) performed on


a quantum system. The standard textbook description of such a process is
described below. The physical variable A(c) is represented in quantum me-
chanics by an observable, namely by a Hermitian operator, which is denoted
as A. The correspondence between the variable A(c) and the operator A will
be discussed below in chapter 4. As we have seen above, it is possible to con-
struct a complete orthonormal basis made of eigenvectors of the Hermitian
operator A having the properties given by Eqs. (2.64), (2.65) and (2.66). In
that basis, the vector state |α of the system can be expressed as
gn

|α = 1 |α = an,i |α |an,i  . (2.80)
n i=1

Even when the state vector |α is given, quantum mechanics does not gener-
ally provide a deterministic answer to the question: what will be the outcome
of the measurement. Instead it predicts that:
1. The possible results of the measurement are the eigenvalues {an } of the
operator A.
2. The probability pn to measure the eigen value an is given by
gn

pn = α| Pn |α = |an,i |α|2 . (2.81)
i=1

Note that the state vector |α is assumed to be normalized.


3. After a measurement of A with an outcome an the state vector collapses
onto the corresponding eigensubspace and becomes

Pn |α
|α →  . (2.82)
α| Pn |α

It is easy to show that the probability to measure something is unity


provided that |α is normalized:
! "
  
pn = α| Pn |α = α| Pn |α = 1 . (2.83)
n n n

Eyal Buks Quantum Mechanics - Lecture Notes 28


2.8. Example - Spin 1/2

We also note that a direct consequence of the collapse postulate is that two
subsequent measurements of the same observable performed one immediately
after the other will yield the same result. It is also important to note that the
above ’standard textbook description’ of the measurement process is highly
controversial, especially, the collapse postulate. However, a thorough discus-
sion of this issue is beyond the scope of this course.
Quantum mechanics cannot generally predict the outcome of a specific
measurement of an observable A, however it can predict the average, namely
the expectation value, which is denoted as A. The expectation value is easily
calculated with the help of Eq. (2.69)
 
A = an pn = an α| Pn |α = α| A |α . (2.84)
n n

2.8 Example - Spin 1/2

Spin is an internal degree of freedom of elementary particles. Electrons, for


example, have spin 1/2. This means, as we will see in chapter 6, that the
state of a spin 1/2 can be described by a state vector |α in a vector space
of dimensionality 2. In other words, spin 1/2 is said to be a two-level system.
The spin was first discovered in 1921 by Stern and Gerlach in an experiment
in which the magnetic moment of neutral silver atoms was measured. Silver
atoms have 47 electrons, 46 out of which fill closed shells. It can be shown
that only the electron in the outer shell contributes to the total magnetic
moment of the atom. The force F acting on a magnetic moment µ moving in
a magnetic field B is given by F = ∇ (µ · B). Thus by applying a nonuniform
magnetic field B and by monitoring the atoms’ trajectories one can measure
the magnetic moment.
It is important to keep in mind that generally in addition to the spin
contribution to the magnetic moment of an electron, also the orbital motion
of the electron can contribute. For both cases, the magnetic moment is related
to angular momentum by the gyromagnetic ratio. However this ratio takes
different values for these two cases. The orbital gyromagnetic ratio can be
evaluated by considering a simple example of an electron of charge e moving
in a circular orbit or radius r with velocity v. The magnetic moment is given
by
AI
µorbital = , (2.85)
c
where A = πr2 is the area enclosed by the circular orbit and I = ev/ (2πr)
is the electrical current carried by the electron, thus
erv
µorbital = . (2.86)
2c

Eyal Buks Quantum Mechanics - Lecture Notes 29


Chapter 2. State Vectors and Operators

This result can be also written as


µ
µorbital = B L , (2.87)

where L = me vr is the orbital angular momentum, and where
e
µB = (2.88)
2me c
is the Bohr’s magneton constant. The proportionality factor µB / is the
orbital gyromagnetic ratio. In vector form and for a more general case of
orbital motion (not necessarily circular) the orbital gyromagnetic relation is
given by
µB
µorbital = L. (2.89)

On the other hand, as was first shown by Dirac, the gyromagnetic ratio
for the case of spin angular momentum takes twice this value
2µB
µspin = S. (2.90)

Note that we follow here the convention of using the letter L for orbital
angular momentum and the letter S for spin angular momentum.
The Stern-Gerlach apparatus allows measuring any component of the
magnetic moment vector. Alternatively, in view of relation (2.90), it can be
said that any component of the spin angular momentum S can be measured.
The experiment shows that the only two possible results of such a measure-
ment are +/2 and −/2. As we have seen above, one can construct a com-
plete orthonormal basis to the vector space made of eigenvectors of any given
observable. Choosing the observable Sz = S · ẑ for this purpose we construct
a basis made of two vectors {|+; ẑ , |−; ẑ}. Both vectors are eigenvectors of
Sz

Sz |+; ẑ = |+; ẑ , (2.91)
2

Sz |−; ẑ = − |−; ẑ . (2.92)
2
In what follow we will use the more compact notation
|+ = |+; ẑ , (2.93)
|− = |−; ẑ . (2.94)
The orthonormality property implied that
+ |+ = − |− = 1 , (2.95)
− |+ = 0 . (2.96)
The closure relation in the present case is expressed as

Eyal Buks Quantum Mechanics - Lecture Notes 30


2.8. Example - Spin 1/2

|+ +| + |− −| = 1 . (2.97)

In this basis any ket-vector |α can be written as

|α = |+ + |α + |− − |α . (2.98)

The closure relation (2.97) and Eqs. (2.91) and (2.92) yield


Sz = (|+ +| − |− −|) (2.99)
2
It is useful to define also the operators S+ and S−
S+ =  |+ −| , (2.100)
S− =  |− +| . (2.101)
In chapter 6 we will see that the x and y components of S are given by

Sx = (|+ −| + |− +|) , (2.102)
2

Sy = (−i |+ −| + i |− +|) . (2.103)
2
All these ket-vectors and operators have matrix representation, which for the
basis {|+; ẑ , |−; ẑ} is given by
 
1
|+ =˙ , (2.104)
0
 
0
|− =˙ , (2.105)
1
 
 01
Sx =˙ , (2.106)
2 10
 
 0 −i
Sy =
˙ , (2.107)
2 i 0
 
 1 0
Sz =
˙ , (2.108)
2 0 −1
 
01
S+ =˙  , (2.109)
00
 
00
S− =˙  . (2.110)
10

Exercise 2.8.1. Given that the state vector of a spin 1/2 is |+; ẑ calculate
(a) the expectation values Sx  and Sz  (b) the probability to obtain a value
of +/2 in a measurement of Sx .

Solution 2.8.1. (a) Using the matrix representation one has

Eyal Buks Quantum Mechanics - Lecture Notes 31


Chapter 2. State Vectors and Operators
  
  01 1
Sx  = +| Sx |+ = 10 =0, (2.111)
2 10 0
  
  1 0 1 
Sz  = +| Sz |+ = 10 = . (2.112)
2 0 −1 0 2
(b) First, the eigenvectors of the operator Sx are found by solving the equa-
tion Sx |α = λ |α, which is done by diagonalization of the matrix represen-
tation of Sx . The relation Sx |α = λ |α for the two eigenvectors is written
in a matrix form as
  ! √1 " ! "
√1
 01 2  2
√1
= , (2.113)
2 10 2
2 √12
  ! √1 " !
√1
"
 01 2  2
=− . (2.114)
2 10 − √12 2 − √12
That is, in ket notation

Sx |±; x̂ = ± |±; x̂ , (2.115)
2
where the eigenvectors of Sx are given by
1
|±; x̂ = √ (|+ ± |−) . (2.116)
2
Using this result the probability p+ is easily calculated


2

1
1
p+ = |+ |+; x̂|2 =

+| √ (|+ + |−)

= . (2.117)
2 2

2.9 Unitary Operators

Unitary operators are useful for transforming from one orthonormal basis to
another.

Definition 2.9.1. An operator U is said to be unitary if U † = U −1 , namely


if U U † = U † U = 1.

Consider two observables A and B, and two corresponding complete and


orthonormal bases of eigenvectors

A |an  = an |an  , am |an  = δ nm , |an  an | = 1 , (2.118)
n

B |bn  = bn |bn  , bm |bn  = δ nm , |bn  bn | = 1 . (2.119)
n

The operator U, which is given by

Eyal Buks Quantum Mechanics - Lecture Notes 32


2.10. Trace

U= |bn  an | , (2.120)
n

transforms each of the basis vector |an  to the corresponding basis vector |bn 

U |an  = |bn  . (2.121)

It is easy to show that the operator U is unitary


 
U †U = |an  bn |bm  am | = |an  an | = 1 . (2.122)
n,m
   n
δnm

The matrix elements of U in the basis {|an } are given by

an | U |am  = an |bm  , (2.123)

and those of U † by

an | U † |am  = bn |am  .

Consider a ket vector



|α = |an  an |α , (2.124)
n

which can be represented as a column vector in the basis {|an }. The nth
element of such a column vector is an |α. The operator U can be employed
for finding the corresponding column vector representation of the same ket-
vector |α in the other basis {|bn }
 
bn |α = bn |am  am |α = an | U † |am  am |α . (2.125)
m m

Similarly, Given an operator X the relation between the matrix elements


an | X |am  in the basis {|an } to the matrix elements bn | X |bm  in the
basis {|bn } is given by

bn | X |bm  = bn |ak  ak | X |al  al |bm 
k,l

= an | U † |ak  ak | X |al  al | U |am  .
k,l
(2.126)

2.10 Trace
Given an operator X and an orthonormal and complete basis {|an }, the
trace of X is given by

Eyal Buks Quantum Mechanics - Lecture Notes 33


Chapter 2. State Vectors and Operators

Tr (X) = an | X |an  . (2.127)
n

It is easy to show that Tr (X) is independent on basis, as is shown below:



Tr (X) = an | X |an 
n

= an |bk  bk | X |bl  bl |an 
n,k,l

= bl |an  an |bk  bk | X |bl 
n,k,l

= bl |bk  bk | X |bl 
  
k,l
δkl

= bk | X |bk  .
k
(2.128)
The proof of the following relation

Tr (XY ) = Tr (Y X) , (2.129)

is left as an exercise.

2.11 Commutation Relation


The commutation relation of the operators A and B is defined as

[A, B] = AB − BA . (2.130)

As an example, the components Sx , Sy and Sz of the spin angular momentum


operator, satisfy the following commutation relations

[Si , Sj ] = iεijk Sk , (2.131)

where

 0 i, j, k are not all different
εijk = 1 i, j, k is an even permutation of x, y, z (2.132)

−1 i, j, k is an odd permutation of x, y, z

is the Levi-Civita symbol. Equation (2.131) employs the Einstein’s conven-


tion, according to which if an index symbol appears twice in an expression,
it is to be summed over all its allowed values. Namely, the repeated index k
should be summed over the values x, y and z:

εijk Sk = εijx Sx + εijy Sy + εijz Sz . (2.133)

Eyal Buks Quantum Mechanics - Lecture Notes 34


2.12. Simultaneous Diagonalization of Commuting Operators

Moreover, the following relations hold


1
Sx2 = Sy2 = Sz2 = 2 , (2.134)
4
3
S2 = Sx2 + Sy2 + Sz2 = 2 . (2.135)
4
The relations below, which are easy to prove using the above definition,
are very useful for evaluating commutation relations
[F, G] = − [G, F ] , (2.136)
[F, F ] = 0 , (2.137)
[E + F, G] = [E, G] + [F, G] , (2.138)
[E, F G] = [E, F ] G + F [E, G] . (2.139)

2.12 Simultaneous Diagonalization of Commuting


Operators

Consider an observable A having a set of eigenvalues {an }. Let gn be the


degree of degeneracy of eigenvalue an , namely gn is the dimension of the
corresponding eigensubspace, which is denoted by Fn . Thus the following
holds

A |an,i  = an |an,i  , (2.140)

where i = 1, 2, · · · , gn , and

an′ ,i′ |an,i  = δ nn′ δ ii′ . (2.141)

The set of vectors {|an,1  , |an,2  , · · · , |an,gn } forms an orthonormal basis for
the eigensubspace Fn . The closure relation can be written as
gn
 
1= |an,i  an,i | = Pn , (2.142)
n i=1 n

where
gn

Pn = |an,i  an,i | . (2.143)
i=1

Now consider another observable B, which is assumed to commute with


A, namely [A, B] = 0.

Claim. The operator B has a block diagonal matrix in the basis {|an,i },
namely am,j | B |an,i  = 0 for n = m.

Eyal Buks Quantum Mechanics - Lecture Notes 35


Chapter 2. State Vectors and Operators

Proof. Multiplying Eq. (2.140) from the left by am,j | B yields

am,j | BA |an,i  = an am,j | B |an,i  . (2.144)

On the other hand, since [A, B] = 0 one has

am,j | BA |an,i  = am,j | AB |an,i  = am am,j | B |an,i  , (2.145)

thus

(an − am ) am,j | B |an,i  = 0 . (2.146)

For a given n, the gn × gn matrix an,i′ | B |an,i  is Hermitian, namely



an,i′ | B |an,i  = an,i | B |an,i′  . Thus, there exists a unitary transformation
Un , which maps Fn onto Fn , and which diagonalizes the block of B in the
subspace Fn . Since Fn is an eigensubspace of A, the block matrix of A in the
new basis remains diagonal (with the eigenvalue an ). Thus, we conclude that
a complete and orthonormal basis of common eigenvectors of both operators
A and B exists. For such a basis, which is denoted as {|n, m}, the following
holds
A |n, m = an |n, m , (2.147)
B |n, m = bm |n, m . (2.148)

2.13 Uncertainty Principle

Consider a quantum system in a state |n, m, which is a common eigenvector


of the commuting observables A and B. The outcome of a measurement of
the observable A is expected to be an with unity probability, and similarly,
the outcome of a measurement of the observable B is expected to be bm
with unity probability. In this case it is said that there is no uncertainty
corresponding to both of these measurements.

Definition 2.13.1. The variance in a measurement


' (of a given observable A
2
of a quantum system in a state |α is given by (∆A) , where ∆A = A−A,
namely
' ( ' ( ) *
(∆A)2 = A2 − 2A A + A2 = A2 − A2 , (2.149)

where
A = α| A |α , (2.150)
) 2*
A = α| A2 |α . (2.151)

Eyal Buks Quantum Mechanics - Lecture Notes 36


2.13. Uncertainty Principle

Example 2.13.1. Consider a spin 1/2 system in a state |α = |+; ẑ. Using
Eqs. (2.99), (2.102) and (2.134) one finds that
' ( ) *
2 2
(∆Sz ) = Sz2 − Sz  = 0 , (2.152)
' ( ) * 1
(∆Sx )2 = Sx2 − Sx 2 = 2 . (2.153)
4
The last example raises the question: can one find a state |α for which
the variance in the measurements of both Sz and Sx vanishes? According to
the uncertainty principle the answer is no.
Theorem 2.13.1. The uncertainty principle - Let A and B be two observ-
ables. For any ket-vector |α the following holds
' (' ( 1
(∆A)2 (∆B)2 ≥ |[A, B]|2 . (2.154)
4
Proof. Applying the Schwartz inequality [see Eq. (2.166)], which is given by

u |u v |v ≥ |u |v|2 , (2.155)

for the ket-vectors


|u = ∆A |α , (2.156)
|v = ∆B |α , (2.157)
and exploiting the fact that (∆A)† = ∆A and (∆B)† = ∆B yield
' (' (
(∆A)2 (∆B)2 ≥ |∆A∆B|2 . (2.158)

The term ∆A∆B can be written as


1 1
∆A∆B = [∆A, ∆B] + [∆A, ∆B]+ , (2.159)
2 2
where
[∆A, ∆B] = ∆A∆B − ∆B∆A , (2.160)
[∆A, ∆B]+ = ∆A∆B + ∆B∆A . (2.161)
While the term [∆A, ∆B] is anti-Hermitian, whereas the term [∆A, ∆B]+ is
Hermitian, namely
([∆A, ∆B])† = (∆A∆B − ∆B∆A)† = ∆B∆A − ∆A∆B = − [∆A, ∆B] ,
 †
[∆A, ∆B]+ = (∆A∆B + ∆B∆A)† = ∆B∆A + ∆A∆B = [∆A, ∆B]+ .
In general, the following holds
+
α| X |α∗ if X is Hermitian
α| X |α = α| X † |α∗ = , (2.162)
− α| X |α∗ if X is anti-Hermitian

Eyal Buks Quantum Mechanics - Lecture Notes 37


Chapter 2. State Vectors and Operators

thus
1 1) *
∆A∆B = [∆A, ∆B] + [∆A, ∆B]+ , (2.163)
2   2  
∈I ∈R

and consequently
1 1
) *
2
|∆A∆B|2 = |[∆A, ∆B]|2 +
[∆A, ∆B]+
. (2.164)
4 4
Finally, with the help of the identity [∆A, ∆B] = [A, B] one finds that
' (' ( 1
2
(∆A) (∆B)2 ≥ |[A, B]|2 . (2.165)
4

2.14 Problems
1. Derive the Schwartz inequality
 
|u |v| ≤ u |u v |v , (2.166)

where |u and |v are any two vectors of a vector space F.
2. Derive the triangle inequality:
  
(u| + v|) (|u + |v) ≤ u |u + v |v . (2.167)

3. Show that if a unitary operator U can be written in the form U = 1+iǫF ,


where ǫ is a real infinitesimally small number, then the operator F is
Hermitian.
4. A Hermitian operator A is said to be positive-definite if, for any vector
|u, u| A |u ≥ 0. Show that the operator A = |a a| is Hermitian and
positive-definite.
5. Show that if A is a Hermitian positive-definite operator then the following
hold
 
|u| A |v| ≤ u| A |u v| A |v . (2.168)

6. Find the expansion of the operator (A − λB)−1 in a power series in λ ,


assuming that the inverse A−1 of A exists.
7. The derivative of an operator A (λ) which depends explicitly on a para-
meter λ is defined to be
dA (λ) A (λ + ǫ) − A (λ)
= lim . (2.169)
dλ ǫ→0 ǫ
Show that
d dA dB
(AB) = B+A . (2.170)
dλ dλ dλ

Eyal Buks Quantum Mechanics - Lecture Notes 38


2.14. Problems

8. Show that
d  −1  dA −1
A = −A−1 A . (2.171)
dλ dλ
9. Let |u and |v be two vectors of finite norm. Show that

Tr (|u v|) = v |u . (2.172)

10. If A is any linear operator, show that A† A is a positive-definite Her-


mitian operator whose trace is equal to the sum of the square moduli of
the matrix
 elements of A in any arbitrary representation. Deduce that
Tr A† A = 0 is true if and only if A = 0.
11. Show that if A and B are two positive-definite observables, then Tr (AB) ≥
0.
12. Show that for any two operators A and L
1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · . (2.173)
2! 3!
13. Show that if A and B are two operators satisfying the relation [[A, B] , A] =
0 , then the relation

[Am , B] = mAm−1 [A, B] (2.174)

holds for all positive integers m .


14. Show that

eA eB = eA+B e(1/2)[A,B] , (2.175)

provided that [[A, B] , A] = 0 and [[A, B] , B] = 0.


15. Proof Kondo’s identity


, -
A, e−βH = e−βH eλH [H, A] e−λH dλ , (2.176)
0

where A and H are any two operators and β is real.


16. Show that Tr (XY ) = Tr (Y X).
17. Consider the two normalized spin 1/2 states |α and |β. The operator
A is defined as

A = |α α| − |β β| . (2.177)

Find the eigenvalues of the operator A.


18. A molecule is composed of six identical atoms A1 , A2 , . . . , A6 which
form a regular hexagon. Consider an electron, which can be localized on
each of the atoms. Let |ϕn  be the state in which it is localized on the
nth atom (n = 1, 2, · · · , 6). The electron states will be confined to the

Eyal Buks Quantum Mechanics - Lecture Notes 39


Chapter 2. State Vectors and Operators

space spanned by the states |ϕn , which is assumed to be orthonormal.


The Hamiltonian of the system is given by

H = H0 + W . (2.178)

The eigenstates of H0 are the six states |ϕn , with the same eigenvalue
E0 . The operator W is described by
W |ϕ1  = −a |ϕ2  − a |ϕ6  ,
W |ϕ2  = −a |ϕ3  − a |ϕ1  ,
..
.
W |ϕ6  = −a |ϕ1  − a |ϕ5  .
(2.179)
Find the eigenvalues and eigen vectors of H. Clue: Consider a solution of
the form
6

|k = eikn |ϕn  . (2.180)
n=1

2.15 Solutions

1. Let

|γ = |u + λ |v , (2.181)

where λ ∈ C. The requirement γ |γ ≥ 0 leads to

u |u + λ u |v + λ∗ v |u + |λ|2 v |v ≥ 0 . (2.182)

By choosing

v |u
λ=− , (2.183)
v |v

one has

v |u u |v
v |u
2
u |u − u |v −

v |u +

v |v ≥ 0 , (2.184)
v |v v |v v |v

thus
 
|u |v| ≤ u |u v |v . (2.185)

Eyal Buks Quantum Mechanics - Lecture Notes 40


2.15. Solutions

2. The following holds


(u| + v|) (|u + |v) = u |u + v |v + 2 Re (u |v)
≤ u |u + v |v + 2 |u |v| .
(2.186)
Thus, using Schwartz inequality one has
 
(u| + v|) (|u + |v) ≤ u |u + v |v + 2 u |u v |v
  2
= u |u + v |v .
(2.187)
3. Since
     
1 = U † U = 1 − iǫF † (1 + iǫF ) = 1 + iǫ F − F † + O ǫ2 , (2.188)
one has F = F † .

4. In general, (|β α|) = |α β|, thus clearly the operator A is Hermitian.
Moreover it is positive-definite since for every |u the following holds
u| A |u = u |a a |u = |a |u|2 ≥ 0 . (2.189)
5. Let
v| A |u
|γ = |u − |v .
v| A |v
Since A is Hermitian and positive-definite the following holds
0 ≤ γ| A |γ
   
u| A |v v| A |u
= u| − v| A |u − |v
v| A |v v| A |v
|u| A |v|2 |u| A |v|2 |u| A |v|2
= u| A |u − − + ,
v| A |v v| A |v v| A |v
(2.190)
thus
 
|u| A |v| ≤ u| A |u v| A |v . (2.191)
Note that this result allows easy proof of the following: Under the same
conditions (namely, A is a Hermitian positive-definite operator) Tr (A) =
0 if and only if A = 0.
6. The expansion is given by
  −1
(A − λB)−1 = A 1 − λA−1 B
 −1 −1
= 1 − λA−1 B A
  −1 2  −1 3 
= 1 + λA B + λA B + λA B + · · · A−1 .
−1

(2.192)

Eyal Buks Quantum Mechanics - Lecture Notes 41


Chapter 2. State Vectors and Operators

7. By definition:
d A (λ + ǫ) B (λ + ǫ) − A (λ) B (λ)
(AB) = lim
dλ ǫ→0 ǫ
(A (λ + ǫ) − A (λ)) B (λ) A (λ + ǫ) (B (λ + ǫ) − B (λ))
= lim + lim
ǫ→0 ǫ ǫ→0 ǫ
dA dB
= B+A .
dλ dλ
(2.193)
8. Taking the derivative of both sides of the identity 1 = AA−1 on has
dA −1 dA−1
0= A +A , (2.194)
dλ dλ
thus
d  −1  dA −1
A = −A−1 A . (2.195)
dλ dλ
9. Let {|n} be a complete orthonormal basis, namely

|n n| = 1 . (2.196)
n

In this basis
! "
 
Tr (|u v|) = n |u v |n = v| |n n| |u = v |u . (2.197)
n n
 †
10. The operator A† A is Hermitian since A† A = A† A, and positive-
definite since the norm of A |u is nonnegative for every |u, thus one
has u| A† A |u ≥ 0. Moreover, using a complete orthonormal basis {|n}
one has
  
Tr A† A = n| A† A |n
n

= n| A† |m m| A |n
n,m
 2
= |m| A |n| .
n,m
(2.198)

11. Let {|b } be a complete orthonormal basis made of eigenvectors of B
(i.e., B |b′  = b′ |b′ ). Using this basis for evaluating the trace one has
 
Tr (AB) = b′ | AB |b′  = b′ b′ | A |b′  ≥ 0 .
 (2.199)
′ ′
  
b b ≥0 ≥0

Eyal Buks Quantum Mechanics - Lecture Notes 42


2.15. Solutions

12. Let f (s) = esL Ae−sL , where s is real. Using Taylor expansion one has

1 df

1 d2 f

f (1) = f (0) + + +··· , (2.200)


1! ds
s=0 2! ds2
s=0

thus

1 df

1 d2 f

eL Ae−L = A + + +··· , (2.201)


1! ds
s=0 2! ds2
s=0

where
df
= LesL Ae−sL − esL Ae−sL L = [L, f (s)] , (2.202)
ds 
d2 f df
= L, = [L, [L, f (s)]] , (2.203)
ds2 ds
therefore
1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · . (2.204)
2! 3!
13. The identity clearly holds for the case m = 1. Moreover, assuming it
holds for m, namely assuming that

[Am , B] = mAm−1 [A, B] , (2.205)

one has
, m+1 -
A , B = A [Am , B] + [A, B] Am
= mAm [A, B] + [A, B] Am .
(2.206)
It is easy to show that if [[A, B] , A] = 0 then [[A, B] , Am ] = 0, thus one
concludes that
, m+1 -
A , B = (m + 1) Am [A, B] . (2.207)

14. Define the function f (s) = esA esB , where s is real. The following holds
df
= AesA esB + esA BesB
ds  
= A + esA Be−sA esA esB

Using Eq. (2.174) one has

Eyal Buks Quantum Mechanics - Lecture Notes 43


Chapter 2. State Vectors and Operators

∞ m
sA (sA)
e B= B
m=0
m!
∞
sm (BAm + [Am , B])
=
m=0
m!
∞  
 s BA + mAm−1 [A, B]
m m
=
m=0
m!
∞ m−1
(sA)
= BesA + s [A, B]
m=1
(m − 1)!
= BesA + sesA [A, B] ,
(2.208)
thus
df
= AesA esB + BesA esB + sesA [A, B] esB
ds
= (A + B + [A, B] s) f (s) .
(2.209)
The above differential equation can be easily integrated since [[A, B] , A] =
0 and [[A, B] , B] = 0. Thus
s2
f (s) = e(A+B)s e[A,B] 2 . (2.210)

For s = 1 one gets

eA eB = eA+B e(1/2)[A,B] . (2.211)

15. Define
, -
f (β) ≡ A, e−βH , (2.212)

−βH
g (β) ≡ e eλH [H, A] e−λH dλ . (2.213)
0

Clearly, f (0) = g (0) = 0 . Moreover, the following holds


df
= −AHe−βH + He−βH A = −Hf + [H, A] e−βH , (2.214)

dg
= −Hg + [H, A] e−βH , (2.215)

namely, both functions satisfy the same differential equation. Therefore
f = g.
16. Using a complete orthonormal basis {|n} one has

Eyal Buks Quantum Mechanics - Lecture Notes 44


2.15. Solutions

Tr (XY ) = n| XY |n
n

= n| X |m m| Y |n
n,m

= m| Y |n n| X |m
n,m

= m| Y X |m
m
= Tr (Y X) .
(2.216)
+
Note that using this result it is easy to show that Tr (U XU ) = Tr (X)
, provided that U is a unitary operator.
17. Clearly A is Hermitian, namely A† = A, thus the two eigenvalues λ1
and λ2 are expected to be real. Since the trace of an operator is basis
independent, the following must hold

Tr (A) = λ1 + λ2 , (2.217)

and
 
Tr A2 = λ21 + λ22 . (2.218)

On the other hand, with the help of Eq. (2.172) one finds that

Tr (A) = Tr (|α α|) − Tr (|β β|) = 0 , (2.219)

and
 
Tr A2 = Tr (|α α |α α|) + Tr (|β β |β β|) − Tr (|α α |β β|) − Tr (|β β |α α|)
= 2 − α |β Tr (|α β|) − β |α Tr (|β α|)
 
= 2 1 − |α |β|2 ,
(2.220)
thus
.
λ± = ± 1 − |α |β|2 . (2.221)

Alternatively, this problem can also be solved as follows. In general, the


state |β can be decomposed into a parallel to and a perpendicular to |α
terms, namely

|β = a |α + c |γ , (2.222)

where a, c ∈ C, the vector |γ is orthogonal to |α, namely γ |α = 0, and


in addition |γ is assumed to be normalized, namely γ |γ = 1. Since |β

Eyal Buks Quantum Mechanics - Lecture Notes 45


Chapter 2. State Vectors and Operators

is normalized one has |a|2 + |c|2 = 1. The matrix representation of A in


the orthonormal basis {|α , |γ} is given by
   2 
α| A |α α| A |γ |c| −ac∗
A=˙ = ≡ Â . (2.223)
γ| A |α γ| A |γ −a∗ c − |c|2

Thus,
 
Tr  = 0 , (2.224)

and
     
2 2 2 2
Det  = − |c| |c| + |a| = − 1 − |α |β| , (2.225)

therefore the eigenvalues are


.
λ± = ± 1 − |α |β|2 . (2.226)

18. Following the clue, we seek a solution to the eigenvalue equation

H |k = Em |k , (2.227)

where
6

|k = eikn |ϕn  . (2.228)
n=1

thus
6
 
(
(

H |k = E0 |k − a eikn


ϕ(n−1)′ +
ϕ(n+1)′ = E |k , (2.229)
n=1

where n′ is the modulus of n divided by 6 (e.g., 1′ = 1, 0′ = 6, 7′ = 1).


A solution is obtained if

e6ik = 1 , (2.230)

or

km = , (2.231)
3
where m = 1, 2, · · · , 6. The corresponding eigen vectors are denoted as
6

|km  = eikm n |ϕn  , (2.232)
n=1

and the following holds

Eyal Buks Quantum Mechanics - Lecture Notes 46


2.15. Solutions

6

( 6

(

H |km  = E0 |km  − aeikm eikm (n−1)


ϕ(n−1)′ − ae−ikm eikm (n+1)
ϕ(n+1)′
n=1 n=1
= (E0 − 2a cos km ) |km  ,
(2.233)
thus

H |km  = Em |k , (2.234)

where

Em = E0 − 2a cos km . (2.235)

Eyal Buks Quantum Mechanics - Lecture Notes 47


3. The Position and Momentum Observables

Consider a point particle moving in a 3 dimensional space. We first treat


the system classically. The position of the particle is described using the
Cartesian coordinates qx , qy and qz . Let
∂L
pj = (3.1)
∂ q̇j
be the canonically conjugate variable to the coordinate qj , where j ∈ {x, y, z}
and where L is the Lagrangian. As we have seen in exercise 4 of set 1, the
following Poisson’s brackets relations hold
{qj , qk } = 0 , (3.2)
{pj , pk } = 0 , (3.3)
{qj , pk } = δ jk . (3.4)
In quantum mechanics, each of the 6 variables qx , qy , qz , px , py and pz is
represented by an Hermitian operator, namely by an observable. It is postu-
lated that the commutation relations between each pair of these observables
is related to the corresponding Poisson’s brackets according to the rule
1
{, } → [, ] . (3.5)
i
Namely the following is postulated to hold
[qj , qk ] = 0 , (3.6)
[pj , pk ] = 0 , (3.7)
[qj , pk ] = iδ jk . (3.8)
Note that here we use the same notation for a classical variable and its
quantum observable counterpart. In this chapter we will derive some results
that are solely based on Eqs. (3.6), (3.7) and (3.8).

3.1 The One Dimensional Case


In this section, which deals with the relatively simple case of a one dimen-
sional motion of a point particle, we employ the less cumbersome notation
Chapter 3. The Position and Momentum Observables

x and p for the observables qx and px . The commutation relation between


these operators is given by [see Eq. (3.8)]

[x, p] = i . (3.9)

The uncertainty principle (2.154) employed for x and p yields


' (' ( 2
2 2
(∆x) (∆px ) ≥ . (3.10)
4

3.1.1 Position Representation

Let x′ be an eigenvalue of the observable x, and let |x′  be the corresponding


eigenvector, namely

x |x′  = x′ |x′  . (3.11)

Note that x′ ∈ R since x is Hermitian. As we will see below transformation


between different eigenvectors |x′  can be performed using the translation
operator J (∆x ).

Definition 3.1.1. The translation operator is given by


 
i∆x p
J (∆x ) = exp − , (3.12)


where ∆x ∈ R.

Recall that in general the meaning of a function of an operator can be


understood in terms of the Taylor expansion of the function, that is, for the
present case
∞  n
1 i∆x p
J (∆x ) = − . (3.13)
n=0
n! 

It is easy to show that J (∆x ) is unitary

J † (∆x ) = J (−∆x ) = J −1 (∆x ) . (3.14)

Moreover, the following composition property holds

J (∆x1 ) J (∆x2 ) = J (∆x1 + ∆x2 ) . (3.15)

Theorem 3.1.1. Let x′ be an eigenvalue of the observable x, and let |x′  be


the corresponding eigenvector. Then the ket-vector J (∆x ) |x′  is a normalized
eigenvector of x with an eigenvalue x′ + ∆x .

Eyal Buks Quantum Mechanics - Lecture Notes 50


3.1. The One Dimensional Case

Proof. With the help of Eq. (3.77), which is given by


dB
[x, B (p)] = i , (3.16)
dp
and which is proven in exercise 1 of set 3, one finds that
∆x
[x, J (∆x )] = i J (∆x ) . (3.17)
i
Using this result one has

xJ (∆x ) |x′  = ([x, J (∆x )] + J (∆x ) x) |x′  = (x′ + ∆x ) J (∆x ) |x′  , (3.18)

thus the ket-vector J (∆x ) |x′  is an eigenvector of x with an eigenvalue x′ +


∆x . Moreover, J (∆x ) |x′  is normalized since J is unitary.
In view of the above theorem we will in what follows employ the notation

J (∆x ) |x′  = |x′ + ∆x  . (3.19)

An important consequence of the last result is that the spectrum of eigenval-


ues of the operator x is continuous and contains all real numbers. This point
will be further discussed below.
The position wavefunction ψα (x′ ) of a state vector |α is defined as:

ψα (x′ ) = x′ |α . (3.20)

Given the wavefunction ψα (x′ ) of state vector |α, what is the wavefunction
of the state O |α, where O is an operator? We will answer this question below
for some cases:
1. The operator O = x. In this case

x′ | x |α = x′ x′ |α = x′ ψα (x′ ) , (3.21)

namely, the desired wavefunction is obtained by multiplying ψα (x′ ) by


x′ .
2. The operator O is a function A (x) of the operator x. Let

A (x) = an xn . (3.22)
n

be the Taylor expansion of A (x). Exploiting the fact that x is Hermitian


one finds that
 
x′ | A (x) |α = an x′ | xn |α = an x′n x′ |α = A (x′ ) ψα (x′ ) .
n
   n
x′n x′ |
(3.23)

Eyal Buks Quantum Mechanics - Lecture Notes 51


Chapter 3. The Position and Momentum Observables

3. The operator O = J (∆x ). In this case

x′ | J (∆x ) |α = x′ | J † (−∆x ) |α = x′ − ∆x |α = ψα (x′ − ∆x ) .


(3.24)

4. The operator O = p. In view of Eq. (3.12), the following holds


   
ip∆x i∆x
J (−∆x ) = exp =1+ p + O (∆x )2 , (3.25)
 
thus
i∆x ′  
2
x′ | J (−∆x ) |α = ψα (x′ ) + x | p |α + O (∆x ) . (3.26)

On the other hand, according to Eq. (3.24) also the following holds

x′ | J (−∆x ) |α = ψα (x′ + ∆x ) . (3.27)

Equating the above two expressions for x′ | J (−∆x ) |α yields

ψα (x′ + ∆x ) − ψα (x′ )
x′ | p |α = −i + O (∆x ) . (3.28)
∆x
Thus, in the limit ∆x → 0 one has
dψα
x′ | p |α = −i . (3.29)
dx′
To mathematically understand the last result, consider the differential
operator
 
˜ d
J (−∆x ) = exp ∆x
dx
 2
d 1 d
= 1 + ∆x + ∆x +··· .
dx 2! dx
(3.30)
In view of the Taylor expansion of an arbitrary function f (x)
df (∆x )2 d2 f
f (x0 + ∆x ) = f (x0 ) + ∆x + +···
 dx

 2! dx2
d

= exp ∆x f

dx x=x0

= J˜ (−∆x ) f
,
x=x0
(3.31)
one can argue that the operator J˜ (−∆x ) generates translation.

Eyal Buks Quantum Mechanics - Lecture Notes 52


3.1. The One Dimensional Case

As we have pointed out above, the spectrum (i.e., the set of all eigen-
values) of x is continuous. On the other hand, in the discussion in chapter
2 only the case of an observable having discrete spectrum has been consid-
ered. Rigorous mathematical treatment of the case of continuous spectrum
is nontrivial mainly because typically the eigenvectors in such a case cannot
be normalized. However, under some conditions one can generalize some of
the results given in chapter 2 for the case of an observable having a continu-
ous spectrum. These generalization is demonstrated below for the case of the
position operator x:
1. The closure relation (2.23) is written in terms of the eigenvectors |x′  as
∞
dx′ |x′  x′ | = 1 , (3.32)
−∞

namely, the discrete sum is replaced by an integral.


2. With the help of Eq. (3.32) an arbitrary ket-vector can be written as
∞ ∞
′ ′ ′
|α = dx |x  x |α = dx′ ψα (x′ ) |x′  , (3.33)
−∞ −∞

and the inner product between a ket-vector |α and a bra-vector β| as
∞ ∞
β |α = ′ ′
dx β |x  x |α = ′
dx′ ψ ∗β (x′ ) ψα (x′ ) . (3.34)
−∞ −∞

3. The normalization condition reads


∞
2
1 = α |α = dx′ |ψα (x′ )| . (3.35)
−∞

4. The orthonormality relation (2.64) is written in the present case as

x′′ |x′  = δ (x′ − x′′ ) . (3.36)

Note that the above orthonormality relation (3.36) is consistent with


the closure relation (3.32). This can be seen by evaluating the operator
12 = 1 × 1 using Eqs. (3.32) and (3.36)
∞ ∞ ∞
2 ′ ′′ ′′ ′′ ′ ′
1 = dx dx |x  x |x  x | = dx′ |x′  x′ | , (3.37)
  
−∞ −∞ δ(x′ −x′′ ) −∞

thus, as expected 12 = 1.

Eyal Buks Quantum Mechanics - Lecture Notes 53


Chapter 3. The Position and Momentum Observables

5. In a measurement of the observable x, the quantity


2 2
f (x′ ) = |x′ |α| = |ψα (x′ )| (3.38)

represents the probability distribution function to find the particle at the


point x = x′ .
6. That is, the probability to find the particle in the interval (x1 , x2 ) is given
by
 x2
p(x1 ,x2 ) = dx′ f (x′ ) . (3.39)
x1

This can be rewritten as

p(x1 ,x2 ) = α| P(x1 ,x2 ) |α , (3.40)

where the projection operator P(x1 ,x2 ) is given by


 x2
P(x1 ,x2 ) = dx′ |x′  x′ | . (3.41)
x1

The operator P(x1 ,x2 ) is considered to be a projection operator since for


every x0 ∈ (x1 , x2 ) the following holds
 x2
P(x1 ,x2 ) |x0  = dx′ |x′  x′ |x0  = |x0  . (3.42)
x1   
δ(x′ −x0 )

7. Any realistic measurement of a continuous variable such as position is


subjected to finite resolution. Assuming that a particle has been mea-
sured to be located in the interval (x′ − δ x /2, x′ + δ x /2), where δ x is the
resolution of the measuring device, the collapse postulate implies that
the state of the system undergoes the following transformation

P(x′ −δx /2,x′ +δx /2) |α


|α →  . (3.43)
α| P(x′ −δx /2,x′ +δx /2) |α

8. Some observables have a mixed spectrum containing both a discrete and


continuous subsets. An example of such a mixed spectrum is the set of
eigenvalues of the Hamiltonian operator of a potential well of finite depth.

3.1.2 Momentum Representation

Let p′ be an eigenvalue of the observable p, and let |p′  be the corresponding


eigenvector, namely

p |p′  = p′ |p . (3.44)

Eyal Buks Quantum Mechanics - Lecture Notes 54


3.2. Transformation Function

Note that p′ ∈ R since p is Hermitian. Similarly to the case of the position


observable, the closure relation is written as

dp′ |p′  p′ | = 1 , (3.45)

and the orthonormality relation as

p′′ |p′  = δ (p′ − p′′ ) . (3.46)

The momentum wavefunction φα (p′ ) of a given state |α is defined as

φα (p′ ) = p′ |α . (3.47)

The probability distribution function to measure a momentum value of p = p′


is
2 2
|φα (p′ )| = |p′ |α| . (3.48)

Any ket-vector can be decomposed into momentum eigenstates as


∞ ∞
′ ′ ′
|α = dp |p  p |α = dp′ φα (p′ ) |p′  . (3.49)
−∞ −∞

The inner product between a ket-vector |α and a bra-vector β| can be
expressed as
∞ ∞
β |α = ′ ′
dp β |p  p |α = ′
dp′ φ∗β (p′ ) φα (p′ ) . (3.50)
−∞ −∞

The normalization condition reads


∞
2
1 = α |α = dp′ |φα (p′ )| . (3.51)
−∞

3.2 Transformation Function

What is the relation between the position wavefunction ψα (x′ ) and its mo-
mentum counterpart φα (p′ )?

Claim. The transformation function x′ |p′  is given by


 ′ ′
1 ip x
x′ |p′  = √ exp . (3.52)
2π 

Eyal Buks Quantum Mechanics - Lecture Notes 55


Chapter 3. The Position and Momentum Observables

Proof. On one hand, according to Eq. (3.44)


x′ | p |p′  = p′ x′ |p′  , (3.53)
and on the other hand, according to Eq. (3.29)

x′ | p |p′  = −i x′ |p′  , (3.54)
∂x′
thus

p′ x′ |p′  = −i x′ |p′  . (3.55)
∂x′
The general solution of this differential equation is
 ′ ′
ip x
x′ |p′  = N exp , (3.56)

where N is a normalization constant. To determine the constant N we employ
Eqs. (3.36) and (3.45):
δ (x′ − x′′ )
= x′ |x′′ 

= dp′ x′ |p′  p′ |x′′ 
∞  
′ 2 ip′ (x′ − x′′ )
= dp |N | exp

−∞
∞
′ ′′
=  |N | 2
dkeik(x −x ) .
−∞
  
2πδ(x′ −x′′ )
(3.57)
Thus, by choosing N to be real one finds that
 ′ ′
1 ip x
x′ |p′  = √ exp . (3.58)
2π 
The last result together with Eqs. (3.32) and (3.45) yield
/∞ ′ ip′ x′
∞ dp e  φα (p′ )
−∞
ψα (x′ ) = x′ |α = dp′ x′ |p′  p′ |α = √ , (3.59)
2π
−∞
/∞ ip′ x′
∞ dx′ e−  ψα (x′ )
−∞
φα (p′ ) = p′ |α = ′ ′ ′ ′
dx p |x  x |α = √ . (3.60)
2π
−∞

That is, transformations relating ψα (x′ ) and φα (p′ ) are the direct and inverse
Fourier transformations.

Eyal Buks Quantum Mechanics - Lecture Notes 56


3.3. Generalization for 3D

3.3 Generalization for 3D


According to Eq. (3.6) the observables qx , qy and qz commute with each
other, hence, a simultaneous diagonalization is possible. Denoting the com-
mon eigenvectors as

*
|r′  =
qx′ , qy′ , qz′ , (3.61)

one has

*
qx |r′  = qx′
qx′ , qy′ , qz′ , (3.62)

*
qy |r′  = qy′
qx′ , qy′ , qz′ , (3.63)

*
qz |r′  = qz′
qx′ , qy′ , qz′ . (3.64)
The closure relation is written as
∞ ∞ ∞
1= dx′ dy ′ dz ′ |r′  r′ | , (3.65)
−∞ −∞ −∞

and the orthonormality relation as

r′ |r′′  = δ (r′ − r′′ ) . (3.66)

Similarly, according to Eq. (3.7) the observables px , py and pz commute


with each other, hence, a simultaneous diagonalization is possible. Denoting
the common eigenvectors as

*
|p′  =
p′x , p′y , p′z , (3.67)

one has

*
px |p′  = p′x
p′x , p′y , p′z , (3.68)

*
py |p′  = p′y
p′x , p′y , p′z , (3.69)

*
pz |p′  = p′z
p′x , p′y , p′z . (3.70)
The closure relation is written as
∞ ∞ ∞
1= dp′x dp′y dp′z |p′  p′ | , (3.71)
−∞ −∞ −∞

and the orthonormality relation as

p′ |p′′  = δ (p′ − p′′ ) . (3.72)

The translation operator in three dimensions can be expressed as


 
i∆ · p
J (∆) = exp − , (3.73)


Eyal Buks Quantum Mechanics - Lecture Notes 57


Chapter 3. The Position and Momentum Observables

where ∆ = (∆x , ∆y , ∆z ) ∈ R3 , and where

J (∆) |r′  = |r′ + ∆ . (3.74)

The generalization of Eq. (3.52) for three dimensions is


 ′ ′
′ ′ 1 ip · r
r |p  = 3/2
exp . (3.75)
(2π) 

3.4 Problems

1. Show that
dA
[p, A (x)] = −i , (3.76)
dx
dB
[x, B (p)] = i , (3.77)
dp
where A (x) is a differentiable function of x and B (p) is a differentiable
function of p.
2. Show that the mean value of x in a state described by the wavefunction
ψ (x), namely
+∞

x = dxψ∗ (x) xψ (x) , (3.78)
−∞

is equal to the value of a for which the expression


+∞

F (a) ≡ dxψ∗ (x + a) x2 ψ (x + a) (3.79)
−∞

obtains a minimum, and that this minimum has the value


) *
Fmin = (∆x)2 = x2 − x2 . (3.80)

3. Consider a Gaussian wave packet, whose x space wavefunction is given


by
 
′ 1 ′ x′2
ψα (x ) = √ exp ikx − 2 . (3.81)
π1/4 d 2d

Calculate
' (' (
a) (∆x)2 (∆p)2
b) p′ |α

Eyal Buks Quantum Mechanics - Lecture Notes 58


3.4. Problems

4. Show that the state |α with wave function


+ √
1/ 2a for |x| ≤ a
x′ |α = (3.82)
0 for |x| > a

the uncertainty in momentum is infinity.


5. Show that
∞
d
p = −i dx′ |x′  x′ | . (3.83)
dx′
−∞

6. Show that
  
1 ip′ · (r′ − r′′ )
3 d3 p′ exp = δ (r′ − r′′ ) . (3.84)
(2π) 

7. Find eigenvectors and corresponding eigenvalues of the operator

O = p + Kx , (3.85)

where K is a real constant, p is the momentum operator, which is canon-


ically conjugate to the position operator x. Calculate the wavefunction
of the eigenvectors.
8. Let |α be the state vector of a point particle having mass m that moves
in one dimension along the x axis. The operator pα is defined by the
following requirements: (1) pα is Hermitian (i.e. p†α = pα ) (2) [x, pα ] = 0
(i.e. pα commutes with the position operator x) and (3)

α| (p − pα )2 |α = min α| (p − O)2 |α , (3.86)


O

where p is the momentum operator (i.e. the minimum value of the quan-
tity α| (p − O)2 |α is obtained when the operator O is chosen to be
pα ).
a) Calculate the matrix elements x′ | pα |x′′  of the operator pα in the
position representation.
b) The operator P is the difference between the ’true’ momentum op-
erator and pα

P = p − pα . (3.87)
' (
Calculate the variance (∆P)2 with respect to the state |α
' (
(∆P)2 = α| P 2 |α − α| P |α2 . (3.88)

Eyal Buks Quantum Mechanics - Lecture Notes 59


Chapter 3. The Position and Momentum Observables

c) Use your results to prove the uncertainty relation (3.10)


' (' ( 2
(∆x)2 (∆p)2 ≥ . (3.89)
4
where
' (
(∆x)2 = α| x2 |α − α| x |α2 , (3.90)

and where
' (
(∆p)2 = α| p2 |α − α| p |α2 . (3.91)

3.5 Solutions
1. The commutator [x, p] = i is a constant, thus the relation (2.174) can
be employed
dxm
[p, xm ] = −imxm−1 = −i , (3.92)
dx
m
dp
[x, pm ] = impm−1 = i . (3.93)
dp
This holds for any m, thus, for any differentiable function A (x) of x and
for any differentiable function B (p) of p one has
dA
[p, A (x)] = −i , (3.94)
dx
dB
[x, B (p)] = i . (3.95)
dp
2. The following holds
+∞

F (a) = dxψ∗ (x + a) x2 ψ (x + a)
−∞
+∞

2
= dx′ ψ∗ (x′ ) (x′ − a) ψ (x′)
−∞
' (
= (x − a)2
) *
= x2 − 2a x + a2 .
(3.96)
The requirement
dF
=0 (3.97)
da

Eyal Buks Quantum Mechanics - Lecture Notes 60


3.5. Solutions

leads to −2 x + 2a = 0, or a = x. At that point one has


' ( ) *
2 2
Fmin = (x − x) = x2 − x . (3.98)

3. The following hold


+∞

x = dx′ ψ∗α (x′ ) x′ ψα (x′ )
−∞
+∞
  ′2 
1 x
= 1/2 dx′ exp − 2 x′
π d d
−∞
=0,
(3.99)

+∞

) *
2
x = dx′ ψ∗α (x′ ) x′2 ψα (x′ )
−∞
+∞
  ′2 
1 x
= 1/2 dx exp − 2 x′2

π d d
−∞

d3 π1/2
1
=
π1/2 d 2
d2
= ,
2
(3.100)

+∞

dψα
p = −i dx′ ψ∗α (x′ )
dx′
−∞
+∞
  ′2   
i x x′
= − 1/2 dx′ exp − 2 ik − 2
π d d d
−∞
i
=− ikdπ1/2
π1/2 d
= k ,
(3.101)

Eyal Buks Quantum Mechanics - Lecture Notes 61


Chapter 3. The Position and Momentum Observables

+∞

) 2* d2 ψα
p = (−i)2 dx′ ψ∗α (x′ )
dx′2
−∞
+∞
  ′2  !! 2 ""
2 1 ′ x x′ 1
= (−i) 1/2 dx exp − 2 ik − 2 − 2
π d d d d
−∞
 
1 1 √ 2d4 k2 + d2
= (−i)2 1/2 − d π
π d 2 d4
! "
2 1
= (k) 1 + 2 ,
2 (dk)
(3.102)
a) thus
! ! " "
' (' ( d2 1 2
(∆x)2 (∆p)2 = (k)2
1+ 2
2
− (k) = .
2 2 (dk) 4
(3.103)

b) Using Eq. (3.60) onehas  


1 ip′ x′
p′ |α = √ dx′ exp − ψα (x′ )
2π 
∞   
1 1 ′ ip′ ′ x′2
= √ √ dx exp ik − x − 2
2π π1/4 d  2d
−∞
√ ! "
d (k − p′ )2 d2
= √ exp − .
π 1/4  22
(3.104)
4. The momentum wavefunction is found using Eq. (3.60)
∞  
′ 1 ′ ip′ x′
φα (p ) = √ dx exp − x′ |α
2π 
−∞
a  
1 ′ ip′ x′
= √ dx exp −
4πa 
−a
0 ′
a sin ap
= ′ .
π ap 
(3.105)
The momentum wavefunction φα )(p′ )*is normalizable, however, the inte-
grals for evaluating both p and p2 do not converge.

Eyal Buks Quantum Mechanics - Lecture Notes 62


3.5. Solutions

5. Using Eqs. (3.29) and (3.32) one has


∞
p |α = dx′ |x′  x′ | p |α
−∞
∞
d
= −i dx′ |x′  x′ |α ,
dx′
−∞
(3.106)
thus, since |α is an arbitrary ket vector, the following holds
∞
d
p = −i dx′ |x′  x′ | . (3.107)
dx′
−∞

6. With the help of Eqs. (3.66), (3.71) and (3.75) one finds that
δ (r′ − r′′ ) = r′ |r′′ 

= d3 p′ r′ |p′  p′ |r′′ 
  ′ 
1 3 ′ ip · (r′ − r′′ )
= d p exp .
(2π)3 
(3.108)
7. Using the identity (2.173), which is given by
1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · . (3.109)
2! 3!
and the identity (3.76), which is given by
dg
[g (x) , p] = i , (3.110)
dx
one finds that
 
g(x) −g(x) dg i dg i dg
e pe = p + i + g (x) , + g (x) , g (x) , +· · · .
dx 2! dx 3! dx
(3.111)

Choosing g (x) to be given by

Kx2
g (x) = (3.112)
2i
yields

UpU † = p + Kx = O , (3.113)

Eyal Buks Quantum Mechanics - Lecture Notes 63


Chapter 3. The Position and Momentum Observables

where the unitary operator U is given by


iKx2
U = e− 2 .

Thus, the vectors |ψ (p′ ), which are define as

|ψ (p′ ) = U |p′  , (3.114)

where |p′  is an eigenvector of p with eigenvalue p′ (i.e. p |p′  = p′ |p′ ),


are eigenvectors of O, and the following holds

O |ψ (p′ ) = p′ |ψ (p′ ) . (3.115)

With the help of Eq. (3.52), which is given by


1 ip′ x′
x′ |p′  = √ e  , (3.116)
2π
one finds that the wavefunction ψ (x′ ; p′ ) = x′ |ψ (p′ ) of the state
|ψ (p′ ) is given by
iKx′2
ψ (x′ ; p′ ) = e− 2 x′ |p′ 
 
1 i ′2
p′ x′ − Kx
= √ e 2
.
2π
(3.117)
8. With the help of Eq. (3.32) one finds that
∞ ∞
pα = dx′ dx′′ |x′  x′ | pα |x′′  x′′ | . (3.118)
−∞ −∞

The requirement [x, pα ] = 0 implies that


∞ ∞

dx dx′′ |x′  x′ | pα |x′′  (x′ − x′′ ) x′′ | = 0 , (3.119)
−∞ −∞

hence x′ | pα |x′′  = 0 unless x′ = x′′ . Thus by using the notation

x′ | pα |x′′  = φα (x′ ) δ (x′ − x′′ ) , (3.120)

the operator pα can be expressed as


∞
pα = dx′ |x′  φα (x′ ) x′ | . (3.121)
−∞

The requirement that pα is Hermitian implies that φα (x′ ) is real.

Eyal Buks Quantum Mechanics - Lecture Notes 64


3.5. Solutions

a) With the help of Eq. (3.83), which is given by


∞
d
p = −i dx′ |x′  x′ | , (3.122)
dx′
−∞

one finds that


∞  
d
p − pα = dx′ |x′  −i ′ − φα (x′ ) x′ | , (3.123)
dx
−∞

hence in terms of the wavefunction ψα (x′ ) = x′ |α of |α one has
∞  
d
(p − pα ) |α = dx |x  −i ′ − φα (x ) ψα (x′ )
′ ′ ′
dx
−∞
∞  
′ ′ d logψα
′ ′
= dx |x  ψα (x ) −i − φα (x ) .
dx′
−∞
(3.124)
Similarly
∞  
d logψ∗α
α| (p − pα ) = dx ′
ψ∗α (x′ ) i ∗
− φα (x ) x′ | , (3.125)

dx′
−∞

and thus
α| (p − pα )2 |α
∞

2


d logψ α

= dx ρ (x )
i ′
+ φα (x )
.
dx
−∞
(3.126)
where
2
ρ (x′ ) = |ψα (x′ )| . (3.127)

The minimum value is obtained when (recall that φα (x′ ) is required


to be real)  
 d logψα d logψ∗α
φα (x′ ) = −
2i dx′ dx′
ψα
 d log ψ∗α
= ,
2i dx′
(3.128)
and thus

Eyal Buks Quantum Mechanics - Lecture Notes 65


Chapter 3. The Position and Momentum Observables

ψα
 d log ψ∗α
x′ | pα |x′′  = δ (x′ − x′′ ) . (3.129)
2i dx′
Note: Comparing this result with the expression for the current den-
sity J associated with the state |α [see Eq. (4.200)] yields the fol-
lowing relation 
 ∗ dψ α
J= Im ψα ′
m dx
 
ρ (x )  d logψα d logψ∗α

= −
m 2i dx′ dx′

ρ (x )
= φ (x′ ) .
m α
(3.130)
b) As can be seen from Eqs. (3.123) and (3.128) the following holds
 
∞ d log ψ α
d 1 ψ ∗
P = i dx′ |x′  − ′ + α 
x′ | , (3.131)
dx 2 dx′
−∞

hence
∞  
dψ 1 dψ dψ∗α
α| P |α = i dx′ −ψ∗α α′ + ψ∗α α′ − ψ
dx 2 dx dx′ α
−∞
∞  
i dψ dψ∗α
=− dx′ ψ∗α α′ + ψ
2 dx dx′ α
−∞
∞
i dρ (x′ )
=− dx′
2 dx′
−∞
=0,
(3.132)
thus'[see Eqs.
( (3.126) and (3.128)]
2
(∆P) = α| P 2 |α
 2 ∞

2
 ′


d logψ α d logψ∗α

= dx ρ (x )
+
2 dx′ dx′

−∞
 2 ∞  2
 ′ ′ d logρ (x′ )
= dx ρ (x ) .
2 dx′
−∞
(3.133)
Note that the result α| P |α = 0 implies that pα and p have the
same expectation value, i.e. α| pα |α = α| p |α. On the other hand,

Eyal Buks Quantum Mechanics - Lecture Notes 66


3.5. Solutions

contrary to p, the operator pα commutes with the position operator


x.
c) '
Using the
( relation
' α|
( pα |α = α| p |α one finds that
2 2
(∆p) − (∆pα ) = α| p2 |α − α| p2α |α
 
= α| (p − pα )2 |α + α| ppα + pα p − 2p2α |α
 
= α| P 2 |α + α| ppα + pα p − 2p2α |α .
(3.134)
As can be see from Eq. (3.128), the following holds
∞  
d logψα
pα =  dx′ |x′  Im x′ | , (3.135)
dx′
−∞

thus  
α| ppα + pα p − 2p2α |α
 α|p x′ 
∞ |  x′ |p|α    

ψα + ψα d logψ d logψα
2 ′  α 
= dx ρ − 2 Im Im
 dx′ dx′
−∞
=0,
(3.136)
and 'therefore
( ' (
(∆p)2 = α| P 2 |α + (∆pα )2
≥ α| P 2 |α
 2 ∞  2
 ′ ′ d logρ (x′ )
= dx ρ (x ) .
2 dx′
−∞
(3.137)
For general real functions f (x′ ) , g (x′ ) : R → R the Schwartz in-
equality (2.166) implies that


2


∞ ∞

′ ′ ′

′ ′ 2 2


dx f (x ) g (x )

≤ dx (f (x )) dx′ (g (x′ )) . (3.138)



−∞

−∞ −∞

Implementingthis inequality for the functions


f (x′ ) = ρ (x′ ) (x′ − x) , (3.139)
 d logρ (x′ )
g (x′ ) = ρ (x′ ) , (3.140)
dx′
where
∞
x = dx′ ρ (x′ ) x′
−∞

Eyal Buks Quantum Mechanics - Lecture Notes 67


Chapter 3. The Position and Momentum Observables

is the expectation value of x, yields




2

/∞ ′


d logρ (x )

∞  2

dx′ ρ (x′ ) (x′ − x) dx′


d logρ (x ) −∞

dx′ ρ (x′ ) ≥ ' ( ,


dx′ (∆x) 2
−∞
(3.141)

where
' ( ∞
2 2
(∆x) = dx′ ρ (x′ ) (x′ − x) (3.142)
−∞

is the variance of x. By integrating by parts one finds that


∞ ∞
′ ′ ′ d logρ (x′ ) dρ (x′ )
dx ρ (x ) (x − x) ′
= dx′ (x′ − x)
dx dx′
−∞ −∞
∞
=− dx′ ρ (x′ )
−∞
= −1 .
(3.143)
Combining these results [see Eqs. (3.137) and (3.141)] lead to
' (' (   2
(∆x)2 (∆p)2 ≥ . (3.144)
2

Eyal Buks Quantum Mechanics - Lecture Notes 68


4. Quantum Dynamics

The time evolution of a state vector |α is postulated to be given by the


Schrödinger equation
d |α
i = H |α , (4.1)
dt
where the Hermitian operator H = H† is the Hamiltonian of the system.
The Hamiltonian operator is the observable corresponding to the classical
Hamiltonian function that we have studied in chapter 1. The time evolution
produced by Eq. (4.1) is unitary, as is shown below:
Claim. The norm α |α is time independent.
Proof. Since H = H† , the dual of the Schrödinger equation (4.1) is given by
d α|
−i = α| H . (4.2)
dt
Using this one has
 
d α |α d α| d |α 1
= |α+α| = (− α| H |α + α| H |α) = 0 . (4.3)
dt dt dt i

4.1 Time Evolution Operator


The time evolution operator u (t, t0 ) relates the state vector at time |α (t0 )
with its value |α (t) at time t:

|α (t) = u (t, t0 ) |α (t0 ) . (4.4)

Claim. The time evolution operator satisfies the Schrödinger equation (4.1).
Proof. Expressing the Schrödinger equation (4.1) in terms of Eq. (4.4)
d
i u (t, t0 ) |α (t0 ) = Hu (t, t0 ) |α (t0 ) , (4.5)
dt
and noting that |α (t0 ) is t independent yield
Chapter 4. Quantum Dynamics
 
d
i u (t, t0 ) |α (t0 ) = Hu (t, t0 ) |α (t0 ) . (4.6)
dt

Since this holds for any |α (t0 ) one concludes that

du (t, t0 )
i = Hu (t, t0 ) . (4.7)
dt
This results leads to the following conclusion:

Claim. The time evolution operator is unitary.

Proof. Using Eq. (4.7) one finds that


 
d u† u du du†
= u† + u
dt dt dt
1  † 
= u Hu − u† Hu
i
=0.
(4.8)

Furthermore, for t = t0 clearly u (t0 , t0 ) = u (t0 , t0 ) = 1. Thus, one concludes
that u† u = 1 for any time, namely u is unitary.

4.2 Time Independent Hamiltonian

A special case of interest is when the Hamiltonian is time independent. In


this case the solution of Eq. (4.7) is given by
 
iH (t − t0 )
u (t, t0 ) = exp − . (4.9)


The operator u (t, t0 ) takes a relatively simple form in the basis of eigenvectors
of the Hamiltonian H. Denoting these eigenvectors as |an,i , where the index
i is added to account for possible degeneracy, and denoting the corresponding
eigenenergies as En one has

H |an,i  = En |an,i  , (4.10)

where

an′ ,i′ |an,i  = δ nn′ δ ii′ . (4.11)

By using the closure relation, which is given by


gn

1= |an,i  an,i | , (4.12)
n i=1

Eyal Buks Quantum Mechanics - Lecture Notes 70


4.3. Example - Spin 1/2

and Eq. (4.9) one finds that


 
iH (t − t0 )
u (t, t0 ) = exp − 1

 gn  
iH (t − t0 )
= exp − |an,i  an,i |
n i=1

 gn  
iEn (t − t0 )
= exp − |an,i  an,i | .
n i=1

(4.13)
Using this results the state vector |α (t) can be written as
|α (t) = u (t, t0 ) |α (t0 )
 gn  
iEn (t − t0 )
= exp − an,i |α (t0 ) |an,i  .
n i=1

(4.14)
Note that if the system is initially in an eigenvector of the Hamiltonian
with eigenenergy En , then according to Eq. (4.14)
 
iEn (t − t0 )
|α (t) = exp − |α (t0 ) . (4.15)


However, the phase factor multiplying |α (t0 ) has no effect on any mea-
surable physical quantity of the system, that is, the system’s properties are
time independent. This is why the eigenvectors of the Hamiltonian are called
stationary states.

4.3 Example - Spin 1/2

In classical mechanics, the potential energy U of a magnetic moment µ in a


magnetic field B is given by

U = −µ · B . (4.16)

The magnetic moment of a spin 1/2 is given by [see Eq. (2.90)]


2µB
µspin = S, (4.17)

where S is the spin angular momentum vector and where
e
µB = (4.18)
2me c

Eyal Buks Quantum Mechanics - Lecture Notes 71


Chapter 4. Quantum Dynamics

is the Bohr’s magneton (note that the electron charge is taken to be negative
e < 0). Based on these relations we hypothesize that the Hamiltonian of a
spin 1/2 in a magnetic field B is given by
e
H=− S·B. (4.19)
me c
Assume the case where
B = Bẑ , (4.20)
where B is a constant. For this case the Hamiltonian is given by
H = ωSz , (4.21)
where
|e| B
ω= (4.22)
me c
is the so-called Larmor frequency. In terms of the eigenvectors of the operator
Sz

Sz |± = ± |± , (4.23)
2
where the compact notation |± stands for the states |±; ẑ, one has

H |± = ± |± , (4.24)
2
namely the states |± are eigenstates of the Hamiltonian. Equation (4.13) for
the present case reads
iωt iωt
u (t, 0) = e− 2 |+ +| + e 2 |− −| . (4.25)
Exercise 4.3.1. Consider spin 1/2 in magnetic field given by B = Bẑ, where
B is a constant. Given that |α (0) = |+; x̂ at time t = 0 calculate (a) the
probability p± (t) to measure Sx = ±/2 at time t; (b) the expectation value
Sx  (t) at time t.
Solution 4.3.1. Recall that [see Eq. (2.102)]
1
|±; x̂ = √ (|+ ± |−) (4.26)
2
(a) Using Eq. (4.25) one finds
p± (t) = |±; x̂| u (t, 0) |α (0)|2

 iωt 
2

1 − 2 iωt

=
(+| ± −|) e |+ +| + e |− −| (|+ + |−)

2
2

 

2

1 iωt iωt
=

e− 2 ± e 2

,
2
(4.27)

Eyal Buks Quantum Mechanics - Lecture Notes 72


4.4. Connection to Classical Dynamics

thus

ωt
p+ (t) = cos2 , (4.28)
2
 
ωt
p− (t) = sin2 . (4.29)
2
(b)Using the results for p+ and p− one has

Sx  = (p+ − p− )
2    
 2 ωt 2 ωt
= cos − sin
2 2 2

= cos (ωt) .
2
(4.30)

4.4 Connection to Classical Dynamics


In chapter 1 we have found that in classical physics, the dynamics of a variable
A(c) is governed by Eq. (1.38), which is given by

dA(c) 1 (c) (c) 2 ∂A(c)


= A ,H + . (4.31)
dt ∂t
We seek a quantum analogy to this equation. To that end, we derive an
equation of motion for the expectation value A of the observable A that
corresponds to the classical variable A(c) . In general, the expectation value
can be expressed as

A = α (t)| A |α (t) = α (t0 )| u† (t, t0 ) Au (t, t0 ) |α (t0 ) = α (t0 )| A(H) |α (t0 ) ,
(4.32)

where u is the time evolution operator and

A(H) = u† (t, t0 ) Au (t, t0 ) . (4.33)

The operator A(H) is called the Heisenberg representation of A. We first


derive an equation of motion for the operator A(H) . By using Eq. (4.7) one
finds that the following holds
du 1
= Hu , (4.34)
dt i
du† 1
= − u† H , (4.35)
dt i
therefore

Eyal Buks Quantum Mechanics - Lecture Notes 73


Chapter 4. Quantum Dynamics

dA(H) du† du ∂A
= Au + u† A + u† u
dt dt dt ∂t
1  †  ∂A
= −u HAu + u† AHu + u† u
i ∂t
1  †  ∂A
= −u Huu† Au + u† Auu† Hu + u† u
i ∂t
1   ∂A(H)
= −H(H) A(H) + A(H) H(H) + .
i ∂t
(4.36)
Thus, we have found that

dA(H) 1 3 (H) (H) 4 ∂A(H)


= A ,H + . (4.37)
dt i ∂t
Furthermore, the desired equation of motion for A is found using Eqs. (4.32)
and (4.37)
5 6
d A 1 ∂A
= [A, H] + . (4.38)
dt i ∂t

We see that the Poisson’s brackets in the classical equation of motion (4.31)
for the classical variable A(c) are replaced by a commutation relation in the
quantum counterpart equation of motion (4.38) for the expectation value A
1
{, } → [, ] . (4.39)
i
Note that for the case where the Hamiltonian is time independent, namely
for the case where the time evolution operator is given by Eq. (4.9), u com-
mutes with H, namely [u, H] = 0, and consequently

H(H) = u† Hu = H . (4.40)

4.5 Symmetric Ordering

What is in general the correspondence between a classical variable and its


quantum operator counterpart? Consider for example the system of a point
particle moving in one dimension. Let x(c) be the classical coordinate and let
p(c) be the canonically conjugate momentum. As we have done in chapter 3,
the quantum observables corresponding to x(c) and p(c) are the Hermitian
operators x and p. The commutation
 relation [x, p] is derived from the cor-
responding Poisson’s brackets x(c) , p(c) according to the rule

1
{, } → [, ] , (4.41)
i

Eyal Buks Quantum Mechanics - Lecture Notes 74


4.5. Symmetric Ordering

namely
1 2
x(c) , p(c) = 1 → [x, p] = i . (4.42)

However,
 what is the quantum operator corresponding to a general func-
tion A x(c) , p(c) of x(c) and p(c) ? This question raises the issue of order-
 
ing. As an example, let A x(c) , p(c) = x(c) p(c) . Classical variables obviously
commute, therefore x(c) p(c) = p(c) x(c) . However, this is not true for quantum
operators xp = px. Moreover, it is clear that both operators xp and px cannot
be considered as observables since they are not Hermitian

(xp) = px = xp , (4.43)

(px) = xp = px . (4.44)
A better candidate to serve as the quantum operator corresponding to the
classical variables x(c) p(c) is the operator (xp + px) /2, which is obtained from
x(c) p(c) by a procedure called symmetric ordering. A general transformation
that produces a symmetric ordered observable
 A (x, p) that corresponds to
a given general function A x(c) , p(c) of the classical variable x(c) and its
canonical conjugate p(c) is given below
∞ ∞  
A (x, p) = A x(c) , p(c) Υ dx(c) dp(c) ,
−∞ −∞
(4.45)
where
∞ ∞
1 (c)
−x)+η(p(c) −p))
e  (ξ(x
i
Υ = 2 dξdη . (4.46)
(2π)
−∞ −∞

This transformation is called the Weyl transformation. The identity


∞
′ ′′
dkeik(x −x ) = 2πδ (x′ − x′′ ) , (4.47)
−∞

implies that
  
1
e  ξ(x −x) dξ = δ x(c) − x ,
i (c)
(4.48)
2π
  
1
e  η(p −p) dη = δ p(c) − p .
i (c)
(4.49)
2π
At first glance these relations
 may
 leadto the (wrong) conclusion that the
term Υ equals to δ x(c) − x δ p(c) − p , however, this is incorrect since x
and p are non-commuting operators.

Eyal Buks Quantum Mechanics - Lecture Notes 75


Chapter 4. Quantum Dynamics

4.6 Problems
1. Consider spin 1/2 in magnetic field given by B = Bẑ, where B is a
constant. At time t = 0 the system is in the state |+; x̂. Calculate Sx ,
Sy  and Sz  as a function of time t.
2. Consider a point particle having mass m moving in one dimension under
the influence of the potential V (x). Let |ψn  be a normalized eigenvector
of the Hamiltonian of the system with eigenvalue En . Show that the cor-
responding wavefunction ψn (x′ ) in the coordinate representation satisfies
the following equation
2 d2 ψn (x′ )
− + V (x′ ) ψ n (x′ ) = En ψn (x′ ) . (4.50)
2m dx′2
3. Consider the Hamiltonian operator
p2
H= + V (r) , (4.51)
2m
where r = (x, y, z) is the vector of position operators, p = (px , py , pz )
is the vector of canonical conjugate operators, and the mass m is a con-
stant. Let |ψn  be a normalizable eigenvector of the Hamiltonian H with
eigenvalue En . Show that
ψn | p |ψn  = 0 . (4.52)
4. Show that in the p representation the Schrödinger equation
d |α
i = H |α , (4.53)
dt
where H is the Hamiltonian
p2
H= + V (r) , (4.54)
2m
can be transformed into the integro-differential equation

d p2
i φα = φα + dp′ U (p − p′ ) φα , (4.55)
dt 2m
where φα = φα (p′ , t) = p′ |α is the momentum wave function and
where
  
i
U (p) = (2π)−3 dr V (r) exp − p · r . (4.56)

5. Consider a particle of mass m in a scalar potential energy V (r). Prove
Ehrenfest’s theorem
d2
m r = − ∇V (r) . (4.57)
dt2

Eyal Buks Quantum Mechanics - Lecture Notes 76


4.6. Problems

6. Show that if the potential energy V (r) can be written as a sum of func-
tions of a single coordinate, V (r) = V1 (x1 ) + V2 (x2 ) + V3 (x3 ), then the
time-independent Schrödinger equation can be decomposed into a set of
one-dimensional equations of the form

d2 ψi (xi ) 2m
+ 2 [Ei − Vi (xi )] ψi (xi ) = 0 , (4.58)
dx2i 

where i ∈ {1, 2, 3}, with ψ (r) = ψ1 (x1 ) ψ2 (x2 ) ψ3 (x3 ) and E = E1 +


E2 + E3 .
7. Show that, in one-dimensional problems, the energy spectrum of the
bound states is always non-degenerate.
8. Let ψn (x) (n = 1, 2, 3, · · · ) be the eigen-wave-functions of a one-
dimensional Schrödinger equation with eigen-energies En placed in order
of increasing magnitude (E1 < E2 < · · · . ). Show that between any two
consecutive zeros of ψn (x), ψn+1 (x) has at least one zero.
9. What conclusions can be drawn about the parity of the eigen-functions
of the one-dimensional Schrödinger equation

d2 ψ (x) 2m
+ 2 (E − V (x)) ψ (x) = 0 (4.59)
dx2 
if the potential energy is an even function of x , namely V (x) = V (−x).
10. Show that the first derivative of the time-independent wavefunction is
continuous even at points where V (x) has a finite discontinuity.
11. A particle having mass m is confined by a one dimensional potential given
by
+
−W if |x| ≤ a
Vs (x) = , (4.60)
0 if |x| > a

where a > 0 and W > 0 are real constants. Show that the particle has
at least one bound state (i.e., a state having energy E < 0).
12. Consider a particle having mass m confined in a potential well given by
+
0 if 0 ≤ x ≤ a
V (x) = . (4.61)
∞ if x < 0 or x > a

The eigenenergies are denoted by En and the corresponding eigen states


are denoted by |ϕn  , where n = 1, 2, · · · (as usual, the states are num-
bered in increasing order with respect to energy). The state of the system
at time t = 0 is given by

|Ψ (0) = a1 |ϕ1  + a2 |ϕ2  + a3 |ϕ3  . (4.62)

(a) The energy E of the system is measured at time  t =0 . What is the
probability to measure a value smaller than 3π 2 2 / ma2 ? (b) Calculate

Eyal Buks Quantum Mechanics - Lecture Notes 77


Chapter 4. Quantum Dynamics
.
the standard deviation ∆E = E 2  − E2 at time t = 0 . (c) the same
as (b), however for any time t > 0 . (d) The energy was measured at
time t and the value of 2π2 2 / ma2 was found. The energy is measured
again at later time t0 > t . Calculate E and ∆E at time t0 .
13. Consider a point particle having mass m in a one dimensional potential
given by
V (x) = −αδ (x) , (4.63)
where δ (x) is the delta function. The value of the parameter α suddenly
changes from α1 at times t < 0 to the value α2 at times t > 0. Both
α1 and α2 are positive real numbers. Given that the particle was in the
ground state at times t < 0, what is the probability p that the particle
will remain bounded at t > 0?
14. Consider a point particle having mass m in a one dimensional potential
given by

V (x) = −αδ (x) , (4.64)


where δ (x) is the delta function, and where α > 0. Let |γ 0  be the
ground state and let E0 be the energy of the ground state. The particle
is prepared in the state
 
ip0 x
|g (p0 ) = exp |γ 0  , (4.65)

where p0 is real and where x is the position operator. Calculate the
probability s (p0 ) that a measurement of energy will yield the result E0 .
15. Consider a point particle having mass m in a one dimensional potential
given by
+
−αδ (x) |x| < a
V (x) = , (4.66)
∞ |x| ≥ a

where δ (x) is the delta function and α is a constant. Let E0 be the energy
of the ground state. Under what conditions E0 < 0?
16. The same as the previous exercise, however for the potential
+
∞ x<0
V (x) = , (4.67)
−αδ (x − x0 ) x ≥ 0
where α is real and x0 is positive.
17. Thomas-Reiche-Kuhn sum rule - Let
p2
H= + V (r) (4.68)
2m
be the Hamiltonian of a particle of mass m moving in a potential V (r).
Show that

Eyal Buks Quantum Mechanics - Lecture Notes 78


4.6. Problems

 2 2
(Ek − El ) |k| x |l| = , (4.69)
2m
k

where the sum is taken over all energy eigen-states of the particle (where
H |k = Ek |k ), and x is the x component of the position vector operator
r (the Thomas-Reiche-Kuhn sum rule).
18. A particle having mass m is confined in a one dimensional potential well
given by
+
0 0<x<a
V (x) = .
∞ else

a) At time t = 0 the position was measured and the result was x = a/2.
The resolution of the position measurement is ∆x , where ∆x << a.
After time τ 1 the energy was measured. Calculate the probability pn
to measure that the energy of the system is En , where En are the
eigenenergies of the particle in the well, and where n = 1, 2, · · · .
b) Assume that the result of the measurement in the previous section
was E2 . At a later time τ 2 > τ 1 the momentum p of the particle
was measured. Calculate the expectation value p.
19. A particle having mass m is in the ground state of an infinite potential
well of width a, which is given by
+
0 0<x<a
V1 (x) = . (4.70)
∞ else

At time t = 0 the potential suddenly changes and becomes


+
0 0 < x < 2a
V2 (x) = , (4.71)
∞ else

namely the width suddenly becomes 2a. (a) Find the probability p to
find the particle in the ground state of the new well. (b) Calculate the
expectation value of the energy H before and after the change in the
potential. ' (
20. Calculate the uncertainties in position (∆x)2 and in momentum
' (
(∆p)2 of the energy eigenstates of a particle having mass m, which
is confined in a one dimensional potential well given by
+
0 0<x<a
V (x) = . (4.72)
∞ else

21. The continuity equation - Consider a point particle having mass m


and charge q placed in an electromagnetic field. Show that

+ ∇J = 0 , (4.73)
dt

Eyal Buks Quantum Mechanics - Lecture Notes 79


Chapter 4. Quantum Dynamics

where

ρ = ψψ∗ (4.74)

is the probability distribution function, ψ (x′ ) is the wavefunction,

 qρ
J= Im (ψ∗ ∇ψ) − A (4.75)
m mc
is the current density, and A is the electromagnetic vector potential.
22. A particle having mass m moves in one dimension under the influence of
the potential V (x′ ). In the range |x′ | > a the potential V (x′ ) vanishes.
Consider a solution to the time independent Schrödinger equation, whose
wavefunction ψ (x′ ) in the range |x′ | > a is taken to be given by
+ ′ ′
A1 eikx + B1 e−ikx x′ < −a
ψ (x′ ) = ′ ′ , (4.76)
A2 eikx + B2 e−ikx x′ > a

where A1 , B1 , A2 , B2 and k are all constants. Find a relation that the


constants A1 , B1 , A2 and B2 must satisfy.
23. A particle having mass m moves in one dimension under the influence of
a rectangular potential barrier given by
+
′ Ub |x′ | ≤ a2
V (x ) = . (4.77)
0 |x′ | > a2

Consider a solution to the time independent Schrödinger equation, whose


wavefunction in the range |x| > a/2 has the form
+ ikx′ ′
′ e + re−ikx x′ < − a2
ψ (x ) = ′ , (4.78)
teikx x′ > a2

where k is a constant. Calculate the transmission and reflection coeffi-


cients t and r respectively.  
24. Calculate the Weyl transformation A (x, p) of the classical variable A x(c) , p(c) =
p(c) x(c) .  
25. Invert Eq. (4.45), i.e. express the variable A x(c) , p(c) as a function of
the operator A (x, p).

4.7 Solutions

1. The operators Sx , Sy and Sz are given by Eqs. (2.102), (2.103) and (2.99)
respectively. The Hamiltonian is given by Eq. (4.21). Using Eqs. (4.38)
and (2.131) one has

Eyal Buks Quantum Mechanics - Lecture Notes 80


4.7. Solutions

d Sx  ω
= [Sx , Sz ] = −ω Sy  , (4.79)
dt i
d Sy  ω
= [Sy , Sz ] = ω Sx  , (4.80)
dt i
d Sz  ω
= [Sz , Sz ] = 0 , (4.81)
dt i
where
|e| B
ω= . (4.82)
me c
At time t = 0 the system is in state
1
|+; x̂ = √ (|+ + |−) , (4.83)
2
thus
 
Sx  (t = 0) = (+| + −|) (|+ −| + |− +|) (|+ + |−) = .
4 2

Sy  (t = 0) = (+| + −|) (−i |+ −| + i |− +|) (|+ + |−) = 0 .
4

Sz  (t = 0) = (+| + −|) (|+ +| − |− −|) (|+ + |−) = 0 .
4
The solution is easily found to be given by
 

Sx  (t) = cos (ωt) , (4.84)
2
 

Sy  (t) = sin (ωt) , (4.85)
2
Sz  (t) = 0 . (4.86)
2. The Hamiltonian operator H is given by
p2
H= + V (x) . (4.87)
2m
Multiplying the relation
H |ψn  = En |ψn  (4.88)
from the left by x′ | yields [see Eqs. (3.23) and (3.29)]
2 d2 ψn (x′ )
− + V (x′ ) ψ n (x′ ) = En ψn (x′ ) , (4.89)
2m dx′2
where
ψn (x′ ) = x′ |ψn  (4.90)
is the wavefunction in the coordinate representation.

Eyal Buks Quantum Mechanics - Lecture Notes 81


Chapter 4. Quantum Dynamics

3. Using [x, px ] = [y, py ] = [z, pz ] = i one finds that


2 
p
[H, r] = ,r
2m
1 , 2 - , 2 - , 2 -
= px , x , py , y , pz , z
2m

= (px , py , pz )
im

= p.
im
(4.91)
Thus
im
ψn | p |ψn  = ψn | [H, r] |ψn 

im
= ψn | (Hr − rH) |ψn 

imEn
= ψn | (r − r) |ψn 

=0.
(4.92)
4. Multiplying Eq. (4.53) from the left by the bra p′ | and inserting the
closure relation

1 = dp′′ |p′′  p′′ | (4.93)

yields

dφα (p′ )
i = dp′′ p′ | H |p′′  φα (p′′ ) . (4.94)
dt
The following hold

p′ | p2 |p′′  = p′2 δ (p′ − p′′ ) , (4.95)

and
 
′ ′′ ′
p | V (r) |p  = dr dr′′ p′ |r′  r′ | V (r) |r′′  r′′ |p′′ 
     ′′ ′′ 
−3 ip′ · r′ ip · r
= (2π) dr′ dr′′ exp − V (r′ ) δ (r′ − r′′ ) exp
 
  ′ ′′ ′

i (p − p ) · r
= (2π)−3 dr′ exp − V (r′ )

= U (p′ − p′′ ) ,
(4.96)

Eyal Buks Quantum Mechanics - Lecture Notes 82


4.7. Solutions

thus the momentum wave functionφα (p′ ) satisfies the following equation

dφα p′2
i = φ + dp′′ U (p′ − p′′ ) φα . (4.97)
dt 2m α
5. The Hamiltonian is given by
p2
H= + V (r) . (4.98)
2m
Using Eq. (4.38) one has
d x 1 1 ), -* px 
= [x, H] = x, p2x = , (4.99)
dt i i2m m
and
d px  1
= [px , V (r)] , (4.100)
dt i
or with the help of Eq. (3.76)
5 6
d px  ∂V
=− . (4.101)
dt ∂x
This together with Eq. (4.99) yield
5 6
d2 x ∂V
m =− . (4.102)
dt2 ∂x
Similar equations are obtained for y and z, which together yield Eq.
(4.57).
6. Substituting a solution having the form
ψ (r) = ψ1 (x1 ) ψ2 (x2 ) ψ3 (x3 ) (4.103)
into the time-independent Schrödinger equation, which is given by
2m
∇2 ψ (r) + [E − V (r)] ψ (r) = 0 , (4.104)
2
and dividing by ψ (r) yield
3 
 
1 d2 ψi (xi ) 2m 2m
2 − 2 Vi (xi ) = − 2 E . (4.105)
i=1
ψi (xi ) dxi  

In the sum, the ith term (i ∈ {1, 2, 3}) depends only on xi , thus each
term must be a constant
1 d2 ψi (xi ) 2m 2m
2 − 2 Vi (xi ) = − 2 Ei , (4.106)
ψ i (xi ) dxi  
where E1 + E2 + E3 = E.

Eyal Buks Quantum Mechanics - Lecture Notes 83


Chapter 4. Quantum Dynamics

7. Consider two eigen-wave-functions ψ1 (x) and ψ2 (x) having the same


eigenenergy E. The following holds
d2 ψ1 2m
+ 2 (E − V (x)) ψ 1 = 0 , (4.107)
dx2 
d2 ψ2 2m
+ 2 (E − V (x)) ψ 2 = 0 , (4.108)
dx2 
thus
1 d2 ψ1 1 d2 ψ2
= , (4.109)
ψ 1 dx2 ψ2 dx2
or
 
d2 ψ1 d2 ψ2 d dψ1 dψ 2
ψ2 2
− ψ 1 2
= ψ2 − ψ1 =0, (4.110)
dx dx dx dx dx
therefore
dψ1 dψ
ψ2 − ψ1 2 = C , (4.111)
dx dx
where C is a constant. However, for bound states

lim ψ (x) = 0 , (4.112)


x→±∞

thus C = 0, and consequently


1 dψ1 1 dψ2
= . (4.113)
ψ 1 dx ψ2 dx
Integrating the above equation yields

log ψ1 = log ψ2 + α , (4.114)

where α is a constant. Therefore

ψ1 = eα ψ 2 , (4.115)

and therefore ψ2 is just proportional to ψ1 (both represent the same


physical state).
8. Consider two eigen-wave-functions ψn (x) and ψn+1 (x) with En < En+1 .
As we saw in the previous exercise, the spectrum is non-degenerate. More-
over, the Schrödinger equation

d2 ψ 2m
+ 2 (E − V (x)) ψ = 0 , (4.116)
dx2 
which the eigen-wave-functions satisfy, is real. Therefore given that ψ (x)
is a solution with a given eigenenergy E, then also ψ ∗ (x) is a solution

Eyal Buks Quantum Mechanics - Lecture Notes 84


4.7. Solutions

with the same E. Therefore, all eigen-wave-functions can be chosen to be


real (i.e., by the transformation ψ (x) → (ψ (x) + ψ∗ (x)) /2). We have
d2 ψn 2m
+ 2 (En − V (x)) ψn = 0 , (4.117)
dx2 
d2 ψn+1 2m
+ 2 (En+1 − V (x)) ψn+1 = 0 . (4.118)
dx2 
By multiplying the first Eq. by ψn+1 , the second one by ψn , and sub-
tracting one has

d2 ψn d2 ψn+1 2m
ψn+1 − ψ n + 2 (En − En+1 ) ψn ψ n+1 = 0 , (4.119)
dx2 dx2 
or
 
d dψ dψ 2m
ψn+1 n − ψn n+1 + [En − En+1 ] ψn ψn+1 = 0 . (4.120)
dx dx dx 2

Let x1 and x2 be two consecutive zeros of ψn (x) (i.e., ψn (x1 ) =


ψn (x2 ) = 0). Integrating from x1 to x2 yields

x2

 x2
dψ dψ

n+1
2m
ψn+1 n − ψn = (E n+1 − E n ) dxψ n ψn+1 .
dx  dx

2    x1
=0 x1 >0
(4.121)

Without lost of generality, assume that ψn (x) > 0 in the range (x1 , x2 ).
Since ψn (x) is expected to be continuous, the following must hold

dψn

>0, (4.122)
dx
x=x1

dψn

<0. (4.123)
dx
x=x2
As can be clearly seen from Eq. (4.121), the assumption that ψn+1 (x) > 0
in the entire range (x1 , x2 ) leads to contradiction. Similarly, the possibil-
ity that ψn+1 (x) < 0 in the entire range (x1 , x2 ) is excluded. Therefore,
ψn+1 must have at least one zero in this range.
9. Clearly if ψ (x) is an eigen function with energy E, also ψ (−x) is an
eigen function with the same energy. Consider two cases: (i) The level E
is non-degenerate. In this case ψ (x) = cψ (−x), where c is a constant.
Normalization requires that |c|2 = 1. Moreover, since the wavefunctions
can be chosen to be real, the following holds: ψ (x) = ±ψ (−x). (ii) The
level E is degenerate. In this case every superposition of ψ (x) and ψ (−x)
can be written as a superposition of an odd eigen function ψodd (x) and
an even one ψ even (x), which are defined by

Eyal Buks Quantum Mechanics - Lecture Notes 85


Chapter 4. Quantum Dynamics

ψodd (x) = ψ (x) − ψ (−x) , (4.124)


ψeven (x) = ψ (x) + ψ (−x) . (4.125)
10. The time-independent Schrödinger equation reads

d2 ψ (x) 2m
+ 2 (E − V (x)) ψ (x) = 0 . (4.126)
dx2 
Assume V (x) has a finite discontinuity at x = x0 . Integrating the
Schrödinger equation in the interval (x0 − ε, x0 + ε) yields


x +ε x0 +ε
dψ (x)

0 2m
= 2 (V (x) − E) ψ (x) = 0 . (4.127)
dx
x0 −ε 
x0 −ε

In the limit ε → 0 the right hand side vanishes (assuming ψ (x) is


bounded). Therefore dψ (x) /dx is continuous at x = x0 .
11. Since Vs (−x) = Vs (x) the ground state wavefunction is expected to be
an even function of x. Consider a solution having an energy E and a
wavefunction of the form

 Ae−γx if x > a
ψ (x) = B cos (kx) if − a ≤ x ≤ a , (4.128)

Aeγx if x < −a

where

−2mE
γ= , (4.129)

and

2m (W + E)
k= . (4.130)

Requiring that both ψ (x) and dψ (x) /dx are continuous at x = a yields

Ae−γa = B cos (ka) , (4.131)

and

−γAe−γa = −kB sin (ka) , (4.132)

or in a matrix form
   
A 0
C = , (4.133)
B 0

where

Eyal Buks Quantum Mechanics - Lecture Notes 86


4.7. Solutions
 
e−γa − cos (ka)
C= . (4.134)
−γe−γa k sin (ka)

A nontrivial solution exists iff Det (C) = 0, namely iff


γ
= tan (ka) . (4.135)
k
This condition can be rewritten using Eqs. (4.129) and (4.130) and the
dimensionless parameters
K = ka , (4.136)

2mW
K0 = a, (4.137)

as
 2
1 1 K
cos2 K = 2 =  2 = . (4.138)
1 + tan K 1 + γk K0

Note, however, that according to Eq. (4.135) tan K > 0. Thus, Eq. (4.135)
is equivalent to the set of equations
K
|cos K| = , (4.139)
K0
tan K > 0 . (4.140)
This set has at least one solution (this can be seen by plotting the func-
tions |cos K| and K/K0 ).
12. Final answers: (a) |a1 |2 + |a2 |2 . (b)
7 ! 3
8 3 "2
2 2 8 
π  9 2 4 2 2
∆E = |an | n − |an | n . (4.141)
2ma2 n=1 n=1
 
(c) The same as at t = 0. (d) E = 2π2 2 / ma2 , ∆E = 0.
13. The Schrödinger equation for the wavefunction ψ (x) is given by
 2 
d 2m
+ 2 E ψ (x) = 0 . (4.142)
dx2 
The boundary conditions at x = 0 are
   
ψ 0+ = ψ 0− , (4.143)
+ −
dψ (0 ) dψ (0 ) 2
− = − ψ (0) , (4.144)
dx dx a0
where
2
a0 = . (4.145)

Eyal Buks Quantum Mechanics - Lecture Notes 87


Chapter 4. Quantum Dynamics

Due to symmetry V (x) = V (−x) the solutions are expected to have


definite symmetry (even ψ (x) = ψ (−x) or odd ψ (x) = −ψ (−x)). For
the ground state, which is expected to have even symmetry, we consider
a wavefunction having the form

ψ (x) = Ae−κ|x| , (4.146)

where A is a normalization constants and where



−2mE
κ= . (4.147)

The parameter κ is real for E < 0. This even wavefunction satisfies the
Schrödinger equation for x = 0 and the boundary condition (4.143). The
condition (4.144) leads to a single solution for the energy of the ground
state
mα2
E=− . (4.148)
22
Thus the normalized wavefunction of the ground state ψ0 (x) is given by
0  mα 

ψ0,α (x) = exp − 2 |x| . (4.149)
2 
The probability p that the particle will remain bounded is given by

 ∞
2


p=

ψ0,α1 (x) ψ0,α2 (x) dx

−∞

  
2
4m2 α1 α2

∞ m (α1 + α2 )

= 4
exp − 2
x dx

 0 
4α1 α2
= .
(α1 + α2 )2
(4.150)
14. The normalized wavefunction of the ground state is given by [see Eq.
(4.149)]
0  mα 

ψ0 (x) = exp − 2 |x| . (4.151)
2 

Thus, the probability s (p0 ) is given by

Eyal Buks Quantum Mechanics - Lecture Notes 88


4.7. Solutions

 
2

ip0 x

s (p0 ) =
γ 0 | exp |γ 0 





2

   


mα  ip0 x 2mα

=
2 exp exp − 2 |x| dx


  

−∞
1
=  2 .
 2 p20
1+ 4m2 α2

(4.152)
15. The Schrödinger equation for the wavefunction ψ (x) is given by
2 
d 2m
+ 2 (E − V ) ψ (x) = 0 . (4.153)
dx2 
The boundary conditions imposed upon ψ (x) by the potential are [see
Eq. (4.127)]
ψ (±a) = 0 , (4.154)
   
ψ 0+ = ψ 0− , (4.155)
dψ (0+ ) dψ (0− ) 2
− = − ψ (0) , (4.156)
dx dx a0
where
2
a0 = . (4.157)

Due to symmetry V (x) = V (−x) the solutions are expected to have
definite symmetry (even ψ (x) = ψ (−x) or odd ψ (x) = −ψ (−x)). For
the ground state, which is expected to have even symmetry, we consider
a wavefunction having the form
+
A sinh (κ (x − a)) x > 0
ψ (x) = , (4.158)
−A sinh (κ (x + a)) x < 0
where A is a normalization constants and where

−2mE0
κ= . (4.159)

The parameter κ is real for E0 < 0. This even wavefunction satisfies Eq.
(4.153) for x = 0 and the boundary conditions (4.154) and (4.155). The
condition (4.156) reads
κa0 = tanh (κa) . (4.160)
Nontrivial (κ = 0) real solution exists only when a > a0 , thus E0 < 0 iff
2
a > a0 = . (4.161)

Eyal Buks Quantum Mechanics - Lecture Notes 89


Chapter 4. Quantum Dynamics

16. For the present case the boundary conditions imposed upon ψ (x) by the
potential are [see Eq. (4.127)]
ψ (0) = 0 , (4.162)
 +  
ψ x0 = ψ x− 0 , (4.163)
 +  −
dψ x0 dψ x0 2
− = − ψ (x0 ) , (4.164)
dx dx a0
where
2
a0 = . (4.165)

Consider a solution having the form
+
A sinh (κx) + B cosh (κx) 0 ≤ x < x0
ψ (x) = , (4.166)
e−κ(x−x0 ) x > x0

where

−2mE0
κ= . (4.167)

The boundary conditions yield
B=0, (4.168)
1 = A sinh (κx0 ) , (4.169)
2
−κ (1 + A cosh (κx0 )) = − , (4.170)
a0
thus
2x0
κx0 (1 + coth (κx0 )) = . (4.171)
a0
Note that for x ≥ 0 the following holds x (1 + coth x) ≥ 1, thus a solution
with E0 < 0 (i.e. a solution with a real positive κ) is possible only if
2x0
≥1, (4.172)
a0
or
2
x0 ≥ . (4.173)
2mα
17. Using Eq. (4.37) one has

dx(H) 1 3 (H) 4
= x ,H , (4.174)
dt i
therefore

Eyal Buks Quantum Mechanics - Lecture Notes 90


4.7. Solutions

dx(H) 1 i (Ek − El )
k| |l = k| x(H) H − Hx(H) |l = k| x(H) |l .
dt i 
(4.175)

Integrating yields
 
(H) (H) i (Ek − El ) t
k| x (t) |l = k| x (t = 0) |l exp . (4.176)

Using this result one has

(Ek − El ) |k| x |l|2
k


2

= (Ek − El )
k| x(H) |l

k

= (Ek − El ) k| x(H) |l l| x(H) |k
k
 
  dx(H) (H) (H) dx(H)
= k| |l l| x |k − k| x |l l| |k
2i dt dt
k
 
  (H) dx(H) dx(H) (H)
= l| x |k k| |l − l| |k k| x |l
2i k dt dt
 dx(H) dx(H) (H)
= l| x(H) − x |l .
2i dt dt
(4.177)
Using again Eq. (4.37) one has

1 3 (H) 4 px
(H)
dx(H)
= x ,H = , (4.178)
dt i m
therefore
  3 4
(Ek − El ) |k| x |l|2 = l| x(H) , p(H)
x |l
2im
k

= i
2im
2
= .
2m
(4.179)
18. The wavefunctions of the normalized eigenstates are given by
0
2 nπx
ψn (x) = sin , (4.180)
a a
and the corresponding eigenenergies are

Eyal Buks Quantum Mechanics - Lecture Notes 91


Chapter 4. Quantum Dynamics

π2 2 n2
En = . (4.181)
2ma2
a) The wavefunction after the measurement is a normalized wavepacket
centered at x = a/2 and having a width ∆x
+ 1


∆x
√ x − 2 ≤ 2
ψ (x) = ∆x . (4.182)
0 else
Thus in the limit ∆x << a

 a
2


∆x 2 nπ
pn =

dxψ∗n (x) ψ (x)

≃ 2 sin . (4.183)
0 a 2
Namely, pn = 0 for all even n, and the probability of all energies with
odd n is equal.
b) Generally, for every bound state in one dimension p = 0 [see Eq.
(4.52)].
19. For a well of width a the wavefunctions of the normalized eigenstates are
given by
0
(a) 2 nπx
ψn (x) = sin , (4.184)
a a
and the corresponding eigenenergies are
2 π 2 n2
En(a) = . (4.185)
2ma2
(a) The probability is given by

 a
2

(a) (2a)
32
p=

dxψ1 (x) ψ1 (x)

= 2 . (4.186)
0 9π
(a)
(b) For times t < 0 it is given that H = E1 . Immediately after the
change (t = 0+ ) the wavefunction remains unchanged. A direct evaluation
(a)
of H using the new Hamiltonian yields the same result H = E1 as for
t < 0. At later times t > 0 the expectation value H remains unchanged
due to energy conservation.
20. The wavefunctions of the normalized eigenstates are given by [see Eq.
(4.181)]
0
2 nπx
ψn (x) = sin , (4.187)
a a
and the corresponding eigenenergies are [see Eq. (4.180)]
π2 2 n2
En = , (4.188)
2ma2

Eyal Buks Quantum Mechanics - Lecture Notes 92


4.7. Solutions

where n = 1, 2, · · · . By symmetry for all states x = a/2. Furthermore,


for all states p = 0 [see Eq. (4.52)]. For the n’th state the following
holds
  
) 2 * 2 a ′ ′2 2 nπx′ a2 2n2 π2 − 3
x = dx x sin = , (4.189)
a 0 a 6n2 π2
thus
' ( ) *  
2 2 1 1
(∆x) = x2 − x = a2 − 2 2
12 2n π

and [see Eq. (3.29)]


' ( ) * 2  nπ 2  a nπx′

nπ
2
(∆p)2 = p2 = dx′ sin2 = , (4.190)
a a 0 a a

thus [compare with the uncertainty principle (3.10)]


' (' ( 2  n2 π 2 
(∆x)2 (∆p)2 = −2 . (4.191)
4 3
21. The Schrödinger equation is given by
d |α
i = H |α , (4.192)
dt
where the Hamiltonian is given by [see Eq. (1.62)]
 2
p− qc A
H= + qϕ . (4.193)
2m
Multiplying from the left by x′ | yields
dψ 1  q 2
i = −i∇− A ψ + qϕψ , (4.194)
dt 2m c
where

ψ = ψ (x′ ) = x′ |α . (4.195)

Multiplying Eq. (4.194) by ψ∗ , and subtracting the complex conjugate of


Eq. (4.194) multiplied by ψ yields
 
dρ 1 q 2  q 2
i = ψ∗ −i∇− A ψ − ψ i∇− A ψ∗ , (4.196)
dt 2m c c
where

ρ = ψψ∗ (4.197)

Eyal Buks Quantum Mechanics - Lecture Notes 93


Chapter 4. Quantum Dynamics

is the probability distribution function. Moreover, the following holds


 q 2  q 2
ψ∗ −i∇− A ψ − ψ i∇− A ψ∗
 c c 
 q 2 iq iq
∗ 2 2 2
= ψ − ∇ + A + ∇A+ A∇ ψ
c c c
  q 2 
2 2 2 iq iq
−ψ − ∇ + A − ∇A− A∇ ψ ∗
c c c
 
= −2 ψ ∗ ∇2 ψ − ψ∇2 ψ∗
iq ∗
+ (ψ ∇Aψ + ψ∗ A∇ψ + ψ∇Aψ∗ + ψA∇ψ∗ )
c
iq
= −2 ∇ (ψ∗ ∇ψ − ψ∇ψ∗ ) + ∇ (ψ∗ Aψ + ψAψ∗ ) .
c
(4.198)
Thus, Eq. (4.196) can be written as


+ ∇J = 0 , (4.199)
dt
where
 qρ
J= Im (ψ∗ ∇ψ) − A. (4.200)
m mc
22. The current density J [see Eq. (4.75)] that is associated with the wave-
′ ′
function ψ (x′ ) = Aeikx + Be−ikx is given by
 
 ∂
J= Im ψ∗ ′ ψ
m ∂x
   ′ ′
 ′ ′

= Im ik A∗ e−ikx + B ∗ eikx Aeikx − Be−ikx
m
   ′ ′

= Im ik |A|2 − |B|2 + AB ∗ e2ikx − A∗ Be−2ikx
m
k  2 
= |A| − |B|2 .
m
(4.201)
Thus for a solution to the time independent Schrödinger equation, for
which the current density ρ = ψψ∗ is time independent, the continuity
equation (4.73) yields the relation

|A1 |2 − |B1 |2 = |A2 |2 − |B2 |2 . (4.202)

23. In this problem the potential is piecewise constant. At the points where
the piecewise constant potential abruptly changes the solution has to
satisfy the requirements that both ψ (x) and dψ/dx′ [see Eq. (4.127)] are

Eyal Buks Quantum Mechanics - Lecture Notes 94


4.7. Solutions

continuous. Consider first a general case, where a given potential is taken


to be given by
+
′ Ul x′ ≤ x0
V (x ) = , (4.203)
Ur x′ > x0

where Ul and Ur are constants, and the wavefunction is expressed as


+ ′ ′
′ Al eikl x + Bl e−ikl x x′ ≤ x0
ψ (x ) = ′ ′ , (4.204)
Ar eikr x + Br e−ikr x x′ > x0

where Al , Bl , Ar and Br are constants, and where the constants kl and


kr is related to the energy of the particle E by [see Eq. (4.50)]
2 kl2
= E − Ul , (4.205)
2m
2 kr2
= E − Ur . (4.206)
2m
The requirements that both ψ (x) and dψ/dx′ [see Eq. (4.127)] are con-
tinuous yield a linear relation between the amplitudes on the left Al and
Bl and those on the right Ar and Br
   
Al Ar
=M , (4.207)
Bl Br

where
 −1  ik x 
eikl x0 e−ikl x0 e r 0 e−ikr x0
M=
kl eikl x0 −kl e−ikl x0 kr eikr x0 −kr e−ikr x0
! "
1 1 + kkrl 1 − kkrl
= Φ (kl x0 ) Φ (−kr x0 ) ,
2 1 − kkrl 1 + kkrl
(4.208)
and where the matrix Φ (θ) is defined by
 −iθ 
e 0
Φ (θ) = . (4.209)
0 eiθ

The above general result (4.207) is employed below for the given piecewise
constant potential (4.77). Assume first that the wave function on the right

side of the barrier in the region x′ > a/2 is given by ψ (x′ ) = eikx , and

on the left side of the barrier in the region x < −a/2 it is given by
′ ′
ψ (x′ ) = Aeikx + Be−ikx . With the help of Eq. (4.207) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 95


Chapter 4. Quantum Dynamics
    
A 1 ka 1 + κk 1 − κk
= Φ − Φ (κa)
B 4 2 1 − κk 1 + κk
    
1 + κk 1 − κk ka 1
× Φ −
1 − κk 1 + κk 2 0
 ika  
i k κ
 
e cos κa  − 2 κ + k sin κa
= 1 κ k ,
2 i k − κ sin κa
(4.210)
where
2 k2
=E, (4.211)
2m
2 2
 κ
= E − Ub , (4.212)
2m
and thus
1 e−ika
t==   , (4.213)
A cos κa − 2i κk + κk sin κa
1
 κ k  −ika
B 2 i k − κ e sin κa
r= =  . (4.214)
A cos κa − 2i κk + κk sin κa
Note that, as is expected from current conservation [see Eq. (4.202)], the
following holds |t|2 + |r|2 = 1.
24. Using Eq. (4.45) one has
   
1
p(c) x(c) e  [ξ(x −x)+η(p(c) −p)]
i (c)
A (x, p) = 2 dξdηdx(c) dp(c) .
(2π)
(4.215)

With the help of Eq. (2.175), which is given by

eA eB = eA+B e(1/2)[A,B] , (4.216)

one has
i i i 1
e−  ξx e−  ηp = e−  (ξx+ηp) e− 2 2 ξη[x,p] , (4.217)

thus
   
1 (c) (c) i ξη
p(c) x(c) e  (ξx +ηp ) e  2 e−  ξx e−  ηp dξdηdx(c) dp(c)
i i i
A (x, p) = 2
(2π)
   
1 (c) η (c)
p(c) x(c) e  [(ξ(x + 2 )+ηp )] e−  ξx e−  ηp dξdηdx(c) dp(c) .
i i i
= 2
(2π)
(4.218)
Changing the integration variable

Eyal Buks Quantum Mechanics - Lecture Notes 96


4.7. Solutions
η
x(c) = x(c)′ − , (4.219)
2
one has
    
1 η  i (ξx(c)′ +ηp(c) ) − i ξx − i ηp
A (x, p) = 2 p(c) x(c)′ − e e  e  dξdηdx(c)′ dp(c)
(2π) 2
    
1 (c) (c)′ η  i ξ(x(c)′ −x) i η(p(c) −p)
= p x − e e dξdηdx(c)′ dp(c) .
(2π)2 2
(4.220)
Using the identity
∞
′ ′′
dkeik(x −x ) = 2πδ (x′ − x′′ ) , (4.221)
−∞

one finds that


  
1
e  ξ(x −x) dξ = δ x(c)′ − x ,
i (c)′
(4.222)
2π
  
1
e  η(p −p) dη = δ p(c) − p ,
i (c)
(4.223)
2π
thus
    
1 η  i η(p(c) −p) (c) 1
e  ξ(x −x) dξ
i (c)′
(c) (c)′ (c)′
A (x, p) = p x − e dηdx dp
2π 2 2π
   
1 (c) (c)′ η  i η(p(c) −p)  
= p x − e dηdx dp δ x(c)′ − x
(c)′ (c)
2π 2
    i (c)
1 η
= p(c) x − e  η(p −p) dηdp(c)
2π 2
   
1 1 η i (c)
e  η(p −p) dη − p(c) e  η(p −p) dηdp(c)
i (c)
= p(c) xdp(c)
2π 2π 2
 
1 η
p(c) e  η(p −p) dηdp(c)
i (c)
= px −
2π 2
   η (p
i (c)
−p)
1  (c) ∂e
= px − p dηdp(c)
2π 2i ∂p(c)
 
 ∂ 1 (c)
dηe  η(p −p) .
i
= px − dp(c) p(c) (c)
2i ∂p 2π
  
δ(p(c) −p)

(4.224)
Integration by parts yields

Eyal Buks Quantum Mechanics - Lecture Notes 97


Chapter 4. Quantum Dynamics
  (c)   
 ∂p
A (x, p) = px − (c)
δ p(c) − p dp(c)
2i ∂p

= px −
2i
[x, p]
= px +
2
xp + px
= .
2
(4.225)
 (c) (c) 
25. Below we derive an expression for the variable A x , p in terms
of the matrix elements of the operator A (x, p) in the basis of position

'
|x . To
eigenvectors
′′


that′′end( we begin by evaluating the matrix element
′ x
′ x
x − 2
A (x, p)
x + 2 using Eqs. (4.218), (3.19) and (4.222)
5

6
′ x′′


′ x′′

x − A (x, p)
x +
2
2
      i
1 η
e  [(ξ(x + 2 )+ηp )]
(c) (c)
(c) (c)
= 2 A x , p
(2π)
5

6
′ x′′

− i ξx − i ηp

′ x′′
× x − e e x + dξdηdx(c) dp(c)
2

2
      i  
1 (c) (c)  [(ξ (x
(c) i ′
+ η2 )+ηp(c) )] −  ξ x − 2
x′′
= A x , p e e
(2π)2
5
6
x′′

′ x′′
× x′ − x + + η dξdηdx(c) dp(c)
2
2
    i ′′ (c) 
1 ′
e  [ξ(x −x )] dξ
i (c)
(c) (c) − x p (c) (c)
= 2 A x , p e dx dp
(2π)
    i ′′ (c)  
1
= A x(c) , p(c) e−  x p dx(c) dp(c) δ x(c) − x′
2π
   i ′′ (c)
1
= A x′ , p(c) e−  x p dp(c) .
2π
i ′′ ′
Applying the inverse Fourier transform, i.e. multiplying by e  x p
and
integrating over x′′ yields
 5

6
′ x′′


′ x′′ i ′′ ′
x −

A (x, p)
x + e  x p dx′′
2
2
   
1 i ′′ ′
A x′ , p(c) dp(c) e  x (p −p ) dx′′ ,
(c)
=
2π
(4.226)
thus with the help of Eq. (4.223) one finds the desired inversion of Eq.
(4.45) is given by

Eyal Buks Quantum Mechanics - Lecture Notes 98


4.7. Solutions
 5

6
′ ′ ′ x′′


′ x′′ i ′′ ′
A (x , p ) = x −
A (x, p)
x +

e  x p dx′′ . (4.227)
2 2

Note that A (x′ , p′ ), which appears on the left hand side of the above
equation (4.227) is a classical variable, whereas A (x, p) on the right hand
side is the corresponding quantum operator. A useful relations can be
obtained by integrating A (x′ , p′ ) over p′ . With the help of Eq. (4.222)
one finds that
  5

6
x′′


′ x′′ i ′′ ′
A (x′ , p′ ) dp′ = dx′′ x′ − A (x, p)
x + e  x p dp′
2

2
= 2π x′ | A (x, p) |x′  .
(4.228)
Another useful relations can be obtained by integrating A (x′ , p′ ) over
x′ .With the help of Eqs. (3.52) and (4.223) one finds that
   5

6
′ ′ ′ ′ x′′


′ x′′ i ′′ ′
A (x , p ) dx = x −
A (x, p)
x +
e  x p dx′′ dx′
2 2
    5 ′′

6
′ x ′′ ′′ ′′′

′′′
′ x′′ i ′′ ′
= x − |p  p | A (x, p) |p  p
x + e  x p dx′′ dx′ dp′′ dp′′′
2 2
   
1 i ′ ′′ ′′′ i x′′ ′′ ′′′
e  x (p −p ) e  2 (−p −p ) p′′ | A (x, p) |p′′′  e  x p dx′′ dx′ dp′′ dp′′′
i ′′ ′
=
2π
  
i x′′ ′′ ′′′
δ (p′′ − p′′′ ) e  2 (−p −p ) p′′ | A (x, p) |p′′′  e  x p dx′′ dp′′ dp′′′
i ′′ ′
=
 
i ′′ ′ ′′
= p′′ | A (x, p) |p′′  e  x (p −p ) dx′′ dp′′

= 2π p′′ | A (x, p) |p′′  δ (p′ − p′′ ) dp′′

= 2π p′ | A (x, p) |p′  .


(4.229)

Eyal Buks Quantum Mechanics - Lecture Notes 99


5. The Harmonic Oscillator

Consider a particle of mass m in a parabolic potential well


1
U (x) = mω2 x2 ,
2
where the angular frequency ω is a constant. The classical equation of motion
for the coordinate x is given by [see Eq. (1.19)]
∂U
mẍ = − = −mω 2 x . (5.1)
∂x
It is convenient to introduce the complex variable α, which is given by
 
1 i
α= x + ẋ , (5.2)
x0 ω
where x0 is a constant having dimension of length. Using Eq. (5.1) one finds
that
   
1 i 1 i 2
α̇ = ẋ + ẍ = ẋ − ω x = −iωα . (5.3)
x0 ω x0 ω
The solution is given by

α = α0 e−iωt , (5.4)

where α0 = α (t = 0). Thus, x and ẋ oscillate in time according to


 
x = x0 Re α0 e−iωt , (5.5)
 
ẋ = x0 ω Im α0 e−iωt . (5.6)
The Hamiltonian is given by [see Eq. (1.34)]

p2 mω 2 x2
H= + . (5.7)
2m 2
In quantum mechanics the variables x and p are regarded as operators satis-
fying the following commutation relations [see Eq. (3.9)]

[x, p] = xp − px = i . (5.8)
Chapter 5. The Harmonic Oscillator

5.1 Eigenstates

The annihilation and creation operators are defined as


0  
mω ip
a= x+ , (5.9)
2 mω
0  
† mω ip
a = x− . (5.10)
2 mω
The inverse transformation is given by
0
  
x= a + a† , (5.11)
2mω
0
mω  
p=i −a + a† . (5.12)
2
The following holds
, - i
a, a† = ([p, x] − [x, p]) = 1 , (5.13)
2
The number operator, which is defined as

N = a† a, (5.14)

can be expressed in terms of the Hamiltonian


N = a† a   
mω ip ip
= x− x+
2 mω mω
 2 
mω p 2 i [x, p]
= +x +
2 m2 ω2 mω
 2 2 2

1 p mω x 1
= + −
ω 2m 2 2
H 1
= − .
ω 2
(5.15)
Thus, the Hamiltonian can be written as
 
1
H = ω N + . (5.16)
2

The operator N is Hermitian, i.e. N = N † , therefore its eigenvalues are


expected to be real. Let {|n} be the set of eigenvectors of N and let {n} be
the corresponding set of eigenvalues

Eyal Buks Quantum Mechanics - Lecture Notes 102


5.1. Eigenstates

N |n = n |n . (5.17)

According to Eq. (5.16) the eigenvectors of N are also eigenvectors of H

H |n = En |n , (5.18)

where the eigenenergies En are given by


 
1
En = ω n + . (5.19)
2

Theorem 5.1.1. Let |n be a normalized eigenvector of the operator N with


eigenvalue n. Then (i) the vector

|n + 1 = (n + 1)−1/2 a† |n (5.20)

is a normalized eigenvector of the operator N with eigenvalue n + 1; (ii) the


vector

|n − 1 = n−1/2 a |n (5.21)

is a normalized eigenvector of the operator N with eigenvalue n − 1

Proof. Using the commutation relations


, - , -
N, a† = a† a, a† = a† , (5.22)
, -
[N, a] = a† , a a = −a , (5.23)
one finds that
, - 
N a† |n = N, a† + a† N |n = (n + 1) a† |n , (5.24)

and

N a |n = ([N, a] + aN ) |n = (n − 1) a |n . (5.25)

Thus, the vector a† |n, which is proportional to |n + 1, is an eigenvector of


the operator N with eigenvalue n + 1 and the vector a |n, which is propor-
tional to |n − 1, is an eigenvector of the operator N with eigenvalue n − 1.
Normalization is verified as follows
, -
n + 1 |n + 1 = (n + 1)−1 n| aa† |n = (n + 1)−1 n| a, a† + a† a |n = 1 ,
(5.26)

and

n − 1 |n − 1 = n−1 n| a† a |n = 1 . (5.27)

Eyal Buks Quantum Mechanics - Lecture Notes 103


Chapter 5. The Harmonic Oscillator

As we have seen from the above theorem the following hold



a |n = n |n − 1 , (5.28)

a† |n = n + 1 |n + 1 . (5.29)

Claim. The spectrum (i.e. the set of eigenvalues) of N are the nonnegative
integers {0, 1, 2, · · · }.

Proof. First, note that since the operator N is positive-definite the eigenval-
ues are necessarily non negative

n = n| a† a |n ≥ 0 . (5.30)

On the other hand, according to Eq. (5.28), if n is an eigenvalue also n − 1


is an eigenvalue, unless n = 0. For the later case according to Eq. (5.28)
a |0 = 0. Therefore, n must be an integer, since otherwise one reaches a
contradiction with the requirement that n ≥ 0.

According to exercise 6 of set 4, in one-dimensional problems the energy


spectrum of the bound states is always non-degenerate. Therefore, one con-
cludes that all eigenvalues of N are non-degenerate. Therefore, the closure
relation can be written as


1= |n n| . (5.31)
n=0

Furthermore, using Eq. (5.29) one can express the state |n in terms of the
ground state |0 as
 † n
a
|n = √ |0 . (5.32)
n!

5.2 Coherent States

As can be easily seen from Eqs. (5.11), (5.12), (5.28) and (5.29), all energy
eigenstates |n have vanishing position and momentum expectation values
n| x |n = 0 , (5.33)
n| p |n = 0 . (5.34)
Clearly these states don’t oscillate in phase space as classical harmonic os-
cillators do. Can one find quantum states having dynamics that resembles
classical harmonic oscillators?

Eyal Buks Quantum Mechanics - Lecture Notes 104


5.2. Coherent States

Definition 5.2.1. Consider a harmonic oscillator having ground state |0.


A coherent state |α with a complex parameter α is defined by

|α = D (α) |0 , (5.35)

where
 
D (α) = exp αa† − α∗ a , (5.36)

is the displacement operator.

In the set of problems at the end of this chapter the following results are
obtained:
• The displacement operator is unitary D† (α) D (α) = D (α) D† (α) = 1.
• The coherent state |α is an eigenvector of the operator a with an eigenvalue
α, namely

a |α = α |α . (5.37)


 †
• For any function f a, a having a power series expansion the following
holds
   
D† (α) f a, a† D (α) = f a + α, a† + α∗ . (5.38)

• The displacement operator satisfies the following relations


|α|2 † ∗ |α|2 ∗ †
D (α) = e− 2 eαa e−α a = e 2 e−α a eαa , (5.39)
√ mω α−α∗ i α+α ∗ ∗2 2
√ x − √m √ p α 4−α
D (α) = e  2 e ω 2 e , (5.40)
 ′∗ ∗ ′

αα − α α
D (α) D (α′ ) = exp D (α + α′ ) . (5.41)
2
• Coherent state expansion in the basis of number states
∞

|α|2 αn
|α = e 2 √ |n . (5.42)
n=0 n!

• The following expectation values hold

Eyal Buks Quantum Mechanics - Lecture Notes 105


Chapter 5. The Harmonic Oscillator
 
2
Hα = α| H |α = ω |α| + 1/2 , (5.43)
 
α| H2 |α = 2 ω2 |α|4 + 2 |α|2 + 1/4 , (5.44)
.
2
∆Hα = α| (∆H) |α = ω |α| , (5.45)
0
2
xα = α| x |α = Re (α) , (5.46)


pα = α| p |α = 2mω Im (α) , (5.47)
. 0
2 
∆xα = α| (∆x) |α = , (5.48)
2mω
. 0
2 mω
∆pα = α| (∆p) |α = , (5.49)
2

∆xα ∆pα = . (5.50)
2
• The wave function of a coherent state is given by
ψα (x′ ) = x′ |α
 ∗2    2 
α − α2  mω 1/4 x′ − xα x′
= exp exp − + i pα .
4 π 2∆xα 
(5.51)
• The following closure relation holds
 
1
1= |α α| d2 α , (5.52)
π

where d2 α denotes infinitesimal area in the α complex plane, namely d2 α =


d {Re α} d {Im α}.
Given that at time t = 0 the oscillator is in a coherent state with para-
meter α0 , namely |ψ (t = 0) = |α0 , the time evolution can be found with
the help of Eqs. (4.14), (5.19) and (5.42)
∞  
|α |2  iEn t αn0
− 20
|ψ (t) = e exp − √ |n
n=0
 n!


|α0 |2 αn
= e−iωt/2 e− 2 exp (−iωnt) √ 0 |n
n=0 n!
∞  n
|α0 |2  α0 e−iωt
= e−iωt/2 e− 2 √ |n
n=0 n!

*
= e−iωt/2
α = α0 e−iωt .
(5.53)

Eyal Buks Quantum Mechanics - Lecture Notes 106


5.3. Problems

In view of Eqs. (5.43), (5.45) (5.48) and (5.49), we see from this results that
Hα , ∆Hα , ∆xα and ∆pα are all time independent. On the other hand, as
can be seen from Eqs. (5.46) and (5.47) the following holds
0
2  
xα = α| x |α = Re α0 e−iωt , (5.54)

√  
pα = α| p |α = 2mω Im α0 e−iωt . (5.55)
These results show that indeed, xα and pα have oscillatory time depen-
dence identical to the dynamics of the position and momentum of a classical
harmonic oscillator [compare with Eqs. (5.5) and (5.6)].

5.3 Problems

1. Calculate the wave functions ψn (x′ ) = x′ |n of the number states |n
of a harmonic oscillator.
2. Show that

   tn
exp 2Xt − t2 = Hn (X) , (5.56)
n=0
n!

where Hn (X) is the Hermite polynomial of order n, which is defined by


 2 n  
X d X2
Hn (X) = exp X− exp − . (5.57)
2 dX 2

3. Show that
 
α(2XY −αX 2 −αY 2 )
∞  α n exp
 Hn (X) Hn (Y ) 1−α2
2
= √ , (5.58)
n=0
n! 1 − α2

where Hn (X) is the Hermite polynomial of order n.


4. Show that for the state |n of a harmonic oscillator
' (' (  1
2
2 2
(∆x) (∆p) = n + 2 . (5.59)
2

5. Consider a free particle in one dimension having mass m. Express the


Heisenberg operator x(H) (t) in terms x(H) (0) and p(H)
' (0). At( time t = 0
2
the system in in the state |ψ0 . Express the variance (∆x) (t) at time
t, where ∆x = x − x, in terms of the following expectation values at
time t = 0

Eyal Buks Quantum Mechanics - Lecture Notes 107


Chapter 5. The Harmonic Oscillator

x0 = ψ0 | x |ψ0  , (5.60)


p0 = ψ0 | p |ψ0  , (5.61)
(xp)0 = ψ0 | xp |ψ0  , (5.62)
(∆x)20 = ψ0 | (x − x0 )2 |ψ0  , (5.63)
(∆p)20 2
= ψ0 | (p − p0 ) |ψ0  . (5.64)
6. Consider a harmonic oscillator of angular frequency ω and mass m.
a) Express the Heisenberg picture x(H) (t) and p(H) (t) in terms x(H) (0)
and p(H) (0). , - , -
b) Calculate the following commutators p(H) (t1 ) , x(H) (t2 ) , p(H) (t1 ) , p(H) (t2 )
, -
and x(H) (t1 ) , x(H) (t2 ) .
7. Consider a particle having mass m confined by a one dimensional poten-
tial V (x), which is given by
+ mω2
2
V (x) = 2 x x>0 , (5.65)
∞ x≤0
where ω is a constant.
a) Calculate the eigenenergies of the) system.
*
b) Calculate the expectation values x2 of all energy eigenstates of the
particle.
8. Calculate the possible energy values of a particle in the potential given
by
mω2 2
V (x) = x + αx . (5.66)
2
9. A particle is in the ground state of harmonic oscillator with potential
energy
mω2 2
V (x) = x . (5.67)
2
Find the probability p to find the particle in the classically forbidden
region.
10. Consider a particle having mass m in a potential V given by
+ mω2 z2 a
− 2 ≤ x ≤ a2 and − a2 ≤ y ≤ a2
V (x, y, z) = 2 , (5.68)
∞ else
where ω and a are positive real constants. Find the eigenenergies of the
system.
11. Consider a harmonic oscillator having angular resonance frequency ω 0 .
At time t = 0 the system’s state is given by
1
|α (t = 0) = √ (|0 + |1) , (5.69)
2

Eyal Buks Quantum Mechanics - Lecture Notes 108


5.3. Problems

where the states |0 and |1 are the ground and first excited states, re-
spectively, of the oscillator. Calculate as a function of time t the following
quantities:
a) x
b) )p *
c) x2
d) ∆x∆p
12. Harmonic oscillator having angular resonance frequency ω is in state
1
|ψ (t = 0) = √ (|0 + |n) (5.70)
2
at time t = 0, where |0 is the ground state and |n is the eigenstate
with eigenenergy ω (n + 1/2) (n is a non zero integer). Calculate the
expectation value x for time t ≥ 0.
13. Consider a harmonic oscillator having mass m and angular resonance
frequency ω . At time t = 0 the system’s state is given by |ψ(0) = c0 |0+
c1 |1 , where |n are the eigenstates with energies En = ω (n +.1/2).
1 
Given that H = ω, |ψ(0) is normalized, and x (t = 0) = 2 mω ,
calculate x (t) at times t > 0.
14. Show that
|α|2 † ∗ |α|2 ∗ †
D (α) = e− 2 eαa e−α a
=e 2 e−α a eαa . (5.71)

15. Show that the displacement operator D (α) is unitary.


16. Show that
∞
|α|2 αn
|α = e− 2 √ |n . (5.72)
n=0 n!

17. Show that the coherent state |α is an eigenvector of the operator a with
an eigenvalue α, namely

a |α = α |α . (5.73)

18. Show that


0 
mω α − α∗
D (α) = exp √ x
 2
   ∗2 
i α + α∗ α − α2
× exp − √ √ p exp .
mω 2 4
(5.74)
 †
19. Show that for any function f a, a having a power series expansion the
following holds
   
D† (α) f a, a† D (α) = f a + α, a† + α∗ . (5.75)

Eyal Buks Quantum Mechanics - Lecture Notes 109


Chapter 5. The Harmonic Oscillator

20. Show that the following holds for a coherent state |α:
 
2
a) α| H |α = ω |α| + 1/2 .
 
4 2
b) α| H2 |α = 2 ω2 |α| + 2 |α| + 1/4 .
.
c) α| (∆H)2 |α = ω |α|.
.
2
d) xα = α| x |α = mω Re (α).

e) pα = α|
. p |α = 2mω. Im (α).
2 
f) ∆xα = α| (∆x) |α = 2mω .
. .
2
g) ∆pα = α| (∆p) |α = mω 2 .
21. Consider a harmonic oscillator of mass m and angular resonance fre-
quency ω. The Hamiltonian is given by

p2 1
H= + mω 2 x2 . (5.76)
2m 2
The system at time t is in a normalized state |α, which is an eigenvector
of the annihilation operator a, thus

a |α = α |α , (5.77)

where the eigenvalue α is a complex number. At time t > 0 the energy of


the system is measured. What are the possible results En and what are
the corresponding probabilities pn (t)?
22. Show that the wave function of a coherent state is given by
ψα (x′ ) = x′ |α
 ∗2    2 
α − α2  mω 1/4 x − xα x
= exp exp − + i pα .
4 π 2∆xα 
(5.78)
23. Show that
 
αα′∗ − α∗ α′
D (α) D (α′ ) = exp D (α + α′ ) . (5.79)
2

24. Show that the following closure relation holds


 
1
1= |α α| d2 α , (5.80)
π

where d2 α denotes infinitesimal area in the α complex plane, namely


d2 α = d {Re α} d {Im α}.
25. Calculate the inner product between two coherent states |α and |β,
where α, β ∈ C.

Eyal Buks Quantum Mechanics - Lecture Notes 110


5.3. Problems

26. A one dimensional potential acting on a particle having mass m is given


by
1
V1 (x) = mω 2 x2 + βmω 2 x . (5.81)
2
a) Calculate the Heisenberg representation of the position operator
x(H) (t) and its canonically conjugate operator p(H) (t).
b) Given that the particle at time t = 0 is in the state |0, where the
state |0 is the ground state of the potential
1
V1 (x) = mω2 x2 . (5.82)
2
Calculate the expectation value x at later times t > 0.
27. A particle having mass m is in the ground state of the one-dimensional
potential well V1 (x) = (1/2) mω 2 (x − ∆x )2 for times t < 0 . At time
t = 0 the potential suddenly changes and becomes V2 (x) = (1/2) mω 2 x2 .
a) Calculate the expectation
' value ( x at times t > 0.
2
b) Calculate the variance (∆x) at times t > 0 , where ∆x = x − x.
c) The energy of the particle is measured at time t > 0 . What are the
possible results and what are the probabilities to obtain any of these
results.
28. Consider a particle having mass m in the ground state of the potential
well Va (x) = (1/2) mω2 x2 for times t < 0 . At time t = 0 the potential
suddenly changes and becomes Vb (x) = gx . (a) Calculate the ' expecta-
(
tion value x at times t > 0 . (b) Calculate the variance (∆x)2 at
times t > 0 , where ∆x = x − x.
29. Consider a particle of mass m in a potential of a harmonic oscillator
having angular frequency ω. The operator S (r) is defined as
3 r  2 4
S (r) = exp a† − a2 , (5.83)
2
where r is a real number, and a and a† are the annihilation and creation
operators respectively. The operator T is defined as
T = S (r) aS † (r) . (5.84)
a) Find an expression for the operator T of the form T = Aa + Ba† ,
where both A and B are constants.
b) The vector state |r is defined as
|r = S † (r) |0 , (5.85)
where |0 is the ground state of the harmonic oscillator. Calculate
the expectation values r| x |r of the operator x (displacement) and
the expectation value r| p |r of the operator p (momentum).

Eyal Buks Quantum Mechanics - Lecture Notes 111


Chapter 5. The Harmonic Oscillator

c) Calculate the variance (∆x)2 of x and the variance (∆p)2 of p.


30. The state |r from the previous exercise, which is called a squeezed state,
can be alternatively defined as a normalized state that satisfies the rela-
tion
Q (r) |r = 0 , (5.86)
where the operator Q (r) is defined by
Q (r) = a cosh r + a† sinh r , (5.87)

r is a real number, and a and a are the annihilation and creation opera-
tors respectively. Based on the above definition calculate the expectation
values r| x |r of the position operator x, the expectation value r| p |r
2
of the momentum operator p, the variance (∆x) of x and the variance
(∆p)2 of p with respect to the state |r.
31. Consider one dimensional motion of a particle having mass m. The Hamil-
tonian is given by
H = ω 0 a† a + ω1 a† a† aa , (5.88)
where
0  
mω 0 ip
a= x+ , (5.89)
2 mω0
is the annihilation operator, x is the coordinate and p is its canonical
conjugate momentum. The frequencies ω 0 and ω 1 are both positive.
a) Calculate the eigenenergies of the system.
b) Let |0 be the ground state of the system. Calculate
i. 0|x|0
ii. 0|p|0
' (
iii. 0| (∆x)2 |0
' (
iv. 0| (∆p)2 |0
32. The Hamiltonian of a system is given by
H = ǫN , (5.90)
where the real non-negative parameter ǫ has units of energy, and where
the operator N is given by
N = b† b . (5.91)
The following holds
b† b + bb† = 1 , (5.92)
b2 = 0 , (5.93)
 † 2
b =0. (5.94)

Eyal Buks Quantum Mechanics - Lecture Notes 112


5.3. Problems

a) Find the eigenvalues of H. Clue: show first that N 2 = N .


b) Let |0 be the ground state of the system, which is assumed to be
non-degenerate. Define
 the two states
|+ = A+ 1 + b† |0 , (5.95a)
 †

|− = A− 1 − b |0 , (5.95b)
where the real non-negative numbers A+ and A− are normalization
constants. Calculate A+ and A− . Clue: show first that b† |0 is an
eigenvector of N .
c) At time t = 0 the system is in the state

|α (t = 0) = |+ , (5.96)

Calculate the probability p (t) to find the system in the state |− at
time t > 0.
 
33. Normal ordering - Let f a, a† be a function of  the† annihilation a and

creation
 a operators.
 The normal ordering of f a, a , which is denoted
by : f a, a† : places the a operators on the right and the a† operators
on the left. Some examples are given below
: aa† : = a† a , (5.97)
: a† a : = a† a , (5.98)
 n  n
: a† a : = a† an . (5.99)
Normal ordering is linear, i.e. : f + g :=: f : + : g :. Show that the projec-
tion operator Pn = |n n|, where |n is an eigenvector of the Hamiltonian
of a harmonic oscillator, can be expressed as
1  † n  
Pn = : a exp −a† a an : . (5.100)
n!
34. Consider a harmonic oscillator of angular frequency ω and mass m. A
time dependent force is applied f (t). The function f (t) is assumed to
vanish f (t) → 0 in the limit t → ±∞. Given that the oscillator was
initially in its ground state |0 at t → −∞ calculate the probability pn
to find the oscillator in the number state |n in the limit t → ∞.
35. The parity operator P is defined by
∞
P= dx′ |x′  −x′ | , (5.101)
−∞

where |x′  is an eigenvector of the position operator x having eigenvalue


x′ , i.e. x |x′  = x′ |x′ . Express the parity operator P as a function of the
number operator N = a† a.
36. Show that
† ,  -
eλa a =: exp eλ − 1 a† a : . (5.102)

Eyal Buks Quantum Mechanics - Lecture Notes 113


Chapter 5. The Harmonic Oscillator

37. Consider a harmonic oscillator having mass m and angular resonance


frequency ω. Show that
 ′2 √ ′ †2

x
exp − 2x2 + 2 xx0 a† − a2
|x′  = 0
1/2
|0 , (5.103)
π1/4 x0
where |x′  is an eigenvector of the position operator x with eigenvalue x′ ,
i.e. x |x′  = x′ |x′ ,
0  
mω ip
a= x+ (5.104)
2 mω
is the annihilation operator, |0 is the ground state and
0

x0 = . (5.105)

38. Show that
   2
1 κ x2 x
√ exp =: exp κ 2 : , (5.106)
1+κ 1 + κ x20 x0
where x is the position operator, κ is real and
0

x0 = . (5.107)

39. Let F (X) be a smooth function of the normalized position operator X

a + a† x
X= √ = , (5.108)
2 x0

where a is the annihilation operator, x is the position operator and


0

x0 = . (5.109)

Show that
d dF
: F (X) : =: : . (5.110)
dX dX
40. Consider a system having Hamiltonian H given by
 k
H = ωa† a + ω 1 a† a , (5.111)

where a and a† are the annihilation and creation operators, both ω and
ω 1 are positive, and where k is integer. At initial time t = 0 the state
of the system is an eigenstate of the operator a with eigenvalue α, i.e.
|ψ (t = 0) = |αc , where a |αc = α |αc .

Eyal Buks Quantum Mechanics - Lecture Notes 114


5.4. Solutions

a) Find a general expression for the state of the system |ψ (t) at time
t > 0.
b) Evaluate |ψ (t) at time t = 2π/ω 1 .
c) Evaluate |ψ (t) at time t = π/ω1 .
d) Evaluate |ψ (t) at time t = π/2ω 1 for the case where k is even.
41. Consider two normalized coherent states |α and |β, where α, β ∈ C.
The operator A is defined as

A = |α α| − |β β| . (5.112)

Find the eigenvalues of the operator A.

5.4 Solutions

1. The Hamiltonian is given by

p2 mω2 x2
H= + .
2m 2
Using Eqs. (3.21), (3.29), (5.9) and (5.10) one has
 

 2 −1/2 ′ ′ 2 dψ n
x | a |n = 2x0 x ψn (x ) + x0 ′ , (5.113)
dx
 
 −1/2 ′ dψ
x′ | a† |n = 2x20 x ψn (x′ ) − x20 n′ , (5.114)
dx
where
0

x0 = . (5.115)

For the ground state |0, according to Eq. (5.28), a |0 = 0, thus
dψ0
x′ ψ0 (x′ ) + x20 =0. (5.116)
dx′
The solution is given by
!  2 "
′ 1 x′
ψ0 (x ) = A0 exp − , (5.117)
2 x0

where the normalization constant A0 is found from the requirement


 ∞
2
|ψ0 (x′ )| dx = 1 , (5.118)
−∞

thus

Eyal Buks Quantum Mechanics - Lecture Notes 115


Chapter 5. The Harmonic Oscillator
 !   "
∞ 2
2 x
|A0 | exp − dx = 1 . (5.119)
−∞ x0
  

πx0

Choosing A0 to be real leads to


!  2 "
′ 1 1 x′
ψ0 (x ) = 1/2
exp − . (5.120)
π1/4 x 2 x0
0

All other wavefunctions are found using Eqs. (5.32) and (5.114)
 n
1 2 d
ψn (x′ ) = √ x ′
− x 0 ψ0 (x′ )
(2x0 )n/2 n! dx′
 n !  2 "
1 1 ′ 2 d 1 x′
= √ x − x0 ′ exp − .
π 1/4 2n n! x0n+1/2 dx 2 x0
(5.121)
Using the notation
 2 n  
X d X2
Hn (X) = exp X− exp − , (5.122)
2 dX 2

the expression for ψn (x′ ) can be rewritten as


 ′2   ′
x
exp − 2x 2 Hn xx0
ψn (x′ ) = 0
1/2 √ n
. (5.123)
π 1/4 x0 2 n!

The term Hn (X), which is called the Hermite polynomial of order n, is


calculated below for some low values of n
H0 (X) = 1 , (5.124)
H1 (X) = 2X , (5.125)
H2 (X) = 4X 2 − 2 , (5.126)
H3 (X) = 8X 3 − 12X , (5.127)
H4 (X) = 16X 4 − 48X 2 + 12 . (5.128)
2. Therelation (5.56),
 which is a Taylor expansion of the function f (t) =
exp 2Xt − t2 around the point t = 0, implies that

dn  

Hn (X) = n exp 2Xt − t2

. (5.129)
dt t=0

2
The identity 2Xt − t2 = X 2 − (X − t) yields

Eyal Buks Quantum Mechanics - Lecture Notes 116


5.4. Solutions

 2  dn  

Hn (X) = exp X exp − (X − t)


. (5.130)
dtn t=0

Moreover, using the relation


d   d  
2 2
exp − (X − t) = − exp − (X − t) , (5.131)
dt dX
one finds that
dn  

  2

Hn (X) = exp X 2 (−1)n exp − (X − t)

dX n t=0
 2 n  
n d 2
= exp X (−1) exp −X .
dX n
(5.132)
Note that for an arbitrary function g (X) the following holds
 
  d   d
− exp X 2 exp −X 2 g = 2X − g, (5.133)
dX dX

and
      
X2 d X2 d
exp X− exp − g = 2X − g, (5.134)
2 dX 2 dX

thus
  n  
X2 d X2
Hn (X) = exp X− exp − . (5.135)
2 dX 2

3. With the help of Eq. (5.132) and the general identity


∞ 0
  π 14 4ca+b 2
exp −ax2 + bx + c dx = e a , (5.136)
a
−∞

according to which the following holds (for the case a = 1, b = 2iX and
c = 0)
∞
  1  
exp −X 2 = √ exp −x2 + 2iXx dx , (5.137)
π
−∞

one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 117


Chapter 5. The Harmonic Oscillator
  n ∞
exp X 2 d  
Hn (X) = √ − exp −x2 + 2iXx dx
π dX
−∞
∞
1 n  
= √ (−2ix) exp X 2 − x2 + 2iXx dx
π
−∞
∞
1 n 2
= √ (−2ix) e(X+ix) dx ,
π
−∞
(5.138)
thus the following holds [see Eq. (5.136)]
∞  α n ∞ ∞ ∞ n
Hn (X) Hn (Y ) 1 2 2  (−2αxy)
2
= dx dy e(X+ix) e(Y +iy)
n=0
n! π n=0
n!
−∞ −∞   
e−2αxy
∞ ∞
1 2 2
= dx e(X+ix) dy e(Y +iy) e−2αxy
π
−∞ −∞
  
√ αx(αx−2iY )
πe
∞
1
dx e−(1−α )x +2i(X−Y α)x+X
2 2 2
= √
π
−∞
 
α(2XY −αX 2 −αY 2 )
exp 1−α2
= √ .
1 − α2
(5.139)
4. With the help of Eqs. (5.9), (5.10), (5.11), (5.12) and (5.13) one finds
n| x |n = 0 , (5.140)
 
n| x2 |n = n| aa† + a† a |n = (2n + 1) , (5.141)
2mω 2mω
n| p |n = 0 , (5.142)
mω mω
n| p2 |n = n| aa† + a† a |n = (2n + 1) , (5.143)
2 2
thus
' (' (  1
2
2 2
(∆x) (∆p) = n + 2 .
2
5. The Hamiltonian is given by
p2
H= . (5.144)
2m

Eyal Buks Quantum Mechanics - Lecture Notes 118


5.4. Solutions

Using Eqs. (4.37) and (5.8) one finds that

dx(H) 1 3 (H) (H) 4 p(H) 3 (H) (H) 4 p(H)


= x ,H = x ,p = , (5.145)
dt i im m
and
dp(H) 1 3 (H) (H) 4
= p ,H =0. (5.146)
dt i
The solution is thus
1 (H)
x(H) (t) = x(H) (0) + p (0) t . (5.147)
m
With the help of Eq. (5.147) one finds that
' ( ) *
(∆x)2 (t) = x2 (t) − (x (t))2
 2
(H) 1 (H)
= ψ0 | x (0) + p (0) t |ψ0 
m
   2
1
− ψ0 | x(H) (0) + p(H) (0) t |ψ0 
m
2
t 2t
= (∆x)20 + 2 (∆p)20 + ((xp)0 − x0 p0 ) .
m m
(5.148)
6. The Hamiltonian is given by

p2 mω2 x2
H= + . (5.149)
2m 2
Using Eqs. (4.37) and (5.8) one finds that

dx(H) 1 3 (H) (H) 4 p(H)


= x ,H = , (5.150)
dt i m
and
dp(H) 1 3 (H) (H) 4
= p ,H = −mω2 x(H) . (5.151)
dt i
a) The solutions of the above equations are given by

sin (ωt) (H)


x(H) (t) = x(H) (0) cos (ωt) + p (0) , (5.152)

and

p(H) (t) = p(H) (0) cos (ωt) − mω sin (ωt) x(H) (0) . (5.153)

Eyal Buks Quantum Mechanics - Lecture Notes 119


Chapter 5. The Harmonic Oscillator

b) Using the expressions for x(H) (t) and p(H) (t) and Eq. (5.8) one finds
that 3 4
p(H) (t1 ) , x(H) (t2 )
3 4
= − (cos (ωt1 ) cos (ωt2 ) + sin (ωt1 ) sin (ωt2 )) x(H) (0) , p(H) (0)
= −i cos (ω (t1 − t2 )) ,
3 4 (5.154)
p(H) (t1 ) , p(H) (t2 )
3 4
= mω (cos (ωt1 ) sin (ωt2 ) − sin (ωt1 ) cos (ωt2 )) x(H) (0) , p(H) (0)
= −imω sin (ω (t1 − t2 )) ,
(5.155)
and 3 4
x(H) (t1 ) , x(H) (t2 )
1 3 4
= (cos (ωt1 ) sin (ωt2 ) − sin (ωt1 ) cos (ωt2 )) x(H) (0) , p(H) (0)

i
=− sin (ω (t1 − t2 )) .

(5.156)
7. Due to the infinite barrier for x ≤ 0 the wavefunction must vanish at
x = 0. This condition is satisfied by the wavefunction of all number
states |n with odd value of n (the states |n are  eigenstates
 of the ’regu-
lar’ harmonic oscillator with potential V (x) = mω 2 /2 x2 ). These wave-
functions obviously satisfy the Schrödinger equation for x > 0.
a) Thus the possible energy values are Ek = ω (2k + 3/2) where k =
0, 1, 2, · · · .
b) The corresponding normalized wavefunctions are given by
+√
2ψ2k+1 (x) x > 0
ψ̃k (x) = , (5.157)
0 x≤0

where ψn (x) is the wavefunction of the number states |n. Thus for
a given k

Eyal Buks Quantum Mechanics - Lecture Notes 120


5.4. Solutions

∞

2
) 2*

x k= dx
ψ̃ k (x)
x2
0
∞


2
=2 dx
ψ2k+1 (x)
x2
0
∞


2
= dx
ψ2k+1 (x)
x2
−∞

= 2k + 1| x2 |2k + 1 ,
(5.158)
thus with the help of Eq. (5.141) one finds that
 
) 2*  3
x k= 2k + . (5.159)
mω 2

8. The potential can be written as

mω2  α 2 α2
V (x) = x+ − . (5.160)
2 mω 2 2mω 2
This describes a harmonic oscillator centered at x0 = −α/mω 2 having
angular resonance frequency ω. The last constant term represents energy
shift. Thus, the eigenenergies are given by

En = ω (n + 1/2) − α2 /2mω2 , (5.161)

where n = 0, 1, 2, · · · .
9. In the classically forbidden region V (x) > E0 = ω/2, namely |x| > x0
where
0

x0 = . (5.162)

Using Eq. (5.120) one finds
 ∞
p=2 |ψ0 (x)|2 dx
x0
 ∞ !  2 "
2 x
= 1/2 exp − dx
π x0 x0 x0
= 1 − erf (1)
= 0.157 .
(5.163)

Eyal Buks Quantum Mechanics - Lecture Notes 121


Chapter 5. The Harmonic Oscillator

10. The answer is [see Eqs. (4.181) and (5.19)]


   
π 2 2 n2x + n2y 1
Enx ,ny ,nz = + ω nz + , (5.164)
2ma2 2

where nx and ny are positive integers and nz is a nonnegative integer.


11. With the help of Eq. (4.14) one has
1 iω0 t  
|α (t) = √ e− 2 |0 + e−iω0 t |1 . (5.165)
2
Moreover, the following hold
0
  
x= a + a† , (5.166)
2mω0
0
mω0  
p=i −a + a† , (5.167)
2

a |n = n |n − 1 , (5.168)

a† |n = n + 1 |n + 1 , (5.169)
, †-
a, a = 1 , (5.170)
thus
a)
0
  
x = α (t)| a + a† |α (t)
2mω 0
0
 1   
= 0| + eiω0 t 1| a + a† |0 + e−iω0 t |1
2mω 0 2
0
 1  iω0 t 
= e + e−iω0 t
2mω 0 2
0

= cos (ω 0 t) .
2mω 0
(5.171)
b)
0
mω 0  
p = i α (t)| −a + a† |α (t)
2
0
mω 0 1    
=i 0| + eiω0 t 1| −a + a† |0 + e−iω0 t |1
2 2
0
mω 0
=− sin (ω 0 t) .
2
(5.172)
c)

Eyal Buks Quantum Mechanics - Lecture Notes 122


5.4. Solutions

) 2*   2
x = α (t)| a + a† |α (t)
2mω 0
   2 , - 
= α (t)| a2 + a† + a, a† + 2a† a |α (t)
2mω 0
 
 1
= 1+2
2mω 0 2

= .
mω 0
(5.173)
d) Similarly
) 2* mω 0  2
p =− α (t)| −a + a† |α (t)
2
mω 0   2 , - 
=− α (t)| a2 + a† − a, a† − 2a† a |α (t)
2
= mω0 ,
(5.174)
thus :
0
cos2 (ω 0 t) sin2 (ω0 t)
∆x∆p =  1 − 1−
2 2
0
 1
= 2 + sin2 (2ω 0 t) .
2 4
(5.175)
12. The state |ψ (t) is given by
    
1 iE0 t iEn t
|ψ (t) = √ exp − |0 + exp − |n , (5.176)
2  

where
 
1
En = ω n + , (5.177)
2

thus, using
0
  
x= a + a† , (5.178)
2mω
and

a |n = n |n − 1 , (5.179)

a† |n = n + 1 |n + 1 , (5.180)
one finds that x (t) = 0 if n > 1, and for n = 1

Eyal Buks Quantum Mechanics - Lecture Notes 123


Chapter 5. The Harmonic Oscillator
0
  
x (t) = ψ (t)| a + a† |ψ (t)
2mω
0

= cos (ωt) .
2mω
(5.181)
13. Since H = ω and |ψ(0) is normalized one has
1
|c0 |2 = |c1 |2 = , (5.182)
2
thus |ψ(0) can be written as
0
1 
|ψ(0) = |0 + eiθ |1 , (5.183)
2
where θ is real. Given that at time t = 0
0
1 
x (t = 0) = , (5.184)
2 mω
one finds using the identities
0
  
x= a + a† , (5.185)
2mω

a |n = n |n − 1 , (5.186)

a† |n = n + 1 |n + 1 , (5.187)
that

2
cos θ = . (5.188)
2
Using this result one can evaluate p (t = 0), where
0
mω  
p=i −a + a† , (5.189)
2
thus
0 0 √
mω mω 2
p (t = 0) = sin θ = ± = ±mωx (t = 0) . (5.190)
2 2 2
Using these results together with Eq. (5.152) yields
0
1 
x (t) = (cos (ωt) ± sin (ωt))
2 mω
0 
 π
= cos ωt ∓ .
2mω 4
(5.191)

Eyal Buks Quantum Mechanics - Lecture Notes 124


5.4. Solutions

14. According to identity (2.175), which states that


1 1
eA+B = eA eB e− 2 [A,B] = eB eA e 2 [A,B] , (5.192)

provided that

[A, [A, B]] = [B, [A, B]] = 0 , (5.193)

one finds with the help of Eq. (5.13) that


 
D (α) = exp αa† − α∗ a
|α|2 † ∗
= e− 2 eαa e−α a
|α|2 ∗ †
=e 2 e−α a eαa .
(5.194)
15. Using Eq. (5.194) one has
|α|2 † ∗ |α|2 ∗ †
D† (α) = e− 2 e−αa eα a
=e 2 eα a e−αa , (5.195)

thus

D† (α) D (α) = D (α) D† (α) = 1 . (5.196)

16. Using Eqs. (5.35), (5.28) and (5.29) one finds that
|α|2 † ∗ |α|2 †
|α = e− eαa e−α a |0 = e−
2 2 eαa |0

|α|2  α
n
= e− 2 √ |n .
n=0 n!
(5.197)
17. Using Eqs. (5.42) and (5.28) one has
∞
|α|2 αn
a |α = e− 2 √ a |n
n=0 n!


|α|2 αn−1
= αe− 2  |n − 1
n=1 (n − 1)!
= α |α .
(5.198)
18. Using Eqs. (5.36), (5.9) and (5.10) one has
0 0 
mω ∗ 1 ∗
D (α) = exp (α − α ) x − i (α + α ) p , (5.199)
2 2mω

thus with the help of Eqs. (2.175) and (5.8) the desired result is obtained

Eyal Buks Quantum Mechanics - Lecture Notes 125


Chapter 5. The Harmonic Oscillator
0 
mω α − α∗
D (α) = exp √ x
 2
   ∗2 
i α + α∗ α − α2
× exp − √ √ p exp .
mω 2 4
(5.200)

19. Using the operator identity (2.173)


1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · , (5.201)
2! 3!
and the definition (5.36)
 
D (α) = exp αa† − α∗ a , (5.202)

one finds that

D† (α) aD (α) = a + α , (5.203)


† † † ∗
D (α) a D (α) = a + α . (5.204)

Exploiting the unitarity of D (α)

D (α) D† (α) = 1
 
it is straightforward to show that for any function f a, a† having a
power series expansion the following holds
   
D† (α) f a, a† D (α) = f a + α, a† + α∗ (5.205)

(e.g., D† a2 D = D† aDD† aD = (a + α)2 ).


20. Using Eq. (5.75) and the following identities
 
† 1
H = ω a a + , (5.206)
2
0
  
x= a + a† , (5.207)
2mω
0
mω  
p=i −a + a† , (5.208)
2
all these relations are easily obtained.
21. Expressing the state |α in the basis of eigenvectors of the Hamiltonian
|n


|α = cn |n , (5.209)
n=0

using

Eyal Buks Quantum Mechanics - Lecture Notes 126


5.4. Solutions

a |α = α |α , (5.210)

and

a |n = n |n − 1 , (5.211)

one finds

 ∞


cn n |n − 1 = α cn |n , (5.212)
n=0 n=0

thus
α
cn+1 = √ cn , (5.213)
n+1
therefore
∞
αn
|α = A √ |n . (5.214)
n=0 n!

The normalization constant A is found by


 n
∞ |α|2 2
1 = |A|2 = |A|2 e|α| . (5.215)
n=0
n!

Choosing A to be real yields


|α|2
A = e− 2 , (5.216)

thus
|α|2 αn
cn = e− 2 √ . (5.217)
n!
Note that this result is identical to Eq. (5.42), thus |α is a coherent
state. The possible results of the measurement are
 
1
En = ω n + , (5.218)
2

and the corresponding probabilities, which are time independent, are


given by
 n
2
2
|α|
pn (t) = |cn |2 = e−|α| . (5.219)
n!

Eyal Buks Quantum Mechanics - Lecture Notes 127


Chapter 5. The Harmonic Oscillator

22. Using the relations


0
2
xα = Re (α) , (5.220)


pα = 2mω Im (α) , (5.221)
Eq. (5.74) can be written as
     ∗2 
i pα x i xα p α − α2
D (α) = exp exp − exp . (5.222)
  4
Using Eqs. (3.12) and (3.19) one finds that
 
i xα p
exp − |x′  = |x′ + xα  ,

thus
     ∗2 
′ ′ i pα x i xα p α − α2
x |α = x | exp exp − exp |0
  4
 ∗2   
α − α2 i pα x′
= exp exp x′ − xα |0 .
4 
(5.223)
Using Eq. (5.120) the wavefunction of the ground state is given by
!  2 "
′ 1 1 x′
x |0 = √ exp − , (5.224)
(2π)1/4 ∆xα 2∆xα

where
0

∆xα = , (5.225)
2mω
thus
  2 
x′ −xα
    exp −
α∗2 − α2 i pα x′ 2∆xα
x′ |α = exp exp 1/4 √
4  (2π) ∆xα
 ∗2    2 
2   x − xα
α −α mω 1/4 x
= exp exp − + i pα .
4 π 2∆xα 
(5.226)
23. Using Eqs. (5.36) and (2.175) this relation is easily obtained.
24. With the help of Eq. (5.42) one has
   
1 1 1 2
|α α| d2 α = |n m| √ e−|α| αn α∗m d2 α .
π π n,m n!m!

Eyal Buks Quantum Mechanics - Lecture Notes 128


5.4. Solutions

(5.227)

Employing polar coordinates in the complex plane α = ρeiθ , where ρ is


non-negative real and θ is real, leads to
   ∞  2π
1 2 1 1 2
|α α| d α = |n m| √ dρρn+m+1 e−ρ dθeiθ(n−m)
π π n,m n!m!
0 0
  
2πδnm

 ∞
2 2
= |n n| dρρ2n+1 e−ρ
n
n!
0
 1
= |n n| Γ (n + 1)
n
n!   
=n!

= |n n|
n
=1.
(5.228)
25. Using Eqs. (5.35) and (5.41) one finds that
β |α = 0| D† (β) D (α) |0
= 0| D (−β) D (α) |0
 
−βα∗ + β ∗ α
= exp 0| D (−β + α) |0
2
 
−βα∗ + β ∗ α
= exp 0 |α − β .
2
(5.229)
Thus, with the help of Eq. (5.42) one has
 
−βα∗ + β ∗ α − |α−β| 2
β |α = exp e 2
2
! "
2
|α| |β|2 ∗
= exp − − + αβ
2 2
! "
|α − β|2 ∗
= exp − + i Im (αβ ) .
2
(5.230)
26. The following holds

Eyal Buks Quantum Mechanics - Lecture Notes 129


Chapter 5. The Harmonic Oscillator

1 1
V1 (x) = mω2 (x + β)2 − mω 2 β 2
2 2
1 1
= mω2 x′2 − mω 2 β 2 ,
2 2
(5.231)
where

x′ = x + β . (5.232)

a) Thus, using Eqs. (5.152) and (5.153) together with the relations
x′(H) (t) = x(H) (t) + β , (5.233)
(H) ′(H)
p (t) = p (t) , (5.234)
one finds
  sin (ωt) (H)
x(H) (t) = x(H) (0) + β cos (ωt) + p (0) − β , (5.235)

 
p(H) (t) = p(H) (0) cos (ωt) − mω sin (ωt) x(H) (0) + β . (5.236)

b) For '
this case (at time t = 0
x(H) (0) = 0 , (5.237)
' (
p(H) (0) = 0 , (5.238)
thus
' (
x(H) (t) = β (cos (ωt) − 1) . (5.239)

27. The state of the system at time t = 0 is given by


 
i∆x
|ψ (t = 0) = exp − p |0 , (5.240)

where |0 is the ground state of the potential V2 . In general a coherent
state with parameter α can be written as
0     ∗2 
mω α − α∗ i α + α∗ α − α2
|α = exp √ x exp − √ √ p exp |0 .
 2 mω 2 4
(5.241)

a) Thus |ψ (t = 0) = |α0 , where


0

α 0 = ∆x . (5.242)
2
The time evolution of a coherent state is given by

*
|ψ (t) = e−iωt/2
α = α0 e−iωt , (5.243)

Eyal Buks Quantum Mechanics - Lecture Notes 130


5.4. Solutions

and the following holds


0
2 , -
x (t) = Re α0 e−iωt = ∆x cos (ωt) , (5.244)

b) According to Eq. (5.48)
' ( 
2
(∆x) (t) = . (5.245)
2mω
c) In general a coherent state can be expanded in the basis of number
states |n
2  αn
|α = e−|α| /2
√ |n , (5.246)
n n!

thus the probability to measure energy En = ω (N + 1/2) at time t


is given by
  n
e−|α0 | α2n
2
1 mω∆2x mω∆2x
Pn = |n|ψ (t)|2 = 0
= exp − .
n! n! 2 2
(5.247)

28. At time t = 0 the following holds


x = 0 , (5.248)
p = 0 , (5.249)
' ( ) * 
(∆x)2 = x2 = , (5.250)
2mω
' ( ) * mω
(∆p)2 = p2 = . (5.251)
2
Moreover, to calculate xp it is convenient to use
0
  
x= a + a† , (5.252)
2mω
0
mω  
p=i −a + a† , (5.253)
2
, †-
a, a = 1 , (5.254)
thus at time t = 0
 
xp = i 0| aa† − a† a |0 = i . (5.255)
2 2
The Hamiltonian for times t > 0 is given by
p2
H= + gx . (5.256)
2m

Eyal Buks Quantum Mechanics - Lecture Notes 131


Chapter 5. The Harmonic Oscillator

Using the Heisenberg equation of motion for the operators x and x2 one
finds
dx(H) 1 , -
= x(H) , H , (5.257)
dt i
dp(H) 1 , -
= p(H) , H , (5.258)
dt i
dx2(H) 1 3 2 4
= x(H) , H , (5.259)
dt i
or using [x, p] = i
dx(H) p(H)
= , (5.260)
dt m
dp(H)
= −g , (5.261)
dt
dx2(H) 1   1  
= x p + p(H) x(H) = 2x(H) p(H) − i , (5.262)
dt m (H) (H) m
thus

p(H) (t) = p(H) (0) − gt , (5.263)

p(H) (0) t gt2


x(H) (t) = x(H) (0) + − , (5.264)
m 2m

it 2 t
x2(H) (t) = x2(H) (0) − + x (t′ ) p(H) (t′ ) dt′
m m 0 (H)
  
it 2 t p(H) (0) t′ gt′2 , -
= x2(H) (0) − + x(H) (0) + − p(H) (0) − gt′ dt′
m m 0 m 2m
it
= x2(H) (0) −
m
 t! p2(H) (0) t′ gt′2 ′2 2 ′3
"
2 p(H) (0) gt g t
+ x(H) (0) p(H) (0) + − p (0) − x(H) (0) gt′ − + dt′
m 0 m 2m (H) m 2m
it
= x2(H) (0) −
! m "
2 p2(H) (0) t2 p(H) (0) gt3 x(H) (0) gt2 p(H) (0) gt3 g2 t4
+ x(H) (0) p(H) (0) t + − − − + .
m 2m 6m 2 3m 8m
(5.265)
Using the initial conditions Eqs. (5.248), (5.249), (5.250), (5.251) and
(5.255) one finds

gt2
x (t) = − , (5.266)
2m

Eyal Buks Quantum Mechanics - Lecture Notes 132


5.4. Solutions

2 g 2 t4
x (t) = , (5.267)
4m2
p (t) = −gt , (5.268)
 
) 2 *  it 2 it ωt2 g2 t4
x (t) = − + + + , (5.269)
2mω m m 2 4 8m
and
' ( ) *  ωt2   
(∆x)2 (t) = x2 (t) − x (t)2 = + = 1 + ω 2 t2 .
2mω 2m 2mω
(5.270)

29. Using the operator identity (2.173), which is given by


1 1
eL Oe−L = O + [L, O] + [L, [L, O]] + [L, [L, [L, O]]] + · · · , (5.271)
2! 3!
for the operators
O=a, (5.272)
r  † 2 
L= a − a2 , (5.273)
2
and the relations
, †-
a, a = 1 , (5.274)

[L, O] = −ra , (5.275)
[L, [L, O]] = r2 a , (5.276)
[L, [L, [L, O]]] = −r3 a† , (5.277)
[L, [L, [L, [L, O]]]] = r4 a , (5.278)
etc., one finds
   
r 2 r4 r3
T = 1+ + +··· a− r + + · · · a† + · · · , (5.279)
2! 4! 3!
a) Thus

T = Aa + Ba† , (5.280)

where
A = cosh r , (5.281)
B = − sinh r . (5.282)
b) Using the0relations
  
x= a + a† , (5.283)
2mω
0
mω  
p=i −a + a† . (5.284)
2

Eyal Buks Quantum Mechanics - Lecture Notes 133


Chapter 5. The Harmonic Oscillator

one finds 0
  
r| x |r = 0| S (r) a + a† S † (r) |0
2mω
0
  
= 0| T |0 + 0| T † |0
2mω
=0,
0 (5.285)
mω  
r| p |r = i 0| S (r) −a + a† S † (r) |0
2
0
  
= − 0| T |0 + 0| T † |0
2mω
=0.
(5.286)
c) Note that S (r) is unitary, namely S † (r) S (r) = 1, since the operator
 2
a2 − a† is anti Hermitian. Thus
   
r| x2 |r = 0| S (r) a + a† a + a† S † (r) |0
2mω
    
= 0| S (r) a + a† S † (r) S (r) a + a† S † (r) |0
2mω
  2
= 0| T + T † |0
2mω
 (A + B)2  2
= 0| a + a† |0
2mω
 (cosh r − sinh r)2
=
2mω
e−2r
= ,
2mω
(5.287)
and
mω  2
r| p2 |r = 0| S (r) a − a† S † (r) |0
2
mω  2
= 0| T − T † |0
2
mω (A − B)2  2
= 0| a − a† |0
2
mω (cosh r + sinh r)2
=
2
mωe2r
= .
2
(5.288)
Thus

Eyal Buks Quantum Mechanics - Lecture Notes 134


5.4. Solutions

e−2r
2
(∆x) = , (5.289)
2mω
2 mωe2r
(∆p) = , (5.290)
2

(∆x) (∆p) = . (5.291)
2
30. With the help of Eqs. (5.9) and (5.10) one finds that
0
−r   
x=e Q (r) + Q† (r) , (5.292)
2mω
0
r mω  
p = −ie Q (r) − Q† (r) , (5.293)
2
thus with the help of Eq. (5.86) one finds that
r| x |r = 0 , (5.294)
r| p |r = 0 . (5.295)
Using the commutation relation
, -  , -
Q (r) , Q† (r) = cosh2 r − sinh2 r a, a† = 1 ,

one obtains
e−2r  2
r| x2 |r = r| Q (r) + Q† (r) |r
2mω
e−2r
= r| Q (r) Q† (r) |r
2mω
e−2r
= ,
2mω
(5.296)
and similarly
mωe2r  2
r| p2 |r = − r| Q (r) − Q† (r) |r
2
mωe2r
= r| Q (r) Q† (r) |r
2
mωe2r
= ,
2
(5.297)
thus
e−2r
(∆x)2 = , (5.298)
2mω
mωe2r
(∆p)2 = . (5.299)
2

Eyal Buks Quantum Mechanics - Lecture Notes 135


Chapter 5. The Harmonic Oscillator

31. Using the commutation relation


, †-
a, a = 1 , (5.300)

one finds
 
H = ω 0 N + ω 1 N 2 − N , (5.301)

where

N = a† a (5.302)

is the number operator.


a) The eigenvectors of N

N |n = n |n , (5.303)

(where n = 0, 1, · · · ) are also eigenvectors of H and the following


holds

H |n = En |n , (5.304)

where
,  -
En =  ω0 n + ω1 n2 − n . (5.305)

Note that
En+1 − En
= ω 0 + 2ω 1 n , (5.306)

thus En+1 > En .
b) Using the relations
0
  † 
x= a +a , (5.307)
2mω 0
0
mω 0  † 
p=i a −a , (5.308)
2
  † † 
x2 = a a + aa + 2N + 1 , (5.309)
2mω 0
2 mω0  † † 
p = −a a − aa + 2N + 1 , (5.310)
√2
a |n = n |n − 1 , (5.311)

a† |n = n + 1 |n + 1 , (5.312)
one finds
i. 0|x|0 = 0
ii. 0|p|0 = 0

Eyal Buks Quantum Mechanics - Lecture Notes 136


5.4. Solutions
' (
2 
iii.0| (∆x) |0 = 2mω0
' (
iv. 0| (∆p)2 |0 = mω0
2
32. The proof of the clue is:
 
N 2 = b† bb† b = b† 1 − b† b b = N . (5.313)

Moreover, N is Hermitian, thus N is a projector.


a) Let |n be the eigenvectors of N and n the corresponding real eigen-
values (N is Hermitian)

N |n = n |n . (5.314)

Using the clue one finds that n2 = n, thus the possible values of n
are 0 (ground state) and 1 (excited state). Thus, the eigenvalues of
H are 0 and ǫ.
b) To verify the statement in the clue we calculate

N b† |0 = b† bb† |0 = b† (1 − N ) |0 = b† |0 , (5.315)

thus the state b† |0 is indeed an eigenvector of N with eigenvalue 1


(excited state). In what follows we use the notation

|1 = b† |0 . (5.316)

Note that |1 is normalized since

1|1 = 0| bb† |0 = 0| (1 − N ) |0 = 0|0 = 1 . (5.317)

Moreover, since |0 and |1 are eigenvectors of an Hermitian operator


with different eigenvalues they must be orthogonal to each other

0|1 = 0 . (5.318)

Using Eqs. (5.316), (5.317) and (5.318) one finds


+|+ = 2 |A+ |2 , (5.319)
2
−|− = 2 |A− | . (5.320)
choosing the normalization constants to be non-negative real num-
bers lead to
1
A+ = A− = √ . (5.321)
2

c) Using N 2 = N one finds

Eyal Buks Quantum Mechanics - Lecture Notes 137


Chapter 5. The Harmonic Oscillator
  ∞  n
iHt 1 iHt
exp − = 1+ −
 n=1
n! 
∞  n
1 iǫt
= 1+N −
n=1
n! 
! "
 1  iǫt n

= 1 + N −1 + −
n=0
n! 
  
iǫt
= 1 + N −1 + exp − .

(5.322)
Thus
 
2

iHt

p0 (t) =
−| exp − |+



  
2
1
iǫt

(0| − 1|) 1 + N −1 + exp − (|0 + |1)

4 

 
2
1
iǫt

1 − exp −
4 

 
ǫt
= sin2 .
2
(5.323)
33. The closure relation (5.31) can be written as


1= |n m| δ n,m . (5.324)
n,m=0

With the help of Eq. (5.32) together with the relation


 n

1 d

ς m

= δ n,m , (5.325)
n! dς ς=0

which is obtained using the Taylor power expansion series of the function
ς m , one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 138


5.4. Solutions


1= |n m| δ n,m
n,m=0
∞  n

1 d

= √ √ |n m| ς m

n,m=0 n! m! dς ς=0
∞  †
n  n

 a a m
d

= |0 0| ς m

n,m=0
n! m! dς ς=0
! ∞  n  n " ! ∞ "

 a† d  m m

dς a ς

= |0 0|

n=0
n! m=0
m!

ς=0
 

= exp a† |0 0| exp (aς)

.
dς ς=0
(5.326)

Denote the normal ordering representation of the operator |0 0| by Z,


i.e.

|0 0| = : Z : . (5.327)

For general functions f , g and h of the operators a and a† it is easy to


show that the following holds

: f g : =: gf : , (5.328)

: f gh : =: f hg : , (5.329)
and

: f (: g : ) : = : f g : . (5.330)

Thus

Eyal Buks Quantum Mechanics - Lecture Notes 139


Chapter 5. The Harmonic Oscillator
 

† d

1= exp a : Z : exp (aς)


 
ς=0
† d

= : exp a Z exp (aς)

:
dς ς=0
 

† d

= : exp a exp (aς) Z

:
dς ς=0
 d n

 a† dς m

(aς)

= : Z
:
n,m
n! m!

ς=0
 d n

  n dς
am
= : a† ς m
Z:
n,m
n!
m!
ς=0
  
δn,m
 
= : exp a† a Z :
 
= : exp a† a ( : Z : ) : ,
(5.331)
and therefore
 
|0 0| = : exp −a† a : . (5.332)

Using again Eq. (5.32) one finds that


1  † n    n
Pn = |n n| = : a exp −a† a a† : . (5.333)
n!
34. The Hamiltonian H, which is given by
p2 mω2 x2
H= + + xf (t) , (5.334)
2m 2
can be expressed in terms of the annihilation a and creation a† operators
[see Eqs. (5.11) and (5.12)] as
  0
† 1   
H = ω a a + + f (t) a + a† . (5.335)
2 2mω
The Heisenberg equation of motion for the operator a is given by [see
Eq. (4.37)]
0
da 1
= −iωa − i f (t) . (5.336)
dt 2mω
The solution of this first order differential equation is given by
0  t
1 ′
a (t) = e−iω(t−t0 )
a (t0 ) − i dt′ e−iω(t−t ) f (t′ ) , (5.337)
2mω t0

Eyal Buks Quantum Mechanics - Lecture Notes 140


5.4. Solutions

where the initial time t0 will be taken below to be −∞. The Heisenberg
operator a† (t) is found from the Hermitian conjugate of the last result.
Let Pn (t) be the Heisenberg representation of the projector |n n|. The
probability pn (t) to find the oscillator in the number state |n at time t
is given by

pn (t) = 0| Pn (t) |0 . (5.338)

To evaluate pn (t) it is convenient to employ the normal ordering repre-


sentation of the operator Pn (5.100). In normal ordering the first term of
Eq. (5.337), which is proportional to a (t0 ) does not contribute to pn (t)
since a (t0 ) |0 = 0 and also 0| a† (t0 ) = 0. To evaluate pn = pn (t → ∞)
the integral in the second term of Eq. (5.337) is evaluate from t0 = −∞
to t = +∞. Thus one finds that
e−µ µn
pn = , (5.339)
n!
where


2
1

∞ ′ iωt′

µ=
dt e f (t )
. (5.340)
2mω −∞

35. As can be seen from the definition of P, the following holds


∞

x | P |ψ = dx′′ x′ |x′′  −x′′ |ψ
−∞
= −x′ |ψ ,
(5.341)
thus the wave function of P |ψ is ψ (−x′ ) given that the wave function
of |ψ is ψ (x′ ). For the wavefunctions ψn (x′ ) = x′ |n of the number
states |n, which satisfy N |n = n |n, the following holds
+
′ −ψ n (x′ ) n odd
ψn (−x ) = , (5.342)
ψn (x′ ) n even

thus
+
− |n n odd
P |n = , (5.343)
|n n even

or P |n = (−1)n |n ,and consequently the parity operator P can be


expressed as a function of N

P = eiπN . (5.344)

Eyal Buks Quantum Mechanics - Lecture Notes 141


Chapter 5. The Harmonic Oscillator

36. Using Eqs. (5.31), (5.32) and (5.332) together with the relation

a† a |n = n |n , (5.345)

yields



eλa a
= eλn |n n|
n=0
∞  † n
 a an
λn
= e √ |0 0| √
n=0 n! n!
∞
eλn  † n  
= a : exp −a† a : an
n=0
n!

 eλn  n  
= : a† exp −a† a an :
n=0
n!
∞
eλn  † n  
=: a a exp −a† a :
n=0
n!
   
=: exp eλ a† a exp −a† a : ,
(5.346)

thus
† , λ  -
eλa a
=: exp e − 1 a† a : . (5.347)

37. The following holds [see Eq. (5.31)]




|x′  = |n n |x′  , (5.348)
n=0

where
 † n
a
|n = √ |0 , (5.349)
n!
thus with the help of Eq. (5.123) and the generating function of Hermite
polynomials (5.56) one finds that (note that x′ |n is real)
 ′2   ′
exp − 2x x ∞ H x  † n
2  n x0 a
′ 0
|x  = 1/2
√ √ |0
π 1/4 x0 n=0 2n n! n!
 ′2 √ ′ †2

x
exp − 2x 2 + 2 xx0 a† − a2
0
= 1/2
|0 .
π1/4 x0
(5.350)

Eyal Buks Quantum Mechanics - Lecture Notes 142


5.4. Solutions

38. Using the relation x |x′  = x′ |x′  and Eq. (3.32) one finds that
∞
  ′2
exp kx2 = dx′ ekx |x′  x′ | . (5.351)
−∞

Eqs. (5.332) and (5.103) yield


1 ′ 2
|x′  x′ | = √ : e−(X −X ) : , (5.352)
πx0

where
a + a† x
X= √ = , (5.353)
2 x0

and where
x′
X′ = . (5.354)
x0
Thus
∞
  1 ′ 2
−X ) +KX ′2
exp kx2 = √ dX ′ : e−(X :
π
−∞
∞
1 ′2
+2X ′ X−X 2
=√ dX ′ : e−(1−K)X : ,
π
−∞

(5.355)

where K = kx20 . With the help of the identity (5.136) this becomes
   
Kx2 1 K x2
exp =√ : exp : . (5.356)
x20 1−K 1 − K x20

Using the notation


K
κ= , (5.357)
1−K
the results can be also expressed as
   2
1 κ x2 x
√ exp 2 =: exp κ 2 : . (5.358)
1+κ 1 + κ x0 x0

39. The orthogonality between number states yields according to Eq. (5.123)

Eyal Buks Quantum Mechanics - Lecture Notes 143


Chapter 5. The Harmonic Oscillator
 ′2   ′  ′
∞ exp − xx2 Hm xx0 Hn xx0
m |n = dx′ √0
π2m m!2n n!x0
−∞
∞  
′ exp −X ′2 Hm (X ′ ) Hn (X ′ )
= dX √
π2m m!2n n!
−∞
= δ nm .
(5.359)
2
Multiplying Eq. (5.56) by the factor e−z Hm (z)


2 2 tn
e−(z−t) Hm (z) = e−z Hm (z) Hn (z) , (5.360)
n=0
n!

and integrating over z


∞
2
dz e−(z−t) Hm (z)
−∞
∞ n ∞
 t 2
= dz e−z Hm (z) Hn (z) ,
n=0
n!
−∞

(5.361)

yields with the help of Eq. (5.359)


∞
2 √
dz e−(z−t) Hm (z) = (2t)m π. (5.362)
−∞
  √
The relations x |x′  = x′ |x′ , X = a + a† / 2 = x/x0 together with
Eq. (5.352) yield
∞  
′ x′
Hn (X) = dx Hn |x′  x′ |
x0
−∞
∞
1 ′ 2
=√ : dX ′ e−(X −X )
Hn (X ′ ) : ,
π
−∞

(5.363)

thus, with the help of Eq. (5.362) one finds that

Hn (X) =: (2X)n : . (5.364)

Eyal Buks Quantum Mechanics - Lecture Notes 144


5.4. Solutions

The last result together with the identity


dHn
= 2nHn−1 (X ′ ) , (5.365)
dX ′
yields
d 1 dHn (X)
: Xn : = n
dX 2 dX
Hn−1 (X)
=n
2n−1
= n : X n−1 : ,
(5.366)
thus
d d n
: X n : =: X : . (5.367)
dX dX
Thus, for a general smooth function F (X) of the operator X the following
holds
d dF
: F (X) : =: : . (5.368)
dX dX
40. Initially, the system is in a coherent state given by Eq. (5.42)
∞
|α|2 αn
|ψ (t = 0) = |αc = e− 2 √ |n . (5.369)
n=0 n!

The notation |αc is used to label coherent states satisfying a |αc =


α |αc .
 k
a) Since a† a commutes with a† a , the time evolution operator is given
by [see Eq. (4.9)]
 
iHt † k
= e−iω1 (a a) t e−iωa at ,

u (t) = exp − (5.370)

thus
|ψ (t) = u (t) |ψ (t = 0)
∞
† k |α| 2
αn
= e−iω1 (a a) t e−iωa at e− 2

√ |n
n=0 n!
∞  −iωt n
† k |α|2  αe
= e−iω1 (a a) t e− 2 √ |n
n=0 n!
∞  
−iωt n
|α|2  αe
= e− 2 √ e−iφn |n ,
n=0 n!
(5.371)

Eyal Buks Quantum Mechanics - Lecture Notes 145


Chapter 5. The Harmonic Oscillator

where

φn = ω 1 tnk . (5.372)

b) At time t = 2π/ω 1 the phase factor φn is given by φn = 2πnk , thus

e−iφn = 1 , (5.373)

and therefore

 6
(


ψ 2π
2πiω


=
αe− ω1 . (5.374)
ω1 c

c) At time t = π/ω1 the phase factor φn is given by φn = πnk . Using


the fact that
+
  0 n is even
mod nk , 2 = , (5.375)
1 n is odd
one has
n
e−iφn = (−1) , (5.376)

and therefore  n

 6 − πiω

 ∞ αe ω1


ψ π
|α|2


= e− 2 √ (−1)n |n
ω1 n=0 n!

(

− πiω
=
−αe ω1 .
c
(5.377)
d) At time t = π/2ω 1 the phase factor φn is given by φn = (π/2) nk .
For the case where k is even one has
+
  0 n is even
mod nk , 4 = , (5.378)
1 n is odd
thus
+
−iφn 1 n is even
e = , (5.379)
−i n is odd
and therefore
 πiω
n

 6 2 
∞ αe− 2ω
1

π |α|

ψ

= e− 2 √ e−iφn |n . (5.380)
2ω 1 n=0 n!

This state can be expressed as a superposition of two coherent states



 6


ψ π 1  − iπ
− πiω
( iπ
− πiω
(

= √ e 4
αe 2ω 1 + e 4
−αe 2ω 1 . (5.381)
2ω 1 2 c c

Eyal Buks Quantum Mechanics - Lecture Notes 146


5.4. Solutions

41. Let {λn } be the set of eigenvalues of A. Clearly A is Hermitian, namely


A† = A, thus the eigenvalues λn are expected to be real. Since the trace
of an operator is basis independent, the following must hold

Tr (A) = λn , (5.382)
n

and
   2
Tr A2 = λn . (5.383)
n

On the other hand, with the help of Eq. (2.172) one finds that

Tr (A) = Tr (|α α|) − Tr (|β β|) = 0 , (5.384)

and
 
Tr A2 = Tr (|α α |α α|) + Tr (|β β |β β|)
− Tr (|α α |β β|) − Tr (|β β |α α|)
= 2 − α |β Tr (|α β|) − β |α Tr (|β α|)
 
= 2 1 − |α |β|2 .
(5.385)
Clearly, A cannot have more than two nonzero eigenvalues, since the
dimensionality of the subspace spanned by the vectors {|α , |β} is at
most 2, and therefore A has three eigenvalues 0, λ+ and λ− , where [see
Eq. (5.230)]
. 
λ± = ± 1 − |α |β|2 = ± 1 − e−|α−β| .
2
(5.386)

Eyal Buks Quantum Mechanics - Lecture Notes 147


6. Angular Momentum

Consider a point particle moving in three dimensional space. The orbital


angular momentum L is given by
 
x̂ ŷ ẑ
L = r × p = det  x y z  ,
px py pz
where r = (x, y, z) is the position vector and where p = (px , py , pz ) is the
momentum vector. In classical physics the following holds:
Claim.
{Li , Lj } = εijk Lk , (6.1)
where

 0 i, j, k are not all different
εijk = 1 i, j, k is an even permutation of x, y, z . (6.2)

−1 i, j, k is an odd permutation of x, y, z
Proof. Clearly, Eq. (6.1) holds for the case i = j. Using Eq. (1.48), which
reads
{xi , pj } = δ ij , (6.3)
one has
{Lx , Ly } = {ypz − zpy , zpx − xpz }
= {ypz , zpx } + {zpy , xpz }
= y {pz , z} px + x {z, pz } py
= −ypx + xpy
= Lz .
(6.4)
In a similar way one finds that {Ly , Lz } = Lx and {Lz , Lx } = Ly . These
results together with Eq. (1.49) complete the proof.
Using the rule (4.41) {, } → (1/i) [, ] one concludes that in quantum
mechanics the following holds:
[Li , Lj ] = iεijk Lk . (6.5)
Chapter 6. Angular Momentum

6.1 Angular Momentum and Rotation


We have seen before that the unitary operator u (t, t0 ) is the generator of time
evolution [see Eq. (4.4)]. Similarly, we have seen that the unitary operator
 
i∆ · p
J (∆) = exp − (6.6)

[see Eq. (3.73)] is the generator of linear translations:

J (∆) |r′  = |r′ + ∆ . (6.7)

Below we will see that one can define a unitary operator that generates ro-
tations.
Exercise 6.1.1. Show that
   
x x
†  
Dẑ (φ) y Dẑ (φ) = Rẑ y  ,
 (6.8)
z z
where
 
iφLz
Dẑ (φ) = exp − , (6.9)

and where
 
cos φ − sin φ 0
Rẑ =  sin φ cos φ 0  . (6.10)
0 0 1

Solution 6.1.1. Equation (6.8) is made of 3 identities:


Dẑ† (φ) xDẑ (φ) = x cos φ − y sin φ , (6.11)
Dẑ† (φ) yDẑ (φ) = x sin φ + y cos φ , (6.12)
Dẑ† (φ) zDẑ (φ) = z . (6.13)
As an example, we prove below the first one. Using the identity (2.173), which
is given by
1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · , (6.14)
2! 3!
one has
Dẑ† (φ) xDẑ (φ)
 2  3
iφ 1 iφ 1 iφ
= x+ [Lz , x] + [Lz , [Lz , x]] + [Lz , [Lz , [Lz , x]]] + · · · .
 2!  3! 
(6.15)

Eyal Buks Quantum Mechanics - Lecture Notes 150


6.1. Angular Momentum and Rotation

Furthermore with the help of


Lz = xpy − ypx , (6.16)
[xi , pj ] = iδ ij , (6.17)
one finds that
[Lz , x] = −y [px , x] = iy ,
2
[Lz , [Lz , x]] = ix [py , y] = − (i) x ,
[Lz , [Lz , [Lz , x]]] = − (i)2 [Lz , x] = − (i)3 y ,
4
[Lz , [Lz , [Lz , [Lz , x]]]] = (i) x ,
..
. (6.18)
thus
   
φ2 φ4 φ3
Dẑ† (φ) xDẑ (φ) = x 1 − + +··· −y φ− + ···
2! 4! 3!
= x cos φ − y sin φ .
(6.19)
The other 2 identities in Eq. (6.8) can be proven in a similar way.

The matrix Rẑ [see Eq. (6.10)] represents a geometrical rotation around
the z axis with angle φ. Therefore, in view of the above result, we refer to the
operator Dẑ (φ) as the generator of rotation around the z axis with angle φ.
It is straightforward to generalize the above results and to show that rotation
around an arbitrary unit vector n̂ axis with angle φ is given by
 
iφL · n̂
Dn̂ (φ) = exp − . (6.20)


In view of Eq. (3.73), it can be said that linear momentum p generates


translations. Similarly, in view of the above equation (6.20), angular momen-
tum L generates rotation. However, there is an important distinction between
these two types of geometrical transformations. On one hand, according to
Eq. (3.7) the observables px , py and pz commute with each other, and con-
sequently translation operators with different translation vectors commute

[J (∆1 ) , J (∆2 )] = 0 . (6.21)

On the other hand, as can be seen from Eq. (6.5), different components of L do
not commute and therefore rotation operators Dn̂ (φ) with different rotations
axes n̂ need not commute. Both the above results, which were obtained from
commutation relations between quantum operators, demonstrate two well
known geometrical facts: (i) different linear translations commute, whereas
(ii) generally, different rotations do not commute.

Eyal Buks Quantum Mechanics - Lecture Notes 151


Chapter 6. Angular Momentum

6.2 General Angular Momentum


Elementary particles carry angular momentum in two different forms. The
first one is the above discussed orbital angular momentum, which is com-
monly labeled as L. This contribution L = r × p has a classical analogue,
which was employed above to derive the commutation relations (6.5) from
the corresponding Poisson’s brackets relations. The other form of angular
momentum is spin, which is commonly labeled as S. Contrary to the orbital
angular momentum, the spin does not have any classical analogue. In a gen-
eral discussion on angular momentum in quantum mechanics the label J is
commonly employed.
L - orbital angular momentum
S - spin angular momentum
J - general angular momentum
In the discussion below we derive some properties of angular momentum
in quantum mechanics, where our only assumption is that the components of
the angular momentum vector of operators J = (Jx , Jy , Jz ) obey the following
commutation relations
[Ji , Jj ] = iεijk Jk . (6.22)
Namely, we assume that Eq. (6.5), which was obtained from the corresponding
Poisson’s brackets relations for the case of orbital angular momentum holds
for general angular momentum.

6.3 Simultaneous Diagonalization of J2 and Jz


As we have seen in chapter 2, commuting operators can be simultaneously
diagonalized. In this section we seek such simultaneous diagonalization of the
operators J2 and Jz , where
J2 = Jx2 + Jy2 + Jz2 . (6.23)
As is shown by the claim below, these operators commute.
Claim. The following holds
, 2 - , - , -
J , Jx = J2 , Jy = J2 , Jz = 0 . (6.24)
Proof. Using Eq. (6.22) one finds that
, 2 - , 2 - , 2 -
J , Jz = Jx , Jz + Jy , Jz
= i (−Jx Jy − Jy Jx + Jy Jx + Jx Jy ) = 0 .
(6.25)
, - , -
In a similar way one can show that J2 , Jx = J2 , Jy = 0.

Eyal Buks Quantum Mechanics - Lecture Notes 152


6.3. Simultaneous Diagonalization of J2 and Jz

The common eigenvectors of the operators J2 and Jz are labeled as |a, b,
and the corresponding eigenvalues are labeled as a2 and b respectively
J2 |a, b = a2 |a, b , (6.26)
Jz |a, b = b |a, b . (6.27)
Recall that we have shown in chapter 5 for the case of harmonic oscillator
that the ket-vectors a |n and a† |n are eigenvectors of the number operator
N provided that |n is an eigenvector of N. Somewhat similar claim can
be made regrading the current problem under consideration of simultaneous
diagonalization of J2 and Jz :

Theorem 6.3.1. Let |a, b be a normalized simultaneous eigenvector of the


operators J2 and Jz with eigenvalues 2 a and b respectively, i.e.
J2 |a, b = a2 |a, b , (6.28)
Jz |a, b = b |a, b , (6.29)
a, b |a, b = 1 . (6.30)
Then (i) the vector

|a, b + 1 ≡ −1 [a − b (b + 1)]−1/2 J+ |a, b (6.31)

where

J+ = Jx + iJy , (6.32)

is a normalized simultaneous eigenvector of the operators J2 and Jz with


eigenvalues 2 a and  (b + 1) respectively, namely
J2 |a, b + 1 = a2 |a, b + 1 , (6.33)
Jz |a, b + 1 = (b + 1)  |a, b + 1 . (6.34)
(ii) The vector

|a, b − 1 ≡ −1 [a − b (b − 1)]−1/2 J− |a, b (6.35)

where

J− = Jx − iJy , (6.36)

is a normalized simultaneous eigenvector of the operators J2 and Jz with


eigenvalues 2 a and  (b − 1) respectively, namely
J2 |a, b − 1 = a2 |a, b − 1 , (6.37)
Jz |a, b − 1 = (b − 1)  |a, b − 1 . (6.38)

Eyal Buks Quantum Mechanics - Lecture Notes 153


Chapter 6. Angular Momentum

Proof. The following holds


 
, - 
J2 (J± |a, b) =  J2 , J± + J± J2  |a, b = a2 (J± |a, b) . (6.39)
  
0

Similarly

Jz (J± |a, b) = ([Jz , J± ] + J± Jz ) |a, b , (6.40)

where

[Jz , J± ] = [Jz , Jx ± iJy ] =  (iJy ± Jx ) = ±J± , (6.41)

thus

Jz (J± |a, b) = (b ± 1)  (J± |a, b) . (6.42)

Using the following relations



J+ J+ = J− J+
= (Jx − iJy ) (Jx + iJy )
= Jx2 + Jy2 + i [Jx , Jy ]
= J2 − Jz2 − Jz ,
(6.43)


J− J− = J+ J−
= (Jx + iJy ) (Jx − iJy )
= Jx2 + Jy2 + i [Jy , Jx ]
= J2 − Jz2 + Jz ,
(6.44)
one finds that

a, b| J+ J+ |a, b = a, b| J2 |a, b − a, b| Jz (Jz + ) |a, b
= 2 [a − b (b + 1)] ,
(6.45)
and

a, b| J− J− |a, b = a, b| J2 |a, b − a, b| Jz (Jz − ) |a, b
= 2 [a − b (b − 1)] .
(6.46)
Thus the states |a, b + 1 and |a, b − 1 are both normalized.

Eyal Buks Quantum Mechanics - Lecture Notes 154


6.3. Simultaneous Diagonalization of J2 and Jz

What are the possible values of b? Recall that we have shown in chapter 5
for the case of harmonic oscillator that the eigenvalues of the number operator
N must be nonnegative since the operator N is positive-definite. Below we
employ a similar approach to show that:

Claim. b2 ≤ a
Proof. Both Jx2 and Jy2 are positive-definite, therefore

a, b| Jx2 + Jy2 |a, b ≥ 0 . (6.47)

On the other hand, Jx2 + Jy2 = J2 − Jz2 , therefore a − b2 ≥ 0.

As we did in chapter 5 for the case of the possible eigenvalues n of the


number operator N , also in the present case the requirement b2 ≤ a restricts
the possible values that b can take:
Claim. For a given value of a the possible values of b are {−bmax , −bmax + 1, · · · , bmax − 1, bmax }
where a = bmax (bmax + 1). Moreover, the possible values of bmax are 0, 1/2, 1, 3/2, 2, · · · .

Proof. There must be a maximum value bmax for which

J+ |a, bmax  = 0 . (6.48)

Thus, also

J+ J+ |a, bmax  = 0 (6.49)

holds. With the help of Eq. (6.43) this can be written as


 2 
J − Jz2 − Jz |a, bmax  = [a − bmax (bmax + 1)] 2 |a, bmax  = 0 . (6.50)

Since |a, bmax  = 0 one has

a − bmax (bmax + 1) = 0 , (6.51)

or

a = bmax (bmax + 1) . (6.52)

In a similar way with the help of Eq. (6.44) one can show that there exists a
minimum value bmin for which

a = bmin (bmin − 1) . (6.53)

From the last two equations one finds that

bmax (bmax + 1) = bmin (bmin − 1) , (6.54)

or

Eyal Buks Quantum Mechanics - Lecture Notes 155


Chapter 6. Angular Momentum

(bmax + bmin ) (bmax − bmin + 1) = 0 . (6.55)


Thus, since bmax − bmin + 1 > 0 one finds that
bmin = −bmax . (6.56)
The formal solutions of Eqs. (6.52) and (6.53) are given by
1 1√
bmax = − ± 1 + 4a , (6.57)
2 2
and
1 1√
bmin = ∓ 1 + 4a . (6.58)
2 2
Furthermore, a is an eigenvalue of a positive-definite operator J2 , therefore
a ≥ 0. Consequently, the only possible solutions for which bmax ≥ bmin are
1 1√
bmax = − + 1 + 4a ≥ 0 , (6.59)
2 2
and
1 1√
bmin = − 1 + 4a = −bmax ≤ 0 . (6.60)
2 2
That is, for a given value of a, both bmax and bmin are uniquely de-
termined. The value bmin is obtained by successively applying the oper-
ator J− to the state |a, bmax  an integer number of times, and therefore
bmax − bmin = 2bmax must be an integer. Consequently, the possible values of
bmax are 0, 1/2, 1, 3/2, · · · .
We now change the notation |a, b for the simultaneous eigenvectors to
the more common notation |j, m, where
j = bmax , (6.61)
m=b. (6.62)
Our results can be summarized by the following relations
J2 |j, m = j (j + 1) 2 |j, m , (6.63)
Jz |j, m = m |j, m , (6.64)

J+ |j, m = j (j + 1) − m (m + 1) |j, m + 1 , (6.65)

J− |j, m = j (j + 1) − m (m − 1) |j, m − 1 , (6.66)
where the possible values j can take are
1 3
j = 0, , 1, , · · · , (6.67)
2 2
and for each given j, the quantum number m can take any of the 2j + 1
possible values
m = −j, −j + 1, · · · , j − 1, j . (6.68)

Eyal Buks Quantum Mechanics - Lecture Notes 156


6.5. Orbital Angular Momentum

6.4 Example - Spin 1/2

The vector space of a spin 1/2 system is the subspace spanned by the ket-
vectors |j = 1/2, m = −1/2 and |j = 1/2, m = 1/2. In this subspace the
spin angular momentum is labeled using the letter S, as we have discussed
above. The matrix representation of some operators of interest in this basis
can be easily found with the help of Eqs. (6.63), (6.64), (6.65) and (6.66):
2
 
2 3 10
S =˙ , (6.69)
4 01
 
 1 0 
Sz =
˙ ≡ σz , (6.70)
2 0 −1 2
 
01
S+ =
˙  , (6.71)
00
 
00
S− =
˙  . (6.72)
10
The above results for S+ and S− together with the identities
S+ + S−
Sx = , (6.73)
2
S+ − S−
Sy = , (6.74)
2i
can be used to find the matrix representation of Sx and Sy
 
 01 
Sx =
˙ ≡ σx , (6.75)
2 10 2
 
 0 −i 
Sy =
˙ ≡ σy . (6.76)
2 i 0 2
The matrices σ x , σy and σz are called Pauli’s matrices, and are related to
the corresponding spin angular momentum operators by the relation

Sk =
˙ σk . (6.77)
2

6.5 Orbital Angular Momentum

As we have discussed above, orbital angular momentum L = r × p refers to


spatial motion. For this case the states |l, m (here, the letter l is used instead
of j since we are dealing with orbital angular momentum) can be described
using wave functions. In this section we calculate these wave functions. For
this purpose it is convenient to employ the transformation from Cartesian to
spherical coordinates

Eyal Buks Quantum Mechanics - Lecture Notes 157


Chapter 6. Angular Momentum

x = r sin θ cos φ , (6.78)


y = r sin θ sin φ , (6.79)
z = r cos θ , (6.80)
where
r≥0, (6.81)
0≤θ≤π, (6.82)
0 ≤ φ ≤ 2π . (6.83)
Exercise 6.5.1. Show that:
1.

r′ | Lz |α = −i x′ |α . (6.84)
∂φ
2.
 
∂ ∂
r′ | L± |α = −i exp (±iφ) ±i − cot θ r′ |α . (6.85)
∂θ ∂φ
3.
 
′ 2 2 1 ∂2 1 ∂ ∂
r | L |α = − 2 2 + sin θ ∂θ sin θ r′ |α . (6.86)
sin θ ∂φ ∂θ
Solution 6.5.1. Using the relations
 
x̂ ŷ ẑ
L = r × p = det  x y z  , (6.87)
px py pz

r′ | r |α = r′ r′ |α , (6.88)



r′ | p |α = ∇ r′ |α , (6.89)
i
[see Eqs. (3.21) and (3.29)] one finds that
 
 ∂ ∂
r′ | Lx |α = y −z ψα (r′ ) , (6.90)
i ∂z ∂y
 
 ∂ ∂
r′ | Ly |α = z −x ψα (r′ ) , (6.91)
i ∂x ∂z
 
 ∂ ∂
r′ | Lz |α = x −y ψα (r′ ) , (6.92)
i ∂y ∂x
where

ψα (r′ ) = r′ |α . (6.93)

Eyal Buks Quantum Mechanics - Lecture Notes 158


6.5. Orbital Angular Momentum

The inverse transformation is given by



r = x2 + y 2 + z 2 , (6.94)
z
cos θ =  , (6.95)
x2 + y 2 + z 2
x
cot φ = . (6.96)
y
1. The following holds
∂ ∂x ∂ ∂y ∂ ∂z ∂
= + +
∂φ ∂φ ∂x ∂φ ∂y ∂φ ∂z
∂ ∂
= −r sin θ sin φ + r sin θ cos φ
∂x ∂y
∂ ∂
= −y +x ,
∂x ∂y
(6.97)
thus using Eq. (6.92) one has


r′ | Lz |α = −i ψ (r′ ) . (6.98)
∂φ α

2. Using Eqs. (6.90) and (6.91) together with the relation L+ = Lx + iLy
one has
i ′ i
r | L+ |α = r′ | Lx + iLy |α
  
∂ ∂ ∂ ∂
= y −z + iz − ix ψα (r′ )
∂z ∂y ∂x ∂z
  
∂ ∂ ∂
= z i − − i (x + iy) ψα (r′ )
∂x ∂y ∂z
  
∂ ∂ ∂
= z i − − ir sin θeiφ ψα (r′ ) .
∂x ∂y ∂z
(6.99)
Thus, by using the identity
∂ ∂x ∂ ∂y ∂ ∂z ∂
= + +
∂θ ∂θ ∂x ∂θ ∂y ∂θ ∂z
 
∂ ∂ ∂
= r cos θ cos φ + sin φ − r sin θ ,
∂x ∂y ∂z
(6.100)
or
 
∂ ∂ ∂ ∂
r sin θ = r cos θ cos φ + sin φ − , (6.101)
∂z ∂x ∂y ∂θ

Eyal Buks Quantum Mechanics - Lecture Notes 159


Chapter 6. Angular Momentum

one finds that


     
i ′ ∂ ∂ iφ ∂ ∂ ∂
r | L+ |α = z i − − ie cot θ x +y − ψα (r′ )
 ∂x ∂y ∂x ∂y ∂θ

 iφ
 ∂  iφ
 ∂ iφ ∂
= i z − e x cot θ − z + ie y cot θ + ie ψα (r′ )
∂x ∂y ∂θ
     
= eiφ i cot θ ze −iφ  ∂  −iφ tan θ + iy  ∂ + i ∂  ψ α (r′ )
 tan θ − x ∂x − cot θ ze    ∂y ∂θ
x−iy x−iy
  
∂ ∂ ∂
= eiφ cot θ y −x +i ψα (r′ )
∂x ∂y ∂θ
 
iφ ∂ ∂
=e i − cot θ ψα (r′ ) .
∂θ ∂φ
(6.102)
In a similar way one evaluates r′ | L− |α. Both results can be expressed
as
 
′ ∂ ∂
r | L± |α = −i exp (±iφ) ±i − cot θ ψα (r′ ) . (6.103)
∂θ ∂φ
3. Using the result of the previous section one has
1
r′ | Lx |α = r′ | (L+ + L− ) |α
2    
i iφ ∂ ∂ −iφ ∂ ∂
= e cot θ −i +e cot θ +i ψα (r′ )
2 ∂φ ∂θ ∂φ ∂θ
 
∂ ∂
= i cos φ cot θ + sin φ ψα (r′ ) .
∂φ ∂θ
(6.104)
Similarly
 
′ ∂ ∂
r | Ly |α = i sin φ cot θ − cos φ ψα (r′ ) , (6.105)
∂φ ∂θ
thus
r′ | L2 |α = r′ | L2x + L2y + L2z |α
 2  2 
2 ∂ ∂ ∂ ∂ ∂2
= − cos φ cot θ + sin φ + sin φ cot θ − cos φ + 2 ψα (r′ )
∂φ ∂θ ∂φ ∂θ ∂φ

  ∂2 ∂ ∂2
= −2 1 + cot2 θ 2 + cot θ ∂θ + ψ α (r′ )
∂φ ∂θ 2
 
1 ∂2 1 ∂ ∂
= −2 + sin θ ψα (r′ ) .
sin2 θ ∂φ2 sin θ ∂θ ∂θ
(6.106)

Eyal Buks Quantum Mechanics - Lecture Notes 160


6.5. Orbital Angular Momentum

Spherical Harmonics. The above exercise allows translating the relations


(6.63) and (6.64), which are given by
L2 |l, m = l (l + 1) 2 |l, m , (6.107)
Lz |l, m = m |l, m , (6.108)
into differential equations for the corresponding wavefunctions
 
1 ∂2 1 ∂ ∂
− + sin θ ψα (r′ ) = l (l + 1) ψα (r′ ) , (6.109)
sin2 θ ∂φ2 sin θ ∂θ ∂θ

−i ψα (r′ ) = mψα (r′ ) , (6.110)
∂φ
where

m = −l, −l + 1, · · · , l − 1, l . (6.111)

We seek solutions having the form

ψα (r′ ) = f (r) Ylm (θ, φ) . (6.112)

We require that both f (r) and Ylm (θ, φ) are normalized


∞
1= drr2 |f (r)|2 , (6.113)
0
π 2π
1= dθ sin θ dφ |Ylm (θ, φ)|2 . (6.114)
0 0

These normalization requirements guarantee that the total wavefunction is


normalized
∞ ∞ ∞
2
1= dx dy dz |ψα (r′ )| . (6.115)
−∞ −∞ −∞

Substituting into Eqs. (6.109) and (6.110) yields


  

1 ∂2 1 ∂ sin θ ∂θ
− + Ylm = l (l + 1) Ylm , (6.116)
sin2 θ ∂φ2 sin θ ∂θ
∂ m
−i Y = mYlm . (6.117)
∂φ l
The functions Ylm (θ, φ) are called spherical harmonics
In the previous section, which discusses the case of general angular mo-
mentum, we have seen that the quantum number m can take any half integer
value 0, 1/2, 1, 3/2, · · · [see Eq. (6.67)]. Recall that the only assumption em-
ployed in order to obtain this result was the commutation relations (6.22).

Eyal Buks Quantum Mechanics - Lecture Notes 161


Chapter 6. Angular Momentum

However, as is shown by the claim below, only integer values are allowed for
the case of orbital angular momentum. In view of this result, one may argue
that the existence of spin, which corresponds to half integer values, is in fact
predicted by the commutation relations (6.22) only.

Claim. The variable m must be an integer.

Proof. Consider a solution having the form

Ylm (θ, φ) = Flm (θ) eimφ . (6.118)

Clearly, Eq. (6.117) is satisfied. The requirement

Ylm (θ, φ) = Ylm (θ, φ + 2π) , (6.119)

namely, the requirement that Ylm (θ, φ) is continuos, leads to

e2πim = 1 , (6.120)

thus m must be an integer.

The spherical harmonics Ylm (θ, φ) can be obtained by solving Eqs. (6.116)
and (6.117). However, we will employ an alternative approach, in which in
the first step we find the spherical harmonics Yll (θ, φ) by solving the equation

L+ |l, l = 0 , (6.121)

which is of first order [contrary to Eq. (6.116), which is of the second order].
Using the identity (6.85), which is given by
 
′ iφ ∂ ∂
r | L+ |α = −ie i − cot θ r′ |α , (6.122)
∂θ ∂φ

one has
 

− l cot θ Fll (θ) = 0 . (6.123)
∂θ

The solution is given by

Fll (θ) = Cl (sin θ)l , (6.124)

where Cl is a normalization constant. Thus, Yll is given by

Yll (θ, φ) = Cl (sin θ)l eilφ . (6.125)

In the second step we employ the identity (6.66), which is given by



J− |j, m = j (j + 1) − m (m − 1) |j, m − 1 , (6.126)

Eyal Buks Quantum Mechanics - Lecture Notes 162


6.6. Problems

and Eq. (6.85), which is given by


 
′ ∂ ∂
r | L± |α = −i exp (±iφ) ±i − cot θ r′ |α , (6.127)
∂θ ∂φ

to derive the following recursive relation


  

e−iφ − − m cot θ Ylm (θ, φ) = l (l + 1) − m (m − 1)Ylm−1 (θ, φ) ,
∂θ
(6.128)

which allows finding Ylm (θ, φ) for all possible values of m provided that
Yll (θ, φ) is given. The normalized spherical harmonics are found using this
method to be given by
:
(−1)l 2l + 1 (l + m)! imφ dl−m
Ylm (θ, φ) = l e (sin θ)−m l−m
(sin θ)2l .
2 l! 4π (l − m)! d (cos θ)
(6.129)

As an example, closed form expressions for the cases l = 0 and l = 1 are


given below
1
Y00 (θ, φ) = √ , (6.130)

0
±1 3
Y1 (θ, φ) = ∓ sin θe±iφ , (6.131)

0
0 3
Y1 (θ, φ) = cos θ . (6.132)

6.6 Problems

1. Let Rı̂ (where i ∈ {x, y, z} ) be the 3×3 rotation matrices (as defined in
the lecture). Show that for infinitesimal angle φ the following holds
 
[Rx̂ (φ) , Rŷ (φ)] = 1 − Rẑ φ2 , (6.133)

where

[Rx̂ (φ) , Rŷ (φ)] = Rx̂ (φ) Rŷ (φ) − Rŷ (φ) Rx̂ (φ) . (6.134)

2. Show that
   
iJz φ iJz φ
exp Jx exp − = Jx cos φ − Jy sin φ . (6.135)
 

Eyal Buks Quantum Mechanics - Lecture Notes 163


Chapter 6. Angular Momentum

3. The components of the Pauli matrix vector σ are given by:


     
01 0 −i 1 0
σx = , σy = , σz = . (6.136)
10 i 0 0 −1

a) Show that

(σ · a) (σ · b) = a · b + iσ · (a × b) , (6.137)

where a and b are vector operators which commute with σ , but not
necessarily commute with each other.
b) Show that
 
iσ · n̂φ φ φ
exp − = 1 cos − iσ · n̂ sin , (6.138)
2 2 2

where n̂ is a unit vector and where 1 is the 2 × 2 identity matrix.


4. Find the eigenvectors and eigenvalues of the matrix σ · n̂ (n̂ is a unit
vector).
5. Consider an electron in a state in which the component of its spin along
the z axis is +/2 . What is the probability that the component of the spin
along an axis z ′ , which makes an angle θ with the z axis, will measured
to be +/2 or −/2 . What is the average value of the component of the
spin along this axis?
6. The 2 × 2 matrix U is given by

1 + iα (σ · n̂)
U= , (6.139)
1 − iα (σ · n̂)

where

σ = σ x x̂ + σ y ŷ + σz ẑ (6.140)

is the Pauli vector matrix,

n̂ = nx x̂ + ny ŷ + nz ẑ (6.141)

is a unit vector, i.e. n̂· n̂ = 1, and nx , ny , nz and α are all real parameters.
Note that generally for a matrix or an operator A1 ≡ A−1 .
a) show that U is unitary.
b) Show that

dU 2i (σ · n̂)
= U. (6.142)
dα 1 + α2
c) Calculate U by solving the differential equation in the previous sec-
tion.

Eyal Buks Quantum Mechanics - Lecture Notes 164


6.6. Problems

7. A particle is located in a box, which is divided into a left and right


sections. The corresponding vector states are denoted as |L and |R
respectively. The Hamiltonian of the system is given by

H = EL |L L| + ER |R R| + ∆ (|L R| + |R L|) . (6.143)

The particle at time t = 0 is in the left section

|ψ (t = 0) = |L . (6.144)

Calculate the probability pR (t) to find the particle in the state |R at
time t.
8. A magnetic field given by

B (t) = B0 ẑ + B1 (x̂ cos (ωt) + ŷ sin (ωt)) (6.145)

is applied to a spin 1/2 particle. At time t = 0 the state is given by

|α (t = 0) = |+; ẑ . (6.146)

Calculate the probability P+− (t) to find the system in the state |−; ẑ
at time t > 0.
9. A magnetic field given by

B (t) = B0 ẑ + g (t) B1 (x̂ cos (ωt) + ŷ sin (ωt)) (6.147)

is applied to a spin 1/2 particle. While B0 , B1 and ω 1 are taken to be


constants, the function g (t) is assumed to be given by


 0 t<0


1 0 ≤ t < τp
g (t) = 0 τp ≤ t < τp + τ0 , (6.148)


 1 τ p + τ 0 ≤ t < 2τ p + τ 0


0 2τ p + τ 0 ≤ t

i.e. two oscillatory magnetic field pulses, both having duration of τ p , are
applied, and the dwell time between these pulses is τ 0 . The normalized
pulse duration αp is defined to be

αp = ω 1 τ p , (6.149)

the normalized dwell time α0 is defined to be

α0 = ∆ωτ 0 , (6.150)

and the normalized detuning δ is defined to be


∆ω
δ= , (6.151)
ω1

Eyal Buks Quantum Mechanics - Lecture Notes 165


Chapter 6. Angular Momentum

where

∆ω = ω 0 − ω ,

and where
|e| B0
ω0 = , (6.152)
me c
|e| B1
ω1 = . (6.153)
me c
At time t = 0 the state is assumed to be given by

|α (t = 0) = |+; ẑ . (6.154)

Calculate the probability P++ (t) to find the system in the state |+; ẑ
at time t > 2τ p + τ 0 . Assume that the normalized detuning is small, i.e.
|δ| ≪ 1, and expand P++ (t) to lowest nonvanishing order in δ for the
case where the normalized pulse duration is taken to be given by
π
αp = . (6.155)
2
10. Find the time evolution of the state vector of a spin 1/2 particle in
a magnetic field along the z direction with time dependent magnitude
B (t) = B (t) ẑ.
11. A magnetic field given by B = B cos (ωt) ẑ, where B is a constant, is
applied to a spin 1/2. At time t = 0 the spin is in state |ψ (t), which
satisfies

Sx |ψ (t = 0) = |ψ (t = 0) , (6.156)
2
Calculate the expectation value Sz  at time t ≥ 0.
12. Consider a spin 1/2 particle. The time dependent Hamiltonian is given
by
4ωSz
H=− , (6.157)
1 + (ωt)2
where ω is a real non-negative constant and Sz is the z component of the
angular momentum operator. Calculate the time evolution operator u of
the system.
13. Consider a spin 1/2 particle in an eigenstate of the operator S · n̂ with
eigenvalue +/2 , where S is the vector operator of angular momentum
and where n̂ is a unit vector. The angle between the unit vector n̂ and
the z axis is θ . Calculate the expectation values
a) S
' z (
b) (∆Sz )2

Eyal Buks Quantum Mechanics - Lecture Notes 166


6.6. Problems

14. An ensemble of spin 1/2 particles are in a normalized state

|ψ = α |+ + β |− ,

where the states |+ and |− are the eigenstates of Sz (the z component
of the angular momentum operator). At what direction the magnetic field
should be aligned in a Stern-Gerlach experiment in order for the beam
not to split.
15. Consider a spin 1/2 particle having gyromagnetic ratio γ in a magnetic
field given by B (t) û . The unit vector is given by

û = (sin θ cos ϕ, sin θ sin ϕ, cos θ) , (6.158)

where θ, ϕ are angles in spherical coordinates. The field intensity is given


by

 0 t<0
B (t) = B0 0 < t < τ . (6.159)

0 t>τ

At times t < 0 the spin was in state |+ , namely in eigenstate of Sz with
positive eigenvalue. Calculate the probability P− (t) to find the spin in
state |− at time t , where t > τ .
16. Consider a spin 1/2 particle. The Hamiltonian is given by

H = ωSx , (6.160)

where ω is a Larmor frequency and where Sx is the x component of the


angular momentum operator. The z component of the angular momen-
tum is measured at the times tn = nT /N where n = 0, 1, 2, · · · , N , N is
integer and T is the time of the last measurement.
a) Find the matrix representation of the time evolution operator u (t)
in the basis of |±; ẑ states.
b) What is the probability psame to get the same result in all N + 1
measurements. Note that the initial state of the particle is unknown.
c) For a fixed T calculate the limit lim psame .
N→∞
17. Consider a spin 1/2 particle. No external magnetic field is applied. Three
measurements are done one after the other. In the first one the z com-
ponent of the angular momentum is measured, in the second one the
component along the direction û is measured and in the third measure-
ment, again the z component is measured. The unit vector û is described
using the angles θ and ϕ

û = (sin θ cos ϕ, sin θ sin ϕ, cos θ) . (6.161)

Calculate the probability psame to have the same result in the 1st and
3rd measurements.

Eyal Buks Quantum Mechanics - Lecture Notes 167


Chapter 6. Angular Momentum

18. Let µ (t) be the expectation value of the magnetic moment associated
with spin 1/2 particle (µ = γS , where S is the angular momentum and
γ is the gyromagnetic ratio). Show that in the presence of a time varying
magnetic field B (t) the following holds

d
µ (t) = γ µ (t) × B (t) . (6.162)
dt
19. The Hamiltonian of an electron of mass m, charge q, spin 1/2, placed in
electromagnetic field described by the vector potential A (r, t) and the
scalar potential ϕ (r, t), can be written as [see Eq. (1.62)]
 2
p − qc A q
H= + qϕ − σ·B, (6.163)
2m 2mc
where B =∇×A. Show that this Hamiltonian can also be written as
1 3  q 42
H= σ· p− A + qϕ . (6.164)
2m c
20. Show that
3 4  
j, m| (∆Jx )2 + (∆Jy )2 |j, m = 2 j 2 + j − m2 . (6.165)

21. Find the condition under which the Hamiltonian of a charged particle in
a magnetic field
1  q 2
H= p− A . (6.166)
2m c
can be written as
1 2 q q2
H= p − p·A+ A2 . (6.167)
2m mc 2mc2
22. Consider a point particle having mass m and charge q moving under the
influence of electric field E and magnetic field B, which are related to
the scalar potential ϕ and to the vector potential A by
1 ∂A
E = −∇ϕ − , (6.168)
c ∂t
and

B=∇×A. (6.169)

Find the coordinates representation of the time-independent Schrödinger


equation H |α = E |α.

Eyal Buks Quantum Mechanics - Lecture Notes 168


6.6. Problems

23. A particle of mass m and charge e interacts with a vector potential


Ax = Az = 0 , (6.170)
Ay = Bx . (6.171)
Calculate the ground state energy. Clue: Consider a wave function of the
form

ψ (x, y, z) = χ (x) exp (iky y) exp (ikz z) . (6.172)

24. Find the energy spectrum of a charged particle having mass m and charge
q moving in uniform and time-independent magnetic field B = Bẑ and
electric field E = Ex̂.
25. Consider a particle having mass m and charge e moving in xy plane under
the influence of the potential U (y) = 12 mω 20 y 2 . A uniform and time-
independent magnetic field given by B = Bẑ is applied perpendicularly
to the xy plane. Calculate the eigenenergies of the particle.
26. Consider a particle with charge q and mass µ confined to move on a circle
of radius a in the xy plane, but is otherwise free. A uniform and time
independent magnetic field B is applied in the z direction.
a) Find the eigenenergies.
b) Calculate the current Jm for each of the eigenstates of the system.
27. The Hamiltonian of a non isotropic rigid rotator is given by

L2x L2y L2
H= + + z , (6.173)
2Ixy 2Ixy 2Iz
where L is the vector angular momentum operator. At time t = 0 the
state of the system is described by the wavefunction

ψ (θ, φ) = A sin θ cos φ , (6.174)

where θ, φ are angles in spherical coordinates and A is a normalization


constant. Calculate the expectation value Lz  at time t > 0 .
28. The eigenstates of the angular momentum operators L2 and Lz with
l = 1 and m = −1, 0, 1 are denoted as |1, −1, |1, 0 and |1, 1.
a) Write the 3 × 3 matrix of the operator Lx in this l,= 1 subspace.
√ -
b) Calculate the expectation value Lx  for the state 12 |1, 1 + 2 |1, 0 + |1, −1 .
c) The same as the previous section for the state √12 [|1, 1 − |1, −1].
d) Write the 3 × 3 matrix representation in this basis of the rotation
operator at angle φ around the z axis.
e) The same as in the previous section for an infinitesimal rotation with
angle dφ around the x axis.
29. Consider a particle of mass m in a 3D harmonic potential
1  
V (x, y, z, ) = mω 2 x2 + y 2 + z 2 . (6.175)
2

Eyal Buks Quantum Mechanics - Lecture Notes 169


Chapter 6. Angular Momentum

The state vector |ψ of the particle satisfy


ax |ψ = αx |ψ , (6.176)
ay |ψ = αy |ψ , (6.177)
az |ψ = αz |ψ , (6.178)
where αx , αy and αz are complex and ax , ay and az are annihilation
operators
0  
mω ipx
ax = x+ , (6.179)
2 mω
0  
mω ipy
ay = y+ , (6.180)
2 mω
0  
mω ipz
az = z+ , (6.181)
2 mω
Let L be the vector operator of the orbital angular momentum.
a) Calculate Lz .
b) Calculate ∆Lz .
30. A rigid rotator is prepared in a state
|α = A (|1, 1 − |1, −1) , (6.182)
where A is a normalization constant, and where the symbol |l, m denotes
an angular momentum state with quantum numbers l and m. Calculate
Lx .
a) ' (
b) (∆Lx )2 .
31. The Hamiltonian of a top is given by
L2x + L2y L2
H= + z , (6.183)
2I1 2I2
where L is the angular momentum vector operator. Let |ψ0  be the
ground state of the system.
a) Calculate the quantity Az (φ), which is defined as
   
iLz φ iLz φ
Az (φ) = ψ 0 | exp H exp − |ψ0  . (6.184)
 
b) Calculate the quantity Ax (φ), which is defined as
   
iLx φ iLx φ
Ax (φ) = ψ0 | exp H exp − |ψ0  . (6.185)
 
32. The wavefunction of a point particle is given by
ψ (r) = (x + y + 2z) f (r) , (6.186)

where f (r) is a function of the radial coordinate r = x2 + y 2 + z 2 .

Eyal Buks Quantum Mechanics - Lecture Notes 170


6.6. Problems

a) In a measurement of L2 what are the possible outcomes and the


corresponding probabilities.
b) The same for a measurement of Lz .
33. Consider a system comprising of two spin 1/2 particles.
a) Show that
, 2 -
S , Sz = 0 , (6.187)

where S = S1 + S2 , Sz = S1z + S2z and where S1 and S2 are the an-


gular momentum vector operators of the first and second spin repet-
itively, i.e. S1 = (S1x , S1y , S1z ) and S2 = (S2x , S2y , S2z ).
b) Find an orthonormal basis of common eigenvectors of S2 and Sz
[recall that the existence , of such- a basis is guaranteed by the result
of the previous section S2 , Sz = 0, see Eqs. (2.147) and (2.148)].
34. Consider a system in a common eigenvector |j, m of the angular momen-
tum operators J2 and Jz . A measurement of the operator Jn̂ = n̂ · J is
being performed, where n̂ = (cos ϕ sin θ, sin ϕ sin θ, cos θ) ' is a unit
( vector.
2
Calculate the expectation value Jn̂  and the variance (∆Jn̂ ) .
35. Consider a harmonic oscillator having angular resonance frequency ω and
mass m. The operator S (ξ, ϕ) is defined by [compare with Eq. (5.83)]
 iϕ †2 
e a − e−iϕ a2
S (ξ, ϕ) = exp ξ , (6.188)
2

where both ξ and ϕ are real and


0  
mω ip
a= x+ (6.189)
2 mω

is the annihilation operator [see Eq. (5.9)]. Show that S (ξ, ϕ) can be
factorized according to
 iϕ 
e †2
S (ξ, ϕ) = exp a tanh ξ
2
 
log (cosh ξ)  † †

× exp − aa + a a
2
 −iϕ 
e 2
× exp − a tanh ξ .
2
(6.190)

36. Show that the operator S (ξ, ϕ) (6.188) satisfies


 
S (ξ, 0) = Q e−ξ , (6.191)

where Q, which is called the squeezing operator, is given by

Eyal Buks Quantum Mechanics - Lecture Notes 171


Chapter 6. Angular Momentum

∞
dx′
Q (µ) = √ |x′ /µ x′ | , (6.192)
µ
−∞

where |x′  is an eigenvector of the position operator x having eigenvalue


x′ , i.e. x |x′  = x′ |x′ .

6.7 Solutions

1. By cyclic permutation of
 
cos φ − sin φ 0
Rẑ =  sin φ cos φ 0  , (6.193)
0 0 1

one has
 
1 0 0
Rx̂ =  0 cos φ − sin φ  , (6.194)
0 sin φ cos φ
 
cos φ 0 − sin φ
Rŷ =  0 1 0  . (6.195)
sin φ 0 cos φ
On one hand
1 − [Rx̂ (φ) , Rŷ (φ)]
 
1 −1 + cos2 φ sin φ − sin φ cos φ
= 1 − cos2 φ 1 sin φ cos φ − sin φ 
sin φ − sin φ cos φ sin φ cos φ − sin φ 1
 2 
1 −φ 0  
=  φ2 1 0  + O φ3 .
0 0 1
(6.196)
On the other hand
   
 2 cos φ2 − sin φ2 0 1 −φ2 0  
Rẑ φ =  sin φ2 cos φ2 0  =  φ2 1 0  + O φ3 , (6.197)
0 0 1 0 0 1

thus
   
1 − [Rx̂ (φ) , Rŷ (φ)] = Rẑ φ2 + O φ3 . (6.198)

Eyal Buks Quantum Mechanics - Lecture Notes 172


6.7. Solutions

2. Using the identity (2.173), which is given by


1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · , (6.199)
2! 3!
and the commutation relations (6.22), which are given by
[Ji , Jj ] = iεijk Jk , (6.200)
one has
   
iJz φ iJz φ
exp Jx exp −
 
 2
iφ 1 iφ
= Jx + [Jz , Jx ] + [Jz , [Jz , Jx ]]
 2! 
 3
1 iφ
+ [Jz , [Jz , [Jz , Jx ]]] + · · ·
3! 
   
1 2 1 3
= Jx 1 − φ + · · · − Jy φ − φ + · · ·
2! 3!
Jx cos φ − Jy sin φ .
(6.201)
3. The components of the Pauli matrix vector σ are given by:
     
01 0 −i 1 0
σx = , σy = , σz = . (6.202)
10 i 0 0 −1
a) The following  holds 
az ax − iay
σ·a= , (6.203)
ax + iay −az
 
bz bx − iby
σ·b = , (6.204)
bx + iby −bz
thus  
az bz + (ax − iay ) (bx + iby ) az (bx − iby ) − (ax − iay ) bz
(σ · a) (σ · b) =
(ax + iay ) bz − az (bx + iby ) az bz + (ax + iay ) (bx − iby )
 
10
= a·b
01
 
01
+i (ay bz − az by )
10
 
0 −i
+i (az bx − ax bz )
i 0
 
1 0
+i (ax by − ay bx )
0 −1
= a · b + iσ · (a × b) .
(6.205)

Eyal Buks Quantum Mechanics - Lecture Notes 173


Chapter 6. Angular Momentum

b) Using (a) one has

(σ · n̂)2 = 1 , (6.206)

thus with the help of the Taylor expansion of the functions cos (x)
and sin (x)
 one finds     
iσ · n̂φ σ · n̂φ σ · n̂φ
exp − = cos − i sin
2 2 2
φ φ
= 1 cos − iσ · n̂ sin .
2 2
(6.207)
4. In spherical coordinates the unit vectors n̂ is expressed as

n̂ = (cos ϕ sin θ, sin ϕ sin θ, cos θ) , (6.208)

thus
 
cos θ sin θe−iϕ
σ · n̂ = . (6.209)
sin θeiϕ − cos θ

The eigen values λ+ and λ− are found solving

λ+ + λ− = Tr (σ · n̂) = 0 , (6.210)

and

λ+ λ− = Det (σ · n̂) = −1 , (6.211)

thus

λ± = ±1 . (6.212)

The normalized eigenvectors can be chosen to be given by


  
cos θ2 exp − iϕ2
|+ =
˙ , (6.213)
sin θ2 exp iϕ2
  
− sin θ2 exp − iϕ 2

|− =
˙ . (6.214)
cos θ2 exp iϕ 2

5. Using Eq. (6.213) one finds the probability p+ to measure +/2 is given
by

   


  cos θ exp − iϕ
2
p+ =

1 0 2  2
= cos2 θ , (6.215)
sin θ2 exp iϕ2

2

and the probability p− to measure −/2 is


θ
p− = 1 − p+ = sin2 . (6.216)
2

Eyal Buks Quantum Mechanics - Lecture Notes 174


6.7. Solutions

The average value of the component of the spin along z ′ axis is thus
 
 θ θ 
cos2 − sin2 = cos θ . (6.217)
2 2 2 2

6. In general, note that all smooth functions of the matrix (σ · n̂) commute,
a fact that greatly simplifies the calculations.
a) The following holds
1
= 1 + iα (σ · n̂) + [iα (σ · n̂)]2 + · · · , (6.218)
1 − iα (σ · n̂)

thus †
1
= 1 − iα (σ · n̂) + [(−i) α (σ · n̂)]2 + · · ·
1 − iα (σ · n̂)
1
= ,
1 + iα (σ · n̂)
(6.219)
therefore
1 + iα (σ · n̂) 1 − iα (σ · n̂)
UU† = =1, (6.220)
1 − iα (σ · n̂) 1 + iα (σ · n̂)

and similarly U † U = 1.
b) Exploiting again the fact that all smooth functions of the matrix
(σ · n̂) commute and using Eq. (6.206) one has
dU [1 − iα (σ · n̂)] (σ · n̂) + [1 + iα (σ · n̂)] (σ · n̂)
=i
dα [1 − iα (σ · n̂)]2
2 (σ · n̂)
=i
[1 − iα (σ · n̂)]2
2 (σ · n̂) 1 + iα (σ · n̂)
=i
[1 − iα (σ · n̂)] [1 + iα (σ · n̂)] 1 − iα (σ · n̂)
2i (σ · n̂)
= U.
1 + α2
(6.221)
c) By integration one
 has  α 
dα′
U = U0 exp 2i (σ · n̂) ′2
0 1+α
 −1

U0 exp 2i (σ · n̂) tan α ,
(6.222)
where U0 is a the matrix U at α = 0. With the help of Eq. (6.138)
one thus finds that
,    -
U = U0 1 cos 2 tan−1 α + iσ · n̂ sin 2 tan−1 α , (6.223)

Eyal Buks Quantum Mechanics - Lecture Notes 175


Chapter 6. Angular Momentum

Using the identities


  1 − α2
cos 2 tan−1 α = , (6.224)
1 + α2
  2α
sin 2 tan−1 α = , (6.225)
1 + α2
and assuming U0 = 1 one finds that

1 − α2 2α
U= + iσ · n̂ . (6.226)
1 + α2 1 + α2
7. In terms of Pauli matrices

H=E
˙ a σ0 + ∆σx + Ed σz , (6.227)

where
EL + ER EL − ER
Ea = , Ed = , (6.228)
2 2
and
     
10 01 1 0
σ0 = , σx = , σz = . (6.229)
01 10 0 −1

Using Eq. (6.138), which is given by


 
iσ · n̂φ φ φ
exp − = cos − iσ · n̂ sin , (6.230)
2 2 2

the time evolution operator u (t) can be calculated


 
iHt
u (t) = exp −

   
iEa σ 0 t i (∆σ x + Ed σz ) t
= exp − exp −
 
  !  "
iEa t iσ · n̂ ∆2 + Ed2 t
= exp − exp − ,
 
(6.231)
where
(∆, 0, Ed )
σ · n̂ = σ ·  , (6.232)
∆2 + Ed2

thus
 !   "
iEa t ∆2 + Ed2 t ∆σ x + Ed σz t ∆2 + Ed2
u (t) = exp − cos −i  sin .
  ∆2 + Ed2 

Eyal Buks Quantum Mechanics - Lecture Notes 176


6.7. Solutions

(6.233)

The probability pR (t) is thus given by


2
pR (t) = |R| u (t) |ψ (t = 0)|
= |R| u (t) |L|2
.  2
∆2
2
t ∆2 + EL −E
2
R

=  EL −ER 2 sin .
∆2 + 
2
(6.234)
8. The Hamiltonian is given by

H = ω 0 Sz + ω1 (cos (ωt) Sx + sin (ωt) Sy ) , (6.235)

where
|e| B0
ω0 = , (6.236)
me c
|e| B1
ω1 = . (6.237)
me c
The matrix representation in the basis {|+ , |−} (where |+ = |+; ẑ
and |− = |−; ẑ) is found using Eqs. (6.70), (6.75) and (6.76)
 
 ω0 ω 1 exp (−iωt)
H=˙ . (6.238)
2 ω 1 exp (iωt) −ω 0

The Schrödinger equation is given by


d
i |α = H |α . (6.239)
dt
It is convenient to express the general solution as
   
iωt iωt
|α (t) = b+ (t) exp − |+ + b− (t) exp |− . (6.240)
2 2

Substituting into the Schrödinger equation yields


 − iωt   
d e 2 0 b+
i iωt
dt 0 e 2 b−
 −iωt
  − iωt  
1 ω0 ω1e e 2 0 b+
= ,
2 ω1 eiωt −ω0
iωt
0 e 2 b−
(6.241)
or

Eyal Buks Quantum Mechanics - Lecture Notes 177


Chapter 6. Angular Momentum
 iωt    − iωt  
iω −e− 2 0 b+ e 2 0 ḃ+
iωt + iωt
2 0 e 2 b− 0 e 2 ḃ−
   − iωt  
−i ω0 ω1 e−iωt e 2 0 b+
= .
ω 1 eiωt
iωt
2 −ω 0 0 e 2 b −

(6.242)
By multiplying from the left by
 iωt 
e 2 0
iωt
0 e− 2

one has
       
iω −1 0 b+ ḃ+ −i ω 0 ω 1 b+
+ = , (6.243)
2 0 1 b− ḃ− 2 ω 1 −ω 0 b−
or
   
ḃ+ Ω b+
i = , (6.244)
ḃ− 2 b−

where
 
∆ω ω 1
Ω= = ∆ωσ z + ω 1 σ x , (6.245)
ω 1 −∆ω

and

∆ω = ω 0 − ω . (6.246)

At time t = 0
   
b+ (0) 1
= . (6.247)
b− (0) 0

The time evolution is found using Eq. (6.138)


    
b+ (t) iΩt b+ (0)
= exp −
b− (t) 2 b− (0)
 
cos θ − i √∆ω sin θ
−i √ ω21 sin θ 2  
1
 ω21 +(∆ω)2 ω1 +(∆ω) 
= ,
−i √ ω21 sin θ 2 cos θ + i √∆ω sin θ 0
2
ω1 +(∆ω) ω1 +(∆ω)2

(6.248)
where
.
ω 21 + (∆ω)2 t
θ= . (6.249)
2

Eyal Buks Quantum Mechanics - Lecture Notes 178


6.7. Solutions

The probability is thus given by


.
2
ω1 2
ω21 + (∆ω)2 t
P+− (t) = 2 2 sin . (6.250)
ω 1 + (∆ω) 2

9. The transformation into the rotating frame reads


   
iωt iωt
|α (t) = b+ (t) exp − |+ + b− (t) exp |− . (6.251)
2 2

For time periods where g (t) is constant the time evolution is governed
by Eq. (6.248). Thus at time t = 2τ p + τ 0 one has
   
b+ (2τ p + τ 0 ) 1
= Mp M0 Mp , (6.252)
b− (2τ p + τ 0 ) 0

where
 
cos θp − iδ
√sin θ2p i sin θp
−√ 2
1+δ 1+δ
Mp =  i sin θ p iδ sin θp
 , (6.253)
−√ 2
cos θp + √
1+δ 1+δ 2

 α0 
e−i 2 0
M0 = α0 , (6.254)
0 ei 2
and where

1 + δ 2 αp
θp = . (6.255)
2
Thus the probability P++ (2τ p + τ 0 ) is given by
P++ (2τ p + τ 0 ) = |b+ (2τ p + τ 0 )|2

! "2
2

iα0 2


iδ sin θ p e sin θ p

=
cos θp −  −

1 + δ2 1 + δ 2


 
2

iδ sin (2θ p ) 1 − eiα0 sin2 θp

=
cos (2θp ) −  +
.

1 + δ2 1 + δ2

(6.256)
For the case where αp = π/2 one has to lowest nonvanishing order in δ
1 − cos α0  
P++ (2τ p + τ 0 ) = + δ sin α0 + O δ 2 . (6.257)
2
Note that the probability P++ (t) is unchanged for t > 2τ p + τ 0 .

Eyal Buks Quantum Mechanics - Lecture Notes 179


Chapter 6. Angular Momentum

10. The Schrödinger equation is given by

d |α
i = H |α , (6.258)
dt
where

H = ωSz , (6.259)

and
|e| B (t)
ω (t) = . (6.260)
me c
In the basis of the eigenvectors of Sz one has

|α = c+ |+ + c− |− , (6.261)

and

i (ċ+ |+ + ċ− |−) = ωSz (c+ |+ + c− |−) , (6.262)

where

Sz |± = ± |± , (6.263)
2
thus one gets 2 decoupled equations

ċ+ = − c+ , (6.264)
2

ċ− = c− . (6.265)
2
The solution is given by
  
i t ′ ′
c± (t) = c± (0) exp ∓ ω (t ) dt
2 0
  t 
i |e| ′ ′
= c± (0) exp ∓ B (t ) dt .
2me c 0
(6.266)
11. At time t = 0
1
|ψ (t = 0) = √ (|+ + |−) . (6.267)
2
Using the result of the previous problem and the notation
eB
ωc = , (6.268)
mc

Eyal Buks Quantum Mechanics - Lecture Notes 180


6.7. Solutions

one finds
     
t t
1  iω c iω c
|ψ (t) = √ exp − cos (ωt′ ) dt′  |+ + exp  cos (ωt′ ) dt′  |−
2 2 2
0 0
    
1 iω c sin ωt iωc sin ωt
= √ exp − |+ + exp |− ,
2 2ω 2ω
(6.269)
thus

Sz  (t) = ψ (t)| Sz |ψ (t) = 0 . (6.270)

12. The Schrödinger equation for u is given by


du
i = Hu , (6.271)
dt
thus
du 4iωSz 1
= u. (6.272)
dt  1 + (ωt)2

By integration one finds


 
t ′
4iωS z dt
u (t) = u (0) exp  
 1 + (ωt′ )2
0
 
4iSz
= u (0) exp tan−1 (ωt) .

(6.273)
Setting an initial condition u (t = 0) = 1 yields
 
4iSz −1
u (t) = exp tan (ωt) . (6.274)


The matrix elements of u (t) in the basis of the eigenstates |± of Sz are
given by
  1 + iωt
+| u (t) |+ = exp 2i tan−1 (ωt) = , (6.275)
1 − iωt
  1 − iωt
−| u (t) |− = exp −2i tan−1 (ωt) = , (6.276)
1 + iωt
+| u (t) |− = −| u (t) |+ = 0 . (6.277)
13. The eigenvector of S · n̂, where n̂ = (sin θ cos ϕ, sin θ sin ϕ, cos θ) with
eigenvalue +/2 is given by [see Eq. (6.213)]

Eyal Buks Quantum Mechanics - Lecture Notes 181


Chapter 6. Angular Momentum

θ ϕ θ ϕ
|+; S · n̂ = cos e−i 2 |+ + sin ei 2 |− . (6.278)
2 2
The operator Sz is written as

Sz = (|+ +| − |− −|) . (6.279)
2
a) Thus
 
 2 θ 2θ 
+; S · n̂| Sz |+; S · n̂ = cos − sin = cos θ . (6.280)
2 2 2 2

b) Since Sz2 is the identity operator times 2 /4 one has


' ( ) * 2   2
(∆Sz )2 = Sz2 − Sz 2 = 1 − cos2 θ = sin2 θ . (6.281)
4 4
14. We seek a unit vector n̂ such that

|ψ = |+; S · n̂ , (6.282)

where |+; S · n̂ is given by Eq. (6.213)


   
θ+ iϕ+ θ+ iϕ+
|+; S · n̂ = cos exp − |++sin exp |− , (6.283)
2 2 2 2

thus the following hold



θ+

ctg =

, (6.284)
2 β

and

ϕ+ = arg (β) − arg (α) . (6.285)

Similarly, by requiring that

|ψ = |−; S · n̂ , (6.286)

where
   
θ− iϕ θ− iϕ−
|−; S · n̂ = − sin exp − − |++cos exp |− , (6.287)
2 2 2 2

one finds

θ−
α

tan =

, (6.288)
2 β
ϕ− = arg (β) − arg (α) + π . (6.289)

Eyal Buks Quantum Mechanics - Lecture Notes 182


6.7. Solutions

15. The Hamiltonian at the time interval 0 < t < τ is given by

H = −γB0 (S · û) , (6.290)

where γ is the gyromagnetic ratio and S is the angular momentum op-


erator. The eigenvectors of S · û with eigenvalue ±/2 are given by [see
Eqs. (6.213) and (6.214)]

θ ϕ θ ϕ
|+; S · û = cos e−i 2 |+ + sin ei 2 |− , (6.291)
2 2
θ ϕ θ ϕ
|−; S · û = − sin e−i 2 |+ + cos ei 2 |− , (6.292)
2 2
Thus in the time interval 0 < t < τ the state vector is given by
   
iγB0 t iγB0 t
|α = |+; S · û +; S · û |+ exp + |−; S · û −; S · û |+ exp −
2 2
   
θ ϕ iγB0 t θ ϕ iγB0 t
= |+; S · û cos ei 2 exp − |−; S · û sin ei 2 exp −
2 2 2 2
   
θ iγB0 t θ iγB0 t
= eiϕ cos2 exp + sin2 exp − |+
2 2 2 2
   
θ θ iγB0 t iγB0 t
+ sin cos exp − exp − |−
2 2 2 2
   
1 + cos θ iγB0 t 1 − cos θ iγB0 t
= eiϕ exp + exp − |+
2 2 2 2
 
γB0 t
+i sin θ sin |−
2
     
γB0 t γB0 t γB0 t
= eiϕ cos + i cos θ sin |+ + i sin θ sin |− .
2 2 2
(6.293)
Thus for t > τ
 
2 2 γB0 τ
P− (t) = sin θ sin . (6.294)
2

An alternative solution - The Hamiltonian in the basis of |± states is


given by
γB0 
H=− (σ · û) , (6.295)
2
where σ is the Pauli matrix vector
     
01 0 −i 1 0
σ1 = , σ2 = , σ3 = . (6.296)
10 i 0 0 −1

Eyal Buks Quantum Mechanics - Lecture Notes 183


Chapter 6. Angular Momentum

The time evolution operator is given by


  
iHt iγB0 t
u (t) = exp − = exp (σ · û) . (6.297)
 2

Using the identity (6.138) one finds


   
γB0 t γB0 t
u (t) = I cos + iû · σ sin
2 2
       
cos γB20 t + i cos θ sin γB20 t i sin θe−iϕ sin γB20 t
=       ,
i sin θeiϕ sin γB20 t cos γB20 t − i cos θ sin γB20 t
(6.298)
thus for t > τ

 
2  

  1

2 2 γB0 τ
P− (t) =

0 1 u (t) = sin θ sin . (6.299)


0
2

16. The matrix representation of the Hamiltonian in the basis of |±; Sz 


states is given by

H=
˙ (x̂ · σ) , (6.300)
2
where σ is the Pauli matrix vector
     
01 0 −i 1 0
σx = , σy = , σz = . (6.301)
10 i 0 0 −1

a) The time evolution operator is given by


   
iHt iωt
u (t) = exp − ˙ exp −
= (x̂ · σ) . (6.302)
 2

Using the identity

exp (iu · σ) = 1 cos α + iû · σ sin α , (6.303)

where u =αû is a three-dimensional real vector and û is a three-


dimensional real unit vector, one finds
ωt ωt
u (t) =
˙ 1 cos − iσ1 sin
 2 2
cos ωt
2 −i sin ωt
2
= .
−i sin ωt2 cos ωt
2
(6.304)

Eyal Buks Quantum Mechanics - Lecture Notes 184


6.7. Solutions

b) Let P++ (t) be the probability to measure Sz = +/2 at time t > 0


given that at time t = 0 the spin was found to have Sz = +/2.
Similarly, P−− (t) is the probability to measure Sz = −/2 at time
t > 0 given that at time t = 0 the spin was found to have Sz = −/2.
These probabilities

are given by
 
2

  1

ωt

P++ (t) =
1 0 u (t) = cos2 , (6.305)
0
2

 
2

  0

ωt
P−− (t) =

0 1 u (t) = cos2 . (6.306)


1
2
Thus, assuming that the first measurement has yielded Sz = +/2
,  T -N
one finds psame = P++ N , whereas assuming that the first mea-
,  T -N
surement has yielded Sz = −/2 one finds psame = P−− N .
Thus in general independently on the result of the first measurement
one has
ωT
psame = cos2N . (6.307)
2N
c) Using   
ωT
psame = exp 2N log cos
2N
! !  2  4 ""
1 ωT 1
= exp 2N log 1 − +O
2 2N N
!  3 "
(ωT )2 1
= exp − +O ,
4N N
(6.308)
one finds

lim psame = 1 . (6.309)


N→∞

This somewhat surprising result is called the quantum Zeno effect or


the ’watched pot never boils’ effect.
17. The eigenvectors of S · û with eigenvalues ±/2 are given by
θ ϕ θ ϕ
|+; û = cos e−i 2 |+ + sin ei 2 |− , (6.310a)
2 2
θ ϕ θ ϕ
|−; û = − sin e−i 2 |+ + cos ei 2 |− , (6.310b)
2 2
where the states |± are eigenvectors of S · ẑ. Let P (σ3 , σ 2 |σ1 ) be the
probability to measure S · û =σ 2 (/2) in the second measurement and to
measure S ·ẑ =σ3 (/2) in the third measurement given that the result of
the first measurement was S · ẑ =σ1 (/2), and where σn ∈ {+, −}. The
following holds

Eyal Buks Quantum Mechanics - Lecture Notes 185


Chapter 6. Angular Momentum

2 2 θ
P (+, +|+) = |+|+; û| |+|+; û| = cos4 , (6.311a)
2
θ
P (+, −|+) = |+|−; û|2 |+|−; û|2 = sin 4 , (6.311b)
2
2 2 θ
P (−, −|−) = |−|−; û| |−|−; û| = cos4 , (6.311c)
2
2 2 θ
P (−, +|−) = |−|+; û| |−|+; û| = sin 4 , (6.311d)
2
thus independently on what was the result of the first measurement one
has
θ θ 1
psame = cos4 + sin 4 = 1 − sin2 θ . (6.312)
2 2 2
18. The Hamiltonian is given by
H = −µ · B . (6.313)
Using Eq. (4.38) for µz one has
d µz  1
= [µz , H]
dt i
γ2
= − Bx [Sz , Sx ] + By [Sz , Sy ]
i
2
= γ By Sx − Bx Sy 
= γ (µ × B) · ẑ .
(6.314)
Similar expressions are obtained for µx and µy that together can be
written in a vector form as
d
µ (t) = γ µ (t) × B (t) . (6.315)
dt
19. Using Eq. (6.137), which is given by
(σ · a) (σ · b) = a · b + iσ · (a × b) , (6.316)
one has
3  q 42  q 2
σ· p− A = p− A + iσ · ((p − qA) × (p − qA))
c c
 q 2 q
= p− A − i σ · (A × p + p × A) .
c c
(6.317)
The z component of the term (A × p + p × A) can be expressed as
(A × p + p × A) · ẑ = Ax py − Ay px + px Ay − py Ax
= [Ax , py ] − [Ay , px ] ,
(6.318)

Eyal Buks Quantum Mechanics - Lecture Notes 186


6.7. Solutions

thus, with the help of Eq. (3.76) one finds that


 
dAx dAy
(A × p + p × A) · ẑ = i − = −i (∇ × A) · ẑ . (6.319)
dy dx

Similar results can be obtained for the x and y components, thus


3  q 42  q 2 q
σ· p− A = p− A − σ·B. (6.320)
c c c
20. Since

j, m| Jx |j, m = j, m| Jy |j, m = 0 , (6.321)

and

Jx2 + Jy2 = J2 − Jz2 , (6.322)

one finds that


3 4
j, m| (∆Jx )2 + (∆Jy )2 |j, m = j, m| J2 |j, m − j, m| Jz2 |j, m
 
= 2 j 2 + j − m2 .
(6.323)
21. The condition is

p·A=A·p, (6.324)

or

[px , Ax ] + [py , Ay ] + [pz , Az ] = 0 , (6.325)

or using Eq. (3.76)

∂Ax ∂Ay ∂Az


+ + =0, (6.326)
∂x ∂y ∂z
or

∇·A=0 . (6.327)

22. The Hamiltonian is given by Eq. (1.62)


 2
p− qc A
H= + qϕ , (6.328)
2m
thus, the coordinates representation of H |α = E |α is given by

r′ | H |α = E r′ |α . (6.329)

Eyal Buks Quantum Mechanics - Lecture Notes 187


Chapter 6. Angular Momentum

Using the notation

r′ |α = ψ (r′ ) (6.330)

for the wavefunction together with Eqs. (3.23) and (3.29) one has

1  q 2
−i∇ − A + qϕ ψ (r′ ) = Eψ (r′ ) . (6.331)
2m c
23. The Hamiltonian is given by
 2  2
p− ec A p2x + p2z py − eBx
c
H= = +
2m 2m 2m
p2x 1  cp 2 p 2
y
= + mω 2c x − + z ,
2m 2 eB 2m
(6.332)
where
eB
ωc = . (6.333)
mc
Using the clue

ψ (x, y, z) = χ (x) exp (iky y) exp (ikz z) (6.334)

one finds that the time independent Schrödinger equation for the wave
function χ (x) is thus given by
  2   
p̂2x 1 2 cky 2 kz2
+ mω c x − χ (x) = E − χ (x) , (6.335)
2m 2 eB 2m

where p̂x = −i∂/∂x, thus the eigenenergies are given by


 
1 2 kz2
En,k = ω c n + + , (6.336)
2 2m
where n is integer and k is real, and the ground state energy is
ω c
En=0,k=0 = . (6.337)
2
24. Using the gauge A = Bxŷ the Hamiltonian is given by [see Eq. (1.62)]
 2
p− qc A
H= − qEx
2m
 2
qBx
2
p + pz 2 py − c
= x + − qEx .
2m 2m
(6.338)

Eyal Buks Quantum Mechanics - Lecture Notes 188


6.7. Solutions

The last two terms can be written as


 2
py − qBx
c p2y 1 3 4
2
− qEx = + mω 2c (x − x0 ) − x20 , (6.339)
2m 2m 2
where
qB
ωc = , (6.340)
mc
and
mc2  qpy 
x0 = qE + B . (6.341)
q2 B2 mc
Substituting the trial wavefunction

ψ (x, y, z) = ϕ (x) exp (iky y) exp (ikz z) , (6.342)

into the three dimensional Schrödinger equation yields a one dimensional


Schrödinger equation
 
p̂2x 1 2 2 1 2 2
2 ky2 + 2 kz2
+ mω c (x − x̃0 ) − mω c x̃0 + ϕ (x) = Eϕ (x) ,
2m 2 2 2m
(6.343)

where p̂x = −i∂/∂x and where


 
mc2 qky
x̃0 = 2 2 qE + B . (6.344)
q B mc

This equation describes a harmonic oscillator with a minimum potential


at x = x̃0 , with added constant terms that give rise to a shift in the
energy level, which are thus given by
 
1 1 2 ky2 + 2 kz2
En,ky ,kz = ωc n + − mω 2c x̃20 +
2 2 2m
  2 2
1 mc E cky E 2 kz2
= ωc n + − 2
− + ,
2 2B B 2m
(6.345)
where n = 0, 1, 2, · · · and where the momentum variables ky and kz can
take any real value.
25. The Schrödinger equation reads
 2 
p̂ − ec A
+ U (y) ψ (x, y) = Eψ (x, y) , (6.346)
2m

Eyal Buks Quantum Mechanics - Lecture Notes 189


Chapter 6. Angular Momentum

where

p̂ = −i∇ .

Employing the gauge A = −Byx̂ one has


 2 
p̂x + ec By p̂2y
+ + U (y) ψ (x, y) = Eψ (x, y) , (6.347)
2m 2m

where p̂x = −i∂/∂x and p̂y = −i∂/∂y. By substituting the trial wave-
function

ψ (x, y) = exp (ikx) χ (y) , (6.348)

one obtains a one dimensional Schrödinger equation for χ (y)


 e 2 
p̂2y c By + k 1 2 2
+ + mω 0 y χ (y) = Eχ (y) , (6.349)
2m 2m 2

or
 
p̂2y 2 k2 1 2 2 eBk
+ + mωc0 y − y χ (y) = Eχ (y) , (6.350)
2m 2m 2 mc

where ω 2c0 ≡ ω 2c + ω 20 and ω c = |e| B/mc. This can also be written as


  2 
p̂2y 1 2 eBk 2 k2 ω 20
+ mωc0 y − 2 2 + χ (y) = Eχ (y) . (6.351)
2m 2 m cω c0 2m ω2c0

This is basically a one-dimensional Schrödinger equation with a parabolic


potential of a harmonic oscillator and the eigenenergies are thus given
by:
 
1 2 k2 ω20
E (n, k) = ω c0 n + + ,
2 2m ω 2c0

where n = 0, 1, 2, · · · and k is real.


26. It is convenient to choose a gauge having cylindrical symmetry, namely
1
A=− r×B. (6.352)
2
For this gauge ∇·A = 0, thus according to Eq. (6.167) the Hamiltonian
is given by

1 2 q q2
H= p − p·A+ A2 . (6.353)
2µ µc 2µc2

Eyal Buks Quantum Mechanics - Lecture Notes 190


6.7. Solutions

The Schrödinger equation in cylindrical coordinates (ρ, z, φ) is given by


(note that A = (ρB/2) φ̂)
     2
2 1 ∂ ∂ψ 1 ∂2ψ ∂2ψ iqB ∂ψ q2 ρB
− ρ + 2 + 2 + + ψ = Eψ .
2µ ρ ∂ρ ∂ρ ρ ∂φ2 ∂z 2µc ∂φ 2µc2 2
(6.354)

The particle is constrained to move along the ring, which is located at


z = 0 and ρ = a, thus the effective one dimensional Schrödinger equation
of the system is given by
2 ∂ 2 ψ iqB ∂ψ q 2 a2 B 2
− + + ψ = Eψ . (6.355)
2µa2 ∂φ2 2µc ∂φ 8µc2

a) Consider a solution of the form


1
ψ (φ) = √ exp (imφ) , (6.356)
2πa

where the pre factor (2πa)−1/2 ensures normalization. The continu-


ity requirement that ψ (2π) = ψ (0) implies that m must be an inte-
ger. Substituting this solution into the Schrödinger equation (6.355)
yields
2 m2 qBm q 2 a2 B 2
Em = − +
2µa2 2µc 8µc2
!   "
2 2
2 qBa 2
1 qBa
= m2 − m+
2µa2 c 4 c
 2
2 qBa2
= m −
2µa2 2c
2
 2
 Φ
= m− ,
2µa2 Φ0
(6.357)
where

Φ = Bπa2 , (6.358)

is the magnetic flux threading the ring and


ch
Φ0 = . (6.359)
q
b) In general the current density is given by Eq. (4.200). For a wave-
function having the form

ψ (r) = α (r) eiβ(r) , (6.360)

Eyal Buks Quantum Mechanics - Lecture Notes 191


Chapter 6. Angular Momentum

where both α and β are real, one has


 q
J = Im [α (∇ (α) + α∇ (iβ))] − (ρA)
µ µc
α2 q 2
= ∇ (β) − α A
µ µc
2 
|ψ| q 
= ∇ (β) − A .
µ c
(6.361)
In the present case one has
ρB φ̂
A= , (6.362)
2
m
∇β = φ̂ , (6.363)
a
and the normalized wavefunctions are
1
ψm (φ) = √ exp (imφ) , (6.364)
2πa
thus
   
1 m q aB  Φ
Jm =  − φ̂ = m− φ̂ . (6.365)
2πaµ a c 2 2πa2 µ Φ0

Note that the following holds


c ∂Em
|Jm | = − . (6.366)
q ∂Φ
27. The Hamiltonian can be written as
L2 − L2z L2
H= + z
2Ixy 2Iz
 
L2 1 1
= + − L2z ,
2Ixy 2Iz 2Ixy
(6.367)
Thus the states |l, m (the standard eigenstates of L2 and Lz ) are eigen-
states of H and the following holds

H |l, m = El,m |l, m , (6.368)

where
  
l (l + 1) 1 1
El,m = 2 + − m2 . (6.369)
2Ixy 2Iz 2Ixy

Using the expression

Eyal Buks Quantum Mechanics - Lecture Notes 192


6.7. Solutions
0
3
Y1±1 (θ, φ) =∓ sin θe±iφ , (6.370)

one finds that
0
2π  −1 
sin θ cos φ = Y1 − Y11 , (6.371)
3
thus the normalized state at t = 0 can be written as
1
|ψ (0) = √ (|1, −1 − |1, 1) . (6.372)
2
Since E1,−1 = E1,1 the state |ψ (0) is stationary. Moreover
ψ (t)| Lz |ψ (t) = ψ (0)| Lz |ψ (0)
1
= ((1, −1| − 1, 1|)) Lz ((|1, −1 − |1, 1))
2
1
= ((1, −1| − 1, 1|)) ((− |1, −1 − |1, 1))
2
=0.
(6.373)
28. With the help of the relations
L+ + L−
Lx = , (6.374)
2
L+ |l, m =  l (l + 1) − m (m + 1) |l, m + 1 , (6.375)

L− |l, m =  l (l + 1) − m (m − 1) |l, m − 1 . (6.376)
one finds
a)
 
010
 
˙√
Lx = 1 0 1 . (6.377)
2 010

b)
  1 
010
 1 √1 1
 2
 1 0 1   √1  =  .
Lx  = √ 2 2 2 2
(6.378)
2 010 1
2

c)
   √1 
010 − 2
  √1 √1

Lx  = √ − 2 0 2
1 0 1 0  = 0 . (6.379)
2 010 √1
2

Eyal Buks Quantum Mechanics - Lecture Notes 193


Chapter 6. Angular Momentum

d)
 
  exp (−iφ) 0 0
iφLz
Dẑ (φ) = exp − ˙ 
= 0 1 0  . (6.380)
 0 0 exp (iφ)

e) In general
   
i (dφ) L · n̂ i (dφ) L · n̂
Dn̂ (dφ) = exp − =1− + O (dφ)2 ,
 
(6.381)

thus
 
1 − i(dφ)

2
0  
 i(dφ) 
˙ 
Dx̂ (dφ) = − 2
√ 1 − i(dφ)
√  + O (dφ)
2 
2
. (6.382)
0 − i(dφ)

2
1

29. Using
Lz = xpy − ypx , (6.383)
0
  
x= ax + a†x , (6.384)
2mω
0
  
y= ay + a†y , (6.385)
2mω
0
mω  
px = i −ax + a†x , (6.386)
2
0
mω  
py = i −ay + a†y , (6.387)
2
one finds
i ,     -
Lz = ax + a†x −ay + a†y − ay + a†y −ax + a†x
2 
= i ax a†y − a†x ay .
(6.388)
a) Thus
 
Lz  = i αx α∗y − α∗x αy . (6.389)

b) Using
, the†commutation
- relations
ax , ax = 1 , (6.390)
, -
ay , a†y = 1 , (6.391)
one finds

Eyal Buks Quantum Mechanics - Lecture Notes 194


6.7. Solutions
) 2*   
Lz = −2 αx , αy , αz | ax α∗y − α∗x ay αx a†y − a†x αy |αx , αy , αz 
3      2 4
2 2 2 2 2
= 2 |αx | 1 + |αy | + |αy | 1 + |αx | − αx α∗y − (α∗x ay ) ,
(6.392)
thus 3      2  2 4
(∆Lz )2 = 2 |αx |2 1 + |αy |2 + |αy |2 1 + |αx |2 + αx α∗y − α∗x αy − αx α∗y − (α∗x ay )2
3     4
2 2 2 2 2 2
= 2 |αx | 1 + |αy | + |αy | 1 + |αx | − 2 |αx | |αy |
 
2 2
= 2 |αx | + |αy | ,
(6.393)
and
.
∆Lz =  |αx |2 + |αy |2 . (6.394)

30. The normalization constant can be chosen to be A = 1/ 2. In general:
L+ + L−
Lx = , (6.395)
2
L+ |l, m =  l (l + 1) − m (m + 1) |l, m + 1 , (6.396)

L− |l, m =  l (l + 1) − m (m − 1) |l, m − 1 . (6.397)
a) The following holds
(L− |1, 1 − L+ |1, −1)
Lx |α = √
2 2
 (|1, 0 − |1, 0)
= =0,
2
(6.398)
thus

Lx  = 0 . (6.399)

b) Using Lx |α = 0 one finds


' ( ) *
(∆Lx )2 = L2x − Lx 2 = 0 − 0 = 0 . (6.400)

31. The Hamiltonian can be expressed as


L2 L2 L2 L2 L2
H= + z − z = + z , (6.401)
2I1 2I2 2I1 2I1 2Ie
where
I1 I2
Ie = . (6.402)
I1 − I2
Thus, the angular momentum states |l, m, which satisfy

Eyal Buks Quantum Mechanics - Lecture Notes 195


Chapter 6. Angular Momentum

L2 |l, m = l (l + 1) 2 |l, m , (6.403)


Lz |l, m = m |l, m , (6.404)
are eigenvector of H

H |l, m = El,m |l, m , (6.405)

where
 
l (l + 1) 2 m2 2 2 2 2 I1
El,m = + = l (l + 1) − m + m . (6.406)
2I1 2Ie 2I1 I2

a) Since [H, Lz ] = 0 one has


   
iLz φ iLz φ
exp H exp − =H, (6.407)
 

thus for the ground state l = m = 0

Az (φ) = ψ 0 | H |ψ0  = E0,0 = 0 . (6.408)

b) The operator Lx can be expressed as

L+ + L−
Lx = . (6.409)
2
In general 
L+ |l, m =  l (l + 1) − m (m + 1) |l, m + 1 , (6.410)

L− |l, m =  l (l + 1) − m (m − 1) |l, m − 1 , (6.411)
thus

L+ |0, 0 = L− |0, 0 = 0 , (6.412)

and consequently
 
iLx φ
exp − |ψ0  = |ψ0  , (6.413)


thus

Ax (φ) = ψ0 | H |ψ0  = E0,0 = 0 . (6.414)

32. The wavefunction of a point particle is given by

ψ (r) = (x + y + 2z) f (r) , (6.415)



where f (r) is a function of the radial coordinate r = x2 + y 2 + z 2 . As
can be see from Eqs. (6.131) and (6.132), which are given by

Eyal Buks Quantum Mechanics - Lecture Notes 196


6.7. Solutions
0
3
Y1±1 (θ, φ)
=∓ sin θe±iφ , (6.416)

0
0 3
Y1 (θ, φ) = cos θ . (6.417)

the following holds
0
2π  
x=r −Y11 + Y1−1 (6.418)
3
0
2π  1 
y = ir Y1 + Y1−1 (6.419)
3
0
4π 0
z=r Y . (6.420)
3 1
and thus
0 
π −1 + i 1 1 + i −1 0
ψ (r) = 2 √ Y1 + √ Y1 + 2Y1 rf (r) . (6.421)
3 2 2

a) In a measurement of L2 the only possible outcome is 22 .


b) In a measurement of Lz the outcome  and − have both probability
1/6, whereas the outcome 0 has probability 2/3.
33. The notation |η1 , η 2  is used to label the common eigenvectors of the
operator S1z , S2z , S21 and S22 , where η1 ∈ {+, −} and η2 ∈ {+, −}. The
following holds [see Eqs. (6.69) and (6.76)]

S1z |η1 , η2  = η1 |η1 , η2  , (6.422)
2

S2z |η1 , η2  = η2 |η1 , η2  , (6.423)
2
and
32
S21 |η1 , η2  = S22 |η1 , η2  = |η1 , η2  . (6.424)
4
a) The following holds

S2 = S21 + S22 + S1 · S2 + S2 · S1 . (6.425)

Any operator of the first particle commutes with any operator of the
second one thus
S2 = S21 + S22 + 2S1 · S2
= S21 + S22 + 2 (S1x S2x + S1y S2y + S1z S2z ) .
(6.426)
In terms of the operators S1± and S2± , which are related to S1x , S2x ,
S1y and S2y by

Eyal Buks Quantum Mechanics - Lecture Notes 197


Chapter 6. Angular Momentum

S1+ + S1− S1+ − S1−


S1x = , S1y = , (6.427)
2 2i
S2+ + S2− S2+ − S2−
S2x = , S2y = , (6.428)
2 2i
2
S is given by

S2 = S21 + S22 + S1+ S2− + S1− S2+ + 2S1z S2z . (6.429)

With
, 2 the- help
, Eqs. (6.24) and (6.41) one finds that -
S , Sz = S21 + S22 + S1+ S2− + S1− S2+ + 2S1z S2z , S1z + S2z
= [S1+ S2− + S1− S2+ , S1z + S2z ]
= [S1+ , S1z ] S2− + [S1− , S1z ] S2+ + S1+ [S2− , S2z ] + S1− [S2+ , S2z ]
=  (−S1+ S2− + S1− S2+ + S1+ S2− − S1− S2+ ) ,
(6.430)
thus
, 2 -
S , Sz = 0 . (6.431)

b) The following holds [see Eqs. (6.71) and (6.72)]


    
|+, + 2000 |+, +
 |+, −    
2  0 1 1 0   |+, − 
S2  
 |−, +  =   0 1 1 0   |−, +  , (6.432)
|−, − 0002 |−, −

and
    
|+, + 100 0 |+, +
 |+, −   0 0 0 0   |+, − 
Sz     
 |−, +  =   0 0 0 0   |−, +  . (6.433)
|−, − 0 0 0 −1 |−, −

It is thus easy to show that the following set of 4 ket vectors


|+, − − |−, +
|S = 0, M = 0 = √ , (6.434)
2
|S = 1, M = 1 = |+, + , (6.435)
|+, − + |−, +
|S = 1, M = 0 = √ , (6.436)
2
|S = 1, M = −1 = |−, − , (6.437)
forms the desired complete and orthonormal basis of common eigen-
vectors of S2 and Sz , and the following holds
S2 |S, M  = S (S + 1) 2 |S, M  , (6.438)
Sz |S, M  = M |S, M . (6.439)
34. The following holds [see Eqs. (6.32) and (6.36)]

Eyal Buks Quantum Mechanics - Lecture Notes 198


6.7. Solutions
 
sin θ e−iϕ J+ + eiϕ J−
Jn̂ = + cos θJz , (6.440)
2
thus [see Eqs. (6.63), (6.64), (6.65) and (6.66)]

j, m| Jn̂ |j, m = m cos θ , (6.441)

and [see Eqs. (6.43) and (6.44)]


sin2 θ (J+ J− + J− J+ )
j, m| Jn̂2 |j, m = j, m| + cos2 θJz2 |j, m
 4 
sin2 θ J2 − Jz2
= j, m| + cos2 θJz2 |j, m
 2 
2
 2

2 sin θ j (j + 1) − m 2 2
= + cos θm ,
2
(6.442)
thus the expectation value is given by Jn̂  = m cos θ and the variance
is given by
' ( j (j + 1) − m2
(∆Jn̂ )2 = 2 sin2 θ . (6.443)
2
35. Define the vector of operators Σ = (Σx , Σy , Σz ), where

a†2 − a2
Σx = , (6.444)
2
a†2 + a2
Σy = −i , (6.445)
2
aa† + a† a
Σz = . (6.446)
2
Using Eq. (5.13), which is given by
, †-
a, a = 1 , (6.447)

one finds that

[Σx , Σy ] = 2iΣz , (6.448)

[Σy , Σz ] = 2iΣx , (6.449)


[Σz , Σx ] = 2iΣy , (6.450)
thus

[Σi , Σj ] = 2iεijk Σk , (6.451)

where i, j, k ∈ {x, y, z}. The operator S (ξ, ϕ) (6.188) can be rewritten as

Eyal Buks Quantum Mechanics - Lecture Notes 199


Chapter 6. Angular Momentum
,  -
S (ξ, ϕ) = exp ξ eiϕ Σ+ + e−iϕ Σ− , (6.452)

where
1 a†2
Σ+ = (Σx + iΣy ) = , (6.453)
2 2
1 a2
Σ− = (Σx − iΣy ) = − . (6.454)
2 2
The vector of Pauli matrices σ = (σx , σy , σz ) satisfies a similar set of com-
mutation relations [σ i , σ j ] = 2iεijk σk as the set (6.451). Thus, all identi-
ties that are derived for the vector of Pauli matrices σ = (σx , σy , σz ) are
applicable for the vector Σ = (Σx , Σy , Σz ) provided that the derivation
uses only the commutation relations [σi , σj ] = 2iεijk σk . With the help
of the identity (6.138) one finds that the 2 × 2 matrix s (ξ, ϕ), which is
defined by [compare with Eq. (6.452)]
,  -
s (ξ, ϕ) = exp ξ eiϕ σ + + e−iϕ σ− , (6.455)

where
 
1 01
σ + = (σ x + iσy ) = , (6.456)
2 00
 
1 00
σ− = (σ x − iσy ) = , (6.457)
2 10

is given by
 
cosh ξ eiϕ sinh ξ
s (ξ, ϕ) = −iϕ . (6.458)
e sinh ξ cosh ξ

Furthermore, with the help of the following matrix identity


 
cosh ξ eiϕ sinh ξ
e−iϕ sinh ξ cosh ξ
 iϕ 
1 e tanh ξ
=
0 1
 − log(cosh ξ) 
e 0
×
0 elog(cosh ξ)
 
1 0
× ,
e−iϕ tanh ξ 1
(6.459)

and the relations σ2+ = σ2− = 0 one has

Eyal Buks Quantum Mechanics - Lecture Notes 200


6.7. Solutions

s (ξ, ϕ)
 
= exp eiϕ tanh ξσ+
× exp (− log (cosh ξ) σ z )
 
× exp e−iϕ tanh ξσ − .
(6.460)

The above expression for s (ξ, ϕ) yields a similar identity for the operator
S (ξ, ϕ)
 iϕ 
e †2
S (ξ, ϕ) = exp a tanh ξ
2
 
log (cosh ξ)  † 
× exp − aa + a† a
2
 −iϕ 
e
× exp − a2 tanh ξ .
2
(6.461)

36. With the help of Eqs. (5.332) and (5.103) one finds that
∞
1
Q (µ) = √ dx′
x0 πµ
−∞
 
x′2 √ x′ † a†2
× exp − 2 2 + 2 a −
2µ x µx0 2
 † 0
×: exp −a a :
 
x′2 √ x′ a2
× exp − 2 + 2 a − .
2x0 x0 2
(6.462)

Since the integrated function is in normal ordering the integration can


be performed while disregarding the nonvanishing commutation relation
between a and a† , namely by treating these operators as if they where
c-numbers
∞
1
Q (µ) = √ dx′
x0 πµ
−∞
   
√ †


1+ 12 x′2
µ
+
2 a+ a
µ
x′

(a+a† )2
2x2 x0 2
×: e 0 : ,
(6.463)

or with the help of the identity (5.136) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 201


Chapter 6. Angular Momentum
 
† 2 2
0 a+ a (a+a† )
µ
2µ −
1+ 12 2
Q (µ) = :e µ :
1 + µ2
0
2µ 1−µ2 a†2 −a2 (1−µ)2
− 1+µ2 aa†
= : e 1+µ2 2
: .
1 + µ2
(6.464)
Thus, using the notation
µ = e−ξ , (6.465)
and the identities
1 − e−2ξ
= tanh ξ , (6.466)
1 + e−2ξ
2e−ξ 1
= , (6.467)
1 + e−2ξ cosh ξ
 2
1 − e−ξ 1
=1− , (6.468)
1 + e−2ξ cosh ξ
one has
0
2µ 2
− tanh ξ a −a
†2
+( cosh
1 †
ξ −1)aa :
Q (µ) = : e 2
1 + µ2
0
1 †
e 2 a : e( cosh ξ −1)aa : e− 2 a .
tanh ξ †2 1 tanh ξ 2
=
cosh ξ
(6.469)
This can be further simplified with the help of Eq. (5.102)
0
1 tanh ξ †2 † tanh ξ 2
Q (µ) = e 2 a e− log(cosh ξ)a a e− 2 a .
cosh ξ
(6.470)
Using also
0
1 1
= e− 2 log(cosh ξ)
cosh ξ
and
1 aa† + a† a
a† a + =
2 2
one has
log(cosh ξ)
Q (µ) = e
tanh ξ †2
2 a
e− 2 (aa† +a† a) e− tanh
2
ξ 2
a
. (6.471)
The last result together with Eq. (6.190) leads to
 
S (ξ, 0) = Q e−ξ . (6.472)

Eyal Buks Quantum Mechanics - Lecture Notes 202


7. Central Potential

Consider a particle having mass m in a central potential, namely a potential


V (r) that depends only on the distance

r = x2 + y 2 + z 2 (7.1)
from the origin. The Hamiltonian is given by
p2
H= + V (r) . (7.2)
2m
Exercise 7.0.1. Show that
[H, Lz ] = 0 , (7.3)
, -
H, L2 = 0 . (7.4)
Solution 7.0.1. Using
[xi , pj ] = iδ ij , (7.5)
Lz = xpy − ypx , (7.6)
one has
, 2 - , - , - , -
p , Lz = p2x , Lz + p2y , Lz + p2z , Lz
, - , -
= p2x , xpy − p2y , ypx
= i (−2px py + 2py px )
=0,
(7.7)
and
, - , - , - , -
r2 , Lz = x2 , Lz + y 2 , Lz + z 2 , Lz
, - , -
= −y x2 , px + y 2 , py x
=0.
(7.8)
Thus Lz commutes with any smooth function of r2 , and consequently
[H, Lz ] = 0. In
, a similar
- way one can show that [H, Lx ] = [H, Ly ] = 0,
and therefore H, L2 = 0.
Chapter 7. Central Potential

In classical physics the corresponding Poisson’s brackets relations hold

{H, Lx } = {H, Ly } = {H, Lz } = 0 , (7.9)

and

H, L2 = 0 . (7.10)

These relations imply that classically the angular momentum is a constant of


the motion [see Eq. (1.40)]. On the other in quantum mechanics, as we have
seen in section 2.12 of chapter 2, the commutation relations
[H, Lz ] = 0 , (7.11)
, -
H, L2 = 0 , (7.12)
imply that it is possible to find a basis for the vector space made of common
eigenvectors of the operators H, L2 and Lz .

7.1 Simultaneous Diagonalization of the Operators H,


L2 and Lz

We start by proving some useful relations:

Exercise 7.1.1. Show that

L2 = r2 p2 − (r · p)2 + ir · p . (7.13)

Solution 7.1.1. The following holds


L2z = (xpy − ypx )2
= x2 p2y + y 2 p2x − xpy ypx − ypx xpy
= x2 p2y + y 2 p2x − xpx ([py , y] + ypy ) − ypy ([px , x] + xpx )
= x2 p2y + y 2 p2x − xpx ypy − ypy xpx + i (xpx + ypy ) .
(7.14)
Using the relation

xpx xpx = x ([px , x] + xpx ) px = −ixpx + x2 p2x , (7.15)

or

ixpx = x2 p2x − xpx xpx , (7.16)

one has

Eyal Buks Quantum Mechanics - Lecture Notes 204


7.1. Simultaneous Diagonalization of the Operators H, L2 and Lz

i
L2z = x2 p2y + y 2 p2x − xpx ypy − ypy xpx + (xpx + ypy )
2
1 
+ x2 p2x − xpx xpx + y 2 p2y − ypy ypy .
2
(7.17)
By cyclic permutation one obtains similar expression for L2x and for L2y . Com-
bining these expressions lead to
L2 = L2x + L2y + L2z
i 1 2 2 
= y 2 p2z + z 2 p2y − ypy zpz − zpz ypy +
(ypy + zpz ) + y py − ypy ypy + z 2 p2z − zpz zpz
2 2
2 2 2 2 i 1 2 2 
+z px + x pz − zpz xpx − xpx zpz + (zpz + xpx ) + z pz − zpz zpz + x2 p2x − xpx xpx
2 2
2 2 2 2 i 1 2 2 
+x py + y px − xpx ypy − ypy xpx + (xpx + ypy ) + x px − xpx xpx + y 2 p2y − ypy ypy
2 2
  
= x2 + y2 + z 2 p2x + p2y + p2z − (xpx + ypy + zpz )2 + i (xpx + ypy + zpz )
= r2 p2 − (r · p)2 + ir · p .
(7.18)
Exercise 7.1.2. Show that
 
′ 2 2 1 ∂2 ′ ′ 1 ′ 2
r | p |α = − r r |α − 2 ′2 r | L |α . (7.19)
r′ ∂r′2  r
Solution 7.1.2. Using the identities
L2 = r2 p2 − (r · p)2 + ir · p , (7.20)
r′ | r |α = r′ r′ |α , (7.21)
and

r′ | p |α = ∇ r′ |α , (7.22)
i
one finds that
r′ | L2 |α = r′ | r2 p2 |α − r′ | (r · p)2 |α + i r′ | r · p |α . (7.23)
The following hold

r′ | r · p |α = −ir′ · ∇ r′ |α = −ir′ r′ |α , (7.24)
∂r′
 2

r′ | (r · p)2 |α = −2 r′ ′ r′ |α
∂r
 
∂2 ∂
= −2 r′2 ′2 + r′ ′ r′ |α ,
∂r ∂r
(7.25)

Eyal Buks Quantum Mechanics - Lecture Notes 205


Chapter 7. Central Potential

r′ | r2 p2 |α = r′2 r′ | p2 |α , (7.26)


thus
  
∂2 2 ∂ 1
r′ | p2 |α = −2 + r′ |α − r′
| L2
|α , (7.27)
∂r′2 r′ ∂r′ 2 r′2
or
 
′ 2 2 1 ∂2 ′ ′ 1
r | p |α = − r r |α − 2 ′2 r′ | L2 |α . (7.28)
r′ ∂r′2  r
The time-independent Schrödinger equation in the coordinates represen-
tation

r′ | H |α = E r′ |α , (7.29)

where the Hamiltonian H is given by Eq. (7.2), can thus be written using the
above results as

′ −2 1 ∂ 2 ′ ′ 1
r | H |α = r r |α − 2 ′2 r | L |α +V (r′ ) r′ |α . (7.30)
′ 2
2m r′ ∂r′2  r

7.2 The Radial Equation


Consider a solution having the form
 
r′ |α = ϕ (r′ ) = R (r′ ) Ylm θ′ , φ′ . (7.31)

With the help of Eq. (6.107) one finds that

r′ | L2 |α = 2 l (l + 1) ϕ (r′ ) . (7.32)

Substituting into Eq. (7.30) yields an equation for R (r)



−2 1 d2 1
rR (r) − 2 l (l + 1) R (r) + V (r) R (r) = ER (r) . (7.33)
2m r dr2 r
The above equation, which is called the radial equation, depends on the quan-
tum number l, however, it is independent on the quantum number m. The
different solutions for a given l are labeled using the index k

−2 1 d2 1
rR kl − l (l + 1) Rkl + V Rkl = ERkl . (7.34)
2m r dr2 r2
It is convenient to introduce the function ukl (r), which is related to Rkl (r)
by the following relation
1
Rkl (r) = ukl (r) . (7.35)
r

Eyal Buks Quantum Mechanics - Lecture Notes 206


7.2. The Radial Equation

Substituting into Eq. (7.34) yields an equation for ukl (r)


 2 2 
− d
+ V eff (r) ukl (r) = Ekl ukl (r) , (7.36)
2m dr2

where the effective potential Veff (r) is given by

l (l + 1) 2
Veff (r) = + V (r) . (7.37)
2mr2
The total wave function is thus given by
1
ϕklm (r) = ukl (r) Ylm (θ, φ) . (7.38)
r
Since the spherical harmonic Ylm (θ, φ) is assumed to be normalized [see Eq.
(6.114)], to ensure that ϕklm (r) is normalized we require that
∞ ∞
2
1= 2
drr |Rkl (r)| = dr |ukl (r)|2 . (7.39)
0 0

In addition solutions with different k are expected to be orthogonal, thus


∞
dru∗k′ l (r) ukl (r) = δ kk′ . (7.40)
0

The wave functions ϕklm (r) represent common eigenstates of the operators
H, Lz and L2 , which are denoted as |klm and which satisfy the following
relations

ϕklm (r′ ) = r′ |klm , (7.41)

and
H |klm = Ekl |klm , (7.42)
L2 |klm = l (l + 1) 2 |klm , (7.43)
Lz |klm = m |klm . (7.44)
The following claim reveals an important property of the radial wavefunc-
tion near the origin (r = 0):

Claim. If the potential energy V (r) does not diverge more rapidly than 1/r
near the origin then

lim u (r) = 0 . (7.45)


r→0

Eyal Buks Quantum Mechanics - Lecture Notes 207


Chapter 7. Central Potential

Proof. Consider the case where near the origin u (r) has a dominant power
term having the form rs (namely, all other terms are of order higher than s,
and thus become negligibly small for sufficiently small r). Substituting into
Eq. (7.36) and keeping only the dominant terms (of lowest order in r) lead
to
−2 l (l + 1) 2 s−2
s (s − 1) rs−2 + r =0, (7.46)
2m 2m
thus s = −l or s = l + 1. However, the solution s = −l for l ≥ 1 must
be rejected since for this case the normalization condition (7.39) cannot be
satisfied as the integral diverges near r = 0. Moreover, also for l = 0 the
solution s = −l must be rejected. For this case ϕ (r) ≃ 1/r near the origin,
however, such a solution contradicts Eq. (7.30), which can be written as

2 2
− ∇ ϕ (r) + V (r) ϕ (r) = Eϕ (r) . (7.47)
2m
since
1
∇2 = −4πδ (r) . (7.48)
r
We thus conclude that only the solution s = l + 1 is acceptable, and conse-
quently limr→0 u (r) = 0.

7.3 Hydrogen Atom

The hydrogen atom is made of two particles, an electron and a proton. It


is convenient to employ the center of mass coordinates system. As is shown
below, in this reference frame the two body problem is reduced into a central
potential problem of effectively a single particle.

Exercise 7.3.1. Consider two point particles having mass m1 and m2 re-
spectively. The potential energy V (r) depends only on the relative coordi-
nate r = r1 − r2 . Show that the Hamiltonian of the system in the center of
mass frame is given by

p2
H= + V (r) , (7.49)

where the reduced mass µ is given by
m1 m2
µ= . (7.50)
m1 + m2

Eyal Buks Quantum Mechanics - Lecture Notes 208


7.3. Hydrogen Atom

Solution 7.3.1. The Lagrangian is given by


m1 ṙ21 m2 ṙ22
L= + − V (r1 − r2 ) . (7.51)
2 2
In terms of center of mass r0 and relative r coordinates, which are given by
m1 r1 + m2 r2
r0 = , (7.52)
m1 + m2
r = r1 − r2 , (7.53)
the Lagrangian is given by
 2  2
m1 ṙ0 + m1m+m2
2
ṙ m2 ṙ0 − m1
m1 +m2 ṙ
L= + − V (r)
2 2
M ṙ20 µṙ2
= + − V (r) ,
2 2
(7.54)
where the total mass M is given by

M = m1 + m2 , (7.55)

and the reduced mass by


m1 m22 + m2 m21 m1 m2
µ= 2 = . (7.56)
(m1 + m2 ) m1 + m2

Note that the Euler Lagrange equation for the coordinate r0 yields that r̈0 = 0
(since the potential is independent on r0 ). In the center of mass frame r0 = 0.
The momentum canonically conjugate to r is given by
∂L
p= . (7.57)
∂ ṙ
Thus the Hamiltonian is given by
p2
H = p · ṙ−L = + V (r) . (7.58)

For the case of hydrogen atom the potential between the electron having
charge −e and the proton having charge e is given by
e2
V (r) = − . (7.59)
r
Since the proton’s mass mp is significantly larger than the electron’s mass
me (mp ≃ 1800me ) the reduced mass is very close to me
me mp
µ= ≃ me . (7.60)
me + mp

Eyal Buks Quantum Mechanics - Lecture Notes 209


Chapter 7. Central Potential

The radial equation (7.36) for the present case is given by


 2 2 
− d
+ V eff (r) ukl (r) = Ekl ukl (r) , (7.61)
2µ dr2
where
e2 l (l + 1) 2
Veff (r) = − + . (7.62)
r 2µr2
In terms of the dimensionless radial coordinate
r
ρ= , (7.63)
a0
where
2
a0 = = 0.53 × 10−10 m , (7.64)
µe2
is the Bohr’s radius, and in terms of the dimensionless parameter
0
Ekl
λkl = − , (7.65)
EI
where
µe4
EI = = 13.6 eV , (7.66)
22
is the ionization energy, the radial equation becomes
 
d2
− 2 + Vl (ρ) + λ2kl ukl = 0 (7.67)

where
2 l (l + 1)
Vl (ρ) = − + . (7.68)
ρ ρ2

10
8
6
4
2

0 1 2 3 4 5
-2
-4
-6
-8
-10

The function Vl (ρ) for l = 0 (solid line) and l = 1 (dashed line).

Eyal Buks Quantum Mechanics - Lecture Notes 210


7.3. Hydrogen Atom

We seek solutions of Eq. (7.67) that represent bound states, for which Ekl
is negative, and thus λkl is a nonvanishing real positive. In the limit ρ → ∞
the potential Vl (ρ) → 0, and thus it becomes negligibly small in comparison
with λkl [see Eq. (7.67)]. Therefore, in this limit the solution is expected to
be asymptotically proportional to e±λkl ρ . To ensure that the solution is nor-
malizable the exponentially diverging solution e+λkl ρ is excluded. Moreover,
as we have seen above, for small ρ the solution is expected to be proportional
to ρl+1 . Due to these considerations we express ukl (r) as

ukl (r) = y (ρ) ρl+1 e−λkl ρ . (7.69)

Substituting into Eq. (7.67) yields an equation for the function y (ρ)
2   
d (l + 1) d 2 (1 − λkl (l + 1))
+ 2 − λ kl + y=0. (7.70)
dρ2 ρ dρ ρ

Consider a power series expansion of the function y (ρ)



<
y (ρ) = cq ρq . (7.71)
q=0

Substituting into Eq. (7.70) yields



< ∞
<
q (q − 1) cq ρq−2 + 2 (l + 1) qcq ρq−2
q=0 q=0

< ∞
<
−2λkl qcq ρq−1 + 2 (1 − λkl (l + 1)) cq ρq−1 = 0 ,
q=0 q=0

(7.72)

thus
cq 2 [λkl (q + l) − 1]
= . (7.73)
cq−1 q (q + 2l + 1)

We argue below that for physically acceptable solutions y (ρ) must be a poly-
nomial function [i.e. the series (7.71) needs to be finite]. To see this note that
for large q Eq. (7.73) implies that
cq 2λkl
lim = . (7.74)
q→∞ cq−1 q
Similar recursion relation holds for the coefficients of the power series expan-
sion of the function e2λkl ρ

<
e2λkl ρ = c̃q ρq , (7.75)
q=0

where

Eyal Buks Quantum Mechanics - Lecture Notes 211


Chapter 7. Central Potential
q
(2λkl )
c̃q = , (7.76)
q!
thus
c̃q 2λkl
= . (7.77)
c̃q−1 q
This observation suggests that for large ρ the function ukl asymptotically
becomes proportional to eλkl ρ . However, such an exponentially diverging so-
lution must be excluded since it cannot be normalized. Therefore, to avoid
such a discrepancy, we require that y (ρ) must be a polynomial function.
As can be see from Eq. (7.73), this requirement is satisfied provided that
λkl (q + l) − 1 = 0 for some q. A polynomial function of order k − 1 is ob-
tained when λkl is taken to be given by
1
λkl = , (7.78)
k+l
where k = 1, 2, 3, · · · . With the help of Eq. (7.73) the polynomial function
can be evaluated. Some examples are given below
yk=1,l=0 (ρ) = c0 , (7.79)
yk=1,l=1 (ρ) = c0 , (7.80)
 ρ
yk=2,l=0 (ρ) = c0 1 − , (7.81)
 2
ρ
yk=2,l=1 (ρ) = c0 1 − . (7.82)
6
The coefficient c0 can be determined from the normalization condition.
As can be seen from Eqs. (7.65) and (7.78), all states having the same
sum k + l, which is denoted as

n=k+l , (7.83)

have the same energy. The index n is called the principle quantum number.
Due to this degeneracy, which is sometimes called accidental degeneracy, it
is common to label the states with the indices n, l and m, instead of k, l and
m. In such labeling the eigenenergies are given by
EI
En = − , (7.84)
n2
where

n = 1, 2, · · · . (7.85)

For a given n the quantum number l can take any of the possible values

l = 0, 1, 2, · · · , n − 1 , (7.86)

Eyal Buks Quantum Mechanics - Lecture Notes 212


7.4. Problems

and the quantum number m can take any of the possible values

m = −l, −l + 1, · · · , l − 1, l . (7.87)

The level of degeneracy of the level En is given by


n−1
  
2 (n − 1) n
gn = 2 (2l + 1) = 2 + n = 2n2 . (7.88)
2
l=0

Note that the factor of 2 is due to spin. The normalized radial wave functions
of the states with n = 1 and n = 2 are found to be given by
 3/2
1
R10 (r) = 2 e−r/a0 , (7.89)
a0
  3/2
r 1 r
R20 (r) = 2 − e− 2a0 , (7.90)
a0 2a0
 3/2
1 r r
R21 (r) = √ e− 2a0 . (7.91)
2a0 3a0
The wavefunction ϕn,l,m (r) of an eigenstate with quantum numbers n, l and
m is given by

ϕnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) . (7.92)

While the index n labels the shell number, the index l labels the sub-shell.
In spectroscopy it is common to label different sub-shells with letters:

l=0 s
l=1p
l=2d
l=3 f
l=4g

7.4 Problems
1. Consider the wave function with quantum numbers n, l, and m of a
hydrogen atom ϕn,l,m (r).
a) Show that the probability current in spherical coordinates r, θ, ϕ is
given by


2

ϕn,l,m (r)

Jn,l,m (r) = m φ̂ , (7.93)


µ r sin θ

where µ is the reduced mass and φ̂ is a unit vector orthogonal to ẑ


and r̂.

Eyal Buks Quantum Mechanics - Lecture Notes 213


Chapter 7. Central Potential

b) Use the result of the previous section to show that the total angular
momentum expectation value is given by L = mẑ.
2. Show that the average electrostatic potential in the neighborhood of an
hydrogen atom in its ground state is given by
   
1 1 2r
ϕ=e + exp − , (7.94)
a0 r a0
where a0 is the Bohr radius.
3. An hydrogen atom is in its ground state. The distance r between the
electron and the proton is measured. Calculate the expectation value r
and the most probable value r0 (at which the probability distribution
function obtains a maximum).
4. Tritium, which is labeled as 3 H, is a radioactive isotope of hydrogen.
The nucleus of tritium contains 1 proton and two neutrons. An atom
of tritium is in its ground state, when the nucleus suddenly decays into
a helium nucleus, with the emission of a fast electron, which leaves the
atom without perturbing the extra-nuclear electron. Find the probability
that the resulting He+ ion will be left in:
a) 1s state.
b) 2s state.
c) a state with l = 0.
5. At time t = 0 an hydrogen atom is in the state

|α (t = 0) = A (|2, 1, −1 + |2, 1, 1) ,

where A is a normalization constant and where |n, l, m denotes the eigen-


state with quantum numbers n, l and m. Calculate the expectation value
x at time t.
6. Find the ground state energy E0 of a particle having mass m in a central
potential V (r) given by
+
0 a≤r≤b
V (r) = , (7.95)
∞ else

where r = x2 + y2 + z 2 .
7. Consider a particle having mass m in a 3D potential given by

V (r) = −Aδ (r − a) , (7.96)



where r = x2 + y 2 + z 2 is the radial coordinate, the length a is a
constant and δ () is the delta function. For what range of values of the
constant A the particle has a bound state.
8. Consider a particle having mass m in a 3D central potential given by
+
−U0 r ≤ r0
V (r) = . (7.97)
0 r > r0

Eyal Buks Quantum Mechanics - Lecture Notes 214


7.5. Solutions

where r = x2 + y 2 + z 2 is the radial coordinate, U0 is real and r0 is
positive. For what range of values of the potential depth U0 the particle
has a bound state.
9. A spinless point particle is in state |γ. The state vector |γ is an eigen-
vector of the operators Lx , Ly and Lz (the x, y and z components of
the angular momentum vector operator). What can be said about the
wavefunction ψ (r′ ) of the state |γ?
10. Consider two (non-identical) particles having the same mass m moving
under the influence of a potential U (r), which is given by
1
U (r) = mω 2 r2 .
2
In addition, the particles interact with each other via a potential given
by
1
V (r1 , r2 ) = mΩ 2 (r1 − r2 )2 ,
2
where r1 and r2 are the (three dimensional) coordinate vectors of the first
and second particle respectively. Find the eigenenergies of the system.

7.5 Solutions
1. In general the current density is given by Eq. (4.200). For a wavefunction
having the form

ψ (r) = α (r) eiβ(r) , (7.98)

where both α and β are real, one has



J= Im [α (∇α + iα∇β)]
µ
α2
= ∇β
µ
 |ψ|2
= ∇β .
µ
(7.99)
a) The wavefunction ϕn,l,m (r) is given by

ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) = Rnl (r) Flm (θ) eimφ , (7.100)

where both Rnl and Flm are real, thus




2

ϕn,l,m (r)

Jn,l,m (r) = ∇ (mφ) . (7.101)


µ

Eyal Buks Quantum Mechanics - Lecture Notes 215


Chapter 7. Central Potential

In spherical coordinates one has


∂ ∂ ∂
∇ = x̂ + ŷ + ẑ
∂x ∂y ∂z
∂ 1 ∂ 1 ∂
= r̂ + θ̂ + φ̂ ,
∂r r ∂θ r sin θ ∂φ
(7.102)
thus


2

ϕn,l,m (r)

Jn,l,m (r) = m φ̂ . (7.103)


µ r sin θ
b) The contribution of the volume element d3 r to the angular momen-
tum with respect to the origin is given by µr × Jn,l,m (r) d3 r. In
spherical coordinates the total angular momentum is given by
 


2
3
ϕn,l,m (r)

L = µr×Jn,l,m (r) d r = m r× ϕ̂d3 r . (7.104)


r sin θ
By symmetry, only the component along ẑ of r × ϕ̂ contributes, thus
L = mẑ . (7.105)
2. The charge density of the electron in the ground state is given by
 


2 e 2r

ρ = −e ϕ1,0,0 (r) = − 3 exp − . (7.106)


πa0 a0
The Poisson’s equation is given by
∇2 ϕ = −4πρ . (7.107)
To verify that the electrostatic potential given by Eq. (7.94) solves this
equation we calculate
1 d2
∇2 ϕ = (rϕ)
r dr2    
2
e d r 2r
= + 1 exp −
r dr2 a0 a0
 
2r
4e exp − a0
=
a30
= −4πρ .
(7.108)
Note also that
lim ϕ (r) = 0 , (7.109)
r→∞

as is required for a neutral atom.

Eyal Buks Quantum Mechanics - Lecture Notes 216


7.5. Solutions

3. The radial wave function of the ground state is given by


 3/2  
1 r
R10 (r) = 2 exp − (7.110)
a0 a0
thus the probability distribution function of the variable r is given by
 3  
2 4 r 2r
f (r) = |rR10 (r)| = exp − . (7.111)
r a0 a0
Thus
 ∞  ∞
3
r = rf (r) dr = 4a0 x3 exp (−2x) dx = a0 . (7.112)
0 0 2
The most probable value r0 is found from the condition
 
df 8r0 2r0
0= = 4 exp − (a0 − r0 ) , (7.113)
dr a0 a0
thus

r0 = a0 . (7.114)

4. The radial wave function of a hydrogen-like atom with a nucleus having


charge Ze is found by substituting e2 by Ze2 in Eqs. (7.89), (7.90) and
(7.91), namely
 3/2
(Z) Z
R10 (r) = 2 e−Zr/a0 , (7.115)
a0
 3/2
(Z) Z Zr
R20 (r) = (2 − Zr/a0 ) e− 2a0 , (7.116)
2a0
 3/2
(Z) Z Zr − 2aZr
R21 (r) = √ e 0 . (7.117)
2a0 3a0
The change in reduced mass is neglected. Therefore
a) For the 1s state
∞ 2  
 7 2a3 2
2 (Z=1) (Z=2) 2 0
Pr (1s) =  drr R10 R10  = 3 = 0.702 .
a0 33
0

b) For the 2s state


∞ 2
  2
(Z=1) (Z=2) 16 a30
Pr (2s) =  drr2 R10 R20  = 6 (2 − 3) = 0.25 .
a0 8
0

Eyal Buks Quantum Mechanics - Lecture Notes 217


Chapter 7. Central Potential

c) For this case the probability vanishes due to the orthogonality be-
tween spherical harmonics with different l.

5. The normalization constant is chosen to be A = 1/ 2. Since both states
|2, 1, −1 and |2, 1, 1 have the same energy the state |α is stationary.
The following holds
ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (7.118a)
 3/2
1 r r
R21 (r) = √ e− 2a0 , (7.118b)
2a0 3a0
0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (7.118c)
2 2π
0
1 1 3
Y1 (θ, φ) = − sin θeiφ , (7.118d)
2 2π
x = r sin θ cos φ . (7.118e)
In general

∞ 1 2π  ′ ∗
′ ′ ′ 3
n l m | x |nlm = dr r R n′ l′ Rnl d (cos θ) dφ sin θ cos φ Ylm
′ Ylm .
0 −1 0
(7.119)

thus
2π
2, 1, 1|x|2, 1, 1 ∝ dφ cos φ = 0 , (7.120)
0
2π
2, 1, −1|x|2, 1, −1 ∝ dφ cos φ = 0 , (7.121)
0
2π
2, 1, 1|x|2, 1, −1 ∝ dφ cos φe−2iφ = 0 , (7.122)
0
2π
2, 1, −1|x|2, 1, 1 ∝ dφ cos φe2iφ = 0 , (7.123)
0

and therefore

x (t) = 0 . (7.124)

6. The radial equation is given by [see Eq. (7.36)]



2 d2 l (l + 1) 2
− + + V (r) uk,l (r) = Ek,l uk,l (r) . (7.125)
2m dr2 2mr2

Eyal Buks Quantum Mechanics - Lecture Notes 218


7.5. Solutions

Since the centrifugal term l (l + 1) 2 /2mr2 is non-negative the ground


state is obtained with l = 0. Thus the ground state energy is [see Eq.
(4.181)]
π 2 2
E0 = . (7.126)
2m (b − a)2
7. The radial equation is given by

2 d2 l (l + 1) 2
− + + V (r) uk,l (r) = Ek,l uk,l (r) . (7.127)
2m dr2 2mr2
The boundary conditions imposed upon u (r) by the potential are
u (0) = 0 , (7.128)
   
u a+ = u a− (7.129)
du (a+ ) du (a− ) 2
− = − u (a) . (7.130)
dr dr a0
where
2
a0 = . (7.131)
mA
Since the centrifugal term l (l + 1) 2 /2mr2 is non-negative the ground
state is obtained with l = 0. We seek a solution for that case having the
form
+
sinh (κr) r<a
u (r) = , (7.132)
sinh (κa) exp (−κ (r − a)) r > a
where

−2mE
κ= . (7.133)

The condition (7.130) yields
2
−κ sinh (κa) − κ cosh (κa) = − sinh (κa) , (7.134)
a0
or
κa0 1
= .
2 1 + coth (κa)
A real solution exists only if
a0
<a, (7.135)
2
or
2
A> . (7.136)
2ma

Eyal Buks Quantum Mechanics - Lecture Notes 219


Chapter 7. Central Potential

8. The radial equation is given by



2 d2 l (l + 1) 2
− + + V (r) uk,l (r) = Ek,l uk,l (r) . (7.137)
2m dr2 2mr2

The boundary condition that is imposed upon u (r) at the origin is u (0) =
0. Since the centrifugal term l (l + 1) 2 /2mr2 is non-negative the ground
state is obtained with l = 0. For that case the solution in the range r ≤ r0
has the form u (r) = sin kr, where k is related to the energy E by

h2 k2
= E + U0 . (7.138)
2m
In the range r > r0 the general solution has the form u (r) = Ae−κr +
Beκr , where

h2 κ2
= −E . (7.139)
2m
A bound state can be obtained provided that E < 0 (to ensure that κ
is real) and B = 0 (to ensure that limr→∞ u (r) = 0; it is assumed that
κ is non-negative). The requirements that both u (r) and du/dr [see Eq.
(4.127)] are continuous at r = r0 yield (for the case B = 0)
sin kr0 = Ae−κr0 , (7.140)
k
cos kr0 = −Ae−κr0 , (7.141)
κ
thus the following must hold
k
tan kr0 = − . (7.142)
κ
Since both k and κ are required to be nonnegative, the above condition
can be satisfied only if tan kr0 ≤ 0, which implies that
0
2m (E + U0 ) π
kr0 = r0 ≥ . (7.143)
h2 2
This together with the requirement that E < 0 yield
0
2mU0 π
r0 ≥ , (7.144)
h2 2
or
π2 h2
U0 ≥ . (7.145)
8mr02

Eyal Buks Quantum Mechanics - Lecture Notes 220


7.5. Solutions

9. The state vector |γ is an eigenvector of the operators Lx , Ly , therefore it


is easy to see that it consequently must be an eigenvector of the operator
[Lx , Ly ] with a zero eigenvalue. Thus, since [Lx , Ly ] = iLz , one has
Lz |γ = 0. Similarly, one finds that Lx |γ = Ly |γ = 0. Therefore, |γ
is also an eigenvector of the operator L2 = L2x + L2y + L2z with a zero
eigenvalue. Therefore the wavefunction has the form
 ′ ′  R (r′ )
ψ (r′ ) = R (r′ ) Yl=0
m=0
θ ,φ = √ , (7.146)

where the radial function R (r′ ) is an arbitrary normalized function.
10. The Lagrangian is given by
 
m ṙ21 + ṙ22 1   1
L= − mω2 r21 + r22 − mΩ 2 (r1 − r2 )2 . (7.147)
2 2 2
In terms of center of mass r0 and relative r coordinates, which are given
by
r1 + r2
r0 = , (7.148)
2
r = r1 − r2 , (7.149)
the Lagrangian is given by
3 2  2 4
m ṙ0 + 12 ṙ + ṙ0 − 12 ṙ
L=
 2 2  2 
1 2 1 1 1
− mω r0 + r + r0 − r − mΩ 2 r2
2 2 2 2
 2 1 2  
m 2ṙ0 + 2 ṙ 1 1 1
= − mω 2 2r20 + r2 − mΩ 2 r2
2 2 2 2
2 2  
M ṙ0 1 µṙ 1
= − M ω2 r20 + − µ ω 2 + 2Ω 2 r2 ,
2 2 2 2
(7.150)
where the total mass M is given by

M = 2m , (7.151)

and the reduced mass [see also Eq. (7.50)] by


m
µ= . (7.152)
2
The Lagrangian L describes two decoupled three dimensional harmonic
oscillators. The first, which is associated with the center of mass motion,
has mass M = 2m and angular resonance frequency ω, whereas the

Eyal Buks Quantum Mechanics - Lecture Notes 221


Chapter 7. Central Potential

second one, which is associated with the relative √ coordinate r, has mass
µ = m/2 and angular resonance frequency ω2 + 2Ω 2 . The quantum
energy eigenvectors are denoted by |n0x , n0y , n0z , nx , ny , nz , where all
six quantum numbers n0x , n0y , n0z , nx , ny and nz are integers, and the
corresponding eigenenergies are given by
 
3
En0x ,n0y ,n0z ,nx ,ny ,nz = ω + n0x + n0y + n0z
2
  
2 2
3
+ ω + 2Ω + nx + ny + nz .
2
(7.153)

Eyal Buks Quantum Mechanics - Lecture Notes 222


8. Density Operator

Consider an ensemble of N identical copies of a quantum system. The en-


semble can be divided into subsets, where all systems belonging to the same
subset have the

same
* state vector. Let N wi be the number of systems having
state vector
α(i) , where

0 ≤ wi ≤ 1 , (8.1)

and where

wi = 1 . (8.2)
i

The state vectors are all assumed to be normalized


'
(

α(i)
α(i) = 1 . (8.3)

Consider a measurement of an observable A, having a set of eigenvalues


{an } and corresponding set of eigenvectors {|an }

A |an  = an |an  . (8.4)

The set of eigenvectors {|an } is assumed to be orthonormal and complete


am |an  = δ nm , (8.5)

|an  an | = 1 . (8.6)
n

Consider a measurement of the observable A done on a system that is ran-


domly chosen
from
* the ensemble. The probability to choose
* a system having
state vector
α(i)
is w
i . Given that the state vector is
α(i) , the expectation
) *
value of A is α(i)
A
α(i) [see Eq. (2.84)]. Thus, the expectation (average)
value of such a measurement done on a system that is randomly chosen from
the ensemble is given by
 '

( 

(
2




A = wi α(i)
A
α(i) = wi
an
α(i)
an . (8.7)
i i n
Chapter 8. Density Operator

Claim. The expectation value can be expressed as

A = Tr (ρA) , (8.8)

where

('

ρ= wi
α(i) α(i)
(8.9)
i

is the density operator.

Proof. Let {|bm } be an orthonormal and complete basis for the vector space

|bm  bm | = 1 . (8.10)
m

The following holds


 '

(

A = wi α(i)
A
α(i)
i
 '

(

= wi α(i)
A |bm  bm
α(i)
i m
 
('

= bm | wi
α(i) α(i)
A |bm 
m i
= Tr (ρA) ,
(8.11)
where

('

ρ= wi
α(i) α(i)
.
i

Below we discuss some basic properties of the density operator:

Exercise 8.0.1. Show that ρ† = ρ.

Solution 8.0.1. Trivial by the definition (8.9).

Exercise 8.0.2. Show that Tr (ρ) = 1.


<
Solution 8.0.2. Using a complete orthonormal basis m |bm  bm | = 1 one
has

Eyal Buks Quantum Mechanics - Lecture Notes 224


! "
 
('


(i) (i)

Tr (ρ) = bm | wi
α α
|bm 
m i
! "
 '

(
(i)

= wi α
|bm  bm |
α(i)
i m
 '
(

= wi α(i)
α(i)
i

= wi
i
=1.
(8.12)

Exercise 8.0.3. Show that for any normalized state |β the following holds

0 ≤ β| ρ |β ≤ 1 . (8.13)

Solution 8.0.3. Clearly, 0 ≤ β| ρ |β since



(' 

'
2

β| ρ |β = wi β
α(i) α(i) |β = wi
α(i) |β
≥ 0 . (8.14)
i i

On the other hand, according to the Schwartz inequality [see Eq. (2.166)],
which is given by
 
|u |v| ≤ u |u v |v , (8.15)

one has

'
 .)
*

(i)


α |β
≤ β |β α(i)
α(i) = 1 . (8.16)
<
Moreover, i wi = 1, thus

'
2

β| ρ |β = wi
α(i) |β
≤ 1 . (8.17)
i
 
Exercise 8.0.4. Show that Tr ρ2 ≤ 1.

Solution 8.0.4. The fact that ρ is Hermitian (i.e., ρ† = ρ) guaranties the


existence of a complete orthonormal basis {|qm } of eigenvectors of ρ, which
satisfy
qm′ |qm  = δ mm′ , (8.18)

|qm  qm | = 1 , (8.19)
m

and

Eyal Buks Quantum Mechanics - Lecture Notes 225


Chapter 8. Density Operator

ρ |qm  = qm |qm  , (8.20)


where the eigenvalues qm are real. Using this basis one has
   
Tr ρ2 = qm | ρ2 |qm  = 2
qm . (8.21)
m m

According to inequality (8.13)


0 ≤ qm = qm | ρ |qm  ≤ 1 , (8.22)
thus
! "2
   2 
Tr ρ2 = qm ≤ qm = (Tr (ρ))2 = 1 . (8.23)
m m

Definition 8.0.1. An ensemble is said to be pure if its density operator can


be expressed as
ρ = |α α| .
 
Exercise 8.0.5. Show that Tr ρ2 = 1 iff ρ represents a pure ensemble.
Solution 8.0.5. (i)Assuming that ρ represents a pure ensemble,
  one has
ρ2 = ρ, thus Tr ρ2 = Tr (ρ) = 1. (ii) Assume that Tr ρ2 = 1. Since ρ
is Hermitian (i.e., ρ† = ρ), there is a complete orthonormal basis {|qm } of
eigenvectors of ρ, such that
qm′ |qm  = δ mm′ , (8.24)

|qm  qm | = 1 , (8.25)
m
and
ρ |qm  = qm |qm  , (8.26)
where the eigenvalues qm are real. Moreover, according to inequality (8.22)
0 ≤ qm ≤ 1 . (8.27)
 
For this basis the assumption Tr ρ2 = 1 yields
   2
1 = Tr ρ2 = qm . (8.28)
m

Moreover, also Tr (ρ) = 1, thus



1= qm . (8.29)
m

Both equalities can be simultaneously satisfied only if


+
1 m = m0
qm = . (8.30)
0 m = m0
For this case ρ = |qm0  qm0 |, thus ρ represents a pure ensemble.

Eyal Buks Quantum Mechanics - Lecture Notes 226


8.2. Quantum Statistical Mechanics

8.1 Time Evolution

Consider a density operator




('

ρ (t) = wi
α(i) (t) α(i) (t)
, (8.31)
i

*
where the state vectors
α(i) (t) evolve in time according to

*
(
d
α(i)

i = H
α(i) , (8.32)
)dt(i)
'

d α

−i = α(i)
H , (8.33)
dt
where H is the Hamiltonian. Taking the time derivative yields
! "
dρ 1 
('

('


(i) (i)

(i) (i)

= wi H
α α (t)
− wi
α (t) α
H , (8.34)
dt i i i

thus
dρ 1
= − [ρ, H] . (8.35)
dt i
This result resembles the equation of motion (4.37) of an observable in the
Heisenberg representation, however, instead of a minus sign on the right hand
side, Eq. (4.37) has a plus sign.
Alternatively, the time evolution of the operator ρ can be expressed in
terms of
the time * evolution operator

u (t,
* t0 ), which relates the state vector
at time
α(i) (t0 ) with its value
α(i) (t) at time t [see Eq. (4.4)]

(
(

(i)


α (t) = u (t, t0 )
α(i) (t0 ) . (8.36)

With the help of this relation Eq. (8.31) becomes

ρ (t) = u (t, t0 ) ρ (t0 ) u† (t, t0 ) . (8.37)

8.2 Quantum Statistical Mechanics

Consider an ensemble of identical copies of a quantum system. Let H be the


Hamiltonian having a set of eigenenergies {Ei } and a corresponding set of
eigenstates {|i}, which forms an orthonormal and complete basis
H |i = Ei |i , (8.38)

|i i| = 1 . (8.39)
i

Eyal Buks Quantum Mechanics - Lecture Notes 227


Chapter 8. Density Operator

Consider the case where the ensemble is assumed to be a canonical ensemble


in thermal equilibrium at temperature T . According to the laws of statistical
mechanics the probability wi to find an arbitrary system in the ensemble in
a state vector |i having energy Ei is given by
1 −βEi
wi = e , (8.40)
Z
where β = 1/kB T , kB is Boltzmann’s constant, and where

Z= e−βEi (8.41)
i

is the partition function.


Exercise 8.2.1. Show that the density operator ρ can be written as
e−βH
ρ= . (8.42)
Tr (e−βH )
Solution 8.2.1. According to the definition (8.9) one has
 1  −βEi
ρ= wi |i i| = e |i i| . (8.43)
i
Z i

Moreover, the following hold


   
Z= e−βEi = i| e−βH |i = Tr e−βH , (8.44)
i i

and
  
e−βEi |i i| = e−βH |i i| = e−βH |i i| = e−βH , (8.45)
i i i

thus
e−βH
ρ= . (8.46)
Tr (e−βH )
As will be demonstrated below [see Eq. (8.446)], the last result for ρ can
also be obtained from the principle of maximum entropy.

8.3 Problems
1. Consider a spin 1/2 in a magnetic field B = Bẑ and in thermal equilib-
rium at temperature T . Calculate S · û , where S is the vector operator
of the angular momentum and where û is a unit vector, which can be
described using the angles θ and φ
û = (sin θ cos φ, sin θ sin φ, cos θ) . (8.47)

Eyal Buks Quantum Mechanics - Lecture Notes 228


8.3. Problems

2. A spin 1/2 particle is an eigenstate of the operator Sy with eigenvalue


+/2.
a) Write the density operator in the basis of eigenvectors of the operator
Sz .
b) Calculate ρn , where n is integer.
c) Calculate the density operator (in the same basis) of an ensemble of
particles, half of them in an eigenstate of Sy with eigenvalue +/2,
and half of them in an eigenstate of Sy with eigenvalue −/2.
d) Calculate ρn for this case.
3. A spin 1/2 is at time t = 0 in an eigenstate of the operator Sθ = Sx sin θ+
Sz cos θ with an eigenvalue +/2, where θ is real and Sx and Sz are
the x and z components, respectively, of the angular momentum vector
operator. A magnetic field B is applied in the x direction between time
t = 0 and time t = T .
a) The z component of the angular momentum is measured at time
t > T . Calculate the probability P+ to measure the value /2.
b) Calculate the density operator ρ of the spin at times t = T .
4. A spin 1/2 electron is put in a constant magnetic field given by B =
Bẑ, where B is a constant. The system is in thermal equilibrium at
temperature T .
a) Calculate the correlation function

Cz (t) = Sz (t) Sz (0) . (8.48)

b) Calculate the correlation function

Cx (t) = Sx (t) Sx (0) . (8.49)

5. Consider a harmonic oscillator .


with frequency ω. Show that the variance
of the number operator ∆N = N 2  − N 2 (where N = a† a ) is given
by
a) ∆N =  0 for energy eigenstates.
b) ∆N = N  for coherent states.
c) ∆N = N  (N  + 1) for thermal states.
6. Consider a harmonic oscillator having angular resonance frequency ω.
The oscillator is in)thermal
* equilibrium at temperature T . Calculate the
expectation value x2 .
7. Consider a harmonic oscillator in thermal equilibrium at temperature T ,
whose Hamiltonian is given by

p2 mω2 x2
H= + . (8.50)
2m 2
Show that the density operator is given by

Eyal Buks Quantum Mechanics - Lecture Notes 229


Chapter 8. Density Operator
 
ρ= d2 α |α α| P (α) , (8.51)

where |α is a coherent state, d2 α denotes infinitesimal area in the α


complex plane,
! "
2
1 |α|
P (α) = exp − , (8.52)
π N  N 

and where N is the expectation value of the number operator N .


8. Consider a harmonic oscillator in thermal equilibrium at temperature T ,
whose Hamiltonian is given by

p2 mω2 x2
H= + . (8.53)
2m 2
Calculate the probability distribution function f (x) of the random vari-
able x.
9. An LC oscillator (see figure) made of a capacitor C in parallel with an
inductor L, is in thermal equilibrium at temperature T . The charge in
the capacitor q is being measured.

L C

a) Calculate the expectation ( q of q.


' value
b) Calculate the variance (∆q)2 .
10. Consider an observable A having a set of eigenvalues {an }. Let Pn be
a projector operator onto the eigensubspace corresponding to the eigen-
value an . A given physical system is initially described by the density op-
erator ρ0 . A measurement of the observable A is then performed. What
is the density operator ρ1 of the system immediately after the measure-
ment?
11. A model that was proposed by von Neumann describes an indirect mea-
surement process of a given observable A. The observable A is assumed
to be a function of the degrees of freedom of a subsystem, which we refer
to as the measured system (MS). The indirect measurement is performed
by first letting the MS to interact with a measuring device (MD), hav-
ing its own degrees of freedom, and then in the final step, performing a
quantum measurement on the MD. The MS is assumed to initially be in
a pure state |α (i.e. its density operator is assumed to initially be given
by ρ0 = |α α|). Let A be an observable operating on the Hilbert space

Eyal Buks Quantum Mechanics - Lecture Notes 230


8.3. Problems

of the MS. The initial state of the MS can be expanded in the basis of
eigenvectors {|an } of the observable A

|α = cn |an  , (8.54)
n

where cn = an |α and where

A |an  = an |an  . (8.55)

For simplicity, the Hamiltonian of the MS is taken to be zero. The MD is


assumed to be a one dimensional free particle, whose Hamiltonian van-
ishes, and whose initial state is labeled by |ψi . The position wavefunction
ψ (x′ ) = x′ |ψi  of this state is taken to be Gaussian having width x0
!  2 "
′ 1 1 x′
ψ (x ) = 1/2
exp − . (8.56)
π 1/4 x 2 x0
0

The interaction between the MS and the MD is taken to be given by

V (t) = −f (t) xA , (8.57)

where f (t) is assumed to have compact support with a peak near the
time of the measurement.
a) Express the vector state of the entire system |Ψ (t) at time t in the
basis of states {|p′  ⊗ |an′ }. This basis spans the Hilbert space of
the entire system (MS and MD). The state |p′  ⊗ |an′  is both, an
eigenvector of A (with eigenvalue an ) and of the momentum p of the
MD (with eigenvalue p′ ).
b) In what follows the final state of the system after the measurement
will be evaluated by taking the limit t → ∞. The outcome of the
measurement of the observable A, which is labeled by A, is deter-
mined by performing a measurement of the momentum variable p of
the MD. The outcome, which is labeled by P, is related to A by
P
A= , (8.58)
pi
where
 ∞
pi = dt′ f (t′ ) . (8.59)

Calculate the probability distribution g (A) of the random variable


A.
c) Consider another measurement that is performed after the entan-
glement between the MS and the MD has been fully created. The

Eyal Buks Quantum Mechanics - Lecture Notes 231


Chapter 8. Density Operator

additional measurement is associated with the observable B, which


is assumed to be a function of the degrees of freedom of the MS only.
Show that the expectation value B̄ of the observable B is given by

B̄ = an′ | BρR |an′  , (8.60)
n′

where the operator ρR , which is called the reduced density operator,


is given by
a 2
 −η2 n′ −an′′
ρR = cn′ c∗n′′ e 2
|an′  an′′ | . (8.61)
n′ ,n′′

12. A particle having mass m moves in the xy plane under the influence of
a two dimensional potential V (x, y), which is given by

mω 2  2 
V (x, y) = x + y 2 + λmω 2 xy , (8.62)
2
where both ω and λ are real constants. Calculate
) *in thermal equilibrium
at temperature T the expectation values x, x2 .
13. Consider a harmonic oscillator having angular resonance
) frequency ω and
*
mass m. Calculate the correlation function G (t) = x(H) (t) x(H) (0) ,
where x(H) (t) is the Heisenberg representation of the position operator,
for the cases where
a) the oscillator is in its ground state.
b) the oscillator is in thermal equilibrium at temperature T .
14. In general, the Wigner function of a point particle moving in one dimen-
sion is given by
∞  5

6
′ ′ 1 ′′ p′ x′′ ′ x′′

′ x′′
W (x , p ) = dx exp i x − ρ x + , (8.63)
2π  2

2
−∞

where ρ is the density operator of the system, and where |x′  represents
an eigenvector of the position operator x having eigenvalue x′ , i.e. x |x′  =
x′ |x′ . As can be seen from Eq. (4.227), the Wigner function is the inverse
Weyl transformation of the density operator divided by the factor of 2π.
Consider the case of a point particle having mass m in a potential of a
harmonic oscillator having angular frequency ω. Calculate the Wigner
function W (x′ , p′ ) for the case where the system is in a coherent sate
|α.
15. A particle having mass m is in the ground state of the one-dimensional
potential well V1 (x) = (1/2) mω 2 (x − ∆x )2 for times t < 0 . At time
t = 0 the potential suddenly changes and becomes V2 (x) = (1/2) mω 2 x2 .
Calculate the Wigner function of the system at times t > 0.

Eyal Buks Quantum Mechanics - Lecture Notes 232


8.3. Problems

16. Consider a point particle having mass m in a potential of a harmonic


oscillator having angular frequency ω. Calculate the Wigner function
W (x′ , p′ ) for the case where the system is in thermal equilibrium at
temperature T .
17. Consider a point particle having mass m in a potential of a harmonic
oscillator having angular frequency ω. Calculate the Wigner function
W (x′ , p′ ) for the case where the system is in the number state |n = 1.
18. The Wigner function of a point particle moving in one dimension is
given by Eq. (8.63). Show that the marginal distributions x′ | ρ |x′  and
p′ | ρ |p′  of the position x and momentum p observables, respectively, are
given by
 ∞
′ ′ −1
x | ρ |x  =  dp′ W (x′ , p′ ) , (8.64)
−∞
 ∞
p′ | ρ |p′  = −1 dx′ W (x′ , p′ ) . (8.65)
−∞

19. Show that for a pure state the Wigner function is bounded by |W (x′ , p′ )| ≤
1/2π. Note that this bound together with Eqs. (8.64) and (8.65) can be
used to demonstrate the uncertainty principle (3.10).
20. Consider a particle having mass m moving along the x axis under the
influence of the potential V (x). Show that the time evolution of the
Wigner function W (8.63) is governed by
  2l
dW <∞
2i ∂ 2l+1 V ∂ 2l+1 W
= {H, W } + 2l+1
, (8.66)
dt l=1 (2l + 1)! ∂ (x′ ) ∂ (p′ )2l+1

where H is the Hamiltonian and {H, W } is the Poisson’s brackets of H


and W .
21. The function W (X ′ , P ′ ) is defined as the inverse Fourier transform of
the function W̃ (ξ, η)
∞ ∞
′ ′ 1 ′
+iηP ′
W (X , P ) = 2 dξdη W̃ (ξ, η) eiξX , (8.67)
(2π)
−∞ −∞

where the function W̃ (ξ, η) is given by

W̃ (ξ, η) = Tr [exp (−iξX − iηP ) ρ] , (8.68)

X and P are dimensionless position and momentum operators, which are


given by

a + a† a − a†
X= √ , P = √ , (8.69)
2 i 2

Eyal Buks Quantum Mechanics - Lecture Notes 233


Chapter 8. Density Operator

and which satisfy [X, P ] = i [see Eq. (5.13)] and ρ is the density operator.
Show that
∞ 5

6
′ 1 ′ ′′ ′ X ′′

′ X ′′ ′′ ′
W (X , P ) = dX X − ρ X + eiX P , (8.70)
2π 2

2
−∞

i.e. show that W (X ′ , P ′ ) is the Wigner function expressed in terms of


the dimensionless variables X ′ and P ′ [whereas Eq. (8.63) is the Wigner
function expressed in terms of the position x′ and momentum p′ vari-
ables].
22. Equation (8.67) can be rewritten as

W (X ′ , P ′ ) = Tr (Υ ρ) , (8.71)

where the operator Υ is given by


∞ ∞
1 ′
−X )+iη(P ′ −P )

Υ (X , P ) = ′
2 dξdη eiξ(X . (8.72)
(2π)
−∞ −∞

Note that the operator Υ given by Eq. (8.72) is the dimensionless version
of the Weyl kernel (4.46), which defines the Weyl transformation (4.45).
Show that
   
X ′ + iP ′ X ′ + iP ′
Υ (X ′ , P ′ ) = π −1 D† − √ PD − √ , (8.73)
2 2
where
 
D (α) = exp αa† − α∗ a (8.74)

is the displacement operator [see Eq. (5.36)], a is the annihilation oper-


ator and
∞
P= dX ′ |X ′  −X ′ | (8.75)
−∞

is the parity operator [see Eq. (5.101)], where |X ′  is an eigenvector of the


dimensionless position operator X having eigenvalue X ′ , i.e. X |X ′  =
X ′ |X ′ .
23. Homodyne Tomography - Consider a point particle having mass m
in a potential of a harmonic oscillator having angular frequency ω. The
normalized homodyne observable Xφ with a real phase φ is defined by

a† eiφ + ae−iφ
Xφ = √ , (8.76)
2

Eyal Buks Quantum Mechanics - Lecture Notes 234


8.3. Problems

where a and a†are annihilation and creation operators [see Eqs. (5.9) and

(5.10)]. Let w Xφ be the normalized probability distribution function
of the observable
 X φ . The technique of homodyne detection can be used
to measure w Xφ′ for any given value of the phase φ.
a) To generalize Eqs. (8.64) and (8.65) show that the following holds
for any real φ

∞
   
w Xφ′ = dPφ′ W Xφ′ cos φ − Pφ′ sin φ, Xφ′ sin φ + Pφ′ cos φ .
−∞
(8.77)

b) Show that the Wigner function


  (8.63) can be extracted from the
measured distributions w Xφ′ for all values of φ.
24. Consider a harmonic oscillator in thermal equilibrium at temperature T ,
whose Hamiltonian is given by

p2 mω2 x2
H= + . (8.78)
2m 2
Calculate the matrix elements x′′ | ρ |x′  of the density operator in the
basis of eigenvectors of the position operator x.
25. Consider a harmonic oscillator having angular resonance frequency ω.
The oscillator is in)thermal* equilibrium at temperature T . Calculate the
expectation value e−iζXφ , where Xφ is given by [see Eq. (8.76)]

a† eiφ + ae−iφ
Xφ = √ , (8.79)
2

a and a† are annihilation and creation operators [see Eqs. (5.9) and (5.10)]
and both
) −iζX * φ and ζ are real. Use your result for the expectation value
e φ
to evaluate
 the Wigner function of the system.
26. Show that when w Xφ′ is φ independent the following holds

∞   
′ 1

W (X , P ) = dζ ζ w̃ (ζ) J0 ζ X ′2 + P ′2 , (8.80)

0
 
where w̃ (ζ) is the (φ independent) Fourier transform of w Xφ′ , i.e.

∞
  ′
w̃ (ζ) = dXφ′ w Xφ′ e−iζXφ . (8.81)
−∞

Eyal Buks Quantum Mechanics - Lecture Notes 235


Chapter 8. Density Operator

27. Consider a point particle having mass m in a potential of a harmonic


oscillator having angular frequency ω. Express the Wigner function
W (X ′ , P ′ ; t) at time t in terms of the Wigner function W (X ′ , P ′ ; 0)
at time t = 0.
28. Let W (X ′ , P ′ ) be the Wigner function of a system whose density oper-
ator is ρ. Express the Wigner function Wα (X ′ , P ′ ) of a system whose
′ † ′
density operator  ′ is displaced
 according to ρα = D (α ) ρD (α ), where
′ † ′∗
D (α ) = exp α a − α a is the displacement operator [see Eq. (5.36)],
and where α′ is complex, in terms of the Wigner function of the undis-
placed system W (X ′ , P ′ ).
29. Consider a weak measurement of the dimensionless position X [see Eq.
(8.69)] of a point particle moving in one dimension. In view of Eq. (8.61),
the reduced density operator of the system after the measurement, which
is labeled as ρR , is expected to be related to the density operator before
the measurement ρ by the following relation
5

6 5

6  ′′  2
′ X ′′

′ X ′′ ′ X ′′

′ X ′′ − ηX
X −
ρR
X + = X −
ρ
X + e 2
,
2 2 2 2
(8.82)
where the dimensionless parameter η characterizes the strength of the
measurement, and where |X ′  represents an eigenvector of the dimen-
sionless position operator X [see Eq. (8.69)] having eigenvalue X ′ , i.e.
X |X ′  = X ′ |X ′ . Express the reduced Wigner function WR (X ′ , P ′ ),
which is given by [see Eq. (8.70)]
∞ 5

6
′ ′ 1 X ′′

′ X ′′ ′′ ′
WR (X , P ) = dX ′′ X′ − ρ X + eiX P , (8.83)
2π 2
R
2
−∞

in terms of the Wigner function W (X ′ , P ′ ) before the measurement.


30. A point particle having mass m is confined by the three dimensional
potential
1
V (r) = mω2 r2 , (8.84)
2

2 2 2
) r2 *= x + y + z and where ω is a real constant. Calculate x
where
and x in thermal equilibrium at temperature T .
31. The entropy σ is defined by

σ = − Tr (ρ log ρ) . (8.85)

Show that σ is time independent.


32. Consider the 2 × 2 matrix ρ, which is given by
1
ρ= (1 + k · σ) , (8.86)
2

Eyal Buks Quantum Mechanics - Lecture Notes 236


8.3. Problems

where k = (kx , ky , kz ) is a three dimensional vector of complex numbers


and where σ = (σ x , σ y , σz ) is the Pauli matrix vector.
a) Under what conditions on k the matrix ρ can represent a valid density
operator of a spin 1/2 particle?
b) Under what conditions on k the matrix ρ can represent a valid density
operator of a spin 1/2 particle in a pure state?
c) Calculate the term Tr (û · σρ), where û is a unit vector, i.e. û · û = 1.
33. The matrix representation in the basis of eigenvectors of Sz of the density
operator of a spin 1/2 particle is given by
1
ρ= (1 + k · σ) , (8.87)
2
where k = (kx , ky , kz ) is a three dimensional vector of real numbers, and
σ = (σx , σy , σ z ) is the Pauli matrix vector. The entropy σ is defined by

σ = − Tr (ρ log ρ) . (8.88)

a) Calculate σ.
b) A measurement of Sz is performed. Calculate the entropy after the
measurement.
34. The maximum entropy principle - The entropy σ is defined by

σ (ρ) = − Tr (ρ log ρ) . (8.89)

Consider the case where the density matrix is assumed to satisfy a set of
contrarians, which are expressed as

gl (ρ) = 0 , (8.90)

where l = 0, 1, · · · L. The functionals gl (ρ) maps the density operator ρ


to a complex number, i,e. gl (ρ) ∈ C.
a) Find an expression for a density matrix that satisfy all these con-
strains, for which the entropy σ obtains a stationary point (maxi-
mum, minimum or a saddle point). Assume that the constrain l = 0
is the requirement that Tr (ρ) = 1, i.e. g0 (ρ) can be taken to be given
by

g0 (ρ) = Tr (ρ) − 1 = 0 . (8.91)

Moreover, assume that the other constrains l = 1, · · · L are the re-


quirements that the expectation values of the Hermitian operators
X1 , X2 · · · , XL are the following real numbers X1 , X2 , · · · ,XL respec-
tively, i.e. gl (ρ) for l ≥ 1 can be taken to be given by

gl (ρ) = Tr (ρXl ) − Xl = 0 . (8.92)

Eyal Buks Quantum Mechanics - Lecture Notes 237


Chapter 8. Density Operator

b) Express ρ for the case of a microcanonical ensemble, for which the


only required constrain is (8.91).
c) Express ρ for the case of a canonical ensemble, for which in addition
to the constrain is (8.91) the expectation value of the Hamiltonian
H is required to have a given value, which is labeled by H.
d) Express ρ for the case of a grandcanonical ensemble, for which in ad-
dition to the constrain is (8.91) the expectation values of the Hamil-
tonian H and of the operator N are required to have given values,
which are labeled by H and N  respectively. The operator N ,
which will be defined in chapter 16, is called the number of particles
operator.
35. Consider a point particle having mass m moving in one dimension under
the influence of the potential V (x). Calculate the canonical partition
function Zc [see Eq. (8.447)] in the classical limit, i.e. in the limit of high
temperature.
36. Let ρ be the density operator of a given system. The system is composed
of two subsystems, each having its own degrees of freedom, which are
labeled as ’1’ and ’2’ (e.g. a system of two particles). Let {|n1 1 } ({|n2 2 })
be an orthonormal basis spanning the Hilbert space of subsystem ’1’
(’2’). The set of vectors {|n1 , n2 }, where |n1 , n2  = |n1 1 |n2 2 , forms an
orthonormal basis spanning the Hilbert space of the combined system.
For a general operator O the partial trace over subsystem ’1’ is defined
by the following relation

Tr1 (O) ≡ 1 n1 | O |n1 1 . (8.93)
n1

Similarly, the partial trace over subsystem ’2’ is defined by



Tr2 (O) ≡ 2 n2 | O |n2 2 . (8.94)
n2

The observable A1 is a given Hermitian operator on the Hilbert space of


subsystem ’1’. Show that the expectation value of a measurement of A1
that is performed on subsystem ’1’ is given by

A1  = Tr1 (ρ1 A1 ) . (8.95)

where the operator ρ1 , which is given by

ρ1 = Tr2 ρ , (8.96)

is called the reduced density operator of subsystem ’1’.


37. Let ρ be the density operator of a given system. The total entropy of the
system σ is given by

σ = − Tr (ρ log ρ) . (8.97)

Eyal Buks Quantum Mechanics - Lecture Notes 238


8.4. Solutions

As in the previous exercise, the system is composed of two subsystems,


which are labeled as ’1’ and ’2’. Let {|n1 1 } ({|n2 2 }) be an orthonormal
basis spanning the Hilbert space of subsystem ’1’ (’2’). The set of vec-
tors {|n1 , n2 }, where |n1 , n2  = |n1 1 |n2 2 , forms an orthonormal basis
spanning the Hilbert space of the combined system. The reduced density
operators ρ1 and ρ2 of subsystems ’1’ and ’2’ respectively are giving by

ρ1 = 2 n2 | ρ |n2 2 = Tr2 ρ , (8.98)
n2

ρ2 = 1 n1 | ρ |n1 1 = Tr1 ρ , (8.99)
n1

and the subsystems’ entropies σ1 and σ2 are given by


σ1 = − Tr1 (ρ1 log ρ1 ) , (8.100)
σ2 = − Tr2 (ρ2 log ρ2 ) . (8.101)
Show that

σ1 + σ2 ≥ σ . (8.102)

38. Consider a spin 1/2 in a magnetic field B = Bẑ and in thermal equilib-
rium at temperature T . Calculate the entropy σ, which is defined by

σ = − Tr (ρ log ρ) , (8.103)

where ρ is the density operator of the system.


39. A spin 1/2 is in a state |H, which satisfies the following relation
 
1 1 2 (Sx + Sz )
|H H| = 1+ √ , (8.104)
2 2 

where 1 is the identity operator, and where Sx and Sz are spin angular
momentum operators. In a measurement of Sz what is the probability
pz+ to obtain the value +/2?
40. Consider a harmonic oscillator of angular frequency ω and mass m in
thermal equilibrium at temperature T . Calculate the entropy σ, which is
defined by

σ = − Tr (ρ log ρ) , (8.105)

where ρ is the density operator of the system.

8.4 Solutions

1. The Hamiltonian is given by

Eyal Buks Quantum Mechanics - Lecture Notes 239


Chapter 8. Density Operator

H = ωSz , (8.106)
where
|e| B
ω= (8.107)
me c
is the Larmor frequency. In the basis of the eigenvectors of Sz

Sz |± = ± |± , (8.108)
2
one has

H |± = ± |± , (8.109)
2
thus
e−Hβ
ρ=
Tr (e−Hβ )
 ωβ  ωβ
e− 2 |+ +| + e 2 |− −|
=  ωβ  ωβ ,
e− 2 +e 2

(8.110)
where β = 1/kB T , and therefore with the help of Eqs. (2.102) and (2.103),
which are given by

Sx = (|+ −| + |− +|) , (8.111)
2

Sy = (−i |+ −| + i |− +|) , (8.112)
2
one has
Sx  = Tr (ρSx ) = 0 , (8.113)
Sy  = Tr (ρSy ) = 0 , (8.114)
and with the help of Eq. (2.99), which is given by

Sz = (|+ +| − |− −|) , (8.115)
2
one has
Sz  = Tr (ρSz )
!  ωβ  ωβ
"
e− 2 |+ +| + e 2 |− −| 
= Tr  ωβ  ωβ (|+ +| − |− −|)
e− 2 + e 2 2
 ωβ  ωβ
 e− 2 − e 2
=
2 e−  ωβ
2
 ωβ
+e 2
 
 ωβ
= − tanh ,
2 2
(8.116)

Eyal Buks Quantum Mechanics - Lecture Notes 240


8.4. Solutions

thus
 
 cos θ ωβ
S · û = − tanh . (8.117)
2 2
2. Recall that
1
|±; ŷ = √ (|+ ± i |−) , (8.118)
2
a) thus
   
1 1   1 1 −i
ρ=
˙ 1 −i = . (8.119)
2 i 2 i 1
b) For a pure state ρn = ρ.
c) For this case
 
 
1  1 10
ρ= |+; ŷ +; ŷ| + |−; ŷ −; ŷ| =˙ , (8.120)
2    2 01
=1

d) and
 
1 10
ρn =
˙ . (8.121)
2n 01

3. The state at time t = 0 is given by


 
cos θ2
|ψ (t = 0) =
˙ , (8.122)
sin θ2

and the one at time t = T is


 
iωT σx
|ψ (t = T ) = exp |ψ (t = 0) , (8.123)
2
where σx is a Pauli matrix, and
eB
ω= . (8.124)
me c
Using the identity
 
iσ · n̂φ φ φ
exp − = cos − iσ · n̂ sin , (8.125)
2 2 2
one finds
   
iσ · n̂φ ωT ωT cos ωT ωT
2 i sin 2
exp − = cos +iσx sin =
˙ , (8.126)
2 2 2 i sin ωT ωT
2 cos 2

Eyal Buks Quantum Mechanics - Lecture Notes 241


Chapter 8. Density Operator

thus
  
cos ωT ωT
2 i sin 2 cos θ2
|ψ (t = T ) =
˙ ωT ωT
i sin 2 cos 2 sin θ2
 
cos ωT θ ωT θ
2 cos 2 + i sin 2 sin 2
= .
i sin 2 cos 2 + cos 2 sin θ2
ωT θ ωT

(8.127)
a) The probabilities to measured ±/2 are thus given by
ωT θ ωT θ
P+ = cos2 cos2 + sin2 sin2
2 2 2 2
1 + cos (ωT ) cos θ
= , (8.128)
2
and
ωT θ ωT θ
P− = cos2 sin 2 + sin2 cos 2
2 2 2 2
1 − cos (ωT ) cos θ
= . (8.129)
2
b) The density operator is given by
ρ11 = P+ ,
ρ22 = P− ,
  
ωT θ ωT θ ωT θ ωT θ
ρ21 = cos cos + i sin sin −i sin cos + cos sin
2 2 2 2 2 2 2 2
sin θ i
= − sin ωT cos θ ,
2 2
ρ12 = ρ∗21 .
4. The Hamiltonian is given by

H = −ωSz , (8.130)

where
eB
ω= , (8.131)
me c
thus, the density operator is given by
   

1  exp 2kB T 
0
 ,
ρ=˙ (8.132)
Z 0 exp − 2kω
BT

where
   
ω ω
Z = exp + exp − . (8.133)
2kB T 2kB T

Eyal Buks Quantum Mechanics - Lecture Notes 242


8.4. Solutions

a) Using
   
iHt iHt
Sz (t) = exp Sz (0) exp − = Sz (0) , (8.134)
 
one finds
) *   2
Cz (t) = Sz2 (0) = Tr ρSz2 (0) = . (8.135)
4
b) The following holds
   
iωSz t iωSz t
Sx (t) = exp − Sx (0) exp
 
= Sx cos (ωt) + Sy sin (ωt) ,
(8.136)
thus ) *
Cx (t) = cos (ωt) Sx2 (0) + sin (ωt) Sy (0) Sx (0) (8.137)
2
cos (ωt) 
= + sin (ωt) Sy (0) Sx (0) .
4
In terms of Pauli matrices    
  
2 exp 2kω T 0 0 −i 0 1
Sy (0) Sx (0) = Tr  B   
4Z 0 exp − 2kω i 0 10
BT
   

2 −i exp 2kB T 0
= Tr    
4Z 0 i exp − 2kωBT
 
i2 ω
=− tanh ,
4 2kB T
(8.138)
thus
 
2 ω
Cx (t) = cos (ωt) − i sin (ωt) tanh . (8.139)
4 2kB T
5. The variance ∆N is given by
a) For an energy eigenstate |n one has

N |n = n |n , (8.140)

thus

N  = n| N |n = n , (8.141)

and
) 2*
N = n| N 2 |n = n2 , (8.142)

Eyal Buks Quantum Mechanics - Lecture Notes 243


Chapter 8. Density Operator

therefore

∆N = 0 . (8.143)

b) For a coherent state |α one has

a |α = α |α , (8.144)

thus
2
N  = α| N |α = α| a† a |α = |α| , (8.145)

and
 
) 2* , - 
N = α| a† aa† a |α = α| a†  a, a† + a† a a |α = |α|2 + |α|4 ,
  
=1
(8.146)

therefore

∆N = N  . (8.147)

c) In general, for a thermal state one has

O = Tr (ρO) , (8.148)

where O is an operator,
1
ρ = e−Hβ , (8.149)
Z 
Z = Tr e−Hβ , (8.150)
and β = 1/kB T and H is the Hamiltonian. For the present case
 
1
H = ω N + , (8.151)
2

thus

Eyal Buks Quantum Mechanics - Lecture Notes 244


8.4. Solutions

N  = Tr (ρN )

<
n| e−Hβ N |n
n=0
= < ∞
n| e−Hβ |n
n=0

<
ne−nωβ
n=0
= ∞
<
e−nωβ
n=0
∞ 
< −nωβ
∂ log e
1 n=0
=−
ω ∂β
e−βω
= ,
1 − e−βω
(8.152)
and ) *  
N 2 = Tr ρN 2

<
n| e−Hβ N 2 |n
n=0
= < ∞
n| e−Hβ |n
n=0

<
n2 e−nωβ
n=0
= <∞
e−nωβ
n=0
  <∞
1 2 ∂2
ω ∂β 2
e−nωβ
n=0
= ∞
<
e−nωβ
n=0
 −βω 
e + 1 e−βω
= 2 ,
(1 − e−βω )
(8.153)
and therefore
) * e−βω
(∆N)2 = N 2 − N2 = 2 = N  (N + 1) . (8.154)
(1 − e−βω )
6. The density operator is given by
1 −Hβ
ρ= e . (8.155)
Z
where

Eyal Buks Quantum Mechanics - Lecture Notes 245


Chapter 8. Density Operator
∞ β ω
   e− 2 1
e−βω(n+ 2 ) =
1
Z = tr e−Hβ = =   , (8.156)
n=0
1 − e−βω 2 sinh ωβ
2

and β = 1/kB T . Thus using


  2  † 2 
x2 = a + a + 2a† a + 1 , (8.157)
2mω
one finds
) 2*  
x = Tr x2 ρ

1 
= n| x2 e−Hβ |n
Z n=0

1  −ω(n+ 12 )β
= e n| x2 |n
Z n=0
∞  
 1  1 −ω(n+ 12 )β
= n+ e
mω Z n=0 2
  ∞
 1 1 d  −ω(n+ 12 )β
= − e .
mω Z ω dβ n=0
(8.158)
However


e−ω(n+ 2 )β = Z ,
1
(8.159)
n=0

thus
) 2* 1 d
x = log Z −1
mω 2 dβ
 
d ωβ
1 dβ sinh 2
=  
mω 2 sinh ωβ
2
 
1 ω ωβ
= coth .
mω 2 2 2
(8.160)
Note that at high temperatures ωβ ≪ 1
) 2* kB T
x ≃ , (8.161)
mω 2
as is required by the equipartition theorem of classical statistical mechan-
ics.

Eyal Buks Quantum Mechanics - Lecture Notes 246


8.4. Solutions

7. In the basis of number states the density operator is given by


<∞
−Hβ e−Hβ |n n|
e n=0
ρ= = <∞
Tr (e−Hβ )
n| e−Hβ |n
n=0

<
e−βω(N+ 2 ) |n n|
1

= n=0
<∞
n| e−βω(N+ 2 ) |n
1

n=0
<∞
e−nβω |n n|
n=0
= ∞
<
e−nβω
n=0

 
= 1 − e−βω e−nβω |n n| ,
n=0
(8.162)
where β = 1/kB T . Thus, N  is given by
N  = Tr (ρN )

 
= 1 − e−βω ne−nβω
n=0

  ∂ 
= −ω 1 − e−βω e−nβω
∂β n=0
  ∂ 1
= −ω 1 − e−βω
∂β 1 − e−βω
e−βω
= .
1 − e−βω
(8.163)
Moreover, the following holds
1
N + 1 = , (8.164)
1 − e−βω
N 
= e−βω , (8.165)
N  + 1
thus, ρ can be rewritten as

Eyal Buks Quantum Mechanics - Lecture Notes 247


Chapter 8. Density Operator

 −βω

ρ = 1−e e−nβω |n n|
n=0
∞  n
1 N
= |n n| .
N  + 1 n=0 N  + 1
(8.166)
To verify the validity of Eq. (8.51) we calculate
 
n| ρ |m = d2 αP (α) n |α α |m . (8.167)

With the help of Eq. (5.42), which is given by

∞
|α|2 αn
|α = e− 2 √ |n , (8.168)
n=0 n!

one finds that


  ! "
1 2 |α|2 2 α
n
α∗m
n| ρ |m = d α exp − e−|α| √ √ . (8.169)
π N N  n! m!

Employing polar coordinates in the complex plane α = reiθ , where r is


non-negative real and θ is real,
∞ 2π
1 −(1+ N
1
)r
2
n+m+1
n| ρ |m = √ dre r dθeiθ(n−m)
π N  n!m!
0 0
  
2πδnm
∞
2δ nm
dre−(1+ N )r r2n+1 ,
1 2
=
N  n!
0
(8.170)
and the transformation of the integration variable
 
1
t = 1+ r2 , (8.171)
N 
 
1
dt = 1 + 2rdr , (8.172)
N 
lead to

Eyal Buks Quantum Mechanics - Lecture Notes 248


8.4. Solutions

∞
δ nm
n| ρ |m =  n+1 dte−t tn
1
N  1 + N n! 0
  
Γ (n+1)=n!
δ nm
=  n+1
1
N  1 + N
N n δ nm
= n+1 ,
(1 + N )
(8.173)
in agreement with Eq. (8.166).
8. The density operator [see Eq. (8.51)] is given by
 
ρ= d2 α |α α| P (α) , (8.174)

where |α is a coherent state, d2 α denotes infinitesimal area in the α


complex plane,
! "
1 |α|2
P (α) = exp − , (8.175)
π N  N 

and where
e−βω
N  = (8.176)
1 − e−βω
is the expectation value of the number operator N. Thus,
 
f (x′ ) = x′ | ρ |x′  = d2 αP (α) x′ |α α |x′  .

By employing the expression for the wave function ψα (x′ ) = x′ |α of a
coherent state which is given by [see Eq. (5.51)]
ψα (x′ ) = x′ |α
 ∗2     
α − α2  mω 1/4 x′ − xα 2 x′
= exp exp − + i pα ,
4 π 2∆xα 
(8.177)
where
0
2
xα = α| x |α = Re (α) , (8.178)

. 0
2 
∆xα = α| (∆x) |α = , (8.179)
2mω

Eyal Buks Quantum Mechanics - Lecture Notes 249


Chapter 8. Density Operator

one finds that


1  mω 1/2
f (x′ ) = x′ | ρ |x′  =
π N π
  ! "   2 
2 |α|2 x′ − xα
× d α exp − exp −2
N 2∆xα
1  mω 1/2
=
π N  π
  2 
  ! "
2 ′
|α|  x 
× d2 α exp − exp −2  . − Re (α) 
N 2

 2

 mω 1/2 
 √x2
′ 


−2
=  π e 1+2N

1 + 2 N 
 2

: 
 √x2
′ 

1 1 mω
= √ 
e−2 1+2N
π mω (1 + 2 N)
 ′ 2
1 − xξ
= √ e ,
ξ π

where
0

ξ= (1 + 2 N) ,

and where
 
e−βω βω
1 + 2 N  = 1 + 2 = coth . (8.180)
1 − e−βω 2
9. Recall that the LC circuit is a harmonic oscillator.
a) In terms of the annihilation and creation operators
0  
Lω ip
a= q+ , (8.181)
2 Lω
0  
† Lω ip
a = q− , (8.182)
2 Lω
one has 0
  
q= a + a† , (8.183)
2Lω
 
1
H = ω a† a + . (8.184)
2

Eyal Buks Quantum Mechanics - Lecture Notes 250


8.4. Solutions

The density operator is given by


1 −βH
ρ= e , (8.185)
Z
where
1
β= , (8.186)
kB T
and

 β ω
e− 2 1
e−βω(n+ 2 ) =
1
Z = Tr e−βH = = , (8.187)
n=0
1−e−βω
2 sinh βω
2

thus
0 ∞
1    
q = Tr (qρ) = n| a + a† e−βH |n = 0 . (8.188)
Z 2Lω n=0

b) Similarly
) 2*  
q = Tr q 2 ρ

 1   2
= n| a + a† e−βH |n
2Lω Z n=0
∞  
1 1  1 −βω(n+ 12 )
= ω n + e
Lω2 Z n=0 2
1 1 dZ
=−
Lω 2 Z dβ
Cω ω
= coth .
2 2kB T
(8.189)
10. In general, ρ0 can be expressed as


('

ρ0 = wi
α(i) α(i)
, (8.190)
i
< ) (i)
(i) *
where 0 ≤ wi ≤ 1, i wi = 1, and where
α

(i) *α = 1. Assume first

that the system is initially in the state α . The probability for this to
be the case is wi . In general, the possible results of a measurement of the
observable A are the eigenvalues {an }. The probability p
n to *measure the
eigenvalue an given that the system is initially in state
α(i) is given by
'

(

pn = α(i)
Pn
α(i) . (8.191)

After a measurement of A with an outcome an the state vector collapses


onto the corresponding eigensubspace and becomes

Eyal Buks Quantum Mechanics - Lecture Notes 251


Chapter 8. Density Operator


(
*

(i) Pn
α(i)

α → )
.

*. (8.192)
α(i)
Pn
α(i)

*
Thus, given that the system is initially in state
α(i) the final density
operator is given by



* ) (i)

' ( Pn
α(i) α
Pn
(i) (i)

(i)
ρ1 = α
Pn
α .)

* .)

*
n α(i)
Pn
α(i) α(i)
Pn
α(i)


('

= Pn
α(i) α(i)
Pn .
n
(8.193)
Averaging over all possible initial states thus yields
 (i)
 

('



ρ1 = wi ρ1 = Pn wi
α(i) α(i)
Pn = Pn ρ0 Pn . (8.194)
i n i n

11. Since [V (t) , V (t′ )] = 0 the time evolution operator from initial time t0
to time t is given by
  
i t ′ ′
u (t, t0 ) = exp − dt V (t )
 t0
 
ipi A
= exp x ,

(8.195)

where
 t
pi = dt′ f (t′ ) . (8.196)
t0

While the initial state of the entire system at time t0 is given by |Ψ (t0 ) =
|ψi  ⊗ |α, the final state at time t is given by

|Ψ (t) = u (t, t0 ) |Ψ (t0 )



= cn Jn |ψi  ⊗ |an  ,
n
(8.197)

where the operator Jn is given by


 
ipi an
Jn = exp x . (8.198)


Eyal Buks Quantum Mechanics - Lecture Notes 252


8.4. Solutions
/
a) By introducing the identity operator dp′ |p′  p′ | = 1MD on the
Hilbert space of the MD, where |p′  are eigenvectors of the momen-
tum operator p, which is canonically conjugate to x, the state |Ψ (t)
can be expressed as
 
|Ψ (t) = cn dp′ p′ | Jn |ψ i  |p′  ⊗ |an  . (8.199)
n

With the help of the general identity (3.76), which is given by


dA
[p, A (x)] = −i , (8.200)
dx
where A (x) is a function of the operator x, one finds that

pJn |p′  = ([p, Jn ] + Jn p) |p′ 


= (pi an + p′ ) Jn |p′  ,
(8.201)

thus the vector Jn |p′  is an eigenvector of p with eigenvalue pi an +p′ .


Moreover, note that this vector, which is labeled as |p′ + pi an  ≡
Jn |p′ , is normalized, provided that |p′  is normalized, since Jn is
unitary. The momentum wavefunction φ (p′ ) = p′ |ψi  of the state
|ψi  is related to the position wavefunction x′ |ψi  by a Fourier trans-
form [see Eq. (3.60)]
/∞ ip′ x′
dx′ e−  x′ |ψ i 
−∞
φ (p′ ) = √
2π
!  2 "
1 1 p′
= 1/2
exp − ,
π1/4 p 2 p0
0
(8.202)

where

p0 = . (8.203)
x0
In terms of φ (p′ ) the state |Ψ (t) thus can be expressed as
 
|Ψ (t) = cn dp′ p′ − pi an |ψi  |p′  ⊗ |an 
n
 
= cn dp′ φ (p′ − pi an ) |p′  ⊗ |an  .
n
(8.204)

Eyal Buks Quantum Mechanics - Lecture Notes 253


Chapter 8. Density Operator

b) The probability distribution g (A) of the random variable A can be


calculated using Eq. (8.204)
 2
g (A) = pi |(an′ | ⊗ pi A|) |Ψ (t)|
n′
 2 2
= pi |cn′ | |φ (pi (A − an′ ))|
n′
η  2
(A−an′ )2
= |cn′ |2 e−η ,
π1/2 n′
(8.205)

where
 ∞
pi x0
η= = dt′ f (t′ ) . (8.206)
p0  t0

The expectation value of A is given by


∞
A = dA′ g (A′ ) A′
−∞

 ∞
η ′′ 2
= |cn′ |2
dA′′ e−(ηA ) (A′′ + an′ )
n′
π1/2
−∞
 2
= |cn′ | an′ .
n′
(8.207)

c) The density operator of the entire system is taken to be given by


ρf = |Ψ (∞) Ψ (∞)| for this case. The additional measurement is
associated with the observable B, which is assumed to be a function
of the degrees of freedom of the MS only. This assumption allows
expressing the expectation value B̄ of the observable B as

B̄ = Tr (Bρf )

= dp′ an′ | ⊗ p′ | Bρf |p′  ⊗ |an′ 
n′

= an′ | BρR |an′  ,
n′
(8.208)

where ρR , which is given by



ρR = dp′ p′ | ρf |p′  , (8.209)

Eyal Buks Quantum Mechanics - Lecture Notes 254


8.4. Solutions

is called the reduced density operator. Note that ρR is an operator


on the Hilbert space of the MS. With the help of the expressions
 
|Ψ (∞) = cn′ dp′′ φ (p′′ − pi an′ ) |p′′  ⊗ |an′  , (8.210)
n
 
Ψ (∞)| = c∗n′′ dp′′′ φ∗ (p′′′ − pi an′′ ) an′′ | ⊗ p′′′ | , (8.211)
n′

one finds that


 
ρR = cn′ cn′′ dp′

n′ ,n′′

× φ (p′ − pi an′ ) φ∗ (p′ − pi an′′ ) |an′  an′′ | .


(8.212)

By employing the transformation of integration variable


2p′ − pi (an′ + an′′ )
x= , (8.213)
2p0
and its inverse
 
pi an′ + an′′
p′ = p0 x + , (8.214)
p0 2

one finds that


  a −a


′′ 2
−η2 n 2 n
dp′ φ (p′ − pi an′ ) φ∗ (p′ − pi an′′ ) = e , (8.215)

thus
a 2
 −η2 n′ −an′′
ρR = cn′ c∗n′′ e 2
|an′  an′′ | . (8.216)
n′ ,n′′

12. It is convenient to employ the coordinate transformation


x+y
x′ = √ , (8.217)
2
x − y
y′ = √ . (8.218)
2
The Lagrangian of the system can be written using these coordinates [see
Eq. (9.189)] as

L = L+ + L− , (8.219)

where

Eyal Buks Quantum Mechanics - Lecture Notes 255


Chapter 8. Density Operator

mẋ′2 mω2
L+ = − (1 + λ) x′2 , (8.220)
2 2
and
mẏ ′2 mω2
L− = − (1 − λ) y ′2 . (8.221)
2 2
Thus, the system is composed of√two decoupled harmonic
√ oscillators with
angular resonance frequencies ω 1 + λ (for x′ ) and ω 1 − λ (for y ′ ). In
thermal equilibrium according to Eq. (8.160) one has
 √ 
) ′2 *  ω 1 + λβ
x = √ coth , (8.222)
2mω 1 + λ 2
 √ 
) ′2 *  ω 1 − λβ
y = √ coth , (8.223)
2mω 1 − λ 2
where β = 1/kB T . Moreover, due to symmetry, the following holds
x′  = y ′  = 0 , (8.224)
x′ y′  = 0 . (8.225)
With the help of the inverse transformation, which is given by
x′ + y ′
x= √ , (8.226)
2
x′ − y ′
y= √ , (8.227)
2
one thus finds

x = 0 , (8.228)

and
) 2 * 1 ) ′2 *
x = x + y′2
2  √   √ 

ω 1+λβ ω 1−λβ
  coth 2 coth 2
= √ + √  .
4mω 1+λ 1−λ
(8.229)
13. Using Eq. (5.152), which is given by

p(H) (0)
x(H) (t) = x(H) (0) cos (ωt) + sin (ωt) , (8.230)

one finds that
' ( sin (ωt) ) *
G (t) = cos (ωt) x2(H) (0) + p(H) (0) x(H) (0) . (8.231)

Eyal Buks Quantum Mechanics - Lecture Notes 256


8.4. Solutions

Using the relations


0
  
x= a + a† , (8.232)
2mω
0
mω  
p=i −a + a† , (8.233)
2
, †-
a, a = 1 , (8.234)
one finds that
  2  † 2 
x2 = a + a + 2a† a + 1 , (8.235)
2mω
px   2  † 2 
=i −a + a −1 . (8.236)
mω 2mω
a) Thus, for the ground state [see Eqs. (5.28) and (5.29)]

 
G (t) = [cos (ωt) − i sin (ωt)] = exp (−iωt) . (8.237)
2mω 2mω
b) The density operator ρ is given by Eq. (8.166)

∞  n
1 N
ρ= |n n| , (8.238)
N  + 1 n=0 N  + 1

where
e−βω
N  = Tr (ρN ) = , (8.239)
1 − e−βω
N = a† a, and where β = 1/kB T . Using the fact that 'ρ is diagonal
) *  2 (
in the basis of number states one finds that a2 = a† = 0.
Combining all these results leads to

G (t) = [(2 N  + 1) cos (ωt) − i sin (ωt)]
2mω 
 βω
= coth cos (ωt) − i sin (ωt) .
2mω 2
(8.240)
14. The wave function of the coherent state |α is given by Eq. (5.51)
ψα (x′ ) = x′ |α
 ∗2     
α − α2  mω 1/4 x′ − xα 2 x′
= exp exp − + i pα .
4 π 2∆xα 
(8.241)
where

Eyal Buks Quantum Mechanics - Lecture Notes 257


Chapter 8. Density Operator
0
2 ′
xα = α| x |α = α , (8.242)


pα = α| p |α = 2mωα′′ , (8.243)
α′ = Re (α) , (8.244)
α′′ = Im (α) , (8.245)
. 0
2 
∆xα = α| (∆x) |α = , (8.246)
2mω
Using the definition (8.63) and the identity
∞ 0
  π 14 4ca+b 2
exp −ax2 + bx + c dx = e a , (8.247)
a
−∞

one has
 ∞  ′ ′′  5
6
1 px x′′
x′′
W (x′ , p′ ) = exp i x′ − |α α

x′ + dx′′
2π −∞  2 2
 mω 1/2  ∞
= π
dx′′
2π −∞
! ′′
"2 ! ′′
"2  
x′ − x −x α x′ + x −x α p′ −p α
− 2
2∆xα − 2
2∆xα +i  x′′
×e
 mω 1/2  ∞
= π
dx′′
2π −∞
′′ 2  
(x′ −x α )2 +( x2 ) p′ −p α
− 2(∆xα )2
+i  x′′
×e ,
(8.248)
thus
 2  2
x′ −x α p′ −p α
′ 1 − 12
′ ∆xα − 12 ∆pα
W (x , p ) = e , (8.249)
π
where [see Eq. (5.49)]
. 0
2 mω 
∆pα = α| (∆p) |α = = . (8.250)
2 2∆xα

*
15. At time t > 0 the system is in a coherent state
α = α0 e−iωt , where [see
Eqs. (5.53) and (5.240)]
0

α0 = ∆x . (8.251)
2
Thus, with the help of Eq. (8.249) one finds that the Wigner function of
the system at time t is given by

Eyal Buks Quantum Mechanics - Lecture Notes 258


8.4. Solutions
 2  2
x′ −x α p′ −p α
′ 1 − 12
′ ∆xα − 12 ∆pα
W (x , p ) = e , (8.252)
π
where
xα = ∆x cos (ωt) , (8.253)
pα = −mω∆x sin (ωt) , (8.254)
0

∆xα = , (8.255)
2mω
0
mω
∆pα = . (8.256)
2
16. The density operator [see Eq. (8.51)] is given by
 
ρ= d2 α |α α| P (α) , (8.257)

where |α is a coherent state, d2 α denotes infinitesimal area in the α


complex plane,
! "
1 |α|2
P (α) = exp − , (8.258)
π N  N 

and where
e−βω
N  = (8.259)
1 − e−βω
is the expectation value of the number operator N. Thus
 
W (x′ , p′ ) = d2 αP (α) Wα (x′ , p′ ) . (8.260)

where
 ∞  5
6
′ 1 ′ p′ x′′ ′ x′′
x′′
Wα (x , p ) = exp i x − |α α

x′ + dx′′ ,
2π −∞  2 2
(8.261)

which is the Wigner function of a coherent state |α, was found to be


given by [see Eq. (8.249)]
 2  2
x′ −x α p′ −p α
′ 1 − 12
′ ∆xα − 12 ∆pα
W (x , p ) = e , (8.262)
π
where

Eyal Buks Quantum Mechanics - Lecture Notes 259


Chapter 8. Density Operator
0
2 ′
xα = α| x |α = α = 2∆xα α′ , (8.263)


pα = α| p |α = 2mωα′′ = 2∆pα α′′ , (8.264)
α′ = Re (α) , (8.265)
α′′ = Im (α) , (8.266)
. 0
2 
∆xα = α| (∆x) |α = , (8.267)
2mω
. 0
2 mω 
∆pα = α| (∆p) |α = = . (8.268)
2 2∆xα
Thus W (x′ , p′ ) is given by
 2  2
  x′ −x α p′ −p α
′ ′ 1 2 − |α|
2
− 12 ∆xα − 12 ∆pα
W (x , p ) = 2 d αe N e
π N 
  ′ ′ 2
  ′ ′ 2
1 α′2 1 x −2∆xα α
′ − N − 2
′′2
− α − 1 p −2∆x αα
= 2 dα e ∆xα
dα′′ e N 2 ∆pα
.
π N 
(8.269)
With the help of the identity (5.136) one thus finds that
  2  ′ 2
x′ p
′ 1′ 1 − 12 2N +1
1
∆xα + ∆p α
W (x , p ) = e , (8.270)
π 2 N  + 1
where
 
e−βω βω
2 N  + 1 = 1 + 2 = coth , (8.271)
1 − e−βω 2

and where β = 1/kB T .


17. With the help of Eq. (5.123) one finds that the wavefunction of the num-
ber state |n = 1 is given by
√ x′  ′2 
x
2 x0 exp − 2x 2
′ ′ 0
ψn=1 (x ) = x |n = 1 = 1/2
, (8.272)
π 1/4 x0

where
0

x0 = . (8.273)

thus
′′ 2 ′′ 2
′′

(x′ − x2 ) ′′

(x′ + x2 )
 ∞ x′ − x2 2x2 x′ + x2 2x2
1 p′ x′′ x0 e 0
x0 e 0
W (x′ , p′ ) = ei  dx′′ ,
π −∞ π1/2 x0

Eyal Buks Quantum Mechanics - Lecture Notes 260


8.4. Solutions

(8.274)

or
 ′ 2 ! "
− xx  ∞ 2
e 0 1 x′ X2 ip′ X2
′ ′
W (x , p ) = − e p0 X− 4
dX , (8.275)
π π1/2 −∞ x0 4

where X = x′′ /x0 and where p0 = /x0 . The integration, which is per-
formed with the help of Eq. (5.136), yields
   
   ′ 2 
2
′ ′ 2 − xx0′ 2 − pp0′ 2 x′ p 1
W (x , p ) = e + − . (8.276)
π x0 p0 2

Note that near the origin the Wigner function W (x′ , p′ ) becomes nega-
tive.
18. The relation (8.64) is proven by
 ∞
−1
 dp′ W (x′ , p′ )
−∞
 ∞5

6  ∞
′ x′′

′ x′′ ′′ 1 p′ x′′
= x −
ρ
x + dx dp′ ei 
−∞ 2 2 2π −∞
  
δ(x′′ )
′ ′
= x | ρ |x  .
(8.277)
With the help of the identities (3.45) and (3.52) W (x′ , p′ ) can be ex-
pressed as
W (x′ , p′ )
 ∞  ∞  ∞ ′′ ′′
p′ x′′ +p′′ (x′ − x )−p′′′ (x′ + x )
1 ′′ ′′ ′′′ i 2 2
= dx dp dp e  p′′ | ρ |p′′′ 
(2π)2  −∞ −∞ −∞
 
 ∞  ∞ x′ (p′′ −p′′′ )
 ∞ x′′ p′ −
p′′

p′′′
1 ′′ ′′′ i ′′ ′′′ 1 ′′ i
2 2
= dp dp e  p | ρ |p  dx e  .
2π −∞ −∞ 2π −∞
 
 

′′ ′′′
δ p′ − p2 − p2

(8.278)
The above result easily leads to (8.65)

Eyal Buks Quantum Mechanics - Lecture Notes 261


Chapter 8. Density Operator
 ∞
−1 dx′ W (x′ , p′ )
−∞
 ∞  ∞  
p′′ p′′′
= dp′′ dp′′′ p′′ | ρ |p′′′  δ p′ − −
−∞ −∞ 2 2
 ∞ x′ (p′′ −p′′′ )
1
dx′ ei 
2π −∞
  
δ(p′′ −p′′′ )
′ ′
= p | ρ |p  .
(8.279)
19. For a pure state ρ = |α α| one finds that the Wigner function is given
by [see Eq. (8.63)]
∞    
1 p′ x′′ x′′ x′′

W (x , p ) = ′
dx′′ ei  ψα x′ − ψ∗α x′ + , (8.280)
2π 2 2
−∞

where ψα (x′ ) = x′ |α is the position wavefunction of |α. Thus with
the help of Schwartz inequality one finds that
∞
′ ′′  
2
′ ′ 2 1

′′
i p x ′ x′′

|W (x , p )| ≤ dx
e ψ x −
(2π)2
α
2

−∞
∞
 
2

x′′

× dx

ψ∗α x′ +
′′
,
2

−∞
(8.281)
thus
1
|W (x′ , p′ )| ≤ . (8.282)

20. The Hamiltonian operator of the system is given by

p2
H= + V (x) , (8.283)
2m
where p is the canonical conjugate operator to the position operator x.
With the help of Eq. (8.35), which reads

dρ 1
= − [ρ, H] , (8.284)
dt i
one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 262


8.4. Solutions

∞ 5

6
dW 1 ip′ x′′ x′′

′ x′′
= dx′′ e  x′ − x +
dt 2π 2
dt
2
−∞
= Sk + Sp ,
(8.285)
where the term
∞ 5

6
1 ′′ ip′ x′′
′ x′′

p2

′ x′′
Sk = − dx e  x − ρ, x + (8.286)
2πi 2
2m
2
−∞

represents the contribution of the kinetic energy, and where the term
∞ 5

6
1 ip′ x′′ x′′


′ x′′
Sp = − dx′′ e  x′ − [ρ, V (x)]
x + (8.287)
2πi 2

2
−∞

represents the contribution of the potential energy. To evaluate the term


Sk , the closure relation (2.23) is employed

1= |φn  φn | , (8.288)
n

where {|φn } is an arbitrary orthonormal basis, in order to obtain the


following relation
5

6
′ x′′

p2

′ x′′
x − ρ, x +
2
2m
2
5

6
1  ′ x′′

2
′ x′′
= x − ρ |φn  φn | p
x +
2m n 2
2
 5

′′


6
1 ′ x
2


′ x′′
− x − p |φn  φn | ρ
x + .
2m n 2
2
(8.289)
By introducing the variables
x′′
x′′′ = x′ − , (8.290)
2
′′
x
x′′′′ = x′ + , (8.291)
2
and employing the relations [see Eq. (3.29)]
5

x′′

2 d2
x′ −
p |φn  = −2 ′′′2 x′′′ |φn  , (8.292)
2 dx

′′
6

x d2
φn | p2

x′ + = −2 ′′′′2 x′′′′ |φn ∗ , (8.293)


2 dx

Eyal Buks Quantum Mechanics - Lecture Notes 263


Chapter 8. Density Operator
<
one finds that (after removing the factor n |φn  φn |)
5

6
′ x′′

p2

′ x′′
x − ρ, x +
2
2m
2
 
2 d2 d2
= − x′′′ | ρ |x′′′′  .
2m dx′′′2 dx′′′′2
(8.294)
Using the relations
d ∂x′ d ∂x′′ d 1 d d
′′′
= ′′′ ′
+ ′′′ ′′
= ′
− ′′ , (8.295)
dx ∂x dx ∂x dx 2 dx dx
d ∂x′ d ∂x′′ d 1 d d
′′′′
= ′′′′ ′
+ ′′′′ ′′ = ′
+ ′′ , (8.296)
dx ∂x dx ∂x dx 2 dx dx
one finds that
d2 d2 d2
− = −2 , (8.297)
dx′′′2 dx′′′′2 dx′ dx′′
and thus
5

6
′ x′′

p2

′ x′′
x − ρ, x +
2
2m
2
5

6
2 d2 ′ x ′′


ρ
x′ + x
′′
=− x − .
m dx′ dx′′ 2

2
(8.298)
Inserting into Eq. (8.286) and integrating by parts with respect to x′′
lead to
∞ 5

6
 ′ ′′
′′ ip x d2 ′ x′′

′ x′′
Sk = dx e x − ρ x +
2πim dx′ dx′′ 2

2
−∞

′ ∞ 5

6
p 1 ′′ ip′ x′′ d ′ x′′

′ x′′
=− dx e  x − ρ x + ,
m 2π dx′ 2

2
−∞
(8.299)
or [see Eq. (8.63)]
p′ ∂W
Sk = − . (8.300)
m ∂x′
To evaluate Sp the following relation is employed
5

6
x′′


′ x′′
x′ − [ρ, V (x)]
x +
2

2
  ′′
   5

6
x x′′ x′′

′ x′′
= V x′ + − V x′ − x′ − ρ x + .
2 2 2

2
(8.301)

Eyal Buks Quantum Mechanics - Lecture Notes 264


8.4. Solutions

The term V (x′ + x′′ /2)−V (x′ − x′′ /2), which represents an odd function
of x′′ , can be Taylor expanded as
     ′′ 2l+1
x′′ x′′ <∞ 1 ∂ 2l+1 V x
V x′ + − V x′ − =2 2l+1
.
2 2 l=0 (2l + 1)! ∂ (x′ ) 2
(8.302)

As can be seen from Eq. (8.63), the following holds

∞ 5

6  2l+1 2l+1
1 2l+1 ip′ x′′ x′′

′ x′′  ∂ W
dx′′ (x′′ ) e  x′ − ρ x + = .
2π 2

2 i ∂ (p′ )2l+1
−∞
(8.303)

Employing the above results to evaluate Eq. (8.287) and separating the
first term from all higher order terms yield
  2l
∂V ∂W <∞
2i ∂ 2l+1 V ∂ 2l+1 W
Sp = + 2l+1
. (8.304)
∂x′ ∂p′ l=1 (2l + 1)! ∂ (x′ ) ∂ (p′ )2l+1

Combining Eqs. (8.300) and (8.304) yields


  2l
dW p′ ∂W ∂V ∂W < ∞
2i ∂ 2l+1 V ∂ 2l+1 W
=− + + , (8.305)
dt m ∂x′ ∂x′ ∂p′ l=1 (2l + 1)! ∂ (x′ )2l+1 ∂ (p′ )2l+1

or in terms of the Poisson’s brackets [see Eqs. (1.37) and (8.283)]


  2l
dW <∞
2i ∂ 2l+1 V ∂ 2l+1 W
= {H, W } + 2l+1
. (8.306)
dt l=1 (2l + 1)! ∂ (x′ ) ∂ (p′ )2l+1

Note that when ∂ 2l+1 V/∂ (x′ )2l+1 = 0 for l ≥ 1 the above result coin-
cides with the classical equation of motion dW/dt = {H, W }. Thus one
concludes that the quantum time evolution of W of a harmonic oscillator
is identical to the classical one.
21. With the help of Eq. (2.175) one finds that
iξη
exp (−iξX − iηP ) = e− 2 e−iηP e−iξX . (8.307)

By evaluating the trace in Eq. (8.68) using an orthonormal basis of di-


mensionless position eigenstates (i.e. X |X ′  = X ′ |X ′ ) one finds that
[see Eq. (3.19) and recall that Tr (AB) = Tr (BA)]

Eyal Buks Quantum Mechanics - Lecture Notes 265


Chapter 8. Density Operator

∞
− iξη
W̃ (ξ, η) = e 2 dX ′ X ′ | ρe−iηP e−iξX |X ′ 
−∞
∞
iξη ′
= e− 2 dX ′ e−iξX X ′ | ρ |X ′ + η
−∞
∞
′′
' η

η(
= dX ′′ e−iξX X ′′ −
ρ
X ′′ + ,
2 2
−∞
(8.308)
thus
∞ ∞
1 ′
+iηP ′
W (X ′ , P ′ ) = dξdη W̃ (ξ, η) eiξX
(2π)2
−∞ −∞
∞ ∞
1 ' η

η ( iηP ′
= dη dX ′′ X ′′ −
ρ
X ′′ + e
2π 2 2
−∞ −∞
∞
1 ′′ ′
× dξe−iξ(X −X )

−∞
  
δ(X ′′ −X ′ )
∞ 5

6
1 ′′ ′ X ′′

′ X ′′ ′′ ′
= dX X − ρ X + eiX P .
2π 2

2
−∞
(8.309)
′ ′
22. For the case X = P = 0 the operator Υ is given by
∞ ∞
1
Υ (0, 0) = dξdη e−iξX−iηP , (8.310)
(2π)2
−∞ −∞

or [see Eq. (2.175) and recall that [X, P ] = i]


∞ ∞
1 iξη
Υ (0, 0) = 2 dξdη e− 2 e−iηP e−iξX , (8.311)
(2π)
−∞ −∞

and thus the following holds

Eyal Buks Quantum Mechanics - Lecture Notes 266


8.4. Solutions

∞ ∞
′ 1 iξη ′
Υ (0, 0) |X  = dξdη e− 2 e−iηP e−iξX |X ′ 
(2π)2
−∞ −∞
∞ ∞
1 1 η ′
= dη e −iηP ′
|X  dξ e−iξ( 2 +X )
2π 2π
−∞ −∞
  
δ( η2 +X ′ )

= π−1 e2iX P
|X ′  ,
(8.312)
or [see Eq. (3.19)]
Υ (0, 0) |X ′  = π −1 |−X ′  , (8.313)
which implies that Υ (0, 0) is related to the parity operator P (8.75) by
Υ (0, 0) = π−1 P . (8.314)
By employing the relations
a + a† a − a†
X= √ , P = √ , (8.315)
2 2i
together with Eq. (8.310) one finds that
∞ ∞
1 −iξ a+a


−iη a−a

P= dξdη e 2 2i

−∞ −∞
∞ ∞
1 − iξ+η
√ a− iξ−η
√ a†
= dξdη e 2 2 .

−∞ −∞
(8.316)
The above result together with Eq. (5.38) yield
   
X ′ + iP ′ X ′ + iP ′
π−1 D† − √ PD − √
2 2
∞ ∞  ′ +iP ′   ′ −iP ′ 
1 − iξ+η
√ a− X √ − iξ−η
√ a† − X √
= dξdη e 2 2 2 2

(2π)2
−∞ −∞
∞ ∞  ′ ′
  ′ −iP ′

1 − iξ+η
√ X+iP
√ −X √
+iP
− iξ−η
√ X−iP
√ −X √
= dξdη e 2 2 2 2 2 2

(2π)2
−∞ −∞
∞ ∞
1 ′
−X )+iη (P ′ −P )
= dξdη eiξ(X ,
(2π)2
−∞ −∞
(8.317)

Eyal Buks Quantum Mechanics - Lecture Notes 267


Chapter 8. Density Operator

thus [see Eq. (8.72)]


   
−1 † X ′ + iP ′ X ′ + iP ′
π D − √ PD − √ = Υ (X ′ , P ′ ) . (8.318)
2 2
23. The normalized homodyne observable Xφ can be expressed in terms of
the dimensionless position and momentum operators X and P , which are
given by

a + a† a − a†
X= √ , P = √ , (8.319)
2 i 2
and which satisfy [X, P ] = i [see Eq. (5.13)], as
X − iP iφ X + iP −iφ
Xφ = e + e
2 2
= X cos φ + P sin φ .
(8.320)

The associated dimensionless momentum operator Pφ is defined as Pφ =


−X sin φ + P cos φ, i.e.
    
Xφ cos φ sin φ X
= , (8.321)
Pφ − sin φ cos φ P

and the inverse transformation is given by


    
X cos φ − sin φ Xφ
= . (8.322)
P sin φ cos φ Pφ

a) The expectation value


) −iζXφ *
e = Tr [exp (−iζXφ ) ρ] (8.323)

is the
 characteristic
 function of the probability distribution
 function
′ ′ ′
Pr Xφ of Xφ , which is denoted below as w Xφ , and thus it is
 to the Fourier transform w̃φ (ζ) of the probability distribution
related
w Xφ′ by

∞
) −iζXφ *   ′
e = dXφ′ w Xφ′ e−iζXφ = w̃φ (ζ) . (8.324)
−∞

On the other hand, the Fourier transform W̃ (ξ, η) of the Wigner


function W (X ′ , P ′ ) is given by [see Eq. (8.68)]

W̃ (ξ, η) = Tr [exp (−iξX − iηP ) ρ] . (8.325)

Eyal Buks Quantum Mechanics - Lecture Notes 268


8.4. Solutions

The comparison between Eq. (8.323) and Eq. (8.325) yields the fol-
lowing relation (recall that Xφ = X cos φ + P sin φ)

W̃ (ζ cos φ, ζ sin φ) = w̃φ (ζ) . (8.326)

Applying the inverse Fourier transform to Eq. (8.324) leads to [see


Eq. (8.67)]
∞
 ′ 1 ′
w Xφ = dζ eiζXφ w̃φ (ζ)

−∞
∞
1 ′
= dζ eiζXφ W̃ (ζ cos φ, ζ sin φ) ,

−∞
(8.327)
and thus by employing the inverse Fourier transform to Eq. (8.67)
∞ ∞
  ′ ′
−iη′ P ′
W̃ ξ ′ , η′ = dX ′ dP ′ e−iξ X W (X ′ , P ′ ) , (8.328)
−∞ −∞

one finds that


∞ ∞ ∞
  1 ′
cos φ+P ′ sin φ−Xφ

w Xφ′ = dζdX ′ dP ′ e−iζ (X ) W (X ′ , P ′ ) .

−∞−∞ −∞
(8.329)

The variable transformation (8.322) leads to


∞ ∞
 ′  
w Xφ = dXφ′′ dPφ′ W Xφ′′ cos φ − Pφ′ sin φ, Xφ′′ sin φ + Pφ′ cos φ
−∞−∞
∞
1 ′′ ′
× dζ e−iζ (Xφ −Xφ )

−∞
  
δ (Xφ′′ −X ′
φ )
∞
 
= dPφ′ W Xφ′ cos φ − Pφ′ sin φ, Xφ′ sin φ + Pφ′ cos φ .
−∞
(8.330)
b) The relation (8.326) allows evaluating the Wigner function, which is
related to its Fourier transformed function W̃ (ξ, η) by Eq. (8.67),
using the so-called inverse Radon transform. In polar coordinates
(8.67) becomes

Eyal Buks Quantum Mechanics - Lecture Notes 269


Chapter 8. Density Operator

∞ π
1 ′
cos φ+P ′ sin φ)

W (X , P ) = ′
dζ dφ |ζ| W̃ (ζ cos φ, ζ sin φ) eiζ (X ,
(2π)2
−∞ 0
(8.331)

thus with the help of Eq. (8.326) one finds that

∞ π
1 ′
cos φ+P ′ sin φ)
′ ′
W (X , P ) = 2 dζ dφ |ζ| w̃φ (ζ) eiζ (X ,
(2π)
−∞ 0
(8.332)

thus [see Eq. (8.324)]

∞ π ∞
1   ′ ′ ′

W (X , P ) = ′
dζ dφ dXφ′ |ζ| w Xφ′ eiζ (X cos φ+P sin φ−Xφ ) .
(2π)2
−∞ 0 −∞
(8.333)

24. The density operator [see Eq. (8.51)] is given by


 
ρ= d2 α |α α| P (α) , (8.334)

where |α is a coherent state, d2 α denotes infinitesimal area in the α


complex plane,
! "
1 |α|2
P (α) = exp − , (8.335)
π N  N 

and where
e−βω
N  = (8.336)
1 − e−βω
is the expectation value of the number operator N . By employing the
expression for the wave function ψα (x′ ) = x′ |α of a coherent state
which is given by [see Eq. (5.51)]
ψα (x′ ) = x′ |α
 ∗2    2 
α − α2  mω 1/4 x′ − xα x′
= exp exp − + i pα ,
4 π 2∆xα 
(8.337)
where

Eyal Buks Quantum Mechanics - Lecture Notes 270


8.4. Solutions
0
2 ′
xα = α| x |α = α , (8.338)


pα = α| p |α = 2mωα′′ , (8.339)

α = Re (α) , (8.340)
α′′ = Im (α) , (8.341)
. 0
2 
∆xα = α| (∆x) |α = , (8.342)
2mω
one finds that
 
x′′ | ρ |x′  = d2 αP (α) x′′ |α α |x′ 
 mω 1/2   ! "
π 2 |α|2
= d α exp −
π N  N 
   2   
x′ − xα x′′ − xα 2 (x′′ − x′ )
× exp − − + i pα
2∆xα 2∆xα 
 mω 1/2    
α′2 + α′′2
= π d2 α exp −
π N  N 
    ′′ 2 
′ ′ 2
X − 2α X − 2α′ ′′ ′′ ′
× exp − − + iα (X − X )
2 2
 mω 1/2   
π ′ 2 N + 1 ′2 ′ ′′ ′ X ′2 + X ′′2
= dα exp − α + (X + X ) α −
π N  N 4
  ′′2

α
× dα′′ exp − + iα′′ (X ′′ − X ′ ) .
N 
(8.343)
where
0
′ 2mω ′
X = x , (8.344)

0
2mω ′′
X ′′ = x . (8.345)

With the help of the identity (5.136) one finds that
: ′
′2 +X ′′2 −N (X ′′ −X ′ )2 + N (X +X )
′′ 2
 mω 1/2 1 −(X ) 2N +1
′′ ′
x | ρ |x  = e 4 ,
π 2 N + 1
(8.346)
or using the identity
(X ′ + X ′′ )2 + (X ′ − X ′′ )2
X ′2 + X ′′2 = , (8.347)
2

Eyal Buks Quantum Mechanics - Lecture Notes 271


Chapter 8. Density Operator

one has
:  
X ′ +X ′′ 2  ′ 
 mω 1/2 1 2 ′′ 2 2N +1
′′ ′ − − X −X
x | ρ |x  = e 2(2N +1) 2 2
.
π 2 N + 1
(8.348)

In terms of x′ and x′′ this result can be written as


 2  2
1 − x′ +x′′ x′ −x′′
−(2N+1)2
x′′ | ρ |x′  = √ e 2ξ 2ξ
, (8.349)
ξ π
where
0

ξ= (2 N + 1) , (8.350)

 
e−βω βω
2 N  + 1 = 1 + 2 = coth , (8.351)
1 − e−βω 2
and where β = 1/kB T . Alternatively, the result can be expressed as
 ′ ′′  2  ′ ′′  2
− tanh( β2ω ) x 2x
+x
−coth( β2ω ) x 2x
−x
e 0 0
x′′ | ρ |x′  = 0   , (8.352)
x0 π coth βω
2

where
0

x0 = . (8.353)

25. The following holds [see Eqs. (8.8), (8.42)]
 
) −iζXφ * Tr e−βH e−iζXφ
e = , (8.354)
Tr (e−βH )
 
where H = ω a† a + 1/2 [see , Eq.- (5.16)]. Using the identity (2.173)
and the commutation relation a, a† = 1 [see Eq. (5.13)] one finds that
the following holds
† †
eiφa a ae−iφa a
= ae−iφ , (8.355)
† †
eiφa a a† e−iφa a
= a† eiφ , (8.356)
and thus Xφ can be expressed as
† †
Xφ = eiφa a X0 e−iφa a
, (8.357)

where X0 , which is given by

Eyal Buks Quantum Mechanics - Lecture Notes 272


8.4. Solutions

a + a†
X0 = √ , (8.358)
2
is related to the position operator x by [see Eq. (5.11)]

x = x0 X0 , (8.359)

where
0

x0 = . (8.360)

The last result implies that
† †
e−iζXφ = eiφa a e−iζX0 e−iφa a
, (8.361)

and thus [see Eq. (8.354) and recall that Tr (AB) = Tr (BA)]
) −iζXφ * ) −iζX0 * ' −iζ √ mω x (
e = e = e  . (8.362)
) −iζX *
In other words, e φ
is found to be independent on φ. The expecta-
' √ mω (
−iζ x
tion value e  can be calculated by employing the expression
for the matrix elements x′′ | ρ |x′  of the density operator ρ given by Eq.
(8.352)
∞ √ mω
) −iζXφ *
e = dx′ x′ | ρe−iζ  x
|x′ 
−∞
 ′ 2
∞ √ mω − tanh( β2ω ) xx
′ e 0
= dx′ e−iζ  x
0  ,
βω
−∞ x0 π coth 2

(8.363)
and where β = 1/kB T . Using the the identity (5.136) one finds that

) −iζXφ * ζ 2 coth( β ω )
2

e =e 4 . (8.364)

In view of Eq. (8.160), according to which


 
1 βω 1 ) *
coth = N  + = X02 , (8.365)
2 2 2

where [see Eq. (8.336)]

e−βω
N  = , (8.366)
1 − e−βω

Eyal Buks Quantum Mechanics - Lecture Notes 273


Chapter 8. Density Operator

the above result can be rewritten as


) −iζX * ζ 2 X02

e φ
= e− 2 . (8.367)
) −iζX *
The factor e φ
is thecharacteristic function of the probability
 dis-

tribution function Pr Xφ′ of Xφ′ , which is denoted below as w Xφ′ ,
and thus it is related
  to the Fourier transform w̃φ (ζ) of the probability
distribution w Xφ′ by [see Eq. (8.324)]

∞
) −iζX *   ′
e φ
= dXφ′ w Xφ′ e−iζXφ = w̃φ (ζ) . (8.368)
−∞

With the help of Eqs. (8.324), (8.326) and (8.367) one finds that the
Fourier transform W̃ (ξ, η) of the Wigner function W (X ′ , P ′ ) satisfy the
following relation for any real φ

 
ζ 2 X02

W̃ (ζ cos φ, ζ sin φ) = e− 2 , (8.369)

and thus
(ξ2 +η2 )X02 
W̃ (ξ, η) = e− 2 . (8.370)

The inverse Fourier transformation [see Eqs. (5.136) and (8.309)] yields
 ∞ ξ2 X02
 iξX ′ 1  ∞ η 2 X02
 iηP ′
′ ′ 1 − −
W (X , P ) = dξ e 2 e dη e 2 e
2π −∞ 2π −∞
′2 ′2
1 − X +P2
2X0
= e
2π X02 
1 X ′2 +P ′2
= e− 2N +1 .
π (2 N  + 1)
(8.371)
It is easy to see that the above result for the Wigner function W (X ′ , P ′ )
for the dimensionless variables X ′ and P ′ is consistent with Eq. (8.270)
for the Wigner function W (x′ , p′ ) for the displacement x′ and momentum
p′ variables.
26. By using Eqs. (8.67) and employing cylindrical coordinates

ξ = ζ cos φ , η = ζ sin φ , (8.372)


X ′ = R′ cos θ , P ′ = R′ sin θ ,

one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 274


8.4. Solutions

∞ π
′ ′ 1 ′
W (X , P ) = dζ ζ dφ W̃ (ζ cos φ, ζ sin φ) eiζR cos(φ−θ)
,
(2π)2
0 −π
(8.373)
 
thus since w Xφ′ is φ independent [see Eqs. (8.324) and (8.326)] one
has [note that contrary to Eq. (8.333), integration over ζ below is taken
to be over positive values only]
∞ π
′ ′ 1 ′
W (X , P ) = dζ ζ w̃ (ζ) dφ eiζR cos(φ−θ)
, (8.374)
(2π)2
0 −π

where
∞
  ′
w̃ (ζ) = dXφ′ w Xφ′ e−iζXφ . (8.375)
−∞

With the help of Jacobi-Anger expansion




exp (iz cos x) = in Jn (z) einx , (8.376)
n=−∞

one finds that


π ∞
 π
iζR′ cos(φ−θ) n ′ −inθ
dφ e = i Jn (ζR ) e dφ einφ
−π n=−∞ −π

= 2πJ0 (ζR ) ,
(8.377)

thus
∞   
′ ′1
W (X , P ) = dζ ζ w̃ (ζ) J0 ζ X ′2 + P ′2 , (8.378)

0

or
/∞  
dz w̃ √ z zJ0 (z)
X ′2 +P ′2
′ ′ 0
W (X , P ) = . (8.379)
2π (X ′2 + P ′2 )

As an example of the usage of Eq. (8.378), consider the case of a harmonic


oscillator having angular resonance frequency ω in thermal equilibrium
at temperature T . For this case [see Eq. (8.367)]

Eyal Buks Quantum Mechanics - Lecture Notes 275


Chapter 8. Density Operator

 
ζ 2 X02

w̃ (ζ) = e− 2 , (8.380)

and thus with the help of the identity


∞
z2 A2 1 − R22
ze− 2 J0 (zR) dz = e 2A , (8.381)
A2
0

one finds that Eq. (8.378) yields


′2 ′2
1 − X +P
W (X ′ , P ′ ) = e 2X 2 
0 , (8.382)
2π X02 

in agreement with Eq. (8.371).


27. The density operator evolves in time according to [see Eq. (8.37)]

ρ (t) = u (t) ρ0 u† (t) , (8.383)

where u (t) is the time evolution operator for the system. The Wigner
function W (X ′ , P ′ ; t) can be expressed according to Eq. (8.333) as
∞ π ∞
1   ′ ′ ′
′ ′
W (X , P ; t) = 2 dζ dφ dXφ′ |ζ| w Xφ′ eiζ (X cos φ+P sin φ−Xφ ) ,
(2π)
−∞ 0 −∞
(8.384)
 
where w Xφ′ is the probability distribution function of the observable
Xφ , which is given by

a† eiφ + ae−iφ
Xφ = √ . (8.385)
2
In terms of the density operator ρ (t) one has [recall that Tr (AB) =
Tr (BA)]
  
* )

w Xφ′ = Tr
Xφ′ Xφ′
ρ (t)

*)

= Tr u† (t)
Xφ′ Xφ′
u (t) ρ0 ,
(8.386)

(
(

where
Xφ′ is an eigenvector of Xφ with eigenvalue Xφ′ , i.e. Xφ
Xφ′ =

(

Xφ′
Xφ′ . With the help of Eqs. (2.173) and (4.9) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 276


8.4. Solutions

† iφ
† + ae−iφ −iω(a† a+ 12 )t
a+ 12 )t a e
u† (t) Xφ u (t) = eiω(a √ e
2
† iφ
† a e + ae−iφ −iωta† a
= eiωta a √ e
2
a† ei(φ+ωt) + ae−i(φ+ωt)
= √
2
= Xφ+ωt ,
(8.387)
thus the following holds

*
*
Xφ+ωt u† (t)
Xφ′ = Xφ′ u† (t)
Xφ′ , (8.388)

(

i.e. u† (t)
Xφ′ is an eigenvector of Xφ+ωt with eigenvalue Xφ′ . This eigen-

(


vector is labeled below as
Xφ+ωt . Using these results and the trigono-
metric identities
cos (x + y) = cos x cos y − sin x sin y , (8.389)
sin (x + y) = cos x sin y + sin x cos y , (8.390)
the Wigner function W (X ′ , P ′ ; t) becomes
∞ π ∞
′ ′ 1
W (X , P ; t) = dζ dφ dXφ′ |ζ|
(2π)2
−∞ 0 −∞

′ *) ′
 ′ ′ ′
× Tr
Xφ+ωt Xφ+ωt
ρ0 eiζ (X cos φ+P sin φ−Xφ )
∞ π ∞
1 ′
= dζ dφ dXφ′ |ζ|
(2π)2
−∞ 0 −∞

′ * ) ′
 iζ (X ′ cos(φ′ −ωt)+P ′ sin(φ′ −ωt)−X ′ )
× Tr
Xφ′ Xφ′
ρ0 e φ ,
(8.391)
thus

W (X ′ , P ′ ; t) = W (X ′ cos (ωt) − P ′ sin (ωt) , X ′ sin (ωt) + P ′ cos (ωt) ; 0) ,


(8.392)

i.e. the time evolution of the Wigner function represents rigid rotation in
phase space at angular velocity ω.
28. The Wigner function W (X ′ , P ′ ) can be expressed as [see Eq. (8.67)]
∞ ∞
′ ′1 ′
+iηP ′
W (X , P ) = dξdη W̃ (ξ, η) eiξX , (8.393)
(2π)2
−∞ −∞

Eyal Buks Quantum Mechanics - Lecture Notes 277


Chapter 8. Density Operator

where

W̃ (ξ, η) = Tr (D (α) ρ) , (8.394)

and where
ξ + iη
α= √ , (8.395)
2i
thus for the displaced system one has [see Eq. (5.41) and recall that
Tr (AB) = Tr (BA)]
∞ ∞
′ ′ 1   ′ ′
Wα (X , P ) = 2 dξdη Tr D (α) D (α′ ) ρD† (α′ ) eiξX +iηP
(2π)
−∞ −∞
∞ ∞
1 ′∗
−α∗ α′ ′
+iηP ′
= dξdη eαα Tr (D (α) ρ) eiξX ,
(2π)2
−∞ −∞
(8.396)
thus using
 ′ ′∗
   ′ ′∗ 
−α
′∗
−α∗ α′ iξX ′ +iηP ′ iξ X ′ − α +α
√ +iη P ′ − α √
eαα e =e 2 2i (8.397)

one finds that

Wα (X ′ , P ′ ) = W (X ′ − Xα′ ′ , P ′ − Pα′ ′ ) , (8.398)

where
α′ + α′∗
Xα′ ′ = √ , (8.399)
2
α′ − α′∗
Pα′ ′ = √ . (8.400)
i 2
29. In general, the convolution theorem states that
∞ ∞
1 ′′ ′′ ′′ iX ′′ P ′
dX f (X ) g (X ) e = dP ′′ F (P ′′ ) G (P ′ − P ′′ ) ,

−∞ −∞
(8.401)
where F (P ′ ) (G (P ′ )) is the Fourier transform of f (X ′′ ) (g (X ′′ )), i.e.
∞
′ 1 ′′ ′
F (P ) = dX ′′ f (X ′′ ) eiX P , (8.402)

−∞
∞
1 ′′
P′
G (P ′ ) = dX ′′ g (X ′′ ) eiX , (8.403)

−∞

Eyal Buks Quantum Mechanics - Lecture Notes 278


8.4. Solutions

thus with the help of the identity


P′ 2 ∞  ′′  2
e−( η ) 1 ′′ − ηX iX ′′ P ′
√ = dX e 2
e , (8.404)
πη 2π
−∞

one finds that the reduced Wigner function WR (X ′ , P ′ ) is related to


W (X ′ , P ′ ) by [see Eq. (8.82)]
∞ P ′ −P ′′ 2

′ ′ ′′ e−( η
′ ′′
)
WR (X , P ) = dP W (X , P ) √ . (8.405)
πη
−∞

As an example, for the case where W (X ′ , P ′ ) is normally distributed


according to
 ′2 ′2

exp − X δ+P2
W (X ′ , P ′ ) = , (8.406)
πδ 2
where δ is a constant, Eq. (8.405) yields
 2
 ′ 2 P′
− Xδ − √
η2 +δ2
e e
WR (X ′ , P ′ ) = √ √  2 . (8.407)
πδ π η + δ2
30. The dynamics along the x direction is governed by the Hamiltonian Hx
of a harmonic oscillator
p2x 1
Hx = + mω 2 x2 . (8.408)
2m 2
) *
By symmetry x = 0. The expectation value x2 was calculated in Eq.
(8.160) and found to be given by
 
) 2* 1 ω ωβ
x = coth . (8.409)
mω 2 2 2
31. In general, for any smooth function f (ρ) of ρ the following holds

f (ρ (t)) = u (t, t0 ) f (ρ (t0 )) u† (t, t0 ) . (8.410)

This can be shown by Taylor expanding the function f (ρ) as a power


series


f (ρ (t)) = an (ρ (t))n , (8.411)
n=0

using Eq. (8.37) and the fact that u† (t, t0 ) u (t, t0 ) = 1, i.e. the unitarity
of the time evolution operator. By using this result for the function ρ log ρ

Eyal Buks Quantum Mechanics - Lecture Notes 279


Chapter 8. Density Operator

together with the general identity Tr (XY ) = Tr (Y X) [see Eq. (2.129)]


one easily finds that σ is time independent. This somewhat surprising
result can be attributed to the fact that the unitary time evolution that is
governed by the Schrödinger equation is symmetric under time reversal.
In the language of statistical mechanics it corresponds to a reversible
process, for which entropy is preserved.
32. Using the definition of the Pauli matrices (6.136) one finds that
 
1 1 + kz kx − iky
ρ= , (8.412)
2 kx + iky 1 − kz
and
 
1 1 + 2kz + k2 2 (kx − iky )
ρ2 = ,
4 2 (kx + iky ) 1 − 2kz + k2

where k2 = kx2 + ky2 + kz2 .


a) Note that for any k the following holds Tr (ρ) = 1. The requirement
that ρ is Hermitian, i.e. the requirement
 that ρ† = ρ, implies that
∗ ∗ ∗
kz = kz and kx − kx + i ky − ky = 0, thus kx , ky and kz are all real.
   
Moreover, the requirement that Tr ρ2 = 1 + k2 /2 ≤ 1 implies
that k2 ≤ 1.  
b) For this case Tr ρ2 = 1, thus k2 = 1.
c) With the help of Eq. (6.137), which is given by

(σ · a) (σ · b) = a · b + iσ · (a × b) , (8.413)

and the fact that all three Pauli matrices have a vanishing trace, one
finds that
1 1
Tr (û · σρ) = Tr (û · σ) + Tr ((û · σ) (k · σ))
2 2
1
= Tr ((û · σ) (k · σ))
2
1 i
= Tr (û · k) + Tr (σ · (û × k))
2 2
1
= Tr (û · k)
2
= û · k .
(8.414)
33. Using the definition of the Pauli matrices (6.136) one finds that
 
1 1 + kz kx − iky
ρ= . (8.415)
2 kx + iky 1 − kz
a) Let λ± be the two eigenvalues of ρ. The following holds

Tr (ρ) = λ+ + λ− = 1 , (8.416)

Eyal Buks Quantum Mechanics - Lecture Notes 280


8.4. Solutions

and
 
Det (ρ) = λ+ λ− = 1 − k2 /4 , (8.417)

where k2 = kx2 + ky2 + kz2 . Thus

1 ± |k|
λ± = , (8.418)
2
and therefore

σ = f (|k|) , (8.419)

where
1−x 1−x 1+x 1+x
f (x) = − log − log . (8.420)
2 2 2 2

0.7

0.6

0.5

0.4
y
0.3

0.2

0.1

0.2 0.4 x 0.6 0.8 1

The function f (x) = − 1−x


2 log 1−x
2 − 1+x
2 log 1+x
2 .

b) As can be seen from Eq. (8.194), after the measurement ρ becomes


diagonal in the basis of eigenvectors of the measured observable,
namely, after the measurement the density matrix is given by
 
1 1 + kz 0
ρc = , (8.421)
2 0 1 − kz

and thus the entropy after the measurement is


1 − kz 1 − kz 1 + kz 1 + kz
σ c = f (kz ) = − log − log . (8.422)
2 2 2 2
34. First, consider a general functional g (ρ) of the density operator having
the form

g (ρ) = Tr (f (ρ)) , (8.423)

Eyal Buks Quantum Mechanics - Lecture Notes 281


Chapter 8. Density Operator

where the function f (ρ) can be Taylor expanded as a power series




f (ρ) = ak ρk , (8.424)
k=0

and where ak are complex constants. Consider an infinitesimal change in


the density operator ρ → ρ + dρ. To first order in dρ the corresponding
change dg in the functional g (ρ) can be expressed as
dg = g (ρ + dρ) − g (ρ)
!∞ "
 3 k
4
k
= Tr ak (ρ + dρ) − ρ
k=0
  

  
2
= Tr  ak ρk−1 dρ + ρk−2 (dρ) ρ + ρk−3 (dρ) ρ2 + · · · + O (dρ) .
  
k=0
k terms
(8.425)
By exploiting the general identity Tr (XY ) = Tr (Y X) the above result
can be simplified (note that generally ρ needs not to commute with dρ)
! ∞ " 
  
dg = Tr ak kρk−1 dρ + O (dρ)2 , (8.426)
k=0

thus to first order in dρ the following holds


 
df
dg = Tr dρ . (8.427)

In the above expression the term df /dρ is calculated by simply taking the
derivative of the function f (x) (where x is considered to be a number)
and substituting x = ρ. Alternatively, the change dg can be expressed in
terms of the infinitesimal change dρnm in the matrix elements ρnm of ρ.
To first order in the infinitesimal variables dρnm one has
 ∂g
dg = dρ . (8.428)
n,m
∂ρnm nm

It is convenient to rewrite the above expression as


¯ · dρ ,
dg = ∇g (8.429)
¯ and of dρ are given
where the vector elements of the nabla operator ∇
by
  ∂
¯
∇ = , (8.430)
n,m ∂ρnm

Eyal Buks Quantum Mechanics - Lecture Notes 282


8.4. Solutions

and
 
dρ n,m = dρnm . (8.431)

Thus, to first order one has


¯ · dρ ,
dσ = ∇σ (8.432)

and
¯ l · dρ ,
dgl = ∇g (8.433)

where l = 0, 1, 2, ...L.
a) In general, the technique of Lagrange multipliers is very useful for
finding stationary points of a function, when constrains are applied.
A stationary point of σ occurs iff for every small change dρ, which is
¯ 0 , ∇g
orthogonal to all vectors ∇g ¯ 1 , ∇g
¯ 2 , ..., ∇g
¯ L (i.e. a change which
does not violate the constrains) one has
¯ · dρ .
0 = dσ = ∇σ (8.434)
¯
 the vector ∇σ belongs to the
This condition is fulfilled only when
¯ ¯ ¯ ¯ L . In other
subspace spanned by the vectors ∇g0 , ∇g1 , ∇g2 , ..., ∇g
words, only when
¯ = ξ 0 ∇g
∇σ ¯ 0 + ξ 1 ∇g
¯ 1 + ξ 2 ∇g
¯ 2 + ... + ξ L ∇g
¯ L, (8.435)

where the numbers ξ 0 , ξ 1 , ..., ξ L , which are called Lagrange multipli-


ers, are constants. By multiplying by dρ the last result becomes

dσ = ξ 0 dg0 + ξ 1 dg1 + ξ 2 dg2 + ... + ξ L dgL . (8.436)

Using Eqs. (8.427), (8.89, (8.91) and (8.92) one finds that
dσ = − Tr ((1 + log ρ) dρ) , (8.437)
dg0 = Tr (dρ) , (8.438)
dgl = Tr (Xl dρ) , (8.439)
thus
! L
" 

0 = Tr 1 + log ρ + ξ 0 + ξ l Xl dρ . (8.440)
l=1

The requirement that the last identity holds for any dρ implies that
L

1 + log ρ + ξ 0 + ξ l Xl = 0 , (8.441)
l=1

thus

Eyal Buks Quantum Mechanics - Lecture Notes 283


Chapter 8. Density Operator
! L
"

−1−ξ0
ρ=e exp − ξ l Xl . (8.442)
l=1

The Lagrange multipliers ξ 0 , ξ 1 , ..., ξ L can be determined from Eqs.


(8.91) and (8.92). The first constrain (8.91) is satisfy by replacing
the factor e−1−ξ0 by the inverse of the partition function Z
! L "
1 
ρ = exp − ξ l Xl . (8.443)
Z
l=1

where
! L
"

Z = Tr − ξ l Xl . (8.444)
l=1

As can be seen from the above expression for Z, the following holds
1   ∂ log Z
Xl  = Tr Xl e−βH = − . (8.445)
Z ∂ξ l

b) For the case of a microcanonical ensemble Eq. (8.443) yields ρ = 1/Z,


i.e. ρ is proportional to the identity operator.
c) For the case of a canonical ensemble Eq. (8.443) yields
1 −βH
ρc = e , (8.446)
Zc
where the canonical partition function Zc is given by
 
Zc = Tr e−βH , (8.447)

and where β labels the Lagrange multiplier associated with the given
expectation value H. By solving Eq. (8.92), which for this case is
given by [see also Eq. (8.445)]

1   ∂ log Zc
H = Tr He−βH = − . (8.448)
Zc ∂β
the Lagrange multiplier β can be determined. Note that the tem-
perature T is defined by the relation β = 1/kB T , where kB is the
Boltzmann’s constant.
d) For the case of a grandcanonical ensemble Eq. (8.443) yields
1 −βH+βµN
ρgc = e , (8.449)
Zgc

where the grandcanonical partition function Zgc is given by

Eyal Buks Quantum Mechanics - Lecture Notes 284


8.4. Solutions
 
Zgc = Tr e−βH+βµN . (8.450)
Here the Lagrange multiplier associated with the given expectation
value N  is given by −βµ, where µ is known as the chemical poten-
tial. The average
 energy
 H is given by
H = Tr Hρgc
 
Tr He−β(H−µN)
=  
Tr e−β(H−µN)
   
Tr − (H − µN ) e−β(H−µN) µ β Tr N e−βH+βµN
=−   + ,
Tr e−β(H−µN) β Tr (e−βH+βµN )
thus
   
∂ log Zgc µ ∂ log Zgc
H = − + . (8.451)
∂β µ β ∂µ β

Similarly, the average number of particles N  is given by


 
  Tr N e−βH+βµN
N  = Tr Nρgc = . (8.452)
Tr (e−βH+βµN )
In terms of the fugacity λ , which is defined by
λ = eβµ , (8.453)
N  can be expressed as
∂ log Zgc
N  = λ . (8.454)
∂λ
35. The Hamiltonian can be expressed as a function of the operators p and
x as
p2
H (p, x) = + V (x) . (8.455)
2m
Evaluating Zc according to Eq. (8.447) by tracing over momentum states
yields
 
Zc = Tr e−βH

= dp′ p′ | e−βH |p′ 
 
= dx′ dp′ p′ |x′  x′ | e−βH |p′  .
(8.456)
In the classical limit the parameter β, which is inversely proportional to
the temperature, is considered as small. Using Eq. (12.120) from chapter
12, which states that for general operators A and B the following holds

Eyal Buks Quantum Mechanics - Lecture Notes 285


Chapter 8. Density Operator
 
eβ(A+B) = eβA eβB + O β 2 , (8.457)

one finds that


p2  
e−βH = e−βV (x) e−β 2m + O β 2 . (8.458)

This result together with Eq. (3.52), which is given by


 ′ ′
′ ′ 1 ip x
x |p  = √ exp , (8.459)
2π 

yield in the classical limit


 
p2
Zc = dx′ dp′ p′ |x′  x′ | e−βV (x) e−β 2m |p′ 
 
′ p′2
= dx ′
dp′ e−βV (x ) e−β 2m p′ |x′  x′ |p′ 
 
1 ′ ′
= dx′ dp′ e−βH(p ,x ) .
2π
(8.460)
Note that the this result can be also obtained by taking the limit  → 0,
for which the operator x and p can be considered as commuting operators
(recall that [x, p] = i), and consequently in this limit e−βH can be
factored in the same way [see Eq. (8.458)].
36. The measurement of the observable A1 is describe by the its extension,
which is given by A1 12 , where 12 is the identity operator on subsystem
’2’. Thus with the help of Eq. (8.8) one finds that
A1  = Tr (ρA1 12 )

= n1 , n2 | ρA1 12 |n1 , n2 
n1 ,n2
! "
 
= 1 n1 | 2 n2 | ρ |n2 2 A1 |n1 1
n1 n2
= Tr1 (ρ1 A1 ) .
(8.461)
37. In terms of the matrix elements ρn1 ,n2 ,m1 ,m2 of the operator ρ, which are
given by

ρ(n1 ,n2 ),(m1 ,m2 ) = n1 , n2 | ρ |m1 , m2  , (8.462)

the matrix elements of ρ1 and ρ2 are given by



(ρ1 )n1 ,m1 = ρ(n1 ,n2 ),(m1 ,n2 ) , (8.463)
n2

Eyal Buks Quantum Mechanics - Lecture Notes 286


8.4. Solutions

and

(ρ2 )n2 ,m2 = ρ(n1 ,n2 ),(n1 ,m2 ) . (8.464)
n1

In general ρ is Hermitian, i.e.


 ∗
ρ(n1 ,n2 ),(m1 ,n2 ) = ρ(m1 ,n2 ),(n1 ,n2 ) , (8.465)

and therefore
 ∗   ∗
(ρ1 )n1 ,m1 = ρ(n1 ,n2 ),(m1 ,n2 )
n2

= ρ(m1 ,n2 ),(n1 ,n2 )
n2
= (ρ1 )m1 ,n1 ,
(8.466)
i.e. ρ1 is also Hermitian, and similarly ρ2 is also Hermitian. Thus the
eigenvalues of ρ1 and ρ2 are all real. Moreover, these eigenvalues represent
probabilities, and therefore they are expected to be all nonnegative and
smaller than unity. In what follows it is assumed that the set of vectors
{|n1 1 } ({|n2 2 }) are chosen to be eigenvectors of the operator ρ1 (ρ2 ).
Thus ρ1 and ρ2 can be expressed as

ρ1 = wn(1)1
|n1 1 1 n1 | , (8.467)
n1

and

ρ2 = wn(2)
1
|n2 2 2 n2 | , (8.468)
n2

(1) (2)
where the eigenvalues satisfy 0 ≤ wn1 ≤ 1 and 0 ≤ wn1 ≤ 1. Similarly, ρ
can be diagonalized as

ρ= wk |k k| , (8.469)
k

where 0 ≤ wk ≤ 1. In terms of these eigenvalues the entropies are given


by

σ1 = − Tr1 (ρ1 log ρ1 ) = − wn(1)
1
log wn(1)
1
, (8.470)
n1

σ2 = − Tr2 (ρ2 log ρ2 ) = − wn(2)
2
log wn(2)
2
, (8.471)
n2

and

Eyal Buks Quantum Mechanics - Lecture Notes 287


Chapter 8. Density Operator

σ = − Tr (ρ log ρ) = − wk log wk . (8.472)
k

As can be seen from Eqs. (8.462), (8.463) and (8.467), the following holds
wn(1)
1
= (ρ1 )n1 ,n1

= ρ(n1 ,n2 ),(n1 ,n2 )
n2

= n1 , n2 | ρ |n1 , n2 
n2

= n1 , n2 |k wk k |n1 , n2  ,
n2 k
(8.473)
thus

wn(1)
1
= wn1 ,n2 , (8.474)
n2

and similarly

wn(2)
2
= wn1 ,n2 , (8.475)
n1

where

wn1 ,n2 = n1 , n2 |k wk k |n1 , n2  . (8.476)
k

Note that
! "
  
wn1 ,n2 = wk k| |n1 , n2  n1 , n2 | |k
n1 ,n2 k n1 ,n2

= wk k |k ,
k
(8.477)
thus the normalization condition k |k = 1 together with the require-
ment that

Tr ρ = wk = 1 , (8.478)
k

imply that

wn1 ,n2 = 1 , (8.479)
n1 ,n2

Eyal Buks Quantum Mechanics - Lecture Notes 288


8.4. Solutions

i.e.

Tr1 ρ1 = wn(1)
1
=1, (8.480)
n1

and

Tr2 ρ2 = wn(2)
2
=1. (8.481)
n2
 
(1) (2)
Consider the quantity y wn1 ,n2 /wn1 wn2 , where the function y (x) is
given by

y (x) = x log x − x + 1 . (8.482)

The following holds


dy
= log x , (8.483)
dx
and
d2 y 1
= , (8.484)
dx2 x
thus the function y (x) has a single stationary point at x = 1, which is a
minima point. Moreover y (1) = 0, thus one concludes that

y (x) ≥ 0 (8.485)
(1) (2)
for x ≥ 0. For x = wn1 ,n2 /wn1 wn2 the inequality (8.485) implies that
wn1 ,n2 wn1 ,n2 wn1 ,n2
(1) (2)
log (1) (2)
− (1) (2)
+1≥0. (8.486)
wn1 wn2 wn1 wn2 wn1 wn2
(1) (2)
Multiplying by wn1 wn2 and summing over n1 and n2 yields
 wn ,n2   
wn1 ,n2 log (1)1 (2) − wn1 ,n2 + wn(1)
1
wn(2)
2
≥ 0 , (8.487)
n1 ,n2 wn1 wn2 n1 ,n2 n1 n2

thus with the help of Eqs. (8.479), (8.480) and (8.481) one finds that
 wn ,n2
wn1 ,n2 log (1)1 (2) ≥0, (8.488)
n1 ,n2 wn1 wn2

and with the help of Eqs. (8.470) and (8.471) that



σ1 + σ2 ≥ − wn1 ,n2 log wn1 ,n2 . (8.489)
n1 ,n2

Eyal Buks Quantum Mechanics - Lecture Notes 289


Chapter 8. Density Operator

Using Eq. (8.476) one obtains



− wn1 ,n2 log wn1 ,n2
n1 ,n2
  2 1
= |n1 , n2 |k| wk log
n1 ,n2
wn1 ,n2
k
! "
  2 wk
= |n1 , n2 |k| wk log
n1 ,n2
wn1 ,n2
k
  2
− |n1 , n2 |k| wk log wk
n1 ,n2 k
! "
  2 wk
= |n1 , n2 |k| wk log
n1 ,n2
wn1 ,n2
k
 
− wk log wk |n1 , n2 |k|2 ,
k n1 ,n2
  
=1
(8.490)
thus

− wn1 ,n2 log wn1 ,n2
n1 ,n2
! "
  2 wk
= |n1 , n2 |k| wk log
n1 ,n2
wn1 ,n2
k
+σ .
(8.491)
Furthermore, according to inequality (8.485) the following holds
 wk
|n1 , n2 |k|2 wk log
wn1 ,n2
k
 wk wk
= |n1 , n2 |k|2 wn1 ,n2 log
wn1 ,n2 wn1 ,n2
k
  
wk
≥ |n1 , n2 |k|2 wn1 ,n2 −1 ,
wn1 ,n2
k
 
|n1 , n2 |k|2 wk − wn1 ,n2 |n1 , n2 |k|2
k k
=0.
(8.492)
These results together with inequality (8.489) yield
σ1 + σ2 ≥ σ . (8.493)

Eyal Buks Quantum Mechanics - Lecture Notes 290


8.4. Solutions

38. With the help of Eq. (8.110), which is given by

e−Hβ
ρ= = p+ |+ +| + p− |− −| , (8.494)
Tr (e−Hβ )
where the probabilities p+ and p− are given by
 ωβ
e∓ 2
p± =  ωβ  ωβ , (8.495)
e− 2 +e 2

ω = |e| B/me c is the Larmor frequency [see Eq. (4.22)] and where β =
1/kB T , one finds that
σ = −p+ log p+ − p− log p−
1 − tanh ωβ
2 1 − tanh ωβ
2 1 + tanh ωβ
2 1 + tanh ωβ
2
=− log − log .
2 2 2 2
(8.496)
39. With the help of Eqs. (6.77) and (8.414) one finds that

H| Sz |H = √ , (8.497)
2 2
thus
1
pz+ − (1 − pz+ ) = √ , (8.498)
2
and therefore
π
pz+ = cos2 . (8.499)
8
40. The density operator is given by [see Eq. (8.162)]

 
ρ = 1 − e−βω e−nβω |n n| , (8.500)
n=0

where β = 1/kB T , thus


σ = − Tr (ρ log ρ)
∞
    
=− 1 − e−βω e−nβω log 1 − e−βω e−nβω
n=0

   
= − 1 − e−βω log 1 − e−βω e−nβω
n=0

 
+βω 1 − e−βω ne−nβω .
n=0
(8.501)

Eyal Buks Quantum Mechanics - Lecture Notes 291


Chapter 8. Density Operator

By using the relations



 1
e−nβω = , (8.502)
n=0
1 − e−βω

and [see Eq. (8.152)]



 ∞
−nβω 1 ∂  −nβω e−βω
ne =− e = 2 , (8.503)
n=0
ω ∂β n=0 (1 − e−βω )

one finds that


  βω
σ = − log 1 − e−βω + . (8.504)
eβω − 1

Eyal Buks Quantum Mechanics - Lecture Notes 292


9. Time Independent Perturbation Theory

Consider a Hamiltonian H0 having a set of eigenenergies {Ek }. Let gk be


the degree of degeneracy of eigenenergy Ek , namely gk is the dimension of
the corresponding eigensubspace , which is denoted by Fk . The set {|k, i}
of eigenvectors of H0 is assumed to form an orthonormal basis for the vector
space, namely

H0 |k, i = Ek |k, i , (9.1)

and

k′ , i′ |k, i = δ kk′ δ ii′ . (9.2)

For a given k the degeneracy index i can take the values 1, 2, · · · , gk . The
set of vectors {|k, 1 , |k, 2 , · · · , |k, gk } forms an orthonormal basis for the
eigensubspace Fk . The closure relation can be written as
gk
 
1= |k, i k, i| = Pk , (9.3)
k i=1 k

where
gk

Pk = |k, i k, i| (9.4)
i=1

is a projector onto eigen subspace Fk . The orthogonality condition (9.2)


implies that

Pk Pk′ = Pk δ kk′ . (9.5)

A perturbation V = λṼ is being added to the Hamiltonian

H = H0 + λṼ , (9.6)

where λ ∈ R. We wish to find the eigenvalues and the eigenvectors of the


Hamiltonian H

H |α = E |α . (9.7)


Chapter 9. Time Independent Perturbation Theory

In many cases finding an analytical solution to the above equation is either


very hard or impossible. In such cases one possibility is to employ numerical
methods. However, another possibility arrises provided that the eigenvalues
and eigenvectors of H0 are known and provided that the perturbation λṼ can
be considered as small, namely, provided the eigenvalues and eigenvectors of
H do not significantly differ from those of H0 . In such a case an approximate
solution can be obtained by the time independent perturbation theory.

9.1 The Level En


Consider the level En of the unperturbed Hamiltonian H0 . Let Pn be the
projector onto the eigensubspace Fn , and let

Qn = 1 − Pn = Pk . (9.8)
k
=n

Equation (9.7) reads

λṼ |α = (E − H0 ) |α . (9.9)

It is useful to introduce the operator R, which is defined as


 Pk
R= . (9.10)
k
=n
E − Ek

Claim. The eigenvector |α of the Hamiltonian H is given by


 −1
|α = 1 − λRṼ Pn |α . (9.11)

Proof. Using Eq. (9.5) it is easy to show that

Pn R = RPn = 0 . (9.12)

Moreover, the following holds


  Pk Pk ′  Pk
Qn R = = =R, (9.13)

E − Ek′ E − Ek
k
=n k
=n k
=n

and similarly

RQn = R . (9.14)

Furthermore, by expressing H0 as
gk
 
H0 = Ek |k, i k, i| = En Pn + Ek P k , (9.15)
k i=1 k
=n

Eyal Buks Quantum Mechanics - Lecture Notes 294


9.1. The Level En

one finds that


 < 
 Pk E − En Pn − k′
=n Ek′ Pk′
R (E − H0 ) =
E − Ek
k
=n
 Pk (E − Ek )
=
E − Ek
k
=n
= Qn ,
(9.16)
and similarly

(E − H0 ) R = Qn . (9.17)

The last two results suggest that the operator R can be considered as the
inverse of E − H0 in the subspace of eigenvalue zero of the projector Pn
(which is the subspace of eigenvalue unity of the projector Qn ). Multiplying
Eq. (9.9) from the left by R yields

λRṼ |α = R (E − H0 ) |α . (9.18)

With the help of Eq. (9.16) one finds that

λRṼ |α = Qn |α . (9.19)

Since Pn = 1 − Qn [see Eq. (9.8)] the last result implies that


 
Pn |α = |α − λRṼ |α = 1 − λRṼ |α , (9.20)

which leads to Eq. (9.11)


 −1
|α = 1 − λRṼ Pn |α . (9.21)

Note that Eq. (9.11) can be expanded as power series in λ


 
|α = 1 + λRṼ + λ2 RṼ RṼ + · · · Pn |α . (9.22)

9.1.1 Nondegenerate Case

In this case gn = 1 and

Pn = |n n| . (9.23)

In general the eigenvector |α is determined up to multiplication by a con-


stant. For simplicity we choose that constant to be such that

Eyal Buks Quantum Mechanics - Lecture Notes 295


Chapter 9. Time Independent Perturbation Theory

Pn |α = |n , (9.24)

namely

n |α = 1 . (9.25)

Multiplying Eq. (9.9), which is given by

λṼ |α = (E − H0 ) |α , (9.26)

from the left by n| yields

n| λṼ |α = n| (E − H0 ) |α , (9.27)

or

n| E |α = n| H0 |α + n| λṼ |α , (9.28)

thus

E = En + n| λṼ |α . (9.29)

Equation (9.22) together with Eq. (9.24) yield


 
|α = 1 + λRṼ + λ2 RṼ RṼ + · · · |n
 |k, i k, i| Ṽ |n
= |n + λ
E − Ek
k
=n
i
  |k, i k, i| Ṽ |k′ , i k′ , i| Ṽ |n
+λ2

(E − Ek ) (E − Ek′ )
k
=n k =

n
i i
+··· .
(9.30)
Substituting Eq. (9.30) into Eq. (9.29) yields
E = En + λ n| Ṽ |n
 n| Ṽ |k, i k, i| Ṽ |n
+λ2
E − Ek
k
=n
i
  n| Ṽ |k, i k, i| Ṽ |k′ , i k′ , i| Ṽ |n
+λ3

(E − Ek ) (E − Ek′ )
k
=n k =

n
i i
+··· .
(9.31)

Eyal Buks Quantum Mechanics - Lecture Notes 296


9.2. Example

Note that the right hand side of Eq. (9.31) contains terms that depend on E.
To second order in λ one finds
 |k, i| V |n|2  
E = En + n| V |n + + O λ3 . (9.32)
En − Ek
k
=n
i

Furthermore, to first order in λ Eq. (9.30) yields


 |k, i k, i| V |n  
|α = |n + + O λ2 . (9.33)
En − Ek
k
=n
i

9.1.2 Degenerate Case

The set of vectors {|n, 1 , |n, 2 , · · · , |n, gn } forms an orthonormal basis for
the eigensubspace Fn . Multiplying Eq. (9.9) from the left by Pn yields

Pn λṼ |α = Pn (E − H0 ) |α , (9.34)

thus with the help of Eq. (9.15) one has

Pn λṼ |α = (E − En ) Pn |α . (9.35)

Substituting Eq. (9.22), which is given by


 
|α = Pn + λRṼ Pn + λ2 RṼ RṼ Pn + · · · |α , (9.36)

into this and noting that Pn2 = Pn yield

Pn λṼ Pn |α + λ2 Pn Ṽ RṼ Pn |α + · · · = (E − En ) Pn |α . (9.37)

Thus, to first order in λ the energy correction E − En is found by solving

Pn V Pn |α = (E − En ) Pn |α . (9.38)

The solutions are the eigenvalues of the gn × gn matrix representation of the


operator V in the subspace Fn .

9.2 Example

Consider a point particle having mass m whose Hamiltonian is given by

H = H0 + V , (9.39)

where

Eyal Buks Quantum Mechanics - Lecture Notes 297


Chapter 9. Time Independent Perturbation Theory

p2 mω2 x2
H0 = + . (9.40)
2m 2
and where
0

V = λω x. (9.41)

The eigenvectors and eigenvalues of the Hamiltonian H0 , which describes a
one dimensional harmonic oscillator, are given by

H0 |n = En |n , (9.42)

where n = 0, 1, , 2 · · · , and where


 
1
En (λ = 0) = ω n + . (9.43)
2

Note that, as was shown in chapter 5 [see Eq. (5.161)], the eigenvectors
and eigenvalues of H can be found analytically for this particular case. For
the sake of comparison we first derive this exact solution. Writing H as
0
p2 mω 2 x2 mω
H= + + λω x
2m 2 
! 0 " 2
p2 mω 2  1
= + x+λ − ωλ2 ,
2m 2 mω 2
(9.44)
one sees that H describes a one dimensional harmonic oscillator (as H0 also
does). The exact eigenenergies are given by
1
En (λ) = En (λ = 0) − ωλ2 , (9.45)
2
and the corresponding exact wavefunctions are
= 0
′ ′ 
x |n (λ) = x + λ |n . (9.46)

Using identity (3.19), which is given by

J (∆x) |x′  = |x′ + ∆x , (9.47)

where J (∆x) is the translation operator, the exact solution (9.46) can be
rewritten as
! 0 "

x′ |n (λ) = x′ | J −λ |n , (9.48)

Eyal Buks Quantum Mechanics - Lecture Notes 298


9.2. Example

or simply as
! 0 "

|n (λ) = J −λ |n . (9.49)

Next we calculate an approximate eigenvalues and eigenvectors using per-


turbation theory. Using the identity
0
  
x= a + a† , (9.50)
2mω
one has
λω  
V = √ a + a† . (9.51)
2
Furthermore, using the identities

a |n = n |n − 1 , (9.52)

a† |n = n + 1 |n + 1 , (9.53)

one has
λω  
m| V |n = √ m| a |n + m| a† |n
2
λω √ √ 
= √ nδ m,n−1 + n + 1δ m,n+1 .
2
(9.54)
Thus En (λ) can be expanded using Eq. (9.32) as
 |k, i| V |n|2  
En (λ) = En + n| V |n + + O λ3
   En − Ek
k
=n
=0 i
2
|n − 1| V |n| |n + 1| V |n|2  
= En + + + O λ3
En − En−1 En − En+1
  2
1 nλ (n + 1) λ2  
= ω n + + ω − ω + O λ3
2 2 2
  2
1 λ  
= ω n + − ω + O λ3 ,
2 2
(9.55)
in agreement (to second order) with the exact result (9.45), and |n (λ) can
be expanded using Eq. (9.30) as

Eyal Buks Quantum Mechanics - Lecture Notes 299


Chapter 9. Time Independent Perturbation Theory

 |k, i k, i| V |n  


|n (λ) = |n + + O λ2
En − Ek
k
=n
i
|n − 1 n − 1| V |n |n + 1 n + 1| V |n  
= |n + + + O λ2
En − En−1 En − En+1
√ √
|n − 1 λω

2
n |n + 1 λω

2
n+1  
= |n + − + O λ2
ω ω
λ λ †  2
= |n + √ a |n − √ a |n + O λ .
2 2
(9.56)
Note that with the help of the following identify
0
mω  
p=i −a + a† , (9.57)
2
the last result can be written as
! 0 "
 ip  
|n (λ) = 1 + λ |n + O λ2 . (9.58)
mω 

Alternatively, in terms of the translation operator J (∆x), which is given by


 
ip∆x
J (∆x) = exp − , (9.59)

one has
! 0 "
  
|n (λ) = J −λ |n + O λ2 , (9.60)

in agreement (to second order) with the exact result (9.49).

9.3 Problems
1. The volume effect: The energy spectrum of the hydrogen atom was
calculated in chapter 8 by considering the proton to be a point particle.
Consider a model in which the proton is instead assumed to be a sphere
of radius ρ0 where ρ0 << a0 (a0 is Bohr’s radius), and the charge of
the proton +e is assumed to be uniformly distributed in that sphere.
Show that the energy shift due to such perturbation to lowest order in
perturbation theory is given by
e2 2
∆En,l = ρ |Rn,l (0)|2 , (9.61)
10 0
where Rn,l (r) is the radial wave function.

Eyal Buks Quantum Mechanics - Lecture Notes 300


9.3. Problems

2. Consider a particle having mass m in a 3D central potential given by


mω 2 r2
V (r) = + gr4 . (9.62)
2

where r = x2 + y2 + z 2 is the radial coordinate, and where ω and g
are both positive. Calculate to lowest nonvanishing order in g the energy
of the ground state.
3. Consider an hydrogen atom. A perturbation given by

V = Ar , (9.63)

where r = x2 + y 2 + z 2 is the radial coordinate and A is a constant is
added.
a) Calculate to first order in A the energy of the ground state.
b) Calculate to first order in A the energy of the first excited state.
4. A weak uniform electric field E = Eẑ, where E is a constant, is applied
to a hydrogen atom. Calculate to 1st order in perturbation theory the
correction to the energy of the
a) level n = 1 (n is the principle quantum number).
b) level n = 2.
5. A particle having mass m and charge q is confined in a 3D infinite po-
tential well of width l, which is given by
+
0 if |x| ≤ l/2 and |y| ≤ l/2 and |z| ≤ l/2
V (x, y, z) = . (9.64)
+∞ elsewhere

A weak uniform electric field E = Eẑ, where E is a constant, is applied.


Calculate the eigenenergies to first order in E.
6. Consider two particles, both having the same mass m, moving in a one-
dimensional potential with coordinates x1 and x2 respectively. The po-
tential energy is given by

1 1
V (x1 , x2 ) = mω 2 (x1 − a)2 + mω 2 (x2 + a)2 + λmω2 (x1 − x2 )2 ,
2 2
(9.65)

where λ is real. Find the energy of the ground state to lowest non-
vanishing order in λ.
7. A particle having mass m is confined in a potential well of width l, which
is given by
+
0 for 0 ≤ x ≤ l
V (x) = . (9.66)
+∞ elsewhere

Find to lowest order in perturbation theory the correction to the ground


state energy due to a perturbation given by

Eyal Buks Quantum Mechanics - Lecture Notes 301


Chapter 9. Time Independent Perturbation Theory
 
l
W (x) = w0 δ x − , (9.67)
2

where w0 is a real constant.


8. Consider a particle having mass m in a two dimensional potential well of
width a that is given by
+
0 if 0 ≤ x ≤ a and 0 ≤ y ≤ a
V (x, y) = . (9.68)
+∞ elsewhere

A perturbation given by
+
w0 if 0 ≤ x ≤ a2 and 0 ≤ y ≤ a
2
W (x, y) = , (9.69)
0 elsewhere

is added.
a) Calculate to lowest non-vanishing order in w0 the energy of the
ground state.
b) The same for the first excite state.
9. Consider a particle having mass m moving in a potential energy given by

mω 2  2 
V (x, y) = x + y 2 + βmω 2 xy , (9.70)
2
where the angular frequency ω is a constant and where the dimensionless
real constant β is assumed to be small.
a) Calculate to first order in β the energy of the ground state.
b) Calculate to first order in β the energy of the first excited state.
10. Consider a harmonic oscillator having angular resonance frequency ω. A
perturbation given by
 
V = β a† a† + aa (9.71)

is added, where a is the annihilation operator and β is a real constant.


Calculate the energies of the system to second order in β.
11. The Hamiltonian of a spin S = 1 is given by
 
H = αSz2 + β Sx2 − Sy2 , (9.72)

where α and β are both constants.


a) Write the matrix representation of H in the basis {|S = 1, m = −1 , |S = 1, m = 0 |S = 1, m = 1}.
b) Calculate (exactly) the eigenenergies and the corresponding eigen-
vectors.
c) For the case β << α use perturbation theory to calculate to lowest
order in α and β the eigenenergies of the system.

Eyal Buks Quantum Mechanics - Lecture Notes 302


9.3. Problems

12. Jaynes-Cummings model - Consider a system composed of a harmonic


oscillator having angular resonance frequency ω r > 0 and a two-level
system. The Hamiltonian of the system is assumed to be given by
H = Hr + Ha + V . (9.73)
The term Hr is the Hamiltonian for the harmonic oscillator [see Eq.
(5.16)]
 
1
Hr = ω r a† a + , (9.74)
2
where a and a† are annihilation and creation operators respectively. The
term Ha is the Hamiltonian for the two-level system
ω a
Ha = Σz , (9.75)
2
where
Σz = |+ +| − |− −| ,
the ket vectors |± represent the two levels and where ω a > 0. The
coupling term between the oscillator and the two-level system is given by
 
V = g a† Σ− + aΣ+ , (9.76)
where
Σ+ = |+ −| , (9.77)
Σ− = |− +| . (9.78)
a) Calculate to lowest non-vanishing order in g the eigenenergies of the
system for the case ω r = ωa .
b) The same for the case ωr = ω a .
c) Consider the unitary transformation
H′ = U † HU , (9.79)
where
g 
U = exp S , (9.80)

the operator S is given by
 
S = a† Σ− − aΣ+ , (9.81)
and where
∆ = ωa − ωr . (9.82)
Calculate H′ to second order in g/∆.

Eyal Buks Quantum Mechanics - Lecture Notes 303


Chapter 9. Time Independent Perturbation Theory

d) Find the exact energy eigenvectors and eigenenergies of H.


e) Find a unitary operator U that diagonalizes H.
f) Use the result of the previous exercise and calculate H′ = U HU † to
forth order in g/∆.
13. Consider a particle having mass m in a two-dimensional potential given
by
1  
V0 = mω 2 x2 + y 2 . (9.83)
2
The following perturbation is added
βω 2
V1 = L , (9.84)
 z
where Lz is the z component of the angular momentum operator.
a) Find to second orders in β the energy of the ground state.
b) Find to first order in β the energy of the first excited level.
14. A particle having mass m moves in a one dimensional potential
+
V0 sin 2πx
l 0≤x≤l .
V (x) = (9.85)
∞ else
Consider the constant V0 to be small. Calculate the system’s eigenenergies
En to first order in V0 .
15. Consider a particle having mass m confined by the one-dimensional po-
tential well, which is given by

∞ x < 0
V (x) = εx 0≤x≤L .
L
∞ x>L
Find to first order in ǫ the energy of the ground state.
16. A particle of mass m is trapped in an infinite 2 dimensional well of width
l
+
0 0 ≤ x ≤ l and 0 ≤ y ≤ l
V (x, y) = . (9.86)
∞ else
A perturbation given by
2 π2
W (x, y) = λ δ (x − lx ) δ (y − ly ) . (9.87)
m
is added, where
0 ≤ lx ≤ l , (9.88)
and
0 ≤ ly ≤ l . (9.89)
Calculate to 1st order in perturbation theory the correction to the energy
of the:

Eyal Buks Quantum Mechanics - Lecture Notes 304


9.3. Problems

a) ground state.
b) first excited state.
17. Consider a rigid rotator whose Hamiltonian is given by

L2x + L2y L2 L2x − L2y


H= + z +λ , (9.90)
2Ixy 2Iz 2Ixy

where L is the angular momentum vector operator. Use perturbation


theory to calculate the energy of the ground state to second order in λ.
18. Consider two particles having the same mass m moving along the x axis.
The Hamiltonian of the system is given by

p21 p2
H= + 2 − αδ (x1 ) − αδ (x2 ) + λδ (x1 − x2 ) , (9.91)
2m 2m
where x1 and x2 are the coordinates of the first and second particle
respectively, p1 and p2 are the corresponding canonically conjugate mo-
mentums, α and λ are both real positive constants and δ () denotes the
delta function. Calculate to first order in λ the energy of the ground state
of the system.
19. In this problem the main results of time independent perturbation theory
are derived using an alternative approach. Consider a general square
matrix

W = D + ΩV , (9.92)

where Ω ∈ R, D is diagonal

D |n0  = λn0 |n0  , (9.93a)


n0 | D = λn0 n0 | , (9.93b)

and we assume that none of the eigenvalues of D is degenerate. The set


of eigenvectors of D is assumed to be orthonormal

n0 |m0  = δ nm , (9.94)

and complete (the dimensionality is assumed to be finite)



1= |n0  n0 | . (9.95)
n

Calculate the eigenvalues of W

W |n = λ |n (9.96)

to second order in Ω.

Eyal Buks Quantum Mechanics - Lecture Notes 305


Chapter 9. Time Independent Perturbation Theory

20. Consider the Hamiltonian H, which is given by

H = H0 + λṼ , (9.97)

where λ ∈ R. The set {|k} of eigenvectors of H0 with corresponding


eigenvalues {Ek }, which satisfy

H0 |k = Ek |k , (9.98)

is assumed to form an orthonormal basis for the vector space, i.e.

k′ |k = δ kk′ . (9.99)

Consider the transformation

HR = eL He−L , (9.100)

where the operator L is assumed to be anti Hermitian, i.e. L† = −L, in


order to ensure that eL is unitary. Show that to second order in λ the
matrix elements k| HR |k′′  are given by
 
λ2  1 1
k| HR |k′′  = Ek δ k,k′ + k| Ṽ |k′′  k′′ | Ṽ |k′  − ,
2 ′′ Ek − Ek′′ Ek′′ − Ek′
k
(9.101)
provided that the following condition is satisfied

λṼ + [L, H0 ] = 0 . (9.102)

Note that to first order in λ the following holds k| HR |k′′  = k| H0 |k′′ ,
thus, in spite of the fact that the perturbation λṼ is first order in λ, the
transformed Hamiltonian HR depends on λ only to second order.
21. Calculate the expectation values of the kinetic energy nlm| T |nlm and
the potential energy nlm| V |nlm of a hydrogen atom in an energy eigen-
state |nlm.
22. Calculate the expectation values nlm| r−2 |nlm, where r is the radial
position coordinate and where |nlm is an energy eigenstate |nlm of a
hydrogen atom.

9.4 Solutions

1. The radial force acting on the electron is found using Gauss’ theorem

 e2
r 2  r > ρ0
Fr (r) = e2 r 3 . (9.103)
 r2 ρ r ≤ ρ0
0

Eyal Buks Quantum Mechanics - Lecture Notes 306


9.4. Solutions

The potential energy V (r) is found by integrating Fr (r) and by requiring


that V (r) is continuous at r = ρ0

e2
 − r  r > ρ0
V (r) = e2 r
2 . (9.104)
 2ρ ρ − 3 r ≤ ρ0
0 0

Thus, the perturbation term in the Hamiltonian is given by



 2 
e   0  r > ρ0
Vp (r) = V (r)− − = 2 . (9.105)
e2 r 2ρ0
r  2ρ0 ρ0 + r − 3 r ≤ ρ0

To first order one has

∆En,l = nlm| Vp |nlm . (9.106)

The wavefunctions for the unperturbed case are given by

ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (9.107)

Since Vp depends on r only, one finds that


∞
∆En,l = drr2 |Rnl (r)|2 Vp (r)
0
ρ0 ! 2 "
2e2 r 2ρ
= 2
drr |Rnl (r)| + 0 −3 .
2ρ0 ρ0 r
0
(9.108)
2
In the limit where ρ0 << a0 the term |Rnl (r)| can approximately be
replaced by |Rnl (0)|2 , thus
ρ0 ! 2 "
2
2 2 e r 2ρ0
∆En,l = |Rnl (0)| drr + −3
2ρ0 ρ0 r
0
e2 ρ20
= |Rnl (0)|2 .
10
(9.109)
2. For the unperturbed case, i.e. when g = 0, the energy eigenvectors are
denoted by |nx , ny , nz , where the quantum numbers nx , ny and nz are
non-negative integers, and the corresponding eigenenergies are given by
 
3
Enx ,ny ,nz = ω + nx + ny + nz . (9.110)
2

Eyal Buks Quantum Mechanics - Lecture Notes 307


Chapter 9. Time Independent Perturbation Theory

With the help of Eqs. (5.11), (5.13), (5.28), (5.29) and (9.31) together
with the relation
 2
r4 = x2 + y 2 + z 2
= x4 + y 4 + z 4 + 2x2 y2 + 2y2 z 2 + 2z 2 x2 ,
(9.111)
one finds that the energy of the ground state Egs is given by
3ω  
Egs = + g 0, 0, 0| r4 |0, 0, 0 + O g2
2
3ω  2  
= + 3g 0, 0, 0| x4 |0, 0, 0 + 6g 0, 0, 0| x2 |0, 0, 0 + O g 2
2
 2
3ω   
= + 15g + O g2 .
2 2mω
(9.112)
3. The wavefunctions for the unperturbed case are given by
ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (9.113)
where for the states relevant to this problem
 3/2
1
R10 (r) = 2 e−r/a0 , (9.114a)
a0
 3/2
1 r
R20 (r) = (2 − r/a0 ) e− 2a0 , (9.114b)
2a0
 3/2
1 r r
R21 (r) = √ e− 2a0 , (9.114c)
2a0 3a0
0
1
Y00 (θ, φ) = , (9.114d)

0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (9.114e)
2 2π
0
0 1 3
Y1 (θ, φ) = cos θ , (9.114f)
2 π
0
1 3
Y11 (θ, φ) = − sin θeiφ , (9.114g)
2 2π
and the corresponding eigenenergies are given by
EI
En(0) = − , (9.115)
n2
where
me e4
EI = . (9.116)
22

Eyal Buks Quantum Mechanics - Lecture Notes 308


9.4. Solutions

The perturbation term V in the Hamiltonian is given by V = Ar. The


matrix elements of V are expressed as
∞ 1 2π  ′ ∗
n′ l′ m′ | V |nlm = A dr r3 Rn′ l′ Rnl d (cos θ) dφ Ylm
′ Ylm
0 −1 0
∞
= Aδ l,l′ δ m,m′ dr r3 Rn′ l′ Rnl .
0
(9.117)
a) Thus, to first order
(0)
E1 = E1 + 100| V |100 + O(A2 ) , (9.118)

where
∞
3Aa0
100| V |100 = A dr r3 R10
2
(r) = . (9.119)
2
0

b) The first excited state is degenerate, however, as can be seen from


Eq. (9.117) all off-diagonal elements are zero. The diagonal elements
are given by
∞
200| V |200 = A dr r3 R220 = 6Aa0 , (9.120a)
0
∞
21m| V |21m = A dr r3 R21 = 5Aa0 . (9.120b)
0
Thus, the degeneracy is lifted
(0)
E2,l=0 = E2 + 6Aa0 + O(A2 ) , (9.121)
(0)
E2,l=1 = E2 + 5Aa0 + O(A2 ) . (9.122)
4. The wavefunctions for the unperturbed case are given by

ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (9.123)

where for the states relevant to this problem

Eyal Buks Quantum Mechanics - Lecture Notes 309


Chapter 9. Time Independent Perturbation Theory
 3/2
1
R10 (r) = 2 e−r/a0 , (9.124)
a0
 3/2
1 r
R20 (r) = (2 − r/a0 ) e− 2a0 , (9.125)
2a0
3/2
1 r − r
R21 (r) = √ e 2a0 , (9.126)
2a0 3a0
0
1
Y00 (θ, φ) = , (9.127)

0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (9.128)
2 2π
0
0 1 3
Y1 (θ, φ) = cos θ , (9.129)
2 π
0
1 3
Y11 (θ, φ) = − sin θeiφ , (9.130)
2 2π
and the corresponding eigenenergies are given by
EI
En(0) = − , (9.131)
n2
where
me e4
EI = . (9.132)
22
The perturbation term V in the Hamiltonian is given by
V = eEz = eEr cos θ . (9.133)
The matrix elements of V are expressed as

∞ 1 2π  ′ ∗
′ ′ ′ 3
n l m | V |nlm = eE dr r Rn′ l′ Rnl d (cos θ) dφ cos θ Ylm
′ Ylm .
0 −1 0
(9.134)
a) Disregarding spin this level is non degenerate. To 1st order
(0) (0)
E1 = E1 + 1, 0, 0| V |1, 0, 0 = E1 ,
since
1
d (cos θ) cos θ = 0 ,
−1

thus the energy of the ground state is unchanged to 1st order.

Eyal Buks Quantum Mechanics - Lecture Notes 310


9.4. Solutions

b) The level n = 2 has degeneracy 4. The matrix of the perturbationV


in thedegenerate subspace is given by 
2, 0, 0| V |2, 0, 0 2, 0, 0| V |2, 1, −1 2, 0, 0| V |2, 1, 0 2, 0, 0| V |2, 1, 1
 2, 1, −1| V |2, 0, 0 2, 1, −1| V |2, 1, −1 2, 1, −1| V |2, 1, 0 2, 1, −1| V |2, 1, 1 
M =  2, 1, 0| V |2, 0, 0 2, 1, 0| V |2, 1, −1 2, 1, 0| V |2, 1, 0 2, 1, 0| V |2, 1, 1  .

2, 1, 1| V |2, 0, 0 2, 1, 1| V |2, 1, −1 2, 1, 1| V |2, 1, 0 2, 1, 1| V |2, 1, 1


(9.135)
Using
1
d (cos θ) cos θ = 0 , (9.136)
−1
1
d (cos θ) cos θ sin θ = 0 , (9.137)
−1
1
d (cos θ) cos θ sin2 θ = 0 , (9.138)
−1
1
d (cos θ) cos3 θ = 0 , (9.139)
−1
2π
dφ e±iφ = 0 , (9.140)
0
one finds
 
0 0γ 0
 0 00 0
M =
 γ∗
 , (9.141)
00 0
0 00 0

where
γ = 2, 0, 0| V |2, 1, 0
∞ 1 2π
 ∗
= eE dr r R2,0 R2,1 d (cos θ) dφ cos θ Y00 Y10
3

0 −1 0
∞    4
eE r r r
= dr 2 − e− a0
8 a0 a0
0
1 2π
1 2
× d (cos θ) cos θ dφ .

−1 0
(9.142)

Eyal Buks Quantum Mechanics - Lecture Notes 311


Chapter 9. Time Independent Perturbation Theory

Using
1
2
d (cos θ) cos2 θ = , (9.143)
3
−1

and  ∞
x4 e−x dx = 24 (9.144)
0
 ∞
x5 e−x dx = 120 (9.145)
0
one finds
γ = 2, 0, 0| V |2, 1, 0
∞    4
eE r r r
= dr 2 − e− a0
24 a0 a0
0
∞
a0 eE
= dx (2 − x) x4 e−x
24
0
= −3a0 eE .
(9.146)
The eigenvalues of the matrix M are 0, 0, 3a0 eE and −3a0 eE. Thus
to 1st order the degeneracy is partially lifted with subspace of dimen-
(0)
sion 2 having energy E2 , and another 2 nondegenerate subspaces
(0)
having energies E2 ± 3a0 eE.
5. For E = 0 the normalized wavefunctions ψ(0) ′ ′ ′
nx ,ny ,nz (x , y , z ) are given by

ψ(0) ′ ′ ′
nx ,ny ,nz (x , y , z )
= x′ , y ′ , z ′ |nx , ny , , nz 
 3/2      
2 nx π x′ + 2l ny π y ′ + 2l nz π z ′ + 2l
= sin sin sin ,
l l l l
(9.147)
and the corresponding eigenenergies are
 
(0)
2 π2 n2x + n2y + n2z
Enx ,ny ,nz = , (9.148)
2ml2
where nx , ny , nz ∈ {1, 2, · · · }. The matrix elements of the perturbation
V = qEz are given by
) ′ ′ ′

′′ ′′ ′′ * 2qE
nx , ny , nz
V
nx , ny , nz = δ n′x ,n′′x δ n′y ,n′′y In′z ,n′′z , (9.149)
l
where

Eyal Buks Quantum Mechanics - Lecture Notes 312


9.4. Solutions
    
l/2
n′′z π z ′ + 2l
′ n′z π z ′ + 2l ′
In′z ,n′′ = dz sin sin z . (9.150)
z
−l/2 l l

Note that In′z ,n′′z = 0 if n′z = n′′z (since for that case the integrand is an
odd function of z ′ ), and thus
) ′ ′ ′

′′ ′′ ′′ *  
nx , ny , nz
V
nx , ny , nz ∝ δ n′x ,n′′x δ n′y ,n′′y 1 − δ n′z ,n′′z . (9.151)

With the )help of the


above

result* it is easy to see that all the matrix
elements n′x , n′y , n′z
V
n′′x , n′′y , n′′z that are needed for first order per-
turbation theory, for both non-degenerate and degenerate energy levels,
vanish, and consequently, to first order in E the energy eigenstates remain
unchanged.
6. To lowest order in perturbation theory the ground state energy is given
by

∞ ∞
 
Egs = ω+λmω 2
dx1 dx2 ϕ20 (x1 − a) ϕ20 (x2 + a) (x1 − x2 )2 +O λ2 ,
−∞ −∞
(9.152)

where ϕ0 (x) is the ground state wavefunction of a particle having mass


m confined by a potential (1/2) mω2 x2 centered at x = 0. Employing the
transformation
x′1 = x1 − a , (9.153)
x′2 = x2 + a , (9.154)
and Eq. (5.141) one finds that
Egs = ω
∞
2 2
+λmω dx′1 ϕ20 (x′1 ) (x′1 + a)
−∞
∞
2
+λmω2 dx′2 ϕ20 (x′2 ) (x′2 + a)
−∞
∞ ∞
2
−2λmω dx′1 ϕ20 (x′1 ) (x′1 + a) dx′2 ϕ20 (x′2 ) (x′2 − a)
−∞ −∞
 
+O λ2
 
  
= ω + 2λmω2 + a2 + 2λmω 2 a2 + O λ2
2mω
   
= ω + λ ω + 4mω 2 a2 + O λ2 .
(9.155)

Eyal Buks Quantum Mechanics - Lecture Notes 313


Chapter 9. Time Independent Perturbation Theory

Note that this problem can be also solved exactly by employing the co-
ordinate transformation
x1 + x2
x+ = √ , (9.156)
2
x1 − x2
x− = √ . (9.157)
2
The inverse transformation is given by
x+ + x−
x1 = √ , (9.158)
2
x+ − x−
x2 = √ . (9.159)
2
The following holds
x21 + x22 = x2+ + x2− , (9.160)
and
ẋ21 + ẋ22 = ẋ2+ + ẋ2− . (9.161)
Thus, the Lagrangian of the system can be written as
 
m ẋ21 + ẋ22
L= − V (x1 , x2 )
 22   
m ẋ+ + ẋ2− 1 √
= − mω 2 x2+ + x2− − 2a 2x− + 2a2 + 4λx2−
2 2
= L+ + L− ,
(9.162)
where
mẋ2+ 1
L+ = − mω2 x2+ , (9.163)
2 2
and
 ! 
√ "2
mẋ2− 1 a 2 8λa 2
L− = − mω 2 (1 + 4λ) x− − +  . (9.164)
2 2 1 + 4λ 1 + 4λ

Thus, the system is composed of two decoupled harmonic oscillators, and


therefore, the exact eigenenergies are given by
   
1 √ 1 4λmω 2 a2
En+ ,n− = ω n+ + +ω 1 + 4λ n− + + , (9.165)
2 2 1 + 4λ
where n+ , n− = 0, 1, 2, · · · . To first order in λ one thus has
   
1 1 , -  
En+ ,n− = ω n+ + +ω n− + +λ ω (2n− + 1) + 4mω 2 a2 +O λ2 .
2 2
(9.166)

Eyal Buks Quantum Mechanics - Lecture Notes 314


9.4. Solutions

7. For w0 = 0 the normalized wavefunctions ψ(0)


n (x) are given by
0
2 nπx′
ψ(0)
n (x) = x′
|n = sin , (9.167)
l l
and the corresponding eigenenergies are
2 π 2 n2
En(0) = . (9.168)
2ml2
The matrix elements of the perturbation are given by
l  
2w0 nπx mπx l
n|W |m = sin sin δ x− dx
l l l 2
0
2w0 nπ mπ
= sin sin .
l 2 2
(9.169)
For the ground state
2w0
1|V |1 = , (9.170)
l
thus
2 π2 2w0  
E1 = 2
+ + O w02 . (9.171)
2ml l
8. For w0 = 0 the normalized wavefunctions ψ(0)
nx ,ny (x, y) are given by

2 nx πx′ ny πy ′
ψ(0) ′ ′ ′ ′
nx ,ny (x , y ) = x , y |nx , ny  = sin sin , (9.172)
a a a
and the corresponding eigenenergies are
 
(0)
2 π2 n2x + n2y
Enx ,ny = , (9.173)
2ma2
where nx = 1, 2, · · · and ny = 1, 2, · · · .
a) The ground state (nx , ny ) = (1, 1) is nondegenerate, thus to first
order in w0
2 π2
E0 = + 1, 1|W |1, 1
ma2
a/2 a/2
2 π2 4w0 πx πy
= + 2 sin2 dx sin2 dy
ma2 a a a
0 0
2 2
 π w0
+ ,
ma2 4
(9.174)

Eyal Buks Quantum Mechanics - Lecture Notes 315


Chapter 9. Time Independent Perturbation Theory

b) The first excite state is doubly degenerate. The matrix of the per-
turbation
 in the correspondingsubspace is given by
1, 2|W |1, 2 1, 2|W |2, 1
2, 1|W |1, 2 2, 1|W |2, 1
 
a/2
/ a/2
/ a/2
/ a/2
/
2 πx 2 2πy πx 2πx 2πy πy
sin a dx sin a dy sin a sin a dx sin a sin a dy 
4w0  0 0 0 0 
= 2  a/2 a/2 a/2 a/2 
a  / 2πx 1πx
/ πy 2πy
/ 2 2πx
/ 2 πy

sin a sin a dx sin a sin a dy sin a dx sin a dy
0 0 0 0
 1 16 
= w0 16 4 9π2 ,
1
9π2 4
(9.175)
To first order in perturbation theory the eigenenergies are found
by adding the eigenvalues of the above matrix to the unperturbed
(0) (0)
eigenenergy E1,2 = E2,1 . Thus, to first order in w0

52 π2 w0 16w0  
E1,± = 2
+ ± 2
+ O w02 . (9.176)
2ma 4 9π
9. For the unperturbed case β = 0 one has
H0 |nx , ny  = ω (nx + ny + 1) |nx , ny  , (9.177)
where nx , ny = 0, 1, 2, · · · . Using the identities
0
  
x= ax + a†x , (9.178)
2mω
0
  
y= ay + a†y , (9.179)
2mω
the perturbation term V1 = βmω 2 xy can be expressed as
ω   
V1 = β ax + a†x ay + a†y .
2
a) For the ground state |0, 0, which is nondegenerate, one has
 |nx , ny | V1 |0, 0|2
E0,0 (β) = ω + 0, 0| V1 |0, 0 +
   E0,0 (0) − Enx ,ny
nx ,ny
=0,0
=0
2
|1, 1| V1 |0, 0|
= ω +
2ω
 2
ωβ
2
= ω −
2ω
 
β2
= ω 1 − .
8
(9.180)

Eyal Buks Quantum Mechanics - Lecture Notes 316


9.4. Solutions

b) The first excited state is doubly degenerate, thus the eigenenergies


are found by diagonalizing the matrix of V1 in the corresponding
subspace
   
1, 0| V1 |1, 0 1, 0| V1 |0, 1 ωβ 0 1
= . (9.181)
0, 1| V1 |1, 0 0, 1| V1 |0, 1 2 10

Thus the degeneracy is lifted and the energies are given by 2ω (1 ± β/4).
Note that this problem can be also solved exactly by employing the
coordinate transformation
x+y
x′ = √ , (9.182)
2
x−y
y′ = √ . (9.183)
2
The inverse transformation is given by
x′ + y ′
x= √ , (9.184)
2
x′ − y ′
y= √ . (9.185)
2
The following hold
x2 + y 2 = x′2 + y ′2 , (9.186)
2 2 ′2 ′2
ẋ + ẏ = ẋ + ẏ , (9.187)
1  ′2 
xy = x − y ′2 . (9.188)
2
Thus, the Lagrangian
  of the system can be written as
m ẋ2 + ẏ2
L= − V (x1 , x2 )
 ′22 
m ẋ + ẏ ′2 mω 2  ′2  βmω 2  ′2 
= − x + y ′2 − x − y ′2
2 2 2
= L+ + L− ,
(9.189)
where
mẋ′2 mω 2
L+ = − (1 + β) x′2 , (9.190)
2 2
and
mẏ ′2 mω 2
L− = − (1 − β) y′2 . (9.191)
2 2
Thus, the system is composed of two decoupled harmonic oscillators,
and therefore, the exact eigenenergies are given by
     
1 1
En+ ,n− = ω 1 + β nx + + 1 − β ny + , (9.192)
2 2
where nx , ny = 0, 1, 2, · · · . To second order in β one thus has

Eyal Buks Quantum Mechanics - Lecture Notes 317


Chapter 9. Time Independent Perturbation Theory
 
nx − ny nx + ny + 1 2  
En+ ,n− = ω nx + ny + 1 + β− β + O β3 .
2 8
(9.193)

10. Using Eqs. (5.28) and (5.29) one finds that


 
m| V |n = β n (n − 1)δ m,n−2 + β (n + 1) (n + 2)δ m,n+2 , (9.194)

thus
 
1  |m| V |n|2  
En (β) = ω n + + n| V |n + + O β3
2    En (0) − Em (0)
m
=n
=0
  2
1 β  
= ω n + + [n (n − 1) − (n + 1) (n + 2)] + O β 3
22ω
   2 
1 β  
= ω n + 1−2 + O β3 .
2 ω
(9.195)
11. In general the subspace of angular momentum states with J = 1 is
spanned by the basis

{|j = 1, m = −1 , |j = 1, m = 0 , |j = 1, m = 1} , (9.196)

and the following holds


j ′ , m′ | Jz |j, m = mδ j′ ,j δ m′ ,m , (9.197)
j ′ , m′ | J2 |j, m = j (j + 1) 2 δ j ′ ,j δ m′ ,m , (9.198)

j ′ , m′ | J± |j, m =  (j ∓ m) (j ± m + 1)δ j ′ ,j δ m′ ,m±1 , (9.199)
J± = Jx ± iJy . (9.200)
In matrix form
 
10 0
Jz =˙ 0 0 0  , (9.201)
0 0 −1
 
100
J2 =˙ 22  0 1 0  , (9.202)
001
 
√ 010
J+ =˙  20 0 1 , (9.203)
000
 
√ 000
J− =˙  21 0 0 . (9.204)
010

Eyal Buks Quantum Mechanics - Lecture Notes 318


9.4. Solutions

a) The Hamiltonian is given by



H = αSz2 + β Sx2 − Sy2
β3 2 2
4
= αSz2 + (S+ + S− ) + (S+ − S− )
4
2 β 2 2

= αSz + S+ + S− .
2
(9.205)
Thus, in matrix
 form    
100 001 000
H= ˙ α2  0 0 0  + β2  0 0 0  +  0 0 0 
001 000 100
 
α0β
= 2  0 0 0  .
β0α
(9.206)
b) The eigenvalues
 and
 eigenvectors
 are given
 by
α0β 1 1
2    2
 0 0 0 0 =  (α + β) 0  ,
 (9.207)
β0α 1 1
    
α0β −1 −1
2  0 0 0   0  = 2 (α − β)  0  , (9.208)
β0α 1 1
    
α0β 0 0
2  0 0 0   1  = 2 × 0  1  . (9.209)
β0α 0 0
c) The Hamiltonian is written
 as H = H0 + V where in matrix form
100
H0 =˙ 2 α  0 0 0  , (9.210)
001
 
001
V =˙ 2 β  0 0 0  . (9.211)
100
0
For the nondegenerate eigenenergy Em=0 = 0 on has to second order
in perturbation expansion
 |1, m′ | V |1, 0|2
0
Em=0 = Em=0 +1, 0| V |1, 0+ 0 0 = 0 . (9.212)
Em=0 − Em ′
m′ =±1

0
For the degenerate eigenenergy Em=±1 = 2 α the perturbation in
the subspace spanned by {|1, −1 , |1, 1} is given in matrix form by
 
01
˙ 2β
Vm=±1 = , (9.213)
10

Eyal Buks Quantum Mechanics - Lecture Notes 319


Chapter 9. Time Independent Perturbation Theory

thus to first order in perturbation expansion

Em=±1 = 2 (α ± β) . (9.214)

12. For the unperturbed case V = 0, the eigenvectors and eigenenergies are
related by
0
(Hr + Ha ) |n, σ = En,σ |n, σ , (9.215)

where n = 0, 1, 2, · · · is the quantum number of the harmonic oscillator,


and σ ∈ {−1, +1} is the quantum number associated with the two-level
particle, and
 
0 1 ωa
En,σ = ωr n + +σ . (9.216)
2 2

a) To second order in perturbation theorem [see Eq. (9.32)]

0
 |n′ , σ′ | V |n, σ|2
En,σ = En,σ + n, σ| V |n, σ + 0 − E0
. (9.217)
En,σ n′ ,σ ′
n′ ,σ ′
=n,σ

Using √
V |n, + = ga† |n, − = g n + 1 |n + 1, − , (9.218)

V |n, − = ga |n, + = g n |n − 1, + , (9.219)
one finds for σ = 
+1 
1 ω a g 2 (n + 1)
En,+1 = ωr n + + +
2 2 ω a − ωr
   2 
g 2
1 ω a + g∆
=  ωr + n+ + ,
∆ 2 2
(9.220)
and for σ = −1  
1 ω a g2 n
En,−1 = ωr n + − −
2 2 ω a − ωr
   
2 g2
g 1 −ω a + ∆
=  ωr − n+ + ,
∆ 2 2
(9.221)
where

∆ = ωa − ωr . (9.222)

For the general case this can be written as


    
g2 1 1 g2
En,σ =  ω r + σ n+ + σω a + . (9.223)
∆ 2 2 ∆

Eyal Buks Quantum Mechanics - Lecture Notes 320


9.4. Solutions

Thus, according to the above result (9.223), the energies of the states
(n, +1) and (n + 1, −), which are degenerate for the case where ω r =
ωa and where g = 0, are given to second order in g by
  
g2 ∆
En,+1 =  ωr + (n + 1) + , (9.224)
∆ 2
and
  
g2 ∆
En+1,−1 =  ωr − (n + 1) − . (9.225)
∆ 2

b) In the degenerate case ω r = ω a ≡ ω the eigenenergies for the case


V = 0 are given by
 
0 1 σ
En,σ = ω n + + , (9.226)
2 2

thus the pairs of states |n, + and |n + 1, − are degenerate. In the


subset of such a pair the perturbation is given by
   √ 
n, +| V |n, + n, +| V |n + 1, − √0 g n + 1
= ,
n + 1, −| V |n, + n + 1, −| V |n + 1, − g n + 1 0
(9.227)

thus to first order in g the eigenenergies are given by


, √ -
E =  ω (n + 1) ± g n + 1 . (9.228)

c) Using Eq. (2.173) one finds that (note that S † = −S)


1
H′ = H + [L, H] + [L, [L, H]] + · · · , (9.229)
2!
where
g  
L= aΣ+ − a† Σ− . (9.230)

Using the commutation relations

[Σz , Σ+ ] = 2Σ+ , (9.231)


[Σz , Σ− ] = −2Σ− , (9.232)
[Σ+ , Σ− ] = Σz , (9.233)
, † -
a, a a = a , (9.234)
, † † -
a , a a = −a† , (9.235)

one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 321


Chapter 9. Time Independent Perturbation Theory
 
  g2 1 + Σz
[L, H] = −g a† Σ− + aΣ+ + 2 + a† aΣz , (9.236)
∆ 2

and thus    
′ g2 †  g2
H =  ωr + Σz a a + ωa + Σz
∆ 2 ∆
 2

 ωr + g∆  3 
g
+ +O .
2 ∆
(9.237)
Note that to second order in g/∆ the states |n, σ are eigenvalues of
H′ , and the following holds

H′ |n, σ = Ẽn,σ |n, σ , (9.238)

where
    
g2 1 1 g2
En,σ =  ωr + σ n+ + σω a + . (9.239)
∆ 2 2 ∆

The above result agrees with Eq. (9.223).


d) Consider the pair of states |n, + and |n + 1, −. The following folds
[see Eq. (9.73)]
∆ √
H |n, + = ω r (n + 1) |n, + + |n, + + g n + 1 |n + 1, − ,
2
(9.240)
and
∆ √
H |n + 1, − = ω r (n + 1) |n + 1, − − |n + 1, − + g n + 1 |n, + ,
2
(9.241)
where

∆ = ωa − ωr , (9.242)

or in a matrix
 form 
|n, +
H
|n + 1, −
     
10 ω n cos θ sin θ |n, +
=  ω r (n + 1) + ,
01 2 sin θ − cos θ |n + 1, −
(9.243)
where 
ω n = ∆2 + 4g 2 (n + 1) , (9.244)

2g n + 1
tan θ = . (9.245)

Thus, the states |n+  and |n− , which are given by [see Eqs. (6.213)
and (6.214)]

Eyal Buks Quantum Mechanics - Lecture Notes 322


9.4. Solutions

θ θ
|n+  = cos |n, + + sin |n + 1, − , (9.246)
2 2
θ θ
|n−  = − sin |n, + + cos |n + 1, − , (9.247)
2 2
are eigenstates of H and the following holds

H |n±  = En± |n±  , (9.248)


where 3 ωn 4
En± =  ω r (n + 1) ±
 2 
0
∆2 2
=  ω r (n + 1) ± + g (n + 1) .
4
(9.249)
The ground state is the state |0, −
H |0, − = Eg |n, − , (9.250)
and the ground state energy is
∆
Eg = − . (9.251)
2
e) The desired unitary operator U is required to satisfy [see Eq. (9.250)]
U |0, − = |0, − , (9.252)
and [see Eqs. (9.246) and (9.247)]
θ θ
U |n, + = |n+  = cos |n, + + sin |n + 1, − , (9.253)
2 2
θ θ
U |n + 1, − = |n−  = − sin |n, + + cos |n + 1, − , (9.254)
2 2
where

2g n + 1
tan θ = . (9.255)

The required transformation can be constructed using the operators
S and N , which are defined by
S = a† Σ− − aΣ+ , (9.256)
and
N = a† a + |+ +| . (9.257)
The following holds
S |0, − = 0 , (9.258)

S |n, + = n + 1 |n + 1, − , (9.259)

S |n + 1, − = − n + 1 |n, + , (9.260)

Eyal Buks Quantum Mechanics - Lecture Notes 323


Chapter 9. Time Independent Perturbation Theory

and
N |n + 1, − = (n + 1) |n, − , (9.261)
N |n, + = (n + 1) |n, + . (9.262)
Thus, the operator I, which is defined by

I = N −1/2 S , (9.263)

satisfies
I |n, + = |n + 1, − , (9.264)
I |n + 1, − = − |n, + . (9.265)
Therefore, Eqs. (9.246) and (9.247) can be rewritten as
|n+  = U |n, + , (9.266)
|n−  = U |n + 1, − , (9.267)
where
θ θ
U = cos + sin I . (9.268)
2 2
Furthermore, with the help of Eq. (9.255) one finds that [note that
I 2 |n, + = − |n, + and I 2 |n + 1, − = − |n + 1, −]
 1/2

I −1 2gN
U = exp tan . (9.269)
2 ∆
To first order in g/∆ the following holds [compare with Eq. (9.80)]
 3 
g g
U = e∆S + O . (9.270)

f) With the help of Eqs. (2.173) and (9.269) one finds that
1 1
H′ = H + [L, H] + [L, [L, H]] + [L, [L, [L, H]]] + · · · , (9.271)
2! 3!
where

L = −Sf (N ) , (9.272)

and where the function f is given by


x−1/2 2gx1/2
f (x) = tan−1
2 ∆ 
g 4x  g 3  g 5 
= − +O .
∆ 3 ∆ ∆
(9.273)
The following holds
 g 2   
8x  g 4 g 6
f 2 (x) = − +O , (9.274)
∆ 3 ∆ ∆

Eyal Buks Quantum Mechanics - Lecture Notes 324


9.4. Solutions

and
 g 3   
g 5
f 3 (x) = +O . (9.275)
∆ ∆
Using the commutation relations
, - , -
a† Σ− + aΣ+ , a† a = − a† Σ− + aΣ+ , |+ +| = −a† Σ− + aΣ+ ,
(9.276)

one finds that

[H, N ] = 0 ,

and using the,commutation relations


-
a† a, a† Σ− − aΣ+ = a† Σ− + aΣ+ , (9.277)

Σz †
, a Σ− − aΣ+ = −a† Σ− − aΣ+ , (9.278)
2
 
, † - 1
a Σ− + aΣ+ , a† Σ− − aΣ+ = 1 + 2 a† a + Σz , (9.279)
2
one finds that
  
 †  † 1
[H, S] = −∆ a Σ− + aΣ+ +g 1 + 2 a a + Σz . (9.280)
2
Thus, the following holds

[L, H] = [H, S] f (N ) , (9.281)

and (note that [S, N ] = 0 and [[H, S] , N ] = 0)

[L, [L, H]] = [[H, S] , S] f 2 (N ) , (9.282)

where
  
1
[[H, S] , S] = −∆ 1 + 2 a† a + Σz
2
 
1  †   
−4g a† a + a Σ− + aΣ+ + 2g a† Σ− + aΣ+ Σz ,
2
(9.283)

and therefore (note that [[[H, S] , S] , N ] = 0)


[L, [L, [L, H]]]
, -
= [[H, S] , S] f 2 (N ) , Sf (N )
= [[[H, S] , S] , S] f 3 (N ) ,
(9.284)

Eyal Buks Quantum Mechanics - Lecture Notes 325


Chapter 9. Time Independent Perturbation Theory

where
[[[H, S] , S] , S]
 
1  †   
= 4∆ a† a + a Σ− + aΣ+ − 2∆ a† Σ− + aΣ+ Σz
2
    
1 1
−4g a† a + 1 + 2 a† a + Σz
2 2
 2
−8g a† Σ− + aΣ+
  
† 1
+2g 1 + 2 a a + Σz Σz .
2
(9.285)
By combining the above results one finds that
4g 4 4g 4 †
−1 H′ = ωr − 3
+ ξΣz − a aΣz a† a
3∆ 3∆3
1 ωr ξ
+ (ωa + ξ) Σz + +
2 2 2
 g 5 
+O .

(9.286)
where
 
g2 4g 2
ξ= 1− . (9.287)
∆ 3∆2

13. Using creation and annihilation operators one has

p2x + p2y 1  
H0 = + mω 2 x2 + y 2 = ω (Nx + Ny + 1) , (9.288)
2m 2
where Nx = a†x ax , Ny = a†y ay , and
βω 2
V = L
 z
βω
= (xpy − ypx )2

βω ,  -2
= i ax a†y − a†x ay
 3  2  2 4
= −βω a2x a†y + a†x a2y − ax a†x a†y ay − a†x ax ay a†y
3  2  2 4
= −βω a2x a†y + a†x a2y − (1 + Nx ) Ny − Nx (1 + Ny ) .
(9.289)
a) For the case β = 0 the ground state |0, 0 is nondegenerate and has
energy E0,0 = ω. Since V |0, 0 = 0 one finds to second order in β

Eyal Buks Quantum Mechanics - Lecture Notes 326


9.4. Solutions

1  |nx , ny | V |0, 0|2  


E0,0 = ω+0, 0| V |0, 0− = ω+O β 3 .
ω nx ,ny
=0,0
n x + n y

(9.290)
b) For the case β = 0 the first excited state is doubly degenerate
H0 |1, 0 = 2ω |1, 0 , (9.291)
H0 |0, 1 = 2ω |0, 1 . (9.292)
Thematrix of V in the basis {|1,  0 , |0, 1} is given by
1, 0| V |1, 0 1, 0| V |0, 1
0, 1| V |1, 0 0, 1| V |0, 1
 
1, 0| [(1 + Nx ) Ny + Nx (1 + Ny )] |1, 0 0
= βω
0 0, 1| [(1 + Nx ) Ny + Nx (1 + Ny )] |0, 1
 
10
= βω .
01
(9.293)
Thus to first order in β the first excited state remains doubly de-
generate with energy 2ω (1 + β). Note - The exact solution can be
found using the transformation
1
ad = √ (ax − iay ) , (9.294)
2
1
ag = √ (ax + iay ) . (9.295)
2
The following holds
3 4 , -
ad , a†d = ag , a†g = 1 ,

1 †  1 † 
a†d ad + a†g ag = ax + ia†y (ax − iay ) + ax − ia†y (ax + iay )
2 2
= a†x ax + a†y ay ,
(9.296)
and
1 †  1 † 
a†d ad − a†g ag = ax + ia†y (ax − iay ) − ax − ia†y (ax + iay )
2  2
= i ax a†y − a†x ay ,
(9.297)
thus
H0 = ω (Nd + Ng + 1) , (9.298)
2
V = βω (Nd − Ng ) , (9.299)
and the exact eigen vectors and eigenenergies are given by
3 4
(H0 + V ) |nd , ng  = ω nd + ng + 1 + β (nd − ng )2 |nd , ng  .
(9.300)

Eyal Buks Quantum Mechanics - Lecture Notes 327


Chapter 9. Time Independent Perturbation Theory

14. For V0 = 0 the wavefunctions ψ(0)


n (x) are given by
0
2 nπx′
ψ(0) ′
n (x) = x |n = sin , (9.301)
l l
and the corresponding eigenenergies are
2 π 2 n2
En(0) = . (9.302)
2ml2
The matrix elements of the perturbation are given by
l
2V0 nπx mπx 2πx
n|V |m = sin sin sin dx . (9.303)
l l l l
0

For the diagonal case n = m


l
2V0 nπx 2πx
n|V |n = sin2 sin dx (9.304)
l l l
0
l/2  nπy  
2V0 nπ  2πy
= sin2 + sin + π dy (9.305)
l l 2 l
−l/2

l/2  2nπy 
2V0 1 − cos l + nπ 2πy
=− sin dy (9.306)
l 2 l
−l/2

=0, (9.307)
(9.308)
since the integrand is clearly an odd function of y. Thus to first order in
V0 the energies are unchanged
2 π2 n2  
En = + O V02 . (9.309)
2ml2
15. For the case ε = 0 the exact wave functions are given by
0  nπx 
(0) 2
ψn (x) = sin , (9.310)
L L
and the corresponding eigenenergies are
2 π 2 n2
En(0) = , (9.311)
2mL2
where n is integer. To first order in ε the energy of the ground state n = 1
is given by

Eyal Buks Quantum Mechanics - Lecture Notes 328


9.4. Solutions
 L  2
(0) ε (0)  
E1 = E1 + dx ψ1 (x) x + O ε2
L 0
  πx 
(0) 2ε L  
= E1 + 2 dx sin2 x + O ε2
L 0 L
(0) ε  
= E1 + + O ε2
2
(9.312)
16. For the case λ = 0 the exact wave functions of the eigenstates are given
by
2 nx πx ny πy
ψ(0)
nx ,ny (x, y) = sin sin , (9.313)
l l l
and the corresponding eigenenergies are
 
(0)
2 π2 n2x + n2y
Enx ,ny = , (9.314)
2ml2
where nx and ny are non-zero integers.
a) The ground state is non degenerate thus to 1st order the energy is
given by
 l l 2
(0) (0)
E0 = E1,1 + ψ1,1 W dxdy
0 0
2 π2
=
ml2
 l l
2 π2 πx 2 πy
+ 2 4λ sin2 sin δ (x − lx ) δ (y − ly ) dxdy
ml 0 0 l l
 
2 π2 πlx πly
= 1 + 4λ sin2 sin2 .
ml2 l l
(9.315)
b) The first excited state is doubly degenerate. The matrix of the per-
turbation
 W in the eigen subspaceis given by
2, 1| W |2, 1 2, 1| W |1, 2
W = ˙
1, 2| W |2, 1 1, 2| W |1, 2
! "
2 π2 sin2 2πll x sin2 πlly sin 2πll x sin πll x sin πll y sin 2πll y
= 4λ
ml2 sin πllx sin 2πll x sin 2πll y sin πll y sin2 πllx sin2 2πll y
! "
2 π2 4 sin2 πllx cos2 πllx sin2 πll y 4 cos πllx sin2 πllx cos πll y sin2 πll y
= 4λ
ml2 4 cos πll x sin2 πllx cos πll y sin2 πll y 4 sin2 πllx sin2 πll y cos2 πlly
πl
! "
16λ2 π 2 sin2 πllx sin2 ly cos2 πllx cos πllx cos πlly
= .
ml2 cos πllx cos πll y cos2 πll y
(9.316)

Eyal Buks Quantum Mechanics - Lecture Notes 329


Chapter 9. Time Independent Perturbation Theory

The eigenvalues of W are

w1 = 0 , (9.317)

and
 
πly πly
16λ2 π2 sin2 πlx
l sin2 l cos2 πlx
l + cos2 l
w2 = . (9.318)
ml2
17. The unperturbed Hamiltonian (λ = 0) can be written as
L2 − L2z L2
H= + z
2Ixy 2Iz
2
 
L 1 1
= + − L2z ,
2Ixy 2Iz 2Ixy
(9.319)
thus the states |l, m (the standard eigenstates of L2 and Lz ) are eigen-
states of H and the following holds

H |l, m = El,m |l, m , (9.320)

where
  
2 l (l + 1) 1 1 2
El,m =  + − m . (9.321)
2Ixy 2Iz 2Ixy

Since the unperturbed Hamiltonian is positive-definite, it is clear that


the state |l = 0, m = 0 is the (nondegenerate) ground state of the system
since its energy vanishes E0,0 = 0. Using
L+ + L−
Lx = , (9.322)
2
L+ − L−
Ly = , (9.323)
2i
one finds that the perturbation term V can be written as

L2+ + L2−
V =λ . (9.324)
4Ixy

To second order in λ the energy of the ground state is found using Eq.
(9.32)

 |l′ , m′ | V |0, 0|2  


E0 = E0,0 +0, 0| V |0, 0+ +O λ3 . (9.325)
E0,0 − El′ ,m′
l′ ,m′
=0,0

Using the relations

Eyal Buks Quantum Mechanics - Lecture Notes 330


9.4. Solutions

L+ |l, m = l (l + 1) − m (m + 1) |l, m + 1 , (9.326)

L− |l, m = l (l + 1) − m (m − 1) |l, m − 1 , (9.327)
it is easy to see that all terms to second order in λ vanish, thus
 
E0 = 0 + O λ3 . (9.328)

18. The Hamiltonian can be written as

H = H1 + H2 + V , (9.329)

where
p21
H1 = − αδ (x1 ) , (9.330)
2m
p22
H2 = − αδ (x2 ) , (9.331)
2m
and

V = λδ (x1 − x2 ) . (9.332)

First consider H1 only. A wavefunction ψ(1) (x1 ) of an eigenstate of H1


must satisfy the following Schrödinger equation
2 
d 2m
+ (E + αδ (x 1 )) ψ(1) (x1 ) = 0 . (9.333)
dx21 2

Integrating around x1 = 0 yields the condition

dψ(1) (0+ ) dψ(1) (0− ) 2mα (1)


− + 2 ψ (0) = 0 . (9.334)
dx1 dx1 
Requiring also that the wavefunction is normalizable leads to
0  mα 
(1) mα
ψ (x1 ) = exp − |x1 | .
2 2
The corresponding eigenenergy is

(1) mα2
E0 = − .
22
The ground state of H2 can be found in a similar way. Thus, the normal-
ized wavefunction of the only bound state of H1 + H2 , which is obviously
the ground state, is given by
mα  mα   mα 
ψ0 (x1 , x2 ) = 2 exp − 2 |x1 | exp − 2 |x2 | , (9.335)
  

Eyal Buks Quantum Mechanics - Lecture Notes 331


Chapter 9. Time Independent Perturbation Theory

and the corresponding energy is given by


mα2
E0 = − . (9.336)
2
Therefore, to first order in λ the energy of the ground state of H is given
by Eq. (9.32)
mα2
Egs = −
2
∞ ∞
 
+λ dx1 dx2 ψ∗0 (x1 , x2 ) δ (x1 − x2 ) ψ0 (x1 , x2 ) + O λ2
−∞ −∞

 mα 2
2 ∞  
mα 4mα  
= − 2 +λ dx1 exp − 2 |x1 | + O λ2
 2 
−∞
mα2 λmα  
=− 2 + + O λ2 .
 22
(9.337)
19. Substituting the expansions
 
|n = |n0  + Ω |n1  + Ω 2 |n2  + O Ω 3 , (9.338)

and
 
λ = λn0 + Ωλn1 + Ω 2 λn2 + O Ω 3 , (9.339)

into Eq. (9.96) and collecting terms having the same order in Ω (up to
second order) yield
(D − λn0 ) |n0  = 0 , (9.340)
(D − λn0 ) |n1  + (V − λn1 ) |n0  = 0 , (9.341)
(D − λn0 ) |n2  + (V − λn1 ) |n1  − λn2 |n0  = 0 . (9.342)
We further require normalization

n|n = 1 , (9.343)

and choose the phase of n0 |n such that

n0 |n ∈ R . (9.344)

Expressing the normalization condition using Eq. (9.338) and collecting


terms having the same order in Ω yield
n0 |n0  = 1 , (9.345)
n0 |n1  + n1 |n0  = 0 , (9.346)
n0 |n2  + n2 |n0  + n1 |n1  = 0 . (9.347)

Eyal Buks Quantum Mechanics - Lecture Notes 332


9.4. Solutions

These results together with Eq. (9.344) yield

n0 |n1  = n1 |n0  = 0 , (9.348)


1
n0 |n2  = n2 |n0  = − n1 |n1  . (9.349)
2
Multiplying Eq. (9.341) by m0 | yields

λn1 m0 |n0  = (λm0 − λn0 ) m0 |n1  + m0 | V |n0  , (9.350)

thus for m = n

λn1 = n0 | V |n0  . (9.351)

Using this result for m = n yields

m0 | V |n0 
m0 |n1  = , (9.352)
λn0 − λm0
thus with the help of Eq. (9.95) one has
 m0 | V |n0 
|n1  = |m0  . (9.353)
m
λn0 − λm0

Multiplying Eq. (9.342) by n0 | yields

λn2 = n0 | V |n1  − λn1 n0 |n1  , (9.354)

or using Eq. (9.353)


 n0 | V |m0  m0 | V |n0 
λn2 = . (9.355)
m
λn0 − λm0

Thus, using this result together with Eq. (9.351) one finds

λ = λn0 + Ω n0 | V |n0 


 n0 | V |m0  m0 | V |n0   
+ Ω2 + O Ω3 .
m
λn0 − λm0
(9.356)

20. The condition [9.102] together with Eq. (9.98) can be used to evaluate
the matrix elements of L
k| Ṽ |k′ 
k| L |k′  = λ . (9.357)
Ek − Ek′
With the help of Eq. (2.173) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 333


Chapter 9. Time Independent Perturbation Theory
3 4 1 3 3 44
HR = H0 +λṼ +[L, H0 ]+ L, λṼ + L, L, H0 + λṼ +· · · . (9.358)
2!
Thus, for the case where condition (9.102) is satisfied the following holds
[note that according to Eq. (9.357) L = O (λ)]
3 4 1  
HR = H0 + L, λṼ + [L, [L, H0 ]] + O λ3
2!
13 4  
= H0 + L, λṼ + O λ3 .
2
(9.359)
where Eq. (9.102) has been employed in the last step. <The last result
together with Eq. (9.357) and the closure relation 1 = k′′ |k′′  k′′ | [see
Eq. (2.23)] lead to
1 3 4  
k| HR |k′′  = k| H0 |k′  + k| L, λṼ |k′  + O λ3
2  
<
λ2 k′′ k| Ṽ |k′′  k′′ | Ṽ |k′  Ek −E
1
′′
1
− E ′′ −E ′  
+ O λ3 .
k k k
= Ek δ k,k′ +
2
(9.360)
21. Consider the Hamiltonian
p2 (1 + λ) e2
H= − , (9.361)
2µ r
where µ is the reduced mass and e is the electron charge. The parameter
λ is a positive constant. The exact eigenenergies are given by Eq. (7.84)
µ (1 + λ)2 e4
En = − . (9.362)
22 n2
On the other hand, perturbation theory yields the following expansion
[see Eq. (9.32)]
µe4 e2  
En = − 2 2
− λ nlm| |nlm + O λ2 . (9.363)
2 n r
By comparing the above results for En one finds that
e2 µe4
− nlm| |nlm = −2 2 2 , (9.364)
r 2 n
thus (recall that V = −e2 /r)
µe4
nlm| V |nlm = −2 , (9.365)
22 n2
and (recall that T + V = H)
µe4
nlm| T |nlm = . (9.366)
22 n2

Eyal Buks Quantum Mechanics - Lecture Notes 334


9.4. Solutions

22. The energy eigenvalues Ekl of the radial equation of the hydrogen atom,
which is given by [see Eq. (7.61)]
 2 2 
− d e2 l (l + 1) 2
− + ukl (r) = Ekl ukl (r) , (9.367)
2µ dr2 r 2µr2

where µ is the reduced mass and e is the electron charge, are [see Eq.
(7.78)]

µe4
Ekl = − , (9.368)
22 (k + l)2
where k = 1, 2, 3, · · · . The quantum number l can formally be treated as a
real number, which is not restricted to take integer values only. Consider
the case where the integer l is replaced by l + ǫ, where 0 ≤ ǫ ≪ 1. While
the exact eigenenergies can still be evaluated by Eq. (7.78)

µe4
Ekl = − , (9.369)
22 (k + l + ǫ)2

perturbation theory yields the following expansion [see Eq. (9.32)]

µe4
Ekl = − + klm| VH |klm + · · · , (9.370)
22 (k + l)2
where the perturbation VH is given by
[(l + ǫ) (l + ǫ + 1) − l (l + 1)] 2
VH =
2µr2
2  
(2l + 1) ǫ
= 2
+ O ǫ2 ,
2µr
(9.371)
By comparing both results for Ekl one finds that

2µ2 e4
klm| r−2 |klm = , (9.372)
4 (2l + 1) (k + l)3
or in terms of the quantum number n = k + l
2
nlm| r−2 |nlm = , (9.373)
a20 (2l + 1) n3
where
2
a0 = . (9.374)
µe2

Eyal Buks Quantum Mechanics - Lecture Notes 335


10. Time-Dependent Perturbation Theory

Recall that the time evolution of a state vector |α is governed by the
Schrödinger equation (4.1)

d |α
i = H |α , (10.1)
dt
where the Hermitian operator H = H† is the Hamiltonian of the system. The
time evolution operator u (t, t0 ) [see Eq. (4.4)] relates the state vector |α (t0 )
at time t0 with its value |α (t) at time t

|α (t) = u (t, t0 ) |α (t0 ) . (10.2)

As we have seen in chapter 4, when the Hamiltonian is time independent


u (t, t0 ) is given by
 
i (t − t0 )
u (t, t0 ) = exp − H . (10.3)


In this chapter we consider the more general case where H is allowed to vary
in time. We first derive a formal expression for the time evolution operator
u (t, t0 ) applicable for general H. Then we present the perturbation theory
expansion of the time evolution operator, and discuss approximation schemes
to evaluate u (t, t0 ).

10.1 Time Evolution

Dividing the time interval (t0 , t) into N sections of equal duration allows
expressing the time evolution operator as
N
>
u (t, t0 ) = u (tn , tn−1 ) , (10.4)
n=1

where

tn = t0 + nǫ , (10.5)
Chapter 10. Time-Dependent Perturbation Theory

and where
t − t0
ǫ= . (10.6)
N
Furthermore, according to the Schrödinger equation (4.7), the following holds
iǫ  
u (tn−1 + ǫ, tn−1 ) = 1 − H (tn ) + O ǫ2 . (10.7)

 
In the limit where N → ∞ higher than first order terms in ǫ, i.e. O ǫ2
terms, are not expected to contribute, thus the time evolution operator can
be expressed as
N
 
> iǫ
u (t, t0 ) = lim 1 − H (tn ) . (10.8)
N→∞ n=1 

10.2 Perturbation Expansion


Consider the case where
H = H0 +λH1 , (10.9)
where λ is real. The perturbation expansion expresses the time evolution
operator u (t, t0 ) of the full Hamiltonian H as
 
u (t, t0 ) = u0 (t, t0 ) + λu1 (t, t0 ) + λ2 u2 (t, t0 ) + O λ3 , (10.10)
where u0 (t, t0 ) is the time evolution of the Hamiltonian H0 . Such an expan-
sion can be very useful for cases where u0 (t, t0 ) can be exactly calculated and
where the parameter λ is small, i.e. |λ| ≪ 1. For such cases only low order
terms in this expansion are needed for approximately evaluating u (t, t0 ).
By employing Eq. (10.8)
N

> iǫ
u (t, t0 ) = lim 1 − (H0 (tn ) + λH1 (tn )) , (10.11)
N→∞ n=1 
one easily obtains the terms u0 , u1 and u2
N
 
> iǫ
u0 (t, t0 ) = lim 1 − H0 (tn ) , (10.12)
N→∞ n=1 

<N iǫ
u1 (t, t0 ) = − lim u0 (t, tn ) H1 (tn ) u0 (tn , t0 )
N→∞ n=1 

i /t ′
=− dt u0 (t, t′ ) H1 (t′ ) u0 (t′ , t0 ) ,
 t0
(10.13)

Eyal Buks Quantum Mechanics - Lecture Notes 338


10.2. Perturbation Expansion

and
N−1
< N
<  ǫ 2
u2 (t, t0 ) = − lim
N→∞ n=1 m=n+1 
× u0 (t, tn ) H1 (tn ) u0 (tn , tm ) H1 (tm ) u0 (tm , t0 )

1 /t ′ /t ′′
=− dt dt
2 t0 t0

× u0 (t, t′ ) H1 (t′ ) u0 (t′ , t′′ ) H1 (t′′ ) u0 (t′′ , t0 ) .


(10.14)

The expansion can be expressed as


 
u (t, t0 ) = u0 (t, t0 ) O (t) + O λ3 , (10.15)

where the operator O (t) is given by

iλ /t ′ λ2 /t /t′
O (t) = 1 − dt H1I (t′ ) − 2 dt′ dt′′ H1I (t′ ) H1I (t′′ ) , (10.16)
 t0  t0 t0

and where H1I (t), which is defined by

H1I (t) ≡ u†0 (t, t0 ) H1 (t) u0 (t, t0 ) , (10.17)

is the so called interaction representation of H1 with respect to u0 .

Exercise 10.2.1. Calculate the expectation value squared |O (t)|2 to low-
est nonvanishing order in λ.

Solution 10.2.1. Since H1 (t) is Hermitian one finds that


|O (t)|2
! ′
"
iλ /t ′ ′ λ2 /t ′ /t ′′ ′ ′′
= 1− dt H1I (t ) − 2 dt dt H1I (t ) H1I (t )
 t0  t0 t0
! ′
"
iλ /t ′ ′ λ2 /t ′ /t ′′ ′′ ′
× 1+ dt H1I (t ) − 2 dt dt H1I (t ) H1I (t )
 t0  t0 t0
! "2
λ2 /t ′
= 1+ 2 dt H1I (t′ )
 t0

λ2 /t ′ /t ′′
− 2 dt dt (H1I (t′ ) H1I (t′′ ) + H1I (t′′ ) H1I (t′ ))
 t0 t0
 3
+O λ ,
(10.18)

Eyal Buks Quantum Mechanics - Lecture Notes 339


Chapter 10. Time-Dependent Perturbation Theory

or
! "2
2 λ2 /t ′ ′
|O (t)| = 1 + 2 dt H1I (t )
 t0

λ2 /t ′ /t ′′
− dt dt H1I (t′ ) H1I (t′′ ) ,
2 t0 t0

(10.19)

thus

2 λ2 /t ′ /t ′′
|O (t)| = 1 − dt dt
2 t0 t0

× [H1I (t′ ) H1I (t′′ ) − H1I (t′ ) H1I (t′′ )] ,


(10.20)

or
2 λ2 /t ′ /t ′′
|O (t)| = 1 − dt dt ∆H1I (t′ ) ∆H1I (t′′ ) , (10.21)
2 t0 t0

where

∆H1I (t) = H1I (t) − H1I (t) . (10.22)

10.3 Transition Probability


Consider the case where the unperturbed Hamiltonian H0 is time indepen-
dent. The eigenvectors of H0 are denoted as |an , and the corresponding
eigenenergies are denoted as En

H0 |an  = En |an  , (10.23)

where

an′ |an  = δ nn′ . (10.24)

In this basis u0 (t, t0 ) is given by


    
i (t − t0 ) iEn (t − t0 )
u0 (t, t0 ) = exp − H0 = exp − |an  an | .
 n

(10.25)

Assuming that initially at time t0 the system is in state |an , what is the
probability to find it later at time t > t0 in the state |am ? The answer to
this question is the transition probability pnm , which is given by

Eyal Buks Quantum Mechanics - Lecture Notes 340


10.3. Transition Probability

pnm = |am | u (t, t0 ) |an |2 . (10.26)

With the help of Eq. (10.16) one finds that


iEm (t−t0 )
e−  am | u (t, t0 ) |an  = am | O (t) |an 
iλ /t ′
= δ nm − dt am | H1I (t′ ) |an 
 t0

λ2 /t ′ /t ′′
− dt dt am | H1I (t′ ) H1I (t′′ ) |an 
2 t0 t0
 3
+O λ ,
(10.27)
thus


iλ /t ′

pnm =
δ nm − dt am | H1I (t′ ) |an 

 t0

2

λ2 /t ′ /t ′′ ′ ′′
 3 

− 2 dt dt am | H1I (t ) H1I (t ) |an  + O λ


.
 t0 t0

(10.28)
In what follows, we calculate the transition probability pnm to lowest non-
vanishing order in λ for the case where n = m, for which the dominant
contribution comes from the term of order λ in Eq. (10.28). For simplicity
the initial time t0 , at which the perturbation is turned on, is taken to be zero,
i.e. t0 = 0. We consider below the following cases:

10.3.1 The Stationary Case

In this case H1 is assumed to be time independent (after being turned on at


t0 = 0). To lowest nonvanishing order in λ Eq. (10.28) yields

t

λ2
/ ′ iω t′
2
pnm = 2
dt e mn
|am | H1 |an |2 , (10.29)


where
Em − En
ω mn = . (10.30)

Using the identity
 Ωt 
/t ′ iΩt′ i Ωt sin 2
dt e = 2e 2 , (10.31)
0 Ω

one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 341


Chapter 10. Time-Dependent Perturbation Theory

4 sin2 ωmn
2
t
2
pnm = 2 2
|am | λH1 |an | . (10.32)
 ω mn

Note that in the limit t → ∞ one finds with the help of Eq. (10.31) that
 
t

4 sin2 Ωt
2

/ iΩt′ ′
2
lim = lim
e dt

t→∞ Ω2 t→∞
0

/t /t ′ ′′
= lim dt′ dt′′ eiΩ(t −t )
t→∞ 0 0
/t
= 2πδ (Ω) dt′
0
= 2πtδ (Ω) .
(10.33)
In this limit pnm is proportional to the time t, i.e. pnm can be expressed as
pnm = wnm t, where wnm is the transition rate, which is given by

wnm = δ (ω mn ) |am | λH1 |an |2 . (10.34)
2
The delta function δ (ωmn ) ensures that energy is conserved in the limit
of long time, and transitions between states having different energies are
excluded. However, such transitions have finite probability to occur for any
finite time interval ∆t. On the other hand, as can be see from Eq. (10.32)
(see also the figure below, which plots the function f (x) = sin2 x/x2 ), the
probability is significant only when ω mn ∆t  1, or alternatively when

∆E∆t   , (10.35)

where ∆E = ω mn .

0.8

0.6

0.4

0.2

-10 -8 -6 -4 -2 0 2 4 x 6 8 10

2
The function f (x) = sin x/x2 .

Eyal Buks Quantum Mechanics - Lecture Notes 342


10.3. Transition Probability

10.3.2 The Near-Resonance Case

In this case H1 is assumed to be given by


′ ′
H1 (t′ ) = Ke−iωt + K† eiωt , (10.36)

where K is an operator that is assumed to be time independent (after being


turned on at t0 = 0), and where the angular frequency ω is a positive constant.
The transition probability is given by [see Eq. (10.28)]

(ωmn −ω)t  

ei 2 sin (ωmn −ω)t
am | λK |an 
4
2
pnm = 2


ω mn − ω
(ω mn +ω)t
 
2
ei 2 sin (ωmn2+ω)t am | λK† |an 

+
,
ω mn + ω

(10.37)
We refer to the case where ω = ω mn as absorption resonance, and to the
case where ω = −ωmn as stimulated emission resonance. Near any of these
resonances ω ≃ ±ω mn the dominant contribution to pnm comes from only
one out of the two terms in Eq. (10.37), thus
 2 (ω −ω)t

 42 sin mn2 2 |am | λK |an |2 ω mn ≃ ω

 (ωmn −ω)
.
pnm ≃ (10.38)
 4 sin2 mn2 2

a | λK† |a 

2 ω

(ω +ω)t

2 (ωmn +ω) m n ≃ −ω

mn

In the long time limit, i.e. in the limit t → ∞, the probability pnm is
found using Eq. (10.33) to be proportional to the time t, i.e. pnm = wnm t,
where the transition rate wnm is given by
?

ω mn ≃ ω

2π 2
 2 δ (ω mn − ω) |am | λK |an |
wnm ≃ 2π

ω mn ≃ −ω

.



2 (10.39)

 2 δ (ω mn + ω) am | λK |an 

In many cases of interest the final state |am  lie in a band of dense states. Let
wn be the total transition rate from the initial state |an . Assume that the
matrix element am | λK |an  does not vary significantly as a function of the
energy Em . For this case the total rate wn can be expressed in terms of the
density of states g (Em ) (i.e. number of states per unit energy) in the vicinity
of the final state |am  [see Eq. (10.39)] as

wn = g (Em ) |am | λK |an |2 , (10.40)

where Em = En + ω. This result is known as the Fermi’s golden rule.

Eyal Buks Quantum Mechanics - Lecture Notes 343


Chapter 10. Time-Dependent Perturbation Theory

10.3.3 H1 is Separable

For this case it is assumed that H1 can be expressed as

H1 (t′ ) = f (t′ ) H̄1 , (10.41)

where f (t′ ) is a real function of time and where H̄1 is time independent
Hermitian operator. To lowest nonvanishing order in λ Eq. (10.28) yields


2
1

/t ′ iωmn t′


2
pnm = 2
dt e f (t )
am | λH̄1 |an 
. (10.42)
 0

10.4 Problems

1. Find the exact time evolution operator u (t, 0) of the Hamiltonian H,


which is given by

H = H0 + Hp , (10.43)

where

H0 = ωa† a , (10.44)
 
Hp = iωζ (t) e2i(ωt−φ) a2 − e−2i(ωt−φ) a†2 , (10.45)

a and a† are the annihilation and creation operators (as defined in chapter
5), ω is positive, φ is real and ζ (t) is an arbitrary real function of time t.
2. Consider a particle having mass m moving under the influence of a one
dimensional potential given by

mω20 x2
V (x) = , (10.46)
2
where the angular resonance frequency ω 0 is a constant. A perturbation
given by

H1 (t′ ) = 2αx cos (ωt′ ) , (10.47)

where the real constant α is assumed to be small, is turned on at time


t = 0. Given that the system was initially at time t = 0 in the ground
state |0 of the unperturbed Hamiltonian, calculate the transition proba-
bility pn0 (t) to the number state |n to lowest nonvanishing order in the
perturbation expansion.

Eyal Buks Quantum Mechanics - Lecture Notes 344


10.4. Problems

3. Repeat the previous exercise with the perturbation

H1 (t′ ) = xf (t′ ) , (10.48)

where the force f (t′ ) is given by


 ′2 
exp − tτ 2
f (t′ ) = α √ , (10.49)
πτ

and where both α and τ are real. Given that the system was initially at
time t → −∞ in the ground state |0 of the unperturbed Hamiltonian,
find the transition probability pn0 to the number state |n in the limit
t → ∞. Compare your approximated result with the exact result given
by Eq. (5.339).
4. Consider a spin 1/2 particle. The Hamiltonian is given by

H = ωSx , (10.50)

where ω is a Larmor frequency and where Sx is the x component of the


angular momentum operator. Given that the spin is initially at time t = 0
in the eigenstate |+; ẑ of the operator Sz (having eigenvalue +/2), what
is the probability p++ (t) to find the spin at the same state |+; ẑ at a
later time t. Compare the exact result with the approximated value that
is obtained from Eq. (10.21).
5. Consider a particle having mass m confined in a potential well given by
+
0 if 0 ≤ x ≤ a
V (x) = . (10.51)
∞ if x < 0 or x > a

The particle is initially at time t → −∞ in the ground state of the well.


A small perturbation
 −( τt )2
λH1 (t) = λ xe , (10.52)
τa
where λ ≪ 1 and where τ is a positive constant having the dimensions
of time, is applied. Calculate the probability to find the particle in the
first excited state in the limit t → ∞.
6. Consider the transition between the energy eigenstates |an  and |am  of
the unperturbed Hamiltonian H0 , which is assumed to be time indepen-
′ ′
dent, due to harmonic perturbation given by H1 (t′ ) = Ke−iωt + K† eiωt
[see Eq. (10.36)], where K is an operator that is assumed to be time in-
dependent (after being turned on at t0 = 0). Calculate to second order in
perturbation theory the transition rate wnm in the long time limit for the
case where the first order contribution vanishes, i.e. for the case where
am | K |an  = 0.

Eyal Buks Quantum Mechanics - Lecture Notes 345


Chapter 10. Time-Dependent Perturbation Theory

10.5 Solutions
1. Expressing the ket vector state as

|ψ = e−iH0 t/ |ψI  , (10.53)

and substituting into the Schrödinger equation, which is given by


d |ψ
i = (H0 + Hp ) |ψ , (10.54)
dt
yield
d |ψI 
i = HI |ψI  . (10.55)
dt
where HI , which is given by

HI = eiH0 t/ Hp e−iH0 t/ , (10.56)

is the interaction picture representation of Hp . With the help of the vector


identity (2.173), which is given by
1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · , (10.57)
2! 3!
and the relations
it , -
H0 , a2 = −2iωta2 , (10.58)

and
it , -
H0 , a†2 = 2iωta†2 , (10.59)

one finds that

eiH0 t/ a2 e−iH0 t/ = a2 e−2iωt , (10.60)


iH0 t/ †2 −iH0 t/ †2 2iωt
e a e =a e , (10.61)

thus
 
HI = iζ (t) e−2iφ a2 − e2iφ a†2 . (10.62)

Since [HI (t) , HI (t′ )] = 0 the solution of Eq. (10.55) is given by


  
i t ′
|ψI (t) = exp − dt HI (t′ ) |ψI (0)
 0
= S (ξ, φ) |ψI (0) ,
(10.63)

Eyal Buks Quantum Mechanics - Lecture Notes 346


10.5. Solutions

where
,  -
S (ξ, φ) = exp ξ e−2iφ a2 − e2iφ a†2 , (10.64)
and where
 t
ξ= dt′ ζ (t′ ) , (10.65)
0

and thus the time evolution operator is thus given by


u (t, 0) = e−iH0 t/ S (ξ, φ) . (10.66)
2. To lowest nonvanishing order in the perturbation expansion one finds
using Eq. (10.38) together with Eqs. (5.11), (5.28) and (5.29) that
(ω −ω)t
2α2 sin2 0 2
pn0 (t) = δ n,1 . (10.67)
mω (ω 0 − ω)2
3. To lowest nonvanishing order in perturbation expansion Eq. (10.42) yields
pn0 = µδ n,1 , (10.68)
where

 ∞
2
1
′ iω0 t′

µ=
dt e f (t )

2mω 0 −∞

α2 1 2 2
= e− 2 ω0 τ .
2mω 0
(10.69)
The exact result is found from Eq. (5.339)
e−µ µn
pn = . (10.70)
n!
To first order in µ both results agree.
4. While the exact result is [see Eq. (6.305)]
ωt
p++ (t) = cos2 , (10.71)
2
Eq. (10.21) yields
1 /t ′ /t ′′
p++ (t) = 1 − dt dt +; ẑ| (ωSx − +; ẑ| ωSx |+; ẑ)2 |+; ẑ
2 0 0
ω 2 +; ẑ| Sx2 |+; ẑ /t ′ /t ′′
= 1− dt dt
2 0 0
 2
ωt
= 1− .
2
(10.72)

Eyal Buks Quantum Mechanics - Lecture Notes 347


Chapter 10. Time-Dependent Perturbation Theory

5. The normalized wavefunctions ψn (x′ ) of the well’s energy eigenstates are


given by
0
′ 2 nπx′
ψn (x ) = sin , (10.73)
a a
and the corresponding eigenenergies are

2 π2 n2
En = , (10.74)
2ma2
where n = 1, 2, · · · . The transition probability is calculated to lowest
nonvanishing order in λ̂ with the help of Eq. (10.42)
 2
t

λ
/ ′ i(E −E )t′ / − t′  2
2
p2,1 =
dt e 2 1
e τ

τa
0


a
2

2 / ′ ′ 2πx′ πx′

dx x sin sin ,
a 0 a a

(10.75)
thus [see Eq. (5.136)
 
256λ2 9π4 2 τ 2
p2,1 = exp − . (10.76)
81π3 8m2 a4

6. For the present case to second order in λ Eq. (10.27) becomes [see Eq.
(10.25)]
iEm (t−t0 )
e− am | u (t, t0 ) |an 


λ  t
/ 2 /t′ ′ ′′
=− 2 dt′ dt′′ eiωml t +iωln t am | H1 (t) |al  al | H1 (t) |an  ,
 l 0 0

(10.77)
where
Em − En
ωmn = , (10.78)

or

Eyal Buks Quantum Mechanics - Lecture Notes 348


10.5. Solutions
iEm (t−t0 )
e−  am | u (t, t0 ) |an 
2 ′
 λ Kml Kln /t /t ′′ i(ωml −ω)t′ +i(ωln −ω)t′′

=− dt dt e
2 0 0
l
 λ2 K∗ Kln /t
lm ′
/t′ ′′ i(ω +ω)t′ +i(ω −ω)t′′
− dt dt e ml ln
2 0 0
l
 λ2 Kml K∗ /t
nl ′
/t′ ′′ i(ωml −ω)t′ +i(ωln +ω)t′′
− dt dt e
2 0 0
l

 λ2 K∗ K∗ /t /t ′′ i(ωml +ω)t′ +i(ωln +ω)t′′
lm nl ′
− dt dt e ,
2 0 0
l
(10.79)
where

Kln = al | K |an  . (10.80)

With the help of the identity


′ ′
/t /t′ ′ ′′ /t ei(Ω1 +Ω2 )t − eiΩ1 t
dt′ dt′′ eiΩ1 t +iΩ2 t = dt′ , (10.81)
0 0 0 iΩ2

one finds that


iEm (t−t0 )
e−  am | u (t, t0 ) |an 
 λ2 Kml Kln /t  
′ i(ωmn −2ω)t′ i(ωml −ω)t′
=− dt e − e
i2 (ω ln − ω) 0
l
 λ2 K∗ Kln /t  
lm ′ iωmn t′ i(ωml +ω)t′
− dt e − e
i2 (ω ln − ω) 0
l
 λ2 Kml K∗ /t  
nl ′ iωmn t′ i(ωml −ω)t′
− dt e − e
l
i2 (ω ln + ω) 0
 λ2 K∗ K∗ /t  
lm nl ′ i(ωmn +2ω)t′ i(ωln +ω)t′
− dt e − e ,
i2 (ω ln + ω) 0
l
(10.82)
or

Eyal Buks Quantum Mechanics - Lecture Notes 349


Chapter 10. Time-Dependent Perturbation Theory
iEm (t−t0 )
e−  am | u (t, t0 ) |an 
2
 λ Kml Kln /t ′
=− 2
dt′ ei(ωmn −2ω)t
l
i (ω ln − ω) 0
 λ2 K∗ K∗ /t ′
− 2
lm nl
dt′ ei(ωmn +2ω)t
i (ω ln + ω) 0
l
 
2 Kml K∗
 λ Kωml Kln
ln −ω
+ ωln +ω
nl
/t ′ i(ωml −ω)t′
+ dt e
i2 0
l
 ∗ 
2 Klm Kln Kml K∗
λ ωln −ω + ωln +ω
nl
/t ′ iωmn t′
− 2
dt e
i 0
l
 λ2 K∗ Kln /t ′
+ lm
dt′ ei(ωml +ω)t
i2 (ω ln − ω) 0
l
 λ2 K∗ K∗ /t ′
+ 2
lm nl
dt′ ei(ωln +ω)t .
i (ω ln + ω) 0
l
(10.83)
By employing the identity [see Eq. (10.33)]

t


/ iΩt′ ′
2
lim
e dt

= 2πtδ (Ω) , (10.84)


t→∞
0

the transition rate wnm can be evaluated in the long time limit. To that
end it is assumed that ω = ±ωml and ω = ±ωln (i.e. it is assumed that the
harmonic perturbation is not at resonance with any first order transition
between the initial |an  or final |am  states and an intermediate state
|al ), and it is further assumed that ω mn > 0 and that ω ≥ 0. Under
/t ′
these assumptions only the terms proportional to 0 dt′ ei(ωmn −2ω)t in
Eq. (10.83) are taken into account, and consequently the transition rate
wnm becomes


2

 λ2 Kml Kln

wnm = 4

δ (ωmn − 2ω) . (10.85)

(ωln − ω)

Eyal Buks Quantum Mechanics - Lecture Notes 350


11. WKB Approximation

The theory of geometrical optics provides an approximated solution to


Maxwell’s equation that is valid for systems whose typical size scales are
much larger than the wavelength λ of electromagnetic waves. In 1926 using a
similar approach the physicists Wentzel, Kramers and Brillouin (WKB) in-
dependently found an approximated solution to the Schrödinger equation in
the coordinate representation for the case where the wavelength associated
with the wavefunction (to be defined below) can be considered as short. Be-
low the WKB approximation is discussed for the time independent and one
dimensional case. This chapter is mainly based on Ref. [3].

11.1 WKB Wavefunction


Consider a point particle having mass m moving under the influence of a one-
dimensional potential V (x). The time independent Schrödinger equation for
the wavefunction ψ (x) is given by [see Eq. (4.50)]
d2 ψ (x) 2m
+ 2 (E − V (x)) ψ (x) = 0 , (11.1)
dx2 
where E is the energy. In terms of the local momentum p (x), which is defined
by

p (x) = 2m (E − V (x)) , (11.2)

the Schrödinger equation becomes


d2 ψ (x)  p 2
+ ψ (x) = 0 . (11.3)
dx2 
Using the notations

ψ (x) = eiW (x)/ , (11.4)

and the relation


!  2 "
d2 ψ (x) i d2 W 1 dW
= − ψ (x) , (11.5)
dx2  dx2  dx
Chapter 11. WKB Approximation

one finds that the Schrödinger equation can be written as


 2
d2 W dW
i 2 − + p2 = 0 . (11.6)
dx dx
In the WKB approach the Plank’s constant  is treated as a small para-
meter. Expanding W as a power series in 

W = W0 + W1 + 2 W2 + · · · (11.7)

one finds that


 2
dW0 d2 W0 dW0 dW1  
− + i 2
− 2 + p2 + O 2 = 0 . (11.8)
dx dx dx dx
The terms of order zero in  yield
 2
dW0
− + p2 = 0 . (11.9)
dx
thus
x
W0 (x) = ± dx′ p (x′ ) , (11.10)
x0

where x0 is a constant.
What is the range of validity of the zero order approximation? As can be
seen by comparing Eq. (11.6) with Eq. (11.9), the approximation W ≃ W0
is valid when the first term in Eq. (11.6) is negligibly small in absolute value
in comparison with the second one, namely when

2


d W

dW
2



. (11.11)
dx dx

It is useful to express this condition in terms of the local wavelength λ (x),


which is given by
2π
λ (x) = . (11.12)
p (x)
By employing the lowest order approximation dW/dx = ±p the condition
(11.11) becomes




≪ 2π . (11.13)

dx

This means that the approximation is valid provided that the change in
wavelength over a distance of one wavelength is small.

Eyal Buks Quantum Mechanics - Lecture Notes 352


11.1. WKB Wavefunction

The terms of 1st order in  of Eq. (11.8) yield an equation for W1


d2 W0  
dW1 i dx2 i d dW0
= dW0
= log . (11.14)
dx 2 dx
2 dx dx

Using Eq. (11.9) one thus has


 
d 1
iW1 − log √ =0. (11.15)
dx p
Therefor, to 1st order in  the wave function is given by

ψ (x) = C+ ϕ+ (x) + C− ϕ− (x) , (11.16)

where
 
x
1 i
ϕ± (x) = √ exp ± dx′ p (x′ ) , (11.17)
p 
x0

and where both C+ and C− are constants.


In general, the continuity equation (4.73), which is given by
dρ dJ
+ =0, (11.18)
dt dx
relates the probability distribution function ρ = |ψ|2 and the current density
J = (/m) Im (ψ∗ dψ/dx) associated with a given one dimensional wave-
function ψ (x). For a stationary ψ (x) the probability distribution function
ρ is time independent, and thus J is a constant. Consider a region where
E > V (x). In such a region, which is classically accessible, the momentum
p (x) is real and positive, and thus the probability distribution function ρ (x)
of the WKB wavefunctions ϕ± (x) is proportional to 1/p. This is exactly
what is expected from a classical analysis of the dynamics, where the time
spent near a point x is inversely proportional to the local classical velocity at
that point v (x) = p (x) /m. With the help of Eq. (4.200) one finds that the
current density J associated with the wavefunction (11.16) is given by
  
  ∗ ∗ ∗ ∗
 dϕ+ dϕ−
J= Im C+ ϕ+ + C− ϕ− C+ + C−
m dx dx
   
 2 ∗ dϕ + 2 ∗ dϕ −
= |C+ | Im ϕ+ + |C− | Im ϕ−
m dx dx
 
dϕ dϕ
+ Im C+ ∗
C− ϕ∗+ − + C+ C− ϕ− +
∗ ∗
.
dx dx
(11.19)
As can be seen from Eq. (11.17), the last term vanishes since ϕ− (x) = ϕ∗+ (x).
Therefor, with the help of Eq. (6.361) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 353


Chapter 11. WKB Approximation

1  
J= |C+ |2 − |C− |2 . (11.20)
m
Thus, the current density J associated with the state ϕ+ (x) is positive,
whereas J < 0 for ϕ− (x). Namely, ϕ+ (x) describes a state propagating from
left to right, whereas ϕ− (x) describes a state propagating in the opposite
direction.

11.2 Turning Point


Consider a point x = a for which E = V (a), namely p (a) = 0 [see Fig.
11.1 (a)]. Such a point is called a turning point since a classical particle that
reaches the point x = a momentarily stops and changes its direction. Near
a turning point the local wavelength λ diverges, and consequently, as can be
seen from Eq. (11.13), the WKB approximation breaks down. Consider the
case where E > V (x) for x > a and where E < V (x) for x < a. In the region
x > a the WKB wave function is expressed using Eq. (11.16), where, for
convenient, the constant x0 is chosen to be a. However, on the other side of
the turning point, namely for x < a, the momentum p (x) becomes imaginary
since E < V (x). Thus, in this region, which is classically forbidden, the wave
function given by Eq. (11.16) contains one exponentially decaying term in
the limit x → −∞ and another exponentially diverging term in the same
limit. To ensure that the wavefunction remains normalizable, the coefficient
of the exponentially diverging term is required to vanish, and thus we seek a
solution having the form
  x 
 C 1
/ ′

 √ exp  dx |p| x<a
ψ (x) =  x|p|  a  x
 . (11.21)
 C / C /
 √+p exp i dx′ p + √−p exp − i dx′ p x > a

a a

Note that the pre-factor 1/ p in the classically forbidden region x < a is

substituted in Eq. (11.21) by 1/ |p|. The ratio between these two factors
in the region x < a is a constant, which is assumed to be absorbed by the
constant C. For given value of C, what are the values of C+ and C− ? It
should be kept in mine that Eq. (11.21) becomes invalid close to the turning
point x = a where the WKB approximation breaks down. Thus, this question
cannot be simply answered by requiring that ψ (x′ ) and its first derivative
are continuous at x = a [e.g., see Eq. (4.127)].
As we have seen above, the WKB approximation breaks down near the
turning point x = a. However the two regions x < a and x > a can be tailored
together by the technique of analytical continuation. In the vicinity of the
turning point, namely for x ≃ a, the potential V (x) can be approximated by

V (x) ≃ V (a) − α (x − a) , (11.22)

Eyal Buks Quantum Mechanics - Lecture Notes 354


11.2. Turning Point

V (x ) − E Im( x )
(a) (b)
Γθ
x θ
a a Re( x )

Im( x ) Im( x )
(c) (d)
Θ+ Θ−

a Re( x ) a Re( x )

Fig. 11.1. (a) The turning point at x = a. (b) The integration trajectory Γθ . The
singly connected region Θ+ (c) and Θ− (d).

where

dV

α=− , (11.23)
dx
x=a
and thus for x ≃ a

p (x) ≃ 2mα (x − a) . (11.24)

Formally, the coordinate x can be considered as complex. Consider a circle


in the complex plane centered at x = a having a radius ρ. The radius ρ > 0
is assumed to be sufficiently large to ensure the validity of the WKB approx-
imation outside it. On the other hand, it is also assumed to be sufficiently
small to allow the employment of the approximation (11.24), namely, for any
point on that circle

x = a + ρeiθ , (11.25)

where θ is real, it is assumed that



p (x) ≃ 2mαρeiθ/2 . (11.26)

Eyal Buks Quantum Mechanics - Lecture Notes 355


Chapter 11. WKB Approximation

We consider below analytical continuation of the wavefunction given by Eq.


(11.21) for the case x > a into a region in the complex plane. Such a region
must excludes the vicinity of the turning point x = a where the WKB approx-
imation breaks down and in addition it is required to be singly connected to
allow analytical continuation. Two such regions are considered below, the first
one, which is labeled as Θ+ (see Fig. 11.1 (c)), excludes the circle |x − a| ≤ ρ
and also excludes the negative imaginary line x = a − ib, whereas the sec-
ond one, which is labeled as Θ− (see Fig. 11.1 (d)), also excludes the circle
|x − a| ≤ ρ and in addition excludes the positive imaginary line x = a + ib,
where in both cases the parameter b is assumed to be real and positive.
To perform the tailoring it is convenient to define the term

i
I± (θ) = ± dx′ p ,

Γθ

where the integration trajectory Γθ [see Fig. 11.1 (b)] contains two sections,
the first along the real axis from x = a to x = a + ρ and the second along the

arc x = a + ρeiθ from θ′ = 0 to θ ′ = θ. With the help of the approximation
(11.26) one finds that
 ρ 
√  θ
i 2mα   ′
I± (θ) = ± dρ′ ρ′ + iρ3/2 dθ′ ei3θ /2 

0 0
  3 
√ 2i e 2 iθ − 1
i 2mα  2 3/2
=± ρ − iρ3/2 
 3 3

2i 2mαρ3/2 32 iθ
=± e
√ 3
2 2mαρ3/2 i(π(1∓ 12 )+ 32 θ)
= e ,
3
(11.27)
thus

2 2mαρ3/2
I± (π) = ± , (11.28)
√ 3 3/2
2 2mαρ
I± (−π) = ∓ . (11.29)
3
The last result allows expressing the analytical continuation of the wavefunc-
tion given by Eq. (11.21) for the case x > a and evaluate its value at the
point x = a − ρ. For the case where the singly connected region Θ+ (Θ− ) is
employed, this is done using integration along the trajectory Γπ (Γ−π ), and
the result is labeled as ψ+ (a − ρ) [ψ− (a − ρ)]

Eyal Buks Quantum Mechanics - Lecture Notes 356


11.2. Turning Point
 √   √ 
2 2mαρ3/2 3/2
C+ exp 3 + C− exp − 2 2mαρ
3
ψ+ (a − ρ) = , (11.30)
(2mαρ)1/4 eiπ/4
 √ 3/2
  √ 
2 2mαρ3/2
C+ exp − 2 2mαρ
3 + C− exp 3
ψ− (a − ρ) = . (11.31)
(2mαρ)1/4 e−iπ/4
Note that the denominators of Eqs. (11.30) and (11.31) are evaluated by

analytical continuation of the factor p [see Eq. (11.26)] along the trajectories
Γπ and Γ−π respectively. On the other hand, according to Eq. (11.21) in the
region x < a one finds by integration along the real axis that
! √ 3
"
C 2 2mαρ 2
ψ (a − ρ) = exp − . (11.32)
(2mαρ)1/4 3

Comparing Eqs. (11.30) and (11.31) with Eq. (11.32) shows that for each of
the two choices Θ+ and Θ− the analytical continuation yields one exponential
term having the same form as the one in Eq. (11.32), and another one, which
diverges in the limit x → −∞. Excluding the diverging terms one finds that
continuity of the non diverging term requires that
C+ C−
C= −iπ/4
= , (11.33)
e eiπ/4
and thus the tailored wavefunction is given by
  x 
 C 1
/ ′
 √
 |p| exp  dx |p| x < a
ψ (x) =  xa  . (11.34)
 2C 1
/ ′ π

 √p cos  dx p − 4 x > a
a

The fact that analytical continuation of the wavefunction in the region x < a
along the trajectory Γπ (Γ−π ) yields only the right to left (left to right) prop-
agating term in the region x > a, and the other term is getting lost along the
way, can be attributed to the limited accuracy of the WKB approximation.
As can be seen from Eq. (11.27), along the integration trajectory Γθ near the
point θ = ±π/3 one term becomes exponentially larger than the other, and
consequently, within the accuracy of this approximation the small term gets
lost.
It is important to keep in mind that the above result (11.34) is obtained
by assuming a particular form for the solution in the region x < a, namely
by assuming that in the classically forbidden region the coefficient of the
exponentially diverging term vanishes. This tailoring role will be employed in
the next section that deals with bound states in a classically accessible region
between two turning points [see Fig. 11.2(a)]. On the other hand, a modified
tailoring role will be needed when dealing with quantum tunneling. For this
case, which will be discussed below, we seek a wave function having the form

Eyal Buks Quantum Mechanics - Lecture Notes 357


Chapter 11. WKB Approximation

V (x ) − E V (x ) − E
(a) (b)

x x
a b a b

Fig. 11.2. The region a ≤ x ≤ b bounded by the two turning points at x = a


and x = b is classically accessible in panel (a), whereas it is classically forbidden in
panel (b) .

  x   
 C+ 1
/ ′ C− 1
/x ′
 √ √
 |p| exp  dx |p| + |p| exp −  dx |p| x < a
ψ (x) = a  x  a . (11.35)
 C i
/ ′ iπ

 √ exp dx p + x > a
p  4
a

Thus, in this case only the term describing propagation from left to right
is kept in the region x > a, and the coefficient of the other term in that
region that describes propagation in the opposite direction is assumed to
vanish. Using the same tailoring technique as in the previous case one find
that C+ = 0 and C− = C, and thus
  
 C 1
/x ′
 √
 |p| exp −  dx |p| x<a
ψ (x) =  x a  . (11.36)
 C i
/ ′ iπ

 p√ exp dx p + x > a
 4
a

11.3 Bohr-Sommerfeld Quantization Rule


Consider a classical accessible region a ≤ x ≤ b bounded by two turning
points at x = a and x = b, namely, consider the case where E > V (x) for
a ≤ x ≤ b and where E < V (x) for x < a and for x > b [see Fig. 11.2(a)].
We seek a normalizable solution, thus the wave function in the classically
forbidden regions x < a and for x > b is assumed to vanish in the limit
x → ±∞. Employing the tailoring role (11.34) with respect to the turning
point at x = a yields the following wave function for the region a ≤ x ≤ b
 x 

2Ca 1 π
ψa (x) = √ cos  dx′ p −  , (11.37)
p  4
a

Eyal Buks Quantum Mechanics - Lecture Notes 358


11.3. Bohr-Sommerfeld Quantization Rule

where Ca is a constant. Similarly, employing the tailoring role (11.34) with


respect to the turning point at x = b yields
 b 

2Cb 1 π
ψb (x) = √ cos  dx′ p −  . (11.38)
p  4
x

The requirement ψa (x) = ψb (x) can be satisfied for any x in the region
a ≤ x ≤ b only if
b
1 π
dx′ p = + nπ . (11.39)
 2
a

where n is integer. Alternatively, this result, which is known as Bohr-


Sommerfeld quantization rule, can be expressed as
@
1 1
dx′ p = n + , (11.40)
2π 2
where
@ b

dx p = 2 dx′ p . (11.41)
a

To normalize the wavefunction ψa (x) = ψb (x) we assume that (a) only


the accessible region a ≤ x ≤ b contributes, since outside this region the
wavefunction exponentially decays; and (b) in the limit of large n the co-
sine term rapidly oscillates and therefore the average of its squared value is
approximately 1/2. Applying these assumptions to ψa (x), which is given by
Eq. (11.37), implies that
b b
′ 2 2 dx′
1≃ dx |ψa (x)| ≃ 2 |Ca | . (11.42)
p
a a

Note that the time period T of classical oscillations between the turning
points x = a and x = b is given by
b
dx′
T =2 . (11.43)
v
a

where v (x) = p (x) /m is the local classical velocity. Thus, by choosing the
pre-factor to be real, one finds that the normalized wavefunction is given by
 x 
0 
m 1 π
ψ (x) = 2 cos  dx′ p −  . (11.44)
pT  4
a

Eyal Buks Quantum Mechanics - Lecture Notes 359


Chapter 11. WKB Approximation

The Bohr-Sommerfeld quantization rule (11.40) can be used to relate the


classical time period T with the energy spacing ∆E = En+1 − En between
consecutive quantum eigenenergies. As can be seen from the validity condition
of the WKB approximation (11.13), the integer n is required to be large to
ensure the validity of Eq. (11.40). In this limit ∆E ≪ E, and thus by taking
the derivative of Eq. (11.40) with respect to energy one finds that
@  −1
′ ∂E
∆E dx = 2π . (11.45)
∂p
In classical mechanics ∂E/∂p is the velocity of the particle v, therefor
@  −1
∂E
dx′ =T , (11.46)
∂p
thus
2π
∆E = . (11.47)
T

11.4 Tunneling
In this case we consider a classical forbidden region a ≤ x ≤ b bounded by
two turning points at x = a and x = b, namely, it is assumed that E < V (x)
for a ≤ x ≤ b and E > V (x) for x < a and for x > b [see Fig. 11.2(b)].
In classical mechanics a particle cannot penetrate into the potential barrier
in the region a ≤ x ≤ b, however such a process is possible in quantum
mechanics. Consider a solution having the form
  x   
 1 i
/ ′ r i
/x ′



√ exp
p  dx p + √p exp −  dx p x<a

  xa   ax 
 / /
C+ C−
ψ (x) = √|p| exp 1 dx′ |p| + √ exp − 1 dx′ |p| a ≤ x ≤ b ,
 |p|

 b  x  b

 t i
/ ′ iπ

 √ exp dx p + 4 x>b
p 
b

where we have introduced the transmission and reflection coefficients t and


r respectively. Such a solution describes an incident wave of unit amplitude
propagating in the region x < a from left to right, a reflected wave having
amplitude r in the same region, and a transmitted wave having amplitude t
in the opposite side of the barrier x > b.
Employing the tailoring role (11.36) yields C+ = 0 and C− = t. Moreover,
employing the tailoring role (11.34) and using the identity
   
x x
1 1
exp − dx′ |p| = τ −1/2 exp − dx′ |p| , (11.48)
 
b a

Eyal Buks Quantum Mechanics - Lecture Notes 360


11.6. Solutions

where
 
b
2
τ = exp − dx′ |p| , (11.49)

a

yield |t| τ −1/2 = 1, thus the transmission probability is given by


 
b
2 2
|t| = τ = exp − dx′ |p| . (11.50)

a

It is important to keep in mind that this approximation is valid only


when τ ≪ 1. One way of seeing this is by noticing that the second tailoring
step, as can be seen from Eq. (11.34), also leads to the conclusion that |r| =
1. This apparently contradicts Eq. (11.50), which predicts a nonvanishing
value for |t|, whereas current conservations, on the other hand, requires that
|t|2 +|r|2 = 1 [see Eq. (4.202)]. This apparent contradiction can be attributed
to limited accuracy of the WKB approximation, however, Eq. (11.50) can be
considered to be a good approximation only provided that τ ≪ 1.

11.5 Problems
1. Calculate the transmission probability τ of a particle having mass m
and energy E through the potential barrier V (x), which vanishes in the
region x < 0 and which is given by V (x) = U − ax in the region x ≥ 0,
where a > 0 and where U > E.
2. Calculate the transmission probability τ of a particle having mass m and
energy E through the potential barrier V (x) = −mω 2 x2 /2, where ω > 0.
Consider the general case without assuming τ ≪ 1.
3. Consider a particle having mass m moving in a one dimensional double
well potential (see Fig. 11.3), which is assumed to be symmetric, i.e.
V (x) = V (−x). In the limit where the barrier separating the two wells
can be considered as impenetrable, each well is characterized by a set of
eigenstates having eigenenergies {En }. To lowest nonvanishing order in
the penetrability of the barrier calculate the eigenenergies of the system.

4. Employ the WKB approximation to derive the eigenenergies of the hy-


drogen atom.

11.6 Solutions
1. The classical turning points are x = 0 and x = (U − E) /α. Thus with
the help of Eq. (11.50) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 361


Chapter 11. WKB Approximation

V (x )

En
x
−a a
Fig. 11.3. Double well potential.

 
√ (U −E)/α
 0
 2 2mα U −E 
τ = exp − dx − x
 α
0
!√ "
4 2m 3
= exp − (U − E) 2 .
3α
(11.51)
2. The factor p/ can be expressed as
.  2 2
p (x) 2m E + mω2 x
= =
 : 
x 2Ex20
= 1+ ,
x20 E0 x2
(11.52)

where x0 = /mω and where E0 = ω. For sufficiently large |x|, namely
for x2 ≫ Ex20 /E0 , one has

p (x) x E
≃ 2+ , (11.53)
 x0 E0 x

where x is assumed to be positive. The corresponding WKB wavefunc-


tions (11.17) in the same limit of large large |x| are given (up to multi-
plication by a constant) by

Eyal Buks Quantum Mechanics - Lecture Notes 362


11.6. Solutions
 
x
1 i
ϕ± (x) =  exp ± dx′ p (x′ )
x0 p/ 
  ! "
/x ′ ′ /x dx′
exp ± x2 dx x exp ±i EE0
i
x′
0
0 x0
≃ 1/2  1/4
x 2Ex20
x0 2 1 + E0 x2
 ±i EE − 12  
x 0 ix2
≃ exp ± 2 .
x0 2x0
(11.54)
Consider a solution having the asymptotic form
 −i EE    i EE − 12  2
 − 12
x ix2
+ r − xx0 ix
0 0
 − exp − 2x 2 exp 2x 2 x → −∞
x0
ψ (x′ ) =  i EE0 − 12  2 0
,

 t xx0 0 ix
exp 2x 2 x→∞
0

(11.55)

where t and r are transmission and reflection coefficients respectively,


which can be related one to another by the technique of analytical con-
tinuation. Consider x as a complex variable
x
= ρeiθ , (11.56)
x0
where ρ > 0 and θ is real. The transmitted term in the limit x → ∞
along the upper semicircle x/x0 = ρeiθ , where 0 ≤ θ ≤ π is given by
 2 2iθ 
 iθ i EE − 12 iρ e
t ρe 0
exp , (11.57)
2

thus for θ = π this term becomes identical to the reflected term at x/x0 =
−ρ, which is given by
 2
i EE − 12 iρ
r (ρ) 0 exp , (11.58)
2

provided that
   
 i E − 1 iρ2 e2iπ E 1 iρ2
t ρeiπ E0 2 exp = r (ρ)i E0 − 2 exp , (11.59)
2 2
or
πE
−ite− E0 = r . (11.60)

Eyal Buks Quantum Mechanics - Lecture Notes 363


Chapter 11. WKB Approximation

Moreover, current conservation requires that |t|2 + |r|2 = 1, thus




2
2
− πE

|t| +
−ite E0
= 1 , (11.61)

2
and therefor the transmission probability τ = |t| is given by
1
τ= 2πE . (11.62)
1 + e− E0

As we have seen above, the analytical continuation of the transmitted


term in the region x → ∞ leads to the reflected term in the region
x → −∞. What about the incident term in the region x → −∞ (the first
term)? Note that this term (the incident one) becomes exponentially
small compared with the reflected term in a section near θ =  3π/4 along

the upper semicircle [due to the exponential factors exp ±ix2 /2x20 ].
Consequently, within the accuracy of the WKB approximation it does
not contribute to the analytically continued value.
3. Consider one of the single-well eigenenergies En . The associated eigen-
state corresponding to the left well is labeled as |n, L and the one corre-
sponding to the right well as |n, R. The effect of finite penetrability of
the barrier can be evaluated using time independent perturbation theory
for the degenerate case [see Eq. (9.38]. For the unperturbed case, where
the barrier separating the two wells can be considered as impenetrable,
the level En is doubly degenerate, and the corresponding eigen space
is spanned by the vectors {|n, L , |n, R}. The projection of the Hamil-
tonian of the system H = p2 /2m + V on this eigen space is represented
by the 2 × 2 matrix Hn , which is given by
 
n, L| H |n, L n, L| H |n, R
Hn = . (11.63)
n, R| H |n, L n, R| H |n, R

Employing the approximations


H |n, L ≃ En |n, L , (11.64)
H |n, R ≃ En |n, R , (11.65)
one finds that
 
1 γ
Hn = En , (11.66)
γ∗ 1

where

γ = n, L |n, R , (11.67)

or in the coordinate representation

Eyal Buks Quantum Mechanics - Lecture Notes 364


11.6. Solutions

∞
γ= dx ϕ∗n,L (x) ϕn,R (x) , (11.68)
−∞

where ϕn,L (x) and ϕn,R (x) are the wavefunctions of the states |n, L
and |n, R respectively, i.e.
ϕn,L (x) = x |n, L , (11.69)
ϕn,R (x) = x |n, R . (11.70)
The main contribution to the overlap integral (11.68) comes from the
classically forbidden region |x| ≤ a, where x = ±a are turning points
(i.e., En = V (a) = V (−a)). With the help of Eq. (11.36) one finds that
! "  
2 1
/x ′ 1
/a ′
a |C| exp −  dx |p| exp −  dx |p|
−a x
γ ≃ dx (11.71)
|p|
−a
  a
a 
2  1 ′  dx
= |C| exp − dx |p| , (11.72)
 |p|
−a −a
(11.73)
where C is the normalization factor
 of the WKB wavefunction, which is
approximately given by C = 2 m/T (T is the time period of classical
oscillations of a particle having energy En in a well) in the limit of large
n [see Eq. (11.44)], thus
/a dx  
4 |p/m| a
−a 1
γ≃ exp − dx′ |p| . (11.74)
T 
−a

Finally, By diagonalizing the matrix Hn one finds that the two eigenen-
ergies are En (1 ± γ).
4. The radial equation for the case of hydrogen is given by [see Eq. (7.61)]
 2 2 
− d
+ Veff (r) ukl (r) = Ekl ukl (r) , (11.75)
2µ dr2

where µ ≃ me is the reduced mass (me is the electron’s mass), and where

e2 l (l + 1) 2
Veff (r) = − + . (11.76)
r 2µr2
The eigenenergies Ekl are calculated using the Bohr-Sommerfeld quanti-
zation rule (11.40)

Eyal Buks Quantum Mechanics - Lecture Notes 365


Chapter 11. WKB Approximation

r2 
1 1
dr 2µ (Ekl − Veff (r)) = k + . (11.77)
π 2
r1

where k is required to be an integer. The points r1,2 are classical turning


points that satisfy

Ekl = Veff (r1,2 ) . (11.78)

Using the notation


r1,2
ρ1,2 = , (11.79)
a0
Ekl
ε=− , (11.80)
EI
where
2
a0 = (11.81)
µe2
is the Bohr’s radius and where
µe4
EI = (11.82)
22
is the ionization energy, Eq. (11.78) becomes
2 l (l + 1)
ε= − , (11.83)
ρ ρ2
thus
1  
ρ1,2 = 1 ± 1 − l (l + 1) ε . (11.84)
ε
Similarly Eq. (11.77) becomes
ρ2 :  
2 l (l + 1) 1
dρ − − ε = π k + , (11.85)
ρ ρ2 2
ρ1

or
 
√ 1
εI = π k + , (11.86)
2
where the integral I, which is given by
ρ2 
(ρ − ρ1 ) (ρ2 − ρ)
I= dρ , (11.87)
ρ
ρ1

Eyal Buks Quantum Mechanics - Lecture Notes 366


11.6. Solutions

can be calculated using the residue theorem


! : "
ρ1 + ρ2 4ρ1 ρ2
I =π 1− . (11.88)
2 (ρ1 + ρ2 )2

Thus the quantization condition (11.86) becomes

Ekl 1
ε=− =  2 . (11.89)
EI
l (l + 1) + k + 12

Comparing with the exact result (7.84) shows that the WKB result is a
good approximation provided that the quantum numbers are large.

Eyal Buks Quantum Mechanics - Lecture Notes 367


12. Path Integration

In this chapter, which is mainly based on Ref. [4], the technique of Feynman’s
path integration is briefly reviewed.

12.1 Charged Particle in Electromagnetic Field

Consider a point particle having mass m and charge q moving under the
influence of electric field E and magnetic field B, which are related to the
scalar potential ϕ and to the vector potential A by
1 ∂A
E = −∇ϕ − , (12.1)
c ∂t
and

B=∇×A. (12.2)

The classical Lagrangian of the system is given by Eq. (1.43)


1 q
L = mṙ2 − qϕ + A · ṙ , (12.3)
2 c
and the classical Hamiltonian is given by Eq. (1.62)
 2
p− qc A
H= + qϕ . (12.4)
2m
The solution of the Euler Lagrange equations (1.8) yields the classical equa-
tion of motion of the system, which is given by Eq. (1.60)
 
1
mr̈ = q E + ṙ × B . (12.5)
c

In what follows, we consider for simplicity the case where both ϕ and A are
time independent. For this case H becomes time independent, and thus the
quantum dynamics is governed by the time evolution operator, which is given
by Eq. (4.9)
Chapter 12. Path Integration
 
iHt
u (t) = exp − . (12.6)


The propagator K (r′b , t; r′a ) is defined by

K (r′b , t; r′a ) = r′b | u (t) |r′a  , (12.7)

where |r′  denotes a common eigenvector of the position operators x, y, and z


with vector of eigenvalues r′ = (x′ , y′ , z ′ ). As can be seen from the definition,
the absolute value squared of the propagator K (r′b , t; r′a ) is the probability
distribution function to find the particle at point r′b at time t given that it
was initially localized at point r′a at time t = 0.
Dividing the time interval (0, t) into N sections of equal duration allows
expressing the time evolution operator as
 N
t
u (t) = u . (12.8)
N

The identity operator in the position representation [see Eq. (3.65)] is given
by

1r = d3 r′ |r′  r′ | . (12.9)

Inserting 1r between any two factors in Eq. (12.8) and using the notation
r′a = r′0 , (12.10)
r′b = r′N , (12.11)
t
ǫ= , (12.12)
N
one finds that

K (r′b , t; r′a ) = r′N | u (ǫ) u (ǫ) u (ǫ) × · · · × u (ǫ) |r′0 


  
= d3 r′1 d3 r′2 × · · · × d3 r′N−1

*)

*)

× r′N | u (ǫ)
r′N−1 r′N−1
u (ǫ)
r′N−2 r′N−2
u (ǫ) × · · · × |r′1  r′1 | u (ǫ) |r′0  ,
(12.13)

thus
#
N−1 N−1
#  
K (r′b , t; r′a ) = d3 r′n K r′m+1 , ǫ; r′m . (12.14)
n=1 m=0

In what follows the limit N → ∞ will be taken,


 and therefor
 it is sufficient
to calculate the infinitesimal propagator K r′m+1 , ǫ; r′m to first order only
in ǫ.

Eyal Buks Quantum Mechanics - Lecture Notes 370


12.1. Charged Particle in Electromagnetic Field

With the help of the relation (12.120), which is given by


 
eǫ(A+B) = eǫA eǫB + O ǫ2 , (12.15)

one has
 
iHǫ
u (ǫ) = exp −

!  2 "  
iǫ p− qc A iǫqϕ  
= exp − exp − + O ǫ2 .
2m 
(12.16)
Equation (12.122), which is given by
    ′2 0 
iǫV2 1 3 ′ ir ǫ ′
exp − = d r exp − i V · r , (12.17)
2m (2πi)3/2 2 m

allows expressing the first term in Eq. (12.16) as


!  2 "   ′2 0 
iǫ p− qc A 1 3 ′ ir ǫ  q  ′
exp − = d r exp − i p− A · r .
2m (2πi)3/2 2 m c
(12.18)

Moreover, with the help of Eq. (12.121), which is given by


 
eǫ(A+B) = eǫB/2 eǫA eǫB/2 + O ǫ3 , (12.19)

one finds that


 0 
ǫ  q 
exp −i p− A · r′
m c
0 ′
  0  0   
ǫ q A·r ǫ ǫ q A · r′
= exp i exp −i p · r′ exp i + O ǫ3/2 .
m c 2 m m c 2
(12.20)
Combining these results yields
  )

K r′m+1 , ǫ; r′m = r′m+1


u (ǫ) |r′m 
  ′2  !0 ,  - "
1 3 ′ ir ǫ q A (r′m ) + A r′m+1 · r′
= d r exp exp i
(2πi)3/2 2 m c 2
   0 
iǫqϕ (r′m ) ) ′
ǫ
× exp − rm+1
exp −i p · r′ |r′m 
 m
 
+O ǫ3/2 .
(12.21)

Eyal Buks Quantum Mechanics - Lecture Notes 371


Chapter 12. Path Integration

In the next step the identity operator in the momentum representation [see
Eq. (3.71)], which is given by

1p = d3 p′ |p′  p′ | , (12.22)

is inserted to the left of the ket vector |r′m . With the help of Eq. (3.75),
which is given by
 ′ ′
1 ip · r
r′ |p′  = 3/2
exp , (12.23)
(2π) 

one finds that


 0 
) ′
ǫ
rm+1
exp −i p · r′ |r′m 
m
 !  "  0 
1 3 ′ ip′ · r′m+1 − r′m ǫ ′ ′
= d p exp exp −i p ·r .
(2π)3  m
(12.24)
Thus, by using Eq. (3.84), which is given by
  ′ ′
1 3 ′ ip · r
3 d p exp = δ (r′ ) , (12.25)
(2π) 

one finds that


 0  ! 0 "
) ′
ǫ ǫ
rm+1
exp −i p · r′ |r′m  = δ r′m+1 − r′m − r′ , (12.26)
m m

and thus
 
   m 3/2 iǫ  
K r′m+1 , ǫ; r′m = exp Lm + O ǫ3/2 , (12.27)
2πiǫ 

where
 r′ ′ 2
m+1 −rm  
m ǫ qA (r′m ) + A r′m+1 r′m+1 − r′m
Lm = −qϕ (r′m )+ · . (12.28)
2 c 2 ǫ
Comparing Eq. (12.28) with the classical Lagrangian of the system, which
is given by Eq. (1.43)
1 q
L = mṙ2 − qϕ + A · ṙ , (12.29)
2 c
shows that Lm is nothing but the Lagrangian at point r′m

Eyal Buks Quantum Mechanics - Lecture Notes 372


12.2. Classical Limit

Lm = L (r′m ) . (12.30)
As we have discussed above, the terms of order ǫ3/2 in Eq. (12.27) are not
expected to contribute to K (r′b , t; r′a ) in the limit of N → ∞. By ignoring
these terms Eq. (12.14) becomes
 N/2 N−1 ! "
Nm # N−1
i t 
′ ′ 3 ′ ′
K (rb , t; ra ) = lim d rn exp L (rm ) .
N→∞ 2πit  N m=0
n=1
(12.31)
Recall that the action in classical physics [see Eq. (1.4)] associated with a
given path is given by

S = dt L . (12.32)

Thus, by defining the integral operator


 r′b  N/2 N−1
#
′ Nm
D [r (t)] = lim d3 r′n , (12.33)
r′a N→∞ 2πit
n=1

the propagator K (r′b , t; r′a ) can be written as


 r′b  
i
K (r′b , t; r′a ) = D [r′ (t′ )] exp Sr′ (t′ ) , (12.34)
r′a 
where
t
Sr′ (t) = dt L [r′ (t)] . (12.35)
0

Equation (12.34), which is known as Feynman’s path integral, expresses the


propagator K (r′b , t; r′a ) in terms of all possible paths r′ (t′ ) satisfying r′ (0) =
r′a and ′ ′ ′ ′
 r (t) = rb , where each path r (t ) contributes a phase factor given by
exp iSr′ (t′ ) / , where Sr′ (t′ ) is the classical action of the path r′ (t′ ).
A note regarding notation: In the above derivation of Eq. (12.34) eigenval-
ues and eigenvectors were denoted with prime (e.g., r′ , |r′ , r′ |, p′ ) to make
them distinguishable from the corresponding operators (e.g., r and p). This
distinction is no longer needed for the rest of this chapter, since no quan-
tum operators are used to evaluate path integrals, and therefore, to make the
notation less cumbersome, we omit the prime notation.

12.2 Classical Limit


Recall that the Hamilton’s principle of least action states that the path taken
by a classical system is the one for which the action S obtains a local mini-
mum. This implies that for any infinitesimal change in the path the resultant

Eyal Buks Quantum Mechanics - Lecture Notes 373


Chapter 12. Path Integration

change in the action δS vanishes (i.e., δS = 0). As we have seen in chapter 1,


this principle leads to Lagrange’s equations of motion (1.8), which are given
by
d ∂L ∂L
= . (12.36)
dt ∂ q̇n ∂qn
While in classical mechanics a definite path is associated with the system’s
dynamics, in quantum mechanics all possible paths are included in Feynman’s
path integral. However, as we show below, in the classical limit the dominant
contribution to the path integral comes only from paths near the classical
one. The classical limit is defined to be the limit where the Plank’s constant
approaches zero  → 0. In this limit the exponent exp (iS/) in the path
integral rapidly oscillates, and consequently contributions from neighboring
paths tend to cancel each other. However, near the classical path, such ’av-
eraging out’ does not occur since according to the principle of least action
δS = 0 for the classical path. Consequently, constructive interference between
neighboring paths is possible near the classical path, and as a result the main
contribution to the path integral in the classical limit comes from the paths
near the classical path.

1
0.8
0.6
0.4
0.2

-4 -2 0 2 x 4
-0.2
-0.4
-0.6
-0.8
-1

Graphical demonstration of the stationary phase approximation The plot


shows the function cos αx2 for the case α = 1. According to the
 in the limit α → ∞, the main contribution
stationary phase approximation,
/∞
to the integral −∞ dx cos αx2 comes from the region near the point
 
x = 0, where d x2 /dx = 0.

12.3 Aharonov-Bohm Effect


Using Eq. (1.43) for the classical Lagrangian of a charged particle in station-
ary electromagnetic field one finds that the classical action (12.35) associated
with a path r (t) in the time interval (0, t) is given by

Eyal Buks Quantum Mechanics - Lecture Notes 374


12.3. Aharonov-Bohm Effect

t  
1 2 q
S= dt mṙ − qϕ + A · ṙ . (12.37)
2 c
0

Consider first the case where the vector potential vanishes, i.e. A = 0. For
this case, the system is said to be conservative, and therefor, as we have seen
in chapter 1 [see Eq. (1.29)], the energy of the system
1
E = mṙ2 + qϕ (12.38)
2
is a constant of the motion (see exercise 5 below). In terms of E the action
S (12.37), which is labeled as S0 for this case where A = 0, can be expressed
as
t  
1 2
S0 = dt mṙ − qϕ
2
0
t
 
= dt −E + mṙ2
0
rb
= −Et + m dr · ṙ .
ra
(12.39)
where ra = r (0) and rb = r (t). Employing Eq. (12.38) again allows rewriting
S0 as
rb
S0 = −Et + dr p (r) , (12.40)
ra

where p (r) is the local classical momentum



p (r) = 2m (E − qϕ (r)) . (12.41)
The phase factor in the path integral corresponding to S0 is given by
 r 
    b
iS0 iEt i
exp = exp − exp  dr p (r) . (12.42)
  
ra

Note the similarity between the second factor in the above equation and
between the WKB wavefunction [see Eq. (11.17)]. In the general case, where
A can be nonzero, the phase factor in the path integral becomes [see Eq.
(12.37)]
 
    rb
iS iS0 iq
exp = exp exp  dr · A . (12.43)
  c
ra

Eyal Buks Quantum Mechanics - Lecture Notes 375


Chapter 12. Path Integration

collector
rb , t
impenetrable
long coil
B≠0

ra ,0
source

Fig. 12.1. Two-slit interference experiment with a very long impenetrable cylinder
placed near the gap between the slits.

12.3.1 Two-slit Interference

Consider a two-slit interference experiment where electrons having charge


q = e are injected from a point source at ra (see Fig. 12.1). A collector at
point rb measures the probability density to detect an electron at that point.
A very long impenetrable cylinder is placed near the gap between the slits in
order to produce a magnetic field inside the cylinder in the direction normal
to the plane of the figure. The field outside the cylinder, however, can be
made arbitrarily small, and in what follows we assume that it vanishes.
The probability density Pb to detect the electron at time t by the collector
located at point rb is given by

Pb = |K (rb , t; ra )|2 , (12.44)

where the propagator (12.34) is given for this case by


 
 rb   
iS0,r(t)  ie 
K (rb , t; ra ) = D [r (t)] exp exp  dr · A , (12.45)
ra  c
r(t)

where the trajectories r (t) satisfy r (0) = ra and r (t) = rb .


How Pb is modified when the magnetic field is turned on, and consequently
the last factor in Eq. (12.45) starts to play a role? To answer this question
it is convenient to divide the sum over all paths into two groups, one for all

Eyal Buks Quantum Mechanics - Lecture Notes 376


12.3. Aharonov-Bohm Effect

paths going through the left slit, and another for all paths going through the
right one. Here we disregard paths crossing a slit more than one time, as their
contribution is expected to be small. In general, the difference Θ12 between
the vector potential phase factor in Eq. (12.45) associated with two different
paths r1 (t) and r2 (t) is given by
 
 
e  
Θ12 =  dr · A− dr · A
c
r1 (t) r2 (t)
@
e
= dr · A ,
c
(12.46)
where the closed path integral is evaluated along the path r1 (t) in the forward
direction from ra to rb , and then along the path r2 (t) in the backward
direction from rb back to ra . This integral can be calculated using Stokes’
theorem [see Eq. (12.2)]
@ 
e e φ
Θ12 = dr · A = ds · B = 2π , (12.47)
c c φ0
where φ is the magnetic flux threaded through the area enclosed by the closed
path, and where
hc
φ0 = (12.48)
e
is the so called flux quantum. While Θ12 vanishes for pairs of paths going
through the same slit, it has the same value Θ12 = 2πφ/φ0 (Θ12 = −2πφ/φ0 )
for all the pairs where r1 (t) goes through the left (right) path and where r2 (t)
goes through the right (left) one. Thus, we come to the somewhat surprising
conclusion that the probability density Pb is expected to be dependent on
the magnetic field. The expected dependence is periodic in the magnetic
flux φ with flux quantum φ0 period. Such dependence cannot be classically
understood, since in this example the electrons can never enter the region in
which the magnetic field B is finite, and thus the Lorentz force vanishes in
the entire region accessible for the electrons outside the impenetrable coil.

12.3.2 Gauge Invariance

Consider the following gauge transformation


A → Ã = A + ∇χ , (12.49)
ϕ → ϕ̃ = ϕ , (12.50)
where χ = χ (r) is an arbitrary smooth and continuous function of r, which
is assumed to be time independent. As can be seen from Eqs. (12.1) and

Eyal Buks Quantum Mechanics - Lecture Notes 377


Chapter 12. Path Integration

(12.2), this transformation leaves E and B unchanged, since ∇× (∇χ) = 0.


In chapter 1 we have seen that such a gauge transformation [see Eqs. (1.44)
and 1.45)] modifies the Lagrangian [see Eq. (1.43)]
q
L → L̃ = L + ∇χ · ṙ , (12.51)
c
and also the action [see Eq. (12.37)]
t
q
S → S̃ = S + dt ∇χ · ṙ
c
0
rb
q
= S+ dr · ∇χ
c
ra
q
= S+ [χ (rb ) − χ (ra )] ,
c
(12.52)
however, the classical motion is unaffected.
In quantum mechanics, the propagator is expressed as a path integral
[see Eq. (12.34)],
 where each path r (t) contributes a phase factor given by
exp iSr(t) / . As can be seen from Eq. (12.52), this phase factor is generally
not singly determined, since it depends on the chosen gauge. This result,
however, should not be considered as paradoxical, since only phase difference
between different paths has any physical meaning. Indeed, as we have seen
above [see Eq. (12.47)], phase difference Θ12 , which determines the relative
phase between two different paths, is evaluated along a closed path, which is
singly determined, and therefore gauge invariant.
Exercise 12.3.1. Given that the wavefunction ψ (r′ , t′ ) solves the Schrödinger
equation with vector A and scalar ϕ potentials, show that the wavefunction
ψ̃ (r′ , t′ ), which is given by
 
′ ′ iqχ (r′ )
ψ̃ (r , t ) = exp ψ (r′ , t′ ) , (12.53)
c
solves the Schrödinger equation with vector à = A + ∇χ and scalar ϕ̃ = ϕ
potentials.
Solution 12.3.1. Using Eq. (3.76) one finds that
   
iqχ iqχ
exp − p exp
c c
   
iqχ iqχ
= exp − p, exp +p
c c
q∇χ
= p+ .
c
(12.54)

Eyal Buks Quantum Mechanics - Lecture Notes 378


12.4. One Dimensional Path Integrals

This result implies that


    
iqχ  q q  iqχ q 
exp − p− A− ∇χ exp = p− A , (12.55)
c c c c c

and therefore the following holds


   
iqχ iqχ
exp − H̃ exp =H, (12.56)
c c

where [see Eq. (1.62)]


 2
p− qc A
H= + qϕ (12.57)
2m
is the Hamiltonian corresponding to the vector potential A, whereas
 2
p− qc A− qc ∇χ
H̃ = + qϕ , (12.58)
2m
is the Hamiltonian corresponding to the vector potential Ã. Thus, one finds
that the state vector

(  

iqχ

ψ̃ = exp |ψ (12.59)
c

solves the Schrödinger equation with H̃, provided that the state vector |ψ
solves the Schrödinger equation with H, and therefore
 
iqχ (r′ )
ψ̃ (r′ , t′ ) = exp ψ (r′ , t′ ) . (12.60)
c

12.4 One Dimensional Path Integrals

Consider a point particle having mass m moving in one dimension along the
x axis under the influence of the potential V (x). The path integral (12.34)
for this case becomes
 N/2 N−1 ! N−1  "
Nm # i t  xm+1 − xm
K (xb , t; xa ) = lim dxn exp L xm , t .
N→∞ 2πit  N m=0
n=1 N
(12.61)

where the Lagrangian is


1
L (x, ẋ) = mẋ2 − V (x) . (12.62)
2

Eyal Buks Quantum Mechanics - Lecture Notes 379


Chapter 12. Path Integration

The solution of the Euler Lagrange equation, which is given by Eq. (1.8)
d ∂L ∂L
= , (12.63)
dt ∂ ẋ ∂x
yields the classical equation of motion of the system
dV
mẍ = − . (12.64)
dx

12.4.1 One Dimensional Free Particle

For this case V (x) = 0.


Exercise 12.4.1. Show that
0 
m im
K (xb , t; xa ) = exp (xb − xa )2 . (12.65)
2πit 2t
Solution 12.4.1. The path integral (12.61) for this case becomes
 N/2 N−1  
imN # imN 
N−1
2
K (xb , t; xa ) = lim − dxn exp (xm+1 − xm ) ,
N→∞ 2πt n=1
2t m=0
(12.66)

or
 
#
 α N/2 N−1 N−1
 2
K (xb , t; xa ) = lim dxn exp −α (xm+1 − xm ) ,
N→∞ π n=1 m=0
(12.67)

where
imN
α=− . (12.68)
2t
/
The first integral dx1 can be calculated using the identity

∞ 3 4 0π 3 α 4
2 2
dx1 exp −α (x2 − x1 ) − α (x1 − x0 ) = exp − (x2 − x0 )2 ,
2α 2
−∞
(12.69)
/
The second integral dx2 can be calculated using the identity

∞ 3 α 4 0 2π 3 α 4
2 2
dx2 exp −α (x3 − x2 ) − (x2 − x0 ) = exp − (x3 − x0 )2 .
2 3α 3
−∞

Eyal Buks Quantum Mechanics - Lecture Notes 380


12.4. One Dimensional Path Integrals

(12.70)
/
Similarly, the nth integral dxn yields
0 
nπ α 2
exp − (xn+1 − x0 ) . (12.71)
(n + 1) α n+1

Therefor, the propagator is given by


0
 α N/2 π 2π (N − 1) π 3 α 4
2
K (xb , t; xa ) = lim ×···× exp − (xb − xa ) ,
N→∞ π 2α 3α Nα N
(12.72)

or
0 
m im 2
K (xb , t; xa ) = exp (xb − xa ) . (12.73)
2πit 2t

As can be seen from the classical equation of motion (12.64), a free particle
moves at a constant velocity. Thus, the classical path satisfying x (0) = xa
and x (t) = xb is given by

(xb − xa ) t′
xc (t′ ) = xa + . (12.74)
t
The corresponding classical action Sc is

m (xb − xa )2
Sc = dt′ L (x, ẋ) = . (12.75)
2t
xc (t′ )

Note that the following holds

d2 Sc m
=− . (12.76)
dxa dxb t
Thus the propagator can be expressed in terms of the classical action Sc as
:  
i d2 Sc i
K (xb , t; xa ) = exp Sc . (12.77)
2π dxa dxb 

As we will see below, a similar expression for the propagator is obtained also
for other cases.

12.4.2 Expansion Around the Classical Path

Motivated by the previous example of a free particle, we attempt below to


relate the propagator for the more general case, where V (x) is allowed to be

Eyal Buks Quantum Mechanics - Lecture Notes 381


Chapter 12. Path Integration

x dependent, with the classical path xc (t′ ) and the corresponding classical
action Sc . Consider a general path x (t′ ) satisfying the boundary conditions
x (0) = xa and x (t) = xb . It is convenient to express the path as

x (t′ ) = xc (t′ ) + δ (t′ ) , (12.78)

where the deviation δ (t′ ) from the classical path xc (t′ ) vanishes at the end
points δ (0) = δ (t) = 0. The action associated with the path x (t′ ) can be
expressed as

S= dt′ L (x, ẋ) , (12.79)
x(t′ )

where the Lagrangian is given by Eq. (12.62).


Expanding S is orders of δ yields

S = Sc + S1 + S2 + · · · , (12.80)

where

Sc = dt′ L (x, ẋ) , (12.81)
 !

"
∂L

∂L

S1 = dt′ δ + δ̇ , (12.82)
∂x
x=xc ∂ ẋ
x=xc
 !


"
2
2
2

∂ L ∂ L ∂ L 2
S2 = dt′
δ2 + 2
δ δ̇ +
δ̇ . (12.83)
∂x2
x=xc ∂x∂ ẋ
x=xc ∂ ẋ2
x=xc
In the general case, higher orders in such an expansion may play an important
role, however, as will be discussed below, in the classical limit the dominant
contribution to the path integral comes from the lowest order terms.

Claim. S1 = 0.

Proof. Integrating by parts the term proportional to δ̇ in the expression for


S1 yields
!

t  !

"
∂L


∂L
d ∂L

S1 = δ
+ dt′

δ. (12.84)
∂ ẋ
x=xc
∂x
x=xc dt ∂ ẋ
x=xc
0

The first term in Eq. (12.84) vanishes due to the boundary conditions δ (0) =
δ (t) = 0, whereas the second one vanishes because xc (t′ ) satisfies the Euler
Lagrange equation (12.63), thus S1 = 0. The fact that S1 vanishes is a direct
consequence of the principle of least action of classical mechanics that was
discussed in chapter 1.

Eyal Buks Quantum Mechanics - Lecture Notes 382


12.4. One Dimensional Path Integrals

Employing the coordinate transformation (12.78) and the expansion of S


around the classical path allows rewriting the path integral (12.61) as

K (xb , t; xa ) = Pc (xb , t; xa ) K (t) , (12.85)

where


iSc
Pc (xb , t; xa ) = exp , (12.86)

  
′ i  3 
K (t) = D [δ (t )] exp S2 + O δ , (12.87)

and where
  N/2 N−1
#
Nm
D [δ (t′ )] = lim dδ n . (12.88)
N→∞ 2πit n=1

The term K (t) is evaluated by integrating over all paths δ (t′ ) satisfying the
boundary conditions δ (0) = δ (t) = 0.

Exercise 12.4.2. Show that


/ ′
dx Pc (xb , t2 ; x′ ) Pc (x′ , t1 ; xa ) K (t1 + t2 )
= . (12.89)
Pc (xb , t1 + t2 ; xa ) K (t1 ) K (t2 )

Solution 12.4.2. As can be seen from the definition of the propagator


(12.7), the following holds
 
dx′ K (xb , t2 ; x′ ) K (x′ , t1 ; xa ) = dx′ xb | u (t2 ) |x′  x′ | u (t1 ) |xa 
= xb | u (t1 + t2 ) |xa 
= K (xb , t1 + t2 ; xa ) .
(12.90)
Requiring that this property is satisfied by the propagator K (xb , t; xa ) that
is given by Eq. (12.85) leads to
/ ′
dx Pc (xb , t2 ; x′ ) Pc (x′ , t1 ; xa ) K (t1 + t2 )
= . (12.91)
Pc (xb , t1 + t2 ; xa ) K (t1 ) K (t2 )

12.4.3 One Dimensional Harmonic Oscillator

For this case the Lagrangian is taken to be given by

mẋ2 mω 1 xẋ mω 2 x2
L (x, ẋ) = + − , (12.92)
2 2 2

Eyal Buks Quantum Mechanics - Lecture Notes 383


Chapter 12. Path Integration

where m, ω and ω 1 are assumed to be real constants. As we will see below,


the term (mω1 /2) xẋ doesn’t affect the dynamics, however, it is taken into
account in order to allows studying the more general case where the La-
grangian contains all possible types of quadratic (in x and ẋ) terms (though,
for simplicity, all coefficients in the Lagrangian are assumed to be time inde-
pendent). Consider a general path x (t′ ) satisfying the boundary conditions
x (0) = xa and x (t) = xb . Using the notation

x (t′ ) = xc (t′ ) + δ (t′ ) ,

the Lagrangian becomes



m  2  
2 2
L (x, ẋ) = ẋc + δ̇ + ω 1 (xc + δ) ẋc + δ̇ − ω (xc + δ) , (12.93)
2

thus the action associated with the path x (t′ ) can be expressed as

S= dt′ L (x, ẋ) = Sc + S1 + S2 , (12.94)
x(t′ )

where
t
m  
Sc = dt′ ẋ2c + ω1 xc ẋc − ω 2 x2c , (12.95)
2
0
t 3 ω1   4
S1 = m dt′ ẋc δ̇ + xc δ̇ + δ ẋc − ω 2 xc δ , (12.96)
2
0
t  2 
m
S2 = dt′ δ̇ + ω1 δ δ̇ − ω 2 δ 2 . (12.97)
2
0

As we have seen above, the principle of least action implies that S1 = 0. Note
that in this case the expansion to second order in δ is exact and all higher
order terms vanish. Thus, the exact solution of this problem will also provide
an approximate solution for systems whose Lagrangian can be approximated
by a quadratic one.

Exercise 12.4.3. Find the classical action Sc of a classical path satisfying


x (0) = xa and x (t) = xb .

Solution 12.4.3. The Euler Lagrange equation (12.63)

d ∂L ∂L
= , (12.98)
dt ∂ ẋ ∂x
for this case yields

Eyal Buks Quantum Mechanics - Lecture Notes 384


12.4. One Dimensional Path Integrals

ẍ = −ω2 x , (12.99)

thus, indeed the term (mω 1 /2) xẋ doesn’t affect the dynamics. Requiring also
boundary conditions x (0) = xa and x (t) = xb leads to

xb sin (ωt′ ) − xa sin (ω (t′ − t))


xc (t′ ) = . (12.100)
sin (ωt)

To evaluate the corresponding action we calculate the following integrals


t
 
dt′ ẋ2c − ω 2 x2c =
0
t 3
ω2 2
dt′ (xb cos (ωt′ ) − xa cos (ω (t′ − t)))
sin2 (ωt)
0
4
2
− (xb sin (ωt′ ) − xa sin ω (t′ − t))

 2 2
 2xa xb
= ω xa + xb cot (ωt) −
sin (ωt)
 
2 ωt
= ω (xa − xb ) cot (ωt) − 2xa xb tan ,
2
(12.101)

and
t
dt′ xc ẋc
0
/t
ω dt′ (xb sin (ωt′ ) − xa sin (ω (t′ − t))) (xb cos (ωt′ ) − xa cos (ω (t′ − t)))
0
=
sin2 (ωt)
x2b − x2a
= ,
2
(12.102)
thus, the action is given by

Sc = dt′ L (x, ẋ)
xc (t′ )
   
mω 2 ωt mω 1 x2b − x2a
= (xa − xb ) cot (ωt) − 2xa xb tan + .
2 2 4
(12.103)

Eyal Buks Quantum Mechanics - Lecture Notes 385


Chapter 12. Path Integration

To evaluate the propagator according to Eq. (12.85) the factor K (t) has
to be determined. This can be done by employing relation (12.89) for the
case where xa = xb = 0
/ ′
dx Pc (0, t2 ; x′ ) Pc (x′ , t1 ; 0) K (t1 + t2 )
= . (12.104)
Pc (0, t1 + t2 ; 0) K (t1 ) K (t2 )
Exercise 12.4.4. Show that
0

K (t) = . (12.105)
2πi sin (ωt)

Solution 12.4.4. By using Eqs. (12.103) and (12.104) one finds that
 3 mω 4 K (t1 + t2 )
dx′ exp i (cot (ωt2 ) + cot (ωt1 )) x′2 = , (12.106)
2 K (t1 ) K (t2 )
thus, using the general integral identity
 ∞ 0
  iπ
dx′ exp iαx′2 = , (12.107)
−∞ α
where α is real, one finds that
:
2πi K (t1 + t2 )
= . (12.108)
mω (cot (ωt2 ) + cot (ωt1 )) K (t1 ) K (t2 )

Alternatively, using the identity


1 sin (ωt1 ) sin (ωt2 )
= , (12.109)
cot (ωt2 ) + cot (ωt1 ) sin (ω (t1 + t2 ))
this can be rewritten as
:
2πi sin (ωt1 ) sin (ωt2 ) K (t1 + t2 )
= . (12.110)
mω sin (ω (t1 + t2 )) K (t1 ) K (t2 )

Consider a solution having the form


0

K (t) = ef (t) , (12.111)
2πi sin (ωt)

where f (t) is an arbitrary function of time. Substituting this into Eq. (12.110)
yields

f (t1 ) + f (t2 ) = f (t1 + t2 ) , (12.112)

thus f (t) = At, where A is a constant. Combining all these results the prop-
agator (12.85) for the present case becomes

Eyal Buks Quantum Mechanics - Lecture Notes 386


12.5. Semiclassical Limit
0

K (xb , t; xa ) = eAt
2πi sin (ωt)
!    "
i mω 2 ωt mω 1 x2b − x2a
× exp (xa − xb ) cot (ωt) − 2xa xb tan + .
 2 2 4
(12.113)

In addition we require that in the limit ω, ω 1 → 0 the above result will


approach the result given by Eq. (12.65) for the propagator of a free particle.
This requirement yields A = 0. Note that
 
ωt   2xa xb
(xa − xb )2 cot (ωt) − 2xa xb tan = x2a + x2b cot (ωt) − .
2 sin (ωt)
(12.114)

For the case ω 1 = 0 the propagator K (xb , t; xa ) becomes


0

K (xb , t; xa ) =
2πi sin (ωt)
 
imω , 2 2
 -
× exp xa + xb cos (ωt) − 2xa xb .
2 sin (ωt)
(12.115)

As can be seen from Eq. (12.103), the following holds

d2 Sc mω
=− , (12.116)
dxa dxb sin (ωt)

thus, similar to the case of a free particle [see Eq. (12.77)], also for the present
case of a harmonic oscillator, the propagator can be expressed in terms of
the classical action Sc as
:  
i d2 Sc i
K (xb , t; xa ) = exp Sc . (12.117)
2π dxa dxb 

12.5 Semiclassical Limit

In the semiclassical limit the Plank’s constant  is considered to be small.


In this limit the dominant contribution to the path integral comes only from
paths near the classical one, which has the least action. This implies that  in
the expansion of S around the classical path (12.80) terms of order O δ 3
can be approximately neglected. Thus, as can be seen from Eq. (12.87), in
this limit [see also Eq. (12.85)] the propagator K (xb , t; xa ) is evaluated by
path integration over the quadratic terms S2 only of the action [see Eq.

Eyal Buks Quantum Mechanics - Lecture Notes 387


Chapter 12. Path Integration

(12.83)]. In the previous section we have exactly calculated the propagator


associated with the quadratic Lagrangian of a harmonic oscillator. The result
was expressed in Eq. (12.117) in terms of the classical action Sc . As can be
seen from Eq. (12.77), the same expression is applicable also for the case
of a free particle. It can be shown that the same form is also applicable
for expressing the propagator K (xb , t; xa ) in the semiclassical limit for the
general case
:  
i d2 Sc i
K (xb , t; xa ) = exp Sc . (12.118)
2π dxa dxb 

The proof of the above result, which requires generalization of the derivation
that led to Eq. (12.117) for the case of a general quadratic Lagrangian, will
not be given here. Another important result, which also is given here without
a proof, generalizes Eq. (12.118) for the case of motion in n spacial dimensions
:    
i d2 Sc i
K (rb , t; ra ) = det exp Sc . (12.119)
2π dra drb 

12.6 Problems
1. Show that
 
eǫ(A+B) = eǫA eǫB + O ǫ2 , (12.120)

where A and B are operators.


2. Show that
 
eǫ(A+B) = eǫB/2 eǫA eǫB/2 + O ǫ3 , (12.121)

where A and B are operators.


3. Show that
    ′2 0 
iǫV2 1 3 ′ ir ǫ ′
exp − = 3/2
d r exp − i V · r , (12.122)
2m (2πi) 2 m

where V is a vector operator.


4. Show that the energy (12.38) is indeed a constant of the motion.
5. Consider a quantum system having time independent Hamiltonian H and
a discrete energy spectrum. Express its partition function Z in terms of
the systems’s propagator K (xb , t; xa ).
6. Consider a one dimensional harmonic oscillator having mass m and res-
onance angular frequency ω in thermal equilibrium at temperature T .
Calculate the matrix elements x′′ | ρ |x′  of the density operator in the
basis of eigenvectors |x′  of the position operator x.

Eyal Buks Quantum Mechanics - Lecture Notes 388


12.7. Solutions

7. Consider a free particle in one dimension having mass m. Calculate the


position wavefunction ψ (x′ , t) at time t given that the position wave-
function ψ (x′ , 0) at time t = 0 is given by
!  ′ 2 "
1 1 x
ψ (x′ , 0) = 1/2
exp − . (12.123)
1/4
π x 2 x 0
0

where x0 is a constant.
8. A particle having mass m is in the ground state of the potential well
V0 (x) = (1/2) mω2 x2 for times t < 0 . At time t = 0 the potential
suddenly changes and becomes V1 (x) = mgx.
a) Calculate the propagator K (xb , t; xa ) from point xa to point xb in
the semiclassical limit for the case where the potential is V1 (x) (i.e.
for the Hamiltonian after the change at t = 0).
b) Use
' the result
( ) of the previous section to calculate the variance
*
(∆x)2 (t) = x2 (t) − x (t)2 of the position operator x at time t.

12.7 Solutions

1. Consider the operator

C (ǫ) = e−ǫA eǫ(A+B) e−ǫB . (12.124)

Clearly, C (0) = 1. Moreover, with the help of Eq. (2.170) one finds that
dC
= −e−ǫA Aeǫ(A+B) e−ǫB + e−ǫA eǫ(A+B) (A + B) e−ǫB − e−ǫA eǫ(A+B) e−ǫB B ,

(12.125)
thus

dC

= −A + (A + B) − B = 0 , (12.126)

ǫ=0

namely
 
C (ǫ) = 1 + O ǫ2 ,

and therefor
 
eǫ(A+B) = eǫA eǫB + O ǫ2 . (12.127)

2. Consider the operator

C (ǫ) = e−ǫB/2 eǫ(A+B) e−ǫB/2 e−ǫA . (12.128)

Eyal Buks Quantum Mechanics - Lecture Notes 389


Chapter 12. Path Integration

As in the previous exercise, it is straightforward (though, somewhat te-


dious) to show that
C (0) = 1 , (12.129)

dC

=0, (12.130)

ǫ=0

d2 C

=0, (12.131)
dǫ2
ǫ=0
thus
 
C (ǫ) = 1 + O ǫ3 , (12.132)

and therefor
 
eǫ(A+B) = eǫB/2 eǫA eǫB/2 + O ǫ3 . (12.133)
3. The proof is trivial using the identity
 ∞ 0
′2 ′ π β 2 /4α
e−αx +βx dx′ = e . (12.134)
−∞ α
4. By taking the time derivative of E one has
dE
= mṙ · r̈ + q∇ϕ·ṙ = ṙ· (mr̈ − qE) . (12.135)
dt
However, according to the equation of motion (1.60) the term in the
brackets vanishes, and therefor dE/dt = 0.
5. Assume that the energy eigenstates of the Hamiltonian H are labeled by
|an  and the corresponding eigenenergies by En , i.e.

H |an  = En |an  , (12.136)


where
an′ |an  = δ nn′ . (12.137)

With the help of the closure relation



1= |an  an | , (12.138)
n

one finds that the propagator K (xb , t; xa ) can be expressed as


 
iHt
K (xb , t; xa ) = xb | exp − |xa 

  
iEn t
= xb |an  exp − an |xa  .
n

(12.139)

Eyal Buks Quantum Mechanics - Lecture Notes 390


12.7. Solutions

Taking xb = xa and integrating over xa yields


 ∞   
iEn t
dxa K (xa , t; xa ) = exp − . (12.140)
−∞ n


Thus, the partition function Z, which is given by Eq. (8.41)



Z= e−βEn , (12.141)
n

where β = 1/kB T , can be expressed as


 ∞
Z= dx′ K (x′ , −iβ; x′ ) . (12.142)
−∞

6. Using Eq. (8.42) one finds that

x′′ | e−βH |x′ 


x′′ | ρ |x′  = , (12.143)
Z
 
where the partition function Z = Tr e−βH can be expressed in terms
of the propagator K (x′′ , t; x′ ) [see Eq. (12.142)]
 ∞
Z= dx′ K (x′ , −iβ; x′ ) . (12.144)
−∞

Furthermore, as can be seen from the definition of the propagator [see


Eq. (12.7)]
iHt
K (x′′ , t; x′ ) = x′′ | e−  |x′  , (12.145)

the following holds

x′′ | e−βH |x′  = K (x′′ , −iβ; x′ ) . (12.146)

Thus, with the help of Eq. (12.115) one finds for the case of a harmonic
oscillator that (recall that sin (ix) = i sinh x and cos (ix) = cosh x)
 ∞
Z= dx′ K (x′ , −iβ; x′ )
−∞
0  ∞  
mω mω [cosh (βω) − 1] x′2
= dx′ exp −
2π sinh (βω) −∞  sinh (βω)
:
1
=
2 [cosh (βω) − 1]
1
= ,
2 sinh βℏω
2
(12.147)

Eyal Buks Quantum Mechanics - Lecture Notes 391


Chapter 12. Path Integration

and therefore one finds, in agreement with Eq. (8.352), that


K (x′′ , −iβ; x′ )
x′′ | ρ |x′  =
Z:
βℏω 2mω
= sinh
2 π sinh (βω)
 
mω , ′2  -
× exp − x + x′′2 cosh (βω) − 2x′ x′′
2 sinh (βω)
 ′ ′′  2  ′ ′′  2
− tanh( β2ω ) x 2x
+x
−coth( β2ω ) x 2x
−x
e 0 0
= 0   ,
x0 π coth βω
2

(12.148)
where
0

x0 = . (12.149)

7. Denoting the state ket vector of the system by |ψ (t) and the time evo-
lution operator by u (t) one has
ψ (x′ , t) = x′ |ψ (t)
= x′ | u (t) |ψ (0)
 ∞
= dx′′ x′ | u (t) |x′′  x′′ |ψ (0)
−∞
 ∞
= dx′′ K (x′ , t; x′′ ) ψ (x′′ , 0) ,
−∞
(12.150)
′ ′′
where the propagator K (x , t; x ) is given by Eq. (12.73)
:  
′ ′′ 1 i (x′ − x′′ )2
K (x , t; x ) = exp , (12.151)
2πiΩtx20 2Ωt x20
and where

Ω= , (12.152)
mx20
thus
:  ∞
′ 1 1
ψ (x , t) = dx′′
1/2
π 1/4 x0 2πiΩtx20 −∞
    ′′ 2  ′ 2 
1 i x i x′ x′′ i x
× exp − 1− − + .
2 Ωt x0 Ωt x20 2Ωt x0
(12.153)

Eyal Buks Quantum Mechanics - Lecture Notes 392


12.7. Solutions

With the help of the identity


∞ 0
1   1 14 4ca+b 2
√ exp −ax2 + bx + c dx = e a , (12.154)
π a
−∞

one finds that


0 !  ′ 2 "
1 1 1 x
ψ (x′ , t) = 1/2
exp − . (12.155)
π1/4 x0 1 + iΩt 2 (1 + iΩt) x0

8. The Lagrangian for times t > 0 is given by


1
L (x, ẋ) = mẋ2 − mgx . (12.156)
2
The Euler Lagrange equation yields the classical equation of motion of
the system

ẍ = −g . (12.157)

The general solution reads

gt2
x = x0 + v0 t − , (12.158)
2
where the constants x0 and v0 are the initial values of the position and
velocity at time t = 0. Given that x = xa at time t = 0 and x = xb at
time t one finds that x0 = xa and
xb − xa gt
v0 = + , (12.159)
t 2
thus the classical trajectory xc (t′ ) is given by
 
′ xb − xa g g
xc (t ) = xa + + t t′ − t′2 . (12.160)
t 2 2

Using the notation

gt2
xt = − , (12.161)
2
the trajectory xc (t′ ) is expressed as

t′ t′2
xc (t′ ) = xa + (xb − xa − xt ) + xt 2 , (12.162)
t t
and the corresponding velocity ẋc (t′ ) is expressed as

Eyal Buks Quantum Mechanics - Lecture Notes 393


Chapter 12. Path Integration

xb − xa − xt 2xt t′
ẋc (t′ ) = + 2 . (12.163)
t t
The Lagrangian along the classical trajectory is given by
1
L (xc , ẋc ) = mẋ2c − mgxc
2
 2
2xt t′ 
m xb −xta −xt + t2 t′ t′2
= − mg xa + (xb − xa − xt ) + xt 2 ,
2 t t
(12.164)
and the corresponding action Sc is given by

Sc = dt′ L (x, ẋ)
xc (t′ )
 2 
t 
 xb −xta −xt + 2xt t′ 
t2 t′ t′2 
=m dt′ − g xa + (xb − xa − xt ) + xt 2

 2 t t  
0
x2t
(xb − xa )2 + 2xt (xb + xa ) − 3
=m
2t
(12.165)
a) In general, the propagator in the semiclassical limit is given by Eq.
(12.118)
:  
i d2 Sc i
K (xb , t; xa ) = exp Sc , (12.166)
2π dxa dxb 

where for the present case Sc is given by Eq. (12.165) and

d2 Sc m
=− , (12.167)
dxa dxb t
thus 0 ! "
2 x2
m im (xb − xa ) + 2xt (xb + xa ) − 3t
K (xb , t; xa ) = exp
2πit  2t
0 ! 2 x2
"
1 1 i (xb − xa ) + 2xt (xb + xa ) − 3t
= exp
x0 2πiωt ωt 2x20
0 ! 2
"
1 1 i 83 x2t + x2a + 2 (2xt − xa ) (xb − xt ) + (xb − xt )
= exp ,
x0 2πiωt ωt 2x20
(12.168)
where

Eyal Buks Quantum Mechanics - Lecture Notes 394


12.7. Solutions
0

x0 = . (12.169)

b) Initially at time t = 0 the wavefunction ψ (x′ ) is given by [see Eq.
(5.120)]
!  2 "
′′ 1 1 x′′
ψ (x , t = 0) = 1/2
exp − , (12.170)
π 1/4 x 2 x0
0

where
0

x0 = . (12.171)

The wave function
 ∞ at time t is evaluated using the propagator

ψ (x , t) = dx′′ K (x′ , t; x′′ ) ψ (x′′ , 0)
−∞
0
1 1 1
=
x0 2πiωt π1/4 x1/2
0
 ∞ 8 2 ′′ 2 ′′ ′ ′
i 3 xt +(x ) +2(2xt −x )(x −xt )+(x −xt )
2  ′′  2
′′ ωt − 12 xx
2x2
× dx e 0 0

−∞
8 2 ′ ′ 2
0 i 3 xt +4xt (x −xt )+(x −xt )
ωt 2x2
1 1 e 0
= 1/2
x0 2πiωt π1/4 x0
x′′ 2
 ∞ (1− ωt
i
)( x0 ) −ix′′ (x′ −xt )
′′ − 2 +
ωtx2
× dx e 0 .
−∞
(12.172)
With the help of the identity
∞ 0
1   1 4a
b2
√ exp −ax2 + bx dx = e , (12.173)
π a
−∞

one finds that


 2 8 ix2 +4ix (x′ −x )
x′ −xt t t t
− 12 1
+ 3
1+iωt x0 2ωtx2
e 0
ψ (x′ , t) = 1/2 √
. (12.174)
π 1/4 x0 1 + iωt

The probability distribution function f (x′ ) = |ψ (x′ , t)|2 to find the


particle near point x′ at time t is thus given by
 2
1 x′ −xt
− 1+(ωt) 2
e x0
f (x′ ) = √ . , (12.175)
πx0 1 + (ωt)2

Eyal Buks Quantum Mechanics - Lecture Notes 395


Chapter 12. Path Integration

therefore f (x′ ) has a Gaussian distribution with a mean value xt and


variance given by
' ( x2  
2 2
(∆x) (t) = 0 1 + (ωt) . (12.176)
2
Note that this result is in agreement with Eq. (5.270).

Eyal Buks Quantum Mechanics - Lecture Notes 396


13. Adiabatic Approximation

The adiabatic approximation can be employed for treating systems having


slowly varying Hamiltonian. This chapter is mainly based on Ref. [5].

13.1 Momentary Diagonalization

The Schrödinger equation (4.1) is given by

d |α
i = H |α . (13.1)
dt
For any given value of the time t the Hamiltonian H (t) is assumed to have
a discrete spectrum

H (t) |n (t) = En (t) |n (t) , (13.2)

where n = 1, 2, · · · , the momentary eigenenergies En (t) are real, and the set
of momentary eigenvectors is assumed to be orthonormal

n (t) |m (t) = δ nm . (13.3)

The general solution can be expanded using the momentary eigenvectors as


a momentary basis

|α (t) = an (t) eiβ n (t) |n (t) . (13.4)
n

The phase factors β n (t) in the expansion (13.4) are chosen to be given by

β n (t) = ξ n (t) + γ n (t) , (13.5)

where the phase factors



1 t ′
ξ n (t) = − dt En (t′ ) (13.6)

are the so-called dynamical phases, and the other phase factors
Chapter 13. Adiabatic Approximation
 t
γ n (t) = i dt′ n (t′ ) |ṅ (t′ ) (13.7)

are the so-called geometrical phases. As we will see below, choosing the phase
factor β n (t) to be given by Eq. (13.5) ensures that the coefficients an (t)
become constants in the adiabatic limit.
Exercise 13.1.1. Show that the term n (t′ ) |ṅ (t′ ) is pure imaginary.
Solution 13.1.1. Note that by taking the derivative with respect to t (de-
noted by upper-dot) of the normalization condition (13.3) one finds that

ṅ |m + n |ṁ = 0 , (13.8)

thus

n |ṁ = − m |ṅ∗ . (13.9)

The last result for the case n = m implies that n (t′ ) |ṅ (t′ ) is pure imagi-
nary, and consequently γ n (t) are pure real.
Substituting Eq. (13.4) into Eq. (13.1) leads to

i ȧn (t) eiβn (t) |n (t)
n

− an (t) β̇ n (t) eiβn (t) |n (t)
n

+i an (t) eiβ n (t) |ṅ (t)
n

= an (t) eiβn (t) En (t) |n (t) .
n
(13.10)
−iβ m (t)
Taking the inner product with m (t)| e yields

 Em (t)
ȧm (t)+iβ̇ m (t) am (t)+ an (t) eiβn (t) e−iβ m (t) m (t) |ṅ (t) = am (t) .
n
i
(13.11)

Since, by definition, the following holds


Em (t)
iβ̇ m (t) = − m (t) |ṁ (t) , (13.12)
i
Eq. (13.11) can be rewritten as

ȧm = − ei(βn (t)−β m (t)) m (t) |ṅ (t) an . (13.13)
n
=m

Eyal Buks Quantum Mechanics - Lecture Notes 398


13.3. Adiabatic Limit

Exercise 13.1.2. Show that for n = m the following holds

m (t)| Ḣ |n (t)
m (t) |ṅ (t) = . (13.14)
En (t) − Em (t)

Solution 13.1.2. Taking the time derivative of Eq. (13.2)

Ḣ |n + H |ṅ = Ėn |n + En |ṅ , (13.15)

and the inner product with m (t)|, where m = n, yields the desired identity.

13.2 Gauge Transformation

The momentary orthonormal basis {|n (t′ )}n , which is made of eigenvec-
tors of H (t), is clearly not singly determined. Consider the following ‘gauge
transformation’ [see for comparison Eq. (12.49)]

|n (t′ ) → |ñ (t′ ) = e−iΛ(t ) |n (t′ ) , (13.16)

where Λ (t′ ) is arbitrary real function of time. The geometrical phase γ n (t),
which is given by Eq. (13.7)
 t
γ n (t) = i dt′ n (t′ ) |ṅ (t′ ) , (13.17)
t0

is transferred to

γ n (t) → γ̃ n (t) = γ n (t) + Λ (t) − Λ (t0 ) . (13.18)

Thus, in general the geometrical phase is not singly determined. However, it


becomes singly determined, and thus gauge invariant, if the path is closed,
namely if H (t) = H (t0 ), since for such a case Λ (t) = Λ (t0 ).

13.3 Adiabatic Limit

In the adiabatic limit the terms m (t) |ṅ (t) are considered to be negligibly
small. As can be seen from Eq. (13.14), this limit corresponds to the case
where the rate of change in time of the Hamiltonian approaches zero. In
this limit the coefficients an (t) become constants [see Eq. (13.13)], and the
solution (13.4) thus becomes

|α (t) = an eiβn (t) |n (t) . (13.19)
n

Eyal Buks Quantum Mechanics - Lecture Notes 399


Chapter 13. Adiabatic Approximation

13.4 The Case of Two Dimensional Hilbert Space

In this case the Hilbert space is two dimensional and the Hamiltonian can be
represented by a 2 × 2 matrix, which is conveniently expressed as a combina-
tion of Pauli matrices

H=h
˙ 0I + h · σ , (13.20)

where I is the 2 × 2 identity matrix, h0 is a real scalar, h = (h1 , h2 , h3 ) is a


three-dimensional real vector, and the components of the Pauli matrix vector
σ are given by
     
01 0 −i 1 0
σ1 = , σ2 = , σ3 = . (13.21)
10 i 0 0 −1

Using the notation h = H ĥ, where H = h · h, and where ĥ is a unit vector,
given in spherical coordinates by

ĥ = (cos ϕ sin θ, sin ϕ sin θ, cos θ) , (13.22)

one finds that


 
cos θ sin θ exp (−iϕ)
H=h
˙ 0I + H . (13.23)
sin θ exp (iϕ) − cos θ

The orthonormal eigenvectors are chosen to be given by [see Eqs. (6.213) and
(6.214)]
  
cos θ2 exp − iϕ2
|+ =˙  , (13.24)
sin θ2 exp iϕ2
  
− sin θ2 exp − iϕ 2
|− =˙  , (13.25)
cos θ2 exp iϕ 2

and the following holds +|+ = −|− = 1, +|− = 0, and

H |± = (h0 ± H) |± . (13.26)

Note that the eigenstates |± are independent of both h0 and H.


The geometrical phase (13.7) can be evaluated by integration along the
path h (t)
 t  h(t)
′ ′ ′
γ n (t) = i dt n (t ) |ṅ (t ) = i dh · n (h)| ∇h |n (h) . (13.27)
0 h(0)

Exercise 13.4.1. Show that


iϕ̂
±| ∇h |± = ∓ ctg θ . (13.28)
2H

Eyal Buks Quantum Mechanics - Lecture Notes 400


13.4. The Case of Two Dimensional Hilbert Space

Solution 13.4.1. Using the expression for a gradient in spherical coordi-


nates (the radial coordinate r in the present case is H)
∂f 1 ∂f 1 ∂f
∇f = r̂ + θ̂ + ϕ̂ , (13.29)
∂r r ∂θ r sin θ ∂ϕ
one finds that
     
θ̂ − sin θ2 exp  − iϕ
2
iϕ̂ − cos θ2 exp − iϕ
2
∇h |+ = + ,
2H cos θ2 exp iϕ 2 2H sin θ sin θ2 exp iϕ
2
(13.30)
thus
  
iϕ̂      − cos θ2 exp − iϕ
+| ∇h |+ = cos θ2 exp iϕ sin θ2 exp − iϕ 2
2H sin θ 2 2 sin θ2 exp iϕ
2
iϕ̂
=− ctg θ .
2H
(13.31)
The term −| ∇h |− is calculated in a similar way.
For the case of a close path, Stock’s theorem can be used to express the
integral in terms of a surface integral over the surface bounded by the close
curve h (t)
@ 
γ n = i dh · n| ∇h |n = i da · (∇ × n| ∇h |n) . (13.32)
S

Exercise 13.4.2. Show that


ih
∇ × ±| ∇h |± = ± . (13.33)
2 |h|3
Solution 13.4.2. Using the general expression for the curl operator in spher-
ical coordinates (again, note that the radial coordinate r in the present case
is H)
 
1 ∂ (sin θAϕ ) ∂Aθ
∇×A = − r̂
r sin θ ∂θ ∂ϕ
 
1 1 ∂Ar ∂
+ − (rAϕ ) θ̂
r sin θ ∂ϕ ∂r
 
1 ∂ (rAθ ) ∂Ar
+ − ϕ̂ ,
r ∂r ∂θ
(13.34)
one finds that
iĥ ∂ cos θ ih
∇ × ±| ∇h |± = ∓ =∓ . (13.35)
2H 2 sin θ ∂θ 2 |h|3

Eyal Buks Quantum Mechanics - Lecture Notes 401


Chapter 13. Adiabatic Approximation

With the help of the last result one thus finds that

1 h 1
γ± = ∓ da · 3 =∓ Ω, (13.36)
2 |h| 2
S

where Ω is the solid angle subtended by the close path h (t) as seen from the
origin. Due to the geometrical nature of the last result, these phase factors
were given the name geometrical phases.

13.5 Transition Probability


The set of equations of motion (13.13) can be rewritten in a matrix form as
d
i |a) = H |a) , (13.37)
dt
where
 
a1
 a2 
|a) =   (13.38)
..
.
is a column vector of the coefficients an ∈ C, and where the matrix elements
of H are given by

Hmn = Hnm = −iei(βn (t)−β m (t)) m (t) |ṅ (t) (13.39)
for the case n = m and Hnn = 0 otherwise.
The inner product between the vectors
   
a1 b1
   
|a) =  a2  , |b) =  b2  , (13.40)
.. ..
. .
is defined by

(a |b) = (b |a)∗ = a∗m bm . (13.41)
m

The set of vectors {|n)} (n = 1, 2, · · · ), having coefficients am = δ nm , forms


an orthonormal basis for the vector space
(n1 |n2 ) = δ n1 n2 . (13.42)
Consider the case where the system is initially at time t0 in the state |n).
What is the probability pnn (t) to find it later at time t > t0 at the same state
|n)? The adiabatic approximation is valid only when pnn ≃ 1. Considering
the matrix H as a perturbation, the probability pnn can be approximated
using time dependent perturbation theory.

Eyal Buks Quantum Mechanics - Lecture Notes 402


13.5. Transition Probability

Exercise 13.5.1. Show that to lowest nonvanishing order the following


holds


2


/t

pnn (t) = 1 −
dt′ Hnm (t′ )
. (13.43)

t0

Solution 13.5.1. By employing Eqs. (10.21) and (10.27) one finds that (re-
call that Hnn = 0)

/t /t
pnn (t) = 1 − dt′ dt′′ (n| H (t′ ) H (t′′ ) |n) . (13.44)
t0 t0
<
Inserting the identity operator 1 = m |m) (m| between H (t′ ) and H (t′′ )

and recalling that Hmn = Hnm lead to

pnn (t) = 1 − pmn (t) , (13.45)
m

where


2

/t


′ ′

pmn (t) =
dt Hnm (t )
. (13.46)

t0

As can be seen from Eq. (13.39), the matrix elements Hnm (t′ ) are pro-
portional to the oscillatory dynamical phase factors
  
i t ′ ′ ′
Hmn ∝ exp (i (ξ n (t) − ξ m (t))) = exp − dt (En (t ) − Em (t )) .

(13.47)

In the adiabatic limit these terms rapidly oscillate and consequently the prob-
abilities pmn (t) are exponentially small. From the same reason, the dominant
contribution to the integral is expected to come from regions where the en-
ergy gap En (t′ )−Em (t′ ) is relatively small. Moreover, it is also expected that
the main contribution to the total ’survival’ probability pnn will come from
those states whose energy Em (t′ ) is close to En (t′ ). Having this is mind, we
study below the transition probability for the case of a two level system. As
we will see below, the main contribution indeed comes from the region near
the point where the energy gap obtains a minimum.

13.5.1 The Case of Two Dimensional Hilbert Space

We calculate below p−+ for the case H=h


˙ · σ, where h (t) is the straight line

h (t) = Ω (0, 1, γt) , (13.48)

Eyal Buks Quantum Mechanics - Lecture Notes 403


Chapter 13. Adiabatic Approximation

where Ω is a positive constant, γ is a real constant, and where the initial time
is taken to be t0 = −∞ and the final one is taken to be t = ∞. In spherical
coordinates h (t) is given by
h (t) = H (0, sin θ, cos θ) , (13.49)
where
.
H = Ω 1 + (γt)2 , (13.50)

cot θ = γt , (13.51)
and where ϕ = π/2. Thus, the energy gap 2H obtains a minimum at time
t = 0. As can be seen from Eqs. (13.24) and (13.25), for any curve lying on
a plane with a constant azimuthal angle ϕ, the following holds

* θ̇

+̇ = |− , (13.52)
2
and therefor

* θ̇
−
+̇ = , (13.53)
2
and

*
*
+
+̇ = −
−̇ = 0 . (13.54)
For the present case one finds using Eq. (13.51) that

* 1 γ
−
+̇ = − . (13.55)
2 1 + (γt)2
This together with Eqs. (13.39) and (13.46) leads to


2


*


/ ′ i(ξ+ (t′ )−ξ− (t′ )) ′

p−+ =
dt e − (t ) +̇ (t )


−∞


 . 
2

/ t′ ′′ 2


∞ exp −2iΩ 0 dt ′′
1 + (γt )


γ / ′
=

dt

1 + (γt′ )2


2 −∞


 / γt′ √ 
2

∞ exp − 2iΩ 1 + τ 2


γ / ′ γ 0 dτ
=

dt


2 −∞ 1 + (γt′ )2


1 , √  √ -2

2

∞ exp − iΩ 2 2

1 / γ τ 1 + τ − ln −τ + 1 + τ

dτ 2

.


2 −∞ 1+τ

(13.56)

Eyal Buks Quantum Mechanics - Lecture Notes 404


13.5. Transition Probability

Exercise 13.5.2. Show that if γ/Ω ≪ 1 then


 

p−+ ≃ exp −π .
γ
Solution 13.5.2. The variable transformation
τ = sinh z , (13.57)
and the identities

1 + τ 2 = cosh z , (13.58)
 1
τ 1 + τ 2 = sinh (2z) , (13.59)
  2

ln −τ + 1 + τ 2 = −z , (13.60)
dτ = cosh z dz , (13.61)
yield

3 1 4

2

∞ exp − iΩ sinh (2z) + z

1 / γ 2

p−+ =

dz
.

(13.62)

2 −∞ cosh z

In the limit γ/Ω ≪ the phase oscillates rapidly and consequently p−+ → 0.
The stationary phase points zn in the complex plane are found from the
condition
 
d 1
0= sinh 2z + z = cosh 2z + 1 , (13.63)
dz 2
thus
 
1
zn = iπ n + , (13.64)
2
where n is integer. Note, however that the term 1/ cosh z has poles at the same
points. Using the Cauchy’s theorem the path of integration can be deformed
to pass close to the point z−1 = −iπ/2. Since the pole at z−1 is a simple one,
the principle value of the integral exists. To avoid passing through the pole at
z−1 a trajectory forming a half circle "above" the pole with radius ε is chosen
were ε → 0. This section gives the dominant contribution which is iπR, where
R is the residue at the pole. Thus the probability p−+ is approximately given
by
 

p−+ ≃ exp −π . (13.65)
γ
The last result can be used to obtain a validity condition for the adia-
batic approximation. In the adiabatic limit p−+ ≪ 1, and thus the condition
πΩ/γ ≫ 1 is required to ensure the validity of the approximation.

Eyal Buks Quantum Mechanics - Lecture Notes 405


Chapter 13. Adiabatic Approximation

13.6 Slow and Fast Coordinates


Consider a system whose Hamiltonian is given by

H = H0 + H1 . (13.66)

The Hamiltonian H0 is assume to depend on a set of coordinates x̄ =


(x1 , x2,... ) and on their canonically conjugate variables p̄ = (p1 , p2,... ), i.e.
H0 = H0 (x̄, p̄). In what follows the coordinates x̄ = (x1 , x2,... ) will be con-
sidered as slow, and thus H0 will be considered as the Hamiltonian of the
slow subsystem. The other part of the system is a fast subsystem, which is
assumed to have a much faster dynamics and its energy spectrum is assumed
to be discrete. The Hamiltonian of the fast subsystem H1 is assumed to para-
metrically depend on the slow degrees of freedom x̄, i.e. H1 = H1 (x̄). This
dependence gives rise to the coupling between the slow and fast subsystems.
An adiabatic approximation can be employed in order to simplify the
equations of motion of the combined system. In what follows, for simplicity,
this method will be demonstrated for the case where the slow subsystem is
assumed to be composed of a set of decoupled harmonic oscillators. For this
case the Hamiltonian H0 is taken to be given by
  p2 ml ω 2l x2l

l
H0 = + , (13.67)
2ml 2
l

where ml and ωl are the mass and angular frequency of mode l, respectively.
The Hamiltonian of the fast subsystem H1 (x̄), which depends paramet-
rically on x̄, has a set of eigenvectors and corresponding eigenvalues for any
given value of x̄

H1 |n (x̄) = εn (x̄) |n (x̄) , (13.68)

where n = 1, 2, · · · . The set of ’local’ eigenvectors {|n (x̄)} is assumed to


form an orthonormal basis of the Hilbert space of the fast subsystem, and
thus the following is assumed to hold

m (x̄) |n (x̄) = δ mn , (13.69)

and [see Eq. (2.23)]



|k (x̄) k (x̄)| = 1F , (13.70)
k

where 1F is the identity operator on the Hilbert space of the fast subsystem.
The state of the entire system ψ (t) at time t is expanded at any point x̄
using the ’local’ basis {|n (x̄)}

ψ (t) = ξ n (x̄, t) |n (x̄) . (13.71)
n

Eyal Buks Quantum Mechanics - Lecture Notes 406


13.6. Slow and Fast Coordinates

In the above expression a mixed notation is being employed. On one hand,


the ket notation is used to denote the state of the fast subsystem (the terms
|n (x̄)). On the other hand, a wavefunction in the position representation
(the terms ξ n (x̄, t)) is employed to denote the state of the slow subsystem.
Substituting the expansion (13.71) into the Schrödinger equation (4.1)


i = Hψ , (13.72)
dt
leads to
 
[H0 + εn (x̄)] ξ n (x̄, t) |n (x̄) = i ξ̇ n (x̄, t) |n (x̄) , (13.73)
n n

where overdot represents time derivative. Projecting m (x̄)| leads to



m (x̄)| H0 ξ n (x̄, t) |n (x̄) + εm (x̄) ξ m (x̄, t) = iξ̇ m (x̄, t) . (13.74)
n

By calculating the term [see Eq. (13.69)]

m (x̄)| p2l ξ n (x̄, t) |n (x̄) = ξ n (x̄, t) m (x̄)| p2l |n (x̄)


+ 2 (pl ξ n (x̄, t)) m (x̄)| pl |n (x̄) + δ mn p2l ξ n (x̄, t) ,
(13.75)

introducing the notation

Am,n;l ≡ − m (x̄)| pl |n (x̄) , (13.76)

and using [see Eq. (13.70)]



m (x̄)| p2l |n (x̄) = m (x̄)| pl |k (x̄) k (x̄)| pl |n (x̄)
k

= −pl Am,n;l + Am,k;l Ak,n;l , (13.77)
k

one obtains

m (x̄)| p2l ξ n (x̄, t) |n (x̄)


! "

= ξ n (x̄, t) −pl Am,n;l + Am,k;l Ak,n;l
k
− 2Am,n;l pl ξ n (x̄, t) + δ mn p2l ξ n (x̄, t) .
(13.78)

With the help of Eqs. (13.67) and (13.74) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 407


Chapter 13. Adiabatic Approximation
! "
 1  
ξ (x̄, t) −pl Am,n;l + Am,k;l Ak,n;l
2ml n n
l k
 1 , -
− −2Am,n;l pl ξ n (x̄, t) + δ mn p2l ξ n (x̄, t)
2ml n
l
 ml ω 2 x2
l l
+ ξ m (x̄, t) + εm (x̄) ξ m (x̄, t) = iξ̇ m (x̄, t) .
2
l
(13.79)
 
Defining the matrices Âl = Am,n;l , (ε̂)m,n = εm δ mn , and the vector
  m,n
ξ̆ = ξ n , the above can be written in a matrix form as
n
? D
 1  2 m ω2 x2 
l l l
·
pl − Âl + + ε̂ ξ̆ = iξ̆ . (13.80)
2ml 2
l

To calculate the off-diagonal matrix elements of Âl we apply pl to Eq.


(13.68) and project m (x̄)|, where m = n

m (x̄)| pl H1 |n (x̄) = m (x̄)| pl εn (x̄) |n (x̄) . (13.81)



Using Eq. (13.69), the definition (13.76) and pl = −i ∂x l
[see Eq. (3.29)] one
finds that [compare with Eq. (13.14)]

m (x̄)| ∂H
∂xl |n (x̄)
1

Am,n;l = i . (13.82)
εn − εm

In the adiabatic approximation the off diagonal elements of Âl [see Eq.
(13.82)] are considered as negligible small. In this case the set of equations
of motion (13.80) becomes decoupled
?   D
 (pl − Am,m;l (x̄))2 ml ω2 x2
l l
iξ̇ m = + + εm (x̄) ξ m . (13.83)
2ml 2
l

As can be seen from the above result (13.83), the adiabatic approximation
greatly simplifies the system’s equations of motion. The effect of the fast
subsystem on the dynamics of the slow one is taken into account by adding a
vector potential Am,m;l (x̄) and a scalar potential εm (x̄) to the Schrödinger
equation of the slow subsystem [compare with Eq. (4.194)]. However, both
potential terms depend on the state m that is being occupied by the fast
subsystem.
Exercise 13.6.1. Show that if m (x̄)| ∂/∂xl |m (x̄) is pure real then

Amm;l (x̄) = 0 . (13.84)

Eyal Buks Quantum Mechanics - Lecture Notes 408


13.7. Problems

Solution 13.6.1. Note that in general the diagonal elements Am,m;l are real
since pl is Hermitian [see Eq. (13.76)]. On the other hand, if m (x̄)| ∂/∂xl |m (x̄)
is pure real then Amm;l (x̄) is pure imaginary, thus for this case Amm;l (x̄) = 0.

13.7 Problems

1. Consider the following ’gauge transformation’


    
cos θ2 exp − iϕ2

*

cos θ2
|+ =˙ → +̃ = ˙ , (13.85)
sin θ2 exp iϕ2
sin θ2 exp (iϕ)
    
− sin θ2 exp − iϕ 2


*

− sin θ2
|− =˙ → −̃ =
˙ . (13.86)
cos θ2 exp iϕ 2
cos θ2 exp (iϕ)
Find an expression for the transformed geometrical phase γ̃ ± (t).
2. Consider a particle having mass m confined by a time dependent potential
well given by
+
0 if 0 ≤ x ≤ a
V (x) = . (13.87)
∞ if x < 0 or x > a

where the width of the well a oscillates in time according to


 
a (t) = a0 1 − α sin2 (ω p t) , (13.88)

where a0 , α and ω p are positive, and where α < 1.


a) Under what condition the adiabatic approximation is expected to be
valid.
b) Calculate the geometrical phases γ n [see Eq. (13.7)] for a cyclic evo-
lution from time t = 0 to time t = π/ω p .
3. Consider a particle of mass m moving in one dimension along the x axis,
whose time-dependent Hamiltonian H (t′ ) is given by

p2 mω 20 x2
H (t′ ) = + + xf (t′ ) , (13.89)
2m 2
where p is the variable canonically conjugate to x, the force f (t′ ) is given
by
 ′2 
exp − tτ 2
f (t′ ) = λ √ , (13.90)
πτ
and ω 0 , λ and τ are real constants. When τ is sufficiently large the prob-
lem can be treated using the adiabatic approximation. Expand the state
of the system |ψ (t) in a basis of momentary eigenvectors of the Hamil-
tonian H (t′ ) and derive the equations of motion for the coefficients in

Eyal Buks Quantum Mechanics - Lecture Notes 409


Chapter 13. Adiabatic Approximation

that basis. The system is initially prepared, at time t → −∞, in the


ground state of the momentary Hamiltonian limt→−∞ H (t). Using the
equations of motion for the coefficients in the momentary basis calcu-
late to lowest nonvanishing order in the adiabatic approximation the
transition probability pn0 to any of the momentary eigenvectors of the
Hamiltonian limt→∞ H (t) in the limit t → ∞.

13.8 Solutions

1. The following holds


 

*

+̃ = exp iϕ |+ , (13.91)
2
 

*

−̃ = exp iϕ |− , (13.92)
2
thus the transformed geometrical phase γ̃ ± (t) [see Eq. (13.18)] becomes

ϕ (t) ϕ (t0 )
γ n (t) → γ̃ n (t) = γ n (t) − + . (13.93)
2 2
2. The momentary eigenenergies of the system are given by

2 π2 n2
En (t) =  2 , (13.94)
2ma20 1 − α sin2 (ω p t)

where n = 1, 2, · · · . The corresponding eigenvectors are denoted by


|n (t).
a) In general, the main contribution to interstate transitions come from
time periods when the energy gap between neighboring eigenener-
gies is relatively small. At any given time the smallest energy gap
between momentary eigenenergies is the one between the two lowest
states H21 (t) = E2 (t) − E1 (t). Furthermore, the main contribution
to the transition probability is expected to come from the regions near
minima points of the energy gap H21 (t). Near the minima point at
time t = 0 the energy gap H21 (t) is given by
32 π 2
H21 (t) =  2
2ma20 1 − α sin2 (ω p t)
32 π2    
= 2 1 + 2α (ωp t)2 + O t3
2ma0
.
32 π2  
= 1 + 4α (ω p t)2 + O t3 .
2ma20
(13.95)

Eyal Buks Quantum Mechanics - Lecture Notes 410


13.8. Solutions

The estimated transition probability for the two dimensional case


is given by Eq. (13.65). In view of the fact that all other energy
gaps between momentary eigenenergies is at least 5/3 larger than
H21 , it is expected that this estimate is roughly valid for the present
case. The requirement that the transition probability given by Eq.
(13.65) is small is taken to be the validity condition for the adiabatic
approximation. Comparing the above expression for H21 (t) near t =
0 with Eq. (13.50) yields the following validity condition

3π3
2ωp α1/2 ≪ . (13.96)
2ma20

b) In general, the term n (t′ ) |ṅ (t′ ) is pure imaginary [see Eq. (13.9)].
On the other hand, the fact that the wavefunctions of one dimensional
bound states can be chosen to be real (see exercise 7 of chapter 4),
implies that n (t′ ) |ṅ (t′ ) is pure real. Thus n (t′ ) |ṅ (t′ ) = 0 and
therefore all geometrical phases vanish.
3. The Hamiltonian H0 ≡ limt→±∞ H (t) is given by

p2 mω 20 x2
H0 = + . (13.97)
2m 2
The eigenvectors |n and eigenenergies En,0 = ω 0 (n + 1/2) of H0 satisfy
the following relation

H0 |n = En,0 |n , (13.98)

where n = 0, 1, , 2 · · · .The momentary Hamiltonian H (t′ ) can be rewrit-


ten as
 2
′ p2 mω 20 f (t′ ) f 2 (t′ )
H (t ) = + x+ 2 − , (13.99)
2m 2 mω 0 2mω20

thus H (t′ ) describes a harmonic oscillator of angular frequency ω 0 hav-


ing a parabolic potential centered at −f (t′ ) /mω 20 , and consequently the
momentary eigenvectors |n (t) of the Hamiltonian H (t′ ) can be chosen
to be coherent states given by [see Eqs. (5.36) and (5.46)]

|n (t′ ) = D (α (t′ )) |0 , (13.100)


 † 
where D (α) = exp αa − α∗ a is the displacement operator, α (t′ ) is
given by

f (t′ )
α (t′ ) = − , (13.101)
21/2 mω20 x0

and

Eyal Buks Quantum Mechanics - Lecture Notes 411


Chapter 13. Adiabatic Approximation
0

x0 = . (13.102)
mω0
The following holds

H (t′ ) |n (t′ ) = En (t′ ) |n (t′ ) , (13.103)

where n = 1, 2, · · · , the momentary eigenenergies En (t′ ) are given by

f 2 (t′ )
En (t′ ) = En,0 − , (13.104)
2mω20

and n (t) |m (t) = δ nm . The adiabatic expansion is given by [see Eq.


(13.4)]
 ′
|ψ (t′ ) = an (t′ ) eiβn (t ) |n (t′ ) , (13.105)
n

where β n (t′ ) = ξ n (t′ ) + γ n (t′ ),


 t′
′ 1
ξ n (t ) = − dt′′ En (t′′ ) , (13.106)

and
 t′
γ n (t′ ) = i dt′′ n (t′′ ) |ṅ (t′′ ) . (13.107)

To lowest nonvanishing order in the adiabatic approximation the transi-


tion probability p0n is given by [see Eqs. (13.14) and (13.46)]


2

∞ n Ḣ |0


/ ′ ′ (t)| (t)

pn0 =
dt′ ei(β0 (t )−βn (t ))
, (13.108)

−∞ En (t) − E0 (t)

where overdot denotes derivative with respect to time. In general the


term n (t′ ) |ṅ (t′ ) can be expressed as [see Eq. (5.37)]
 
n (t′ ) |ṅ (t′ ) = 0| D† (α (t′ )) α̇a† − α̇∗ a D (α (t′ )) |0
= α̇α∗ − α̇∗ α
α̇
= 2i |α|2 Im ,
α
(13.109)
thus for the current case, for which α (t′ ) is real, n (t′ ) |ṅ (t′ ) = 0, and
consequently the geometrical phase γ n (t) vanishes. Furthermore, for the
current case Ḣ = xf˙, and thus [see Eqs. (5.11), (5.28) and (5.29)]

n (t)| Ḣ |0 (t) = 2−1/2 x0 f˙δ n,1 . (13.110)

Eyal Buks Quantum Mechanics - Lecture Notes 412


13.8. Solutions

With the help of the above results one finds that




2
x20


/

′ iω0 t′ ˙

pn0 = 2 2
dt e f
δ n,1 ,
2 ω 0
−∞

or

pn0 = µδ n,1 , (13.111)

where
λ2 1 2 2
µ= e− 2 ω0 τ . (13.112)
2mω 0
Note that the above result is identical to (10.68). Note also that the exact
result of this problem is given by [see Eq. (5.339)]

e−µ µn
pn0 = . (13.113)
n!

Eyal Buks Quantum Mechanics - Lecture Notes 413


14. The Quantized Electromagnetic Field

This chapter discusses the quantization of electromagnetic (EM) field for the
relatively simple case of a free space cavity.

14.1 Classical Electromagnetic Cavity

Consider an empty volume surrounded by conductive walls having infinite


conductivity. The Maxwell’s equations (in Gaussian units) are given by
1 ∂E
∇×B = , (14.1)
c ∂t
1 ∂B
∇×E = − , (14.2)
c ∂t
∇·E = 0 , (14.3)
∇·B = 0 , (14.4)
where c = 2.99 × 108 m s−1 is the speed of light in vacuum. In the Coulomb
gauge the vector potential A is chosen such that

∇·A =0 , (14.5)

and the scalar potential ϕ vanishes in the absence of sources (charge and
current). In this gauge both electric and magnetic fields E and B can be
expressed in terms of A only as [see Eqs. (1.41) and (1.42)]

1 ∂A
E=− , (14.6)
c ∂t
and

B=∇×A. (14.7)

Exercise 14.1.1. Show that


1 ∂2A
∇2 A = . (14.8)
c2 ∂t2
Chapter 14. The Quantized Electromagnetic Field

Solution 14.1.1. The gauge condition (14.5) and Eqs. (14.6) and (14.7)
guarantee that Maxwell’s equations (14.2), (14.3), and (14.4) are satisfied

1 ∂ (∇ × A) 1 ∂B
∇×E=− =− , (14.9)
c ∂t c ∂t
1 ∂ (∇ · A)
∇·E=− =0, (14.10)
c ∂t
∇ · B = ∇ · (∇ × A) = 0 , (14.11)
where in the last equation the general vector identity ∇ · (∇ × A) = 0 has
been employed. Substituting Eqs. (14.6) and (14.7) into the only remaining
nontrivial equation, namely into Eq. (14.1), leads to

1 ∂2A
∇ × (∇ × A) = − . (14.12)
c2 ∂t2
Using the vector identity

∇ × (∇ × A) = ∇ (∇ · A) − ∇2 A , (14.13)

and the gauge condition (14.5) one finds that

1 ∂2A
∇2 A = . (14.14)
c2 ∂t2
Exercise 14.1.2. Consider a solution having the form

A = q (t) u (r) , (14.15)

where q (t) is independent on position r and u (r) is independent on time t.


Show that q (t) and u (r) must satisfy

∇2 u+κ2 u = 0 , (14.16)

and
d2 q
+ω 2κ q = 0 , (14.17)
dt2
where κ is a constant and where

ω κ = cκ . (14.18)

Solution 14.1.2. The gauge condition (14.5) leads to

∇·u =0 . (14.19)

From Eq. (14.8) one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 416


14.1. Classical Electromagnetic Cavity

1 d2 q
q∇2 u = u . (14.20)
c2 dt2
Multiplying by an arbitrary unit vector n̂ leads to
 2 
∇ u · n̂ 1 d2 q
= 2 . (14.21)
u · n̂ c q dt2
The left hand side of Eq. (14.21) is a function of r only while the right hand
side is a function of t only. Therefore, both should equal a constant, which is
denoted as −κ2 , thus

∇2 u+κ2 u = 0 , (14.22)

and
d2 q
+ω 2κ q = 0 , (14.23)
dt2
where

ω κ = cκ . (14.24)

Exercise 14.1.3. Show that the general solution can be expanded as



A= qn (t) un (r) . (14.25)
n

where the set {un } forms a complete orthonormal basis spanning the vector
space of all solutions of Eq. (14.16) satisfying the proper boundary conditions
on the conductive walls having infinite conductivity.
Solution 14.1.3. Equation (14.16) should be solved with the boundary con-
ditions of a perfectly conductive surface. Namely, on the surface S enclosing
the cavity we have B · ŝ = 0 and E × ŝ = 0, where ŝ is a unit vector normal
to the surface. To satisfy the boundary condition for E we require that u be
normal to the surface, namely, u = ŝ (u · ŝ) on S. This condition guarantees
also that the boundary condition for B is satisfied. To see this we calculate
the integral of the normal component of B over some arbitrary portion S ′ of
S. Using Eq. (14.7) and Stoke’s’ theorem one finds that
 
(B · ŝ) dS = q [(∇ × u) · ŝ] dS
S′ ′
@S
=q u · dl ,
C
(14.26)
where the close curve C encloses the surface S ′ . Thus, since u is normal to
the surface, one finds that the integral along the close curve C vanishes, and
therefore

Eyal Buks Quantum Mechanics - Lecture Notes 417


Chapter 14. The Quantized Electromagnetic Field

(B · ŝ) dS = 0 . (14.27)
S′

Since S ′ is arbitrary we conclude that B · ŝ =0 on S. Each solution of Eq.


(14.16) that satisfies the boundary conditions is called an eigen mode. As can
be seen from Eq. (14.23), the dynamics of a mode amplitude q is the same
as the dynamics of a harmonic oscillator having angular frequency ω κ =
cκ. Inner product between different solutions of Eq. (14.16) that satisfy the
boundary conditions can be defined as

u1 , u2  ≡ (u1 · u2 ) dV , (14.28)
V

where the integral is taken over the volume of the cavity. Using Eq. (14.16)
one finds that
 
 2   
κ2 − κ21 (u1 · u2 ) dV = u1 · ∇2 u2 − u2 · ∇2 u1 dV . (14.29)
V V

Using Green’s theorem one finds that


 
 2 
κ2 − κ21 (u1 · u2 ) dV = (u1 · [(ŝ · ∇) u2 ] − u2 · [(ŝ · ∇) u1 ]) dS .
V S
(14.30)
Using Eq. (14.19), the boundary condition u = ŝ (u · ŝ) on S, and writing ∇ =
ŝ (ŝ · ∇) − ŝ× (ŝ × ∇), we find that the right hand side of (14.30) vanishes.
Thus, solutions with different κ2 are orthogonal to each other. Let {un } be a
complete orthonormal basis spanning the vector space of all solutions of Eq.
(14.16) satisfying the boundary conditions. For any two vectors in this basis
the orthonormality condition is

un , um  = (un · um ) dV = δ n,m . (14.31)
V

Using such a basis we can expand the general solution as



A= qn (t) un (r) . (14.32)
n

Exercise 14.1.4. Show that the total electric energy in the cavity is given
by
1  2
UE = q̇ , (14.33)
8πc2 n n

and the total magnetic energy is given by


1  2 2
UB = κ q . (14.34)
8π n n n

Eyal Buks Quantum Mechanics - Lecture Notes 418


14.1. Classical Electromagnetic Cavity

Solution 14.1.4. Using Eqs. (14.6),(14.7) and (14.25) one finds that the
fields are given by
1
E=− q̇n un , (14.35)
c n

and

B= qn ∇ × un . (14.36)
n

The total energy of the field is given by UE +UB , where UE (UB ) is the energy
associated with the electric (magnetic) field, namely,

1
UE = E2 dV , (14.37)
8π V
and

1
UB = B2 dV . (14.38)
8π V

Using Eqs. (14.35) and (14.31) one finds that


1  2
UE = q̇ , (14.39)
8πc2 n n

and using Eq. (14.36) one finds that



1 
UB = qn qm (∇ × un ) · (∇ × um ) dV . (14.40)
8π n,m V

The last integral can be calculated by using the vector identity


∇ · (F1 × F2 ) = (∇ × F1 ) · F2 − F1 · (∇ × F2 ) , (14.41)
applied to un × (∇ × um ), thus
(∇ × un )·(∇ × um ) = ∇·(un × (∇ × um ))+un ·[∇ × (∇ × um )] . (14.42)
Using the divergence theorem and the fact that un and (∇ × um ) are or-
thogonal to each other on S one finds that the volume integral of the first
term vanishes. To calculate the integral of the second term it is convenient
to use the identity
∇ × (∇ × um ) = ∇ (∇ · um ) − ∇2 um . (14.43)
This together with Eqs. (14.19), (14.16), and (14.31) lead to
1  2 2
UB = κ q . (14.44)
8π n n n

Eyal Buks Quantum Mechanics - Lecture Notes 419


Chapter 14. The Quantized Electromagnetic Field

The Lagrangian of the system is given by [see Eq. (1.16)]


 
1  q̇n2 ω 2n qn2
L = UE −UB = − , (14.45)
4πc2 n 2 2

where the symbol overdot is used for derivative with respect to time, and
where ω n = cκn . The Euler-Lagrange equations, given by
 
d ∂L ∂L
− =0, (14.46)
dt ∂ q̇n ∂qn

lead to Eq. (14.23).


The variable canonically conjugate to qn is [see Eq. (1.20)]
∂L 1
pn = = q̇n . (14.47)
∂ q̇n 4πc2
The classical Hamiltonian HF of the field is given by [see Eq. (1.22)]
   4πc2 p2 1 ω 2n qn2

n
HF = pn q̇n − L = + . (14.48)
n n
2 4πc2 2

The Hamilton-Jacobi equations of motion, which are given by


∂HF
q̇n = = 4πc2 pn , (14.49)
∂pn

∂HF ω2
ṗn = − = − n2 qn , (14.50)
∂qn 4πc
lead also to Eq. (14.23). Note that, as expected, the following holds

HF = UE +UB , (14.51)

namely the Hamiltonian expresses the total energy of the system.

14.2 Quantum Electromagnetic Cavity

The coordinates qn and their canonically conjugate variables pn are regarded


as Hermitian operators satisfying the following commutation relations [see
Eqs. (3.5) and (4.41)]

[qn , pm ] = iδ n,m , (14.52)

and

[qn , qm ] = [pn , pm ] = 0 . (14.53)

Eyal Buks Quantum Mechanics - Lecture Notes 420


14.2. Quantum Electromagnetic Cavity

In general, the Heisenberg equation of motion (4.37) of an operator A(H) is


given by

dA(H) 1 3 (H) (H) 4 ∂A(H)


= A , HF + . (14.54)
dt i ∂t
Thus, with the help of Eq. (14.48) one finds that

q̇n = 4πc2 pn , (14.55)

and
ω2n
ṗn = − qn . (14.56)
4πc2
It is useful to introduce the annihilation and creation operators
0  
iωn t ωn 4πic2 pn
an = e qn + , (14.57)
8πc2  ωn
0  
ωn 4πic2 pn
a†n = e−iωn t qn − . (14.58)
8πc2  ωn
The phase factor eiωn t in the definition of an is added in order to make it
time independent. The inverse transformation is given by
:
2πc2   −iωn t 
qn = e an + eiωn t a†n , (14.59)
ωn
0
ω n  −iωn t 
pn = i 2
−e an + eiωn t a†n . (14.60)
8πc
The commutation relations for the these operators are derived directly from
Eqs. (14.52) and (14.53)
, -
an , a†m = δ n,m , (14.61)
, † †-
[an , am ] = an , am = 0 . (14.62)
The Hamiltonian (14.48) can be expressed using Eqs. (14.59) and (14.60) as
  
† 1
HF = ω n an an + . (14.63)
n
2

The eigenstates are the photon-number states |s1, s2 , ..., sn , ..., which satisfy
[see chapter 5]
  
1
HF |s1, s2 , · · · , sn , · · ·  = ωn sn + |s1, s2 , · · · , sn , · · ·  . (14.64)
n
2

The following holds [see Eqs. (5.28) and (5.29)]

Eyal Buks Quantum Mechanics - Lecture Notes 421


Chapter 14. The Quantized Electromagnetic Field

an |s1, s2 , · · · , sn , · · ·  = sn |s1, s2 , · · · , sn − 1, · · ·  , (14.65)

a†n |s1, s2 , · · · , sn , · · ·  = sn + 1 |s1, s2 , · · · , sn + 1, · · ·  . (14.66)
The non-negative integer sn is the number of photons occupying mode n.
The vector potential A (14.25) becomes
:
 2πc2   
A (r, t) = e−iωn t an + eiωn t a†n un (r) . (14.67)
n
ωn

14.3 Periodic Boundary Conditions

Consider the case where the EM field is confined to a finite volume V , which
for simplicity is taken to have a cube shape with edge L = V −1/3 . The
eigen modes and eigen frequencies of the EM field are found in exercise 1
of this chapter for the case where the walls of the cavity are assumed to
have infinite conductance [see Eq. (14.89)]. It is however more convenient to
assume instead periodic boundary conditions, since the spatial dependence
of the resulting eigen modes [denoted by un (r)], can be expressed in terms of
exponential functions, rather than trigonometric functions [see Eqs. (14.82),
(14.83) and (14.84)]. For this case Eq. (14.63) becomes
  
† 1
HF = ω k ak,λ ak,λ + , (14.68)
2
k,λ

and Eq. (14.67) becomes


:
 2πc2   
A (r, t) = ǫ̂k,λ ei(k·r−ωk t) ak,λ + ǫ̂∗k,λ e−i(k·r−ωk t) a†k,λ , (14.69)
ωk V
k,λ

where the eigen frequencies are given by ω k = c |k|. In the limit of large
volume the discrete sum over wave vectors k can be replaced by an integral
  ∞  ∞  ∞
V
→ dkx dky dkz , (14.70)
k (2π)3 −∞ −∞ −∞

and the commutation relations between field operators become


3 4
[ak,λ , ak′ ,λ ] = a†k,λ , a†k′ ,λ = 0 , (14.71)
3 4  
ak,λ , a†k′ ,λ′ = δ λ,λ′ δ k − k′ . (14.72)

The sum over λ contains two terms corresponding to two polarization


vectors ǫ̂k,λ , which are normalized to unity and mutually orthogonal, i,e.
ǫ̂∗k,λ · ǫ̂k,λ′ = δ λ,λ′ . Furthermore, the Coulomb gauge condition requires that

Eyal Buks Quantum Mechanics - Lecture Notes 422


14.4. Problems

ǫ̂k,λ ·k = ǫ̂∗k,λ ·k = 0, i.e. the polarization vectors are required to be orthogonal


to the wave vector k. Linear polarization can be represented by two mutually
orthogonal real vectors ǫ̂k,1 and ǫ̂k,2 , which satisfy ǫ̂k,1 × ǫ̂k,2 = k/ |k|. For
the case of circular polarization the polarization vectors can be chosen to be
given by
1
ǫ̂k,+ = − √ (ǫ̂k,1 + iǫ̂k,2 ) , (14.73)
2
1
ǫ̂k,− = √ (ǫ̂k,1 − iǫ̂k,2 ) . (14.74)
2
For this case of circular polarization the following holds
ǫ̂∗k,λ · ǫ̂k,λ′ = δ λ,λ′ , (14.75)
k
ǫ̂∗k,λ × ǫ̂k,λ′ = iλ δ λ,λ′ . (14.76)
|k|
where λ ∈ {+, −}.

14.4 Problems

1. Find the eigen modes and eigen frequencies of a cavity having a pizza
box shape with volume V = L2 d.
2. Consider two perfectly conducting metallic plates placed in parallel to
each other. The gap between the plates is d and the temperature is as-
sumed to be zero. Calculate the force per unit area acting between the
plates (the Casimir force).
3. Find the average energy per unit volume of the electromagnetic field in
thermal equilibrium at temperature T .
4. Consider an electromagnetic cavity having a set of normal modes. The
waveform of mode n is denoted by un (r), the angular frequency by ω n ,
the annihilation operator by an , and the creation operator by a†n . The
electric filed operator at point r and time t can be expressed as [see Eqs.
(14.6) and (14.67)]

E (r, t) = E(−) (r, t) + E(+) (r, t) , (14.77)

where

E(−) = − 2πωn ieiωn t un (r) a†n , (14.78)
n

E(+) = 2πω n ie−iωn t un (r) an . (14.79)
n

The correlation function G(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) of degree l is de-


fined by

Eyal Buks Quantum Mechanics - Lecture Notes 423


Chapter 14. The Quantized Electromagnetic Field
' (
G(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) ≡ E (−) (r1 ) · · · E (−) (rl ) E (+) (rl+1 ) · · · E (+) (r2l ) .
(14.80)

The normalized coherence function g(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) of de-


gree l is defined by
G(l) (r1 , · · · , rl ; rl+1 , · · · , r2l )
g (l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) ≡ 2l 
. (14.81)
>
G(1) (rs ; rs )
s=1

Consider the case where all modes in the cavity are in their ground state
except of a single mode, which is in a number state with m photons.
Calculate g(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) for such a state.

14.5 Solutions
1. We seek solutions of Eq. (14.16) satisfying the boundary condition that
the tangential component of u vanishes on the walls. Consider a solution
having the form
0
8
ux (r) = ax cos (kx x) sin (ky y) sin (kz z) , (14.82)
V
0
8
uy (r) = ay sin (kx x) cos (ky y) sin (kz z) , (14.83)
V
0
8
uz (r) = az sin (kx x) sin (ky y) cos (kz z) . (14.84)
V
While the boundary condition on the walls x = 0, y = 0, and z = 0 is
guaranteed to be satisfied, the boundary condition on the walls x = L,
y = L, and z = d yields
nx π
kx = , (14.85)
L
ny π
ky = , (14.86)
L
nz π
kz = , (14.87)
d
where nx , ny and nz are integers. This solution clearly satisfies Eq.
(14.16), where the eigen value κ is given by
.
κ = kx2 + ky2 + kz2 . (14.88)

Alternatively,
. using the notation n = (nx , ny , nz ) one has κ = (π/L) n,
where n = n2x + n2y + n2z . Using Eq. (14.24) one finds that the angular
frequency of a mode characterized by the vector of integers n is given by

Eyal Buks Quantum Mechanics - Lecture Notes 424


14.5. Solutions
0   n 2  n 2
n 2 x y z
ωn = cπ + + . (14.89)
L L d
In addition to Eq. (14.16) and the boundary condition, each solution has
to satisfy also the transversality condition ∇ · u = 0 (14.19), which in
the present case reads
k·a= 0 , (14.90)
where k = (kx , ky , kz ) and a = (ax , ay , az ). Thus, for each set of integers
{nx , ny , nz } there are two orthogonal modes (polarizations), unless nx =
0 or ny = 0 or nz = 0. In the latter case, only a single solution exists.
The inner product between two solutions u1 and u1 having vectors of
integers n1 = (nx1 , ny1 , nz1 ) and n2 = (nx2 , ny2 , nz2 ), and vectors of
amplitudes a1 = (ax1 , ay1 , az1 ) and a2 = (ax2 , ay2 , az2 ), respectively, can
be calculated using Eq. (14.28)

u1 , u2  = (u1 · u2 ) dV
V

= (ux1 ux2 + uy1 uy2 + uz1 uz2 ) dV .
V
(14.91)
The following holds

ux1 ux2 dV
V
8
= ax1 ax2
V
 L n π  n π 
x1 x2
× cos x cos x dx
0 L L
 L n π  n π 
y1 y2
× sin y sin y dy
0 L L
 d n π  n π 
z1 z2
× sin z sin z dz ,
0 d d
8 L2 d
= ax1 ax2 δ nx1 ,nx2 δ ny1 ,ny2 δ nz1 ,nz2 .
V 8
(14.92)
Similar results are obtained for the contribution of the y and z compo-
nents. Thus
u1 , u2  = (a1 · a2 ) δ nx1 ,nx2 δ ny1 ,ny2 δ nz1 ,nz2 , (14.93)
and therefore the vectors of amplitudes a are required to be normalized,
i.e. to satisfy a · a = 1, in order to ensure that the solutions u are
normalized.

Eyal Buks Quantum Mechanics - Lecture Notes 425


Chapter 14. The Quantized Electromagnetic Field

2. Employing the results of the previous exercise, the eigen frequencies ωn


are taken to be given by Eq. (14.89), where L is assumed to be much
larger than d. As can be seen from Eq. (14.64), each mode contributes
energy of ω n /2 to the total energy of the ground state of the system,
which is denoted by E (d). Let E (∞) be the ground state energy in the
limit where d → ∞ and let U (d) = E (d)−E (∞) be the potential energy
of the system. Formally, both E (d) and E (∞) are infinite, however, as
we will show below, the divergence can be regulated when evaluating the
difference U (d). The assumption that L is large allows substituting the
discrete sums over nx and ny by integrals when evaluating E (d) and
E (∞). Moreover the discrete sum over nz is substituted by an integral
in the expression for E (∞). The prime on the summation symbol over
nz in the expression for E (d) below implies that a factor of 1/2 should
be inserted if nz = 0, when only one polarization exists (see previous
exercise). Using these approximations and notation one has
U (d) = E (d) − E (∞)
 2  ′  ∞  ∞ 0  πn 2
L z
= c dkx dky kx2 + ky2 +
π nz 0 0 d
 2  ∞  ∞  ∞ .
L d
−c dkx dky dkz kx2 + ky2 + kz2 .
π π 0 0 0
(14.94)
.
In polar coordinates u = kx2 + ky2 and θ = tan−1 (ky /kx ) one has
dkx dky = ududθ, thus
 2  ′  ∞ 0  πn 2
L π z
U (d) = c du u u2 +
π 2 n 0 d
z
 2   ∞
L dπ ∞ 
−c du u dkz u2 + kz2 .
π π2 0 0
(14.95)
Changing the integration variables
 2
ud
x= , (14.96)
π
kz d
Nz = , (14.97)
π
leads to

Eyal Buks Quantum Mechanics - Lecture Notes 426


14.5. Solutions
! ′  "
 ∞
π 2 cL2
U (d) = F (nz ) − dNz F (Nz )
4d3 nz 0
! ∞  ∞ "
π 2 cL2 1 
= F (0) + F (nz ) − dNz F (Nz ) ,
4d3 2 n =1 0
z

(14.98)
where the function F (ξ) is given by
 ∞ .  ∞
2 √
F (ξ) = dx x + ξ = dy y . (14.99)
0 ξ2

Formally, the function F (ξ) diverges. However, the following physical


argument can be employed in order to regulate this divergency. The as-
sumption that the walls of the cavity perfectly conduct is applicable at
low frequencies. However, any metal becomes effectively transparent in
the limit of high frequencies. Thus, the contribution to the ground state
energy of high frequency modes is expected to be effectively d indepen-
dent, and consequently U (d) is expected to become finite. Based on this
argument the divergency in F (ξ) is removed by introducing a cutoff func-
tion f (y) into the integrand in Eq. (14.99)
 ∞

F (ξ) = dy yf (y) . (14.100)
ξ2

While near y = 0 (low frequencies) the cutoff function is assumed to be


given by f (y) = 1, in the limit of large y (high frequencies) the function
f (y) is assumed to approach zero sufficiency fast to ensure that F (ξ) is
finite for any ξ. Moreover, it is assumed that F (∞) → 0. In this case the
Euler-Maclaurin summation formula, which is given by

 ∞
1 < 1 ′ 1 ′′′
F (0) + F (n) − dN F (N) = − F (0) + F (0) + · · · ,
2 n=1 0 12 720
(14.101)

can be employed to evaluate U (d). The following holds


 
F ′ (ξ) = −2ξ 2 f ξ 2 , (14.102)

thus for small ξ [where the cutoff function is assumed to be given by


f (y) = 1] F ′′ (ξ) = −4ξ and F ′′′ (ξ) = −4, and therefore

π2 cL2
U (d) = − . (14.103)
720d3
The force per unit area (pressure) P (d) is found by taking the derivative
with respect to d and by dividing by the area L2

Eyal Buks Quantum Mechanics - Lecture Notes 427


Chapter 14. The Quantized Electromagnetic Field

π2 c
P (d) = − . (14.104)
240d4
The minus sign indicates that the force is attractive.
3. The average energy U in thermal equilibrium is given by Eq. (8.448),
which is given by
∂ log Zc
U =− , (14.105)
∂β
 
where Zc = Tr e−βH is the canonical partition function, H is the
Hamiltonian [see Eq. (14.64)], β = 1/kB T and kB is the Boltzmann’s
constant. The partition function is found by summing over all photon-
number states |s1, s2 , ...


Zc = s1, s2 , ...| e−βH |s1, s2 , ...
s1, s2 ,...=0

 <
ωn (sn + 12 )
= e−β n

s1, s2 ,...=0
! ∞
"
# 
−βωn (sn + 12 )
= e
n sn =0
! "
# 1
= ,
n 2 sinh βω
2
n

(14.106)
where n labels the cavity modes. Using the last result one finds that
∂ log Zc
U =−
∂β
 
1
 ∂ log 2 sinh β2ωn
=−
n
∂β
 ω n βωn
= coth .
n
2 2
(14.107)
It is easy to see that the above sum diverges since the number of modes
in the cavity is infinite. To obtain a finite result we evaluate below the
difference Ud = U (T ) − U (T = 0) between the energy at temperature
T and the energy at zero temperature, which is given by (recall that
coth (x) → 1 in the limit x → ∞)

Eyal Buks Quantum Mechanics - Lecture Notes 428


14.5. Solutions

 ω n  βωn

Ud = coth −1
n
2 2
 ωn
= .
n
eβωn − 1
(14.108)
The angular frequencies ω n of the modes are given by Eq. (14.89). For
simplicity a cubical cavity having volume V = L3 is considered. For this
case Ud is given by (the factor of 2 is due to polarization degeneracy)
∞ 
 ∞ 

αn
Ud = 2kB T αn − 1
. (14.109)
nx =0 ny =0 nz =0
e
.
where n = n2x + n2y + n2z , and where the dimensionless parameter α is
given by
πβc
α= . (14.110)
L
In the limit where α ≪ 1 (macroscopic limit) the sum can be approxi-
mated by the integral
∞
4π αn
Ud = 2kB T dn n2 αn
8 e −1
0
∞
πkB T x3 dx
= ,
α3 ex − 1
0
  
π4
15

(14.111)
thus the energy per unit volume is given by

Ud π2 (kB T )4
= . (14.112)
V 153 c3
4. When only a single mode in the cavity is excited the normalized coher-
ence function g(l) (r1 , · · · , rl ; rl+1 , · · · , r2l ) becomes [see Eqs. (14.78) and
(14.79) and the definition of g (l) ]
' l (
a† al
(l)
g (r1 , · · · , rl ; rl+1 , · · · , r2l ) = l
, (14.113)
a† a

where a and a† are the annihilation and creation operators


√ of the excited
cavity mode. With the help of the relation a |m = n |m − 1 [see Eq.

Eyal Buks Quantum Mechanics - Lecture Notes 429


Chapter 14. The Quantized Electromagnetic Field

(5.28)] one finds that for the given single mode m photon state g(l) is
given by
+ m!
l m ≥ l
g(l) = (m−l)!m . (14.114)
0 m<l

Eyal Buks Quantum Mechanics - Lecture Notes 430


15. Light Matter Interaction

In this chapter the transitions between atomic states that result from inter-
action with an electromagnetic (EM) field are discussed.

15.1 Hamiltonian
Consider an atom in an EM field. The classical Hamiltonian HF of the EM
field is given by Eq. (14.48). For the case of hydrogen, and in the absence
of EM field, the Hamiltonian of the atom is given by Eq. (7.2). In general,
the classical Hamiltonian of a point particle having charge e and mass me
in an EM field having scalar potential ϕ and vector potential A is given by
Eq. (1.62). In the Coulomb gauge the vector potential A is chosen such that
∇ · A = 0, and the scalar potential ϕ vanishes provided that no sources
(charge and current) are present. The EM field is assumed to be sufficiently
small to allows employing the following approximation
 e 2 e
p− A ≃ p2 −2 A · p , (15.1)
c c
where p is the momentum vector. Recall that in the Coulomb gauge the vector
operators p and A satisfy the relation p · A = A · p, as can be seen from
Eqs. (6.167) and (6.324). These results and approximation allow expressing
the Hamiltonian of the system as
H = H0 + HF + Hp , (15.2)
where H0 is the Hamiltonian of the atom in the absence of EM field, and
where Hp , which is given by
e
Hp = − A·p , (15.3)
me c
is the coupling Hamiltonian between the atom and the EM field.
In the quantum case The Hamiltonian HF of the EM field is given by Eq.
(14.68)
  
1
HF = ω k a†k,λ ak,λ + , (15.4)
2
k,λ
Chapter 15. Light Matter Interaction

and the vector potential A is given by Eq. (14.69)


:
 2πc2   
A (r, t) = ǫ̂k,λ ei(k·r−ωk t) ak,λ + ǫ̂∗k,λ e−i(k·r−ωk t) a†k,λ . (15.5)
ωk V
k,λ

15.2 Transition Rates

While the Hamiltonian Hp is considered as a perturbation, the unperturbed


Hamiltonian is taken to be H0 + HF . The eigenvectors of H0 + HF are la-
beled as |{sk,λ } , η. While the integers sk,λ represent the number of photons
occupying each of the modes of the EM field, the index η labels the atomic
energy eigenstate. The following holds

H0 |{sk,λ } , η = Eη |{sk,λ } , η ,

where Eη is the energy of the atomic state, and


  
1
HF |{sk,λ } , η = ω k sk,λ + |{sk,λ } , η . (15.6)
2
k,λ

15.2.1 Spontaneous Emission

Consider the case where the system is initially in a state |i = |{sk,λ = 0} , ηi ,
for which all photon occupation numbers are zero, and the atomic state is la-
beled by the index ηi . The final state is taken to be |f = a†k,λ |{sk,λ = 0} , ηf ,
for which one photon is created in mode k, λ, and the atomic state is labeled
by the index η f . To lowest nonvanishing order in perturbation theory the
transition rate wi,f is given by Eq. (10.34)

wi,f = δ (ω k − ωi,f ) |f| Hp |i|2 , (15.7)
2
 
where ω i,f = Eηi − Eηf /. With the help of Eqs. (14.69) and (15.3) wi,f
can be rewritten as
 2 2 2

2
e 4π c

wi,f = δ (ω k − ω i,f )
f| ǫ̂∗k,λ · pe−ik·r a†k,λ |i
. (15.8)
me c ωk V

As can be seen from Eq. (7.2), the following holds


1
[H0 , r] = (−i) p , (15.9)
me
thus

Eyal Buks Quantum Mechanics - Lecture Notes 432


15.2. Transition Rates

4π2 e2 ωk

2

wi,f = δ (ω k − ω i,f )
f| ǫ̂∗k,λ · re−ik·r a†k,λ |i

V
4π2 e2 ωk
= δ (ω k − ω i,f ) |Mi,f |2 ,
V
(15.10)
where the atomic matrix element Mi,f is given by

Mi,f = η f | ǫ̂∗k,λ · re−ik·r |ηi  . (15.11)

15.2.2 Stimulated Emission and Absorption

The process of spontaneous emission of a photon in mode k, λ can be labeled


as (i,sk,λ ) → (f,sk,λ + 1), where sk,λ = 0. In the case of stimulated emission,
on the other hand, the initial photon occupation is assumed to be nonzero,
(e)
i.e. sk,λ ≥ 1. Let w(i,sk,λ )→(f,sk,λ +1),λ be the rate of emission of photons in
mode k, λ, given that the initial photon occupation number is sk,λ . With
the help of Eq. (14.66) the expression for the case of spontaneous emission
(15.10) can be easily generalized for arbitrary initial photon occupation sk,λ

(e) 4π2 e2 ω k (sk,λ + 1)


w(i,sk,λ )→(f,sk,λ +1),λ = δ (ωk − ωi,f ) |Mi,f |2 . (15.12)
V
Note that for the case of emission it is assumed that the energy of the
atomic
  i is larger than the energy of the atomic state f, i.e. ω i,f =
state
Eηi − Eηf / > 0.
(a)
Absorption is the reverse process. Let w(i,sk,λ )→(f,sk,λ −1),λ be the rate of
absorption of photons in mode k, λ, given that the initial photon occupation
number is sk,λ . With the help of Eq. (14.65) one finds using a derivation
similar to the one that was used above to obtain Eq. (15.12) that

(a) 4π 2 e2 ω k sk,λ
w(i,sk,λ )→(f,sk,λ −1),λ = δ (ω k + ω i,f ) |Mi,f |2 . (15.13)
V
Note that in this case it is assumed that ω i,f < 0.
The emission (15.12) and absorption (15.13) rates provide the contribu-
(e) (a)
tion of a single mode of the EM field. Let dΓ(i,s)→(f,s+1),λ /dΩ (dΓ(i,s)→(f,s−1),λ /dΩ)
be the total emitted (absorbed) rate in the infinitesimal solid angle dΩ having
polarization λ. For both cases s denotes the photon occupation number of the
initial state. To calculate these rates the contributions from all modes in the
EM field should be added. In the limit of large volume the discrete sum over
wave vectors k can be replaced by an integral according to Eq. (14.70). Note
that both the emitted and absorbed radiations are expected to be isotropic.
Thus, by using the relation ωk = ck, where k = |k|, one finds that

Eyal Buks Quantum Mechanics - Lecture Notes 433


Chapter 15. Light Matter Interaction

(e)
dΓ(i,s)→(f,s+1),λ  ∞
V (e)
= dk k2 w(i,sk,λ )→(f,sk,λ +1),λ
dΩ (2π)3 0
 ∞
e2 (s + 1) 2
= |Mi,f | dx x3 δ (x − ω i,f )
2πc3 0
αfs (s + 1) ω3i,f 2
= |Mi,f | .
2πc2
(15.14)
2
where αfs = e /c ≃ 1/137 is the fine-structure constant. In a similar way,
one finds for the case of absorption that
(a)
dΓ(i,s)→(f,s+1),λ αfs sω3i,f
= |Mi,f |2 . (15.15)
dΩ 2πc2

15.2.3 Selection Rules

While the size of an atom aatom is on the order of the Bohr’s radius a0 =
0.53×10−10 m (7.64), the energy difference Eηi −Eηf is expected to be on the
order of the ionization
 energy of hydrogen atom EI = 13.6 eV (7.66). Using
the relation ωk = Eηi − Eηf / = ck one finds that aatom k ≃ 10−3 . Thus,
to a good approximation the term e−ik·r in the expression for the matrix
element Mi,f can be replaced by unity

Mi,f ≃ η f | ǫ̂∗k,λ · r |ηi  . (15.16)

This approximation is called the dipole approximation.


The atomic energy eigenstates |η can be chosen to be also eigenvectors
of the angular momentum operators Lz and L2 . It is convenient to employ
the notation |k, l, m, σ to label these states, where k, l and m are orbital
quantum numbers and where σ labels the spin state. As can be seen from
Eqs. (7.42), (7.43) and (7.44) the following holds
H0 |k, l, m, σ = Ekl |k, l, m, σ , (15.17)
L2 |k, l, m, σ = l (l + 1) 2 |k, l, m, σ , (15.18)
Lz |k, l, m, σ = m |k, l, m, σ . (15.19)
Since it is assumed that no magnetic field is externally applied, the eigenen-
ergies Ekl are taken to be independent on the quantum numbers m and σ.
Radiation transitions between a pair of states |ki , li , mi , σ i  and |kf , lf , mf , σ f 
can occur only when the corresponding matrix element (15.16) is nonzero.
This requirement yields some conditions known as selection rules. The first
one refers to the spin quantum number σ. Note that Mi,f is a matrix element
of an orbital operator (15.16), and consequently it vanishes unless σf = σ i , or
alternatively, unless ∆σ = σf − σ i = 0. It is important to keep in mind that
this selection rule is valid only when spin-orbit interaction can be neglected.

Eyal Buks Quantum Mechanics - Lecture Notes 434


15.2. Transition Rates

Exercise 15.2.1. Show that the selection rule for the magnetic quantum
number m is given by

∆m = mf − mi ∈ {−1, 0, 1} . (15.20)

Solution 15.2.1. Using the relations Lz = xpy − ypx and [xi , pj ] = iδ ij
it is easy to show that [Lz , z] = 0 and [Lz , x ± iy] = ± (x ± iy). The first
relation together with Eq. (15.19) imply that
0 = kf , lf , mf , σf | [Lz , z] |ki , li , mi , σi 
=  (mf − mi ) kf , lf , mf , σf | z |ki , li , mi , σi  ,
(15.21)
whereas the second relation together with Eq. (15.19) imply that
kf , lf , mf , σ f | [Lz , x ± iy] |ki , li , mi , σ i 
=  (mf − mi ) kf , lf , mf , σf | x ± iy |ki , li , mi , σ i 
= ± kf , lf , mf , σf | (x ± iy) |ki , li , mi , σ i  ,
(15.22)
thus

 (mf − mi ∓ 1) kf , lf , mf , σf | x ± iy |ki , li , mi , σi  = 0 . (15.23)

Therefore Mi,f = 0 [see Eq. (15.16)] unless ∆m ∈ {−1, 0, 1}. The transition
∆m = 0 is associated with linear polarization in the z direction, whereas
the transitions ∆m = ±1 are associated with clockwise and counterclockwise
circular polarizations respectively.

Exercise 15.2.2. Show that the selection rule for the quantum number l is
given by

∆l = lf − li ∈ {−1, 1} . (15.24)

Solution 15.2.2. Using Eq. (15.40), which is given by


, 2 , 2 --  
L , L , r = 22 rL2 +L2 r , (15.25)

together with Eq. (15.18) yield


, , --
kf , lf , mf , σ f | L2 , L2 , r |ki , li , mi , σ i 
= 24 (lf (lf + 1) + li (li + 1)) kf , lf , mf , σ f | r |ki , li , mi , σ i 
= 4 (lf (lf + 1) − li (li + 1))2 kf , lf , mf , σf | r |ki , li , mi , σ i  ,
(15.26)
thus with the help of the identity

Eyal Buks Quantum Mechanics - Lecture Notes 435


Chapter 15. Light Matter Interaction

(lf (lf + 1) − li (li + 1))2 − 2 (lf (lf + 1) + li (li + 1))


3 4
2
= (li + lf ) (li + lf + 2) (li − lf ) − 1 ,
(15.27)
one finds that
3 4
(li + lf ) (li + lf + 2) (li − lf )2 − 1 kf , lf , mf , σf | r |ki , li , mi , σi  = 0 . (15.28)

Since both li and lf are non negative integers, and consequently li +lf +2 > 0,
one finds that kf , lf , mf , σf | r |ki , li , mi , σ i  can be nonzero only when li =
lf = 0 or |∆l| = 1. However, for the first possibility, for which li = mi = lf =
mf = 0, the wavefunctions of both states |ki , li , mi , σi  and |kf , lf , mf , σ f  is a
function of the radial coordinate r only [see Eq. (6.130)], and consequently
kf , lf , mf , σ f | r |ki , li , mi , σi  = 0. Therefore the selection rule is given by ∆l ∈
{−1, 1}.

15.3 Semiclassical Case


Consider the case where one mode of the EM field, which has angular fre-
quency ω and polarization vector ǫ̂, is externally driven to a coherent state
|α, where |α| ≫ 1. In the semiclassical approximation the annihilation op-
erator of the driven mode a is substituted by the complex constant α (and
the operator a† by α∗ ). Furthermore, all other modes are disregarded. When
ω is close to a specific transition frequency ω a = (E+ − E− ) / between two
atomic states, which are labeled by |+ and |−, the atom can be approxi-
mately considered to be a two level system. In the dipole approximation the
matrix element +| Hp |− is given by [see Eqs. (15.3), (15.9) and (15.16)]
  −iωt 
+| Hp |− = Ωe − Ω ∗ eiωt , (15.29)
2
where
0

Ω = −2ieω a αǫ̂ · +| r |− . (15.30)
ωV
It is convenient to express the complex frequency Ω as Ω = ω 1 e−iθ , where
both ω 1 and θ are real. The frequency ω1 is called the Rabi frequency. Due
to selection rules the diagonal matrix elements of Hp vanish.
The Schrödinger equation is given by
d
i |ψ = H |ψ , (15.31)
dt
where the matrix representation in the basis {|+ , |−} of the Hamiltonian
H is given by [see Eq. (15.29)]

Eyal Buks Quantum Mechanics - Lecture Notes 436


15.3. Semiclassical Case
  −i(ωt+θ) 

 i(ωt+θ)ωa −i(ωt+θ)  ω1 e − ei(ωt+θ)
H=
˙ . (15.32)
2 ω1 e −e −ω a

It is convenient to express the general solution as


   
iωt iωt
|ψ (t) = b+ (t) exp − |+ + b− (t) exp |− . (15.33)
2 2

Substituting into the Schrödinger equation yields [see Eq. (6.244)]


     
d b+ 1  ∆ω  ω 1 e−iθ − ei(2ωt+θ) b+
i = ,
dt b− 2 ω1 eiθ − e−i(2ωt+θ) −∆ω b−
(15.34)

where

∆ω = ω a − ω . (15.35)

In the rotating wave approximation the rapidly oscillating terms e±i(2ωt+θ)


are disregarded, since their influence in the long time limit is typically neg-
ligible. This approximation is equivalent to the assumption that the second
term in Eq. (15.29) can be disregarded. Furthermore, the phase factor θ can
be eliminated by resetting the time zero point accordingly. Thus, the Hamil-
tonian can be taken to be given by
 
 ω a ω1 e−iωt
H= ˙ , (15.36)
2 ω1 eiωt −ωa

and the equation of motion in the rotating frame can be taken to be given
by
    
d b+ 1 ∆ω ω1 b+
i = . (15.37)
dt b− 2 ω1 −∆ω b−

The time evolution is found using Eq. (6.138) [see also Eq. (6.248)]
 
b+ (t)
b− (t)
 
cos θ − i √∆ω sin θ
−i √ ω21 sin θ 2  
2
ω1 +(∆ω) 2
ω1 +(∆ω) b+ (0)
=  ,
−i √ ω21 sin θ 2 cos θ + i √∆ω 2
sin θ
2
b− (0)
ω1 +(∆ω) ω1 +(∆ω)

(15.38)
where
.
ω 21 + (∆ω)2 t
θ= . (15.39)
2

Eyal Buks Quantum Mechanics - Lecture Notes 437


Chapter 15. Light Matter Interaction

15.4 Problems
1. Show that
, 2 , 2 --  
L , L , r = 22 rL2 +L2 r . (15.40)

2. Consider an atom having a set of orthonormal energy eigenstates {|ηn }.


The oscillator strength fnm associated with the transition between state
|ηn  to state |ηm  is defined by
2me ω n,m 2
fnm = |ηf | r |η i | . (15.41)
3
Show that

fn,n′ = 1 . (15.42)
n′

3. Calculate the lifetime of all states of hydrogen atom having principle


quantum number n = 2.
4. Consider a hydrogen atom that is initially at time t → −∞ in its ground
state. An electric field in the z direction given by
τ2
E (t) = E0 ẑ , (15.43)
τ 2 + t2
where τ is a constant having the dimension of time, is externally applied.
Calculate the probability p2p to find the atom in the sub-shell 2p at time
t → ∞.
5. Consider a particle having mass m and charge q moving in a one dimen-
sional harmonic oscillator having angular resonance frequency ω. Cal-
culate using the dipole approximation the rate of spontaneous emission
from the number state |n to the ground state |0.
6. A hydrogen atom is initially in its ground state. An electric field given
by E0 cos (ωt), where both E0 and ω are constants, is externally applied.
Assume that ω > EI , where EI is the ionization energy of the atom.
Calculate the rate of ionization.

15.5 Solutions
1. Using the relations [Lx , z] = −iy, [Ly , z] = ix and [Lz , z] = 0 one finds
that
, 2 - , 2 - , 2 -
L , z = Lx , z + Ly , z
= i (−Lx y − yLx + Ly x + xLy )
= iV · ẑ ,
(15.44)

Eyal Buks Quantum Mechanics - Lecture Notes 438


15.5. Solutions
, -
where V = r × L − L × r. Thus the following holds L2 , r = iV. With
the help of the identities
[Lx , Vz ] = −Lx [Lx , y] − [Lx , y] Lx + [Lx , Ly ] x + x [Lx , Ly ] = −iVy ,
[Ly , Vz ] = − [Ly , Lx ] y − y [Ly , Lx ] + Ly [Ly , x] + [Ly , x] Ly = iVx ,
[Lz , Vz ] = − [Lz , Lx y] − [Lz , yLx ] + [Lz , Ly x] + [Lz , xLy ] = 0 ,
one finds that
, 2 , 2 -- , -
L , L , z = i L2 , Vz
= 2 (Lx Vy + Vy Lx − Ly Vx − Vx Ly )
= 2 (L × V − V × L) · ẑ ,
(15.45)
thus
, 2 , 2 --
L , L , r = 2 (L × V − V × L)
= 2 (L× (r × L) − L× (L × r) − (r × L) ×L+ (L × r) ×L)
 
= 22 rL2 +L2 r .
(15.46)
2. Trivial by the Thomas-Reiche-Kuhn sum rule (4.69).
(se)
3. The rate of spontaneous emission Γi→f,λ in general is given by Eq. (15.14)

(se) αfs ω 3i,f


Γi→f,λ = 4π |Mi,f |2 . (15.47)
2πc2
The transition frequency ω i,f between the levels n = 2 to n = 1 is given
by [see Eq. (7.66)]
 
me e4 1 1
ωi,f = − 2+ 2 , (15.48)
23 2 1

thus in terms of the Bohr’s radius a0 [see Eq. (7.64)] one has

(se) 33 α5fs me c2
Mi,f
2
Γi→f,λ =

28 
a0
.

The matrix element of z = r cos θ and of u± = 2−1/2 (x ± iy) =


2−1/2 r sin θe±iφ are given by [see Eq. (7.92)]
n′ , l′ , m′ | z |n, l, m
∞ 1 2π  ′ ∗
= dr r Rn′ l′ Rnl d (cos θ) dφ cos θ Ylm
3
′ Ylm ,
0 −1 0
(15.49)

Eyal Buks Quantum Mechanics - Lecture Notes 439


Chapter 15. Light Matter Interaction

and
n′ , l′ , m′ | u± |n, l, m
0 ∞ 1 2π  ′ ∗
1
= dr r Rn′ l′ Rnl d (cos θ) dφ sin θe±iφ Ylm
3
′ Ylm ,
2
0 −1 0
(15.50)
where n, l and m are the quantum members of hydrogen’s energy eigen-
vectors. The final state is taken to be the ground state (n, l, m) = (1, 0, 0).
In the dipole approximation the transition (2, 0, 0) → (1, 0, 0) is forbidden
due to the selection rule ∆l ∈ {−1, 1}. Using the identities
 3/2
1
R10 (r) = 2 e−r/a0 , (15.51)
a0
 3/2
1 r
R20 (r) = (2 − r/a0 ) e− 2a0 , (15.52)
2a0
 3/2
1 r r
R21 (r) = √ e− 2a0 , (15.53)
2a0 3a0
0
1
Y00 (θ, φ) = , (15.54)

0
1 3
Y1−1 (θ, φ) = sin θe−iφ , (15.55)
2 2π
0
1 3
Y10 (θ, φ) = cos θ , (15.56)
2 π
0
1 3
Y11 (θ, φ) = − sin θeiφ , (15.57)
2 2π
where a0 is Bohr’s radius [see Eq. (7.64)], one finds for the radial part
that
∞ √
3 27 6
dr r R10 R21 = a0 , (15.58)
35
0

and for the angular part that

Eyal Buks Quantum Mechanics - Lecture Notes 440


15.5. Solutions

1 2π
 ∗ 1
d (cos θ) dφ cos θ Y00 Y10 = √ , (15.59)
3
−1 0
1 2π
 ∗
d (cos θ) dφ cos θ Y00 Y1±1 = 0 , (15.60)
−1 0
1 2π
 ∗
d (cos θ) dφ sin θe±iφ Y00 Y10 = 0 , (15.61)
−1 0

and
1 2π
 ∗
d (cos θ) dφ sin θe−iφ Y00 Y1−1 = 0 , (15.62)
−1 0
0 1 2π 0
1 −iφ
 0 ∗ 1 1
d (cos θ) dφ sin θe Y0 Y1 = − , (15.63)
2 3
−1 0
0 1 2π 0
1 iφ
 0 ∗ −1 1
d (cos θ) dφ sin θe Y0 Y1 = , (15.64)
2 3
−1 0
1 2π
 ∗
d (cos θ) dφ sin θeiφ Y00 Y11 = 0 . (15.65)
−1 0

Thus, by combining all these results one finds that the inverse lifetime of
the states (2, 1, −1), (2, 1, 0) and (2, 1, 1) is given by

7
2
33 α5fs me c2
2 √

Γ (se) =


35 2
, (15.66)
28 

whereas the lifetime of the state (2, 0, 0) is infinite (in the dipole approx-
imation).
4. The probability p2pm to find the atom in the state |n = 2, l = 1, m is
calculated using Eq. (10.42) together with Eq. (7.84)


2
e2 E02 τ 2


/ 3E
′ i 4 I t′ τ
2
p2pm =
dt e
|2, 1, m| z |1, 0, 0| . (15.67)
2
−∞ 2 ′2
τ +t

where
µe4
EI = (15.68)
22
is the ionization energy. The following holds

Eyal Buks Quantum Mechanics - Lecture Notes 441


Chapter 15. Light Matter Interaction


/ 3EI ′ τ 1 ∞/ dxeix
dt′ ei 4 t =  
−∞ τ2 +t′2 Ω −∞ 1 + x 2


/ dxeix
=Ω ,
−∞ (x − iΩ) (x + iΩ)
(15.69)
where
3EI τ
Ω= , (15.70)
4
thus with the help of the residue theorem one finds that

/ 3EI ′ τ
dt′ ei 4 t = πe−Ω . (15.71)
−∞ τ2 + t′2

The matrix element 2, 1, m| z |1, 0, 0 is calculated with the help of Eq.
(15.49)
∞ 1 2π
2, 1, m| z |1, 0, 0 = 3
dr r R21 R10 d (cos θ) dφ cos θ (Y1m )∗ Y00
0 −1 0
√ 7
2 2a0
= δ m,0 ,
35
(15.72)
where
2
a0 = (15.73)
µe2
is the Bohr’s radius, thus
 2
215 eE0 a0 τ 3EI τ
p2pm = 10 π e− 2 δ m,0 . (15.74)
3 
5. The oscillator is assumed to move along the z direction. The rate of
(se)
spontaneous emission Γ|n→|0,λ with polarization λ into solid angle dΩ
is given in the dipole approximation by [see Eqs. (15.14) and (15.16)]

(se) q 2 (nω)3  2
dΓ|n→|0,λ = 3
|0| z |n|2 ǫ̂∗k,λ · ẑ dΩ , (15.75)
2πc
where ǫ̂∗k,λ is the polarization unit vector. With the help of Eqs. (5.11),
(5.28) and (5.29) one finds that

(se) q 2 (nω)3 n  2
dΓ|n→|0,λ = 3
δ n,1 ǫ̂∗k,λ · ẑ dΩ . (15.76)
2πc 2mω

Eyal Buks Quantum Mechanics - Lecture Notes 442


15.5. Solutions

Integrating over dΩ in spherical coordinates θ and φ with the help of the


relation
  π  1

dΩ cos2 θ = dφ d (cos θ) cos2 θ = , (15.77)
−π −1 3

and summing over the two orthogonal polarization yields the total rate
of spontaneous emission

(se) 2q 2 ω2
Γ|n→|0 = δ n,1 . (15.78)
3mc3
′ ′
6. The wave function of the final state |k′  has the form r′ |k′  = V −1/2 eik ·r ,
where V is the systems’s volume. The perturbation that is induced by
the applied electric field can be expressed as H1 (t) = Ke−iωt + K† eiωt ,
where
eE0 r · û
K= , (15.79)
2
r = r (sin θ cos φ, sin θ sin φ, cos θ) is the position vector operator and û =
(sin θ0 cos φ0 , sin θ0 sin φ0 , cos θ 0 ) is a unit vector in the direction of the
applied electric field. The matrix element Mk′ = k′ | K |n = 1, l = 0, m = 0
corresponding to the transition from the ground state |n = 1, l = 0, m = 0
[see Eq. (7.92)] to the final state |k′  is given by (the z axis is taken to
be in the direction of k′ )

−3/2 ∞ 1 2π
π−1/2 eE0 a0 V −1/2 2 ′
Mk′ = dr r d (cos θ) dφ e−ik r cos θ e−r/a0 r·û ,
2
0 −1 0
(15.80)

where

r · û = r sin θ sin θ0 cos (φ − φ0 ) + r cos θ cos θ0 , (15.81)

thus
∞ 1
−3/2 ′
1/2
Mk′ = π eE0 a0 V −1/2 cos θ0 dr e −r/a0 3
r d (cos θ) e−ik r cos θ cos θ
0 −1
  
′ ′
(
i eik r (k′ r+i)+e−ik r (k′ r−i) )
(k′ r)2

−3/2 16a40 k′ a0
= π1/2 eE0 a0 V −1/2 cos θ0  3 .
i (k′ a0 )2 + 1
(15.82)

Eyal Buks Quantum Mechanics - Lecture Notes 443


Chapter 15. Light Matter Interaction

The rate of ionization w is obtained by summing over k′ [see Eq. (10.39)]


2π 
w= δ (∆Ek′ − ω) |Mk′ |2 , (15.83)
 ′
k

where
2 k′2
∆Ek′ = + EI (15.84)
2me

is the change in the energy of the electron and where EI = me e4 /22 is


the ionization energy of the atom [see Eq. (7.66)]. Replacing the sum by
an integral according to (14.70) yields
∞  
256e2 E02 a30 ′ 2 k′2 (k′ a0 )4
w= dk δ + EI − ω  6 , (15.85)
3 2me
0 (k′ a0 )2 + 1

thus

256e2 me E02 a40 (k0 a0 )3


w=  6 , (15.86)
33
(k0 a0 )2 + 1

where

2me (ω − EI )
k0 = . (15.87)

Note that for a given amplitude E0 the rate w obtains its maximum
value, which is given by [see Eq. (7.64)]

27 3 E02 a30
wmax = , (15.88)
16 
when the angular frequency ω is chosen such that k0 a0 = 3−1/2 .

Eyal Buks Quantum Mechanics - Lecture Notes 444


16. Identical Particles

This chapter reviews the identical particles postulate of quantum mechanics


and second quantization formalism. It is mainly based on the first chapter of
Ref. [6].

16.1 Basis for the Hilbert Space


Consider a system containing some integer number N of identical particles.
For the single particle case, where N = 1, the state of the system |α can be
expanded using an orthonormal basis {|ai }i that spans the single particle
Hilbert space. Based on the single particle basis {|ai }i we wish to construct
a basis for the Hilbert space of the system for the general case, where N can
be any integer. This can be done in two different ways, depending on whether
the identical particles are considered to be distinguishable or indistinguishable
(see example in Fig. 16.1).
Suppose that the particles can be labelled by numbers as billiard balls.
In this approach the particles are considered as distinguishable. For this case
a basis for the Hilbert space of the many-particle system can be constructed
from all vectors having the form |1 : i1 , 2 : i2 , · · · , N : iN . The ket vector
|1 : i1 , 2 : i2 , · · · , N : iN  represents a state having N particles, where the
particle that is labelled by the number m (m = 1, 2, · · · , N ) is in the single
particle state |aim . Each ket vector |1 : i1 , 2 : i2 , · · · , N : iN  can be char-
acterized by a vector of occupation numbers n̄ = (n1 , n2 , · · · ), where ni is
the number of particles occupying the single particle state |ai . Let gn̄ be
the number of different ket-vectors having the form |1 : i1 , 2 : i2 , · · · , N : iN 
that are characterized by the same vector of occupation numbers n̄. It is easy
to show that
N!
gn̄ = > , (16.1)
ni !
i
<
where N = i ni is the number of particles.
Alternatively, the particles can be considered as indistinguishable. In this
approach all states having the same vector of occupation numbers n̄ rep-
resent the same physical state, and thus should be counted only once. In
Chapter 16. Identical Particles

other words, when the particles are considered as indistinguishable the sub-
space corresponding to any given vector of occupation numbers n̄ is rather
than being gn̄ - fold degenerate (as in the approach where the particles are
considered to be distinguishable) is taken to be nondegenerate. The identical
particle postulate of quantum mechanics states that identical particles should
be considered as indistinguishable. Consequently, a basis for the Hilbert space
of the many-particle system can be constructed from the set of ket vectors
{|n̄}n̄ . The ket vector |n̄ represents a state that is characterized by a vector
of occupation numbers n̄ = (n1 , n2 , · · · ), where the integer ni is the number
of particles that are in the single particle state |ai . Such a basis is considered
to be both orthonormal, i.e.
n̄1 |n̄2  = δ n̄1 ,n̄2 , (16.2)
where δ n̄1 ,n̄2 = 1 if n̄1 = n̄2 and δ n̄1 ,n̄2 = 0 otherwise, and complete

|n̄′  n̄′ | = 1 . (16.3)
n̄′

It it convenient to introduce the creation operators a†i . With analogy with


the case of a harmonic oscillator [see Eq. (5.32)] and the case of EM field [see
Eqs. (14.65) and (14.66)] the state |n̄ is expressed as
1  n1  n2
|n̄ = √ a† a†2 · · · |0 , (16.4)
n1 !n2 ! · · · 1
where |0 represents the state where all occupation numbers are zero. Equa-
tion (16.4) suggests that the creation operators a†i maps a given state to a
state having additional particle in the single particle quantum state |ai . The
operator a†i is the Hermitian conjugate of the annihilation operator ai . The
number operator Ni is defined by
Ni = a†i ai . (16.5)
In addition to the above discussed principle of indistinguishability, the
identical particle postulate of quantum mechanics also states that all particles
in nature are divided into two type: Bosons and Fermions. Moreover, while
for the case of Bosons, the creation and annihilation operators satisfy the
following commutation relations
3 4
[ai , aj ] = a†i , a†j = 0 , (16.6)
3 4
ai , a†j = δ ij , (16.7)
for the case of Fermions the following holds
3 4
[ai , aj ]+ = a†i , a†j = 0 , (16.8)
+
3 4

ai , aj = δ ij , (16.9)
+

Eyal Buks Quantum Mechanics - Lecture Notes 446


16.1. Basis for the Hilbert Space

(a)

2 …
1 3
state 1 state 2 state 3 state 4
n1=0 n2=2 n3=1 n4=0

1 …
3 2
state 1 state 2 state 3 state 4
n1=0 n2=2 n3=1 n4=0

3 …
2 1
state 1 state 2 state 3 state 4
n1=0 n2=2 n3=1 n4=0

(b)

state 1 state 2 state 3 state 4


n1=0 n2=2 n3=1 n4=0

<
Fig. 16.1. In this example the number of particles is N = i ni = 3, where the
occupation numbers are given by n̄ = (n1 , n2 , n3 , n4 , · · · ) = (0, 2, 1, 0, · · · ). When
the particles are considered as distinguishable
> [see panel (a)] the corresponding
subspace is gn̄ degenerate, where gn̄ = N!/ ni ! = 3. On the other hand, when
i
the particles are considered as indistinguishable [see panel (b)], the corresponding
subspace is nondegenerate.

3 4
where ai , a†j denotes anti-commutation, i.e.
+

[A, B]+ = AB + BA (16.10)


for general operators A and B.
Exercise 16.1.1. Show that for both Bosons and Fermions
[Ni , Nj ] = 0 . (16.11)

Eyal Buks Quantum Mechanics - Lecture Notes 447


Chapter 16. Identical Particles

Solution 16.1.1. For Bosons this result is trivial [see Eqs. (16.6) and (16.7)].
It is also trivial for Fermions when i = j. Finally, for Fermions when i = j
one has

Ni Nj = a†i ai a†j aj = −a†i a†j ai aj = a†i a†j aj ai = −a†j a†i aj ai = a†j aj a†i ai = Nj Ni .
(16.12)

16.2 Bosons

Based on Eqs. (16.2), (16.4), (16.6) and (16.7) a variety of results can be
obtained:

Exercise 16.2.1. Show that for Bosons


3  n 4  n−1
ai , a†i = n a†i . (16.13)

Solution 16.2.1. Trivial by Eq. (2.174), which states that for any operators
A and B

[A, B n ] = nB n−1 [A, B] , (16.14)

and by Eq. (16.7).

Exercise 16.2.2. Show that for Bosons

ai |0 = 0 . (16.15)

Solution 16.2.2. The norm of the vector ai |0 can be expressed with the
help of Eqs. (16.4) and (16.7)
3 4
0| a†i ai |0 = 0| a†i , ai + ai a†i |0
= − 0 |0 + 0, 0, · · · , ni = 1, 0, · · · |0, 0, · · · , ni = 1, 0, · · ·  ,
(16.16)
thus with the help of the normalization condition (16.2) one finds that
0| a†i ai |0 = 0 and therefore ai |0 = 0.

Exercise 16.2.3. Show that for Bosons

Ni |n̄ = ni |n̄ . (16.17)

Solution 16.2.3. With the help of Eqs. (16.4), (16.13) and (16.15) one finds
that

Eyal Buks Quantum Mechanics - Lecture Notes 448


16.3. Fermions

Ni |n̄ = a†i ai |n̄


1  n1  n2  ni
= √ a†1 a†2 · · · a†i ai a†i · · · |0
n1 !n2 ! · · ·
1  n1  n2 3  ni 4  ni 
= √ a†1 a†2 · · · a†i ai , a†i + a†i ai · · · |0
n1 !n2 ! · · ·
1  n1  n2  ni −1
= √ a†1 a†2 · · · a†i ni a†i · · · |0
n1 !n2 ! · · ·
= ni |n̄ .
(16.18)
Exercise 16.2.4. Show that for Bosons

ai |n1, n2 , · · · , ni , · · ·  = ni |n1, n2 , · · · , ni − 1, · · ·  , (16.19)

a†i |n1, n2 , · · · , ni , · · ·  = ni + 1 |n1, n2 , · · · , ni + 1, · · ·  . (16.20)
Solution 16.2.4. Equation (16.20) follows immediately from Eqs. (16.4) and
(16.6). Moreover, with the help of Eqs. (16.4), (16.13) and (16.15) one finds
that
1  n1  n2  ni
ai |n1, n2 , · · · , ni , · · ·  = √ a†1 a†2 · · · ai a†i · · · |0
n1 !n2 ! · · ·
1  n1  n2 3  ni 4  ni 
= √ a†1 a†2 · · · ai , a†i + a†i ai · · · |0
n1 !n2 ! · · ·
ni  n1  n2  ni −1
= √ a†1 a†2 · · · a†i · · · |0
n1 !n2 ! · · ·

= ni |n1, n2 , · · · , ni − 1, · · ·  .

16.3 Fermions
 2
The anti-commutation relations (16.8) for the case i = j yields a†i = 0. As
can be seen from Eq. (16.4), this implies that the only possible occupation
numbers ni are 0 and 1. This result is known as the Pauli’s exclusion principle,
according to which no more than one Fermion can occupy a given single
particle state. For Fermions Eq. (16.4) can be written as (recall that 0! =
1! = 1)
 n1  n2
|n̄ = a†1 a†2 · · · |0 , (16.21)

where ni ∈ {0, 1}.


Exercise 16.3.1. Show that for Fermions
ai |0 = 0 . (16.22)

Eyal Buks Quantum Mechanics - Lecture Notes 449


Chapter 16. Identical Particles

Solution 16.3.1. The norm of the vector ai |0 can be expressed with the
help of Eqs. (16.21) and (16.9)
3 4
0| a†i ai |0 = 0| a†i , ai − ai a†i |0
+
= 0 |0 − 0, 0, · · · , ni = 1, 0, · · · |0, 0, · · · , ni = 1, 0, · · ·  ,
(16.23)
thus with the help of the normalization condition (16.2) one finds that
0| a†i ai |0 = 0 and therefore ai |0 = 0.

Exercise 16.3.2. Show that for Fermions

Ni |n̄ = ni |n̄ , (16.24)

where Ni = a†i ai .

Solution 16.3.2. Using Eqs. (16.8), (16.9) and (16.21) one finds that
 n1  n2
Ni |n̄ = a†i ai a†1 a†2 · · · |0
<
2 nj  n1  n2  ni
= (−1) j<i a†1 a†2 · · · a†i ai a†i · · · |0
 n1  n2  ni
= a†1 a†2 · · · a†i ai a†i · · · |0 .
(16.25)
For the case ni = 0 this
 yields
ni [see
3 Eq. 4 (16.22)] Ni |n̄ = 0, whereas for the

case ni = 1 one has ai ai = ai , ai −a†i ai = 1−a†i ai , thus Ni |n̄ = |n̄.

+
Both cases are in agreement with Eq. (16.24).

Exercise 16.3.3. Show that for Fermions


<
nj
ai |n1, n2 , · · · , ni , · · ·  = ni (−1)j<i |n1, n2 , · · · , ni − 1, · · ·  , (16.26)
<
nj
a†i |n1, n2 , · · · , ni , · · ·  = (1 − ni ) (−1)j<i |n1, n2 , · · · , ni + 1, · · ·  (16.27)
.

Solution 16.3.3. According to Eq. (16.8) a†i a†j = −a†j a†i . For i = j this
 2
yields a†i = 0. These relations together with Eq. (16.21) leads to Eq.
(16.27) (note that 1 − ni = 1 if ni = 0 and 1 − ni =0 if ni = 1). Similarly,
ni
Eq. (16.26) is obtained by using the identity ai a†i = 1 − a†i ai and by
considering both possibilities ni = 0 and ni = 1.

Eyal Buks Quantum Mechanics - Lecture Notes 450


16.4. Changing the Basis

16.4 Changing the Basis


In the previous section the creation a†i and annihilation ai operators were
defined based on a given single particle orthonormal basis {|ai }i . Consider
an alternative single particle basis {|bj }j , which is made of eigenvectors of

the single particle observable BSP , i.e. the following holds BSP = BSP and

BSP |bj  = β j |bj  , (16.28)

where β j is the single particle eigenvalue corresponding to the eigenvector


|bj . Moreover, this basis is assumed to be both orthonormal, i.e.

bj |bj ′  = δ j,j ′ , (16.29)

and complete, i.e.



|bj  bj | = 1 . (16.30)
j

Exploiting the completeness of the original single particle orthonormal basis


{|ai }i , i.e. the fact that

|ai  ai | = 1 , (16.31)
i

allows expressing the eigenvector |bj  as



|bj  = ai |bj  |ai  . (16.32)
i

The single particle state |ai  can be expressed in the notation of many
particle states as a†i |0, whereas the single particle state |bj  can be expressed
as b†j |0, where the operator b†j , which is the creation operator of the single
particle state |bj , is given by [see Eq. (16.32)]

b†j = ai |bj  a†i . (16.33)
i

The creation operator b†j is the Hermitian conjugate of the annihilation op-
erator

bj = bj |ai  ai . (16.34)
i

An important example is the case where the single particle observable is


taken to be the position observable r. For this case Eq. (16.32) becomes
 ∗
|r′  = ψi (r′ ) |ai  , (16.35)
i

Eyal Buks Quantum Mechanics - Lecture Notes 451


Chapter 16. Identical Particles

where |r′  is a single particle position eigenvector, and where ψi (r′ ) = r′ |ai 
is the wavefunction of the single particle state |ai .
Expressing the single particle state |ai  in the notation of many particle
states as a†i |0 allows expressing the single particle state |r′  in the notation
of many particle states as Ψ † (r′ ) |0 [see Eq. (16.35)], where the operator
Ψ † (r′ ), which is given by
 ∗
Ψ † (r′ ) = ψi (r′ ) a†i , (16.36)
i

is the Hermitian conjugate of the quantized field operator Ψ (r′ ), which is


given by

Ψ (r′ ) = ψi (r′ ) ai . (16.37)
i

Note that while ψi (r′ ) is a wave function, Ψ (r′ ) is an operator on the Hilbert
space of the many particle system.
, -
Exercise 16.4.1. Calculate Ψ (r′ ) , Ψ † (r′′ ) ∓ , where [A, B]∓ = AB ∓ BA
for general operators A and B, and where the minus sign is used for Bosons
and the plus sign for Fermions.

Solution 16.4.1. With he help of Eqs. (16.7) and (16.9) one finds that
, -  3 4
Ψ (r′ ) , Ψ † (r′′ ) ∓ = ψi (r′ ) ψ∗i′ (r′′ ) ai , a†i′

i,i′

= ψi (r′ ) ψ∗i (r′′ )
i

= r′ |ai  ai |r′′ 
i
= r′ |r′′  ,
(16.38)
thus [see Eq. (3.66)]
, -
Ψ (r′ ) , Ψ † (r′′ ) ∓ = δ (r′ − r′′ ) . (16.39)

Similarly, one finds that


[Ψ (r′ ) , Ψ (r′′ )]∓ = 0 , (16.40)
, -
Ψ † (r′ ) , Ψ † (r′′ ) ∓ = 0 . (16.41)

Exercise 16.4.2. Show that



d3 r′ ρ (r′ ) = N , (16.42)

Eyal Buks Quantum Mechanics - Lecture Notes 452


16.5. Many Particle Observables

where

ρ (r′ ) = Ψ † (r′ ) Ψ (r′ ) , (16.43)

and where

N= Ni (16.44)
i

The operator ρ (r′ ) is called the number density operator, and the operator
N is called the total number of particles operator.

Solution 16.4.2. Using the definition of Ψ (r′ ) one finds that


  † 
d3 r′ Ψ † (r′ ) Ψ (r′ ) = ai′ ai d3 r′ ψ∗i′ (r′ ) ψi (r′ )
i,i′

= a†i ai
i
=N .
(16.45)

16.5 Many Particle Observables

Observables of a system of identical particles must be defined and must be


represented by Hermitian operators in a way that is consistent with the prin-
ciple of indistinguishability. Below we consider both, one-particle observables
and two-particle observables, and discuss their representation as operators
on the Hilbert space of the many-particle system.

16.5.1 One-Particle Observables

Consider a single particle observable such as the observable BSP , which was
introduced in the previous section [see Eqs. (16.28), (16.29), (16.30)]. It is
convenient to employ the single particle basis {|bj }j , which is made of single-
particle eigenvectors of BSP that satisfy BSP |bj  = β j |bj  [see Eq. (16.28)],
in order to construct creation b†j and annihilation bj operators. In the many-
particle case, the same physical variable that BSP represents for the single
particle case is represented by the operator B, which is given by

B= β j b†j bj . (16.46)
j

This can be seen by recalling that the operator b†j bj represents the number
of particles in the single particle state |bj  and that β j is the corresponding

Eyal Buks Quantum Mechanics - Lecture Notes 453


Chapter 16. Identical Particles

eigenvalue. With the help of Eqs. (16.28), (16.29), (16.30) and (16.33) (16.34)
the operator B can be expressed in terms of the operators a†i and ai

B= ai′ | BSP |ai  a†i′ ai . (16.47)
i,i′

16.5.2 Two-Particle Observables

Consider two-body interaction that is represented by an Hermitian operator


VTP on the Hilbert space of two-particle states. A basis for this Hilbert space
can be constructed using a given orthonormal basis for the single particle
Hilbert space {|bj }j . When the two particles are considered as distinguish-
able the basis of the Hilbert space of the two-particle states can be taken
to be {|j, j ′ }j,j′ . The ket vector |j, j ′  represents a state for which the first
particle is in single particle state |bj  and the second one is in state |bj ′ .
Assume the case where the single particle basis vectors |bj  are chosen in
such a way that diagonalizes VTP , i.e.
VTP |j, j ′  = vj,j ′ |j, j ′  , (16.48)
where the eigenvalue vj,j ′ is given by
vj,j ′ = j, j ′ | VTP |j, j ′  . (16.49)
In the many-particle case, the same physical variable that VTP represents
for the two-particle case is represented by the operator V , which is given by
1
V = vj,j ′ b†j b†j ′ bj′ bj . (16.50)
2 j,j′

To see that the above expression indeed represents the two particle interaction
consider the expectation value n̄| V |n̄ with respect to the many body state
|n̄ = |n1, n2 , · · · . The following holds [see Eqs. (16.6) , (16.7), (16.8) and
(16.9)]
b†j b†j ′ bj ′ bj = ±b†j b†j′ bj bj ′
3 4 
† † †
= ±bj bj′ , bj ± bj bj′ bj′

 3 4 
† † †
= ±bj ∓ bj , bj′ ± bj bj ′ bj ′

 
= ±bj ∓δ j,j ′ ± bj b†j ′ bj′

= −Nj δ j,j ′ + Nj Nj ′ ,
(16.51)
where the upper sign is used for Bosons and the lower one for Fermions. Thus
V can be rewritten as

Eyal Buks Quantum Mechanics - Lecture Notes 454


16.6. Hamiltonian

1
V = vj,j ′ Nj (Nj ′ − δ j,j′ ) . (16.52)
2 ′
j,j

Separating the terms for which j = j ′ from the terms for which j = j ′ yields
 1
V = vj,j′ Nj Nj ′ + vj,j Nj (Nj − 1) , (16.53)
2 j
j<j ′

thus the matrix element n̄| V |n̄ is given by


  nj (nj − 1)
n̄| V |n̄ = nj nj ′ vj,j ′ + vj,j . (16.54)
2
j<j ′ j

While the factor nj nj ′ represents the number of particle pairs occupying


single particle states j and j ′ for the case j = j ′ , the factor nj (nj − 1) /2
represents the number of particle pairs occupying the same single particle
states j. Thus the above expression for V (16.50) properly accounts for the
two-particle interaction.
With the help of Eqs. (16.29), (16.30) and (16.33) (16.34) the operator V
can be expressed in terms of the operators a†i and ai
1  
V = ai′ , ai′′ |j, j ′  j, j ′ | VTP |j, j ′  j ′ , j |ai′′′ , ai′′′′  a†i′ a†i′′ ai′′′ ai′′′′
2 i′ ,i′′ ,i′′′ ,i′′′′ j,j ′
1  
= ai′ , ai′′ |j, j ′  j, j ′ | VTP |j, j ′  j, j ′ |ai′′′′ , ai′′′  a†i′ a†i′′ ai′′′ ai′′′′
2 i′ ,i′′ ,i′′′ ,i′′′′ j,j ′
 
1  
= ai′ , ai′′ |  |j, j ′  j, j ′ | VTP |ai′′′′ , ai′′′  a†i′ a†i′′ ai′′′ ai′′′′ ,
2 i′ ,i′′ ,i′′′ ,i′′′′ ′
j,j

(16.55)
thus
1 
V = ai′ , ai′′ | VTP |ai′′′′ , ai′′′  a†i′ a†i′′ ai′′′ ai′′′′ . (16.56)
2 i′ ,i′′ ,i′′′ ,i′′′′

16.6 Hamiltonian
Consider the case where the single-particle Hamiltonian is given by

HSP = TSP + USP , (16.57)

where the operator TSP , which is given by


p2SP
TSP = , (16.58)
2m

Eyal Buks Quantum Mechanics - Lecture Notes 455


Chapter 16. Identical Particles

where pSP is the single-particle momentum vector operator and where m is


the mass of a particle, is the single-particle kinetic energy operator, and
where the operator USP (r′ ) the the single-particle potential energy. The
many-particle kinetic energy operator is found using Eq. (16.47)
1 
T = ai′ | p2SP |ai  a†i′ ai . (16.59)
2m i,i′

The matrix element ai′ | p2SP |ai  can be written using the wavefunctions
ψi (r′ ) = r′ |ai  [recall Eq. (3.29), according to which r′ | p |α = −i∇ψα
for a general state |α]

2
ai′ | p2SP |ai  = d3 r′ (∇ψ∗i′ (r′ )) · (∇ψi (r′ )) . (16.60)
2m

Thus, in terms of the quantized field operator Ψ (r′ ) [see Eqs. (16.36) and
(16.37)] the operator T can be expressed as

2
T = d3 r′ ∇Ψ † (r′ ) · ∇Ψ (r′ ) . (16.61)
2m
Integration by parts yields an alternative expression

2
T =− d3 r′ Ψ † (r′ ) ∇2 Ψ (r′ ) . (16.62)
2m
Similarly, the many-particle potential energy operator is found using Eq.
(16.47) [recall Eq. (3.23), according to which r′ | f (r) |α = f (r′ ) ψα (r′ ) for
a general state |α and for a general function f (r)]

U = ai′ | USP (r′ ) |ai  a†i′ ai
i,i′

= d3 r′ USP (r′ ) Ψ † (r′ ) Ψ (r′ ) .
(16.63)
In addition, consider the case where the particles interact with each other
via a two-particle potential VTP (r1 , r2 ). The corresponding many-particle
interaction operator is found using Eq. (16.56). The two-particle matrix ele-
ments of VTP are given by
ai′ , ai′′ | VTP |ai′′′′ , ai′′′ 
 
= d r 3 ′
d3 r′′ ψ∗i′ (r′ ) ψ∗i′′ (r′′ ) VTP (r′ , r′′ ) ψi′′′′ (r′ ) ψ i′′′ (r′′ ) ,
(16.64)
thus

Eyal Buks Quantum Mechanics - Lecture Notes 456


16.6. Hamiltonian
 
1
V = d3 r′ d3 r′′ VTP (r′ , r′′ ) Ψ † (r′ ) Ψ † (r′′ ) Ψ (r′′ ) Ψ (r′ ) . (16.65)
2
Combining all these results yields the total many-particle Hamiltonian

2
H= d3 r′ ∇Ψ † (r′ ) · ∇Ψ (r′ )
2m

+ d3 r′ USP (r′ ) Ψ † (r′ ) Ψ (r′ )
 
1
+ d3 r′ d3 r′′ VTP (r′ , r′′ ) Ψ † (r′ ) Ψ † (r′′ ) Ψ (r′′ ) Ψ (r′ ) .
2
(16.66)

Exercise 16.6.1. Show that the Heisenberg equation of motion for the field
operator Ψ (r′ ) is given by
d
i Ψ (r′ , t)
 dt 
2 2
= − ∇ + USP (r′ ) Ψ (r′ , t)
2m

+ d3 r′′ VTP (r′ , r′′ ) Ψ † (r′′ , t) Ψ (r′′ , t) Ψ (r′ , t) .
(16.67)
Note that in the absence of two-particle interaction the above equation for
the field operator Ψ (r′ , t) is identical to the single-particle Schrödinger equa-
tion for the single particle wavefunction ψ (r′ ). Due to this similarity the
many-particle formalism of quantum mechanics is sometimes called second
quantization.

Solution 16.6.1. The Heisenberg equation of motion [see Eq. (4.37)] is given
by

i = − [H, Ψ ]− . (16.68)
dt
For general operators A, B and C the following holds
[AB, C]− = A [B, C]± ∓ [A, C]± B
= A [B, C]± − [C, A]± B .
(16.69)
Below we employ this relation for evaluating commutation relations. For
Fermions the upper sign (anti-commutation) is chosen, whereas for Bosons
the lower one is chosen (commutation). With the help of Eqs. (16.39), (16.40)
and (16.41) one finds (for both Bosons and for Fermions) that

Eyal Buks Quantum Mechanics - Lecture Notes 457


Chapter 16. Identical Particles

2 , -
[T, Ψ (r′ )]− = − d3 r′′ Ψ † (r′′ ) ∇2 Ψ (r′′ ) , Ψ (r′ ) −
2m

2
= d3 r′′ δ (r′ − r′′ ) ∇2 Ψ (r′′ )
2m
2 2
= ∇ Ψ (r′ ) ,
2m
(16.70)
and


, -
[U, Ψ (r )]− = d3 r′′ USP (r′′ ) Ψ † (r′′ ) Ψ (r′′ ) , Ψ (r′ ) −

= − d3 r′′ USP (r′′ ) δ (r′ − r′′ ) Ψ (r′′ ) (16.71)

= −USP (r′ ) Ψ (r′ ) .


(16.72)
Similarly
[V, Ψ (r′ )]−
 
1 3 ′′
, -
= d r d3 r′′′ VTP (r′′ , r′′′ ) Ψ † (r′′ ) Ψ † (r′′′ ) Ψ (r′′′ ) Ψ (r′′ ) , Ψ (r′ ) −
2
 
1 , -
= d3 r′′ d3 r′′′ VTP (r′′ , r′′′ ) Ψ † (r′′ ) Ψ † (r′′′ ) , Ψ (r′ ) − Ψ (r′′′ ) Ψ (r′′ )
2
 
1 3 ′′
=− d r d3 r′′′ VTP (r′′ , r′′′ ) Ψ † (r′′ ) δ (r′ − r′′′ ) Ψ (r′′′ ) Ψ (r′′ )
2
 
1
− d3 r′′ d3 r′′′ VTP (r′′ , r′′′ ) δ (r′ − r′′ ) Ψ † (r′′′ ) Ψ (r′′′ ) Ψ (r′′ )
2

= − d3 r′′ VTP (r′ , r′′ ) Ψ † (r′′ ) Ψ (r′′ ) Ψ (r′ )
(16.73)
where in the last step it was assumed that VTP (r′′ , r′ ) = VTP (r′ , r′′ ). Com-
bining these results lead to Eq. (16.67).

16.7 Momentum Representation

In the momentum representation the Hamiltonian is constructed using a


single-particle basis made of momentum eigenvectors |p′ . The wavefunc-
′ ′
tions of these single-particle states are proportional to eik ·r [see Eq. (3.75)],
where
p′
k′ = . (16.74)


Eyal Buks Quantum Mechanics - Lecture Notes 458


16.7. Momentum Representation

These wavefunctions can be normalized when the volume of the system is


taken to be finite. For simplicity, consider the case where the particles are
confined within a volume V = L3 having a cubic shape. The normalized wave
functions are taken to be given by
1 ′ ′
r′ |k′  = ψk′ (r′ ) = √ eik ·r , (16.75)
V
where |k′  labels a momentum eigenvector having an eigenvalue k′ . The
requirement that the wavefunctions ψk′ (r′ ) satisfy periodic boundary con-
ditions, i.e. ψk (r′ ) = ψk (r′ + Lx̂) = ψk (r′ + Lŷ) = ψk (r′ + Lẑ), yields a
discrete set of allowed values of the wave vector k

k= (nx , ny , nz ) , (16.76)
L
where nx , ny and nz are all integers. The orthonormality condition reads
 
1 ′ ′′ ′
d3 r′ ψ∗k′′ (r′ ) ψk′ (r′ ) = d3 r′ ei(k −k )·r
V V V
= δ k′ ,k′′ .
(16.77)
In the momentum representation the many-particle kinetic energy T is
given by [see Eq. (16.59)]
1  ′′ 2
T = k | pSP |k′  a†k′′ ak′
2m ′ ′′
k ,k

2  ′2 †
= k ak′ ak′ ,
2m ′
k
(16.78)
the many-particle potential energy U is given by [see Eq. (16.63)]

U= Uk′ −k′′ a†k′′ ak′ , (16.79)
k′ ,k′′

where
Uk′ −k′′ = k′′ | USP (r′ ) |k′ 

1 ′ ′′ ′
= d3 r′ USP (r′ ) ei(k −k )·r ,
V V
(16.80)
and the many-particle interaction operator is given by [see Eq. (16.56)]
1 
V = k′ , k′′ | VTP |k′′′′ , k′′′  a†k′ a†k′′ ak′′′ ak′′′′ , (16.81)
2
k′ ,k′′ ,k′′′ ,k′′′′

Eyal Buks Quantum Mechanics - Lecture Notes 459


Chapter 16. Identical Particles

where
 
1 ′′′′
−k′ )·r′ i(k′′′ −k′′ )·r′′
′ ′′ ′′′′
k , k | VTP |k , k  = 2 ′′′ 3 ′
d r d3 r′′ VTP (r′ , r′′ ) ei(k e .
V V V
(16.82)

The assumption that VTP (r′ , r′′ ) is a function of the relative coordinate r =
r′ − r′′ only, together with the coordinates transformation
r′ + r′′
r0 = , (16.83)
2
r = r′ − r′′ , (16.84)
yields (note that r′ = r0 + r/2 and r′′ = r0 − r/2)
k′ , k′′ | VTP |k′′′′ , k′′′ 
  i(k′′′′ −k′ −k′′′ +k′′ )·r
1 ′′′′ ′ ′′′ ′′
= 2 d3 r0 ei(k −k +k −k )·r0 d3 r VTP (r0 + r/2, r0 − r/2) e 2
V V V
 i(k′′′′ −k′ −k′′′ +k′′ )·r
1
= δ k +k ,k +k
′ ′′ ′′′ ′′′′ d3 r vTP (r) e 2 ,
V V
(16.85)
where

vTP (r) = VTP (r0 + r/2, r0 − r/2) . (16.86)

Thus the only allowed processes for this case are those for which the total
momentum is conserved, i.e. k′ + k′′ = k′′′ + k′′′′ . Using the notation

q = k′′ − k′′′ = k′′′′ − k′ , (16.87)

one can express V as


1 
V = vq a†k′ a†k′′ ak′′ −q ak′ +q , (16.88)
2 ′ ′′
k ,k ,q

where

1 iq·r
vq = d3 r vTP (r) e 2 . (16.89)
V V

16.8 Spin

In addition to spatial (orbital) degrees of freedom, the particles may have


spin. We demonstrate below the inclusion of spin for the case of momentum
representation. The basis for single-particle states is taken to be {|k′ , σ′ }k′ ,σ′ ,

Eyal Buks Quantum Mechanics - Lecture Notes 460


16.9. The Electron Gas

where the quantum number σ indicates the spin state. The single-particle
orthonormality condition reads

k′′ , σ′′ |k′ , σ ′  = δ k′ ,k′′ δ σ′ ,σ′′ . (16.90)

The commutation (for Bosons) and anti-commutation (for Fermions) rela-


tions [see Eqs. (16.6), (16.7), (16.8) and (16.9)] become
3 4
[ak′ ,σ′ , ak′′ ,σ′′ ]± = a†k′ ,σ′ , a†k′′ ,σ′′ =0, (16.91)
±
3 4
ak′ ,σ′ , a†k′′ ,σ′′ = δ k′ ,k′′ δ σ′ ,σ′′ , (16.92)
±

For the example above, the Hamiltonian becomes


2  ′2 † 
H= k ak′ ,σ′ ak′ ,σ′ + Uk′ −k′′ a†k′′ ,σ′ ak′ ,σ′
2m ′ ′ ′ ′′ ′
k ,σ k ,k ,σ
1 
+ vq a†k′ ,σ′ a†k′′ ,σ′′ ak′′ −q,σ′′ ak′ +q,σ′ .
2
k′ ,k′′ ,q,σ ′ ,σ ′′
(16.93)

16.9 The Electron Gas

Consider a free (i.e. noninteracting) gas of N ≫ 1 electrons occupying volume


V. The Hamiltonian is given by [see Eq. (16.93)]

2  ′2 †
H= k ak′ ,σ′ ak′ ,σ′ . (16.94)
2m ′ ′
k ,σ

In the momentum representation the single particle state |k′ , σ has a wave-
function given by [see Eq. (16.75)]
1 ′ ′
r′ |k′ , σ = √ eik ·r , (16.95)
V
and thus the quantized field operator Ψσ (r′ ) is given by [see Eq. (16.37)]
1  ik′ ·r′
Ψσ (r′ ) = √ e ak′ ,σ . (16.96)
V k′

The single particle state |k′ , σ has energy given by

2 k′2
ǫk′ = , (16.97)
2m
where m is the electron mass [see Eq. (16.94)].

Eyal Buks Quantum Mechanics - Lecture Notes 461


Chapter 16. Identical Particles

The allowed values of k′ are determined by boundary conditions. Consider


for simplicity the case where the gas is confined in a cube (having edge length
of V 1/3 ). Imposing periodic boundary conditions on the wavefunction of the
single particle states |k′ , σ leads to the requirement (16.76). Thus, the density
of states per spin in k′ space is V/8π 3 .
In the ground state |ϕ0  all single particle states for which |k′ | ≤ kF are
singly occupied, whereas all single particle states for which |k′ | > kF remain
empty, i.e.
#
|ϕ0  = a†k′ ,σ′ |0 . (16.98)
|k′ |≤kF , σ ′

The Fermi wave vector is chosen such that the number of single particle states
for which |k′ | ≤ kF is N. Since the density of states per spin in k′ space is
V/8π3 one finds that
V 4 3
2 πk = N , (16.99)
8π3 3 F
thus
3π 2 N
kF3 = . (16.100)
V
The Fermi energy ǫF is the corresponding energy
2 kF2
ǫF = . (16.101)
2m
The density of states D (ǫ) per spin and per unit volume is given by
1
D (ǫ) = δ (ǫ − ǫk′ ) . (16.102)
V ′
k

where ǫk′ is given by Eq. (16.97). By replacing the sum by an integral one
finds that
 
1 2 k′2
D (ǫ) = δ ǫ−
V ′ 2m
k
∞  
1 V 2 k′2
= 4π dk′ k′2 δ ǫ −
V 8π3 2m
0
 3/2 ∞
1 2m √
= 2 dǫ′ ǫ′ δ (ǫ − ǫ′ )
4π 2
0
m √
= 2 3 2mǫ .
2π 
(16.103)

Eyal Buks Quantum Mechanics - Lecture Notes 462


16.10. Problems

The ground state energy is given by


ǫF 5/2
23/2 m3/2 VǫF
E0 = 2V dǫ′ D (ǫ′ ) ǫ′ = , (16.104)
5π2 3
0

or [see Eq. (16.100) and (16.101)]

3N 2 kF2
E0 = . (16.105)
5 2m

16.10 Problems

1. Find the many-particle interaction operator V for the case where the
two-particle potential is a constant VTP (r1 , r2 ) = V0 .
2. The same for the Coulomb interaction
e2
VTP (r1 , r2 ) = . (16.106)
|r1 − r2 |
3. Show that

+ ∇J = 0 , (16.107)
dt
where ρ (r′ ) = Ψ † (r′ ) Ψ (r′ ) is the the number density operator [see Eq.
(16.43)] and where the current density operator J is given by
 , † ′   -
J (r′ ) = Ψ (r ) ∇Ψ (r′ ) − ∇Ψ † (r′ ) Ψ (r′ ) . (16.108)
2im
4. Consider two identical Bosons having mass m in a one dimensional po-
tential U (x) well given by
+
0 if 0 ≤ x ≤ L
U (x) = . (16.109)
∞ else

The particles interact with each other via a two-particle interaction given
by VTP = −V0 Lδ (x1 − x2 ), where V0 is a constant. Calculate the ground
state energy to lowest nonvanishing order in V0 .
5. By definition, an ideal gas is an ensemble of non-interacting identical
particles. The set of single particle eigenenergies is denoted by {εi }. Cal-
culate the average energy H and the average number of particles N  in
thermal equilibrium as a function of the temperature T and the chemical
potential µ for the case of
a) Fermions.
b) Bosons.

Eyal Buks Quantum Mechanics - Lecture Notes 463


Chapter 16. Identical Particles

6. Bogoliubov transformation - Consider the transformation


bk = uk ak + vk a†−k , (16.110)

where ak (a†−k ) is the annihilation (creation) operator corresponding to


the single particle state |k (|−k), and where uk and vk are real coeffi-
cients. The state |Vb  is defined by the condition
bk |Vb  = 0 . (16.111)
a) For the case of Fermions, under what conditions the operators bk
and b†k can be considered as annihilation and creation operators?
Evaluate the expectation value Vb | a†k ak |Vb .
b) The same for Bosons.
7. Find the eigenenergies of the Hamiltonian
 3  4
H= ǫk′ a†k′ ak′ + λ ak′ + a†k′ , (16.112)
k′

where ak′ and a†k′ are Boson annihilation and creation operators corre-
sponding to the single particle state |k′ , and where ǫk′ and λ are real
coefficients.
8. Consider a system of identical spinless Bosons, whose Hamiltonian is
given by
 ξ ′  † † 
H= ǫk′ a†k′ ak′ + k
ak′ a−k′ + ak′ a−k′ , (16.113)
′ ′
2
k k

where summation is over momentum single particle states having wave


vector k′ , and ak′ and a†k′ are annihilation and creation operators, re-
spectively. The coefficients ǫk′ and ξ k′ are assumed to be even functions
of k′ , i.e. ǫ−k′ = ǫk′ and ξ −k′ = ξ k′ . Calculate the eigenenergies of H.
9. Find eigenvectors and eigenvalues of the quantized field operator Ψ (r)
for the case of Bosons. Evaluate the expectation values with respect to
the number operator N and with respect with the Hamiltonian of the
many body system (with one-particle and two-particle interactions).
10. Consider a neutral helium atom having 2 electrons and a nucleus having
charge +2e. Calculate the energy of the ground state. Assume that the
Coulomb interaction between the electrons can be considered as small
and calculate the energy correction due to this interaction to lowest non-
vanishing order in perturbation theory. Ignore spin—orbit coupling and
hyperfine interaction.
11. Consider the state
 

|γ = dr dr′′ F (r′ , r′′ ) Ψ † (r′ ) Ψ † (r′′ ) |0 , (16.114)

where Ψ (r′ ) is the Bosonic quantized field operator, |0 represents the
state where all occupation numbers are zero, and F (r′ , r′′ ) is complex.

Eyal Buks Quantum Mechanics - Lecture Notes 464


16.11. Solutions

a) Find a condition that the function F (r′ , r′′ ) must satisfy in order to
ensure that the state |γ is normalized.
b) Consider the case where F (r′ , r′′ ) can be expressed as F (r′ , r′′ ) =
Af1 (r′ ) f2 (r′′ ), where A is a normalization constant (which is chosen
such that γ |γ = 1) and where both functions f1 () and f2 () are
normalized according to
 
′ 2 2
1 = dr |f1 (r )| = dr′ |f2 (r′ )| .

Evaluate the function


g (r′ ) = γ| ρ (r′ ) |γ , (16.115)
where ρ (r′ ) = Ψ † (r′ ) Ψ (r′ ).
c) Calculate the total number of particles

Nγ = dr′ g (r′ ) . (16.116)

12. Consider a free (i.e. noninteracting) gas of N ≫ 1 electrons occupying


volume V. Calculate the correlation function
Cσ (r′ − r′′ ) = ϕ0 | Ψσ† (r′ ) Ψσ (r′′ ) |ϕ0  , (16.117)
where |ϕ0  is the ground state of the N electrons gas, Ψσ (r) is the quan-
tized field operator and σ stands for a spin state.
13. Calculate the ground state energy of electron gas containing N ≫ 1
electrons filling a volume V. Consider the Coulomb interaction between
electrons as weak and calculate the energy shift due to this interaction to
lowest non-vanishing order in perturbation theory. Assume that the vol-
ume V contains a uniform background of positive charge density +eN/V
(without the positive background the system is expected to be unstable
due to the repulsive nature of the Coulomb interaction).

16.11 Solutions
1. In general V is given by Eq. (16.88) where for this case
vq = V0 δ q,0 , (16.118)
thus
V0  † †
V = a ′ a ′′ ak′′ ak′
2 ′ ′′ k k
k ,k
 
V0  † 3 † 4

= a ′ ak′′ ak′′ , ak′ + ak′ ak′′ ak′′ .
2 ′ ′′ k −
k ,k
(16.119)

Eyal Buks Quantum Mechanics - Lecture Notes 465


Chapter 16. Identical Particles

With the help of Eq. (16.69) one finds that [see also Eqs. (16.6), (16.7),
(16.8) and (16.9)]
3 4 3 4
a†k′′ ak′′ , ak′ = a†k′′ [ak′′ , ak′ ]± − ak′ , a†k′′ ak′′
− ±
= −δ k′ ,k′′ ak′ ,
(16.120)
[for Fermions the upper sign (anti-commutation) is taken, whereas for
Bosons the lower one is taken (commutation)], thus
N (N − 1)
V = V0 , (16.121)
2
where N is the total number of particles operator. Note that N (N − 1) /2
is the number of interacting pairs in the system.
2. For this case the Fourier transform f (q) of the function 1/ |r| is needed

1
= d3 q f (q) eiq·r . (16.122)
|r|
Applying the Laplace operator ∇2 and using the identity
1
∇2 = −4πδ (r) (16.123)
|r|
yield

−4πδ (r) = − d3 q f (q) |q|2 eiq·r , (16.124)

thus with the help of the identity


∞
dk eikx = 2πδ (x) , (16.125)
−∞

one finds that


1
f (q) = , (16.126)
2π2 q 2
where q = |q|, and therefore

1 1 1
= 2 d3 q 2 eiq·r (16.127)
|r| 2π q
With the help of this result one finds that V is given by [see Eqs. (16.88)
and (16.89)]
1  4πe2 † †
V = a ′ a ′′ ak′′ −q ak′ +q . (16.128)
2V ′ ′′ q2 k k
k ,k ,q

Eyal Buks Quantum Mechanics - Lecture Notes 466


16.11. Solutions

3. With the help of Eq. (16.67) and its Hermitian conjugate one finds that
dρ dΨ † (r′ ) dΨ (r′ )
= Ψ (r′ ) + Ψ † (r′ )
dt dt dt
1 2 , † ′   -
=− Ψ (r ) ∇ Ψ (r′ , t) − ∇2 Ψ † (r′ , t) Ψ (r′ ) ,
2
i 2m
(16.129)

where the assumptions USP (r′ ) = USP (r′ ) and VTP

(r′ , r′′ ) = VTP (r′ , r′′ )
have been made, thus

+ ∇J = 0 . (16.130)
dt
Note the similarity between this result and the continuity equation that
is satisfied by a single-particle wavefunction [see Eq. (4.73)].
4. For the unperturbed case, i.e. when V0 = 0, the single-particle wavefunc-
tions of the normalized eigenstates are given by
0
2 jπx
ψj (x) = sin , (16.131)
L L
where j = 1, 2, · · · , and the corresponding single-particle eigenenergies
are
2 π2 j 2
εj = . (16.132)
2mL2
For this case the ground state is the many-particle state |GS = |n1 = 2, n2 = 0, n3 = 0, · · · ,
i.e. the state for which both particles are in the j = 1 single-particle state.
In perturbation theory to first order in V0 the energy of this state is given
by [see Eq. (9.32)]
 
E = 2ε1 + GS| V |GS + O V02 , (16.133)

where the many-particle interaction operator V is given by Eq. (16.56).


The matrix element GS| V |GS is given by
1
GS| V |GS = 1, 1| VTP |1, 1 GS| a†1 a†1 a1 a1 |GS
2
1  3 4
= 1, 1| VTP |1, 1 GS| a†1 a1 a†1 − a1 , a†1 a1 |GS
2
N1 (N1 − 1)
= 1, 1| VTP |1, 1 GS| |GS
2
= 1, 1| VTP |1, 1 ,
(16.134)
where the two-particle matrix element 1, 1| VTP |1, 1 is given by

Eyal Buks Quantum Mechanics - Lecture Notes 467


Chapter 16. Identical Particles
 L  L
1, 1| VTP |1, 1 = dx1 dx2 ψ1 (x1 ) ψ1 (x2 ) VTP (x1 , x2 ) ψ1 (x1 ) ψ1 (x2 )
0 0
 L
= −V0 L dx1 ψ41 (x1 )
0
3
= − V0 ,
2
(16.135)
thus
2 π 2 3  
E= 2
− V0 + O V02 . (16.136)
mL 2
5. The grandcanonical partition function [see Eq. (8.450)] is evaluated by
summing over all many-particle states
 
Zgc = Tr e−βH+βµN

= n1, n2 , · · · , ni , · · · | e−βH+βµN |n1, n2 , · · · , ni , · · ·  ,
n1, n2 ,···

(16.137)
where

H= εi a†i ai , (16.138)
i

N= a†i ai , (16.139)
i
and β = 1/kB T , thus one finds that
#
Zgc = e−βni (εi −µ) . (16.140)
i ni

and
! "
 
−βni (εi −µ)
log Zgc = log e . (16.141)
i ni

a) In this case the summation over ni includes only two terms ni = 0


and ni = 1, thus
  
log Zgc = log 1 + e−β(εi −µ) . (16.142)
i

The average energy


 is found
 using  Eq. (8.451)

∂ log Zgc µ ∂ log Zgc
H = − +
∂β µ β ∂µ β
 εi e−β(εi −µ)
= ,
i
1 + e−β(εi −µ)
(16.143)

Eyal Buks Quantum Mechanics - Lecture Notes 468


16.11. Solutions

whereas the average number of particles is found using Eq. (8.454)

∂ log Zgc  e−β(εi −µ)


N  = λ = , (16.144)
∂λ i
1 + e−β(εi −µ)

In terms of the Fermi-Dirac function fFD (ε), which is given by


1
fFD (ε) = , (16.145)
exp [β (ε − µ)] + 1
these results can be rewritten as

H = εi fFD (εi ) , (16.146)
i

and

N  = fFD (εi ) . (16.147)
i

b) In this case the summation over ni includes all integers ni =


0, 1, 2, · · · , thus
  
1
log Zgc = log . (16.148)
i
1 − e−β(εi −µ)

The average energy is found using Eq. (8.451)



H = εi fBE (εi ) ,
i

whereas the average number of particles is found using Eq. (8.454)



N  = fBE (εi ) , (16.149)
i

where
1
fBE (ε) = (16.150)
exp [β (ε − µ)] − 1
is the Bose-Einstein function .
6. The operators ak and a†k satisfy [see Eqs. (16.6), (16.7), (16.8) and (16.9)]
3 4
[ak′ , ak′′ ]± = a†k′ , a†k′′ =0, (16.151)
±
3 4
ak′ , a†k′′ = δ k′ ,k′′ . (16.152)
±

Similarly, The operators bk and b†k can be considered as annihilation and


creation operators provided that they satisfy

Eyal Buks Quantum Mechanics - Lecture Notes 469


Chapter 16. Identical Particles
3 4
[bk′ , bk′′ ]± = b†k′ , b†k′′ =0, (16.153)
±
3 4
bk′ , b†k′′ = δ k′ ,k′′ . (16.154)
±

Using the definition (16.110) together with Eqs. (16.6) and (16.8) these
conditions become
3 4 3 4
vk′ uk′′ a†−k′ , ak′′ + uk′ vk′′ ak′ , a†−k′′ =0, (16.155)
± ±
3 4 3 4
vk′ uk′′ a−k′ , a†k′′ + uk′ vk′′ a†k′ , a−k′′ =0, (16.156)
± ±
3 4 3 4
uk′ uk′′ ak′ , a†k′′ + vk′ vk′′ a†−k′ , a−k′′ = δ k′ ,k′′ . (16.157)
± ±
Note that by inverting the transformation between the operators ak , a−k ,
a†k and a†−k and the operators bk , b−k , b†k and b†−k , which can be expressed
in matrix form as [see Eq. (16.110)]
    
bk uk 0 0 v k ak
 b−k   0 u−k v−k 0   a−k 
 † =  
 bk   0 vk uk 0   a†k  , (16.158)
b†−k v−k 0 0 u−k a†−k
one finds that
    
ak u−k 0 0 −vk bk
 a−k 
  b−k
1  0 uk −v−k 0   
 † =  †  .
 ak  uk u−k − vk v−k  0 −vk u−k 0   bk 
a†−k −v−k 0 0 uk b†−k
(16.159)
This result together with Eq. (16.111) imply that the expectation value
Vb | a†k ak |Vb  is given by
 2
vk
Vb | a†k ak |Vb  = Vb | b−k b†−k |Vb  , (16.160)
uk u−k − vk v−k
thus for both Bosons and Fermions [see Eq. (16.154)]
 2
† vk
Vb | ak ak |Vb  = . (16.161)
uk u−k − vk v−k
a) For the case of Fermions one finds using Eq. (16.9) that the condi-
tions (16.155), (16.156) and (16.157) become (recall that [A, B]+ =
[B, A]+ )
(vk′ uk′′ + uk′ vk′′ ) δ k′ ,−k′′ = 0 , (16.162)
(vk′ uk′′ + uk′ vk′′ ) δ k′ ,−k′′ = 0 , (16.163)
(uk′ uk′′ + vk′ vk′′ ) δ k′ ,k′′ = δ k′ ,k′′ , (16.164)

Eyal Buks Quantum Mechanics - Lecture Notes 470


16.11. Solutions

thus
vk u−k + uk v−k = 0 , (16.165)
2 2
uk + vk = 1 . (16.166)
These conditions are guarantied to be satisfied provided uk and vk
are expressed using a single real parameter θ k as
uk = cos θk , vk = sin θk , (16.167)
u−k = cos θk , v−k = − sin θ k . (16.168)
For this case Eq. (16.159) becomes
    
ak cos θ k 0 0 − sin θk bk
 a−k   0 cos θk sin θk 0   b−k 
 † =   †  , (16.169)
 ak   0 − sin θ k cos θ k 0   bk 

a−k sin θk 0 0 cos θk b†−k

and Eq. (16.161) becomes

Vb | a†k ak |Vb  = sin2 θk . (16.170)

b) For the case of Bosons one finds using Eq. (16.7) that the condi-
tions (16.155), (16.156) and (16.157) become (recall that [A, B]− =
− [B, A]− )
−vk u−k + uk v−k = 0 , (16.171)
2 2
uk − vk = 1 . (16.172)
These conditions are guarantied to be satisfied provided that uk and
vk are expressed using a single real parameter θk as
uk = cosh θk , vk = sinh θk , (16.173)
u−k = cosh θk , v−k = sinh θk . (16.174)
For this case Eq. (16.161) thus becomes

Vb | a†k ak |Vb  = sinh2 θk . (16.175)

7. Consider the unitary transformation [see for comparison Eq. (9.49)]

H̄k′ = eLk′ Hk′ e−Lk′ , (16.176)

where
3  4
Hk′ = ǫk′ a†k′ ak′ + λ ak′ + a†k′ , (16.177)

and where
 
Lk′ = −λ ak′ − a†k′ . (16.178)

With the help of Eq. (2.173), which is given by


1 1
eL Ae−L = A + [L, A] + [L, [L, A]] + [L, [L, [L, A]]] + · · · , (16.179)
2! 3!

Eyal Buks Quantum Mechanics - Lecture Notes 471


Chapter 16. Identical Particles

and the identities


3 4
ak′ , a†k′ ak′ = ak′ , (16.180)
3 4
a†k′ , a†k′ ak′ = −a†k′ , (16.181)
3 4
ak′ − a†k′ , ak′ + a†k′ = 2 , (16.182)

one finds that


 
H̄k′ = Hk′ − λǫk′ ak′ + a†k′ − 2ǫk′ λ2 + ǫk′ λ2
 
= ǫk′ a†k′ ak′ − λ2 .
(16.183)
Thus, the unitary transformation

H̄ = U † HU , (16.184)

where
! "

U = exp − Lk′ , (16.185)
k′

which yields
  
H̄ = ǫk′ a†k′ ak′ − λ2 , (16.186)
k′

can be employed for diagonalization of H. Let |n̄ be a number state,


which satisfy

a†k′ ak′ |n̄ = nk′ |n̄ , (16.187)

where nk′ is the number of particles in single-particle state |k′ . The


following holds
HU |n̄ = UU † HU |n̄
= U H̄ |n̄
  
= ǫk′ nk′ − λ2 U |n̄ ,
k′
(16.188)
thus the eigenvectors of H are the vectors U |n̄ and the corresponding
eigenenergies are given by
  
En̄ = ǫk′ nk′ − λ2 . (16.189)
k′

Eyal Buks Quantum Mechanics - Lecture Notes 472


16.11. Solutions

8. By employing the Bogoliubov transformation [see Eqs. (16.159), (16.173)


and (16.174)]
    
ak′ u−k′ 0 0 −vk′ bk′
 a−k′ 
  b−k  ,
1  0 uk′ −v−k′ 0   ′ 
 † = 
 ak′  uk′ u−k′ − vk′ v−k′  0 −vk′ u−k′ 0   b†k′ 
a†−k′ −v−k′ 0 0 uk′ b†−k′
(16.190)

where
uk′ = u−k′ = cosh θ k′ , (16.191)
vk′ = v−k′ = sinh θk′ , (16.192)
the identities
sinh (2θ k′ ) = 2 sinh θk′ cosh θ k′ , (16.193)
cosh (2θ k′ ) = sinh2 θk′ + cosh2 θk′ , (16.194)
cosh (2θk′ ) + 1
cosh2 θk′ = , (16.195)
2
cosh (2θk′ ) − 1
sinh2 θk′ = , (16.196)
2
3 4
and the commutation relation bk′ , b†k′ = 1, one finds that

H= (ǫk′ cosh (2θk′ ) − ξ k′ sinh (2θk′ )) b†k′ bk′
k′
 cosh (2θk′ ) − 1 ξ k′ sinh (2θk′ )
+ ǫk′ −
2 2
k′
 ξ ′ cosh (2θ k′ ) − ǫk′ sinh (2θk′ )  
 
+ k
bk′ b−k′ + b†k′ b†−k′ .

2
k
(16.197)
To eliminate the mixed terms bk′ b−k′ and b†k′ b†−k′ the factors θk′ are
chosen such that

ξ k′ cosh (2θk′ ) − ǫk′ sinh (2θk′ ) = 0 . (16.198)

Using the identities


1
cosh2 (2θk′ ) = , (16.199)
1 − tanh2 (2θk′ )
tanh2 (2θ k′ )
sinh2 (2θk′ ) = , (16.200)
1 − tanh2 (2θk′ )
one finds that for that choice H becomes diagonal

Eyal Buks Quantum Mechanics - Lecture Notes 473


Chapter 16. Identical Particles
:  2
 ξ k′
H= ǫk′ 1− b†k′ bk′
ǫk′
k′
: 
 ǫk′  2
 1− ξ k′
+ − 1 ,
2 ǫk′
k′

(16.201)
and thus, the eigenenergies are given by
:  2
 ξ k′
En̄ = ǫk′ 1 − nk′
ǫk′
k′
: 
 ǫk′  2
+  1 − ξ k′ − 1 ,

2 ǫk′
k

(16.202)
where the nonnegative integer nk′ is the number of so-called quasi parti-
cles in state k′ .
9. Consider the state |α (r′ ), which is defined by

|α (r′ ) = Dα(r′ ) |0 , (16.203)

where α (r′ ) ∈ C and where the operator Dα(r′ ) is given by [see for com-
parison Eq. (5.36)]
/
dr′ (α(r′ )Ψ † (r′ )−α∗ (r′ )Ψ (r′ ))
Dα(r′ ) = e . (16.204)

For general operators A and B the following holds [see Eq. (2.175)]
1 1
eA+B = eA eB e− 2 [A,B] = eB eA e 2 [A,B] , (16.205)

provided that

[A, [A, B]] = [B, [A, B]] = 0 . (16.206)

Moreover, with the help of Eq. (16.39) one finds that


  
′ ′ † ′ ′ ∗ ′ ′
dr α (r ) Ψ (r ) , − dr α (r ) Ψ (r )
 
, -
= dr′ dr′′ α (r′ ) α∗ (r′′ ) Ψ (r′′ ) , Ψ † (r′ )

2
= dr′ |α (r′ )| ,
(16.207)
thus [see for comparison Eq. (5.39)]

Eyal Buks Quantum Mechanics - Lecture Notes 474


16.11. Solutions
/ / / 2
dr′ α(r′ )Ψ † (r′ ) − dr′ α∗ (r′ )Ψ (r′ ) − 12 dr′ |α(r′ )|
Dα(r′ ) = e e e
/ / / 2
dr′ α∗ (r′ )Ψ (r′ ) dr′ α(r′ )Ψ † (r′ ) 1
dr′ |α(r′ )|
= e− e e2 .
(16.208)
Using the last result (16.208) it is easy to show that Dα(r′ ) is unitary
† †
Dα(r′ ) Dα(r′ ) = Dα(r′ ) Dα(r′ ) = 1 , (16.209)

and thus |α (r′ ) is normalized. With the help of Eq. (16.208) together
with the relation Ψ (r) |0 = 0 one finds that
/ 2 /
1
dr′ |α(r′ )| dr′ α(r′ )Ψ † (r′ )
|α (r′ ) = e− 2 e |0 . (16.210)

To show that |α (r′ ) is an eigenvector


, / of the quantized
 field -operator Ψ (r)
the commutation relation exp dr′ α (r′ ) Ψ † (r′ ) , Ψ (r) is evaluated
below. For general operators A and B and for a smooth function f (A)
the following holds
df
[f (A) , B] = [A, B] , (16.211)
dA
provided that [[A, B] , A]/ = 0 [see Eq. (2.174)]. Using this general result
[with f (A) = eA , A = dr′ α (r′ ) Ψ † (r′ ) and B = Ψ (r)] together with
Eq. (16.39) yields
3 / ′ 4 /
 
′ † ′ ′ ′ † ′
e dr α(r )Ψ (r ) , Ψ (r) = e dr α(r )Ψ (r ) dr′ α (r′ ) Ψ † (r′ ) , Ψ (r)
/

dr′ α(r′ )Ψ † (r′ )
= −e dr′ α (r′ ) δ (r − r′ )
/
dr′ α(r′ )Ψ † (r′ )
= −e α (r) ,
(16.212)
The last result together with the relation Ψ (r) |0 = 0 can be used to
show that the state |α (r′ ) is an eigenvector of Ψ (r) with eigenvalue α (r)
Ψ (r) |α (r′ )
/ ′ 2
/
= Ψ (r) e− 2 dr |α(r )| e dr α(r )Ψ (r ) |0
1 ′ ′ ′ † ′

/ ′ 2
 / ′ 3 / 4
= e− 2 dr |α(r )| e dr α(r )Ψ (r ) Ψ (r) + Ψ (r) , e
′ ′ † ′
1
dr′ α(r′ )Ψ † (r′ )
|0
/ 2 /
− 12 dr |α(r
′ ′
)| e ′
dr α(r )Ψ ′ † ′
(r ) |0 ,
= α (r) e
(16.213)
that is

Ψ (r) |α (r′ ) = α (r) |α (r′ ) . (16.214)

Eyal Buks Quantum Mechanics - Lecture Notes 475


Chapter 16. Identical Particles

The expectation value with respect to the number operator N [see Eqs.
(16.42) and (16.43)] is given by

α (r )| N |α (r ) = d3 r′ α (r′ )| Ψ † (r′ ) Ψ (r′ ) |α (r′ )
′ ′


= d3 r′ |α (r)|2 ,

whereas the expectation value with respect to the Hamiltonian H [see


Eq. (16.66)] is given by

′ ′ 2
α (r )| H |α (r ) = d3 r′ ∇α∗ (r) · ∇α (r)
2m

+ d3 r′ USP (r′ ) |α (r)|2
 
1 3 ′ 2 2
+ d r d3 r′′ VTP (r′ , r′′ ) |α (r′′ )| |α (r′ )| .
2
(16.215)
10. First consider the unperturbed problem where the Coulomb interaction
between the electrons is disregarded. The single-electron Hamiltonian is
obtained by substituting the factor e2 in the Hamiltonian of a hydro-
gen atom by Ze2 , where for helium Z = 2. The single electron energy
eigenstates |n, l, m, σ are chosen to be also eigenvectors of the single
electron angular momentum operators Lz and L2 [see Eqs. (7.42), (7.43)
and (7.44)]. While n, l and m are orbital quantum numbers, σ labels the
spin state. The single electron eigenenergies are given by [see Eq. (7.84)]

Z 2 EI
En = − , (16.216)
n2
where [see Eq. (7.66)]

me e4
EI = , (16.217)
22
and where me is the electron’s mass. The position wavefunction ψ n,l,m (r)
of a single-electron energy eigenstates having orbital quantum numbers
n, l and m is given by [see Eq. (7.92)]
(Z)
ψnlm (r, θ, φ) = Rnl (r) Ylm (θ, φ) , (16.218)
(Z)
where the radial wavefunction Rnl (r) is obtained by substituting e2
by Ze2 in the radial wave function Rnl (r) of hydrogen [see Eqs. (7.115),
(7.116) and (7.117)]. The ground state |Υ  (when the Coulomb interaction
between the electrons is disregarded) is given by [see Eq. (16.21)]

Eyal Buks Quantum Mechanics - Lecture Notes 476


16.11. Solutions

|Υ  = a†n=1,l=0,m=0,σ=− a†n=1,l=0,m=0,σ=+ |0 , (16.219)

where a†n,l,m,σ are creation operators and where |0 represents the state
where all occupation numbers are zero. The energy of the unperturbed
ground state is −2 × 22 EI = −8EI [see Eq. (16.216)]. The Coulomb in-
teraction between the electrons is described by the two-particle operator
[see Eq. (16.106)]

e2
VTP (r1 , r2 ) = . (16.220)
|r1 − r2 |

In the many-particle case the two-electron interaction is represented by


the operator V , which is given by Eq. (16.56). To first order in pertur-
bation theory the energy of the ground state [see Eq. (9.32)] is given by
−8EI + Υ | V |Υ , where [see Eq. (7.115)]
 

2 e2
Υ | V |Υ  = dr1 dr2
ψ1,0,0 (r1 ) ψ1,0,0 (r2 )

|r1 − r2 |

 
R(Z) (r ) R(Z) (r )
2 e2

= dr1 dr2
10 √ 1 10 √ 2


4π 4π
|r1 − r2 |
 6
e2 a20   4(r1 +r2 )
e− a0
= dr1 dr2
π2 |r1 − r2 |
= αEI ,
(16.221)
the dimensionless factor α is given by
  4(r1 +r2 )
27 1 e− a0
α= 2 5 dr1 dr2 , (16.222)
π a0 |r1 − r2 |

and where
2
a0 = (16.223)
me e2
is the Bohr’s radius [see Eq. (7.64)]. The integration over r2 is performed
in spherical coordinated, where the z axis is chosen in the direction of
the vector r1

Eyal Buks Quantum Mechanics - Lecture Notes 477


Chapter 16. Identical Particles

 ∞ 1 2π
27 1 −
4r1 4r2 d (cos θ2 )
α= 2 5 dr1 e a0
dr2 r22 e− a0  dφ2
π a0 r12 + r22 − 2r1 r2 cos θ2
0 −1 0
    
r1 +r2 −|r1 −r2 | 2π
r1 r2
 
7 ∞ r1 ∞
2 1 4r
− a1 1
4r
− a2
4r
− a2
= 4π 4π dr1 r12 e 0 dr2 r22 e 0 + dr2 r2 e 0 
π 2 a50 r1
0 0 r1
5
= ,
2
(16.224)
thus the ground state energy is −8EI + Υ | V |Υ  = − (11/2) EI . Note
that the fact the energy correction Υ | V |Υ  is comparable with the un-
perturbed value of −8EI suggests that the accuracy of the first order
perturbation approximation is relatively poor.
11. With the help of the commutation relations (16.39), (16.40) and (16.41)
one finds that
   
′ ′′ ′′′
γ |γ = dr dr dr dr′′′′ F (r′ , r′′ ) F ∗ (r′′′ , r′′′′ )

× 0| Ψ (r′′′′ ) Ψ (r′′′ ) Ψ † (r′ ) Ψ † (r′′ ) |0


  
= dr′ dr′′ dr′′′′ F (r′ , r′′ ) F ∗ (r′ , r′′′′ ) 0| Ψ (r′′′′ ) Ψ † (r′′ ) |0
   
′ ′′ ′′′
+ dr dr dr dr′′′′ F (r′ , r′′ ) F ∗ (r′′′ , r′′′′ ) 0| Ψ (r′′′′ ) Ψ † (r′ ) Ψ (r′′′ ) Ψ † (r′′ ) |0
   
2
= dr′ dr′′ |F (r′ , r′′ )| + F (r′ , r′′ ) F ∗ (r′′ , r′ ) .
(16.225)
a) The condition is
   
2
1 = dr′ dr′′ |F (r′ , r′′ )| + F (r′ , r′′ ) F ∗ (r′′ , r′ ) . (16.226)

b) The normalization
  condition for this case reads
 
2 2
1 = |A| dr′ dr′′ |f1 (r′ ) f2 (r′′ )| + f1 (r′ ) f2 (r′′ ) f1∗ (r′′ ) f2∗ (r′ )
 
= |A|2 1 + |γ 12 |2 ,
(16.227)
where

γ 12 = dr′ f1 (r′ ) f2∗ (r′ ) . (16.228)

The following holds [see Eqs. (16.39), (16.40) and (16.41)]

Eyal Buks Quantum Mechanics - Lecture Notes 478


16.11. Solutions
   
1
g (r′′′′′ ) = dr′ dr′′ dr′′′ dr′′′′ f1 (r′ ) f2 (r′′ ) f1∗ (r′′′ ) f2∗ (r′′′′ )
1 + |γ 12 |2
× 0| Ψ (r′′′′ ) Ψ (r′′′ ) Ψ † (r′′′′′ ) Ψ (r′′′′′ ) Ψ † (r′ ) Ψ † (r′′ ) |0
|f1 (r′′′′′ )|2 + |f2 (r′′′′′ )|2 + γ 12 f1∗ (r′′′′′ ) f2 (r′′′′′ ) + γ ∗12 f1 (r′′′′′ ) f2∗ (r′′′′′ )
= .
1 + |γ 12 |2
(16.229)
c) The number of particles is given by

Nγ = dr′ g (r′ ) = 2 . (16.230)

12. The correlation function Cσ (r′ − r′′ ) is given by


1  i(k′′ ·r′′ −k′ ·r′ )
Cσ (r′ − r′′ ) = e ϕ0 | a†k′ ,σ ak′′ ,σ |ϕ0  , (16.231)
V ′ ′′
k ,k

where |ϕ0  is the ground state of the free electron gas [see Eq. (16.98)],
thus
1  ik′ ·(r′′ −r′ )
Cσ (r′ − r′′ ) = e . (16.232)
V ′
|k |≤kF

For N ≫ 1 the summation can be approximately substituted by inte-


gration over the Fermi sphere having radius kF [see Eq. (16.100)]. In
spherical coordinates in which the z axis is taken to be in the direction
of the vector r′′ − r′ one has
 
V 2π kF ′ ′2 1
d (cos θ) eik cos θ|r −r | , (16.233)
′ ′′ ′
′ ′′
Cσ (r − r ) = 3 dk k
8π V 0 −1

thus

1 sin (kF |r′′ − r′ |) − kF |r′′ − r′ | cos (kF |r′′ − r′ |)


Cσ (r′ − r′′ ) = .
2π2 |r′′ − r′ |3
(16.234)

With the help of Eq. (16.100) the result can be expressed as

3N sin (kF |r′′ − r′ |) − kF |r′′ − r′ | cos (kF |r′′ − r′ |)


Cσ (r′ − r′′ ) = .
2V (kF |r′′ − r′ |)3
(16.235)

13. First consider the unperturbed case, where the electron-electron Coulomb
interaction is disregarded. The ground state

Eyal Buks Quantum Mechanics - Lecture Notes 479


Chapter 16. Identical Particles
#
|ϕ0  = a†k′ ,σ′ |0 (16.236)
|k′ |≤kF , σ′
 
is given by Eq. (16.98), and its energy E0 = (3N/5) 2 kF2 /2m by Eq.
(16.105), where kF is the Fermi wave vector. To first order in perturbation
(1)
theory the energy of the ground state becomes EGS = E0 + ∆E, where
the energy shift ∆E due to electron-electron Coulomb interaction is given
by [see Eqs. (9.32), (16.65), (16.93) and (16.106)]
 
1 
∆E = 3 ′
d r d3 r′′ VTP (r′ , r′′ ) ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0  ,
2 ′ ′′
σ ,σ
(16.237)

where
e2
VTP (r′ , r′′ ) = . (16.238)
|r′ − r′′ |

With the help of the expansion (16.96) and the commutation relations
(16.91) and (16.92) one finds that
ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0 
1  ′′ ′′′ ′′ ′ ′′′′ ′
= 2 ei(k −k )·r ei(k −k )·r ϕ0 | a†k′′′′ ,σ′ a†k′′′ ,σ′′ ak′′ ,σ′′ ak′ ,σ′ |ϕ0 
V ′ ′′ ′′′ ′′′′
k ,k ,k ,k
1  ′′
−k′ )·r′′ i(k′ −k′′′′ )·r′
= − 2 δ σ′ ,σ′′ ei(k e ϕ0 | a†k′′′′ ,σ′ ak′′ ,σ′ |ϕ0 
V
k′ ,k′′ ,k′′′′
1  ′′
−k′′′ )·r′′ i(k′ −k′′′′ )·r′
+ ei(k e ϕ0 | a†k′′′′ ,σ′ ak′ ,σ′ a†k′′′ ,σ′′ ak′′ ,σ′′ |ϕ0  .
V2
k′ ,k′′ ,k′′′ ,k′′′′
(16.239)
The only nonvanishing terms in the second line are those for which either
k′ = k′′′′ and k′′ = k′′′ or k′ = k′′′ and k′′ = k′′′′ . For the second case
the two possibilities σ′ = σ ′′ and σ ′ = σ′′ are separately considered

Eyal Buks Quantum Mechanics - Lecture Notes 480


16.11. Solutions

ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0 
1  ′ ′ ′′  ′′ ′ ′′
= − 2 δ σ′ ,σ′′ eik ·(r −r ) e−ik ·(r −r )
V ′ ′′
k |k |≤kF
1 
+ 2 ϕ | Nk′ ,σ′ Nk′′ ,σ′′ |ϕ0 
V ′ ′′ 0
k ,k
1  ′ ′′ ′ ′′
+ 2 δ σ′ ,σ′′ ei(k −k )·(r −r ) ϕ0 | a†k′′ ,σ′ ak′ ,σ′ a†k′ ,σ′ ak′′ ,σ′ |ϕ0 
V ′ ′′
k ,k
1  ′ ′′ ′ ′′
+ 2 (1 − δ σ′ ,σ′′ ) ei(k −k )·(r −r ) ϕ0 | a†k′′ ,σ′ ak′ ,σ′ a†k′ ,σ′′ ak′′ ,σ′′ |ϕ0  ,
V ′ ′′ k ,k
(16.240)
thus
ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0 
1  ′ ′ ′′  ′′ ′ ′′
= − 2 δ σ′ ,σ′′ eik ·(r −r ) e−ik ·(r −r )
V ′ ′′
k |k |≤kF
1 
+ 1
V2
|k′′ |,|k′′ |≤kF
1 
+ 2
δ σ′ ,σ′′ 1
V ′
|k |≤kF
1  ik′ · r′ −r′′  −ik′′ · r′ −r′′
+ 2
δ σ′ ,σ′′ e ( ) e ( )
V
k′ |k′′ |≤kF


2



1
′ ′ ′′

ik ·(r −r )

− 2 δ σ′ ,σ′′

e
V


|k′ |≤kF

1 
+ 2 (1 − δ σ′ ,σ′′ ) 1.
V ′ |k |≤kF

(16.241)
For N ≫ 1 the single summation terms are negligibly small
ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0 


2



1 1
′ ′ ′′

ik ·(r −r )

= 2 1 − 2 δ σ′ ,σ′′

e
V V
,
|k′′ |,|k′′ |≤kF
|k′ |≤kF

(16.242)
or [see Eqs. (16.232) and (16.235)]

Eyal Buks Quantum Mechanics - Lecture Notes 481


Chapter 16. Identical Particles

ϕ0 | Ψσ†′ (r′ ) Ψσ†′′ (r′′ ) Ψσ′′ (r′′ ) Ψσ′ (r′ ) |ϕ0 
 2
N 2
= − δ σ′ ,σ′′ |C (r′ − r′′ )| ,
2V
(16.243)
where
3N sin (kF |r′′ − r′ |) − kF |r′′ − r′ | cos (kF |r′′ − r′ |)
C (r′ − r′′ ) = .
2V (kF |r′′ − r′ |)3
(16.244)

With the help of the above result one obtains


 2  
1 N
∆E = d3 r′ d3 r′′ VTP (r′ , r′′ )
2 V
 
3 ′ 2
− d r d3 r′′ VTP (r′ , r′′ ) |C (r′ − r′′ )| .
(16.245)
The first term of (16.245) represents the electrostatic energy due to
electron-electron interaction. However, as is argued below, in the pres-
ence of positive charge density +eN/V this term should be disregarded.
This can be seen by noticing that the self electrostatic energy of the
positive background is identical to the first term of (16.245), whereas
the electrostatic energy due to interaction
/ / between the electrons and the
positive background is − (N/V)2 d3 r′ d3 r′′ VTP (r′ , r′′ ), thus these two
contributions exactly cancels the first term of (16.245). The second term,
which is commonly called the exchange energy, can be evaluated using
(1)
Eq. (16.235). Thus the energy of the ground state EGS (to first order in
perturbation theory) is given by
(1) 3N 2 kF2
EGS =
5 2m
  2
e2 3N (sin (kF |r′ |) − kF |r′ | cos (kF |r′ |))2
−V d3 r′ ′
|r | 2V (kF |r′ |)6
 2  ∞
3N 2 kF2 9πVe2 NV (sin x − x cos x)2
= − 2 dx ,
5 2m kF 0 x5
  
1/4

(16.246)
or [see Eq. (16.100)]
 2 2 
(1) 3  kF 3kF e2
EGS = N − . (16.247)
5 2m 4π

Eyal Buks Quantum Mechanics - Lecture Notes 482


17. Superconductivity

In this chapter two models are discussed, the London’s model, in which a
macroscopic wavefunction is introduced to describe the state of a supercon-
ductor, and the model by Bardeen, Cooper and Schrieffer (BCS), which pro-
vides an insight on the underlying microscopic mechanisms that are respon-
sible for superconductivity.

17.1 Macroscopic Wavefunction


In this chapter the London’s equations are derived from the assumption that
the state of a superconductor can be describe using a macroscopic wavefunc-
tion.

17.1.1 Single Particle in Electromagnetic Field


Consider a single particle having charge q and mass m in electromagnetic
field characterized by the scalar potential ϕ and the vector potential A. The
electric field E and the magnetic field B are given by (in Gaussian units) [see
Eqs. (1.41) and (1.42)]
1 ∂A
E = −∇ϕ − , (17.1)
c ∂t
and
B=∇×A, (17.2)
8 −1
where c = 2.99 × 10 m s is the speed of light in vacuum. Let r = (x, y, z)
be the position vector of the particle in Cartesian coordinates. The variable
vector canonically conjugate to the position vector r is given by [see Eq.
(1.61)]
q
p = mṙ+ A . (17.3)
c
The classical equation of motion is given by [see Eq. (1.60)]
 
1
mr̈ = q E + ṙ × B . (17.4)
c
Chapter 17. Superconductivity

The Schrödinger Equation. The Hamiltonian of the system is given by


[see Eq. (1.62)]
 2
p− qc A
H= + qϕ . (17.5)
2m
The Schrödinger equation for the wavefunction ψ (r′ , t′ ) is given by [see Eq.
(4.194)]
dψ 1  q 2
i = −i∇− A ψ + qϕψ . (17.6)
dt 2m c
The continuity Equation. The continuity equation expresses the proba-
bility conservation law [see Eq. (4.73)]

+∇·J=0 , (17.7)
dt
where
ρ = ψψ∗ (17.8)
is the probability distribution function and
 qρ
J= Im (ψ∗ ∇ψ) − A (17.9)
m mc
is the current density [see Eq. (4.200)]. For a wavefunction having the form
ψ = ρ1/2 eiθ , (17.10)
where θ is real, one has [see Eq. (6.361)]
ρ  q 
J= ∇θ− A . (17.11)
m c
Gauge Invariance. Consider the following gauge transformation [see Eqs.
(12.49) and (12.50)]
A → Ã = A + ∇χ , (17.12)
ϕ → ϕ̃ = ϕ , (17.13)
where χ = χ (r) is an arbitrary smooth and continuous function of r, which
is assumed to be time independent. This transformation leaves E and B
unchanged [see Eqs. (12.1) and (12.2)], however, the wavefunction is trans-
formed according to the following rule. Given that the wavefunction ψ (r′ , t′ )
solves the Schrödinger equation with vector A and scalar ϕ potentials, the
transformed Schrödinger equation with vector à = A+∇χ and scalar ϕ̃ = ϕ
potentials is solved by the transformed wavefunction ψ̃ (r′ , t′ ), which is given
by [see Eq. (12.53)]
 
iqχ (r′ )
ψ̃ (r′ , t′ ) = exp ψ (r′ , t′ ) . (17.14)
c

Eyal Buks Quantum Mechanics - Lecture Notes 484


17.1. Macroscopic Wavefunction

17.1.2 Drude Model

Consider a conductor containing charge carriers having charge q and mass m


in electromagnetic field. The density of charge carriers (i.e. number per unit
volume) is n. Scattering is taken into account in the Drude model by adding
a damping term to the classical equation of motion (17.4)
   
1 1
m r̈+ ṙ = q E + ṙ × B , (17.15)
τ tr c

where τ tr is the so-called scattering time. For simplicity the applied magnetic
field is assumed to vanish, i.e. B = 0. In terms of the current density vector
J, which is related to the the velocity vector v = ṙ by the relation
J
v= , (17.16)
qn
Eq. (17.15) yields
 
m ∂J 1
+ J =E. (17.17)
q 2 n ∂t τ tr

Maxwell’s Equations. The macroscopic Maxwell’s equations (in Gaussian


units) for the electric field E, electric displacement D, magnetic induction
B and magnetic field H in the presence of external charge density ρext and
external current density Jext are given by
4π 1 ∂D
∇×H = Jext + , (17.18)
c c ∂t
1 ∂B
∇×E = − , (17.19)
c ∂t
∇ · D = 4πρext , (17.20)
∇·B = 0 . (17.21)
For an isotropic and linear medium the following relations hold
D = E+4πP , (17.22)
D = ǫE , (17.23)
P = χe E , (17.24)
B = H + 4πM , (17.25)
B = µH , (17.26)
M = χm H , (17.27)
where P is the electric polarization, ǫ = 1 + 4πχe is the permittivity (di-
electric constant of the medium), χe is the electric susceptibility, M is the
magnetization, µ = 1 + 4πχm is the permeability and χm is the magnetic
susceptibility.

Eyal Buks Quantum Mechanics - Lecture Notes 485


Chapter 17. Superconductivity

Fourier Expansion. In general, a scalar function f (r, t) can be Fourier


expanded as
 
f (r, t) = dq dω ei(q·r−ωt) f (q, ω) , (17.28)

and a vector function F (r, t) can be Fourier expanded as


 
F (r, t) = dq dω ei(q·r−ωt) F (q, ω) . (17.29)

A vector function F (r, t) can be decomposed into longitudinal and trans-


verse parts with respect to the wave vector q according to

F = FL + FT , (17.30)

where the longitudinal part is given by FL = (n̂q · F) n̂q , the transverse one
is given by FT = (n̂q × F) × n̂q , and where n̂q = q/ |q| is a unit vector in
the direction of q. The following holds ∇ · F = ∇ · FL and ∇ × F = ∇ × FT .
Recall that for a general scalar φ and a vector A the following holds
∇ · (φA) = φ∇ · A + A · ∇φ ,
∇ × (φA) = φ∇ × A − A × ∇φ ,
thus
 
∇·F = dq dω ei(q·r−ωt) iq · FL (q, ω) , (17.31)

and
 
∇×F= dq dω ei(q·r−ωt) iq × FT (q, ω) . (17.32)

With the help of the above relations the Maxwell’s equations (17.18),
(17.19), (17.20) and (17.21) can be Fourier transformed into
4π iω
iq × HT (q, ω) = Jext (q, ω) − D (q, ω) , (17.33)
c c
ω
q × ET (q, ω) = B (q, ω) , (17.34)
c
iq · DL (q, ω) = 4πρext (q, ω) , (17.35)
q · BL (q, ω) = 0 . (17.36)
Dielectric Constant. While the external charge density ρext is related to
D by the relation ∇ · D = 4πρext [see Eq. (17.20)], the induced charge
density ρind , which is defined as the change in charge density with respect to
the unperturbed case, is related to the electric polarization by the relation
∇·P =−ρind , and the total charge density ρind +ρext is related to the electric
field E by the relation

Eyal Buks Quantum Mechanics - Lecture Notes 486


17.1. Macroscopic Wavefunction

∇ · E = 4π (ρind + ρext ) , (17.37)

which can be Fourier transformed into

iq · EL (q, ω) = 4π (ρind (q, ω) + ρext (q, ω)) . (17.38)

The longitudinal dielectric function ǫ (q, ω), which is defined by [compare


with Eq. (17.23)]

|DL (q, ω)|


ǫ (q, ω) ≡ , (17.39)
|EL (q, ω)|

is thus given by [see Eqs. (17.35) and (17.38)]

ρext (q, ω)
ǫ (q, ω) = . (17.40)
ρext (q, ω) + ρind (q, ω)
For a general longitudinal field FL the following holds ∇ × FL = 0, and thus
both DL and EL can be expressed in terms of scalar potentials as
DL (r, t) = −∇ϕext (r, t) , (17.41)
EL (r, t) = −∇ϕ (r, t) . (17.42)
Consequently, one finds that the longitudinal dielectric function ǫ (q, ω) can
alternatively be expressed as [see Eq. (17.39)]

ϕext (q, ω)
ǫ (q, ω) = . (17.43)
ϕ (q, ω)
For the case where the induced charge is carried by electrons having mass
me , charge qe , density ne and scattering time τ tr,e Eq. (17.17) becomes
 
me ∂Je 1
+ Je = E , (17.44)
qe2 ne ∂t τ tr,e
where the electron current density Je is related to the induced electron charge
density ρind = ρe by the continuity equation (17.7)
dρe
+ ∇ · Je = 0 . (17.45)
dt
Applying ∇· to Eq. (17.44) and using Eqs. (17.45) and (17.37) lead to

d2 ρe 1 dρe
2
+ = −ω 2p,e (ρe + ρext ) , (17.46)
dt τ tr,e dt
where ω p,e , which is given by

4πqe2 ne
ω 2p,e = , (17.47)
me

Eyal Buks Quantum Mechanics - Lecture Notes 487


Chapter 17. Superconductivity

is the so-called electron plasma frequency. Consider the


 case where ρext is
taken to oscillate according to ρext = Re ρext,0 e−iωt , where the angular
frequency ω is a real constant.
 Seeking a solution to Eq. (17.46) having the
form ρe = Re ρe,0 e−iωt yields

ω 2p,e
ρe,0 = ρ , (17.48)
1
ω 2 + iω τ tr,e − ω 2p,e ext,0

thus the frequency dependent dielectric constant ǫe (ω) is given by [see Eq.
(17.40)]

ω 2p,e iωτ tr,e


ǫe (ω) = 1 − . (17.49)
ω 2 iωτ tr,e − 1

Alternatively, in terms of the skin depth δ e , which is given by


:
c 2
δe = , (17.50)
ω p,e ωτ tr,e

the dielectric constant ǫe (ω) can be expressed as

2ic2 1
ǫe (ω) = 1 + 2 2 1 − iωτ . (17.51)
δe ω tr,e

Complex Conductivity. Consider  the case


 where the applied electric field
−iωt
E is taken to be given byE = Re E 0 e . Seeking a solution of Eq. (17.44)
having the form Je = Re Je,0 e−iωt yields

Je,0 = σ e (ω) E0 , (17.52)

where
σe,0
σe (ω) = , (17.53)
1 − iωτ tr,e

is the complex conductivity σ (ω) and where

qe2 ne τ tr,e
σe,0 = . (17.54)
me
In terms of the complex conductivity σe (ω) the dielectric constant ǫe (ω) can
be expressed as [see Eq. (17.49)]

4πσ e (ω)
ǫe (ω) = 1 + i . (17.55)
ω

Eyal Buks Quantum Mechanics - Lecture Notes 488


17.1. Macroscopic Wavefunction

17.1.3 The Macroscopic Quantum Model

The macroscopic quantum model is based on the hypothesis that some prop-
erties of a superconducting media can be described by a single wavefunction
ψs (r′ , t′ ). It is assumed that the local density of superconducting charge car-
riers n⋆s is related to the wavefunction by
2
n⋆s = |ψs (r′ , t′ )| . (17.56)

In the presence of an electromagnetic field the time evolution of ψs (r′ , t′ ) is


governed by the Schrödinger equation [see Eq. (17.6)]
 2
dψ s 1 q⋆
i = −i∇− s A ψs + qs⋆ ϕψs . (17.57)
dt 2m⋆s c

where m⋆s and qs⋆ are the mass and charge respectively of a superconducting
charge carrier. Furthermore, it is assumed that the current density carried
by a superconductor having a macroscopic wavefunction given by [see Eq.
(17.10)]
 ′ ′
ψs (r′ , t′ ) = n⋆s (r′ , t′ )eiθ(r ,t ) , (17.58)

is given by
 
qs⋆ n⋆s (r′ , t′ ) q⋆
Js = ∇θ− s A . (17.59)
m⋆s c

Note that while J in Eq. (17.11) represents probability current density, Js in


a superconductor represents charge current density.

17.1.4 London Equations

London equations can be derived from the macroscopic quantum model by


assuming that the superconducting charge carriers density n⋆s is constant.
2nd London Equation. By taking the curl of Eq. (17.59) and employing
Eq. (17.2) one obtains the second London equation, which reads

qs⋆2 n⋆s
∇ × Js = − B. (17.60)
m⋆s c

Taking the curl of Eq. (17.18), assuming ǫ = µ = 1 and employing Eqs.


(17.60), (17.19) and (17.21) together with the general vector identity

∇ × ∇ × B = ∇ (∇ · B) − ∇2 B (17.61)

lead to

Eyal Buks Quantum Mechanics - Lecture Notes 489


Chapter 17. Superconductivity

1 1 ∂2B
∇2 B = 2 B + c2 ∂t2 , (17.62)
λL
where
:
m⋆s c2
λL = (17.63)
4πn⋆s qs⋆2

is the London penetration depth in Gaussian units (λL = m⋆s /µ0 n⋆s qs⋆2 in
SI units). In terms of the superconducting plasma frequency ωp,s , which is
given by
4πn⋆s qs⋆2
ω 2p,s = , (17.64)
m⋆s
the London penetration depth can be expressed as
c
λL = . (17.65)
ω p,s
For time independent B the solution of Eq. (17.62) yields an exponential
decay of B with characteristic decay length given by the London penetration
depth λL . Thus, except of a region having characteristic width λL near the
surfaces the magnetic field inside a superconductor vanishes (even in the
presence of an externally applied magnetic filed). This expulsion of a magnetic
field from a superconductor, which is called the Meissner effect , represents the
perfect diamagnetism of superconductors. As can be seen from Eq. (17.18), in
the absence of time dependent electric field (i.e. when displacement currents
vanish) the expulsion of a magnetic field also implies that the supercurrent
density Js also vanishes deep inside a superconductor.
1st London Equation. When the superconducting density of charge carri-
ers n⋆s is assumed to be a constant Eq. (17.57) becomes [see Eq. (17.58)]
 2
dθ 1 qs⋆
− = ∇θ− A + qs⋆ ϕ , (17.66)
dt 2m⋆s c
or [see Eq. (17.59)]
dθ m⋆
− = ⋆2 s ⋆2 J2s + qs⋆ ϕ . (17.67)
dt 2qs ns
Applying ∇ to the above leads to
d∇θ m⋆
− = ⋆2 s ⋆2 ∇J2s + qs⋆ ∇ϕ . (17.68)
dt 2qs ns
Taking the time derivative of Eq. (17.59) and employing Eq. (17.1) together
with the last result yield the first London equation

Eyal Buks Quantum Mechanics - Lecture Notes 490


17.1. Macroscopic Wavefunction
 
m⋆s ∂Js 1
+ ⋆ ⋆ ∇J2s =E. (17.69)
qs⋆2 n⋆s ∂t 2qs ns
Typically in superconductors the electric field E on the right hand side
of Eq. (17.69) can be neglected in comparison with the term proportional to
∇J2s on the left

hand
side of Eq. (17.69). The factor ∇J2s can be estimated by
2
the relation
∇J2s
≃ 2 |Js | /l0 , where l0 is a length scale that characterizes
the spacial variations of the current density Js . Moreover, the ratio |Js | /l0
can be estimated from the second London equation (17.60)
|Js | q ⋆2 n⋆
≃ s ⋆ s |B| . (17.70)
l0 ms c
Combining these results allows estimating the term proportional to ∇J2s on
the left hand side of Eq. (17.69)
m⋆s 1


∇J2s
≃ 1 |vs | |B| , (17.71)
qs n⋆s
⋆2 2qs⋆ n⋆s c
where [see Eq. (17.16)]
Js
vs = (17.72)
qs⋆ n⋆s
is the velocity of superconducting charge carriers. In view of the classical
equation of motion (17.4) the above estimate shows that the ratio between
|E| and the the term proportional to ∇J2s in Eq. (17.69) represents the ratio
between electric and magnetic forces acting on the superconducting charges.
Typically in metals electric forces are strongly suppressed due to screening,
and consequently can be neglected in comparison with magnetic forces. Ne-
glecting the E term in Eq. (17.69) leads to
∂Js 1
+ ∇J2s = 0 . (17.73)
∂t 2qs⋆ n⋆s
Homogeneous solutions (i.e. position independent solutions) of the first
London equation (17.69) satisfy
m⋆s ∂Js
=E. (17.74)
qs⋆2 n⋆s ∂t
Comparing the above relation with Eq. (17.17) of the Drude model yields
agreement provided that the scattering time τ tr is assumed to be infinite.
From this point of view the first London equation represents the nullification
of resistance in superconductors. Equation (17.74) implies that the supercon-
ducting conductivity σs (ω) for the homogeneous case is given by [see Eq.
(17.53)]
qs⋆2 n⋆s
σs (ω) = i . (17.75)
ωm⋆s

Eyal Buks Quantum Mechanics - Lecture Notes 491


Chapter 17. Superconductivity

Flux Quantization. Consider a close curve C inside a superconductor. In-


tegrating Eq. (17.59), which is given by
 
qs⋆ n⋆s qs⋆
Js = ∇θ− A , (17.76)
m⋆s c
along the curve yields
@  @ @ 
q ⋆ n⋆ q⋆
dr · Js = s ⋆s  dr · ∇θ− s dr · A . (17.77)
C ms C c C

The assumption that Ethe superconducting wavefunction ψs = n⋆s eiθ is con-
tinuous implies that dr · ∇θ = 2nπ, where n is integer. The integral over
A can be calculated using Stokes’ theorem [see Eqs. (12.2) and (12.47)]
@
dr · A = φC , (17.78)
C
/
where φC = ds · B is the magnetic flux threaded through the area enclosed
by the closed path C. With these results Eq. (17.77) becomes
@  
hq ⋆ n⋆ φ
dr · Js = s ⋆ s n− C , (17.79)
C ms φs
where
hc
φs = (17.80)
qs⋆
is the so called superconducting flux quantum (in Gaussian units).
As was shown above, the second London equation implies that the super-
current density Js vanishes deep inside a superconductor. Consider a close
curve C inside a superconductor and assume that the distance between any
point on C and the nearest surface is much larger than the London penetra-
tion depth λL . For such a curve the left hand side of Eq. (17.79) vanishes,
and consequently

φC = nφs , (17.81)

i.e. the magnetic flux is quantized in units of the superconducting flux quan-
tum.
Two Fluid Model. In the limit of vanishing temperature only supercon-
ducting charge carriers are present in a superconductor. However, at finite
temperature also normally conducting charge carriers may be present. Con-
sequently, the total complex conductivity σ (ω) [see Eq. (17.52)] is taken to
be given by

σ (ω) = σ n (ω) + σs (ω) , (17.82)

Eyal Buks Quantum Mechanics - Lecture Notes 492


17.2. The Josephson Effect

where the normal conductivity σ n (ω) is assumed to be given by the Drude


formula (17.53)

qe2 nn τ tr,e 1
σn (ω) = , (17.83)
me 1 − iωτ tr,e

where nn is the density of normal charge carriers, and the super conductivity
σ s (ω) is assumed to be given by Eq. (17.75)

qs⋆2 n⋆s
σs (ω) = i . (17.84)
ωm⋆s

The electronic dielectric constant ǫe (ω) [see Eq. (17.55)] in the two fluid
model is thus given by

4πσ n (ω) 4πσs (ω)


ǫe (ω) = 1 + i +i , (17.85)
ω ω
or in terms of the skin depth δ [see Eq. (17.50)] and the London penetration
depth λL [see Eq. (17.63)]
2i 1 1
ǫe (ω) = 1 + 2 − , (17.86)
(δ e k) 1 − iωτ tr,e (λL k)2

where k = ω/c. Note that 1/δ 2 ∝ nn [see Eq. (17.50)], whereas 1/λ2L ∝ ns
[see Eq. (17.63)]. Note also that the ratio between these characteristic length
scales is given by
:
λL m⋆s nn q 2 ωτ tr
= . (17.87)
δL m n⋆s qs⋆2 2

17.2 The Josephson Effect

A Josephson junction is formed between two superconductors that are weakly


connected to each other. Electrons can flow between the superconductors by
crossing a barrier. In this chapter the Josephson relations are derived from
the model that was discussed in the previous chapter.

17.2.1 The First Josephson Relation

In terms of the superconducting flux quantum φs [see Eq. (17.80)] Eq. (17.76)
can be rewritten as
qs⋆ n⋆s 
Js = ∇θGI , (17.88)
m⋆s

Eyal Buks Quantum Mechanics - Lecture Notes 493


Chapter 17. Superconductivity

where

∇θGI = ∇θ − A. (17.89)
φs
The phase factor θGI is commonly called the gauge invariant phase.
Consider an integral over Js (17.88) along a path going through a Joseph-
son junction from point r1 on the interface between the first superconductor
and the barrier to point r2 on the interface between the second superconduc-
tor and the barrier. The path is assumed to be parallel to the direction of
the current density Js . The total current flowing through the junction I is
expected to be a function of the phase difference Θ, which is obtained from
integrating ∇θGI
 r2 
2π r2
Θ= dr · ∇θGI = θ (r2 ) − θ (r1 ) − dr · A . (17.90)
r1 φs r1

Furthermore, by symmetry, I is expected to be an odd function of Θ. In



addition, since the macroscopic wavefunction ψs = n⋆s eiθ is unaffected when
θ is replaced by θ+2π [see Eq. (17.58)], I is expected to be a periodic function
of Θ with a period of 2π, i.e. I can be expressed as


I= In sin (nΘ) , (17.91)
n=1

where In are all constants. As will be shown below [see Eq. (17.313)], for the
case where the superconductors are weakly coupled all terms except for the
first one can be neglected. For that case the general current -phase relation
(17.91) becomes the so-called first Josephson relation, which is given by

I = Ic sin Θ , (17.92)

where Ic is the so-called critical current of the junction.

17.2.2 The Second Josephson Relation

The time derivative of Θ (17.90), which is denoted by Θ̇, can be expressed


with the help of Eq. (17.67) as
 
1 m⋆s  2 2
 ⋆ 2π r2 ∂A
Θ̇ = − Js (r2 ) − Js (r1 ) + qs (ϕ (r2 ) − ϕ (r1 )) − dr· .
 2qs⋆2 n⋆2
s φ s r1 ∂t
(17.93)

The assumption Js (r2 ) = Js (r1 ) leads to


  
2πc r2 1 ∂A
Θ̇ = dr · −∇ϕ − . (17.94)
φs r1 c ∂t

Eyal Buks Quantum Mechanics - Lecture Notes 494


17.2. The Josephson Effect

With the help of Eq. (17.1) this becomes


2πc
Θ̇ = V , (17.95)
φs
where V , which is given by
 r2
V = dr · E , (17.96)
r1

is the voltage across the Josephson junction. As will be shown below, the
elementary superconducting charge carrier is a pair of electrons, i.e. qs⋆ = 2e,
and consequently Eq. (17.95) becomes
2eV
Θ̇ = . (17.97)

The above result is the second Josephson relation.

17.2.3 The Energy of a Josephson Junction

Let I (t) and V (t) be the current through and voltage across a Josephson
junction, respectively, at time t. The energy UJ of the junction can be eval-
uated by integrating the work done by the source
 t
UJ = dt′ I (t′ ) V (t′ ) . (17.98)

With the help of the first (17.92) and second (17.97) Josephson relation this
becomes

Ic Θ
UJ = dΘ′ sin Θ′ , (17.99)
2e
thus up to a constant UJ is given by

UJ = −EJ cos Θ , (17.100)

where
Ic φ Ic
EJ = = s . (17.101)
2e 2πc
The energy UJ (17.100) can be expressed as [see Eq. (17.313)]
:  2
I
UJ = −EJ 1 − . (17.102)
Ic

To second order in I this becomes [see Eq. (17.314)]

Eyal Buks Quantum Mechanics - Lecture Notes 495


Chapter 17. Superconductivity

LJ I 2  
UJ = −EJ + + O I4 , (17.103)
2
where
φs
LJ = (17.104)
2πcIc
is the so-called Josephson inductance. Note, however, that an inductor-like
behavior of a Josephson junction is expected only when I ≪ Ic .

17.3 RF SQUID

A radio frequency (RF) superconducting quantum interference device (SQUID)


is made of a superconducting loop interrupted by a Josephson junction (see
Fig. 17.1). Consider a close curve C going around the loop. The requirement
that the phase θ of the macroscopic wavefunction is continues reads
@
2nπ = dr · ∇θ , (17.105)
C

where n is integer. The section of the close curve C inside the superconductor
is denoted by C − and the integral through the junction is denoted as an
integral from point r1 to point r2 . With the help of Eq. (17.76) the above
condition becomes
 r2  
m⋆ m⋆ 2π
2nπ = ⋆ s⋆ dr · Js + ⋆ s⋆ dr · Js + dr · A . (17.106)
qs ns  r1 qs ns  C− φs C

Consider the case where the curve is chosen such that the supercurrent density
Js vanishes everywhere on the curve C − (i.e. inside the superconductor the
distance between any point on C − and the nearest surface is much larger than
the London penetration depth λL ). For this case Eq. (17.106) becomes

2πφ
2nπ = Θ + , (17.107)
φs

where
 r2  r2
m⋆s
Θ= dr · Js = dr · ∇θGI (17.108)
qs n⋆s 

r1 r1

is the gauge invariant phase difference across the junction [see Eqs. (17.88)
and (17.90)] and where
@
φ = dr · A (17.109)
C

Eyal Buks Quantum Mechanics - Lecture Notes 496


17.3. RF SQUID

Fig. 17.1. RF SQUID is made of a superconducting loop interrupted by a Joseph-


son junction.

is the magnetic flux threaded through the area enclosed by the closed path
C [see Eq. (17.78)].
The junction’s critical current is labeled by Ic . It is assumed that the
junction has capacitance, which is denoted by CJ . Consider the case where a
magnetic flux that is denoted by φe is externally applied by a magnetic field,
which is applied perpendicularly to the plane of the loop. The total magnetic
flux φ threading the loop is given by
φ = φe + ΛIs , (17.110)
where Is is the circulating current flowing in the loop and Λ is the self induc-
tance of the loop.

17.3.1 Lagrangian
The Lagrangian of the system [see Eq. (1.16)] can be expressed as a function
of the dimensionless flux coordinate Φ, which is defined by
2πφ
Φ= , (17.111)
φs
and its time derivative Φ̇. According to Faraday’s law of induction the voltage
across the capacitor (in Gaussian units) is

Eyal Buks Quantum Mechanics - Lecture Notes 497


Chapter 17. Superconductivity

φ̇
V =− , (17.112)
c
and therefore the kinetic energy of the system T is the capacitance energy
2
CJ φ̇ CJ φ2s Φ̇2
T = = . (17.113)
2c2 8π2 c2
The potential energy U has two contributions, the inductive energy (in
Gaussian units)

ΛIs2 (φ − φe )2 φ2 (Φ − Φe )2
= = s , (17.114)
2c 2Λc 8π 2 Λc
where
2πφe
Φe = (17.115)
φs

is the normalized external flux, and the Josephson energy UJ [see Eqs.
(17.100) and (17.107)]

φs Ic
UJ = − cos Φ . (17.116)
2πc
Thus the Lagrangian L = T − U is given by

CJ φ2s Φ̇2 φ2s (Φ − Φe )2 φs Ic


L= − + cos Φ , (17.117)
8π2 c2 8π 2 cΛ 2πc
or in a dimensionless form by
! "
Λ Φ̇2
L = E0 − u (Φ; Φe ) , (17.118)
LJ ω 2p

where the energy constant E0 is given by

φ2s
E0 = , (17.119)
8π2 Λc
the junction’s plasma frequency ω p is given by
0 0
c 2ecIc
ωp = = , (17.120)
LJ CJ CJ

where LJ = φs /2πcIc is the Josephson inductance [see Eq. (17.104)], the


dimensionless potential u (Φ; Φe ) is given by

u (Φ; Φe ) = (Φ − Φe )2 − 2β L cos Φ , (17.121)

Eyal Buks Quantum Mechanics - Lecture Notes 498


17.3. RF SQUID

and the dimensionless parameter β L is given by

2πΛIc
βL = . (17.122)
φs

The resulting Euler - Lagrange equation of motion (1.8) is given by


 
d ∂L ∂L
= , (17.123)
dt ∂ Φ̇ ∂Φ

thus
Λ Φ̈
+ Φ − Φe + β L sin Φ = 0 . (17.124)
LJ ω 2p

With the help of Eqs. (17.107), (17.110) and (17.112) the equation of motion
can be rewritten as

Is = Ic sin Θ + CJ V̇ . (17.125)

The above equation states that the circulating current Is equals the sum of
the current Ic sin Θ through the Josephson junction and the current CJ V̇
through the capacitor.

17.3.2 Flux Quantum Bit

Consider the case where the externally applied flux φe is chosen to be close
to a half integer value in units of the superconducting flux quantum φs . To
study this case the normalized externally applied flux Φe (17.115) is taken to
be given by

Φe = π + Φe,r , (17.126)

where Φe,r ≪ 1. Using the notation

Φ = π + Φr , (17.127)

the dimensionless potential u (17.121) becomes

u = (Φr − Φe,r )2 + 2β L cos Φr . (17.128)

Consider that case where Φe,r = 0 (i.e. when Φe = π). As can be seen
from Eq. (17.110), for this case the relative coordinate Φr is proportional to
the circulating current Is
2πΛ
Φr = Is . (17.129)
φs

Eyal Buks Quantum Mechanics - Lecture Notes 499


Chapter 17. Superconductivity

Fig. 17.2. The LC resonator that is coupled to the RF SQUID allows readout.

Moreover, when Φe,r = 0 one finds to second order in Φr that the potential u
is given by
 
u = 2β L + (1 − β L ) Φ2r + O Φ4r . (17.130)
Thus if β L > 1 the point Φr = 0 becomes a local maxima point of u. The
corresponding potential barrier centered at Φr = 0 (i.e. at Φ = π) separates
two symmetric potential wells on the right and on the left. At sufficiently low
temperatures only the two lowest energy levels are expected to be occupied.
In this limit the Hamiltonian of the system can be expressed in the basis of
the states | and |, that represent localized states in the left and right
well, respectively, having opposite circulating currents. Such a device can be
used as an artificial two-level system, i.e. as a quantum bit (qubit in short).

17.3.3 Qubit Readout


Readout of the state of an RF SQUID can be performed by inductively cou-
pling the superconducting loop to an LC resonator (see Fig. 17.2), which is
made of an inductor and a capacitor in parallel having inductance L and
capacitance C respectively. The mutual inductance between the RF SQUID
and the resonator is denoted by M . Detection is performed by injecting a
monochromatic input current Iin into the LC resonator at a frequency close
to the resonance frequency and measuring the output voltage Vout (see Fig.
17.2).
The total magnetic flux φ threading the SQUID loop is given by [compare
with Eq. (17.110)]
φ = φe + φi , (17.131)

Eyal Buks Quantum Mechanics - Lecture Notes 500


17.3. RF SQUID

where the term φi represents the flux generated by both, the circulating
current in the RF SQUID Is and by the current in the inductor of the LC
resonator IL

φi = Is Λ + M IL , (17.132)

where Λ is the self inductance of the loop. Similarly, the magnetic flux ϕ in
the inductor of the LC resonator is given by

ϕ = IL L + MIs . (17.133)

Eqs. (17.132) and (17.133) can be rewritten in a matrix form


    
φi Λ M Is
= . (17.134)
ϕ M L IL

Inverting the above relation allows expressing the currents Is and IL in terms
of φi = φ − φe and ϕ

φi Mϕ
Is = − , (17.135)
Λ (1 − K 2 ) ΛL (1 − K 2 )
ϕ Mφi
IL = − , (17.136)
L (1 − K ) ΛL (1 − K 2 )
2

where the dimensionless constant K is given by


M
K=√ . (17.137)
ΛL
Lagrangian. The Lagrangian of the system L = T − U [see Eq. (1.16)] is
expressed below as a function of the coordinates Φ and ϕ and their time
derivatives Φ̇ and ϕ̇. The contributions to the total kinetic energy T are the
capacitance energy of the Josephson junction that is given by Eq. (17.113)
and the capacitance energy of the capacitor in the LC resonator, thus T is
given by

CJ φ2s Φ̇2 C ϕ̇2


T = + 2 . (17.138)
8π 2 c2 2c
The inductive energy UI stored in the RF SQUID loop and the lumped in-
ductor L is calculated using Eqs. (17.135) and (17.136)

Eyal Buks Quantum Mechanics - Lecture Notes 501


Chapter 17. Superconductivity
  
1   Λ M Is
UI = Is IL
2c M L IL
 1 M
 
1   − ΛL φi
= φ ϕ Λ
2c (1 − K 2 ) i M
− ΛL L
1
ϕ
φ2i 2φ ϕM 2

Λ − ΛLi
+ ϕL
=
2c (1 − K 2 )
 2

ϕ 2 φ i − L
= +
2cL 2cΛ (1 − K 2 )
 2
2 2πMϕ
2 2
Cω e ϕ φ s Φ − Φ e − φs L
= + ,
2c 8π cΛ (1 − K 2 )
2

(17.139)
where
1
ωe = √ (17.140)
LC
is the LC angular resonance frequency. The total potential energy U is given
by
Iin ϕ φs Ic
U = UI − − cos Φ , (17.141)
c 2πc
where the term −Iin ϕ/c is the potential energy of the current source and
− (φs Ic /2πc) cos Φ is the Josephson energy [see Eq. (17.116)].
With the help of the above relations one finds that the Lagrangian of the
system L = T − U can be expressed as

L = L0 + L1 , (17.142)

where L0 , which is given by

C ϕ̇2 Cω2e ϕ2 Iin ϕ


L0 = − + , (17.143)
2c2 2c c
is the Lagrangian of the driven LC resonator. The Lagrangian of the super-
conducting loop L1 is given by [see Eqs. (17.119), (17.120) and (17.122)]
 2
2 2πMϕ
2 2
CJ φs Φ̇ φs Φ − Φ e − φs L φ Ic
L1 = − + s cos Φ
8π2 c2 8π 2 cΛ (1 − K 2 ) 2πc
! "
Λ Φ̇2
= E0 − uK (Φ; Φe,eff ) ,
LJ ω 2p
(17.144)

Eyal Buks Quantum Mechanics - Lecture Notes 502


17.3. RF SQUID

where the dimensionless potential uK (Φ; Φe,eff ) is given by [compare with Eq.
(17.121)]
2
(Φ − Φe,eff )
uK (Φ; Φe,eff ) = − 2β L cos Φ , (17.145)
1 − K2
and where the effective external flux Φe,eff is given by

2πM ϕ
Φe,eff = Φe + . (17.146)
φs L

Note that L1 depends on the effective external flux Φe,eff , which, in turn,
depends on the coordinate ϕ of the LC resonator [see Eq. (17.146)]. This
dependence gives rise to the coupling between the LC resonator and the RF
SQUID.
The Euler - Lagrange equations (1.8) are given by
 
d ∂L ∂L
= , (17.147)
dt ∂ Φ̇ ∂Φ
 
d ∂L ∂L
= , (17.148)
dt ∂ ϕ̇ ∂ϕ
thus
φs (Φ−Φe )
CJ φs Φ̈ 2π − Mϕ L
=− − Ic sin Φ , (17.149)
2πc Λ (1 − K 2 )

and
φ (Φ−Φ )
C ϕ̈ ϕ− MΛ
s

e

=− + Iin . (17.150)
c L (1 − K 2 )

With the help of Eqs. (17.135) and (17.136) the equations of motion can be
rewritten as

CJ φ̈
Is = Ic sin Θ − , (17.151)
c
and
C ϕ̈
Iin = + IL . (17.152)
c
While Eq. (17.151) expresses the law of current conservation in the SQUID
loop, Eq. (17.152) expresses the same law in the LC resonator.

Eyal Buks Quantum Mechanics - Lecture Notes 503


Chapter 17. Superconductivity

Hamiltonian. The variables canonically conjugate to Φ and ϕ are given by


[see Eq. (1.20)]
∂L 2E0 ΛΦ̇
Q= = . (17.153)
∂ Φ̇ LJ ω2p
∂L C ϕ̇
q= = 2 . (17.154)
∂ ϕ̇ c
The Hamiltonian is given by [see Eq. (1.22)]

H = QΦ̇ + q ϕ̇ − L = H0 + H1 , (17.155)

where
c2 q 2 Cω 2e ϕ2 Iin ϕ
H0 = + − , (17.156)
2C 2c c
and where
LJ ω2p Q2
H1 = + E0 uK (Φ; Φe,eff ) . (17.157)
4E0 Λ
Quantization is achieved by regarding the variables {Φ, Q, ϕ, q} as Her-
mitian operators satisfying the following commutation relations [see Eqs.
(3.6), (3.7) and (3.8)]

[Φ, Q] = [ϕ, q] = i , (17.158)

and

[ϕ, Φ] = [ϕ, Q] = [q, Φ] = [q, Q] = 0 . (17.159)

Adiabatic Approximation. The dynamics of the superconducting loop is


typically much faster than the one of the LC resonator. This can be exploited
for simplifying the system’s equations of motion by employing the adiabatic
approximation (see section 13.6). Similarly to L1 , the Hamiltonian H1 of the
superconducting loop depends on the effective external flux Φe,eff , which in
turn depends on the coordinate ϕ of the LC resonator [see Eq. (17.146)].
As a basis for expanding the general solution we use the eigenvectors of
the following Schrödinger equation

H1 |n (ϕ) = εn (ϕ) |n (ϕ) , (17.160)

where ϕ is treated here as a parameter (rather than a degree of freedom).


The local eigenvectors are assumed to be orthonormal

m (ϕ) |n (ϕ) = δ nm . (17.161)

The eigenenergies εn (ϕ) and the associated wavefunctions ψn (Φ, ϕ) are


found by solving the following Schrödinger equation [see Eq. (4.50)]

Eyal Buks Quantum Mechanics - Lecture Notes 504


17.3. RF SQUID

d2 ψn εn
−β J + uK (Φ; Φe,eff ) ψn = ψ , (17.162)
dΦ2 E0 n
where
 2
LJ ω p
βJ = . (17.163)
Λ 2E0

The state of the entire system ψ (t) at time t is expanded at any point ϕ
using the ’local’ basis {|n (ϕ)} [see Eq. (13.71)]

ψ (t) = ξ n (ϕ, t) |n (ϕ) . (17.164)
n

In the adiabatic approximation the time evolution of the coefficients ξ n is gov-


erned by the following set of decoupled equations of motion [see Eq. (13.83)]

[H0 + εn (ϕ)] ξ n = iξ̇ n . (17.165)

Note that in the present case the geometrical vector potential [given by Eq.
(13.76) for the case m = m] vanishes since the wavefunctions ψn (Φ, ϕ) can
be chosen to be real [see Eq. (13.84)].
Two level approximation. Consider again the case where the externally
applied flux φe is chosen to be close to a half integer value in units of the
superconducting flux quantum φs . The potential uK (17.145) can be expressed
as

(Φr − Φe,eff,r )2
uK = + 2β L cos Φr , (17.166)
1 − K2
where Φe,eff,r and Φr are defined by [see Eqs. (17.126) and (17.127)]
2πMϕ
Φe,eff = Φe + = π + Φe,eff,r , (17.167)
φs L
Φ = π + Φr . (17.168)
 
In what follows we focus on the case where |Φe,eff,r | ≪ 1 and β L 1 − K 2 >
1. In this case the potential uK given by Eq. (17.166) contains two wells sepa-
rated by a barrier near Φr = 0 (see Fig. 17.3). As was discussed above, at low
temperatures only the two lowest energy levels are expected to contribute. In
this limit the Hamiltonian H1 can be expressed in the basis of the states |
and |, representing localized states in the left and right well respectively
having opposite circulating currents. In this basis, H1 is represented by the
2 × 2 matrix
 
ηΦe,eff,r ∆
H1 =
˙ . (17.169)
∆ −ηΦe,eff,r

Eyal Buks Quantum Mechanics - Lecture Notes 505


Chapter 17. Superconductivity

The real parameters η and ∆ (having units of energy) can be determined


by solving numerically the Schrödinger equation (17.162). It is convenient to
express η in terms of the parameter Icc
φs Icc
η= . (17.170)
2πc
The physical meaning of the parameter Icc , which has units of current, will
be discussed below.
Using the notation
∆ 2πc∆
tan α = = , (17.171)
ηΦe,eff,r φs Icc Φe,eff,r
H1 can be rewritten as
: 2  
φs Icc Φe,eff,r cos α sin α
H1 =˙ + ∆2 . (17.172)
2πc sin α − cos α
The eigenvectors and eigenenergies are denoted as
H1 |± = ε± |± , (17.173)
where [compare with Eqs. (6.213) and (6.214)]
   
cos α2 − sin α2
|+ =˙ ; |− =
˙ , (17.174)
sin α2 cos α2
and where
: 2
φs Icc Φe,eff,r
ε± = ± + ∆2
2πc
  2 
1 φs Icc Φe,eff,r  
= ±∆ 1 + + O Φ4e,eff,r .
2 2πc∆
(17.175)

Exercise 17.3.1. Show that the expectation value of the circulating current
Is when the qubit is in state n is given by
∂εn
n| Is |n = −c . (17.176)
∂φe
Solution 17.3.1. With the help of Eqs. (17.110), (17.132), (17.135), (17.145)
and (17.157) one finds that
∂H1
−c = Is , (17.177)
∂φe
thus
∂H1 ∂εn
n| Is |n = −c n| |n = −c . (17.178)
∂φe ∂φe

Eyal Buks Quantum Mechanics - Lecture Notes 506


17.3. RF SQUID

Fig. 17.3. Eigenstates of H1 . (a)-(c) The first 3 lowest energy states for the case
Φe,eff,r = 0. (d) The energy of the two lowest states vs. Φe,eff,r .

Eyal Buks Quantum Mechanics - Lecture Notes 507


Chapter 17. Superconductivity

With the help of Eqs. (17.175) and (17.176) one finds that the expectation
value of the circulating current Is corresponding to the eigenvectors |+ and
|− is given by
φs Icc Φe,eff,r
2πc∆
±| Is |± = ∓Icc 0  2 . (17.179)
φs Icc Φe,eff,r
1+ 2πc∆

In the limit where φs Icc |Φe,eff,r | /2πc∆ → ∞ (i.e. when Φe,eff is far from the
point Φe,eff = π) the eigenvectors of H1 become the states | and | [see
Eqs. (17.171) and (17.174)]. Thus, in view of the above result (17.179), it is
evident the the constant Icc represents the absolute value of the circulating
current associated with the states | and |.
In the adiabatic approximation the dynamics of the LC resonator is gov-
erned by the Hamiltonian H± , which is given by

H± = H0 + ε± (ϕ) , (17.180)

when the qubit is the state |± [see Eq. (17.165)]. With the help of Eqs.
(17.156), (17.167) and (17.175) one finds that

c2 q 2 Iin ϕ
H± = + U± (ϕ) − , (17.181)
2C c
where
: 2
Cω 2e ϕ2 φs Icc Φe,eff,r
U± (ϕ) = ± + ∆2 , (17.182)
2c 2πc

and where
2πMϕ
Φe,eff,r = Φe − π + . (17.183)
φs L
For simplicity, consider the case where Φe = π. For that case one finds to
second order in M that [see Eqs. (17.140) and (17.175)]
C 2 2  
U± (ϕ) = ω ϕ ± ∆ + O M4 , (17.184)
2c ±
where
0
2
M M Icc
ω± = ωe 1± . (17.185)
L c∆
Thus the effective angular resonance frequency of the LC resonator ω ± de-
pends on the state being occupied by the qubit. This dependence allows
reading out the qubit state by measuring the resonance frequency of the LC
resonator.

Eyal Buks Quantum Mechanics - Lecture Notes 508


17.4. BCS Model

17.4 BCS Model

This chapter briefly discusses the BCS microscopic model of superconductiv-


ity.

17.4.1 Phonon Mediated Electron-Electron Interaction

Let ρ (r′ ) be the electron density in a medium having volume V. Classi-


cally, the two-particle Coulomb interaction VTP (r1 , r2 ) = e2 / |r1 − r2 | [see
Eq. (16.106)] gives rise to energy V given by [see Eqs. (16.65) and (16.43) for
comparison with the analogous second-quantization expression]
 
1
V = d3 r′ d3 r′′ VTP (r′ , r′′ ) ρ (r′ ) ρ (r′′ ) . (17.186)
2
With the help of the Fourier expansion

1 ′ ′
ρ (r′ ) = 3/2 √ d3 q′ ρ (q′ ) eiq ·r (17.187)
(2π) V

and Eqs. (4.47) and (16.127) one finds that [see Eq. (16.128) for comparison
with the analogous second-quantization expression]

1 4πe2
V = d3 q 2 ρ (q) ρ (−q) . (17.188)
2V q
The effect of induced charges in the medium (i.e. screening) can be taken
into account by dividing by the dielectric constant of the medium ǫ [see Eq.
(17.43)]

1 4πe2
V = d3 q 2 ρ (q) ρ (−q) . (17.189)
2V q ǫ
Ionic Dielectric Constant. The expression for the electronic dielectric con-
stant ǫe (ω) given by Eq. (17.49) is evaluated by calculating the induced elec-
tron charge density ρe due to external charge density ρext . Below, however,
we focus on the induced ion charge density ρi associated with lattice vibra-
tions and the resultant ionic dielectric constant ǫi . To account for the effect of
lattice vibrations on electron-electron interaction both contributions to the
total induced charge density ρe + ρi have to be taken into account. With
analogy to Eq. (17.46), the equation of motion for the ion induced charge
density ρi is given by

d2 ρi 1 dρi
+ = −ω 2p,i (ρe + ρi + ρext ) , (17.190)
dt2 τ tr,i dt

where ω p,i , which is given by

Eyal Buks Quantum Mechanics - Lecture Notes 509


Chapter 17. Superconductivity

4πqi2 ni
ω 2p,i = , (17.191)
mi
is the ion plasma frequency and where mi , qi and ni are the ionic mass, charge
and density, respectively. Note that ωp,i ≪ ωp,e since the ion mass mi is much
larger than electron mass me .
Thomas-Fermi Approximation. The electron density ne is assumed to be
a constant provided that the media is homogeneous. It can be evaluated by
summing the Fermi-Dirac function fFD (ǫi ) over all electronic states having
energies ǫi [see Eq. (16.147)]
1 1 1
ne (µ) = fFD (ǫi ) = , (17.192)
V i V i exp [β (ǫi − µ)] + 1

where V is the volume, β −1 = kB T is the thermal energy and where µ is


the chemical potential. The external charge density ρext together with the
induced charge density ρe + ρi give rise to a spatially varying scalar potential
ϕ (r), which is related to the total charge density by the Poisson’s equation

∇2 ϕ = −4π (ρe + ρi + ρext ) . (17.193)

In the Thomas-Fermi approximation, which is valid provided that ϕ (r) is a


slowly varying function of position on the length scale of electron wavelength,
the local value of electron density is taken to be given by ne (µ − qe ϕ (r)),
and consequently the electron induced charge density ρe (r) is taken to be
given by

ρe (r) = qe [ne (µ − qe ϕ (r)) − ne (µ)] . (17.194)

For sufficiently small ϕ this becomes


∂ne
ρe (r) = −qe2 ϕ (r) . (17.195)
∂µ
When the thermal energy kB T is much smaller than the Fermi energy ǫF the
factor ∂ne /∂µ is approximately the density of states at the Fermi energy ǫF ,
which is given by [see Eq. (16.103)]

∂ne m2 vF
≃ 2e 3 , (17.196)
∂µ π 

where vF is the so-called Fermi velocity (which is defined by the relation


∂ǫk′ /∂k′ = vF , where k′ is the wave number and where the derivative is
taken at ǫk′ = ǫF ). For this case Eq. (17.195) becomes

m2e vF
ρe (r) = −qe2 ϕ (r) . (17.197)
π2 3

Eyal Buks Quantum Mechanics - Lecture Notes 510


17.4. BCS Model

Combing Eqs. (17.193) and (17.197) yields the following relation between the
electron induced charge density ρe and the total charge density ρe + ρi + ρext

∇2 ρe = kTF
2
(ρe + ρi + ρext ) , (17.198)

where

2 4πqe2 m2e vF
kTF = . (17.199)
π2 3
Effective Electron-Electron Interaction. Consider the case where ρext 
is taken to oscillate in time and space according to Re ρext,0 ei(q·r−ωt) ,
where both the real angular frequency ω and the real vector q are con-
stants.
 In steady state the induced
 charges
 oscillate according to ρe =
Re ρe,0 ei(q·r−ωt) and ρi = Re ρi,0 ei(q·r−ωt) . Substituting into Eqs. (17.190)
and (17.198) yields
iω  
−ω2 ρi,0 − ρ = −ω2p,i ρe,0 + ρi,0 + ρext,0 , (17.200)
τ tr,i i,0

and
 
−q 2 ρe,0 = kTF
2
ρe,0 + ρi,0 + ρext,0 . (17.201)

The solution is given by


 
ω2 1+ ωτi
2 tr,i
−kTF ω2p,i
ρext,0
ρe,0 = !   " , (17.202)
ω2 1+ ωτi
2 )
(q 2 + kTF tr,i
−1 2
+ kTF
ω2p,i

and
q 2 ρext,0
ρi,0 = !   " , (17.203)
ω2 1+ ωτi
(q 2 + 2 )
kTF tr,i
−1 + 2
kTF
ω2p,i

thus the ionic dielectric constant ǫi (q, ω) is given by [see Eq. (17.40)]
1 ρe,0 + ρi,0 + ρext,0
=
ǫi (q, ω) ρext,0
 
ω2 1+ ωτi
2 tr,i
q ω2p,i
= !   " .
ω2 1+ ωτi
2 )
(q 2 + kTF tr,i
−1 2
+ kTF
ω2p,i

(17.204)

Eyal Buks Quantum Mechanics - Lecture Notes 511


Chapter 17. Superconductivity

For the case where ωτ tr,i ≫ 1 this becomes

1 q2 ω2
= 2 2 2 2 , (17.205)
ǫi (q, ω) q + kTF ω − Ωp,i

where

2 q2
Ωp,i = ω 2p,i 2 . (17.206)
q 2 + kTF

The expression for the Coulomb energy (17.189) together with the result
for the ionic dielectric constant ǫi (q, ω) (17.205) lead to the effective inter-
action coefficient for a pair of electrons having wave vectors k′ and k′′ and
energies ǫk′ and ǫk′′ respectively

4πe2 4πe2 ω2
vk′ ,k′′ = = 2 ω2 − Ω2 , (17.207)
q 2 ǫi (q, ω) q 2 + kTF p,i

where q = k′′ − k′ and where ω = (ǫk′′ − ǫk′ ) /.


The fact that ǫ−1i (q, ω) becomes negative when ω < Ωp,i indicates that
the effective (i.e. phonon mediated) electron-electron interaction becomes at-
tractive in the limit of low frequencies. The characteristic energy interval
Ωp,i in which the interaction becomes attractive is of the order of the so-
called Debye energy ǫD , which represents the largest energy of an acoustic
phonons in the lattice.

17.4.2 The Hamiltonian

In the BCS model the Hamiltonian of electrons in a superconducting metal


is taken to be given by
  
H= (ǫk′ − ǫF ) a†k′ ,↑ ak′ ,↑ + a†k′ ,↓ ak′ ,↓
k′
ǫF −ǫD <ǫk′ ,ǫk′′ <ǫF +ǫD
g 
− a†k′′ ,↑ a†−k′′ ,↓ a−k′ ,↓ ak′ ,↑ .
V
k′ ,k′′
(17.208)
where ↑ labels spin up state, ↓ labels spin down state, ǫk′ is the energy of
both single particle states |k′ , ↑ and |k′ , ↓ and ǫF is the Fermi energy [see
Eq. (16.101)]. The coupling constant g > 0 gives rise for an effective electron-
electron attracting interaction. The interaction is assume to couple pairs of
electrons whose energies are inside an energy interval of width 2ǫD around
the Fermi energy ǫF .
As can be seen from the comparison with the more general many-particle
interaction operator V given by Eq. (16.93), the BCS Hamiltonian contains

Eyal Buks Quantum Mechanics - Lecture Notes 512


17.4. BCS Model

only interaction terms that represents annihilation (the factor a−k′ ,↓ ak′ ,↑ )
and creation (the factor a†k′′ ,↑ a†−k′′ ,↓ ) of electrons pairs having zero total an-
gular momentum. Moreover, the summation is restricted only to the energy
interval of width 2ǫD in which attractive interaction is expected, and the effec-
tive interaction coefficients are all assumed to be identical [see for comparison
Eq. (17.207)].

17.4.3 Bogoliubov Transformation

As will be seen below, the Hamiltonian can be approximately diagonalized


by employing the Bogoliubov transformation [see Eqs. (16.110), (16.153) and
(16.154)]
   iφ  
bk,↑ e k cos θk 0 0 e−iφk sin θ k ak,↑
 b−k,↓   0 eiφk cos θk −e−iφk sin θk 0   a−k,↓ 
 † = iφk −iφk
 †  ,
 bk,↑   0 e sin θk e cos θk 0   ak,↑ 

b−k,↓ iφk
−e sin θk 0 0 e−iφk
cos θk a†−k,↓
(17.209)
where both θ k and φk are real. The inverse transformation is given by
   −iφ  
ak,↑ e k cos θk 0 0 −e−iφk sin θk bk,↑
 a−k,↓   0 e−iφk cos θk e−iφk sin θ k 0   b−k,↓ 
 † = iφk iφk
 †  .
 ak,↑   0 −e sin θ k e cos θk 0   bk,↑ 
† iφk iφk
a−k,↓ e sin θk 0 0 e cos θ k b†−k,↓
(17.210)
The following holds
a†k,↑ ak,↑ = sin2 θ k (1 − N−k,↓ ) + cos2 θk Nk,↑
sin (2θ k )  
+ bk,↑ b−k,↓ + b†−k,↓ b†k,↑ ,
2
(17.211)
and
a†−k,↓ a−k,↓ = sin2 θk (1 − Nk,↑ ) + cos2 θ k N−k,↓
sin (2θk )  
+ bk,↑ b−k,↓ + b†−k,↓ b†k,↑ ,
2
(17.212)
where

Nk,σ = b†k,σ bk,σ (17.213)

is number operator (with respect to the bk,σ and b†k,σ operators ).

Eyal Buks Quantum Mechanics - Lecture Notes 513


Chapter 17. Superconductivity

Using the above results together with Eqs. (16.153) and (16.154) one finds
that the Hamiltonian can be rewritten as

H=2 (ǫk′ − ǫF ) sin2 θk′
k′

+ (ǫk′ − ǫF ) cos (2θk′ ) (Nk′ ,↑ + N−k′ ,↓ )
k′
  
+ (ǫk′ − ǫF ) sin (2θk′ ) bk′ ,↑ b−k′ ,↓ + b†−k′ ,↓ b†k′ ,↑
k′
ǫF −ǫD <ǫk′ ,ǫk′′ <ǫF +ǫD
g 
− Bk† ′′ Bk′ ,
V
k′ ,k′′
(17.214)
where
Bk′ = a−k′ ,↓ ak′ ,↑
e−2iφk′ sin (2θk′ )
= (N−k′ ,↓ + Nk′ ,↑ − 1)
2 
+e−2iφk′ sin2 θk′ b†−k′ ,↓ b†k′ ,↑ − cos2 θk′ bk′ ,↑ b−k′ ,↓ .
(17.215)
Formally, the operator Bk′ can be written as

Bk′ = Bk′  + δBk′ , (17.216)

where Bk′  is the expectation value of Bk′  in thermal equilibrium and


where

δBk′ = Bk′ − Bk′  . (17.217)

In the mean field approximation the coupling term Bk† ′′ Bk′ is approximated
by
' ( ' (
Bk† ′′ Bk′ ≃ Bk† ′′ Bk′  + Bk† ′′ δBk′ + δBk† ′′ Bk′  , (17.218)

i.e. the term δBk′ , which represents the deviation from the expectation value,
is considered as small. Employing this approximation, using Eq. (17.215)
and disregarding all constant terms [the' first(line in Eq. (17.214) and terms
containing only expectation values, e.g. Bk† ′′ Bk′ ] result in the mean field
Hamiltonian HMF , which is give by

Eyal Buks Quantum Mechanics - Lecture Notes 514


17.4. BCS Model

ǫF −ǫD <ǫk′ <ǫF +ǫD


  2iφ ′  
 ∆e k + ∆∗ e−2iφk′ sin (2θk′ )
HMF = (ǫk′ − ǫF ) cos (2θk′ ) −
2
k′
× (Nk′ ,↑ + N−k′ ,↓ )
ǫF −ǫD <ǫk′ <ǫF +ǫD
 , -
+ (ǫk′ − ǫF ) sin (2θk′ ) + ∆∗ e−2iφk′ cos2 θk′ − ∆e2iφk′ sin2 θk′
k′
×bk′ ,↑ b−k′ ,↓
ǫF −ǫD <ǫk′ <ǫF +ǫD
 , -
+ (ǫk′ − ǫF ) sin (2θk′ ) + ∆e2iφk′ cos2 θk′ − ∆∗ e−2iφk′ sin2 θk′
k′
† †
×b−k′ ,↓ bk′ ,↑ .
(17.219)
where
ǫF −ǫD <ǫk′ <ǫF +ǫD
g 
∆= Bk′  . (17.220)
V
k′

Note that all terms outside the energy interval ǫF − ǫD < ǫk′ < ǫF + ǫD were
disregarded in the above expression for HMF .
Diagonalization is achieved when the last two sums in Eq. (17.219) vanish,
i.e. when the terms ∆e2iφk′ are real and positive and when the parameters
θ k′ are chosen such that the following condition is satisfied

(ǫk′ − ǫF ) sin (2θk′ ) + |∆| cos (2θ k′ ) = 0 . (17.221)

For this case


 
1 |∆|
θk′ = tan−1 − , (17.222)
2 ǫk′ − ǫF
i.e.
|∆|
sin (2θk′ ) = − . , (17.223)
(ǫk′ − ǫF )2 + |∆|2
ǫk′ − ǫF
cos (2θk′ ) = . , (17.224)
2 2
(ǫk′ − ǫF ) + |∆|
and
7
8
8 1 − √ ǫk′ −ǫF
9 (ǫk′ −ǫF )2 +|∆|2
sin (θ k′ ) = , (17.225)
2
7
8
8 1 + √ ǫk′ −ǫF
9 (ǫk′ −ǫF )2 +|∆|2
cos (θ k′ ) = , (17.226)
2

Eyal Buks Quantum Mechanics - Lecture Notes 515


Chapter 17. Superconductivity

and thus HMF becomes diagonal



HMF = ηk′ Nk′ ,σ , (17.227)
k′ ,σ

where
.
2 2
ηk′ = (ǫk′ − ǫF ) + |∆| . (17.228)

17.4.4 The Energy Gap


The value of the energy gap |∆| can be determined from Eq. (17.220). Let
nk′ ,σ denotes the expectation value of the operator Nk′ ,σ , i.e.
Nk′ ,σ  = nk′ ,σ . (17.229)
In thermal equilibrium at temperature T the following holds [see Eqs.
(16.144), (16.145), (17.227) and (17.228)]
1
nk′ ,σ = , (17.230)
eβηk′ + 1
where β = 1/kB T', and where( kB is Boltzmann’s constant. Moreover, in ther-
mal equilibrium b†−k′ ,↓ b†k′ ,↑ = bk′ ,↑ b−k′ ,↓  = 0 and thus Bk′  is given by
[see Eq. (17.215) and recall that |∆| = ∆e2iφk′ ]
∆ (1 − n−k′ ,↓ − nk′ ,↑ )
Bk′  = . , (17.231)
2 (ǫk′ − ǫF )2 + |∆|2
and thus Eq. (17.220) can be expressed as
ǫF −ǫD <ǫk′ <ǫF +ǫD
g  1 − n−k′ ,↓ − nk′ ,↑
1= . , (17.232)
2V
k′ (ǫk′ − ǫF )2 + |∆|2
or [see Eqs. (17.228) and (17.230)]
ǫF −ǫD <ǫk′ <ǫF +ǫD  −1
g  1 − 2 eβηk′ + 1
1=
2V ηk′
k′
 
ǫF −ǫD <ǫk′ <ǫF +ǫD tanh βη k′
g  2
= .
2V ′
ηk′
k
(17.233)
Replacing the sum by an integral leads to

 β ǫ′2 +|∆|2
gD0 ǫD ′ tanh 2
1= dǫ . , (17.234)
2 −ǫD
ǫ + |∆|2
′2

where D0 is the density of states per unit volume.

Eyal Buks Quantum Mechanics - Lecture Notes 516


17.4. BCS Model

Zero Temperature. For the case of zero temperature, where all occupation
numbers vanish, i.e. nk′ ,σ = 0, Eq. (17.234) becomes

gD0 ǫD dǫ′
1= 
2 −ǫD ǫ′2 + ∆20

gD0 ǫD + ǫ2D + ∆20
= log  ,
2 −ǫD + ǫ2D + ∆20
(17.235)
where ∆0 stands for the value of |∆| at zero temperature. The assumption
∆0 ≪ ǫD leads to

gD0 4ǫ2
1= log D , (17.236)
2 ∆20

thus
 
1
∆0 = 2ǫD exp − . (17.237)
gD0

Critical Temperature. The energy gap |∆| vanishes when T = Tc , where


Tc is the critical temperature. For this case Eq. (17.234) becomes
 β ǫ′
gD0 ǫD ′ tanh c2
1= dǫ
2 −ǫD ǫ′
 c2 D
β ǫ
tanh x
= gD0 dx ,
0 x
(17.238)
where β c = 1/kB Tc . Integration by parts (note that limx→0 tanh x log x = 0)
yields
!  βc2ǫD "
β c ǫD β c ǫD log x
1 = gD0 tanh log − dx . (17.239)
2 2 0 cosh2 x

For the case of weak coupling, for which


β c ǫD
≫1, (17.240)
2
one has
  ∞ 
β ǫD log x
1 ≃ gD0 log c − dx . (17.241)
2 0 cosh2 x
Using the identity

Eyal Buks Quantum Mechanics - Lecture Notes 517


Chapter 17. Superconductivity
 ∞
log x 4
− dx 2 = log + CE , (17.242)
0 cosh x π
where CE ≃ 0.577 is Euler’s constant, one finds that [see Eq. (17.237)]
eCE
kB Tc = ∆0 ≃ 0.566 × ∆0 . (17.243)
π
General Temperature. The energy gap |∆| at temperature T can be nu-
merically evaluated from Eq. (17.234). To a good approximation the solution
can be expressed by the following analytical relation
:  3
T
|∆| ≃ ∆0 1 − . (17.244)
Tc

17.4.5 The Ground State


<
The ground state |Ψ0  of the mean field Hamiltonian HMF = k′ ,σ η k′ Nk′ ,σ
(17.227) is a state for which all occupation numbers vanish, i.e. Nk′ ,σ |Ψ0  =
b†k,σ bk,σ |Ψ0  = 0, and therefore bk,σ |Ψ0  = 0 for all k and σ. Moreover, |Ψ0 
is required to be normalized, i.e. Ψ0 |Ψ0  = 1.
Claim. The ground state |Ψ0  is given by
#
|Ψ0  = Kk′ |0 , (17.245)
k′

where
 
Kk′ = eiφk′ cos θk′ − e−iφk′ sin θk′ a†k′ ,↑ a†−k′ ,↓ . (17.246)
3 4
Proof. By employing the fact that Kk† ′′ , Kk′ = 0 provided that k′ = k′′ and
the relation
Kk† ′ Kk′ = cos2 θk′ + sin2 θk′ a−k′ ,↓ ak′ ,↑ a†k′ ,↑ a†−k′ ,↓
 
− sin θk′ cos θk′ e−2iφk′ a†k′ ,↑ a†−k′ ,↓ + e2iφk′ a−k′ ,↓ ak′ ,↑ ,
(17.247)
one finds that |Ψ0  is indeed normalized as required
# † #
Ψ0 |Ψ0  = 0| Kk′′ Kk′ |0
k′′ k′
#
= 0| Kk† ′ Kk′ |0
k′
#3   4
= 0| cos2 θ k′ + sin2 θ k′ 1 − a†k′ ,↑ ak′ ,↑ 1 − a†−k′ ,↓ a−k′ ,↓ |0
k′
=1.
(17.248)

Eyal Buks Quantum Mechanics - Lecture Notes 518


17.4. BCS Model

Moreover, using the relations [see Eq. (17.209)]


 
bk′ ,↑ Kk′ = eiφk′ cos θk′ ak′ ,↑ − e−iφk′ sin θ k′ a†−k′ ,↓
 
× eiφk′ cos θ k′ − e−iφk′ sin θk′ a†k′ ,↑ a†−k′ ,↓ ,
(17.249)
and
 
b−k,↓ Kk′ = eiφk′ cos θ k′ a−k′ ,↓ − e−iφk′ sin θk′ a†k′ ,↑
 
× eiφk′ cos θk′ − e−iφk′ sin θ k′ a†k′ ,↑ a†−k′ ,↓ ,
(17.250)
one finds that
#
bk′ ,↑ |Ψ0  = bk′ ,↑ Kk′′ |0
k′′
#
= Kk′′ bk′ ,↑ Kk′ |0
k′′
=k′
#  
=− Kk′′ sin θ k′ cos θk′ ak′ ,↑ a†k′ ,↑ + 1 a†−k′ ,↓ |0
k′′
=k′
#
= Kk′′ sin θk′ cos θ k′ a†k′ ,↑ ak′ ,↑ a†−k′ ,↓ |0
k′′
=k′
=0,
(17.251)
and similarly
b−k,↓ |Ψ0  = 0 .
Alternatively, the ground state |Ψ0 , which is given by Eq. (17.245), can
be expressed as
# 
|Ψ0  = C0 1 − γ k′ a†k′ ,↑ a†−k′ ,↓ |0 , (17.252)
k′

where C0 is a normalization constant, which is given by


#
C0 = eiφk′ cos θk′ , (17.253)
k′

and where [see Eqs. (17.225) and (17.226)]


γ k′ = e−2iφk′ tan θk′
7
8
8 1 − √ ǫk′ −ǫ2F
−2iφk′ 8 (ǫk′ −ǫF ) +|∆|2
=e 9 .
1 + √ ǫk′ −ǫ2F 2
(ǫk′ −ǫF ) +|∆|

(17.254)

Eyal Buks Quantum Mechanics - Lecture Notes 519


Chapter 17. Superconductivity
 2
Furthermore, since a†k′ ,↑ a†−k′ ,↓ = 0 the following holds
! "

|Ψ0  = C0 exp − γ k′ a†k′ ,↑ a†−k′ ,↓ |0 . (17.255)
k′

17.4.6 Pairing Wavefunction

For a general function of position γ (r′′′ ) the following holds


 
dr ′
dr′′ γ (r′′ − r′ ) Ψ↑† (r′ ) Ψ↓† (r′′ )
 
1  † ′′ ′′ ′ ′
= ak′ ,↑ a†−k′′ ,↓ dr′ dr′′ γ (r′′ − r′ ) eik ·r −ik ·r
V ′ ′′
k ,k
 †  
1 ′′ ′ ′
dr′ ei(k −k )·r
′′ ′′′
= ak′ ,↑ a†−k′′ ,↓ dr′′′ γ (r′′′ ) eik ·r
V
k′ ,k′′   
=δk′ ,k′′
 
′ ′′′
= a†k′ ,↑ a†−k′ ,↓ dr′′′ γ (r′′′ ) eik ·r ,
k′
(17.256)

where Ψσ (r ) is quantized field operators [see Eq. (16.96)]. In view of the
above result the ground state |Ψ0 , which is given by Eq. (17.255), can be
expressed as
  
′ ′′ ′′ ′ † ′ † ′′
|Ψ0  = C0 exp dr dr γ (r − r ) Ψ↑ (r ) Ψ↓ (r ) |0 , (17.257)

where the function γ (r′′ − r′ ), which is called the pairing wavefunction, sat-
isfy

′ ′′′
dr′′′ γ (r′′′ ) eik ·r = −γ k′ , (17.258)

where γ k′ is given by Eq. (17.254).


The energy region near ǫF in which γ k′ changes significantly has a char-
acteristic width given by the energy gap ∆0 [see Eq. (17.254)]. The corre-
sponding region in k′ space has thus a characteristic size given by ∆0 /vF ,
where vF is the so-called Fermi velocity (which is defined by the relation
∂ǫk′ /∂k′ = vF , where the derivative is taken at ǫk′ = ǫF ). Consequently the
paring wavefunction γ (r′′′ ) is expected to have a characteristic ’size’ given
by ξ, where
vF
ξ= , (17.259)
π |∆|
is the so-called BCS coherence length.

Eyal Buks Quantum Mechanics - Lecture Notes 520


17.5. The Josephson Effect

17.5 The Josephson Effect

Consider the global transformation ak′ ,σ → ak′ ,σ e−iΘ/2 and a†k′ ,σ → a†k′ ,σ eiΘ/2 ,
where Θ is real. Such a transformation leaved the Hamiltonian (17.208) un-
changed, however, the factor Bk′ is transformed according to Bk′ → Bk′ e−iΘ
[see Eq. (17.215)] and the energy gap ∆ is transformed according to [see Eq.
(17.220)]

∆ → ∆e−iΘ . (17.260)

Moreover, the ground state |Ψ0  is modified [see Eq. (17.252)] and becomes
|Ψ0  → |Ψ (Θ), where
#
|Ψ (Θ) = Kk′ (Θ) |0 , (17.261)
k′

where the operator Kk′ (Θ) is given by


 
Kk′ (Θ) = eiφk′ cos θk′ − eiΘ e−iφk′ sin θk′ Bk† ′ . (17.262)

As can be seen from Eq. (17.261), the vector state |Ψ (Θ) becomes identical
to the ground state |Ψ0  (17.252) when Θ = 2nπ, where n is integer. In view
of the fact that the pair creation operator Bk† ′ = a†k′ ,↑ a†−k′ ,↓ (17.215) in Eq.
(17.262) is multiplied by the factor eiΘ one may argue that the phase Θ can
be considered as the phase of Cooper pairs.

Claim. The state |Ψ (Θ) (17.263) can be alternatively expressed as

|Ψ (Θ) = e−inP Θ |Ψ0  , (17.263)

where |Ψ0  is the BCS ground state (17.252) and where


1 †
nP = a ′ ak′ ,σ (17.264)
2 ′ k ,σ
k ,σ

is the so-called pair number operator.

Proof. On one hand


 
a†k′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ Kk′ (Θ) |0
 
= −eiΘ e−iφk′ sin θ k′ a†k′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ a†k′ ,↑ a†−k′ ,↓ |0
3    4
= −eiΘ e−iφk′ sin θ k′ a†k′ ,↑ a†−k′ ,↓ 1 − a†k′ ,↑ ak′ ,↑ + a†k′ ,↑ a†−k′ ,↓ 1 − a†−k′ ,↓ a−k′ ,↓ |0
 
= −eiΘ e−iφk′ sin θ k′ a†k′ ,↑ a†−k′ ,↓ + a†k′ ,↑ a†−k′ ,↓ |0 ,
(17.265)

Eyal Buks Quantum Mechanics - Lecture Notes 521


Chapter 17. Superconductivity

and thus
# 1  † #
nP Kk′ (Θ) |0 = ak′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ Kk′′ (Θ) |0
k′
2 k′ k′′
 #
=− Kk′′ (Θ) eiΘ e−iφk′ sin θ k′ a†k′ ,↑ a†−k′ ,↓ |0 .
k′ k′′
=k′
(17.266)
On the other hand
∂ #  # ∂Kk′
−i Kk′′ (Θ) |0 = −i Kk′′ (Θ) |0 , (17.267)
∂Θ k′′ k′ k′′
=k′
∂Θ

and therefore [see Eq. (17.262)]



nP |Ψ (Θ) = −i |Ψ (Θ) , (17.268)
∂Θ
>
where |Ψ (Θ) = Kk′ (Θ) |0 [see Eq. (17.261)]. The above result together
k′
with the Taylor expansion formula for the exponential function [see Eq.
(3.31)] lead to

e−inP Θ |Ψ (0) = |Ψ (Θ) , (17.269)

which proofs the claim (17.263).

17.5.1 The Second Josephson Relation

Consider the case where a voltage V is applied to a superconductor. The


added energy of µ = eV per electron, where e is the electron charge, can be
taken into account by adding a term to the Hamiltonian of the system, which
becomes

H (V ) = HMF + 2µnP , (17.270)

where HMF is given by Eq. (17.227) and where the pair number operator nP
is given by Eq. (17.264). As will be shown below, the added term 2µnP gives
rise to time dependence of the complex energy gap ∆ [see Eq. (17.220)].
Claim. The following holds
d∆
i = 2µ∆ . (17.271)
dt
Proof. With the help of the Heisenberg equation of motion (4.38) one finds
that
d Bk′ 
i = [Bk′ , H (V )] . (17.272)
dt

Eyal Buks Quantum Mechanics - Lecture Notes 522


17.5. The Josephson Effect

thus [see Eqs. (17.211), (17.212), (17.215) and (17.227)]


d Bk′  '3 4(
i = ηk′ e−2iφk′ sin2 θk′ b†−k′ ,↓ b†k′ ,↑ − cos2 θk′ bk′ ,↑ b−k′ ,↓ , Nk′ ,↑ + N−k′ ,↓
dt
e−2iφk′ sin2 (2θk′ ) '3 4(
+µ Nk′ ,↑ + N−k′ ,↓ , bk′ ,↑ b−k′ ,↓ + b†−k′ ,↓ b†k′ ,↑
2 '3 4(
+µe −2iφk′
cos (2θk′ ) sin2 θk′ b†−k′ ,↓ b†k′ ,↑ − cos2 θk′ bk′ ,↑ b−k′ ,↓ , Nk′ ,↑ + N−k′ ,↓
'3 4(
+µe−2iφk′ sin (2θk′ ) b†−k′ ,↓ b†k′ ,↑ , bk′ ,↑ b−k′ ,↓ .
(17.273)
With the help of the commutation relations
3 4
b†−k′ ,↓ b†k′ ,↑ , Nk′ ,↑ + N−k′ ,↓ = −2b†−k′ ,↓ b†k′ ,↑ , (17.274)
3 4
b†−k′ ,↓ b†k′ ,↑ , bk′ ,↑ b−k′ ,↓ = Nk′ ,↑ + N−k′ ,↓ − 1 , (17.275)

one finds that


d Bk′ 
i = µe−2iφk′ sin (2θk′ ) N−k′ ,↓ + Nk′ ,↑ − 1 , (17.276)
dt
and therefore [see Eqs. (17.223) and (17.231)]

d Bk′ 
i = 2µ Bk′  . (17.277)
dt
Thus, the complex energy gap ∆, which is given by [see Eq. (17.220)]
g 
∆= Bk′  , (17.278)
V ′
k

satisfies Eq. (17.271).

For a fixed µ the solution of Eq. (17.271) reads

∆ (t) = ∆ (0) e−iΘ(t) , (17.279)

where the phase factor Θ (t) is given by

2µt 2eV t
Θ (t) = = . (17.280)
 
Taking the time derivative (which is denoted by overdot) yields, in agreement
with Eq. (17.97), the second Josephson relation
2eV
Θ̇ = . (17.281)


Eyal Buks Quantum Mechanics - Lecture Notes 523


Chapter 17. Superconductivity

17.5.2 The Energy of a Josephson Junction

Consider a system composed of two superconductors that are separated one


from the other by a thin insulating layer, which serves as a tunneling barrier.
The Hamiltonian of the system is assumed to be given by

H = H1 + H2 + HT , (17.282)

where H1 and H2 are the Hamiltonians of the two decoupled superconductors


and where the tunneling Hamiltonian HT is taken to be given by
  
HT = tk′ ,k′′ a†1,k′ ,↑ a2,k′′ ,↑ + a†1,−k′ ,↓ a2,−k′′ ,↓
k′ ,k′′
 
+t∗k′ ,k′′ a†2,k′′ ,↑ a1,k′ ,↑ + a†2,−k′′ ,↓ a1,−k′ ,↓ ,
(17.283)
where the annihilation operators of the first (second) superconductor are
labeled by a1,k′ ,σ (a2,k′′ ,σ ). With the help of Eq. (17.210) one finds that HT
can be expressed as

HT = Hk′ ,k′′ , (17.284)
k′ ,k′′

where
 
Hk′ ,k′′ = Ak′ ,k′′ b†1,−k′ ,↓ b†2,k′′ ,↑ − b†1,k′ ,↑ b†2,−k′′ ,↓
+A∗k′ ,k′′ (b2,k′′ ,↑ b1,−k′ ,↓ − b2,−k′′ ,↓ b1,k′ ,↑ )
 
+Bk′ ,k′′ b1,−k′ ,↓ b†2,−k′′ ,↓ + b1,k′ ,↑ b†2,k′′ ,↑
 
+Bk∗ ′ ,k′′ b2,−k′′ ,↓ b†1,−k′ ,↓ + b2,k′′ ,↑ b†1,k′ ,↑ ,
(17.285)
the coefficients Ak′ ,k′′ and Bk′ ,k′′ are given by
Ak′ ,k′′ = τ k′ ,k′′ cos θ1,k′ sin θ2,k′′ + τ ∗k′ ,k′′ sin θ1,k′ cos θ2,k′′ , (17.286)
Bk′ ,k′′ = τ k′ ,k′′ sin θ1,k′ sin θ2,k′′ − τ ∗k′ ,k′′ cos θ1,k′ cos θ2,k′′ , (17.287)
and where

τ k′ ,k′′ = ei(φ1,k′ −φ2,k′′ ) tk′ ,k′′ . (17.288)

We employ below time independent perturbation theory to calculate the


correction δE to the system’s energy to lowest nonvanishing order in the tun-
neling coefficients |tk′ ,k′′ |. The averaged total energy change δE is evaluated
by summing over all basis states of the combined system and multiplying the
energy change of each state by its thermal occupation probability. As can be

Eyal Buks Quantum Mechanics - Lecture Notes 524


17.5. The Josephson Effect

seen from Eq. (9.32) δE vanishes to first order in |tk′ ,k′′ |. To second order in
|tk′ ,k′′ | the correction δE is found to be given by
  
2 nk′ nk′′ (1 − nk′ ) (1 − nk′′ )
δE = 2 |Ak′ ,k′′ | −
ηk′ + ηk′′ η k′ + η k′′
k′ ,k′′
  
2 nk′ (1 − nk′′ ) (1 − nk′ ) nk′′
+2 |Bk′ ,k′′ | + ,
′ ′′
η k′ − ηk′′ ηk′′ − ηk′
k ,k
(17.289)
where ηk′ is given by Eq. (17.228) and nk′ is given by Eq. (17.230). With the
help of Eqs. (17.223), (17.286), (17.287) and (17.288) one finds that
2 2 
|Ak′ ,k′′ | = |τ k′ ,k′′ | cos2 θ1,k′ sin2 θ 2,k′′ + sin2 θ 1,k′ cos2 θ2,k′′
! "
1 t2k′ ,k′′ ∆∗1 ∆2
+ Re ,
2 η 1,k′ η2,k′′
(17.290)
and
 
|Bk′ ,k′′ |2 = |τ k′ ,k′′ |2 sin2 θ1,k′ sin2 θ 2,k′′ + cos2 θ1,k′ cos2 θ2,k′′
! "
1 t2k′ ,k′′ ∆∗1 ∆2
− Re .
2 η1,k′ η2,k′′
(17.291)
In what follows it will be assumed, for simplicity, that all tunneling ampli-
tudes tk′ ,k′′ are identical. Moreover, the two superconductors will be assumed
to be of the same type, i.e. |∆1 | = |∆2 | ≡ |∆|. For this case all the terms
t2k′ ,k′′ ∆∗1 ∆2 can be expressed as

t2k′ ,k′′ ∆∗1 ∆2 = T |∆|2 eiΘ , (17.292)

where T = |tk′ ,k′′ |2 and where Θ is the relative phase difference between
the two superconductors. The energy correction δE can be expressed as a
function of Θ as

δE = (δE)0 − EJ cos Θ , (17.293)

where (δE)0 is independent on Θ and where


 T |∆|2  1 − nk′′ − nk′ nk′ − nk′′

EJ = +
′ ′′
η1,k′ η2,k′′ ηk′ + ηk′′ ηk′ − ηk′′
k ,k
 T |∆|2 (1 − 2nk′′ ) ηk′ − (1 − 2nk′ ) ηk′′
= .
η1,k′ η2,k′′ η2k′ − η2k′′
k′ ,k′′
(17.294)

Eyal Buks Quantum Mechanics - Lecture Notes 525


Chapter 17. Superconductivity

Replacing the sum by an integral leads to


 ∞ ∞ βη2
 |∆|2 dǫ1 dǫ2 tanh 2 η1 − tanh βη2 1 η2
EJ = , (17.295)
πe2 RN 0 0 η1 η2 η21 − η22

where VD0 is the density of states,


.
2
ηn = ǫ2n + |∆| , (17.296)

[see Eq. (17.228)] and where RN , which is given by


RN = , (17.297)
4πe2 V 2 D02 T

is the so-called normal state resistance.


The variable transformation
ηn = |∆| cosh θn , (17.298)
ǫn = |∆| sinh θn , (17.299)
leads to
 
 |∆| β |∆|
EJ = I , (17.300)
πe2 RN 2

where the function I (x) is given by


 ∞ ∞
tanh (x cosh θ 2 ) cosh θ1 − tanh (x cosh θ1 ) cosh θ2
I (x) = dθ1 dθ 2 .
0 0 cosh2 θ1 − cosh2 θ2
(17.301)
In the limit of zero temperature the integral can be evaluated using the
variable transformation
θ1 + θ2
θp = , (17.302)
2
θ1 − θ2
θm = , (17.303)
2
which together with the identities
cosh θ1 + cosh θ2 = 2 cosh θp cosh θm , (17.304)
cosh θ1 − cosh θ2 = 2 sinh θ p sinh θm , (17.305)
 ∞

=π, (17.306)
−∞ cosh θ

lead to

Eyal Buks Quantum Mechanics - Lecture Notes 526


17.5. The Josephson Effect
 ∞ ∞
 |∆| dθ1 dθ2
EJ = 2
πe RN 0 cosh θ 1 + cosh θ 2
 ∞ 0 ∞
 |∆| dθ1 dθ2
= 2
4πe RN −∞ −∞ cosh θ1 + cosh θ2
 ∞  ∞
 |∆| dθp dθ m
= 2
,
4πe RN −∞ cosh θp −∞ cosh θm
(17.307)
thus
π |∆|
EJ = . (17.308)
4e2 RN
For arbitrary temperature the result is
π |∆| β |∆|
EJ = tanh . (17.309)
4e2 RN 2

17.5.3 The First Josephson Relation

As was shown above [see Eq. (17.293)], the energy of a Josephson junction
UJ having phase Θ relative to the energy when the phase vanishes is given
by [compare with Eq. (17.100)]

UJ = −EJ cos Θ . (17.310)

Let I (t) and V (t) be the current through and voltage across a Josephson
junction, respectively, at time t. Assume that initially at time t = 0 the
phase Θ vanishes. Energy conservation leads to the requirement that
 t
UJ = dt′ I (t′ ) V (t′ ) . (17.311)
0

With the help of the second Josephson relation Θ̇ = (2e/) V (17.281) and
Eq. (17.310) this becomes

 Θ
−EJ cos Θ = dΘ′ I (t′ ) . (17.312)
2e 0
Taking the derivative with respect to Θ leads, in agreement with Eq. (17.92),
to the first Josephson relation

I = Ic sin Θ , (17.313)

where the so-called critical current Ic is given by


2eEJ 2πcEJ
Ic = = , (17.314)
 φs

Eyal Buks Quantum Mechanics - Lecture Notes 527


Chapter 17. Superconductivity

where
hc
φs = (17.315)
2e
is the superconducting flux quantum, which is identical to the superconduct-
ing flux quantum given by Eq. (17.80) provided that the charge qs⋆ is taken to
be 2e. Note also that for the ’normal’ flux quantum φ0 given by Eq. (12.48)
the charge of elementary carrier is e.

17.6 Problems
1. Rotating Superconductor - Consider a superconductor rotating at
angular frequency Ω around the z axis. In the presence of an externally
applied magnetic field B calculate the magnetic field deep inside the
superconductor.
2. Consider a conductor containing charge carriers having charge q and
mass m. The density of charge carriers at point r is n (r) and the current
density is J (r). Contrary to the case of a normal metal, it is assumed
that all charge carriers at point r move at the same velocity v, which is
related to J by the relation [see Eq. (17.16)]
J
v= . (17.316)
qn
Show that in steady state this assumption leads to the 2nd London equa-
tion [see Eq. (17.62)]
1
∇2 H = H, (17.317)
λ2L
where H is the magnetic field and where
:
mc2
λL = . (17.318)
4πnq 2
3. Consider the so-called gradiometer RF SQUID seen in Fig. 17.4. The
junction’s critical current is labeled by Ic . It is assumed that the junc-
tion has capacitance, which is denoted by CJ . Consider the case where a
magnetic flux that is denoted by φe1 (φe2 ) is externally applied to the up-
per (lower) loop. Let Λ1 (Λ2 ) be the self inductance of the upper (lower)
loop. Derive an equation of motion for the system.
4. Consider the so-called DC SQUID device seen in Fig. 17.5. The Joseph-
son junctions on both arms of the DC SQUID have critical currents Ic1
and Ic2 respectively and both have the same capacitance CJ . The self
inductance of the loop is denoted as Λ. A bias current Ib is externally
injected and a magnetic flux φe is the externally applied to the loop. Find
equations of motion that govern the device’s dynamics.

Eyal Buks Quantum Mechanics - Lecture Notes 528


17.6. Problems

Fig. 17.4. Gradiometer RF SQUID.

Fig. 17.5. The DC SQUID.

Eyal Buks Quantum Mechanics - Lecture Notes 529


Chapter 17. Superconductivity
' (
2
5. Calculate nP  and (∆nP ) with respect to the BCS ground state |Ψ0 ,
where nP is the pairs number operator (17.264).
6. Calculate the energy density of states for elementary excitations in a
superconductor.
7. Find the time evolution of the operators ak,↑ (t) and a†−k,↓ (t) at time
t [see the Hamiltonian (17.208)] for given initial conditions ak,↑ (0) and
a†−k,↓ (0) at time t = 0.

17.7 Solutions
1. In classical mechanics a mass particle in a rotating frame experiences
a force perpendicular to its velocity called the Coriolis force. For the
present case the Coriolis force FΩ is taken to be given by

FΩ = 2m⋆s v × Ω , (17.319)

where Ω = Ωẑ is the rotation vector and where v = ṙ is the velocity


vector. Additional force perpendicular to the velocity, which is acting in

the presence of a magnetic field B, is the Lorentz force FL = qcs v × B
[see Eq. (17.4)]. From this point of view the effect of rotation can be taken
into account by replacing the magnetic field B by an effective magnetic
field Beff given by
2m⋆s c
Beff = B + Ω. (17.320)
qs⋆
With this approach Eq. (17.62) (for time independent B) becomes
 
2 1 2m⋆s c
∇ B= 2 B+ ⋆ Ω . (17.321)
λL qs

Thus the magnetic field deep inside the superconductor is given by


− (2m⋆s c/qs⋆ ) Ω.
2. The total energy of the system in steady state is given by E = T + UH ,
where

nmv2
T = dV
V 2
is the kinetic energy and where

1
UH = H2 dV (17.322)
8π V
is the magnetic energy [see Eq. (14.38)]. With the help of the Maxwell’s
equation (17.18) and Eq. (17.16) E can be expressed in terms of H as

Eyal Buks Quantum Mechanics - Lecture Notes 530


17.7. Solutions
 3 4
1
E= λ2L (∇ × H)2 + H2 dV . (17.323)
8π V
Let δH be an infinitesimally small change in H, and let δE be the corre-
sponding change in the energy. The requirement that E obtains a mini-
mum value leads to

1 , 2 -
0 = δE = λ (∇ × H) · (∇ × δH) + H · δH dV . (17.324)
4π V L

With the help of the general vector identity [see Eq. (14.41)]

∇ · (F1 × F2 ) = (∇ × F1 ) · F2 − F1 · (∇ × F2 ) , (17.325)

one finds (for the case where F1 and F2 are taken to be given by F1 =
∇ × H and F2 = δH) that

(∇ × H) · (∇ × δH) = (∇ × (∇ × H)) · δH − ∇ · (∇ × H × δH) .


(17.326)

The vector identity ∇ × (∇ × H) = ∇ (∇ · H) − ∇2 H together with the


Maxwell’s equation (17.21) lead to
 
(∇ × H) · (∇ × δH) = − ∇2 H · δH − ∇ · (∇ × H × δH) . (17.327)

The volume integral over the second term on the right hand side can be
expressed as a surface integral using the divergence theorem. However,
when boundary conditions of δH = 0 on the surfaces are applied the
surface integral vanishes. Thus Eq. (17.324) becomes

1  2 2 
0 = δE = −λL ∇ H + H · δH dV . (17.328)
4π V

The requirement that δE vanishes for arbitrary (small) δH leads to Eq.


(17.317). The assumption that B = µH [see Eq. (17.26)], i.e. the assump-
tion that the medium is isotropic and linear, implies that in steady state
Eq. (17.317) is equivalent to the 2nd London equation (17.62).
3. The total magnetic flux φ1 (φ2 ) threading the upper (lower) loop is given
by [see Eq. (17.110) and Fig. 17.6]
φ1 = φe1 + Λ1 Is1 , (17.329)
φ2 = φe2 + Λ2 Is2 , (17.330)
where φ1 (φ2 ) is the total magnetic flux in the upper (lower) loop, Is1 (Is2 )
is the circulating current flowing in the upper (lower) loop and mutual
inductance between the loops is disregarded. The requirement that the
phase of the macroscopic wavefunction is continues in both the upper
and lower loops yields the following relations [see Eq. (17.107)]

Eyal Buks Quantum Mechanics - Lecture Notes 531


Chapter 17. Superconductivity

2πφ1
Θ+ = 2n1 π , (17.331)
φs
2πφ2
−Θ + = 2n2 π , (17.332)
φs
where Θ is the gauge invariant phase difference across the junction, φs is
the flux quantum and where both n1 and n2 are integers. The Lagrangian
of the system can be expressed as a function of the dimensionless flux
coordinates Φ, which is defined by [see Eqs. (17.331) and (17.332)]
   
φ1 φ2
Φ = 2π − n1 = −2π − n2 = −Θ , (17.333)
φs φs

and its time derivative Φ̇ [see Eq. (17.117)]

CJ φ2s Φ̇2 φ2s (Φ − Φe1 )2 φ2s (Φ + Φe2 )2 φs Ic


L= − − + cos Φ , (17.334)
8π2 c2 8π 2 cΛ1 8π2 cΛ2 2πc
where
2πφe1
Φe1 = − 2πn1 , (17.335)
φs
2πφe2
Φe2 = − 2πn2 , (17.336)
φs
are the normalized external fluxes. Using the notation
1 1 1
= + , (17.337)
Λ0 Λ1 Λ2
Λ0 Φe1 Λ0 Φe2
Φe0 = − , (17.338)
Λ1 Λ2
the Lagrangian can be expressed as

CJ φ2s Φ̇2 φ2s (Φ − Φe0 )2 φs Ic


L= − + cos Φ + CG , (17.339)
8π2 c2 8π2 cΛ0 2πc
where the constant CG is given by
 
φ2s Λ 0
Λ1 Φ 2
e1 + Λ0 2
Λ2 Φe2 − Φ 2
e0
CG = − . (17.340)
8π 2 cΛ0
The resulting Euler - Lagrange equation of motion (1.8) is given by
 
d ∂L ∂L
= , (17.341)
dt ∂ Φ̇ ∂Φ

thus

Eyal Buks Quantum Mechanics - Lecture Notes 532


17.7. Solutions

Fig. 17.6. Gradiometer RF SQUID.

CJ φs Φ̈ φ (Φ − Φe0 )
=− s − Ic sin Φ . (17.342)
2πc 2πΛ0
Note that [see Eqs. (17.329) and (17.330)]

2πΛ0 (Is1 − Is2 )


Φ − Φe0 = , (17.343)
φs

thus with the help of Eqs. (17.97) and (17.333) the equation of motion
(17.342) can be expressed as a current conservation law [compare with
Eq. (17.125)]

Is1 − Is2 = Ic sin Θ + CJ V̇ , (17.344)

where V is the voltage across the Josephson junction.


4. The total self inductance Λ of the loop (which is assumed to be equally
divided between both arms) is taken into account by adding an inductor
having inductance Λ/2 to each arm (see Fig 17.7). The total magnetic
flux threading the DC SQUID loop is given by
ΛIs
φ = φe + , (17.345)
2
where φe is the externally applied flux,

Eyal Buks Quantum Mechanics - Lecture Notes 533


Chapter 17. Superconductivity

Is = I1 − I2 (17.346)

is the circulating current in the loop, and where I1 and I2 are the to-
tal currents flowing in the upper and lower arms respectively. The bias
current is given by Ib = I1 + I2 . In general the Josephson current IJk in
each junction (k = 1, 2) is related to the critical current Ick and to the
Josephson phase γ k by [see Eq. (17.92)]

IJk = Ick sin γ k . (17.347)

The voltage across each junction is given by [see (17.97)]


φs
Vk = γ̇ , (17.348)
2πc k
where φs is the flux quantum [see Eq. (17.80)]. The condition that the
phase around the loop is single-valued reads
2πφ
2πn = γ 1 − γ 2 + , (17.349)
φs
where n is integer [see Eq. (17.107)]. By using this and Eq. (17.345) one
finds that
 
φs 2πφe
Is = − γ1 − γ2 + − 2πn . (17.350)
πΛ φs
The total voltage across the DC SQUID, which is denoted as VS is given
by
Λ dI1 Λ dI2
VS = V1 + = V2 + , (17.351)
2c dt 2c dt
and therefore
 
1 Λ dI1 dI2
VS = V1 + V2 + +
2 2c dt dt

1 d φs ΛIb
= (γ + γ 2 ) + .
2 dt 2π 1 2c
(17.352)

To model the current biasing an inductor having inductance LB is added


in parallel to the SQUID (see Fig. 17.7). As will be shown below, in the
limit where LB ≫ Λ the added inductor allows modeling current biasing.
The requirement that the voltage across the DC SQUID is the same as
the one across the inductor LB yields [see Eq. (17.352)]

dφB 1 d φs ΛIb
= (γ 1 + γ 2 ) + , (17.353)
dt 2 dt 2π 2c

Eyal Buks Quantum Mechanics - Lecture Notes 534


17.7. Solutions

where φB is the magnetic flux in the inductor LB . In what follows the bias
current Ib is assumed to be a constant. Integrating the above equation
yields
φs
φB = (γ + γ 1 ) + φB0 , (17.354)
4π 1
where the constant φB0 is taken to be given by φB0 = −LB Ib . The
Lagrangian of the closed system L = T − U [see Eq. (1.16)] is expresses
as a function of the coordinates γ 1 and γ 2 and their time derivatives γ̇ 1
and γ̇ 2 , where the kinetic energy T is given by
 
CJ φ2s γ̇ 21 + γ̇ 22
T = , (17.355)
8π 2 c2
and the potential energy U by [see Eq. (17.350)]

φs (Ic1 cos γ 1 + Ic2 cos γ 2 )


U =−
2πc
 2
φs γ 1 − γ 2 + 2πφ
2
φ
e
− 2πn φ2B
s
+ + .
8π2 cΛ 2cLB
(17.356)

The assumption that LB ≫ Λ together with the assumption that


φs (γ 1 + γ 1 ) /4π ≪ LB Ib [see Eq. (17.354)] implies that the term φ2B /2cLB
can be replaced by −φs (γ 1 + γ 2 ) Ib /4πc (note that constant terms in U
can be disregarded). The Euler-Lagrange set of two equations [see Eq.
(1.8)]
 
d ∂L ∂L
= , (17.357)
dt ∂ γ̇ 1 ∂γ 1
 
d ∂L ∂L
= , (17.358)
dt ∂ γ̇ 2 ∂γ 2

yields
 
2 2πφe
CJ φ2s γ̈ 1 φs Ic1 sin γ 1 φs γ 1 − γ 2 + φs − 2πn φ Ib
2 2
=− − 2
+ s ,
4π c 2πc 4π cΛ 4πc
(17.359)
 
2 2πφe
CJ φ2s γ̈ 2 φs Ic2 sin γ 2 φs γ 1 − γ 2 + φs − 2πn φ Ib
=− + + s ,
4π 2 c2 2πc 4π 2 cΛ 4πc
(17.360)

or [see Eqs. (17.348) and (17.350)]

Eyal Buks Quantum Mechanics - Lecture Notes 535


Chapter 17. Superconductivity

Fig. 17.7. The DC SQUID. The inductor LB is added to model current biasing.

dV1
Ic1 sin γ 1 + CJ = I1 , (17.361)
dt
dV2
Ic2 sin γ 2 + CJ = I2 . (17.362)
dt

5. The following holds [see Eqs. (17.245) and (17.247)]

Eyal Buks Quantum Mechanics - Lecture Notes 536


17.7. Solutions

Ψ0 | a†k′′ ,σ ak′′ ,σ |Ψ0 


# †
= 0| Kk′ Kk′ Kk† ′′ a†k′′ ,σ ak′′ ,σ Kk′′ |0
k′
=k′′
   
= 0| e−iφk′′ cos θk′′ − eiφk′′ sin θ k′′ a−k′′ ,↓ ak′′ ,↑ a†k′′ ,σ ak′′ ,σ eiφk′′ cos θ k′′ − e−iφk′′ sin θk′′ a†k′′ ,↑ a†−k′′ ,↓ |0
= sin2 θ k′′ 0| a−k′′ ,↓ ak′′ ,↑ a†k′′ ,σ ak′′ ,σ a†k′′ ,↑ a†−k′′ ,↓ |0
= sin2 θ k′′ ,
(17.363)
thus

Ψ0 | nP |Ψ0  = sin2 θk′ . (17.364)
k′

 2
Similarly, since a†k′′ ,σ ak′′ ,σ = a†k′′ ,σ ak′′ ,σ one finds that
1    
Ψ0 | n2P |Ψ0  = Ψ0 | a†k′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ a†k′′ ,↑ ak′′ ,↑ + a†−k′′ ,↓ a−k′′ ,↓ |Ψ0 
4 ′ ′′
k ,k
1
= Ψ0 | a†k′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ + 2a†k′ ,↑ ak′ ,↑ a†−k′ ,↓ a−k′ ,↓ |Ψ0 
4 ′
k
1    
+ Ψ0 | a†k′ ,↑ ak′ ,↑ + a†−k′ ,↓ a−k′ ,↓ a†k′′ ,↑ ak′′ ,↑ + a†−k′′ ,↓ a−k′′ ,↓ |Ψ0 
4 ′ ′′
k
=k
1 2 
= sin θ k′ + sin4 θk′ + sin2 θk′ sin2 θ k′′ ,
2 ′ ′ ′′
k k
=k
(17.365)
thus
Ψ0 | (∆nP )2 |Ψ0  = Ψ0 | n2P |Ψ0  − (Ψ0 | nP |Ψ0 )2
1 2  
= sin θ k′ + sin4 θk′ + sin2 θk′ sin2 θ k′′ − sin2 θk′ sin2 θ k′′
2 ′ ′ ′′ ′ ′′
k k
=k k ,k
1 2  
= sin θ k′ 1 − sin2 θk′
2 ′
k
1 2
= sin θ k′ cos2 θk′ .
2 ′
k

6. With the help of Eqs. (16.102) and (17.228) one finds that the density of
states D (ǫ) per unit volume (volume is labeled by V) is given by [compare
with Eq. (16.103)]

Eyal Buks Quantum Mechanics - Lecture Notes 537


Chapter 17. Superconductivity

1
D (ǫ) = δ (ǫ − ηk′ )
V ′
k
∞  . 
1 2V ′ ′2 2 2
= 4π dk k δ ǫ − (ǫk′ − ǫF ) + |∆| .
V 8π3
0
(17.366)
Assuming that the energy ǫk′ of an electron having wave vector k′ is
given by [see Eq. (16.97)]
2 k′2
ǫk′ = , (17.367)
2m
one finds that
∞ 0  . 
′ ǫ′ ′ 2 2
D (ǫ) = DF dǫ δ ǫ − (ǫ − ǫF ) + |∆|
ǫF
0
 . 1/2
ǫ2 − |∆|2 ǫ
= DF 1 +  . ,
ǫF
ǫ2 − |∆|2
(17.368)
where
21/2 m3/2 √
DF = ǫF (17.369)
π2 3
is the normal phase density of states per unit volume at the Fermi energy,
which is labeled by ǫF . For the case where ǫ ≪ ǫF and |∆| ≪ ǫF one has
ǫ
D (ǫ) = DF . . (17.370)
2 2
ǫ − |∆|

7. The time evolution of the operators bk,↑ (t) and b†−k,↓ (t) is governed by
[see Eqs. (4.37) and (17.227)]
dbk,↑  3 4
= −i−1 ηk′ bk,↑ , b†k′ ,σ bk′ ,σ , (17.371)
dt ′ k ,σ

db†−k,↓  3 4
= −i−1 ηk′ b†−k,↓ , b†k′ ,σ bk′ ,σ , (17.372)
dt
k′ ,σ

thus
dbk,↑ 3 4
= −i−1 ηk bk,↑ , b†k,↑ bk,↑ , (17.373)
dt
† 3 4
db−k,↓
= −i−1 η−k b†−k,↓ , b†−k,↓ b−k,↓ . (17.374)
dt

Eyal Buks Quantum Mechanics - Lecture Notes 538


17.7. Solutions

With the help of the identity (16.69) one finds that


dbk,↑
= −i−1 ηk bk,↑ , (17.375)
dt
db†−k,↓
= i−1 η−k b†−k,↓+ , (17.376)
dt
thus
  ! −1
" 
bk,↑ (t) e−i ηk t
0 bk,↑ (0)
= . (17.377)
b†−k,↓ (t) b†−k,↓ (0)
−1
0 ei ηk t

The transformation (17.210), according to which


   
ak,↑ bk,↑
† = MB † , (17.378)
a−k,↓ b−k,↓

where
 
e−iφk cos θk −e−iφk sin θk
MB = ,
eiφk sin θk eiφk cos θk

leads to
  ! −1
"  
ak,↑ (t) e−i ηk t 0 −1 ak,↑ (0)
= MB MB , (17.379)
a†−k,↓ (t) a†−k,↓ (0)
−1
0 ei ηk t

thus [see Eqs. (17.223), (17.224) and (17.228)]


 
   
cos ηk t − i ǫkη′ −ǫ sin ηk t
ηk t
ak,↑ (t)
F
i |∆| e−2iφk
sin ak,↑ (0)
=
′k  η k′    .
a†−k,↓ (t) i |∆|
η ′e
2iφk
sin ηk t cos ηk t + i sin ηk t ǫkη′ −ǫ F a†−k,↓ (0)
k ′ k

(17.380)

Eyal Buks Quantum Mechanics - Lecture Notes 539


18. Open Quantum Systems

This chapter is mainly based on the book [7].

18.1 Classical Resonator


Consider a classical mechanical resonator having mass m and resonance fre-
quency ω 0 . The resonator is driven by an external force Fex that is given
by
 
Fex = F0 cos (ω p t) = F0 Re e−iωp t , (18.1)
where F0 is a real constant. The equation of motion is given by
mẍ + mω20 x = Fex . (18.2)
In steady state we seek a solution having the form
 
x = Re Ae−iωp t , (18.3)
where A is a complex constant. Substituting such a solution into the equation
of motion (18.2) yields
1 F0
A= 2 . (18.4)
m ω0 − ω 2p
This result is clearly nonphysical since it diverges at resonance ω p = ω0 . This
can be fixed by introducing a damping term in the equation of motion
mẍ + mγ ẋ + mω 20 x = Fex , (18.5)
where γ is the damping rate. For this case the steady state amplitude becomes
finite for any driving frequency
1 F0
A= . (18.6)
m ω20 − ω 2p − iωp γ

However, also (18.5) is a nonphysical equation of motion. The equipartition


theorem of classical statistical mechanics predicts that in equilibrium at tem-
perature T the following holds
Chapter 18. Open Quantum Systems

) 2 * kB T
x = . (18.7)
mω20

However, as can be seen from Eq. (18.5), when F0 = 0 the steady state
solution is given by x (t) = 0, contradicting thus the equipartition theorem.
This can be fixes by introducing yet another term f (t) in the equation of
motion representing fluctuating force

mẍ + mγ ẋ + mω 20 x = f (t) + Fex . (18.8)

The fluctuating
) * force has vanishing mean f (t) = 0, however its variance
is finite f 2 (t) > 0. In exercise 1 below the autocorrelation function of the
fluctuating force f (t) is found to be given by (18.153)

f (t) f (t + t′ ) = 2mγkB T δ (t′ ) . (18.9)

Similarly to the classical case, also in the quantum case nonphysical be-
havior is obtained when damping is disregarded. This happens not only for
the above discussed example of a driven resonator. For example, recall that
for a general quantum system driven by a periodic perturbation the time
dependent perturbation theory predicts in the long time limit constant rates
of transition between states [e.g., see Eq. (10.38)]. Such a prediction can
yield correct steady state population of quantum states only when damping
is taken into account.
Damping and fluctuation in a quantum system can be taken into account
by introducing a thermal bath, which is assumed to be weakly coupled to the
system under study. Below this technique is demonstrated for two cases. In
the first one, the system under study (also referred to as the closed system)
is a mechanical resonator, and in the second one it is taken to be a two level
system. In both cases the open system is modeled by assuming that the closed
system is coupled to a thermal bath in thermal equilibrium.

18.2 Quantum Resonator Coupled to Thermal Bath

Consider a mechanical resonator having mass m and resonance frequency ω 0 .


The resonator is coupled to a thermal bath, which is modeled as an ensemble
of harmonic oscillators.

18.2.1 The closed System

First, we consider the isolated resonator. The Hamiltonian is given by [see


Eqs. (5.9), (5.10), (5.11), (5.12), (5.13) and (5.16)]

Eyal Buks Quantum Mechanics - Lecture Notes 542


18.2. Quantum Resonator Coupled to Thermal Bath

p2 1
H0 = + mω20 x2
2m  2 
† 1
= ω0 a a + ,
2
(18.10)
where
0  
mω0 ip
a= x+ , (18.11)
2 mω 0
0  
mω0 ip
a† = x− , (18.12)
2 mω 0
and where
, †-
a, a = 1 . (18.13)

The inverse transformation is


0
  † 
x= a +a , (18.14)
2mω 0
0
mω 0  † 
p=i a −a . (18.15)
2

18.2.2 Coupling to Thermal Bath

Damping is taken into account using a model containing a reservoir of har-


monic oscillators interacting with the resonator. The total Hamiltonian is
given by

Ht = H0 + Hr + V , (18.16)

where H0 is given by Eq. (18.10), Hr is the Hamiltonian of the thermal bath,


which is assumed to be a dense ensemble of harmonic oscillators
  
1
Hr = ω k b†k bk + , (18.17)
2
k

and V is a coupling term


  ∗
V = a λk b†k + a†  λk bk , (18.18)
k k

where λk are coupling constants. The bath operators satisfy regular harmonic
oscillator commutation relations
3 4 , - 3 4
[a, bk ] = a, b†k = a† , bk = a† , b†k = 0 , (18.19)

Eyal Buks Quantum Mechanics - Lecture Notes 543


Chapter 18. Open Quantum Systems
3 4
bk , b†l = δ k,l , (18.20)
and
3 4
[bk , bl ] = b†k , b†l = 0 . (18.21)

Exercise 18.2.1. Show that


ȧ + (iω 0 + γ) a = F (t) , (18.22)
where γ is a constant and where the fluctuating term F (t) is given by
 ∗
F (t) = −i λk exp (−iω k t) bk (0) . (18.23)
k

Solution 18.2.1. In general, the Heisenberg equation of motion of an oper-


ator O is given by Eq. (4.37)
i ∂O
Ȯ = − [O, H] + . (18.24)
 ∂t
Using Eq. (18.24) one finds
 ∗
ȧ = −iω0 a − i λ k bk , (18.25)
k

ȧ† = iω0 a† + i λk b†k , (18.26)
k

ḃk = −iω k bk − iλk a , (18.27)


and
ḃ†k = iωk b†k + iλ∗k a† . (18.28)
The solution of Eq. (18.27) is given by
bk (t) = exp [−iω k (t − t0 )] bk (t0 )
 t
− iλk dt′ exp [−iω k (t − t′ )] a (t′ ) .
t0
(18.29)
Choosing the initial time to be given by t0 = 0 and substituting Eq. (18.29)
into Eq. (18.25) yield
 t 
ȧ + iω 0 a + dt′ a (t′ ) |λk |2 exp [−iωk (t − t′ )]
0 k

= −i λ∗k exp (−iω k t) bk (0) .
k
(18.30)

Eyal Buks Quantum Mechanics - Lecture Notes 544


18.2. Quantum Resonator Coupled to Thermal Bath

The states of the thermal bath are assumed to be very dense, thus one can
replace the sum over k with an integral
 2
|λk | exp [−iω k (t − t′ )]
k
 ∞
2
≃ dΩ |λ (Ω)| exp [−iΩ (t − t′ )] ,
−∞
(18.31)

where λ (Ω) is the density of states. Assuming λ (Ω) is a smooth function


near Ω = ω 0 one finds that
 t 
dt′ a (t′ ) |λk |2 exp [−iωk (t − t′ )]
0 k
 t  ∞
′ ′ 2
⋍ dt a (t ) |λ (ω 0 )| dΩ exp [−iΩ (t − t′ )]
0 −∞
  
2πδ(t−t′ )
2
= π |λ (ω 0 )| a (t) .
(18.32)

Thus using the notation

γ = π |λ (ω 0 )|2 , (18.33)

one has

ȧ + (iω 0 + γ) a = F (t) , (18.34)

ȧ† + (−iω 0 + γ) a† = F † (t) , (18.35)


where

F (t) = −i λ∗k exp (−iω k t) bk (0) , (18.36)
k

F † (t) = i λk exp (iω k t) b†k (0) . (18.37)
k

The fluctuation terms F (t) and F † (t) represent noisy force acting on the
resonator.

From Eqs. (18.34), (18.35), (18.14), and (18.15) one finds that

ṗ + γp + mω 20 x = f (t) , (18.38)

where

Eyal Buks Quantum Mechanics - Lecture Notes 545


Chapter 18. Open Quantum Systems
0
mω 0 , † -
f (t) = i F (t) − F (t) . (18.39)
2
In classical mechanics the momentum p is given by p = mẋ. Using this
substitution the equation of motion for the quantum operator p (18.38) takes
a form analogues to the classical equation of motion of a mechanical resonator
having damping rate γ and influenced by a force f (t)

mẍ + mγ ẋ + mω 20 x = f (t) . (18.40)

18.2.3 Thermal Equilibrium

Exercise 18.2.2. Show that


) † *
F (t) F (t + t′ ) = 2γ n̂0 δ (t′ ) , (18.41)
) *
F (t) F † (t + t′ ) = 2γ (n̂0 + 1) δ (t′ ) , (18.42)
and
) *
F (t) F (t + t′ ) = F † (t) F † (t + t′ ) = 0 , (18.43)

where
1
n̂0 = , (18.44)
eβω0 −1
and where β = 1/kB T .
Solution 18.2.2. The modes of the thermal bath are assumed to be in ther-
mal equilibrium. In general, thermal averaging of an operator Ok , associated
with mode #k in the thermal bath, is given by [see Eqs. (8.8) and (8.42)]

Ok  = Tr (ρk Ok ) , (18.45)

where the density operator ρk is given by


1 −βHr,k
ρk = e , (18.46)
Z
where
 
Z = Tr e−βHr,k , (18.47)
 
1
Hr,k = ω k b†k bk + , (18.48)
2
and β = 1/kB T . Using these expressions one finds that [see Eq. (8.152)]
' ( 1
b†k (t) bk (t) = βωk ≡ n̂k . (18.49)
e −1

Eyal Buks Quantum Mechanics - Lecture Notes 546


18.2. Quantum Resonator Coupled to Thermal Bath

Using Eq. (18.20) one finds that


' (
bk (t) b†k (t) = n̂k + 1 . (18.50)

In a similar way one also finds that


' ( ) * ' (
bk  = b†k = b2k = b†2 k =0. (18.51)

Moreover, using the full bath Hamiltonian Hr one can easily show that
' (
bk bl  = b†k b†l = 0 , (18.52)
' (
b†k (t) bl (t) = δ kl n̂k , (18.53)

and
' (
bk (t) b†l (t) = δ kl (n̂k + 1) . (18.54)

The fluctuating forces are given by Eqs. (18.36) and (18.37). We calculate
below some correlation functions of these forces. Using Eq. (18.51) one finds
) *
F (t) = F † (t) = 0 . (18.55)

Using Eq. (18.53) one finds that


) † * 
F (t) F (t + t′ ) = |λk |2 exp (−iω k t′ ) n̂k . (18.56)
k

Replacing the sum over k with an integral, as in Eq. (18.31), and taking into
account only modes that are nearly resonant with the cavity mode one finds
) † *
F (t) F (t + t′ ) = 2γ n̂0 δ (t′ ) , (18.57)

where
1
n̂0 = . (18.58)
eβω0 − 1
Similarly
) *
F (t) F † (t + t′ ) = 2γ (n̂0 + 1) δ (t′ ) , (18.59)

and
) *
F (t) F (t + t′ ) = F † (t) F † (t + t′ ) = 0 . (18.60)
) † *
Exercise 18.2.3. Show that the expectation value a a in steady state is
given by
) † *
a a = n̂0 . (18.61)

Eyal Buks Quantum Mechanics - Lecture Notes 547


Chapter 18. Open Quantum Systems

Solution 18.2.3. Multiplying Eq. (18.34) by the integration factor e(iω0 +γ)t
yields
d  (iω0 +γ)t 
ae = F (t) e(iω0 +γ)t . (18.62)
dt
The solution is given by
 t

a (t) = a (t0 ) e(iω0 +γ)(t0 −t)
+ dt′ F (t′ ) e(iω0 +γ)(t −t) . (18.63)
t0

Steady state is established when γ (t − t0 ) ≫ 1. In this limit the first term


becomes exponentially small (recall that γ is positive), i.e. effect of initial
condition on the value of a at time t0 becomes negligible. Thus in steady
state the solution becomes
 t

a (t) = dt′ F (t′ ) e(iω0 +γ)(t −t) , (18.64)
t0

and the Hermitian conjugate is given by


 t


a (t) = dt′ F † (t′ ) e(−iω0 +γ)(t −t) . (18.65)
t0

With the help of Eq. (18.57) one finds that


 t  t
) † * ) * ′′ ′
a a = dt′ dt′′ F † (t′′ ) F (t′ ) e(−iω0 +γ)(t −t) e(iω0 +γ)(t −t)
t0 t0
 t

= 2γn̂0 dt′ e2γ (t −t)
t
 0 
= n̂0 1 − e−2γ(t−t0 ) .
(18.66)
The assumption γ (t − t0 ) ≫ 1 allows writing this result as
) † *
a a = n̂0 . (18.67)
) † *
The last result a a = n̂0 verifies that the resonator reached thermal
equilibrium in steady state. Similarly, the next exercise shows that in the
classical limit the equipartition theorem of classical statistical mechanics is
satisfied.
) *
Exercise 18.2.4. Calculate x2 in steady state.
, -
Solution 18.2.4. According to Eq. (18.14) and a, a† = 1 the following
holds

Eyal Buks Quantum Mechanics - Lecture Notes 548


18.3. Two Level System Coupled to Thermal Bath

) 2*  ) †  *
x = a + a a† + a
2mω0
 ) *
= a†2 + a2 + a† a + aa†
2mω0
 ) *
= a†2 + a2 + 2a† a + 1 .
2mω0
(18.68)
) †2 * ) 2 *
As can be seen from Eq. (18.60), a = a = 0. Thus, with the help of
Eq. (18.67) one has
) 2* 
x = (2n̂0 + 1)
2mω0
 βω 0
= coth ,
2mω0 2
(18.69)
in agreement with Eq. (8.160). In the classical limit where kB T ≫ ω0 one
has
) 2 * kB T
x = , (18.70)
mω20
in agreement with the classical equipartition theorem.

18.3 Two Level System Coupled to Thermal Bath


In this section we discuss a two level system (TLS) coupled to thermal baths,
and obtain the Bloch equations.

18.3.1 The Closed System


The Hamiltonian Hq of the closed system can be represented by a 2×2 matrix
[see Eq. (15.36)], which in turn can be expressed in terms of Pauli matrices
(6.136)

Hq =
˙ Ω (t) · σ , (18.71)
2
where Ω (t) is a 3D real vector, and where the components of the Pauli matrix
vector σ are given by
     
01 0 −i 1 0
σx = , σy = , σz = . (18.72)
10 i 0 0 −1
Let P = σ be the vector of expectation values P = (σ x  , σy  , σz ). We
refer to this vector as the polarization vector. With the help of Eq. (4.38),
which is given by

Eyal Buks Quantum Mechanics - Lecture Notes 549


Chapter 18. Open Quantum Systems
5 6
d A 1 ∂A
= [A, H] + , (18.73)
dt i ∂t

and Eq. (6.137) one finds that


dPz 1
= [σ z , Hq ]
dt i
1
= (Ωx [σ z , σx ] + Ωy [σz , σy ])
2i
= (Ωx σ y  − Ωy σx )
= (Ω × P) · ẑ .
(18.74)
Similar expressions are obtained for Px and Py that together can be written
in a vector form as [see also Eq. (6.162)]

dP
= Ω (t) × P . (18.75)
dt
The time varying ’effective magnetic field’ Ω (t) is taken to be given by

Ω (t) = ω 0 ẑ + ω 1 (t) . (18.76)

While ω 0 , which is related to the energy gap ∆ separating the TLS states by
ω 0 = ∆/, is assumed to be stationary, the vector ω1 (t) is allowed to vary
in time, however, it is assumed that |ω 1 (t)| ≪ ω0 .

18.3.2 Coupling to Thermal Baths

As we did in the previous section, damping is taken into account using a model
containing reservoirs having dense spectrum of oscillator modes interacting
with the TLS. Furthermore, since the ensembles are assumed to be dense,
summation over modes is done with continuos integrals. The Hamiltonian H
of the entire system is taken to be given by

H = Hq

+ dω ωa†1 (ω) a1 (ω)

+ dω ωa†2 (ω) a2 (ω)
 0  
Γ1 iφ1
+  dω e σ + a1 (ω) + e−iφ1 a†1 (ω) σ −

 0
Γϕ  iφ2 
+  dω e σz a2 (ω) + e−iφ2 a†2 (ω) σ z ,

(18.77)

Eyal Buks Quantum Mechanics - Lecture Notes 550


18.3. Two Level System Coupled to Thermal Bath

where Hq is the Hamiltonian for the closed TLS,


   
01 00
σ+ = , σ− = , (18.78)
00 10

and the real coupling parameters Γ1 , Γϕ , φ1 and φ2 are assumed to be fre-


quency independent. The bath modes are boson modes satisfying the usual
Bose commutation relations
3 4
ai (ω) , a†i (ω ′ ) = δ (ω − ω ′ ) , (18.79)
[ai (ω) , ai (ω ′ )] = 0 , (18.80)

where i = 1, 2. While the coupling to the first bath (with coupling constant
Γ1 ) gives rise to TLS decay through spin flips, the coupling to the second
bath (with coupling constant Γϕ ) gives rise to pure dephasing.

Exercise 18.3.1. Show that


dσz 1
= [σz , Hq ] − Γ1 (1 + σ z )
dt i
2  
+ −iσ+ V1 + iV1† σ− ,

(18.81)

and
 
dσ+ 1 Γ1
= [σ+ , Hq ] − + Γϕ σ +
dt i 2
i3 †  4
+ −V1 σz + 2σ + Vϕ + Vϕ† ,

(18.82)

where
0 
Γ1 iφ1
V1 =  e dωe−iω(t−t0 ) a1 (t0 , ω) , (18.83)

and
0 
Γϕ iφ2
Vϕ =  e dωe−iω(t−t0 ) a2 (t0 , ω) . (18.84)

Solution 18.3.1. With the help of the identities

[σz , σ + ] = 2σ + , (18.85)
[σz , σ − ] = −2σ− , (18.86)
[σ+ , σ − ] = σz , (18.87)

Eyal Buks Quantum Mechanics - Lecture Notes 551


Chapter 18. Open Quantum Systems

one finds that the Heisenberg equation of motion (4.38) for σ z is given by
dσz 1
= [σz , Hq ]
dt i0

Γ1
− 2i dω eiφ1 σ + a1 (ω)

0 
Γ1
+ 2i dω e−iφ1 a†1 (ω) σ − ,

(18.88)
for σ + by
dσ+ 1
= [σ+ , Hq ]
dt i
0 
Γ1
−i dω e−iφ1 a†1 (ω) σ z

0 
Γϕ
+ 2i dω eiφ2 σ + a2 (ω)

0 
Γϕ
+ 2i dω e−iφ2 a†2 (ω) σ+ ,

(18.89)
for a1 (ω) by
0
da1 (ω) Γ1 −iφ1
= −iωa1 (ω) − i e σ− , (18.90)
dt 2π
and for a2 (ω) by
0
da2 (ω) Γϕ −iφ2
= −iωa2 (ω) − i e σz . (18.91)
dt 4π
Integrating the equations of motion for the bath operators a1 (ω) and a2 (ω)
yields
a1 (ω) = e−iω(t−t0 ) a1 (t0 , ω)
0 
Γ1 −iφ1 t ′ −iω(t−t′ )
−i e dt e σ− (t′ ) ,
2π t0
(18.92)
and
a2 (ω) = e−iω(t−t0 ) a2 (t0 , ω)
0 
Γϕ −iφ2 t ′ −iω(t−t′ )
−i e dt e σ z (t′ ) .
4π t0
(18.93)

Eyal Buks Quantum Mechanics - Lecture Notes 552


18.3. Two Level System Coupled to Thermal Bath

We now substitute these results into the Eqs. (18.88) and (18.89) and make
use of the following relations


dω e−iω(t−t ) = 2πδ (t − t′ ) , (18.94)

 t
1
δ (t − t′ ) f (t′ ) dt′ = sgn (t − t0 ) f (t) . (18.95)
t0 2
where sgn(x) is the sign function
+
+1 if x > 0
sgn (x) = , (18.96)
−1 if x < 0.

to obtain
dσz 1
= [σz , Hq ]
dt i0

Γ1
− 2i dωeiφ1 σ + e−iω(t−t0 ) a1 (t0 , ω)

− Γ1 σ + σ−
0 
Γ1
+ 2i dωe−iφ1 eiω(t−t0 ) a†1 (t0 , ω) σ−

− Γ1 σ + σ− ,
(18.97)

and
dσ+ 1
= [σ+ , Hq ]
dt i
0 
Γ1
−i dωe−iφ1 eiω(t−t0 ) a†1 (t0 , ω) σz

Γ1
+ σ + σz
20

Γϕ
+ 2i dωeiφ2 σ+ e−iω(t−t0 ) a2 (t0 , ω)

Γϕ
+ σ+ σ z
20

Γϕ
+ 2i dωe−iφ2 eiω(t−t0 ) a†2 (t0 , ω) σ+

Γϕ
− σz σ + .
2
(18.98)

Thus, by making use of the following relations

Eyal Buks Quantum Mechanics - Lecture Notes 553


Chapter 18. Open Quantum Systems

1
σ+ σ − = (1 + σ z ) , (18.99)
2
1
σ− σ + = (1 − σ z ) , (18.100)
2
σz σ + = −σ + σz = σ + , (18.101)

one finds that


dσz 1
= [σ z , Hq ] − Γ1 (1 + σz )
dt i
2 
+ −iσ+ V1 + iV1† σ− ,

(18.102)
and
 
dσ+ 1 Γ1
= [σ + , Hq ] − + Γϕ σ+
dt i 2
i3 †  4
+ −V1 σ z + 2 σ+ Vϕ + Vϕ† σ + .

(18.103)

18.3.3 Thermal Equilibrium

Using Eq. (18.51) one finds


' ( ) *
V1  = V1† = Vϕ  = Vϕ† = 0 . (18.104)

Using Eqs. (18.53) and (18.54), the relation




dω e−iω(t−t ) = 2πδ (t − t′ ) , (18.105)

and assuming the case where the dominant contribution to the TLS dynamics
comes from the bath modes near frequency ω0 (recall that ω 0 = ∆/, where
∆ is the energy gap separating the TLS states), one finds that
' (
V1† (t′ ) V1 (t)
  ' (
Γ1 ′ ′
= 2 dω dω′ e−iω (t−t ) a†1 (t0 , ω) a1 (t0 , ω ′ )


Γ1 ′
= 2 dωe−iω(t−t ) n (ω)

≃ 2 Γ1 n̂0 δ (t − t′ ) ,
(18.106)

where n̂0 is given by [see Eq. (18.58)]

Eyal Buks Quantum Mechanics - Lecture Notes 554


18.3. Two Level System Coupled to Thermal Bath

1
n̂0 = . (18.107)
eβω0 − 1
Similarly
' (
V1 (t) V1† (t′ ) = 2 Γ1 (n̂0 + 1) δ (t − t′ ) , (18.108)

) † ′ * Γϕ
Vϕ (t ) Vϕ (t) = 2 n̂0 δ (t − t′ ) , (18.109)
2
) * Γϕ
Vϕ (t) Vϕ† (t′ ) = 2 (n̂0 + 1) δ (t − t′ ) , (18.110)
2
and
' (
V1 (t′ ) V1 (t) = V1† (t′ ) V1† (t)
) *
Vϕ (t′ ) Vϕ (t) = Vϕ† (t′ ) Vϕ† (t) = 0 .
(18.111)

18.3.4 Correlation Functions

Equation of motion for the polarization vector P can be obtained by tak-


ing the expectation value of Eqs. (18.102) and (18.103). Using the iden-
tity σ ± = (1/2) (σx ± iσy ) and the notation P± = (1/2) (Px ± iPy ) and
û± = (1/2) (x̂ ± iŷ) one finds that
Ṗz = (Ω (t) × P)z − Γ1 (1 + Pz )
2 ' (
+ −i σ+ V1  + i V1† σ− ,

(18.112)
and
 
Γ1
Ṗ+ = (Ω (t) × P)+ − + Γϕ P+
2
i3 ' (  ) *4
+ − V1† σz + 2 σ + Vϕ  + Vϕ† σ+ ,

(18.113)
where the subscripts z and + denote the components of the vector Ω (t) × P
in the ẑ and û+ directions respectively. However, Eqs. (18.112) and (18.113)
contain product terms between bath operators and TLS operators (e.g. the
term σ+ V1  in Eq. (18.112). To lowest order such terms can be evaluated
by assuming that these operators are independent, e.g. σ + V1  ≃ σ+  V1 .
However, this approach, which yields vanishing contribution of all such terms
is too crude. Below we employ a better approximation to evaluate the expec-
tation value of such terms. In the first step Eqs. (18.102) and (18.103) are
formally integrated. This yields the following results

Eyal Buks Quantum Mechanics - Lecture Notes 555


Chapter 18. Open Quantum Systems

σz (t) = −1 + (1 + σ z (0)) e−Γ1 t


 t 
1 2 ′
+ dt′ [σ z , Hq ] + −iσ+ V1 + iV1† σ− eΓ1 (t −t) ,
0 i 
(18.114)
and
Γ1
σ + (t) = σ+ (0) e−( 2 +Γϕ )t
 t 
1 i †   ( Γ1 +Γϕ )(t′ −t)
+ dt′ [σ + , Hq ] + −V1 σz + 2 σ + Vϕ + Vϕ† σ+ e 2 .
0 i 
(18.115)
In the second step these expressions for the TLS operators are substituted
into Eqs. (18.112) and (18.113). In this final step, correlations are disregarded
(e.g. the expectation value ' of a term
( having ' the ( form σ+ V1† V1 is evaluated
using the approximation σ+ V1† V1 ≃ σ +  V1† V1 ). The expectation values
of bath operators are calculated with the help of the results of the previous
section. This approach yields the following results

1 t ′ ( Γ21 +Γϕ )(t′ −t) ' † ′ (
σ + V1  = dt e V1 (t ) V1 (t) σz (t′ )
i 0
iΓ1 n̂0
=− Pz ,
2
(18.116)
' ( iΓ n̂
1 0
V1† σ− = Pz , (18.117)
2
' (
V1† σz = −iΓ1 n̂0 P+ , (18.118)
and
) *
σ + Vϕ  + Vϕ† σ + = iΓϕ n̂0 P+ , (18.119)
thus
Ṗz = (Ω (t) × P)z − Γ1 [1 + (2n̂0 + 1) Pz ] , (18.120)
and
 
Γ1
Ṗ+ = (Ω (t) × P)+ − + Γϕ (2n̂0 + 1) P+ . (18.121)
2
A similar equation can be obtained for Ṗ− , which together with Eq. (18.121)
can be written as
 
Γ1
Ṗx = (Ω (t) × P)x − + Γϕ (2n̂0 + 1) Px , (18.122)
2
 
Γ1
Ṗy = (Ω (t) × P)y − + Γϕ (2n̂0 + 1) Py . (18.123)
2

Eyal Buks Quantum Mechanics - Lecture Notes 556


18.4. Problems

18.3.5 The Bloch Equations

Consider the case where ω 1 (t) = 0, i.e. Ω (t) = ω 0 ẑ [see Eq. (18.76)]. For
this case Eqs. (18.120) and (18.121) become
Ṗz = −Γ1 [1 + (2n̂0 + 1) Pz ] , (18.124)
  
Γ1
Ṗ± = ±iω0 − + Γϕ (2n̂0 + 1) P± . (18.125)
2
In the long time limit the solution is given by P± (t → ∞) = 0 and
Pz (t → ∞) = Pz0 , where [see Eq. (18.58)]
 
1 βω 0
Pz0 = − = − tanh . (18.126)
(2n̂0 + 1) 2

Note that Eq. (18.126) is in agreement with the Boltzmann distribution law
of statistical mechanics, according to which in thermal equilibrium the prob-
ability to occupy a state having energy ǫ is proportional to exp (−βǫ) (recall
that Pz is the probability to occupy the upper state of the TLS minus the
probability to occupy the lower one). In terms of the decay times T1 and T2 ,
which are defined by

T1 = Γ1−1 (2n̂0 + 1)−1 , (18.127)


 −1
Γ1
T2 = + Γϕ (2n̂0 + 1)−1 , (18.128)
2

the equations of motion for the general case, which are known as optical
Bloch equations, are given by
Px
Ṗx = (Ω (t) × P)x − , (18.129)
T2
Py
Ṗy = (Ω (t) × P)y − , (18.130)
T2
Pz − Pz0
Ṗz = (Ω (t) × P)z − . (18.131)
T1

18.4 Problems

1. Calculate the autocorrelation function f (t) f (t + t′ ) of the classical


fluctuating force f (t), which was introduced into the classical equation
of motion (18.8) of a mechanical resonator. The autocorrelation function
should yield a result consisting with the equipartition theorem.
2. Calculate the autocorrelation function f (t) f (t + t′ ), where the quan-
tum operator f (t) is given by Eq. (18.39).

Eyal Buks Quantum Mechanics - Lecture Notes 557


Chapter 18. Open Quantum Systems

3. Consider a one-dimensional mechanical resonator having mass m, reso-


nance frequency ω 0 and damping rate γ in thermal equilibrium at temper-
ature T . Calculate the expectation value of the autocorrelation function
g (τ) of the resonator’s coordinate x, which is defined by
1
g (τ ) = x (t) x (t + τ ) + x (t + τ) x (t) , (18.132)
2
in steady state.
4. Consider a TLS having energy gap ∆. A perturbation, which is externally
applied, induces transitions between the states having rate ΓT . Calculate
the polarization vector P in steady state.
5. Magnetic resonance - A time dependent magnetic field given by

B (t) = B0 ẑ + B1 (x̂ cos (ωt) + ŷ sin (ωt)) (18.133)

is applied to a spin 1/2 particle.


a) Use the Bloch equations to determine the polarization Pz in steady
state.
b) The polarization Px in steady state can be expressed as

Px = 2ω 1 [cos (ωt) χ′ (ω) + sin (ωt) χ′′ (ω)] , (18.134)

where χ′ (ω) and χ′′ (ω) are respectively the real and imaginary parts
of the magnetic susceptibility χ (ω) (i.e. χ (ω) = χ′ (ω) + iχ′′ (ω)).
Note that the term proportional to χ′ (ω) is ’in phase’ with respect
to the driving magnetic filed in the x direction [recall that Bx =
B1 cos (ωt)], whereas the second term, which is proportional to χ′′ (ω)
is ’out of phase’ [i.e. proportional to sin (ωt)] with respect to Bx .
Calculate χ (ω).

18.5 Solutions

1. In the absence of any externally applied driving force, i.e. when Fex = 0,
the classical equation of motion is given by (18.8)

mẍ + mγ ẋ + mω20 x = f (t) , (18.135)

where f (t) represents a random force acting on the resonator due to the
coupling with the thermal bath at temperature T . Bellow we consider
statistical properties of the fluctuating function x(t). However, since some
of the quantities we define may diverge, we consider a sampling of the
function x(t) in the finite time interval (−τ /2, τ /2), namely
+
x(t) −τ /2 < t < τ /2
xτ (t) = . (18.136)
0 else

Eyal Buks Quantum Mechanics - Lecture Notes 558


18.5. Solutions

The equipartition theorem requires that


1 ) * 1
mω 20 x2 = kB T , (18.137)
2 2
where
 +∞
) 2* 1
x ≡ lim dt x2τ (t) . (18.138)
τ →∞ τ −∞

Introducing the Fourier transform (FT)


 ∞
1
xτ (t) = √ dω xτ (ω)e−iωt , (18.139)
2π −∞
one finds that
 ∞  ∞
) 2* 1
x = dω xτ (ω) dω ′ xτ (ω′ )
2π −∞ −∞
 +∞
1 ′
× lim dte−i(ω+ω )t
τ →∞ τ
 −∞  
2πδ(ω+ω′ )
 ∞
1
= lim dω xτ (ω)xτ (−ω) .
τ →∞ τ −∞
(18.140)

Since x(t) is real xτ (−ω) = x∗τ (ω). In terms of the power spectrum Sx (ω),
which is defined as
1
Sx (ω) = lim |xτ (ω)|2 , (18.141)
τ →∞ τ
one has
 ∞
) 2*
x = dω Sx (ω) . (18.142)
−∞

Next, we take the FT of Eq. (18.135)


 
−mω 2 − imωγ + mω20 x(ω) = f(ω) , (18.143)
where
 ∞
1
f(t) = √ dωf (ω)e−iωt . (18.144)
2π −∞

Taking the absolute value squared yields


Sf (ω)
Sx (ω) = 3
2
4 . (18.145)
2
m2 (ωγ) + (ω 20 − ω2 )

Eyal Buks Quantum Mechanics - Lecture Notes 559


Chapter 18. Open Quantum Systems

Integrating Eq. (18.145) leads to


 ∞  ∞
1 Sf (ω)
dω Sx (ω) = 2 dω 2 2 . (18.146)
−∞ m −∞ (ωγ) + (ω 20 − ω2 )

Assuming that in the vicinity of ω0 , i.e. near the peak of the integrand,
the spectral density Sf (ω) is a smooth function on the scale of the width
of the peak γ, and also assuming that ω 0 ≫ γ, one approximately finds
that
 ∞  ∞
1 dω
dω Sx (ω) ≃ Sf (ω 0 ) 2 2 2
−∞ m 2
−∞ (ωγ) + (ω 0 − ω 2 )

Sf (ω 0 ) ∞ dα
=
m ω 0 −∞ (αγ/ω 0 ) + (1 − α2 )2
2 3 2

Sf (ω 0 ) ∞ dα

m ω 0 −∞ (αγ/ω 0 )2 + 1
2 3
  
πω0 /γ
π
= 2 2 Sf (ω0 ) .
m γω 0
(18.147)

This together with Eqs. (18.137) and (18.142) yields

mγkB T
Sf (ω0 ) = , (18.148)
π
thus Eq. (18.145) becomes

γkB T 1
Sx (ω) = . (18.149)
mπ (ωγ)2 + (ω 20 − ω2 )2

At the peak ω = ω 0 one finds


kB T
Sx (ω0 ) = . (18.150)
mπγω 20

The correlation function C (t′ ) of the fluctuating force f is defined as:


 +∞
′ ′ 1
C (t ) = f (t) f (t + t ) ≡ lim dt f (t) f (t + t′ ) . (18.151)
τ →∞ τ −∞

Using Eq. (18.144) one finds

Eyal Buks Quantum Mechanics - Lecture Notes 560


18.5. Solutions
  ∞
1 1 +∞
C (t′ ) = lim dt dωf(ω)e−iωt
2π τ →∞ τ −∞ −∞
 ∞
′ ′ −iω′ (t+t′ )
× dω f (ω )e
−∞

1 1 ∞ ′
= lim dωe−iωt f (ω)
2π τ →∞ τ −∞
 ∞  +∞

× ′ ′
dω f (ω ) dte−i(ω+ω )t
−∞
 −∞  
2πδ(ω+ω′ )
 ∞
1 ′
= lim dωe−iωt f(ω)f(−ω)
τ →∞ τ −∞
 ∞

= dωe−iωt Sf (ω) .
−∞
(18.152)
Using Eq. (18.148) and assuming as before that Sf (ω) is smooth function
near ω = ω 0 allow determining the coefficient C (t′ )

′ mγkB T ∞ ′
C (t ) = dωe−iωt = 2mγkB T δ (t′ ) . (18.153)
π
 −∞  
2πδ(t′ )

2. Using the definition (18.39) and Eqs. (18.57), (18.59) and (18.60) one has
mω0
f (t) f (t + t′ ) = −
), 2 -, -*
× F † (t) − F (t) F † (t + t′ ) − F (t + t′ )
= mγω0 (2n̂0 + 1) δ (t′ )
eβω0 + 1 ′
= mγω0 δ (t )
eβω0 − 1
 
βω 0
= mγω0 coth δ (t′ ) .
2
(18.154)

In the classical limit where kB T ≫ ω 0 one finds that


f (t) f (t + t′ ) = 2mγkB T δ (t′ ) , (18.155)
in agreement with Eq. (18.153).
3. With the help of Eqs. (5.11), (18.64) and (18.65) the autocorrelation
function can be expressed in terms of the noise operator F (t′ ) [see
Eq. (18.36)], which satisfy the correlation relations (18.57), (18.59) and
(18.60)

Eyal Buks Quantum Mechanics - Lecture Notes 561


Chapter 18. Open Quantum Systems

g (τ) = Re x (t) x (t + τ )
 )  *
= Re a (t) + a† (t) a (t + τ ) + a† (t + τ )
2mω
 ) *
= Re a (t) a† (t + τ ) + a† (t) a (t + τ)
2mω
 t  t
 ′ ′′ ) *
= Re dt′
dt′′ e(iω0 +γ)(t −t) e(−iω0 +γ)(t −t−τ ) F (t′ ) F † (t′′ )
2mω t0 t0
 t  t
 ′ ′′ ) *
+ Re dt′ dt′′ e(−iω0 +γ)(t −t) e(iω0 +γ)(t −t−τ ) F † (t′ ) F (t′′ ) .
2mω t0 t0
 t
γ (n̂0 + 1) ′ ′
= Re dt′ e(iω0 +γ)(t −t) e(−iω0 +γ)(t −t−τ )
mω t0
 t
γ n̂0 ′ ′
+ Re dt′ e(−iω0 +γ)(t −t) e(iω0 +γ)(t −t−τ ) ,
mω t0
(18.156)
where [see Eq. (18.58)]
1
n̂0 = , (18.157)
eβω0 − 1
thus
 (2n̂0 + 1) 1 − e−2γ(t−t0 )
g (τ ) = cos (ω 0 τ ) e−γτ
mω 2
 coth βω 1 − e−2γ(t−t0 )
= 2
cos (ω 0 τ ) e−γτ .
mω 2
(18.158)
In steady state, i.e. for γ (t − t0 ) ≫ 1, the autocorrelation function g (τ )
becomes
 βω
g (τ ) = coth cos (ω 0 τ ) e−γτ . (18.159)
2mω 2
4. The Bloch equation (18.131) for this case becomes

Pz − Pz0
Ṗz = −ΓT Pz − , (18.160)
T1
thus in steady state
Pz0
Pz = . (18.161)
1 + ΓT T1
Clearly, by symmetry, Px = Py = 0 in steady state.
5. The Hamiltonian of the closed system is given by

Eyal Buks Quantum Mechanics - Lecture Notes 562


18.5. Solutions


˙ Ω (t) · σ ,
Hq = (18.162)
2
where

Ω (t) = ω 0 ẑ+ω1 (cos (ωt) x̂ + sin (ωt) ŷ) , (18.163)

and where [see Eq. (4.22)]


|e| B0
ω0 = , (18.164)
me c
|e| B1
ω1 = . (18.165)
me c
In terms of the vectors û± = (1/2) (x̂ ± iŷ) the vector Ω (t) is given by
 
Ω (t) = ω 0 ẑ+ω1 e−iωt û+ + eiωt û− . (18.166)

In terms of T1 , T2 and Pz0 Eqs. (18.120) and (18.121) become


Pz − Pz0
Ṗz = (Ω (t) × P)z − , (18.167)
T1
and
P+
Ṗ+ = (Ω (t) × P)+ − . (18.168)
T2
With the help of the identities

ẑ × û± = ∓iû± , (18.169)


û+ × û+ = û− × û− = 0 , (18.170)
û+ × û− = −i (1/2) ẑ , (18.171)

Eqs. (18.167) and (18.168) become


 
iω 1 eiωt P+ − e−iωt P− Pz − Pz0
Ṗz = − , (18.172)
2 T1
and
P+
Ṗ+ = −iω0 P+ + iω 1 e−iωt Pz − . (18.173)
T2
By employing the transformation into the rotating frame [see for com-
parison Eq. (6.240)]
P+ (t) = e−iωt PR+ (t) , (18.174)
P− (t) = eiωt PR− (t) , (18.175)
Eqs. (18.172) and (18.173) become

Eyal Buks Quantum Mechanics - Lecture Notes 563


Chapter 18. Open Quantum Systems

iω 1 (PR+ − PR− ) Pz − Pz0


Ṗz = − , (18.176)
2 T1
and
PR+
ṖR+ = i (ω − ω0 ) PR+ + iω 1 Pz − . (18.177)
T2

a) In steady state, i.e. when Ṗz = ṖR+ = 0, one has


iω 1 (PR+ − PR− ) Pz − Pz0
= , (18.178)
2 T1
PR+
i (ω − ω 0 ) PR+ + iω1 Pz = , (18.179)
T2

thus (recall that PR+ = PR− )

1 + T22 (ω − ω 0 )2
Pz = Pz0 . (18.180)
1 + T22 (ω − ω 0 )2 + ω21 T1 T2

b) In steady state one has

iT2 ω 1 [1 + iT2 (ω − ω0 )]
PR+ = Pz0 , (18.181)
1 + T22 (ω − ω 0 )2 + ω21 T1 T2

thus
iT2 ω 1 [1 + iT2 (ω − ω 0 )]
P+ = 2 Pz0 e−iωt . (18.182)
1+ T22 (ω − ω0 ) + ω21 T1 T2

Using the relations


Px = 2ω 1 [cos (ωt) χ′ (ω) + sin (ωt) χ′′ (ω)]
 
= ω1 e−iωt χ (ω) + eiωt χ∗ (ω) ,
(18.183)
and

Px = P+ + P− , (18.184)

one finds that


iT2 [1 + iT2 (ω − ω 0 )]
χ (ω) = Pz0 . (18.185)
1 + T22 (ω − ω 0 )2 + ω 21 T1 T2

Eyal Buks Quantum Mechanics - Lecture Notes 564


References

1. Claude Cohen-Tannoudji, Bernard Diu, Franck Laloe, Quantum Mechanics, Wi-


ley, New York (1977).
2. J.J.Sakurai, Modern Quantum Mechanics, Addison-Wesley, New York (1994).
3. L. D. Landau and L. M. Lifshitz, Quantum Mechanics Non-Relativistic Theory,
Oxford: Pergamon Press (1977).
4. L. S. Schulman, Techniques and Applications of Path Integration, , Wiley, New
York (1981).
5. M. V. Berry, Quantal phase-factors accompanying adiabatic changes, Proc. Roy.
Soc. London A 392, 45-57 (1984).
6. Alexander L. Fetter and John Dirk Walecka, Quantum Theory of Many-Particle
Systems, Mcgraw-Hill (1971).
7. Howard Carmichael, An open systems approach to quantum optics, Springer
(August 1993).
Index

action, 1 Fermi-Dirac function, 469


adiabatic approximation, 397 Fermion, 446
Aharonov-Bohm effect, 374 Feynman’s path integral, 373
angular momentum, 149 fine-structure constant, 434
flux quantum, 377, 492, 528
Bogoliubov transformation, 464, 513 fugacity, 285
Bohr’s magneton, 30, 72
Bohr’s radius, 210 gauge invariance, 377, 484
Bohr-Sommerfeld quantization rule, gauge transformation, 399
358 generalized force, 4
Bose-Einstein function, 469 geometrical phase, 398
Boson, 446, 448 gyromagnetic ratio, 30
bra-vector, 17
Hamilton’s formalism, 1
canonically conjugate, 5 Hamilton-Jacobi equations, 5
central potential, 203 Heisenberg representation, 73
chemical potential, 285 Hermitian adjoint, 22
closure relation, 18 Hydrogen atom, 208
coherence length, 520
collapse postulate, 29 ideal gas, 463
commutation relation, 34 identical particles, 445
commuting operators, 35 inner product, 15
conservative system, 4
continuity equation, 353 Jaynes-Cummings model, 303
Coulomb gauge, 415, 431 Josephson effect, 493
current density, 353 Josephson inductance, 496

DC SQUID, 528 ket-vector, 17


degeneracy, 24 kinetic energy, 4
density operator, 223
Dirac’s notation, 17 Lagrangian, 1
Drude model, 485 Larmor frequency, 72
dual correspondence, 19 linear vector space, 15
London Equations, 489
Ehrenfest’s theorem, 76 London penetration depth, 490
eigenvalue, 23
eigenvector, 23 macroscopic quantum model, 489
equipartition theorem, 548 magnetic moment, 29
Euler-Lagrange equations, 2 matrix representation, 20
expectation value, 29 Maxwell’s equations, 415, 485
Meissner effect, 490
Fermi’s golden rule, 343 momentum representation, 54
Index

momentum wavefunction, 55 scattering time, 485


Schr dinger equation, 3, 69
norm, 16 Schwartz inequality, 37, 38
normal ordering, 113 second quantization, 457
number density operator, 453 semiclassical limit, 387
number operator, 102 shell, 213
sperical harmonics, 161
observable, 22 spin, 29
operator, 17 spin 1/2, 71, 157
optical Bloch equations, 557 SQUID, 496
orbital angular momentum, 149, 157 state vector, 15
orthogonal, 16 stationary state, 71
orthonormal basis, 16 Stern-Gerlach, 30
outer product, 18 superconductivity, 483
symmetric ordering, 74
path integration, 369
thermal bath, 542
Pauli’s exclusion principle, 449 thermal equilibrium, 546
Planck’s constant, 3 Thomas-Fermi approximation, 510
plasma frequency, 488, 498 Thomas-Reiche-Kuhn sum rule, 78
Poisson’s brackets, 7 time evolution operator, 69
position representation, 50 time-dependent perturbation theory,
position wavefunction, 51 337
positive-definite, 38 tomography, 234
potential energy, 4 trace, 33
principle of least action, 2 transformation function, 55
projector, 25 translation operator, 50
pure ensemble, 226 tunneling, 360
turning point, 354
quantized field operator, 452
quantum bit, 499 uncertainty principle, 36, 233
quantum measurement, 28 unitary, 32
quantum statistical mechanics, 227
vector potential, 415, 431
radial equation, 206
reduced mass, 208 Weyl transformation, 75
rotation, 150 WKB approximation, 351

Eyal Buks Quantum Mechanics - Lecture Notes 568

Você também pode gostar