Está en la página 1de 375

LECTURE NOTES

ON QUANTUM MECHANICS

Dr. Shun-Qing Shen

Department of Physics

The University of Hong Kong

September 2004
CT Contents

0.1 General Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

1 Fundamental Concepts 1

1.1 Relation between experimental interpretations and theoretical inferences 2

1.1.1 Experimental facts . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.1.2 Theoretical development . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Materials from Britannica Online . . . . . . . . . . . . . . . . . . . . . 3

1.2.1 Photoelectric effect . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.2 Frank-Hertz experiment . . . . . . . . . . . . . . . . . . . . . . 7

1.2.3 Compton effect . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

1.3 The Stern-Gerlach Experiment . . . . . . . . . . . . . . . . . . . . . . . 11

1.3.1 The Stern-Gerlach experiment . . . . . . . . . . . . . . . . . . . 12

1.3.2 Sequential Stern-Gerlach Experiment . . . . . . . . . . . . . . . 14

1.3.3 Analogy with Polarization of Light . . . . . . . . . . . . . . . . 16

ii
CONTENTS – MANUSCRIPT

1.4 Dirac Notation and Operators . . . . . . . . . . . . . . . . . . . . . . . 16

1.5 Base kets and Matrix Representation . . . . . . . . . . . . . . . . . . . 21

1.5.1 Eigenkets of an Observable . . . . . . . . . . . . . . . . . . . . . 21

1.5.2 Eigenkets as Base kets: . . . . . . . . . . . . . . . . . . . . . . . 22

1.5.3 Matrix Representation: . . . . . . . . . . . . . . . . . . . . . . . 24

1.6 Measurements, Observables & The Uncertainty Relation . . . . . . . . 26

1.6.1 Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

1.6.2 Spin1/2 system . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

1.6.3 Probability Postulate . . . . . . . . . . . . . . . . . . . . . . . . 29

1.6.4 S and S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

1.6.5 The Algebra of Spin Operators . . . . . . . . . . . . . . . . . . 33

1.6.6 Algebra of the Pauli matrices . . . . . . . . . . . . . . . . . . . 34

1.6.7 Observable . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

1.7 Change of Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

1.7.1 Transformation Operator . . . . . . . . . . . . . . . . . . . . 38

1.7.2 Transformation Matrix . . . . . . . . . . . . . . . . . . . . . . . 39

1.7.3 Diagonalization . . . . . . . . . . . . . . . . . . . . . . . . . . 41

1.8 Position, Momentum, and Translation . . . . . . . . . . . . . . . . . . . 43

1.8.1 Continuous Spectra . . . . . . . . . . . . . . . . . . . . . . . . . 43

1.8.2 Some properties of the −function. . . . . . . . . . . . . . . . . 44

iii
CONTENTS – MANUSCRIPT

1.8.3 Position Eigenkets and Position Measurements . . . . . . . . . . 45

1.8.4 Translation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

1.9 The Uncertainty Relation . . . . . . . . . . . . . . . . . . . . . . . . . 52

2 Quantum Dynamics 62

2.1 Time Evolution and the Schrödinger Equation . . . . . . . . . . . . . . 62

2.1.1 Time Evolution Operator . . . . . . . . . . . . . . . . . . . . . 62

2.1.2 The Schrödinger Equation. . . . . . . . . . . . . . . . . . . . . . 65

2.1.3 Time Dependence of Expectation Value: Spin Precession. . . . . 69

2.2 The Schrodinger versus the Heisenberg Picture . . . . . . . . . . . . . . 72

2.2.1 Unitary operators . . . . . . . . . . . . . . . . . . . . . . . . . . 72

2.2.2 Two Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . 73

2.2.3 The Heisenberg Equation of Motion. . . . . . . . . . . . . . . . 75

2.2.4 How to construct a Hamiltonian . . . . . . . . . . . . . . . . . . 76

2.3 Simple Harmonic Oscillator. . . . . . . . . . . . . . . . . . . . . . . . . 78

2.3.1 Eigenvalue and eigenstates . . . . . . . . . . . . . . . . . . . . . 78

2.3.2 Time Development of the Oscillator . . . . . . . . . . . . . . . . 85

2.3.3 The Coherent State . . . . . . . . . . . . . . . . . . . . . . . . 87

2.4 Schrodinger Wave Equation: Simple Harmonic Oscillator . . . . . . . . 89

2.5 Propagators and Feynman Path Integrals . . . . . . . . . . . . . . . . . 91

2.5.1 Propagators in Wave Mechanics. . . . . . . . . . . . . . . . . . . 91

iv
CONTENTS – MANUSCRIPT

2.5.2 Propagator as a Transition Amplitude . . . . . . . . . . . . . . 97

2.5.3 Path Integral as the Sum Over Paths . . . . . . . . . . . . . . . 98

2.5.4 Feynman’s Formalism . . . . . . . . . . . . . . . . . . . . . . . 99

2.6 The Gauge Transformation and Phase of Wave Function . . . . . . . . 102

2.6.1 Constant Potential . . . . . . . . . . . . . . . . . . . . . . . . . 102

2.6.2 Gauge Transformation in Electromagnetism . . . . . . . . . . . 105

2.6.3 The Gauge Transformation . . . . . . . . . . . . . . . . . . . . . 108

2.6.4 The Aharonov-Bohm Effect . . . . . . . . . . . . . . . . . . . . 111

2.6.5 Magnetic Monopole . . . . . . . . . . . . . . . . . . . . . . . . . 114

2.7 Interpretation of Wave Function. . . . . . . . . . . . . . . . . . . . . . 116

2.7.1 What’s Ψ ()? . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

2.7.2 The Classical Limit . . . . . . . . . . . . . . . . . . . . . . . . . 119

2.8 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121

2.8.1 One dimensional square well potential . . . . . . . . . . . . . . 121

2.8.2 A charged particle in a uniform magnetic field . . . . . . . . . . 125

3 Theory of Angular Momentum 128

3.1 Rotation and Angular Momentum . . . . . . . . . . . . . . . . . . . . . 128

3.1.1 Finite versus infinitesimal rotation . . . . . . . . . . . . . . . . 129

3.1.2 Orbital angular momentum . . . . . . . . . . . . . . . . . . . . 136

3.1.3 Rotation operator for spin 1/2 . . . . . . . . . . . . . . . . . . . 138

v
CONTENTS – MANUSCRIPT

3.1.4 Spin precession revisited . . . . . . . . . . . . . . . . . . . . . . 140

3.2 Rotation Group and the Euler Angles . . . . . . . . . . . . . . . . . . . 141

3.2.1 The Group Concept . . . . . . . . . . . . . . . . . . . . . . . . 141

3.2.2 Orthogonal Group . . . . . . . . . . . . . . . . . . . . . . . . . 142

3.2.3 “Special”? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

3.2.4 Unitary Unimodular Group . . . . . . . . . . . . . . . . . . . . 144

3.2.5 Euler Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

3.3 Eigenvalues and Eigenkets of Angular Momentum . . . . . . . . . . . . 148

3.3.1 Representation of Rotation Operator . . . . . . . . . . . . . . . 156

3.4 Schwinger Oscillator Model. . . . . . . . . . . . . . . . . . . . . . . . . 158

3.4.1 Spin 1/2 system . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

3.4.2 Two-spin—1/2 system . . . . . . . . . . . . . . . . . . . . . . . . 163

3.4.3 Explicit Formula for Rotation Matrices. . . . . . . . . . . . . . . 165

3.5 Combination of Angular Momentum and Clebsh-Gordan Coefficients . 167

3.5.1 Clebsch-Gordan coefficients . . . . . . . . . . . . . . . . . . . . 170

3.6 Spin Correlation Measurements and Bell’s Inequality . . . . . . . . . . 176

3.6.1 Spin singlet state . . . . . . . . . . . . . . . . . . . . . . . . . . 176

3.6.2 Einstein’s Locality Principle and Bell’s inequality . . . . . . . . 177

3.6.3 Quantum Mechanics and Bell’s Inequality . . . . . . . . . . . . 179

3.7 PROBLEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

vi
CONTENTS – MANUSCRIPT

4 Symmetries in Physics 183

4.1 Symmetries and Conservation Laws . . . . . . . . . . . . . . . . . . . . 184

4.1.1 Symmetry in Classical Physics . . . . . . . . . . . . . . . . . . . 184

4.1.2 Symmetry in Quantum Mechanics . . . . . . . . . . . . . . . . . 185

4.1.3 Degeneracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

4.1.4 Symmetry and symmetry breaking . . . . . . . . . . . . . . . . 187

4.1.5 Summary: symmetries in physics . . . . . . . . . . . . . . . . . 191

4.2 Discrete Symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

4.2.1 Parity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193

4.2.2 The Momentum Operator . . . . . . . . . . . . . . . . . . . . . 195

4.2.3 The Angular Momentum . . . . . . . . . . . . . . . . . . . . . . 196

4.2.4 Parity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197

4.2.5 Lattice Translation . . . . . . . . . . . . . . . . . . . . . . . . . 200

4.2.6 A more realistic example: a 1D system . . . . . . . . . . . . . . 202

4.3 Permutation Symmetry and Identical Particles . . . . . . . . . . . . . . 204

4.3.1 Identical particles . . . . . . . . . . . . . . . . . . . . . . . . . . 204

4.3.2 Two electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210

4.3.3 The helium atom . . . . . . . . . . . . . . . . . . . . . . . . . . 210

4.4 Time Reversal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210

4.4.1 Classical cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210

vii
CONTENTS – MANUSCRIPT

4.4.2 Time reversal operator Θ . . . . . . . . . . . . . . . . . . . . . . 212

4.4.3 Time reversal for a spin 1/2 system . . . . . . . . . . . . . . . . 214

5 Approximation Methods for Bound States 217

5.1 The Variation Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

5.1.1 Expectation value of the energy . . . . . . . . . . . . . . . . . . 218

5.1.2 Particle in a one-dimensional infinite square well . . . . . . . . . 220

5.1.3 Ground State of Helium Atom . . . . . . . . . . . . . . . . . . . 221

5.2 Stationary Perturbation Theory: Nondegenerate Case . . . . . . . . . . 223

5.2.1 Statement of the Problem . . . . . . . . . . . . . . . . . . . . . 223

5.2.2 The Two-State Problem . . . . . . . . . . . . . . . . . . . . . . 225

5.2.3 Formal Development of Perturbation . . . . . . . . . . . . . . . 226

5.3 Application of the Perturbation Expansion . . . . . . . . . . . . . . . . 230

5.3.1 Simple harmonic oscillator . . . . . . . . . . . . . . . . . . . . . 231

5.3.2 Atomic hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . . 233

5.3.3 Quantum well . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237

5.4 Stationary Perturbation Theory: Degenerate Case . . . . . . . . . . . . 237

5.4.1 Revisited two-state problem . . . . . . . . . . . . . . . . . . . . 237

5.4.2 The basic procedure of degenerate perturbation theory . . . . . 239

5.4.3 Example: Zeeman Effect . . . . . . . . . . . . . . . . . . . . . . 242

5.4.4 Spin-Orbit interaction . . . . . . . . . . . . . . . . . . . . . . . 243

viii
CONTENTS – MANUSCRIPT

5.4.5 Example: First Order Stark Effect in Hydrogen . . . . . . . . . 244

5.5 The Wentzel-Kramers-Brillouin (WKB) approximation . . . . . . . . . 247

5.6 Time-dependent Problem: Interacting Picture and Two-State Problem 247

5.6.1 Time-dependent Potential and Interacting Picture . . . . . . . . 247

5.6.2 Time-dependent Two-State Problem . . . . . . . . . . . . . . . 250

5.7 Time-dependent Perturbation Problem . . . . . . . . . . . . . . . . . . 254

5.7.1 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . 254

5.7.2 Time-independent perturbation . . . . . . . . . . . . . . . . . . 256

5.7.3 Harmonic perturbation . . . . . . . . . . . . . . . . . . . . . . . 257

5.7.4 The Golden Rule . . . . . . . . . . . . . . . . . . . . . . . . . . 258

6 Collision Theory 260

6.1 Collisions in one- and three-dimensions . . . . . . . . . . . . . . . . . . 261

6.1.1 One-dimensional square potential barriers . . . . . . . . . . . . 261

6.1.2 Datta-Das spin field transistor . . . . . . . . . . . . . . . . . . . 265

6.2 Collision in three dimensions . . . . . . . . . . . . . . . . . . . . . . . . 265

6.3 Scattering by Spherically Symmetric Potentials . . . . . . . . . . . . . 270

6.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276

6.4.1 Scattering by a square well . . . . . . . . . . . . . . . . . . . . . 276

6.4.2 Scattering by a hard-sphere potential . . . . . . . . . . . . . . . 278

6.4.3 Identical Particles and Scattering . . . . . . . . . . . . . . . . . 280

ix
CONTENTS – MANUSCRIPT

6.5 Applications to interactions with the classical field . . . . . . . . . . . . 281

6.5.1 Absorption and stimulation emission . . . . . . . . . . . . . . . 281

6.5.2 Electric dipole approximation . . . . . . . . . . . . . . . . . . . 283

6.5.3 Photoelectric effect . . . . . . . . . . . . . . . . . . . . . . . . . 284

6.6 Approximate Collision Theory . . . . . . . . . . . . . . . . . . . . . . . 286

6.6.1 The Lippman-Schwinger Equation . . . . . . . . . . . . . . . . . 286

6.6.2 The Born Approximation . . . . . . . . . . . . . . . . . . . . . . 291

6.6.3 The higher-order Born approximation . . . . . . . . . . . . . . . 292

6.6.4 Optical Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 292

6.6.5 Application: from Yukawa potential to Coloumb potential . . . 294

7 Selected Topics 296

7.1 Relativistic Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . . 296

7.1.1 The Schrodinger Equation . . . . . . . . . . . . . . . . . . . . . 296

7.1.2 Dirac Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 298

7.1.3 Spin of Dirac particles . . . . . . . . . . . . . . . . . . . . . . . 302

7.1.4 The Zeeman coupling . . . . . . . . . . . . . . . . . . . . . . . . 303

7.1.5 Exact solution in Coloumb potential: Hydrogen atom . . . . . . 304

7.1.6 Spin-orbit coupling . . . . . . . . . . . . . . . . . . . . . . . . . 304

7.1.7 Spin-orbital coupling and spintronics . . . . . . . . . . . . . . . 306

7.1.8 Spin transverse force . . . . . . . . . . . . . . . . . . . . . . . . 306

x
CONTENTS – MANUSCRIPT

7.1.9 Spin-orbit coupling and Datta-Das field-effect transistor . . . . . 306

7.1.10 Graphene: 2D massless Dirac quasi-particles . . . . . . . . . . . 306

7.1.11 Topological Insulator and Dirac particels . . . . . . . . . . . . . 318

7.2 Quantum Statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319

7.2.1 Density Operator and Ensembles . . . . . . . . . . . . . . . . . 320

7.2.2 Quantum Statistical Mechanism . . . . . . . . . . . . . . . . . . 322

7.2.3 Quantum Statistics . . . . . . . . . . . . . . . . . . . . . . . . . 326

7.2.4 Systems of non-interaction particles . . . . . . . . . . . . . . . . 327

7.2.5 Bose-Einstein Condensation . . . . . . . . . . . . . . . . . . . . 331

7.2.6 Free fermion gas . . . . . . . . . . . . . . . . . . . . . . . . . . 333

7.3 Quantum Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335

7.3.1 Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336

7.3.2 Quantum Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . 338

7.3.3 Laughlin’s Theory . . . . . . . . . . . . . . . . . . . . . . . . . 340

7.3.4 Charged particle in the presence of a magnetic field . . . . . . . 342

7.3.5 Landau Level and Quantum Hall Effect . . . . . . . . . . . . . . 346

7.4 Quantum Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347

7.4.1 Spin Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . 349

7.4.2 Two-Site Problem . . . . . . . . . . . . . . . . . . . . . . . . . . 351

7.4.3 Ferromagnetic Exchange (  0) . . . . . . . . . . . . . . . . . 354

xi
CONTENTS – MANUSCRIPT

7.4.4 Antiferromagnetic Exchange . . . . . . . . . . . . . . . . . . . . 356

A 0.1 General Information

New Course Code: PHYS6504

• Aim/Following-up: The course provides an introduction to advanced tech-

niques in quantum mechanics and their application to several selected topics in

condensed matter physics.

• Contents: Dirac notation and formalism, time evolution of quantum systems,

angular momentum theory, creation and annihilation operators (the second quan-

tization representation), symmetries and conservation laws, permutation symme-

try and identical particles, quantum statistics, non-degenerate and degenerate

perturbation theory, time-dependent perturbation theory, the variational method

• Prerequisites: PHYS2323 and PHYS3332 or equivalent

• Co-requisite: Nil

• Teaching: 36 hours of lectures and tutorial classes

• Duration: One semester (1st semester)

• Assessment: One three-hour examination (70%) and course assessment (30%)

xii
CONTENTS – MANUSCRIPT

• Textbook: J. J. Sakurai, Modern Quantum Mechanics (Addison-Wesley, 1994)

• Web page: http://bohr.physics.hku.hk/~phys6504/ All lectures notes in pdf

files can be download from the site.

• References: L. Schiff, Quantum Mechanics (McGraw-Hill, 1968, 3nd ed.); Richard

Feynman, Robert B. Leighton, and Matthew L. Sands, Feynman Lectures on

Physics Vol. III, (Addison-Wesley Publishing Co., 1965); L. D. Landau and E.

M. Lifshitz, Quantum Mechanics

• Time and Venue:

14:00 -15:55: Tuesday/CYP-105

14:00 -14:55: Thursday/MW-714

xiii
CONTENTS – MANUSCRIPT

Numerical values of some physical quantities

xiv
CONTENTS – MANUSCRIPT

~ = 1054 × 10−27 erg-sec (Planck’s constant divided by 2

 = 480 × 10−10 esu (magnitude of electron charge)

 = 0911 × 10−27 g (electron mass)

 = 1672 × 10−24 g (proton mass)

0 = ~2 2 = 529 × 10−9 cm (Bohr radius)

2 0 = 272eV (twice binding energy of hydrogen)

 = 300 × 1010 cm/sec (speed of light)

~2 = 137 (reciprocal fine structure constant)

~2 = 0927 × 10−20 erg/oersted (Bohr magneton)

2 = 0511MeV (electron rest energy)

2 = 938MeV (proton rest energy)

1eV = 1602 × 10−12 erg

1 eV/c = 12 400Å

1eV = 11 600K

xv
CONTENTS – MANUSCRIPT

xvi
CN Chapter 1

CT Fundamental Concepts

At the present stage of human knowledge, quantum mechanics can be regarded as the

fundamental theory of atomic phenomena. The experimental data on which it is based

are derived from physical events that almost entirely beyond the range of direct human

perception. It is not surprising that the theory embodies physical concepts that are

foreign to common daily experience.

The most traditional way to introduce the quantum mechanics is to follow the

historical development of theory and experiment — Planck’s radiation law, the Einstein-

Debye’s theory of specific heat, the Bohr atom, de Broglie’s matter wave and so forth

— together with careful analysis of some experiments such as diffraction experiment

of light, the Compton effect, and Franck-Hertz effect. In this way we can enjoy the

experience of physicists of last century to establish the theory. In this course we do not

1
CHAPTER 1 – MANUSCRIPT

follow the historical approach. Instead, we start with an example that illustrates the

inadequacy of classical concepts in a fundamental way.

A 1.1 Relation between experimental interpretations

and theoretical inferences

Schiff’s book has a good introduction to this theory. Here we just list several experi-

ments which played key roles in development of quantum theory.

B 1.1.1 Experimental facts

• Electromagnetic wave/light: Diffraction (Young, 1803; Laue, 1912))

• Electromagnetic quanta /light: Black body radiation (Planck, 1900); Photoelec-

tric effect (Einstein, 1904); Compton effect (1923); Combination Principle (Rita-

Rydberg, 1908)

• Discrete values for physical quantities: Specific heat (Einstein 1907, Debye 1912);

Franck-Hertz experiment (1913); Stern-Gerlach experiment (1922)

B 1.1.2 Theoretical development

• Maxwell’s theory for electromagnetism: Electromagnetic wave (1864)

2
CHAPTER 1 – MANUSCRIPT

• Planck’s theory for black body radiation:  = ~: Electromagnetic quanta (1900)

• de Broglie’s theory:  = : Wave-Particle Duality, (1924)

• Bohr’s atom model, (1913)

• Birth of quantum mechanics: Heisenberg’s theory (1926); Schrodinger’s theory

(1926)

A 1.2 Materials from Britannica Online

B 1.2.1 Photoelectric effect

The phenomenon in which charge particles are released from a material when it absorbs

radiant energy. The photoelectric effect commonly is thought of as the ejection of

electrons from the surface of a metal plate when light falls on it. In the broad sense,

however, the phenomenon can take place when the radiant energy is in the region of

visible or ultraviolet light, X rays, or gamma rays; when the material is a solid, liquid, or

gas; and when the particles released are electrons or ions (charged atoms or molecules).

The photoelectric effect was discovered in 1887 by a German physicist, Hein-

rich Rudolf Hertz, who observed that ultraviolet light changes the lowest voltage at

which sparking takes place between given metallic electrodes. At the close of the 19th

century, it was established that a cathode ray (produced by an electric discharge in a

3
CHAPTER 1 – MANUSCRIPT

Figure 1.1:

rarefied-gas atmosphere) consists of discrete particles, called electrons, each bearing an

elementary negative charge. In 1900 Philipp Lenard, a German physicist, studying the

electrical charges liberated from a metal surface when it was illuminated, concluded

that these charges were identical to the electrons observed in cathode rays. It was fur-

ther discovered that the current (given the name photoelectric because it was caused

by light rays), made up of electrons released from the metal, is proportional to the

intensity of the light causing it for any fixed wavelength of light that is used. In 1902 it

was proved that the maximum kinetic energy of an electron in the photoelectric effect

is independent of the intensity of the light ray and depends on its frequency.

The observations that (1) the number of electrons released in the photoelec-

tric effect is proportional to the intensity of the light and that (2) the frequency, or

4
CHAPTER 1 – MANUSCRIPT

wavelength, of light determines the maximum kinetic energy of the electrons indicated

a kind of interaction between light and matter that could not be explained in terms of

classical physics. The search for an explanation led in 1905 to Albert Einstein’s funda-

mental theory that light, long thought to be wavelike, can be regarded alternatively as

composed of discrete particles (now called photons ), equivalent to energy quanta.

In explaining the photoelectric effect, Einstein assumed that a photon could

penetrate matter, where it would collide with an atom. Since all atoms have electrons,

an electron would be ejected from the atom by the energy of the photon, with great

velocity. The kinetic energy of the electron, as it moved through the atoms of the

matter, would be diminished at each encounter. Should it reach the surface of the

material, the kinetic energy of the electron would be further reduced as the electron

overcame and escaped the attraction of the surface atoms. This loss in kinetic energy is

called the work function, symbolized by omega (). According to Einstein, each light

quantum consists of an amount of energy equal to the product of Planck’s universal

constant () and the frequency of the light (indicated by the Greek ). Einstein’s theory

of the photoelectric effect postulates that the maximum kinetic energy of the electrons

ejected from a material is equal to the frequency of the incident light times Planck’s

constant, less the work function. The resulting photoelectric equation of Einstein can

be expressed by  =  − , in which  is the maximum kinetic energy of the ejected

electron,  is a constant, later shown to be numerically the same as Planck’s constant,

5
CHAPTER 1 – MANUSCRIPT

 is the frequency of the incident light, and  is the work function.

The kinetic energy of an emitted electron can be measured by placing it in an

electric field and measuring the potential or voltage difference (indicated as V) required

to reduce its velocity to zero. This energy is equal to the product of the potential

difference and an electron’s charge, which is always a constant and is indicated by ;

thus,  =  .

The validity of the Einstein relationship was examined by many investigators

and found to be correct but not complete. In particular, it failed to account for the fact

that the emitted electron’s energy is influenced by the temperature of the solid. The

remedy to this defect was first formulated in 1931 by a British mathematician, Ralph

Howard Fowler, who, on the assumption that all electrons with energies greater than

the work function would escape, established a relationship between the photoelectric

current and the temperature: the current is proportional to the product of the square

of the temperature and a function of the incident photon’s energy. The equation is

 =  2 (), in which  is the photoelectric current,  and  are constants, and ()

is an exponential series, whose numerical values have been tabulated; the dimensionless

value  equals the kinetic energy of the emitted electrons divided by the product of the

temperature and the Boltzmann constant of the kinetic theory:  = ( − ) ; in

which  is the argument of the exponential series.

6
CHAPTER 1 – MANUSCRIPT

B 1.2.2 Frank-Hertz experiment

in physics, first experimental verification of the existence of discrete energy states in

atoms, performed (1914) by the German-born physicists James Franck and Gustav

Hertz .

Franck and Hertz directed low-energy electrons through a gas enclosed in an

electron tube. As the energy of the electrons was slowly increased, a certain critical

electron energy was reached at which the electron stream made a change from almost

undisturbed passage through the gas to nearly complete stoppage. The gas atoms were

able to absorb the energy of the electrons only when it reached a certain critical value,

indicating that within the gas atoms themselves the atomic electrons make an abrupt

transition to a discrete higher energy level. As long as the bombarding electrons have

less than this discrete amount of energy, no transition is possible and no energy is

absorbed from the stream of electrons. When they have this precise energy, they lose it

all at once in collisions to atomic electrons, which store the energy by being promoted

to a higher energy level.

B 1.2.3 Compton effect

increase in wavelength of X rays and other energetic electromagnetic radiations that

have been elastically scattered by electrons; it is a principal way in which radiant energy

is absorbed in matter. The effect has proved to be one of the cornerstones of quantum

7
CHAPTER 1 – MANUSCRIPT

Figure 1.2:

8
CHAPTER 1 – MANUSCRIPT

mechanics, which accounts for both wave and particle properties of radiation as well as

of matter.

The American physicist Arthur Holly Compton explained (1922; published

1923) the wavelength increase by considering X rays as composed of discrete pulses,

or quanta, of electromagnetic energy, which he called photons. Photons have energy

and momentum just as material particles do; they also have wave characteristics, such

as wavelength and frequency. The energy of photons is directly proportional to their

frequency and inversely proportional to their wavelength, so lower-energy photons have

lower frequencies and longer wavelengths. In the Compton effect, individual photons

collide with single electrons that are free or quite loosely bound in the atoms of matter.

Colliding photons transfer some of their energy and momentum to the electrons, which

in turn recoil. In the instant of the collision, new photons of less energy and momentum

are produced that scatter at angles the size of which depends on the amount of energy

lost to the recoiling electrons.

Because of the relation between energy and wavelength, the scattered photons

have a longer wavelength that also depends on the size of the angle through which the X

rays were diverted. The increase in wavelength or Compton shift does not depend on the

wavelength of the incident photon. The Compton effect was discovered independently

by the physical chemist Peter Debye in early 1923.


∆ = 
(1 − cos )

9
CHAPTER 1 – MANUSCRIPT

Figure 1.3:

10
CHAPTER 1 – MANUSCRIPT

Figure 1.4: The Stern-Gerlach Experiment


0 = 
= 242631058 × 10−12 m = 0024̊

A 1.3 The Stern-Gerlach Experiment

We start with the Stern-Gerlach experiment to introduce some basic concepts of quan-

tum mechanics. The two-state problem can be regarded as the most quantum. A lot of

important discoveries are related to it. It is worthy studying very carefully. We shall

repeat to discuss the problem throughout this course.

11
CHAPTER 1 – MANUSCRIPT

B 1.3.1 The Stern-Gerlach experiment

Oven: silver atoms (Ag) are heated in the oven. The oven has a small hole through

which some of the silver atoms escape to form an atomic beam.

Ag: (Electron configuration: 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6 4d10 5s1 ). The outer

shell has only ONE electron (5s1 ) . The atom has an angular momentum., which is due

solely to the spin of 47 electron.

Collimating slit: change the diverging Ag beam to a parallel beam.

Shaped Magnet: N and S are north and south poles of a magnet. The knife-

edge of S results in a much stronger magnetic field at the point P than Q . i.e. the

magnet generates an inhomogeneous magnetic field.

Role of the inhomogeneous field: to change the direction of the Ag beam. The

interaction energy is  = − · B The force experienced by the atoms is

 
 = ( · B) ≈ 
 

If the magnetic field is uniform , i.e. 
= 0, the Ag beam will not change its direction.

In the field different magnetic moments experience different forces, and the atoms with

different magnetic moments will change different angles after the beams pass through

the shaped magnet. Suppose the length of the shaped magnet . It takes a time 

12
CHAPTER 1 – MANUSCRIPT

Figure 1.5: Beam from the Stern-Gerlach apparatus: (a) is expected from classical

physics, while (b) is actually observed experimentally.

for particles to go though the magnets. The angle is about

  
∆   
 


 
= = 2
=  2 ∝  
   

They will reach at different places on the screen.

Classically: all values of  =  cos  (0    ) would be expected to realize

between || and − ||  It has a continuous distribution.

 are observed (electron


Experimentally: only two values of z component of 


spin  ∝ )

Consequence: the spin of electron has two discrete values along the magnetic

13
CHAPTER 1 – MANUSCRIPT

field



⎨ ~2
 =


⎩ −~2

~ = 10546 × 10−27 erg.s = 65822 × 10−16 eV.s

− − − − Planck’s constant divided by 2

It should be noted that the constant cannot be determined accurately from this exper-

iment.

B 1.3.2 Sequential Stern-Gerlach Experiment

SGẑ stands for an apparatus with inhomogeneous magnetic field in z direction, and

SGx in x direction

Case(a): no surprising!

Case(b):  + beam is made up of 50%  + and 50%  −?  − beam?

( +) → 12 ( +) + 12 ( −)?

( −) → 12 ( +) + 12 ( −)?

Case(c): Since  − is blocked at the first step, why is  + beam made up of

both  + and  − beams?

( +) → 12 ( +) + 12 ( −)?

( −) → 12 ( +) + 12 ( −)?

14
CHAPTER 1 – MANUSCRIPT

Figure 1.6: Sequential Stern-Gerlach experiment

15
CHAPTER 1 – MANUSCRIPT

Consequences of S-G experiment:

• Spin space cannot be described by a 3-dimensional vector.

• The magnetic moment of atom or spin is discrete or quantized.

• We cannot determine both  and  simultaneously. More precisely, we can

say that the selection of  + beam by SGx̂ completely destroys any previous

information about  .

B 1.3.3 Analogy with Polarization of Light

It is very helpful if you compare the situation with the analogy of the polarized light

through a Polaroid filter. The polarized light wave is described by a complex function,

E = 0 x̂ cos( − )

0 h (−) (−)
i
E = 12 x̂e ± ye
2

Please refer to Sakurai’s book, P.6—10.

A 1.4 Dirac Notation and Operators

The Stern-Gerlach experiment shows that the spin space is not simply a 3-D vector

space and lead to consider a complex vector space. There exists a good analogy with

polarization of light. (Refer to Sakurai’s book). In this section we formulate the basic

16
CHAPTER 1 – MANUSCRIPT

Figure 1.7:

17
CHAPTER 1 – MANUSCRIPT

mathematics of vector spaces as used in quantum mechanics. The theory of linear

algebra has been known to mathematician before the birth of quantum mechanics, but

the Dirac notation has many advantages. At the early time of quantum mechanics

this notation was used to unify Heisenberg’s matrix mechanics and Schrodinger’s wave

mechanics. The notation in this course was first introduced by P. A. M. Dirac.

Ket space: In quantum mechanics, a physical state, for example, a silver

atom with definite spin orientation, is represented by a state vector in a complex vector

space, denoted by |i, a ket. The state ket is postulated to contain complete informa-

tion about physical state. The dimensionality of a complex vector space is specified

according to the nature of physical system under consideration.

An observable can be represented by an operator A

A(|i) = A |i

Eigenket and eigenvalue:

A(|i) =  |i

Eigenstate: the physical state corresponding to an eigenket, |i.

Two kets can be added to form a new ket: |i + |i = |i

One of the postulates is that |i and  |i with the number  6= 0 represent

the same physical state.

18
CHAPTER 1 – MANUSCRIPT

Bra space: a dual correspondence to a ket space. We postulate that corre-

sponding to every ket there exist a bra. The names come from the word “bracket” →

“bra-c-ket”.

1. There exists a one-to one correspondence between a ket space and a bra space.

|i ⇐⇒ h|

2. The bra dual to  |i (c is a complex number)

 |i ⇐⇒ ∗ h|

Inner product: In general, this product is a complex number.

h|i = (h|) · (|i)

Two properties of the inner product:

1. h|i and h|i are complex conjugates of each other.

h|i = (h|i)∗

2. the postulate of positive definite metric.

h|i ≥ 0

19
CHAPTER 1 – MANUSCRIPT

where the equality sign holds only if |i is a null ket. From a physicist’s point

of view this postulate is essential for the probabilistic interpretation of quantum

mechanics. Question: what happens if we postulate h|i ≤ 0?

Orthogonality:

h|i = 0

Normalization: all vectors can be normalized!


µ ¶ 12
1
|̃i = |i , h̃|̃i = 1
h|i

The norm of |i : (h|i)12

Operator: an operator acts on a ket from the left side, x |i  and the result-

ing product is another ket. An operator acts on a bra from the right side, h| xand

the resulting product is another bra.

• — Equality: x = y if x |i = y |i for an arbitrary ket

— Null operator: x |i = 0 for arbitrary |i

— Hermitian operator: x = x†

— x |i ←→ h| x†

Multiplication

• — Noncommutative: xy 6= yx

20
CHAPTER 1 – MANUSCRIPT

— Associative: x(yz) = (xy)z = xyz

— Hermitian adjoint: (xy)† = y† x† (Prove it!)

y |i ⇒ h| y†

¡ ¢
xy |i = x (y |i) ⇒ h| y† x† = h| y† x†

Outer product: |i h| is an operator!

The associative axiom of multiplication: the associative property holds as long

as we are dealing with “legal” multiplications among kets, bras, and operators.

1. (|i h|) · |i = |i (h| |i) ;

2. (h|) · (x |i) = (h| x) · (|i) ≡ h |x| i ;

­ ¯ ¯ ®∗
3. h |x| i = h| · (x |i) = {(h| x† ) · |i}∗ ≡  ¯x† ¯  

A 1.5 Base kets and Matrix Representation

B 1.5.1 Eigenkets of an Observable

The operators for observables must be Hermitian as physical quantities must be real.

1). The eigenvalues of a Hermitian operator A are real.

21
CHAPTER 1 – MANUSCRIPT

G Proof. First, we recall the definitions

A |i =  |i ⇐⇒ h| A† = h| ∗

h| A |i =  ⇔ h| A† |i = ∗

If A = A† =⇒  = ∗ 

2). The eigenkets of Hermitian A corresponding to different eigenvalues are

orthogonal.

G Proof.

A |i =  |i =⇒ h0 |A| i =  h0 |i

h0 | A† = h0 | 0 =⇒ h0 |A| i = 0 h0 |i

We have

( − 0 ) h0 |i = 0 =⇒ h0 |i = 0

It is conventional to normalize
⎧ all eigenkets |0 i of A so that {|0 i} form an orthogonal


⎨ 1 if 0 = 00
0
set: h”| i =  0 00 = 


⎩ 0 otherwise

B 1.5.2 Eigenkets as Base kets:

1. All normalized eigenkets of an observable A form a complete and orthonormal


¯ ®
set. {¯() }. The number of the eigenkets is equal to the dimensionality of the

22
CHAPTER 1 – MANUSCRIPT

complex vector space.

2. An arbitrary ket in the ket space can be expanded in terms of the eigenkets of A:

P ¯ ® ­ ®
|i =  ¯() =⇒ () | = 

µ ¶
P¯ ®­ ® P ¯¯ ® ­ () ¯
|i = ¯() () | = ()
  ¯ |i
 

3. The completeness relation (or closure):

P ¯¯ ® ­ () ¯
()  ¯=1

(1 means the identity operator. For an arbitary ket |i  1 |i = |i)

P¯ ®­ ¯ P¯ ®­ ¯
( ¯() () ¯)( ¯() () ¯)
 

P ¯¯ () ® ­ () ¯¯ () ® ­ () ¯¯ P ¯¯ () ® ­ ¯


=     =    () ¯
 

P ¯¯ ® ­ () ¯ P¯ ®­ ¯ P¯ ®­ ¯
= ()  ¯ =⇒ ( ¯() () ¯)( ¯() () ¯ − 1) = 0
  

This relation is very essential to develop the general formalism of quantum me-

chanics.

¯ ®­ ¯
4. Projection operator:  = ¯() () ¯

¯ ®­ ¯ ¯ ®
 |i = ¯() () ¯ i =  ¯()

23
CHAPTER 1 – MANUSCRIPT
¯ ®
 selects that portion of the ket |i parallel to ¯() :  ·  =  such that

 ’s eigenvalues are either 1 or 0.

¯ ®­ ¯ ¯ ®­ ¯
 ·  = (¯() () ¯)(¯() () ¯)

¯ ®­ ®­ ¯
= ¯() () |() () ¯

¯ ®­ ¯
= ¯() () ¯ = 

B 1.5.3 Matrix Representation:

Having specified the base kets, we show how to represent an operator, say X, by a

square matrix.

P¯ ®­ ¯ P¯ ®­ ¯
X = ( ¯() () ¯)X( ¯() () ¯)
 

P ¯¯ () ® ­ () ¯¯ ¯¯ () ® ­ () ¯¯


=   X  


P ­ () ¯¯ ¯¯ () ® ¯¯ () ® ­ () ¯¯ P ¯ ®­ ¯


= (  X  )   =  ¯() () ¯
 

Let us assume that


­ () ¯
 ¯ = (0 · · ·  0 1 · · · )

24
CHAPTER 1 – MANUSCRIPT

and ⎛ ⎞
⎜ 0 ⎟
⎜ ⎟
⎜ .. ⎟
⎜ . ⎟
⎜ ⎟
⎜ ⎟
¯ () ® ⎜ ⎟
¯ =⎜ 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 ⎟
⎜ ⎟
⎜ ⎟
⎝ .. ⎠
.
­ ¯ ¯ ®
Thus the outer product of () ¯ and ¯() gives a new operator in the form of square

matrix, ⎛ ⎞
⎜ 0 ··· 0 ··· ⎟
⎜ ⎟
⎜ .. . . .. .. ⎟
⎜ . . . . ⎟
¯ () ® ­ () ¯ ⎜ ⎟
¯  ¯ =⎜ ⎟
⎜ ⎟
⎜ 0 ··· 1 ··· ⎟
⎜ ⎟
⎜ ⎟
⎝ .. .. .. . . ⎠
. . . .
­ ¯ ®
The inner product () ¯ () = 1 Thus X can be expressed in a matrix form,
⎛ ⎞
­ (1) ¯ ¯ (1) ® ­ (1) ¯ ¯ (2) ®
⎜  ¯ X ¯  ¯ X ¯ ··· ⎟
⎜ ⎟
⎜ ­ ¯ ¯ ® ­ ¯ ¯ ® ⎟

X = ⎜  ¯ X ¯ ¯ ¯ ··· ⎟
(2) (1) (2)
 X  (2)
⎟
⎜ ⎟
⎝ .. .. ... ⎠
. .
¯ ®
For an operator A which eigenstates are ¯()
⎛ ⎞
(1)
⎜  0 ··· ⎟
⎜ ⎟
⎜ ⎟
A=⎜
⎜ 0 (2) ··· ⎟

⎜ ⎟
⎝ . .. ... ⎠
.. .

Basic Matrix Operations:

25
CHAPTER 1 – MANUSCRIPT

1. Transpose: C = A =⇒  = 

2. Addition: C = A + B =⇒  =  + 

3. Scalar-matrix multiplication: C = A =⇒  = 

P

4. Matrix-matrix multiplication: C = AB =⇒  =  · 
=1

A 1.6 Measurements, Observables & The Uncertainty

Relation

B 1.6.1 Measurements

“Measurement” has special meaning in quantum mechanics. It is a physical operation

procedure, and is quite different from that in classical physics. In the classic case the

measurement reflects the state of an object, such as position, velocity, and energy.

The process of measurement does not change the state of the object. For instance we

measure your body height, which does not change your body height. In the quantum

case the process of measurement changes the state of the object. If we want to measure

the spin state of electron we have to let electron go through the Stern-Gerlach apparatus

(of course we have other methods.). The magnetic field changes the spin state, and

select one of the eigenstates along the direction of the magnetic field. This is not a

particularly easy subject for a beginner. Let us first read the words of the great master,

26
CHAPTER 1 – MANUSCRIPT

P. A. M. Dirac, for guidance.

“A measurement always causes the system to jump into an eigenstate of the

dynamic variable that is being measured .”

–P.A.M. Dirac

Interpretation word by word:

1. Before the measurement of observable A is made, the system is assumed to be

represented by some linear combination of the eigenstates

P¯ ®­ ¯ P ¯ ®
|i = ( ¯() () ¯ i =  ¯() 
 

2. When the measurement is performed, the system is “thrown into” one of the
¯ ®
eigenstates, say ¯(1)  of the observable

¯ (1) ®
|i A measurement ¯
−−−−−−−−−−−→  

3. A measurement usually changes the state. When the measurement causes |i to
¯ ®
change into ¯(1) it is said that A is measured to be (1) !

Before the measurement the system is in |i ; after a measurement the system
¯ ®
is in ¯(1) 

B 1.6.2 Spin1/2 system

We are ready to derive the spin 1/2 operator we encountered in the Stern-Gerlach

experiment. The eigenvalues of S are ±~2. We denote the corresponding eigenkets

27
CHAPTER 1 – MANUSCRIPT

as |+i and |−i, respectively,

h+|+i ≡ h−|−i ≡ 1; h+|−i = 0

|+i and |−i form a complete (?) and orthonormal (?) set of basis. From the complete-

ness relation: 1 = (|+i h+|) + (|−i h−|), we have


⎫ ⎫
⎪ ⎪
S |+i = + ~2 |+i ⎪
⎬ S |+i h+| = + ~2 |+i h+| ⎪



⎪ ⎪

S |−i = − 2 |−i ⎭
~
S |−i h−| = − ~2 |−i h−| ⎭

~
S (|+i h+| + |−i h−|) = (|+i h+| − |−i h−|)
2
~
S = (|+i h+| − |−i h−|)
2

In general any operator for an observable can be expressed in terms of its eigenvalues

and eigenstates,

A | i =  | i

X
A =  | i h | 

Let us construct two operators:

S+ = ~ |+i h−|

S− = ~ |−i h+|

28
CHAPTER 1 – MANUSCRIPT

From the definition, we get





⎨ S+ |+i = ~ |+i h−|+i = 0


⎩ S+ |−i = ~ |+i h−|−i = ~ |+i




⎨ S− |+i = ~ |−i h+|+i = ~ |−i


⎩ S− |−i = ~ |−i h+|−i = 0

Let ⎛ ⎞ ⎛ ⎞
⎜ 1 ⎟ ⎜ 0 ⎟
|+i ≡ ⎜
⎝ ⎠
⎟ ; |−i ≡ ⎜ ⎟
⎝ ⎠
0 1

On the base, we have


⎛ ⎞ ⎛ ⎞ ⎛ ⎞
~⎜ 1 0 ⎟ ⎜ 0 1 ⎟ ⎜ 0 0 ⎟
S = ⎜ ⎟ ; S+ = ~ ⎜ ⎟ ; S− = ~ ⎜ ⎟
2⎝ ⎠ ⎝ ⎠ ⎝ ⎠
0 −1 0 0 1 0

B 1.6.3 Probability Postulate


¯ ®
The probability for jumping into some particular ¯() is given by

¯ ® ¯­ ®¯2
the probability for ¯() :  = ¯ () | ¯

P ¯¯ ® ­ () ®
provided that |i is normalized (h|i = 1). (|i = ()  | ).

X ¯ ®­ ® X
h|i = h ¯() () | = 1 = 
 

29
CHAPTER 1 – MANUSCRIPT

The expectation value of A with respect to state |i is defined as

P ­ () ® ­ () ®­ ®
hAi ≡ h |A| i = |  |A| () () |


P ¯­ ®¯2 P
= () × ¯ |() ¯ = () 
 

­ () ®
 |A| () =   

¯ ®
where  is the probability in the state ¯() .

Selective measurement (or filtration)

We imagine a measurement process with a device that selects only one of

the eigenket and rejects all others. Mathematically, we can say that such a selective

measurement amounts to applying the projection operator.

B 1.6.4 S and S

We are now in the position to determine the eigenkets of S and S and the operators

from the results of sequential Stern-Gerlach experiments.(see Fig.1.3)

Case (b): The eigenstates of S and S can be expressed in terms of the

eigenstates of S  Similar to S , assume the two eigenstates of S are | +i and | −i

with eigenvalues + ~2 and − ~2  respectively,

~
S = (| +i h +| − | −i h −|)
2

The beam of S + goes through SGz and splits into two beams with the same intensity,

30
CHAPTER 1 – MANUSCRIPT

| +i =  | +i +  | −i 

From experiment =⇒ |h+| +i|2 = |h+| −i|2 = 12

From experiment =⇒ |h−| +i|2 = |h−| −i|2 = 12





⎨ | +i = 121 1
|+i + 212 1 |−i
2
=⇒

⎪ 0
⎩ | −i = 121 1
|+i + 212 1 |−i
2

h + | −i = 0

1 1 0
h+|+i + h−|−i −1 +1 = 0 ⇒  01 =  +  1
2 2

~
S = (| +i h +| − | −i h −|)
2

~ −1
= ( |+i h−| + 1 |−i h+|)
2

Likewise,
~ −2
S = ( |+i h−| + 2 |−i h+|)
2

How to determine  1 and  2 ?

From experiment |h ± | +i|2 = |h ± | −i|2 = 12

=⇒  1 −  2 = ±2

We just can determine the difference of  1 and  2. For convenience we take S to be

real such that,

 1 = 0 and  2 = 2

31
CHAPTER 1 – MANUSCRIPT

Therefore,

1
| ±i = (|+i ± |−i);
212

1
| ±i = (|+i ±  |−i)
212

and

~
S = (|+i h−| + |−i h+|)
2

~
S = (− |+i h−| +  |−i h+|)
2

In the form of matrix.


⎛ ⎞ ⎛ ⎞ ⎛ ⎞
~⎜ 0 1 ⎟ ⎜ 0 − ⎟ ⎜ 1 0 ⎟
S = ⎜ ⎟ ; S = ~ ⎜ ⎟ ; S = ~ ⎜ ⎟
2⎝ ⎠ 2⎝ ⎠ 2⎝ ⎠
1 0  0 0 −1

The relations between S  S and S± :

S+ = S + S

S− = S − S

Define
~
S ≡ 
2

where the dimensionless   are called the Pauli matrices:


⎛ ⎞ ⎛ ⎞ ⎛ ⎞
⎜ 0 1 ⎟ ⎜ 0 − ⎟ ⎜ 1 0 ⎟
σ = ⎜

⎟ ; σ = ⎜
⎠ ⎝
⎟ ; σ = ⎜
⎠ ⎝
⎟

1 0  0 0 −1

The eigenvalues of these three matrixes are ±1 (Please check it after class!)

32
CHAPTER 1 – MANUSCRIPT

B 1.6.5 The Algebra of Spin Operators

1. The commutation relations

[S  S ] = −[S  S ] = ~S ;

[S  S ] =  ~S


⎛ ⎞
⎜  =  =  = 1 ⎟
⎜ ⎟
⎝ ⎠
 =  =  = −1
 is an antisymmetric tensor.

Eample:

[S  S ] = S · S − S · S

~ ~
= (|+i h−| + |−i h+|) (− |+i h−| +  |−i h+|)
2 2

~ ~
− (− |+i h−| +  |−i h+|) (|+i h−| + |−i h+|)
2 2

~2 ~2
= ( |+i h+| −  |−i h−|) − (− |+i h+| +  |−i h−|)
4 4

= ~S ;

2. The anticommutation relations:

1
{S  S } = ~2  
2

33
CHAPTER 1 – MANUSCRIPT

3. The increasing and decreasing operators: S± = S ± S

[S  S± ] = [S  S ] ± [S  S ] = S ± (−S )

= ±S + S ≡ ±S±

[S−  S+ ] = −2S

4. The square operator of S


⎛ ⎞
3 ⎜ 1 0 ⎟ ⎟ = 3 ~2
S2 ≡ S2 + S2 + S2 = ~2 ⎜
⎝ ⎠ 4
4
0 1
µ ¶
3 1 1
= × 1+
4 2 2

Note: The spin 1/2 operator is neither fermion nor boson.

B 1.6.6 Algebra of the Pauli matrices

    =   +   

B 1.6.7 Observable

Compatible Observables: Observables A and B are defined to be compatible when

the corresponding operators commute,

[A B] = 0

34
CHAPTER 1 – MANUSCRIPT

and incompatible when

[A B] 6= 0

Degeneracy: Suppose there are two or more linear independent eigenkets of

A having the same eigenvalues then the eigenvalues of the two eigenkets are said to be

degenerate.

If a ket |i is an eigenket for both A and B, i.e.

A |i =  |i B |i =  |i

the ket |i is a simultaneous eigenket of A and B. We can denote it by |i = | i e.g.

[S2  S ] = 0

3 2 ~
S2 |±i = ~ |±i ; S |±i = ± |±i
4 2

3 2
S2 ( |+i +  |−i) = ~ ( |+i +  |−i)
4

~
S ( |+i +  |−i) = ( |+i −  |−i)
2

Theorem 1 Suppose that A and B are compatible observable, and eigenvalues of A

are nondegenerate. Then the matrix elements h” |B| 0 i are all diagonal. (A |i =  |i

¯ ®
G Proof. Suppose A is diagonal on the basis ket {¯() }

¯ ® ¯ ®
A ¯( = () ¯()

35
CHAPTER 1 – MANUSCRIPT

and all () 6= () if  6=  A and B are compatible, [A B] = 0

­ () ®
 |[A B]| ()

­ () ®
=  |[AB − BA]| ()

­ ®
= (() − () ) () |B| () = 0

­ () ®
=⇒  |B| () = 0 if  6= 

If A and B are compatible, we can always diagonalize them on one set of basis

ket simultaneously no matter whether A has degenerate eigenvalues!

Measurement of A and B when [A B] = 0

1. Nondegenerate

|i −→ | i

2. Degenerate

P ¯ ®
|i −→ () ¯0  ()

¯ ®
−→ ¯0  ()

Whether or not there is degeneracy, A measurements and B measurements do

not interfere. We can measure A and B accurately at the same time!

Incompatible Observables

36
CHAPTER 1 – MANUSCRIPT

Incompatible observables do not have a complete set of simultaneous eigenket.

Two imcompatible observables mean [A B] 6= 0

G Proof. Proof: Assume A and B are incompatible and {|0  0 i} is a complete set of

simultaneous eigenkets. Then




AB |0  0 i = A0 |0  0 i = 0 0 |0  0 i ⎪


BA |0  0 i = B0 |0  0 i = 0 0 |0  0 i ⎪

(AB − BA) |0  0 i = [A B] |0 0 i = 0 for all {0 0 }

=⇒ [A B] = 0

This is in contradiction to the incompatible assumption!

Accidently, there may exist an simultaneous eigenket for incompatible A and

B. For instance consider a state with the angular momentum  = 0. Although the

angular momentum opertors L and L do not commute, the state is a simultaneous

eigenstate of these two operators.

A 1.7 Change of Basis

EXP Example 1

Spin 1/2 system:

37
CHAPTER 1 – MANUSCRIPT

Basis I:{| +i  | −i}

~
S = (| +i h +| − | −i h −|)
2

~
S = (| +i h −| − | −i h +|)
2

~
S = (− | +i h −| +  | −i h +|)
2

Basis II: {| +i  | −i}

~
S = (| +i h +| − | −i h −|)
2

~
S = (| +i h −| − | −i h +|)
2

~
S = (− | +i h −| +  | −i h +|)
2

Connection of two basis

1 1
| ±i = √ | +i ± √ | −i
2 2
⎛ ⎞ ⎛ ⎞⎛ ⎞
⎜ | +i ⎟ ⎜ √1 √1 ⎟ ⎜ | +i ⎟
⎜ ⎟ = ⎜ 2 2 ⎟⎜ ⎟
⎝ ⎠ ⎝ ⎠⎝ ⎠
| −i √1 − √12 | −i
2

B 1.7.1 Transformation Operator

Suppose we have two incompatible observables A and B. The ket space can be viewed

as being spanned either by the set {|0 i} or by the set {|0 i}. i.e. A representation or

B representation. Changing the set of base kets is referred to as a change of basis or

change of representation. The two bases are connected by a transformation operator.

38
CHAPTER 1 – MANUSCRIPT

Theorem 2 Given two complete and orthogonal sets of base kets, there exist a unitary
¯ ® ¯ ®
operator U such that ¯() = U ¯() 

Unitary operator: UU+ = U+ U ≡ 1

G Proof. This theorem is proved by an explicit construction of u: the summation of the

out-product of two sets of eigenket forms an operator

P ¯¯ ® ­ () ¯
U= ()  ¯

1)
¯ ® P¯ ®­ ® ¯ ®
U ¯() = ¯() () |() = ¯()
 | {z }

2)

P ¯¯ ® ­ () ¯ () ® ­ () ¯


UU+ = ()  ¯  ¯


P ¯¯ ® ­ () ¯
= ()  ¯=1

B 1.7.2 Transformation Matrix

In the matrix representation, the operator U can be expressed in the form of a matrix
¯ ®
in the base ket {¯() }

­ ® ­ ®
U = { () |U| () } = { () |() }

39
CHAPTER 1 – MANUSCRIPT

For an arbitrary ket in two representations

A:
P ¯¯ () ® ­ () ®
|i =   |

B:
P ¯¯ () ® ­ () ®
|i =   |

Relation of two representations:

­ ® P ­ () () ® ­ () ®


() | =  |  |

P ­ () ¯¯ † ¯¯ () ® ­ () ®


=  U   |

Relation of an operator in two representations

­ () ®
 |X| ()

P ­ () () ® ­ () ®­ ®


=  |  |X| () () |()


P ­ () ¯¯ † ¯¯ () ® ­ () ®­ ®


=  U   |X| () () |U| ()


­ () ¯ † ¯ ®
=  ¯ U X U ¯()

=⇒ X = U† X U

− − − − − − A similarity transformation

40
CHAPTER 1 – MANUSCRIPT

Trace of an operator: the summation of the diagonal elements

P ­ () ¯¯ ¯¯ () ®
Tr(X) ≡  X 

P ­ () () ® ­ () ®­ ®


=  |  |X| () () |()


P ­ () () ® ­ () () ® ­ () ®


=  |  |  |X| ()


P ­ () () ® ­ () ®


=  |  |X| ()


P ­ () ®
=  |X| ()

The trace of X is independent of representation. We can also prove:

Tr(XY) = Tr(YX); Tr(U† XU) = Tr(X)

B 1.7.3 Diagonalization

When we know the matrix elements of B in the base {|0 i} how do we obtain the

eigenvalues and eigenket of B?

¯ ® ¯ ®
B ¯() =  ¯()

41
CHAPTER 1 – MANUSCRIPT
¯ ®
In the base ket {¯() } the operator is expressed as

¯ ®­ ¯ ¯ ®­ ¯
B ¯() () ¯ =  ¯() () ¯

X¯ ®­ ¯ X ¯ ®­ ¯
B ¯() () ¯ =  ¯() () ¯
 
X ¯ ®­ ¯
B =  ¯() () ¯

In the Dirac bra-ket notation, we rewrite it as

P ­ () ®­ ® ­ ®
 |B| () () |() =  () |() 

Furthermore, in the matrix form,


⎛ ⎞⎛ ⎞ ⎛ ⎞
() ()
⎜ 11 12 · · · ⎟ ⎜ 1 ⎟ ⎜ 1 ⎟
⎜ ⎟⎜ ⎟ ⎜ ⎟
⎜ ⎟⎜ ⎟ ⎜ ⎟
⎜   · · · ⎟ ⎜  () ⎟ =  ⎜  ()
() ⎟
⎜ 21 22 ⎟⎜ 2 ⎟ ⎜ 2 ⎟
⎜ ⎟⎜ ⎟ ⎜ ⎟
⎝ . .. . . ⎠ ⎝ .. ⎠ ⎝ . ⎠
.. . . . ..

where
­ ®
 = () |B| ()

and
() ­ ®
 = () |()

As we know from linear algebra, the eigenvalues are determined by

det(B − 1) = 0

EXP Example 2

42
CHAPTER 1 – MANUSCRIPT
⎛ ⎞
~⎜ 0 1 ⎟
S = ⎜ ⎟
2⎝ ⎠
1 0

⎛ ⎞
~
⎜ − 2 ⎟
det(S − 1) = 0 ⇒ det ⎜

⎟=0

~
2
−

~2 ~
2 − = 0⇒=±
4 2

Eigenvectors?

A 1.8 Position, Momentum, and Translation

B 1.8.1 Continuous Spectra

When the eigenvalues of an observable is continuous, the dimensionality of the vector

space is infinite. Denote the eigenvalue-eigenket relation by

̂ |i =  |i

43
CHAPTER 1 – MANUSCRIPT

We generalize of a vector space with finite dimension to that with infinite dimension

discrete continuous

h0 |”i =  0 ” h 0 |”i = ( 0 − j)


P R
|0 i h0 | = 1  |i h| = 1
P ¯¯ () ® ­ () ¯¯ R
|i =   i |i =  |i h| i

h” |A| 0 i = 0  0 ” h” ||  0 i =  0 ( 0 − j)

1. The Kronecker symbol  is replaced by Dirac’s  -function.

2. The discrete sum over the eigenvalues is replaced by an integral over a continuous

variable.

B 1.8.2 Some properties of the −function.


R
1. () = 1

2. () = (−)

3.  0 () = − 0 (−)

4. () = 0

5. 0 () = −(−)

6. () = 1 (−)

7. (2 − 2 ) = 12 (( − ) + ( − ))

44
CHAPTER 1 – MANUSCRIPT

8. ( − ) ·  () = ( − )()

R
9. ( − )( − ) = ( − )

Representation of the −function

sin 
() = lim
→∞ 

1 2  2
() = lim −
→0  12 

Normalization in terms of the −function

 () =  ·

R
+∞ R
+∞
∗ () () = 2 −( − )
−∞ −∞

R
+ sin( −  )
= 2 lim −( − ) = 2 2 lim
→∞ − →+∞ ( −  )

= 2 2( −  ) = ( −  )


⇒ 2 2 = 1 =⇒  = 1 2

B 1.8.3 Position Eigenkets and Position Measurements

To extend the idea of a filtering process to measurements of observables exhibiting

continues spectra, we consider the position operator in one dimension. The eigenkets

|0 i of the position operator x satisfying

x |0 i = 0 |0 i

45
CHAPTER 1 – MANUSCRIPT

are postulated to form a complete set. The state ket for an arbitrary physical states

can be expanded in terms of {|0 i}


Z
|i = 0 |0 i h0 |i

Suppose we place a very tiny detector that clicks only when the particle is precisely

at 0 and nowhere else after the detector clicks, we can say the state in question is

represented by |0 i. In practice the best the detector can do is to locate the particle

∆ ∆
within a narrow range (0 − 2
 0 + 2
) (∆ is very small). When the detector clicks,

the state ket changes abruptly as follows


Z  + ∆

0 0
R
2
|i = 0 | i h |i measurement
−−−−−−−−−→ 0 |0 i h0 |i 
−∞  − ∆
2

Assume that h0 |i does not change appreciably within the narrow interval, the prob-

ability that the detector clicks is given by

|h |i|2 ∆

Obviously,
R
+∞
|h|i|2  = 1 = h|i 
−∞

In the wave mechanics, the physical state is represented by a wave function. Thus h|i

is the wave function in the position space for the state |i
R
Wave function in position space: h0 |i in |i = 0 |0 i h0 |i is the wave

function in the position space. We assume that the position eigenkets |0 i are complete.

46
CHAPTER 1 – MANUSCRIPT

It can be three dimensional, i.e. 0 in |0 i stand for three components of the position

vectors. In this word the position is an observable, and [   ] = 0

B 1.8.4 Translation

Suppose we start with a state that is well located around 0 , and introduce an translation

operator T() such that

T() |i = | + i 

Several properties of T():

1. Identity operation: T( = 0) = 1

2. Associative: T(1 )T(2 ) = T(1 + 2 )

3. Inverse operation: T(−)T() = 1 ⇐⇒ T(−) = T−1 ()

4. Probability conservation: T()† T() = 1

From the definition we have

h| T()† = h + | 

­ ¯ ¯ ®
 ¯T()† T()¯  = h + | + i = h|i 

This leads to the property (4). We now consider an infinitesimal translation  =  → 0.

We demonstrate that if T() is taken to be

T() ≈ 1 − K + 0(()2 )

47
CHAPTER 1 – MANUSCRIPT

where K is an Hermitian operator K† = K then all properties listed are satisfied.

Let us check them one by one.

1.  = 0 :

T( = 0) = 1

2. T(1 )T(2 ) = T(1 + 2 ) :

(1 − K1 )(1 − K2 )

= 1 − K(1 + 2 ) + 0(1 2 )

3. T(−)T() = 1

(1 + K)(1 − K) = 1 − K2 ()2 = 1

4. T()† T() = 1

(1 − K)† (1 − K) = (1 + K† )(1 − K)

= 1 + (K† − K) + 0(()2 )

48
CHAPTER 1 – MANUSCRIPT

Commutation relation of x and K:

xT() |i = x | + i = ( + ) | + i

T()x |i = T() |i = T() |i =  | + i

=⇒ (xT() − T()x) |i =  | + i

=⇒ (−xK + Kx) |i =  | + i ≈  |i

=⇒ [x K] = 

x and K do not commute with each other!

Momentum Operator in the Position Basis.

Translation Operator for an infinitesimal small displacement is defined as

(T() = 1 − K). Let T() act on an arbitrary state

Z Z
|i → T() |i = T() |i h|i =  | + i h|i

Z
=  |i h − |i  +  → 

Z Z
 
=  |i (h|i − h|i ) =  |i (1 −  ) h|i
 
Z Z
also =⇒  |i h| T() |i =  |i h| (1 − K) |i

Z

=  |i (h|i − K h|i ) =⇒ K = −


49
CHAPTER 1 – MANUSCRIPT

Consider an ideal wave function

1
h|i = 12
+x~
(2)

which has the momentum 



K h|i = h|i
~

The physical significance of K is that K is an operator for the momentum with a

proportionality factor. The factor is the Planck’s constant over 2 : ~ = 2

L. de Broglie’s relation (1924):

2 
=
 ~

de Broglie wave is also called matter wave, any aspect of the behavior or

properties of a material object that varies in time or space in conformity

with the mathematical equations that describe waves. By analogy with the

wave and particle behavior of light that had already been established exper-

imentally, the French physicist Louis de Broglie suggested (1924) that par-

ticles might have wave properties in addition to particle properties. Three

years later the wave nature of electrons was detected experimentally. Ob-

jects of everyday experience, however, have a computed wavelength much

smaller than that of electrons, so their wave properties have never been de-

tected; familiar objects show only particle behavior. De Broglie waves play

50
CHAPTER 1 – MANUSCRIPT

an appreciable role, therefore, only in the realm of subatomic particles. The

response of the wave properties of a particle to an external force follows a

basic law of quantum mechanics that, in its mathematical form, is known

as the Schrödinger equation.

The momentum operator in the position space


p = −~ 


The commutation relation:

[x p] = ~

 
[x p] () = (xp − px) () = −~  () + ~ (())
 

 
= −~ () + ~ () + ~ ()
 

Therefore

[x p]() = ~ ()

In the momentum space,

[x p]() = ~()

[x p]() = (xp − px)()


µ ¶
 
= ~  − ~ () = ~()
 

51
CHAPTER 1 – MANUSCRIPT


x =~ 


A 1.9 The Uncertainty Relation

The commutation relation [x p] = ~ shows that the position and the momentum are

incompatible. This relation leads to the famous uncertainty relation

­ ®­ ®
(∆x)2 (∆p)2 ≥ ~2 4

Let A ad B be observables. Define ∆A ≡ A − hAi. The expectation value of h(∆A)2 i

is known as the dispersion of A. For any pair of operators, A and B, we have

­ ®­ ®
(∆A)2 (∆B)2 ≥ |h[A B]i|2 4

The uncertainty relation can be understood from the postulate of measurement. For

an eigenstate of A,
­ ®
(∆A)2 = 0

If a simultaneous eigenstate of A and B, we have

­ ®­ ®
(∆A)2 (∆B)2 = 0

Before presenting a proof of the uncertainty relationship, I introduce several examples

first.

52
CHAPTER 1 – MANUSCRIPT

EXP Example 3

A plane wave with a momentum 0 :

1
0 () = exp[0 ~]
(2~)12
¯ ¯2
The probability to find x is a constant in the whole space, ¯0 ()¯ =

1
2~

 = 0

­ 2®
hi → 0;  → +∞

EXP Example 4

A particle at  = 0 :

0 () = ( − 0 )

In the momentum space,

0 () = ( − 0 )

Z
 (−0 )~
= 
2~

() = 0 ~

53
CHAPTER 1 – MANUSCRIPT

 = 0

­ 2®
hi → 0;  → +∞

EXP Example 5

A Gaussian wave package:

1
() = 12
exp[−2 220 ]
 14 0

The particle is mainly confined in the regime,  ∼ 0 . In the k

space,
Z
1
() = ()−~ 
(2)12
Z
1 1
= 12
exp[−2 220 − ~]
(2)12  14 0
Z µ ¶2
1 1 1  0 1 ³ 0 ´2
= 12
exp[− + ] exp[− ]
(2)12  14 0 2 0 ~ 2 ~
√ 14 12 Z µ ¶2
2 0 1 1  0 1 ³ 0 ´2
= √ exp[− + ] exp[− ]
(2)12  12 20 2 0 ~ 2 ~

0
= 12
exp[−2 20 2~2 ]
 14 0

The particle is mainly confined in the regime,  ∼ ~0  Thus

 ∼ ~

The proof of the uncertainty relationship

54
CHAPTER 1 – MANUSCRIPT

G Proof.

­ ® ­ ® ­ ®
(∆A)2 = (A − hAi)2 = A2 − 2A hAi + hAi2

­ 2®
= A − hAi2 ≥ 0

1) The Schwarz inequality

h|i · h|i ≥ |h|i|2

where the equality only holds if |i = |i 




(h| + ∗ h|)(|i +  |i) ≥ 0 ⎪



taking  = − h|i ⎭
h|i

=⇒ h|i h|i ≥ |h|i|2

To simplify the problem, we can take h|i = h|i = 1 and |h|i|2 ≤ 1 Let

|i = ∆A |i

|i = ∆B |i

­ ®­ ®
=⇒ (∆A)2 (∆B)2 ≥ |h∆A∆Bi|2

where we have used the hermiticity of A and B since they are observables,

∆A = (∆A)† 

55
CHAPTER 1 – MANUSCRIPT

2)

1
∆A∆B = (∆A∆B − ∆B∆A)
2

1
+ (∆A∆B + ∆B∆A)
2

1 1
= [∆A ∆B] + {∆A ∆B}
2 2

1 1
= [A B] + {∆A ∆B}
2 2

Anti-Hermitian of [A B]:

[A B]† = (AB − BA)†

= (AB)† − (BA)†

= B† A† − A† B†

= −[A B]

Hermitian of {∆A ∆B}

({∆A ∆B})+ = {∆A ∆B}

56
CHAPTER 1 – MANUSCRIPT

3) a. The Hermitian operator has only purely real expectation value





⎨ H |i =  |i

H = H =⇒ ;


⎩ h| H† = h| ∗




⎨ h |H| i = 
=⇒ =⇒  = ∗

⎪ ­ ¯ ¯ ®
⎩  ¯H† ¯  = ∗

b. The anti-Hermitian operator has only purely imaginary expectation value

Q† = −Q

Q |i =  |i  h| Q† = ∗ h|

­ ¯ ¯ ®
h |Q| i =   ¯Q† ¯  =  ∗

=⇒  = − ∗

Therefore

1 1
h∆A∆Bi = h[A B]i + (h{∆A ∆B}i)
2 2

= + Im h∆A∆Bi + Re h∆A∆Bi

1 1
|h∆A∆Bi|2 = |h[A B]i|2 + |h{∆A ∆B}i|2
4 4

1
≥ |h[A B]i|2
4
­ ®­ ® 1
(∆A)2 (∆B)2 ≥ (h[A B]i)2
4
57
CHAPTER 1 – MANUSCRIPT

For the position and momentum operators,

[x p] = ~

­ ®­ ®
(∆x)2 (∆p)2 ≥ ~2 4

Since the position and momentum operators are incompatible, [x p] 6= 0 we cannot

find a complete set of simultaneous eigenkets for x and p such that the two observables

can be measured at the same time. For instance for an eigenstate of x, x |i =  |i, it

is not an eigenstate of the momentum. It can be spanned in the momentum space,


Z Z
p |i = p  |i h|i =  |i h|i 

You may have an expectation value for p.

EXP Example 6

Momentum Determination Experiment

We assume that the particle is an atom in an excited state,

which will give off a photon that has the frequency  0 if the atom is at

rest. Because of the Doppler effect, the motion of the atom toward

the observer with speed  means that the observed frequency is

given approximately by

 
 ≈  0 (1 + ) =⇒  = ( − 1)
 0

58
CHAPTER 1 – MANUSCRIPT

Accurate measurement of the momentum  by measurement of

the frequency  requires a relatively long time  ; the minimum error

in the frequency measurement can be shown to be ∆ ≈ 1  The

instant at which the photon is emitted is uncertain by  , at this

instant, the momentum of the atom decreases by ~ , and its

velocity decreases by ~ . This makes the subsequent position

~
of the atom uncertain by the amount ∆ ≈ 
·   Since the later

the photon is emitted, the longer the atom has the higher velocity

and the farther it will have traveled. This position uncertainty arises

entirely because of the finiteness of  . If  were zero, and we knew

the velocity and the velocity change on emission of photon, we would

know where the atom is at each instant; it is because  is finite that

we do not know when the velocity changed and hence where the

atom is later times. The momentum uncertainty is obtained:


∆ ≈ ∆ ≈ ( − 1)
0

 
≈ ∆ ≈
0  0

In the nonrelativistic case considered here    1 =⇒  ≈  0

Therefore

∆ · ∆ ∼ ~

59
CHAPTER 1 – MANUSCRIPT

EXP Example 7

2 × 2 matrix: Suppose a 2 × 2 matrix X is written as

 = 0  0 +  · 
⎛ ⎞ ⎛ ⎞
⎜ 0 0 ⎟ ⎜ 0  ⎟
= ⎜

⎟+⎜
⎠ ⎝


0 0  0
⎛ ⎞ ⎛ ⎞
⎜ 0 − ⎟ ⎜  0 ⎟
+⎜

⎟+⎜
⎠ ⎝


 0 0 −
⎛ ⎞
⎜ 0 +   −  ⎟
= ⎜



 +  0 − 

where 0 and  are real.

Eigenvalues of X:

det( − ) = 0

60
CHAPTER 1 – MANUSCRIPT

⎛ ⎞
⎜ 0 +  −   −  ⎟
det ⎜



 +  0 −  − 

= (0 +  − )(0 −  − ) − ( −  )( +  )

= (0 − )2 − 2 − 2 − 2 = 0

¡ ¢12
=⇒ ± = 0 ± 2 + 2 + 2

61
CN Chapter 2

CT Quantum Dynamics

This chapter is devoted exclusively to the dynamic development of state kets and/or

observables. In other words, we are concerned here with the quantum mechanical

analogue of Newton’s equation of motion.

A 2.1 Time Evolution and the Schrödinger Equation

B 2.1.1 Time Evolution Operator

We now discuss how a physical state evolves with time. Suppose we have a physical

system whose state ket at 0 is represented by |i. Let us denote the ket corresponding

to the state at some later time  by

| 0 : i 

62
CHAPTER 2 – MANUSCRIPT

Our main task is to study the time evolution of a state ket:

|i = | 0 : 0 i =⇒ | 0 : i 

As in the case of translation, we assume that the two kets are connected by the time-

evolution operator

| 0 : i = U( 0 ) |i 

Several properties of U:

1. The identity

U(0  0 ) = 1

2. The unitarity

U† ( 0 )U( 0 ) = 1

3. The composition:

U(2  0 ) = U(2  1 )U(1  0 )

Because time is assumed to be continuous, we consider an infinitesimal time-

evolution operator

| 0 ; 0 + i = U(0 +  0 ) |i

As

lim U(0 +  0 ) = U(0  0 ) = 1


→0

63
CHAPTER 2 – MANUSCRIPT

we expand the operator in terms of 


U(0 +  0 ) = U(0  0 ) + U( 0 )|=0  + 


= 1 − Ω̂ + 

Just like the operator K in the translation operator T(), Ω̂ is a Hermitian operator.

What’s the physical meaning of Ω̂?

We consider a simple example. In the position space

h| 0  i = h|U( 0 )|i

⇓ ˜˜˜ ⇓

1
­ ®
(2)12
(−(−0 )) ˜˜˜˜˜˜˜ (2)112  |−Ω(−0 ) |

 → 0 +  =⇒ Ω → 

In the old quantum theory, the time-frequency  is postulated to be related to energy

 = ~ (Planck-Einstein Relation)

In the classical mechanics, the energy is related to the Hamiltonian operator H , which

is also the generator of time-evolution. It is then natural to relate or define Ω̂ to the

Hamiltonian operator H:

Ω̂ = H~

Thus

64
CHAPTER 2 – MANUSCRIPT

H
(0 +  0 ) = 1 −   + 
~

~ in Ω̂ = H~ and K = p~: The two constants must be the same which can be checked

by comparing the quantum mechanical equation of motion with the classical equation

of motion, say
x p
=
 

B 2.1.2 The Schrödinger Equation.

We are now in the position to derive the fundamental differential equation for the time-

evolution operator U( 0 ) and a physical state |i  From the composition property of

U

U( +  0 ) = U( +  )U( 0 )


µ ¶
H
= 1 −   U( 0 )
~

U( +  0 ) − U( 0 ) 


= − HU( 0 )
 ~

Take  → 0+  we obtain

~ U( 0 ) = HU( 0 )


65
CHAPTER 2 – MANUSCRIPT

Multiplying a state ket from the right sides


~ U( 0 ) |i = HU( 0 ) |i



~ | 0 ; i = H | 0 ; i 


Here we list several formal solutions to the Schrodinger equation for the time-evolution

operator U( 0 )

Case A: The Hamiltonian operator is independent of time,


~ U( 0 ) = HU( 0 )


U( 0 )
~  = H
U( 0 )


~ ln U( 0 ) = H

Z 

ln U( 0 ) − ln U(0  0 ) = − H
~ 0


U( 0 ) = exp[− H( − 0 )]
~

Strictly speaking, the present derivation is not correct completely since U is an operator,

not number. However the final solution is exact,


∙ ¸2
  1 
exp[− H( − 0 )] = 1 − H( − 0 ) + − H( − 0 ) + · · ·
~ ~ 2 ~

Case B: The Hamiltonian H is time-dependent but the H0  at different times

commute
Z  Z 

 ln U( 0 ) = − H()
0 ~ 0

66
CHAPTER 2 – MANUSCRIPT
∙ Z ¸
 
U( 0 ) = exp − H()
~ 0

Case C: The Hamiltonian H is time-dependent but the H0 s at different times

do not commute, i.e.

[H(1 ) H(2 )] 6= 0
Z  Z  Z 1
 (−)2
U( 0 ) = 1 − 1 H(1 ) + 1 2 H(1 )H(2 ) + · · ·
~ 0 ~2 0 0

Z  Z 1 Z  −1
P∞ −
= 1+ ( ) 1 2   × H(1 )H(2 ) · · · H( )
=1 ~ 0 0 0

Energy Eigenkets:

To be able to evaluate the effect of time-evolution operator on a general ket

|i, we consider a special base ket used in expanding |i  Assume that we know all

energy eigenkets of an time-independent H with eigenvalues 

¯ ® ¯ ®
H ¯() =  ¯()

¯ ®
{¯() } forms a complete and orthogonal set of base kets. Therefore

U( 0 ) = exp(−H( − 0 )~)

P ¯¯ () ® ­ () ¯¯ − H (−0 ) ¯¯ () ® ­ () ¯¯


=    ~  


P  ¯ () ® ­ () ¯
= − ~
(−0 ) ¯  ¯

¯ ®
Once the expansion of the initial ket |i in terms of {¯() } is known,

P ¯¯ ® ­ () ¯ P ¯ ®
|i = ()  ¯ i =  ¯() 
 

67
CHAPTER 2 – MANUSCRIPT

The time-evolution operator enables us to solve any initial time problem,

H
| 0 ; i = − ~ (−0 ) |i

P  ¯ () ®
=  − ~
(−0 ) ¯ 


In other words

 ( = 0 ) =  →  () =  − ~
(−0 )

The phase of  () changes with time, but its modulus remains unchanged. If all

eigenvalues are not degenerate, the relative phases of  do vary with time and the

state ket varies with time.

A special case: if the initial state happens to be one of energy eigenkets, say
¯ ®
|i = ¯()  we have

 ¯ () ®
| 0 ; i = − ~
(−0 ) ¯ 

The state remains unchanged. Thus the basic task in quantum dynamics is reduced

to find a complete set of energy eigenket and eigenvalues. It is very helpful if we can

find a maximal set of commuting observables A, B, C,...which also commute with the

Hamiltonian

[A B] = [B C] = [A C] = · · · = 0

[A H] = [B H] = [C H] = · · · = 0

68
CHAPTER 2 – MANUSCRIPT

In the case, the energy eigenket can be characterized by the indices {  } ≡ {}

For example, ⎧



⎪ A |i =  |i 






⎨ B |i =  |i 



⎪ ···






⎩ H |i =  |i

Therefore it is of fundamental importance to find a complete set of mutually compatible

observables that also commute with H.

B 2.1.3 Time Dependence of Expectation Value: Spin Preces-

sion.

It is instructive to study how the expectation value of an observable changes as a

function of time. To end this we treat an example here: spin precession. We start with

a Hamiltonian of a spin 1/2 system with magnetic moment  = ~  subjected to

an external magnetic field B = e

 
=− S·B=− S = S
   

As ⎛ ⎞
~⎜ 1 0 ⎟
S = ⎜ ⎟
2⎝ ⎠
0 −1

69
CHAPTER 2 – MANUSCRIPT

then the corresponding eigenvalues of Ĥ are

~
± = ±  (=  )
2

Since the Hamiltonian is time-independent, the time-evolution operator for the state

ket

( 0) = −S ~ 

The eigenkets of S are


⎛ ⎞
⎜ 1 ⎟ ~
|+i = ⎜
⎝ ⎠
⎟ for  = +
2
0
⎛ ⎞
⎜ 0 ⎟ ~
|−i = ⎜
⎝ ⎠
⎟ for  = −
2
1

1
=⇒  |±i = ± ~ |±i .
2

i.e., |±i are also the energy eigenkets with eigenvalues ±~2 Suppose at  = 0

|i = + |+i + − |−i 

At a later time t,

| 0 : i =  ( 0 ) |i

 
= + − 2 |+i + − + 2 |−i 

70
CHAPTER 2 – MANUSCRIPT

Specifically, (1) Let us suppose the initial state |i is the spin up state |+i i,e. + = 1

and − = 0 then

| 0 : i = − 2 |+i

This is a stationery state ! (2) Suppose the initial state |i is the spin state | +i :

1 1
| +i = |+i + |−i
212 212

At a later time t
1  1 
| 0 : i = − 2 |+i +  2 |−i
212 212

The probability for the system to be found in the states | ±i = √1 (|+i ± |−i):
2

h + | 0 : i

1 1  
= (h+| + h−|) (− 2 |+i +  2 |−i)
212 212

1 −   
= ( 2 +  2 ) = cos
2 2

h − | 0: i = − sin
2


|h + | 0: i|2 = cos2 ;
2


|h − | 0: i|2 = sin2 
2

The total probability is conserved,

|h + | 0: i|2 + |h − | 0: i|2 = 1

71
CHAPTER 2 – MANUSCRIPT

The expectation value of S is

hS i = h 0:  |S |  0: i

~ ~
= |h + | 0: i|2 − |h − | 0: i|2
2 2

~
= cos 
2

Similarly, we have

~
hS i = sin ;
2

hS i = 0

Question: what happens if we take B = e ?

A 2.2 The Schrodinger versus the Heisenberg Picture

B 2.2.1 Unitary operators

Unitary operators are used for many different purposes in quantum mechanics. It

satisfies that U† U = 1 Under the unitary transformation that changes the state kets

it is important to keep in mind that the inner product of a state bra and a state ket

remains unchanged,
­ ®
h|i = |U† U| 

72
CHAPTER 2 – MANUSCRIPT

Another mathematical observation is that

¡ ¢ ¡ ¢
h| U† X (U |i) = h| U† XU |i

that follows from the associative axiom of multiplication. This identity suggests two

different approaches to evaluate the expectation values of physical observables under

the unitary transformation.

B 2.2.2 Two Approaches

When we are concerned with a time-dependent of a physical observable X there are

two approaches:

Approach I:

|i → | 0 : i = U( 0 ) | 0 i

­ ¯ ¯ ®
hXi = h 0 : | X | 0 : i =  ¯U† XU¯ 

Approach II:

|i → |i

X → U† XU ≡ X()

­ ¯ ¯ ®
hXi = h |X()| i =  ¯U† XU¯ 

The expectation values of two approaches are the same!

hXi = hXi 

73
CHAPTER 2 – MANUSCRIPT

Approach I (The Schrodinger Picture): The state ket varies with time while

operators do not change

|i → U |i  X → X

Recall the Schrodinger equation for the wave function, i.e., the state.

Approach II (The Heisenberg Picture): The operators vary with time while

the ket does not change

|i → |i  X → U† XU

Recall the Heisenberg equation for operators.

We denote the operator X() in the Heisenberg picture. U() = U( 0 ) The

relation of an observable in two pictures

X() () = U† ()X() U()

The state ket

| 0 : i = | 0 = 0i ;

| 0 : i = U() | 0 = 0i = U() | 0 : i 

74
CHAPTER 2 – MANUSCRIPT

B 2.2.3 The Heisenberg Equation of Motion.

We now derive the fundamental equation of motion in the Heisenberg picture. Assume

that X() does not depend explicitly on time

 ()  £ † ¤
X = U ()X() U()
 
µ ¶
 † 
= U () X() U() + U† ()X() U()
 

1 † 1
= − U ()HX() U() + U† ()X() HU()
~ ~

1 † 1
= + U ()X() U()U† ()HU() − U† ()HU()U† ()X() U()
~ ~

1 ()
= [X  H() ]
~

 ()
~ X = [X()  H() ]


This is the Heisenberg equation of motion! In the case that H is time-independent or

the H ’s at different times commute we have

H() = U† HU = H

Thus
 ()
~ X = [X()  H]


75
CHAPTER 2 – MANUSCRIPT

B 2.2.4 How to construct a Hamiltonian

1. For a physical system with classical analogue we assume the Hamiltonian to be

of the same form as in classical physics; we merely replace the classical 0 and

0 by the corresponding operators x and p in quantum mechanics. With this

assumption we can reproduce the correct classical equation in the classical limit.

It is required that the Hamiltonian in quantum mechanics should be Hermitian.

2. When the physical system in question has no classical analogues, we can only

guess the structure of the Hamiltonian. We try various forms until we get the

Hamiltonian that leads to results agreeing with empirical observation.

Two useful formulas:


[x   (p)] = ~ 


and

[p  (x)] = −~ 


where F and G are functions that can be expanded in powers of 0  and 0  respectively.

EXP Example 1

A free particle: The Hamiltonian to describe the motion of a free

particle of mass  is taken to be of the same form as in classical

76
CHAPTER 2 – MANUSCRIPT

mechanics:
p2 1
H= = (p2 + p2 + p2 )
2 2

We look at the observables  and  (they are operators, NOT num-

bers)
 1
p = [p  H] = 0
 ~
 1 p p (0)
x = [x  H] = =
 ~  

So
p (0)
x () = x (0) + 

We know

[x (0) x (0)] = 0

but
~
[x () x (0)] = − 

EXP Example 2

A particle in a potential.

p2
H= + V()
2

77
CHAPTER 2 – MANUSCRIPT

The Heisenberg equation of motion leads to

p 1 
= [p V()] = − V()
 ~ 

x 1 p
= [x H] =
 ~ 

Furthermore,

2 x 1 x 1 p 1 p
2
= [  H] = [  H] ≡
 ~  ~   

2 x
= −∇ ()
2

Newtonian Equation?

The expectation value of this equation is known as the Ehrenfest

theorem. It has the classical mechanics analogue. The reason is

that the Planck constant disappears in the equation.

A 2.3 Simple Harmonic Oscillator.

B 2.3.1 Eigenvalue and eigenstates

We consider a simple harmonic oscillator in a one-dimensional case. The basic Hamil-

tonian is
2 
= +  2 2
2 2
78
CHAPTER 2 – MANUSCRIPT

where  and  are Hermitian since they are physically dynamic variables. The quantum

condition is

[ ] =  −  = ~

As the first step to solve the problem, we evaluate the commutator basket of  and 

with 
1 ~
[ ] = [ 2 ] = 
2 
 2
[ ] = [ 2 ] = −~ 2 
2

Adding the two equations, we have

~
[ +  ] = ~ 2  + 

1 1
= ~ 2 ( +  )
 2 2

where  is a constant. Here we take

1
=±


We have
 
[ ±   ] = ±~( ±  )
 

As
  2~
[ +   −  ]=
  

we take
³  ´12  ³  ´12 
= ( +  ); † = ( −  )
2~  2~ 
79
CHAPTER 2 – MANUSCRIPT

such that

[ † ] = 1

Furthermore,
    1
†  = ( −  )( +  )= −
2~   ~ 2

Thus
1
 = ~( + )
2

where

 = † 

Obviously,

[ ] = 0

  and  can be diagonalized simultaneously. We denote the eigenkets of  by its

eigenvalues n. So

 |i =  |i

and
1
 |i = ( + )~ |i
2

What’s ?

We first note that

[ ] = †  − †  = [ † ] = 

80
CHAPTER 2 – MANUSCRIPT

[†  ] = −†

As a result,

© ª
† |i = [ † ] + †  |i

= ( + 1)† |i 

Likewise,

 |i = ( − 1) |i

Thus † |i is also an eigenket of  with eigenvalue  + 1 (increased by one).  |i is

an eigenket of  with eigenvalue  − 1 (decreased by one). So we call a the decreasing

operator and † the increasing operator. From the point of view of energy, the increase

(decrease) of n by one amounts to the creation (annihilation) of one quantum unit of

energy. Hence the terms “creation and annihilation operator” for † and  are deemed

appropriate. Since  |i is an eigenket with  − 1 . We write

 |i =  | − 1i

(h| + )( |i) = h − 1| (∗  ) | − 1i

­ ¯ + ¯ ®
 ¯ ¯  =  = | |2 ≥ 0

81
CHAPTER 2 – MANUSCRIPT

We take  real. then,

 |i = 12 | − 1i

Likewise,

† |i = ( + 1)12 | + 1i

Suppose we keep on applying a to both sides of the equation,

2 |i = [( − 1)]12 | − 2i ;

3 |i = [( − 1)( − 2)]12 | − 3i 

The sequence of the eigenket must terminate at  = 0 as  ≥ 0 If  is a non-integer,

the sequence shall not terminate and leads to eigenkets with negative eigenvalues. This

is in contradiction with  ≥ 0 . To conclude n is a nonnegative integer. The lowest

energy state, i.e. the ground state of the oscillator is

1
=0 = ~
2

The eigenstate eigenvalue

1
|0i 2
~

3
|1i = † |0i 2
~

1 5
|2i = 212
(† )2 |0i 2
~

1 7
|3i = (3!)12
(† )3 |0i 2
~
.. ..
. .

1
|i = (!)12
(† ) |0i ( + 12 )~

82
CHAPTER 2 – MANUSCRIPT

Position and Momentum


µ ¶12
~
 = ( + + )
2
µ ¶12
~
 =  (− + + )
2

From the orthonormality conditions

h0 || i = 12  0 −1

­ 0 ¯ †¯ ®
 ¯ ¯  = [ + 1]12  0 +1

In the matrix form, ⎛ ⎞


⎜ 0 1 0 0 ··· ⎟
⎜ ⎟
⎜ ⎟
⎜ 0 0 212 0 ··· ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
=⎜
⎜ 0 0 0 312 ··· ⎟

⎜ ⎟
⎜ ⎟
⎜ 0 0 0 0 ··· ⎟
⎜ ⎟
⎜ ⎟
⎝ .. .. .. .. ... ⎠
. . . .
⎛ ⎞
⎜ 0 0 0 0 ··· ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 0 0 0 ··· ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
 =⎜

⎜ 0 212 0 0 ··· ⎟ ⎟
⎜ ⎟
⎜ ⎟
⎜ 0 0 312 0 ··· ⎟
⎜ ⎟
⎜ ⎟
⎝ .. .. .. .. . . ⎠
. . . . .

83
CHAPTER 2 – MANUSCRIPT

The matrix elements of  and  are


µ ¶12
0 ~
h || i =
2

(12  0 −1 + ( − 1)12  0 +1 )


µ ¶12
0 ~
h || i = 
2

(−12  0 −1 + ( − 1)12  0 +1 )

In the position representation: the wave function

We first derive h|0i

 |0i = 0

=⇒ h || 0i = 0

Z
=⇒ 0 h || 0 i h0 |0i = 0

Recall
³  ´12 
= ( +  )
2~ 

h || 0 i = ( − 0 )


h || 0 i = −~( − 0 )


~ 
=⇒ ( + ) h|0i = 0
 

84
CHAPTER 2 – MANUSCRIPT

Figure 2.1: Energy eigenfunctions for the first six states of the harmonic oscillator.

Take ~ = 20  The solution of the equation is

1 1 2
h|0i = 1 12
exp(− 2 )
 4 0 2 0

In general
* ¯ ¯ + * ¯ ¯ +
¯ († ) ¯ ¯ († ) ¯
¯ ¯ ¯ ¯
h|i = ¯ ¯0 = ¯ ¯  h|0i
¯ (!) ¯
12 ¯ (!) ¯
12

− +1 2
0 2
2   − 220

= 1 ( − 0 ) 
 4 (2 !)12 

B 2.3.2 Time Development of the Oscillator

So far we have not discussed the time evolution of oscillator state ket or observables

like  and . Everything we did is supposed to hold at some instant of time, say  = 0.

85
CHAPTER 2 – MANUSCRIPT

Figure 2.2: The position probability density for the state n=10 of a harmonic oscillator

(solid curve) and for a classical oscillator of the same total energy (dashed line).

Now we come to see the time evolution of  and . In the Heisenberg picture,

 1
= [ ] = − 2 ;
 ~

 1 
= [ ] = 
 ~ 

2  1  1
2
= [  ] = [−2  ] = − 2 ;
 ~  ~

2  1  1 
2
= [  ] = [  ] = −2 ;
 ~  ~ 

Equivalently,


= − =⇒ () = (0)−


†
= † =⇒ † () = † (0)

86
CHAPTER 2 – MANUSCRIPT

 (0) −
() + () = ((0) +  )
 

 (0) 
() − () = ((0) −  )
 
1
() = (0) cos  + (0) sin 


() = −(0) sin  + (0) cos 

For any energy eigenket, we have





⎨ h |()| i = 0


⎩ h |()| i = 0

For a superposition of energy eigenstates such as

|i = 0 |0i + 1 |1i

we have
∙ ¸12
~ ∗  ∗ −
h |()| i = ( 1  + 0 1  )
2 0

= 212 0 |0 1 | cos( + )

It is noted that the average value of the position () in an energy eigenstate is always

zero.

B 2.3.3 The Coherent State

We have seen that an energy eigenstates does not behave like the classical oscillator —

in the sense of oscillating expectation values for  and  — no matter how large n may

87
CHAPTER 2 – MANUSCRIPT

be.

How can we construct a superposition of energy eigenstates that most closely

imitates the classical oscillator? In wave-function language, we want a wave packet that

bounces back and forth without spreading in shape. It turns out that a coherent state

does the desired job.

A coherent state:

 |i =  |i

Properties of |i

P∞
1. The coherent state |i = =0 () |i

| ()|2 is of the Poisson type about some mean value n̄

̄
|()|2 = ( exp(−̄))
!

2. It can be obtained by translating the oscillator ground state by some finite dis-

tance.

3. It satisfies the minimum uncertainty relation at all time

­ ®­ ®
(∆())2 (∆ ())2 = ~2 4

4.  = (̄)12

88
CHAPTER 2 – MANUSCRIPT

A 2.4 Schrodinger Wave Equation: Simple Harmonic

Oscillator

The Schrodinger equation (see p66 in Schiff’s book)


µ ¶
2  2 2
Ψ = +   Ψ = Ψ
2 2

To solve the problem we first introduce the dimensionless variable  = 0 


µ ¶
~2 2  2 2 2
− +  0  Ψ = Ψ
220  2 2
µ ¶
2 2 2 4 2 220
− 2+ 0  Ψ = Ψ
 ~2 ~2
µ ¶
2 2 2
− 2 + Ψ = Ψ = Ψ
 ~

We take
~
20 = 


As the potential approaches infinite when  → +∞ we require the wavefunction ap-

proach to zero. From the equation we have

2
Ψ →   − 2

We assume
2
Ψ → ()− 2

Then
00 0
 − 2 + ( − 1) = 0

89
CHAPTER 2 – MANUSCRIPT

The solutions are the Hermite polynomials,

00 0
 − 2 + 2 = 0

such that
1
 = 2 + 1;  = ( + )~  = 0 1 2 
2

Alternatively, we may write the Hamiltonian in the form

µ ¶
2 2
− 2 +  Ψ = Ψ

∙µ ¶µ ¶ µ ¶µ ¶¸
1    
+ − + − + Ψ = Ψ
2    

µ ¶
1 
 = √ + ;
2 
µ ¶
† 1 
 = √ − ;
2 

[ † ] = 1

¡ † ¢
2  + 1 Ψ = Ψ

The ground state

Ψ0 = 0

90
CHAPTER 2 – MANUSCRIPT

A 2.5 Propagators and Feynman Path Integrals

B 2.5.1 Propagators in Wave Mechanics.

Time-evolution problem with a time-independent Hamiltonian can be solved once we

expand the initial ket in terms of the eigenkets of an observable that commute with H.

| 0  i


= exp(− ( − 0 )) | 0 i
~
P  ­ ¯ ®
= exp(− ( − 0 )) |0 i 0 ¯ ̇0
{0 } ~

P 
= exp(− 0 ( − 0 )) |0 i h0 | 0 i
{0 } ~

Multiplying both sides by h| on the left, we have

h| 0 : i

P 
= exp(− 0 ( − 0 )) h|0 i h0 | i
0 ~
Z
P 
= 0 exp(− 0 ( − 0 )) h|0 i h0 |0 i h0 | 0 i
0 ~
Z
= 0 (  0  00 ) h0 | 0 i

where
P 
(  0  00 ) = h|0 i h0 |0 i − ~ 0 (−0 )
0

91
CHAPTER 2 – MANUSCRIPT

Denote by  ( ) the wave-function

 ( ) ≡ h | 0 : i

we have
Z
 ( ) ≡ 0 (  0  00 ) (0  )

We call the kernel of the integral operator  as the propagator in the wave mechanics.

In any given problem the propagator depends only on the potential and is independent

of the initial and final wave-function. It can be constructed once the energy eigenvalues

and eigen functions are given. Clearly the time evolution of the wave function is com-

pletely predicted if  is known and  (0  ) is given initially. The only peculiar feature,

if any, is that when a measurement intervenes, the wave function changes abruptly in

an uncontrollable way into one of the eigen functions of observable being measured.

Two Properties of :

1. For   0 ,  satisfies the Schrodinger’s time-dependent wave equation in the

variables 0 an  with  and 0 fixed.


~ (  0  00 ) = (  0  )


G Proof.
P 
(  0  00 ) = h|0 i h0 |0 i − ~ 0 (−0 )
0

92
CHAPTER 2 – MANUSCRIPT

 P 
~ (  0  00 ) = h|0 i h0 |0 i 0 − ~ 0 (−0 )
 0

P 
= h| 0 − ~ 0 (−0 ) |0 i h0 |0 i
0

P 
= h| − ~ (−0 ) |0 i h0 |0 i
0


= h| − ~ (−0 ) |0 i

= ()(  0  )

1.

lim (  0  00 ) = ( − 0 )


→0

G Proof.

P 0
lim (  0  00 ) = lim h|0 i h0 |0 i − ~
(−0 )
→0 →0 0

P
= lim h|0 i h0 |0 i = h|0 i = ( − 0 )
→0 0

The propagator K is simply the Green’s function for the time-dependent wave equation

satisfying

~2 2 
(− ∇ + () − ~ )(  0  00 ) = −~( − 0 )( − 0 ) (2.1)
2 

93
CHAPTER 2 – MANUSCRIPT

With the boundary condition

(  0  00 ) = 0 for   0

The -function ( − 0 ) is needed on the right hand side of the Eq.(2.1) because K

varies discontinuously at  = 0 

EXP Example 3

A free particle in 1D

The Hamiltonian is expressed as

P2
H=
2

The momentum P̂ commutes with H. Hence | i is a simultaneous

eigenket of the operators P̂ and Ĥ.

P |i =  |i

2
H |i = |i
2

94
CHAPTER 2 – MANUSCRIPT

Thus

(  0  00 )
∙ ¸
P 0 2
= h|i h| i exp − ( − 0 )
 2~
Z ∙ ¸
  0 2
= exp  −  − ( − 0 )
2~ ~ ~ 2~

Z+∞ ∙ ¸
0  − 0 ( − 0 ) 2 ( − 0 )2
= exp − ( − ) +
2~ 2~  − 0 2~( − 0 )
−∞

Z+∞ ∙ ¸
  − 0 ( − 0 ) 2 ( − 0 )2
= lim exp − || −  ( − ) +
→0+ 2~ 2~  − 0 2~( − 0 )
−∞

∙ ¸12
 ( − 0 )2
= exp( )
2~( − 0 ) 2~( − 0 )

Certain space and time integral derivable from  are of considerable.

Without loss of generality, we set 0 = 0 in the following

95
CHAPTER 2 – MANUSCRIPT

(1)  = 0

Z
() = (   0)

Z
P  
=  h|i h|i − ~


Z
P  
=  h|i h|i − ~

P  
= h|i − ~

= Tr(−~ )

(2) Let us consider the Laplace-Fourier transform of G(t).

Z∞
̃() ≡ − ()~−~ ( → 0+ )
0

Z∞ X
= −  −(− +)~
0 

X 1
= ( → 0+ )

 −  + 

The complete energy spectrum is exhibited as simply poles of ̃()

in the complex plane.

EXP Example 4

96
CHAPTER 2 – MANUSCRIPT

The propergator for the simple harmonic oscillator

r
00 0 
(  ;   0 ) =
2~ sin[( − 0 )]
∙ ¸
 {(002 + 02 ) cos ( − 0 ) − 200 0 }
× exp
2~ sin[( − 0 )]

B 2.5.2 Propagator as a Transition Amplitude

To gain further insight into the physical meaning of the propagator, we rewrite it in an

alternative form, which is related to the concept of transition amplitude.

P
(  0  0 ) = h|i h|0 i − (−0 )~

X D ¯¯   ¯¯ E D ¯¯  0 ¯¯ E
=  ¯− ~ ¯   ¯ ~ ¯ 0

X D ¯¯  ¯¯ E D ¯¯ 0 ¯¯ E
=  ¯− ~ ¯   ¯ ~ ¯ 0

= h |0  0 i

Where | i and |0  0 i are to be understood as an eigenket and of the position operators

in the Heisenberg picture. Roughly speaking, h |0  0 i is the amplitude for the particle

to go from a space-time point (0 0 ) to another one ( )

97
CHAPTER 2 – MANUSCRIPT

Figure 2.3: Paths in x-t plane.

B 2.5.3 Path Integral as the Sum Over Paths

Without loss of generality we restrict ourselves to one-dimensional problem. The entire

time interval between  = 1 and  =  is divided into  − 1 equal parts.

 − 1
 − −1 = ∆ = ( = 1 )
 −1

Exploiting the composition property we obtain

h   |1  1 i

Z Z Z
=  −1 −2  2

× h   | −1  −1 i h−1  −1 |−2  −2 i

 h2  2 |1  1 i

98
CHAPTER 2 – MANUSCRIPT

Before proceeding further, it is profitable to review here how paths appear in classical

mechanics. Suppose we have a particle subjected to a force field, say gravitional field.

 () ⇔ 

(1  1 ) ⇔ ( 0)
µ ¶12
2
(   ) ⇔ (0 )

The classical Lagrangian is written as

1
 ( ̇) = ̇2 −  ()
2

According to Hamiltonian’s principle, the unique path is determined by minimizing the

action
Z
  ( ̇) = 0
1

Lagraingian equation:

  
− =0
   ̇

B 2.5.4 Feynman’s Formalism

The basic difference between classical mechanics and quantum mechanics should be

apparent: In classical mechanics, a definite path in  −  plane is associated with the

particle’s motion; in contrast, in quantum mechanics all possible paths must play roles

99
CHAPTER 2 – MANUSCRIPT

including those which do not bear any resemblance to the classical path. Yet the

difference should disappear when ~ → 0. To formulate Feynman’s approach, let us go

back to h   |−1  −1 i with ∆ =  − −1 (∆ → 0). We write

1
h   |−1  −1 i = exp [(  − 1)~]
(∆)

where
Z
(  − 1) =  ( ̇)
−1

and (∆) is determined by

h   |−1  −1 i | =−1 = ( − −1 )

Because at any given time the position kets in the Heisenberg picture form a complete

set, it is legitimate to insert the identity operator written as


Z
 | i h | = 1

at any time we desire. For example,


Z
­    ® ­ ®­ ®
   |   =     |   |  

Z
­ ® ­ ®
= 0    |0  0 h0  0 | i  |  

= 

and so on. If we somehow guess the form of h2  2 |1  1 i for a finite time interval by

compounding the appropriate transition amplitude for infinitesimal time interval. This

100
CHAPTER 2 – MANUSCRIPT

property leads to formulate a space-time approach to quantum mechanics. Our task is

to evaluate the ∆ → 0 limit of (  − 1). As ∆ → 0

Z
1
(  − 1) = ( ̇2 −  ())
2
−1

1  − −1 2  − −1
= ∆( ( ) −( ))
2 ∆ 2

For a free-particle case,  = 0,

h   |−1  −1 i | =−1 = ( − −1 )

we obtain
1 ³  ´12
=
(∆) 2~∆

where we have used


Z∞ ³
 ´12   2
 2~ ∆  = 1
2~∆
−∞

and
³  ´12   2
lim  2~ ∆ = ()
∆→0 2~∆

To summarize, as ∆ → 0 we obtain

³  ´12   (−1)
h   |−1  −1 i =  ~
2~∆

101
CHAPTER 2 – MANUSCRIPT

The final expression for the transition amplitude with  − 1 is

Z Y

  −1 (−1)
h   |1  1 i = lim ( ) 2 −1 −2 2  ~
→+∞ 2~∆
=2

Z

h   |1  1 i ≡ D((1 )) ∫1  (̇)~

1

This is Feynman’s path integral!

Motivated by Dirac, Feynman attempted a new formulation of quantum me-

chanics based on the concept of paths. The ideas we borrowed from the conventional

form of the quantum mechanics are (1) the superposition principle (used in summing

the contribution from various alternate paths), (2) the composition property of the

transition amplitude, and (3) classical correspondence in the ~ → 0 limit.

A 2.6 The Gauge Transformation and Phase of Wave

Function

B 2.6.1 Constant Potential

In classical mechanics, it is well known that the zero point of potential energy is of no

physical significance. The force that appears in Newton’s second law depends only on

102
CHAPTER 2 – MANUSCRIPT

the gradient of the potential; an additive constant is clearly irrelevant. e.g.

  −
 = :

− 5  − 5 

when  →  + 0
 
 = → =
 

Let us look at the time evolution of a Schrodinger picture state ket subject to some

potential

 () =⇒  () + 0

2 2
 = +  () =⇒  = +  () + 0
2 2

^
| 0 : i =⇒ | 0 : i =?

To be precise we consider a time-independent 


| 0 : i = − ~ (−0 ) | 0 i

Thus
0 (−0 )
^
| 0 : i = − ~ | 0 : i

For a stationary state, if the time-dependence computed with  () is

−(−0 )~

103
CHAPTER 2 – MANUSCRIPT

Figure 2.4: Quantum-mechanical interference to detect a potential difference.

then the corresponding time dependence computed with  () + 0 is

−(+0 )(−0 )~

Even though the choice of the absolute scale of the potential is arbitrary,

potential difference are of non-trivial physical significance and in fact, can be detected

in a very striking way. A beam of charged particles is split into two parts, each of

which enters a metallic cage. A particle in the beam can be visualized as a wave packet

whose dimension is much smaller than the dimension of the cage. Inside the cage the

potential is spatially uniform. Hence the particle in the cage experience no force. If we

desire, a finite potential difference between the two cages can be maintained by turning

on a switch. The final state which the two beams reach at interference region is

1 1
| i = | 0 : i + | 0 : i
212 212

Thus

h | i = 1 + cos [( −  )~]

104
CHAPTER 2 – MANUSCRIPT

This is an observable effect! This is a purely quantum mechanical effect and has no

classical analogue.

B 2.6.2 Gauge Transformation in Electromagnetism

We first recall the Maxwell’s equations






⎪ ∇ · D = 4;






⎨ ∇ × H = 4 J + 1 D
;
  



⎪ ∇ · B = 0;






⎩ ∇ × E = − 1 B
 

D = E + 4P

H = B − 4M

We introduce a scalar potential  and a vector potential A





⎨ B=5×A


⎩ E=−5− 1 
A
 

Then the Maxwell’s equations are reduced to





⎨ 52  − 12  22  = −4
 

⎪ 1 2
⎩ 52 A − A = − 4 
2 2 

with
1 
5·A+ =0
 
105
CHAPTER 2 – MANUSCRIPT

Under the gauge transformation.





⎨ A → A + 5Λ


⎩ →− 1 Λ
 

with
1 2
52 Λ − Λ=0
2 2

the Maxwell’s equations keep unchanged! In quantum mechanics, the Hamiltonian for

a particle of electric charge  (   0 for electron in Sakurai’s book), subjected to the

electromagnetic field is taken from classical physics to be

1 
 = (p − A)2 + 
2 

1   2
= (p2 − p · A − A · p + 2 A2 ) + 
2   

1 2 2
6= (p2 − p · A + 2 A2 ) + 
2  

because p and A do not commute, [p A] 6= 0. It is straightforward to examine that the

Hamiltonian in this form is Hermitian,  =  †  To study the dynamics of a charged

particle subjected to  and A, let us first proceed in the Heisenberg picture:

 1 1 
 = [  ] = ( −  )
 ~  

where  =    Define
x 
≡ =p− A
 

106
CHAPTER 2 – MANUSCRIPT

We call that p is canonical momentum and  is kinetic (or mechanical) momentum.

The commutation relation of   and   is

~  
[     ] = − (  −  )
  

~
= −  (5 × A)

Hence the Hamiltonian can be rewritten as

2
= + 
2

We now study the Schrodinger’s equation with  and  in the position space.

D ¯  ¯ E
¯ 2¯
 ¯(p − A()) ¯  0 : 

Z Z D ¯ ¯ ED ¯ ¯ E
¯  ¯ ¯  ¯
= 1 2  ¯(p − A())¯ 1 1 ¯(p − A()) |2 i h2 ¯  0 : 
 


= (−~ 5 − A())2 h| 0 : i

Thus the Schrodinger’s equation is derived as

1  
[ (−~ 5 − ())2 + ] h| 0 : i = ~ h| 0 : i
2  

Denote

( ) ≡ h| 0 : i

and

 = ∗ ()()

107
CHAPTER 2 – MANUSCRIPT

The time derivative of the density is

  
 = ( ∗ ) + ∗  = − 5 ·j
  


+5·j=0


where
~ 
j= Im(∗ 5 ) − A ||2
 

For

 = 12 ~

 
j = (5 − A)
 

B 2.6.3 The Gauge Transformation

In quantum mechanics, we consider the gauge transformation





⎨ →


⎩  →  + 5Λ

e.g. Consider a uniform magnetic field in z direction

 
B =  ̂ = (  −  )̂
 

To derive the field, we can choose

(1).  = − 12   = 12 

108
CHAPTER 2 – MANUSCRIPT

(2).  = −  = 0

These two vector potentials are connected by a gauge transformation


A → A + 5( )
2

Let us denote by |i the state ket in the presence of A, and |̃i the state ket in the

presence of A + 5( 
2
) The Schrodinger’s equations for these two state kets are

1  
[ (p − A)2 + ] | 0 : i = ~ | 0 : i
2  

1   ^  ^
[ (p − A − 5 Λ)2 + ]| 0 : i = ~ | 0 : i
2   

^
What’s the relation of | 0 : i and | 0 : i ?

We come to construct an operator G such that

|̃i = G |i

with GG † =1

1   
G †[ (p − A − 5 Λ)2 + ]G | 0 : i = ~ | 0 : i
2   

We require

(1). h|x|i = h̃|x|̃i :

G † G =

109
CHAPTER 2 – MANUSCRIPT
­ ® D E
(2). |p −  A| = ̃|p −  Ã|̃ :

³   ´ ³  ´
G† p − A − 5 Λ G = p − A
  


G † [p G] = + 5 Λ


G † (−~∇G) = + 5 Λ


−~∇ ln G = + 5 Λ

h  i
G = exp  Λ
~

Therefore, if we take

|̃i =  ~  Λ |i

then

 1   
− ~  Λ ( (p − A − 5 Λ)2 + ) ~  Λ
2  

1 
= (p − A)2 + 
2 

The two wave-function are related via


̃( ) =  ~  Λ ( )

110
CHAPTER 2 – MANUSCRIPT

Figure 2.5:

In the form  = 12 exp[~]we have

A → A + ∇Λ

 → 


 → + Λ

B 2.6.4 The Aharonov-Bohm Effect

Consider a particle of charge e going above or below a very long impenetrable cylinder,

as shown in Fig.

Inside the cylinder is a magnetic field parallel to the cylinder axis, taken to be

normal to the plane of Fig. So the particle paths above and below enclose a magnetic

flux.

Assume the magnetic field





⎨  ̂ if   
B=


⎩ 0 if   

111
CHAPTER 2 – MANUSCRIPT

Hence the particle does not experience the Lorentz force outside cylinder.



⎨ (0 1  0), if   
2
A=


⎩ (0 1 2  0), if   
2

Our object is to study how the probability of finding the particle in the interference

region B depends on the magnetic flux. For pedagogical reason we prefer to use the

Feynman path-integral method to attack this problem.

In the presence of the magnetic field, the Lagrangian can be obtained

1 1 2 
(0) 2
 =  ̇ → ̇ + ̇ · A
2 2 

where ̇ = x The corresponding change in the action is given by

 R
 (0) (  − 1) →  (0) (  − 1) + ̇ · A
 −1

 R 
=  (0) (  − 1) +  · A
 −1

where  is the differential line element along the path segment. So when we consider

the entire contribution from 1 to  We have

Q
exp( (0) (  − 1)~)

Q  R
=⇒ exp( (0) (  − 1)~) exp( A · )
 ~ −1

The entire transition amplitude is


R
 R

  
A·
R (0)
  ~(1)
 ~
A· R (0)
  ~(1)
 ~
[() ]  × 1
+ [() ]  × 1

 

112
CHAPTER 2 – MANUSCRIPT

The probability for finding the particle in the interference region B depends on the

modulus square of the entire transition amplitude and hence on the phase difference

between the contribution from the paths going above and below. i.e.

 ∝ 1 + cos(1 − 2 )

where
( )
 R R
1 − 2 = A ·  − A · 
~ 1 1

I

= A · 
~
ZZ

= ∇ × A · 
~


= 
~ 

where  stands for the magnetic flux inside the impenetrable cylinder. This means

that as we change the magnetic field strength, there is a sinusoidal component in the

probability for observing the particle in region B with a period given by a fundamental

unit of magnetic flux, namely

2~
= 4135 × 10−7 Gauss-cm2
||

We emphasize that the interference effect discussed here is a purely quantum mechani-

cal. Classically the motion of a charged particle is determined by the Newton’s second

law supplemented by the force law of Lorentz. In this problem the magnetic field is

113
CHAPTER 2 – MANUSCRIPT

confined in the cylinder, and the particle experiences no force. However the vector

potential exists in the outside of cylinder. It seems that the vector potential A is more

fundamental than the magnetic field B. It is to be noted that the observable effect

depends on the magnetic flux not A directly. This effect has been observed experimen-

tally.

B 2.6.5 Magnetic Monopole

Suppose there is a point magnetic monopole situated at the origin,


B= r̂
2

which satisfies the equation

∇ · B =4 (r)

The solution for the vector potential,

B = ∇×A
∙ ¸
1  
= r̂ ( sin ) −
 sin   
∙ ¸
1 1  
+θ̂ − ( )
 sin   
∙ ¸
1  
+φ̂ ( ) −
  

If there is no singularity in the vector potential, we would have

∇ · (∇ × A) = 0

114
CHAPTER 2 – MANUSCRIPT

This violates the assumption,


Z
B · S =4

The solution I
∙ ¸
 (1 − cos )
A= φ̂
 sin 

The solution is singular at  = . We construct a pair of potentials


∙ ¸
()  (1 − cos )
A = φ̂
 sin 

for    − , and
∙ ¸
()  (1 + cos )
A =− φ̂
 sin 

for    We can get a solution for B, but A is doubly valued. The difference of the

vector potential is
2
A() − A() = − φ̂
 sin 

In the form of the gradient,

2
∇Λ = ∇(−2 ) = − φ̂
 sin 

The the wave functions in two different vector potential will differ by
µ ¶
() −2 
Ψ = exp Ψ() 
~

The wave function must be single valued because once we choose particular gauge,

the expansion of the state ket in terms of the position eigenkets must be unique. Let

115
CHAPTER 2 – MANUSCRIPT

us examine the behavior of the wave function Ψ() , on the equator  = 2 with a

definite radius r, which is a constant. When we increase the azimuthal angle  along

the equator from 0 to 2, Ψ() as well as Ψ() must return to its original value because

the two functions must be single-valued. This is possible only if

2
= ±
~

~ ~
 = ±  ±2 ···
2 || 2 ||

The elementary magnetic charge must be quantized.

We should emphasize that quantum mechanics does not require the existence

of magnetic monopole.

A 2.7 Interpretation of Wave Function.

B 2.7.1 What’s Ψ ()?

Since
Z Z
| 0 : i =  |i h| 0 : i =  |i Ψ ( )

Ψ ( ) is regarded as an expansion coefficient of | 0 : i in terms of the position

eigenkets {|i}. The quantity

( ) = |Ψ ( )|2

116
CHAPTER 2 – MANUSCRIPT

is defined as the probability density. So ( ) is the probability that the particle

appears in a narrow range round  at time . This is the so-called the probabilistic

interpretation of wave function.

The Continuity Equation.


( )

∙ ¸ ∙ ¸
 ∗ ∗ 
= Ψ ( ) Ψ( ) + Ψ ( ) Ψ( )
 

1 1
= − ( ∗ Ψ∗ )Ψ + Ψ∗ Ψ
~ ~

Assume

~2 2
 = − ∇ +
2

 1 1
( ) =  ~(Ψ∗ (∇2 Ψ) − Ψ(∇2 Ψ∗ )) + ( −  ∗ )Ψ∗ Ψ
 2 ~

117
CHAPTER 2 – MANUSCRIPT

As

Ψ∗ ∇2 Ψ = ∇(Ψ∗ ∇Ψ) − (∇Ψ∗ )(∇Ψ)

Ψ∇2 Ψ∗ = ∇(Ψ∇Ψ∗ ) − (∇Ψ)(∇Ψ∗ )

 ~ 1
 =  ∇(Ψ∗ ∇Ψ − Ψ∇Ψ∗ ) + Im( ) · 
 2 ~

 1
+∇·j = Im( ) · 
 ~

~ ∗ ~
j = − (Ψ ∇Ψ − Ψ∇Ψ∗ ) = Im(Ψ∗ ∇Ψ)
2 

–—The probability flux.

When  is purely real, we have


+∇·j=0


This is the conservation law for probability.

Question: what happens if V is not real? When  is complex, which is of-

ten used for nuclear reaction for particles absorbed by nuclei, it is accounted for the

disappearance of particles.

Except for the probability density, the wave function contains more physics

than expected. Let us write it as

Ψ( ) = ( )12 exp[( )~]

118
CHAPTER 2 – MANUSCRIPT

with   0 and  real.


Ψ∗ ∇Ψ = 12 ∇12 + ∇
~


=⇒ j = ∇

The spatial variation of the phase of the wave function characterizes the probability

flux; the stronger the phase variation, the more intense the flux. The direction of j at

some point  is seen to be normal to the surface of a constant that goes through that

point. Since j is a flux, j can be rewritten as

j = (∇) ≡ v

This physical meaning is still unclear. It is not a velocity traditionally!

B 2.7.2 The Classical Limit

Substituting Ψ = 12 ~ into a Schrodinger equation leads to

~2 2 1 
− (∇2 12 + ∇12 · ∇) − 2 12 |∇|2 + ( )12 ∇2  + 12 
2 ~ ~ ~

12  
= ~( + 12 )
 ~ 

Taking ~ → 0 , we obtain

1 
|∇( )|2 +  () + ( ) = 0
2 

119
CHAPTER 2 – MANUSCRIPT

Figure 2.6: One-dimensional square well potential with (a) perfectly rigid walls (b)

finite potential.

This is the Hamilton-Jacobi equation in classical mechanics. In this sense, the Schrodinger

equation goes back to the classical mechanics in the limit ~ → 0 In the classical me-

chanics ( ) stands for Hamiltonian principal function. In a stationary state with

time dependent − ~  the Hamilton’s principle function S is separable

( ) =  () − 

120
CHAPTER 2 – MANUSCRIPT

A 2.8 Examples

B 2.8.1 One dimensional square well potential

C Perfect rigid wall





⎨ 0 || ≤ 
 () =


⎩ +∞ ||  

Since the potential is infinite at ||   the wave function must vanish at the points

 = ± The wave function for || ≤  is simply

~2 2 
− = 
2 2

Introduce a dimensionless quantity  = 

2 22
2  = − 2
 = −2 
 ~

where  = (22 ~2 )12  The general solution is

() =  sin  +  cos 

Application of the boundary conditions at  = ±, i.e.,  = 1, gives

 sin  +  cos  = 0

− sin  +  cos  = 0

There are two possible classes of solution

121
CHAPTER 2 – MANUSCRIPT

(1).  = 0 and cos  = 0

1
 = ( + ) (2.2)
2

(2 + 1)
 =  cos  =  cos (2.3)
2

~2 2 (2 + 1)2 2 ~2
 = = (2.4)
2 2 82

(2)  = 0 and sin = 0

 =  (2.5)


 =  sin  =  sin (2.6)

~2 2 2 2 ~2
 = = (2.7)
2 2 22

C Finite potential step





⎨ 0 || ≤ 
 () =


⎩ 0  ||  

The wave equation at ||  


~2 2 
− = 
2 2

The wave equation at ||  

~2 2 
− + 0  = 
2 2

~2 2 
− = ( − 0 )
2 2
122
CHAPTER 2 – MANUSCRIPT

The general solution for   0





⎨ −    
() =


⎩ +    

where  = [2(0 − )]12 ~

For   0



⎨ 1 sin  + 1 cos    
() =


⎩ 2 sin  + 2 cos    

The wave equation at ||   is the same as that in the rigid wall potential, and the

general solution is

() =  sin  +  cos 

where  = (2~2 )12  We now impose on the solutions (1) and (2) the requirements

that u and du/dx be continuous at x=±.

 sin  +  cos  = −

 sin  −  cos  = −

− sin  +  cos  = −

 sin  +  cos  = −

123
CHAPTER 2 – MANUSCRIPT

from which we obtain

2 sin  = ( − )−

2 cos  = −( − )−

2 cos  = ( + )−

2 sin  = ( + )−

Two classes of solutions

(1)  = 0 and  = 

 tan  = 

(2)  = 0 and  = −

 cot  = −

C Parity

The Schrodinger equation

~2 2 ()
− +  ()() = ()
2 2

If the potential is symmetric about  = 0,  () =  (−), we have

~2 2 (−)
− +  ()(−) = (−)
2 2

124
CHAPTER 2 – MANUSCRIPT

The () and (−) are solutions of the same wave equation with the same eigenvalues

E. Addition or subtraction these two equations gives () ± (−) the solution of the

Schrodinger equation. To simplify the notation, () = ±(−) Such wave function

are said to have even or odd parity. Note that

cos  = cos(−)

sin  = − sin(−)

B 2.8.2 A charged particle in a uniform magnetic field

Consider a charged particle in a uniform magnetic field. Assume the particle is confined

in a two-dimensional plane, and the magnetic field is perpendicular to the plane, say

along the z direction. The Hamiltonian is given by

1 
= ( − )2
2 

with

 =  ̂ = (   −   )̂

 

To derive the field, we can choose

(1).  = −  = 0

(2).  = − 12   = 12 

125
CHAPTER 2 – MANUSCRIPT

C Case (1):  = −  = 0

1 h  2  2
i
 = ( −  ) + ( −  )
2  
∙ ¸
1  2 2
= ( − ) + ( )
2 

Notice that [  ] = 0 i.e. p is a good quantum number. The general form of the

wavefunction is

Ψ = ()0 

The Schrodinger equation Ψ = Ψ is reduced to

∙ ¸
1  2 2
(0 − ) + ( ) () = ()
2 

Alternatively

∙ ¸
2 2  2 2
+ ( − 0 ) () = ()
2 22
0
with 0 = 
 Thus the problem is reduced to the simple harmonic oscillator.

C Case (2).  = − 12   = 12 

1 h   i
 = ( −  )2 + ( −  )2
2  
∙ ¸
1  2  2
= ( − ) + ( + )
2 2 2

126
CHAPTER 2 – MANUSCRIPT

In this case both p and p are not good quantum numbers. We cannot solve this

problem like in the case (1). Take


  =  − ;
2


  =  + 
2
 ~
[     ] = −2~ = −
2 

1 £ 2 ¤
 =   + 2
2

1
= [(  +   )( −   ) + (  −  )(  +   )]
4

2~
[  −      +  ] =

Take
r

 = (  −  )
2~
r
† 
 = (  +   )
2~

Thus [ † ] = 1

~ ¡ † ¢
 =   + †
2
µ ¶ µ ¶
~ † 1 † 1
=  + ⇔ ~   +
 2 2

Find the wave function for the energy eigenstates.

127
CN Chapter 3

CT Theory of Angular Momentum

This chapter is concerned with a systematic treatment of angular momentum and re-

lated topics.

A 3.1 Rotation and Angular Momentum

We consider the rotation in the space of a physical system in a state represented by

a ket |i or the wave function  (r) We describe a rotation by a linear operator ,

which is so defined that any vector r is rotated into the new vector  · r. The rotation

changes the ket |i into the new ket |0 i or change the wave function  (r) into the

wave function 0 (r0 ), which means

hr|i = hr|i

128
CHAPTER 3 – MANUSCRIPT
­ ¯ + ¯ ®
r ¯D ()D()¯  = hr|i

hr|0 i = hr|i

0 (r) =  (r)

B 3.1.1 Finite versus infinitesimal rotation

From elementary physics, we know that

(1). Rotations about the same axis commute.

      
 ( ) ( ) =  ( ) ( ) =  ( + ) =  ( )
6 3 3 6 3 6 2

(2). Rotation about different axes do not commute.

   
 ( ) ( ) 6=  ( ) ( )
2 2 2 2

Consider a vector V with three components    and   When we rotate

the vector, we obtain a new vector, V with    and  . These two vectors are

connected by a 3 × 3 matrix
⎛ ⎞ ⎛ ⎞⎛ ⎞
⎜  ⎟ ⎜    ⎟ ⎜  ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎜  ⎟=⎜  ⎟⎜  ⎟
⎜  ⎟ ⎜    ⎟⎜  ⎟
⎜ ⎟ ⎜ ⎟⎜ ⎟
⎝ ⎠ ⎝ ⎠⎝ ⎠
    

129
CHAPTER 3 – MANUSCRIPT

Figure 3.1: Example to illustrate the noncommunativity of finite rotation.

The requirement that the rotated vector  be real when the vector  are real means

that the elements of  are real. The length of the vector  and  do not change,

which means
⎛ ⎞
⎜  ⎟
⎜ ⎟
⎜ ⎟
(     ) ⎜
⎜ 
⎟ = 2 + 2 + 2

⎜ ⎟
⎝ ⎠


  =     =   

  =  = 1


where “T” stands for a transpose of a matrix:  =   To be definite we consider a

130
CHAPTER 3 – MANUSCRIPT

rotation about  axis by an angle 

 = 

 =  cos  −  sin  (= 0 cos( + ))

 =  sin  +  cos  (= 0 sin( + ))

where  = 0 cos  and  = 0 sin  In the matrix form, we obtain


⎛ ⎞
⎜ cos  − sin  0 ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜ sin  cos  0 ⎟


⎜ ⎟
⎝ ⎠
0 0 1

Similarly ⎛ ⎞
⎜ 1 0 0 ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜ 0 cos  − sin 


⎜ ⎟
⎝ ⎠
0 sin  cos 
⎛ ⎞
⎜ cos  0 sin  ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜ 0 1 0 ⎟

⎜ ⎟
⎝ ⎠
− sin  0 cos 

We are particularly interested in an infinitesimal form, from which a great deal can be

131
CHAPTER 3 – MANUSCRIPT
2
learned about the structure of . Take sin  ≈  cos  ≈ 1 − 2  we have
⎛ ⎞
⎜ 1 0 0 ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜ 0 1 − 2 2 − ⎟

⎜ ⎟
⎝ ⎠
0  1 − 2 2

= 1 + () + (2 ) + · · ·
⎛ ⎞
⎜ 1 − 2 2 0  ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜ 0 1 0 ⎟

⎜ ⎟
⎝ ⎠
− 0 1 − 2 2
⎛ ⎞
⎜ 1 − 2 2 − 0 ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜  1 − 2 2 0 ⎟

⎜ ⎟
⎝ ⎠
0 0 1
⎛ ⎞⎛ ⎞ ⎛ ⎞⎛
⎜ 1 0 0 ⎟ ⎜ 1 − 2 2 0 ⎟ ⎜ 1 − 2 2 0
  ⎟⎜ 1 0
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎜ 0 1 − 2 2 − ⎟⎜ 0 1 ⎟−⎜
0 0 1 0 ⎟ ⎜ 0 1 − 2 2
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎜ ⎟⎜ ⎟ ⎜ ⎟⎜
⎝ ⎠⎝ ⎠ ⎝ ⎠⎝
0  1 − 2 2 − 0 1 − 2 2 − 0 1 − 2 2 0  1
⎛ ⎞
¡ ¢
⎜ 0 −2  −  − 12 2 + 1 ⎟
⎜ ⎟
⎜ ¡ ¢ ⎟
⎜ 2 0  −  − 12 2 + 1 ⎟
⎜ ⎟
⎜ ⎟
⎝ ¡ 1 2 ¢ ¡1 2 ¢ ¡ 1 2 ¢2 ¡ 1 2 ¢ ¡1 2 ¢ ⎠
 −  −2 + 1  +  2 − 1 −2 + 1 + −2 + 1 2 − 1

132
CHAPTER 3 – MANUSCRIPT
⎛ ⎞
2
⎜ 0 − 0 ⎟
⎜ ⎟
⎜ ⎟
= ⎜ 2 0 0 ⎟


⎜ ⎟
⎝ ⎠
0 0 0
Elementary matrix manipulation leads to

 () () −  () () =  (2 ) − 1

If we just remain the quantities of first order in , any two rotations about different

axes commute because  (2 ) ≈ 1

In quantum mechanics

Because rotations affect physical systems, the state ket corresponding to a

rotated system is expected to look different from the state ket corresponding to the

original system. Given a rotation operator , characterized by a 3 × 3 matrix , we

associated with an operator D() in the appropriate ket space such that

|i = D() |i

where |i and |i stand for the kets of the rotated and unrotated systems, respectively.

To construct the rotation operator, it is again fruitful to examine first its properties

under an infinitesimal rotation

J·n
D(n) = 1 −  
~

we define  the - component of the angular momentum. A finite rotation can be

133
CHAPTER 3 – MANUSCRIPT

obtained by compounding successively infinitesimal rotation about the same axis

 
D () = lim [D ( )]
→+∞ 

  
= lim (1 −  · )
→+∞ ~ 
∙ ¸− ~ 
 
= lim (1 −  · )~(− )
→+∞ ~ 


= exp(− )
~
1
Definition: lim (1 + ) ≡ 
→+∞ 

In order to obtain the angular momentum commutation relations we need

more concepts. For every , there exists a rotation operator D() in the appropriate

ket space

 =⇒ D()

We postulate that D() has the same group properties as R.

1. Identity:

 · 1 =  =⇒ D() · D(0) = D()

2. Closure:

1 · 2 = 3 =⇒ D(1 ) · D(2 ) = D(3 )

3. Inverse:

−1 = 1 =⇒ D() · D(−1 ) = D(0) = 1

134
CHAPTER 3 – MANUSCRIPT

4.Associativity:

1 (2 3 ) = (1 2 )3

D(1 ) · D(2 3 ) = D(1 2 ) · D(3 )

As

 () () −  () () =  (2 ) − 1

we examine

D ()D () − D ()D () = D (2 ) − 1

  2 2   2 2
= (1 −  − )(1 −  − )
~ 2~2 ~ 2~2

  2 2   2 2
−(1 −  − )(1 −  − )
~ 2~2 ~ 2~2

= −[   ]2 ~2 = D (2 ) − 1

 2
D (2 ) = 1− 
~

=⇒ [   ] = ~

Similarly, we obtain

[   ] = ~ 

135
CHAPTER 3 – MANUSCRIPT

or

J × J = ~J

Comparing with a conventional vector v

v × v = 0

B 3.1.2 Orbital angular momentum

If the vector  is of infinitesimal length and only quantities of first order in  are

retained, the relation  =  may be written

 =  +  × 

where ⎛ ⎞
⎜ 1 −  ⎟
⎜ ⎟
⎜ ⎟
=⎜
⎜  1 − ⎟

⎜ ⎟
⎝ ⎠
−  1

136
CHAPTER 3 – MANUSCRIPT

To derive the formula we have used the infinitesimal rotation, for example,
⎛ ⎞
⎜ cos  − sin  0 ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜ sin  cos  0 ⎟

⎜ ⎟
⎝ ⎠
0 0 1
⎛ ⎞
⎜ 0 − 0 ⎟
⎜ ⎟
⎜ ⎟
= 1+⎜ 
⎜  0 0 ⎟

⎜ ⎟
⎝ ⎠
0 0 0

and the infinitesimal rotations about any axis are commutable.

Now we wish to find a transformation |i that changes the ket |i into the

ket |i

|i = D() |i

h |i = h |i

=⇒ h| D+ () |i = h |i

=⇒ h |i = h |i

­ ®
=⇒ h |i = −1  |

137
CHAPTER 3 – MANUSCRIPT

Thus we have

D() () =  (−1 )

≈  ( −  × )

≈  () − ( × ) · 5 ()


=⇒ D() = 1 −  ·  + · · ·
~

where

L=r×P

 
 =  −  (= −~( −  ))
 

 
 =  −  (= −~( −  ))
 

 
 =  −  (= −~( −  ))
 

These operators satisfy

[   ] = ~ 

B 3.1.3 Rotation operator for spin 1/2

We have already checked that

[   ] = ~  

138
CHAPTER 3 – MANUSCRIPT

Consider a rotation by a finite angle  about the z-axis. If the ket of a spin 1/2 system

before rotation is given by |i  the ket after rotation is given by

|i = D () |i

with

D () = − ~ 

To see the physical meaning of D (), we compute

D+ () D ()

  
~ 
=  ~   − ~ 
=  ~ (|+i h−| + |−i h+|) − ~ 
2

~ ³    
´
=  2 |+i h−|  2 + − 2 |−i h+| − 2
2

~
= [(|+i h−| + |−i h+|) cos  +  (|+i h−| − |−i h+|) sin ]
2

=  cos  −  sin 
⎛ ⎞ ⎛ ⎞
⎜  ⎟ ⎜  cos  −  sin  ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
D () ⎜  ⎟ D () = ⎜
† ⎜ ⎟
⎜  sin  +  cos 


⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
 
⎛ ⎞⎛ ⎞
⎜ cos  − sin  0 ⎟ ⎜  ⎟
⎜ ⎟⎜ ⎟
⎜ ⎟⎜ ⎟
⎜ ⎟ ⎜
= ⎜ sin  cos  0 ⎟ ⎜  ⎟ ⎟
⎜ ⎟⎜ ⎟
⎝ ⎠⎝ ⎠
0 0 1 

139
CHAPTER 3 – MANUSCRIPT

Compare with

 = 

B 3.1.4 Spin precession revisited

The Hamiltonian is


 = − S·B


= S

where  = ||  The time evolution operator is given by

() = exp [−~] = exp [−  2] 

For an arbitrary state |i, at a later time t

| i = () |i

= ()(|+i h+| + |−i h−|) |i

= −2 |+i h+|i + +2 |−i h−|i

The operators change as

hS i = hS i=0 cos  − hS i=0 sin 

hS i = hS i=0 cos  + hS i=0 sin 

hS i = hS i=0

140
CHAPTER 3 – MANUSCRIPT

Conclusion: the period for the state ket is twice as long as the period for spin precession:

 precession = 2;

 state ket = 4

A 3.2 Rotation Group and the Euler Angles

B 3.2.1 The Group Concept

The branch of mathematics that is appropriate for a full treatment of symmetry is the

theory of groups. Here we give a few basic definitions: A set of objects , , ,... form

a group if a process can be defined that enables us to combine any two of the objects,

such as  and , to form an object , and if the following conditions are satisfied:

1. All results of combination, such as , are members of the group.

2. The group contains an identity or unit member  that has the properties  =

 = , where  is any member of the group.

3. Each member  has an inverse −1 also in the group, such that −1 = −1  = .

4. Group combination is associative, so that

() = ()

141
CHAPTER 3 – MANUSCRIPT

The members of the group are called elements. Though we frequently refer to

the combination as “multiplication”, this does not mean ordinary multiplication. For

example, the set of integers, positive, negative, and zero, form a group if the law of

combination is ordinary addition.

A group is abelian if multiplication is commutative, so that  =  for all

pair of elements. Otherwise the group is non-abelian.

Two groups are said to be isomorphic to each other if there is a unique one-

to-one correspondence between elements of the two groups such that products of cor-

responding elements correspond to each other.

Examples:

(1). The elements are 1 and −1 and the law of combination is multiplication.

(2). {  2    = }:  is the root of  = 1

B 3.2.2 Orthogonal Group

We discussed rotations of a vector. The rotated vector and unrotated vector are con-

nected by a 3 × 3 real and orthogonal matrix

 = 

All rotation matrices form a group.

1. Combination of two 1 and 2 is a new matrix 1 2

142
CHAPTER 3 – MANUSCRIPT

2. Identity:  = 1

3. The inverse matrix:  = −1 and   =  = 

4. Associativity:(1 2 )3 = 1 (2 3 )

This group is named SO(3), where S stands for special, O stands for orthog-

onal, 3 for three dimension. If the vector is n-dimensional, the R form a SO(n) group.

B 3.2.3 “Special”?

Consider ⎛ ⎞
⎜ cos  − sin  0 ⎟
⎜ ⎟
⎜ ⎟
 () = ⎜
⎜ sin  cos  0 ⎟

⎜ ⎟
⎝ ⎠
0 0 1
Its determinant is

det( ()) = 1

In fact,

det() = 1 =⇒ ””

−1 is not one of elements in SO(N). We postulate that D() has the same group

properties of R. The determinant of D() is


µ ¶
J·n
det[D()] = det exp[− ]
~
∙ ¸
J·n
= exp det(− ) = 1
~

143
CHAPTER 3 – MANUSCRIPT

since

det[ ] = 0

B 3.2.4 Unitary Unimodular Group

Another rotation we discussed is for the spin 1/2 system, where the rotation matrices

are 2 × 2 We can write a unitary unimodular matrix as


⎛ ⎞
⎜   ⎟
( ) = ⎜



−∗ ∗

where a and b are complex satisfying ||2 + ||2 = 1 All these matrices form a group.

(1) Closure:

(1  1 )(2  2 ) = (1 2 − 1 ∗2  1 2 + ∗2 1 )

where

|1 2 − 1 2 |2 + |1 2 + ∗2 1 |2 = 1

(2) Identity: ⎛ ⎞
⎜ 1 0 ⎟
(1 0) = ⎜



0 1

(3) Inverse:

† ( )( ) = 1

(4) Associative: multiplications of matrices are associative.

144
CHAPTER 3 – MANUSCRIPT

If we set


Re() = cos
2


Re() = − sin
2


Im() = − sin
2


Im() = − sin
2

the ( ) can be rewritten as

·

( ) =⇒ − 2

As det[( )] = 1, this group is called SU(2) group. SU(2) and SO(3) have a two-to-one

correspondence: In SU(2). a rotation by 2 produces −1. a rotation by 4 produces 1.

In SO(3) , a rotation 2 produces 1. More generally, ( ) and (− −) correspond

to one matrix in R (in SO(3))

B 3.2.5 Euler Rotations

To describe a rotation, usually we need three parameters: two parameters for the

rotation axis and one parameter for the rotation angle. In classical mechanics, an

arbitrary rotation of a rigid body can be accomplished in three steps, known as Euler

rotations.

145
CHAPTER 3 – MANUSCRIPT

Figure 3.2: Euler rotations.

146
CHAPTER 3 – MANUSCRIPT

Three steps:

(a) : R () y → y0

(b) : R0 () z → z0

(c) : R0 () y’ → y00

In terms of 3×3 orthogonal matrices the product of the three operations can be written

as

(  ) ≡ 0 ()0 () ()

There appear both rotations about body axes and about the space-fixed axes. This

is rather inconvenient! It is desirable to express the body-axis rotations in terms of

space-fixed axis rotation. Fortunately, we have

0 () =  () ()−1 ()

0 () = 0 () ()−1


0 ()

Therefore,

(  ) =  () () ()

As an example we calculate D(  ) for a spin 1/2 system.

(  ) =  () () ()

3 2 3
= − 2

− 2

− 2

147
CHAPTER 3 – MANUSCRIPT

Recall that
 ·   
exp(− ) = cos −  ·  sin
2 2 2

To prove the identity we use the ( · )2 = 1


⎛ ⎞
 
⎜ cos 2
−  sin 2
0 ⎟
(  ) = ⎜



0 cos 2 +  sin 2
⎛ ⎞ ⎛ ⎞
   
⎜ cos 2 − sin 2 ⎟ ⎜ cos 2 −  sin 2 0 ⎟

·⎝ ⎟ · ⎜ ⎟
⎠ ⎝ ⎠
sin 2 cos 2 0 cos 2 +  sin 2
⎛ ⎞
− + − −
⎜  cos 2 −
2 sin 2 ⎟
2

= ⎝ ⎟
− +

− 2 sin 2 − 2 cos 2

A 3.3 Eigenvalues and Eigenkets of Angular Momen-

tum

The commutation relation between three components of J are already derived:

[   ] = ~

[   ] = ~

[   ] = ~

148
CHAPTER 3 – MANUSCRIPT

These relation are often rewritten in a more compact form

 ×  = ~

In this section we work out the eigenvalues and eigenkets of angular momentum. To

this end, we introduce a new set of operators:

(1).  2 =   +   +  

(2). ± =  ±  , + = (− )†

 2 commutes with all three  ( =   )

[ 2   ] = 0

As    and  do not commute, we cannot diagonalize    and  simultaneously.

However we can choose one  to be diagonalized with  2 . By convention, we choose

 . We denote the eigenvalues of  2 and  by  and , respectively

 2 | i =  | i

 | i =  | i

To determine the values of  and , it is convenient to work with the ladder operator,

± .

[   ] = ~

[   ] = −~ ⇒ [  ± ] = ±~

149
CHAPTER 3 – MANUSCRIPT

[   ±  ] = ±~( ±  )

The commutation relations become

[  ± ] = ±~±

[+  − ] = 2~

[±   2 ] = 0

First of all, we have

­ 0 0 ¯ 2¯ ®
   ¯ ¯   =  0 0 ;

h0  0 | |  i =  0  0

Then

 ± | i = {[  ± ] + ±  } | i

= (±~± + ± ) | i

= ( ± ~)± | i

From this relation, we conclude

± | i = ± ( ) |  ± ~i 

150
CHAPTER 3 – MANUSCRIPT

Note that the bra of ± | i are

h | (± )† = h | ∓ = h  ± ~| ±∗ ( )

h | ∓ ± | i = |± ( )|2

 2 can be expressed in terms of ± and 

1
2 = (+ − + − + ) + 2
2

= − + + 2 +  ~ = + − + 2 −  ~

 = |+ ( )|2 + ( + ~) = |− ( )|2 + ( − ~)

|+ ( )|2 =  − ( + ~) ≥ 0

|− ( )|2 =  − ( − ~) ≥ 0

If | i is one of eigenkets, then

± | i = ± ( ) |  ± ~i

If + ( ) or − ( ) is not equal to zero, |  ± ~i is also one of eigenkets. Suppose

that we keep on applying ± to both side of the equation above. We can obtain

numerical eigenkets with smaller and smaller or larger and larger  ± ~ until the

sequence terminates at some max and min such that





⎨ + ( max ) = 0


⎩ − ( min ) = 0

151
CHAPTER 3 – MANUSCRIPT

So

|+ ( max )|2 =  − max (max + ~) = 0

|− ( min )|2 =  − min (min − ~) = 0

 = max (max + ~) = min (min − ~)

=⇒ max = −min

Clearly, we must be able to reach | max i by applying + successively to | min i a

finite number of times. i.e.

| max i ∝ (+ ) | min i ∝ | min + ~i

We obtain

max = min + ~

As a result,

max = ~
2

and
 
= ( + 1)~2
2 2

It is conventional to define


max = −min = ~ = ~
2

152
CHAPTER 3 – MANUSCRIPT

and

| i =⇒ | i

such that

 2 | i = ( + 1)~2 | i

 | i = ~ | i

with  = − − + 1 

Take ± ( ) real, we have

+ | i = [( + 1) − ( + 1)]12 ~ | + 1i ;

− | i = [( + 1) − ( − 1)]12 ~ | − 1i 

The eigenkets {| i} form a basis for angular momentum operator

h | 0  0 i =   0  0 

In the form of matrix, elements of ±

h 0 0 |± | i = ~ [( + 1) − ( ± 1)]12   0  0 ±1 

As a result,

153
CHAPTER 3 – MANUSCRIPT

¿ ¯ ¯ À
0 0
¯ ¯
0 0 ¯ + + − ¯
h  | | i =   ¯ ¯ 
2

~
= [( + 1) − ( + 1)]12   0  0 +1
2

~
+ [( + 1) − ( − 1)]12   0  0 −1
2

h 0 0 | | i

~
= [( + 1) − ( + 1)]12   0  0 +1
2

~
− [( + 1) − ( − 1)]12  0  0 −1
2

h0 | | i = ~ 0 ;

­ 0 ¯ 2¯ ®
 ¯ ¯  = ( + 1)~2  0

Our choice of a representation in which  2 and  are diagonal has led to discrete

sequences of values for the corresponding labels j and m. The infinite matrices thus

obtained are most conveniently handled by breaking them up into an infinite set of

finite matrices, each of which is characterized by a particular value of  and has 2 + 1

rows and columns.

154
CHAPTER 3 – MANUSCRIPT

(1)  = 12:  = ±12


⎛ ⎞ ⎛ ⎞
⎜ 0 1 ⎟ ⎜ 0 − ⎟
 = ⎜
~ ⎟  = ~ ⎜ ⎟

2 ⎠ 2 ⎝ ⎠
1 0  0
⎛ ⎞ ⎛ ⎞
⎜ 1 0 ⎟ 2 3 2⎜ 1 0 ⎟
 = ~ ⎜ ⎟  = ~ ⎜ ⎟
2 ⎝ ⎠ 4 ⎝ ⎠
0 −1 0 1

(2)  = 1  = ±1 0
⎛ ⎞ ⎛ ⎞
⎜ 0 1 0 ⎟ ⎜ 0 − 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
 = 212 ⎜ 1 0 1 ⎟
~ ⎜
⎟  =
~
212
⎜ − 0 − ⎟
⎜ ⎟
⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
0 1 0 0  0
⎛ ⎞ ⎛ ⎞
⎜ 1 0 0 ⎟ ⎜ 1 0 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
 = ~ ⎜
⎜ 0 0 0

⎟  = 2~ ⎜ 0 1 0 ⎟
2 2⎜

⎜ ⎟ ⎜ ⎟
⎝ ⎠ ⎝ ⎠
0 0 −1 0 0 1

(3)  = 32   = ± 32  ± 12
⎛ ⎞
12
⎜ 0 3 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 312 0 2 0 ⎟
~⎜ ⎟
 = ⎜ ⎟
2⎜
⎜ 0


⎜ 2 0 312 ⎟
⎜ ⎟
⎝ ⎠
0 0 312 0

155
CHAPTER 3 – MANUSCRIPT
⎛ ⎞
12
⎜ 0 −3 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ −312 0 −2 0 ⎟
~⎜ ⎟
 = ⎜ ⎟
2⎜



⎜ 0 2 0 −312 ⎟
⎜ ⎟
⎝ ⎠
0 0 312 0
⎛ ⎞
⎜ 3 0 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 0 −1 0 0 ⎟
~⎜ ⎟
 = ⎜



2⎜ ⎟
⎜ 0 0 −1 0 ⎟
⎜ ⎟
⎝ ⎠
0 0 0 −3
⎛ ⎞
⎜ 1 0 0 0 ⎟
⎜ ⎟
⎜ ⎟
⎜ 0 1 0 0 ⎟
15 2 ⎜ ⎟
 = ~ ⎜
2 ⎟
4 ⎜ ⎜ 0 0 1

⎜ 0 ⎟

⎜ ⎟
⎝ ⎠
0 0 0 1

B 3.3.1 Representation of Rotation Operator

Have obtained the matrix elements of  and ± , we are now in position to study the


matrix elements of rotation operator D()

­ ¯ ¯ ®
D0 () = 0 ¯−·̂~ ¯ 
()

Since [ 2   ] = 0 we have

[D()  2 ] = 0

156
CHAPTER 3 – MANUSCRIPT

So D() | i is still an eigenket of  2 with the same eigenvalue ( + 1)~2

 2 D() |i = ( + 1)~2 D() |i

We just consider the matrix elements of D() with the same . For every  there are

2 + 1 values of m. D() should be a (2 + 1) × (2 + 1) square matrix. For a rotated

ket
X X ()
() |i = |0 i h0 |()| i = 0 () |0 i
0 0

As is well known, the Euler angles may be used to characterize the most general rotation.

We have

D ¯  ¯ E
() ¯    ¯
0 (  ) = 0 ¯− ~  − ~  − ~  ¯ 

D ¯  ¯ E
0 ¯ ¯
= ( −) 0 ¯− ~  ¯ 

For  = 12 ⎛ ⎞
~⎜ 0 − ⎟
 = ⎜ ⎟
2⎝ ⎠
 0

⎛ ⎞
nD ¯ ¯ Eo ⎜ cos

− sin 

0 ¯ −  ¯ 2 2
 ¯ ~ ¯  = ⎜



sin 2 cos 2

 
= cos  0 −  sin  
2 2

157
CHAPTER 3 – MANUSCRIPT

To derive the matrix we can make use of the identity: ( })2 = 14

For  = 1, ⎛ ⎞
⎜ 0 − 0 ⎟
⎜ ⎟
~ ⎜ ⎟
 = 12 ⎜  0 − ⎟


2 ⎜ ⎟
⎝ ⎠
0  0
⎛ ⎞

⎜ cos2 2
1
− 212 sin  sin2 2 ⎟
D ¯ ¯ E ⎜ ⎟
0 ¯ −  ¯
⎜ ⎟
{  ¯ ~ ¯  } = ⎜ 1
⎜ 212 sin  cos  1
− 212 sin  ⎟

⎜ ⎟
⎝ ⎠
sin2 2 1
212
sin  cos2 2

To derive the matrix we can make use of the identity: ( })3 =  }

A 3.4 Schwinger Oscillator Model.

There exists a very interesting connection between the algebra of angular momentum

and the algebra of two independent simple harmonic oscillators. Let us consider two

types of oscillator: plus and minus type. The creation and annihilation operator are

denoted by ± and †± , respectively,

[+  †+ ] = 1 [−  †− ] = 1

These two oscillators are uncoupled,

[+  †− ] = [−  †+ ] = 0

158
CHAPTER 3 – MANUSCRIPT

Define the number operators

+ = †+ + ; − = †− −

We have

[+  + ] = −+ ; [−  − ] = −−

[+  †+ ] = + ; [−  †− ] = †−

Since [+  − ] = 0 , we define the simultaneous eigenket of + and − by |+  − i

such that
à !à !
(†+ )+ (†− )−
|+  − i = |0 0i
(+ !)12 (− !)12

with

± |0 0i = 0

†+ |+  − i = (+ + 1)12 |+ + 1 − i

− |+  − i = (− )12 |+  − − 1i

We define

+ ≡ ~†+ − ; − ≡ ~†− +

~ † 1 ~
 ≡ (+ + − †− − ) = [+  − ] = (+ − − )
2 2~ 2

159
CHAPTER 3 – MANUSCRIPT

We can prove that these operators satisfy the angular momentum relations

[  ± ] = ±~±

Define the total number operator  by

 = + + − 

we can prove

1
 2 = 2 + (+ − + − + )
2

~2 ~2
= (+ − − )2 + (†+ − †− + + †− + †+ − )
4 2

~2 2
= [ + −2 − 2+ − + 2+ (1 + − ) + 2− (1 + + )]
4 +

 
= ( + 1)~2 
2 2

Therefore, we have

+ + − + + −
 2 |+  − i = ~2 ( + 1) |+  − i
2 2

+ |+  − i = ~†+ − |+  − i

= ~ [(+ + 1)− ]12 |+ + 1 − − 1i

− |+  − i = ~ [+ (− + 1)]12 |+ − 1 − + 1i

~
 |+  − i = (+ − − ) |+  − i
2
160
CHAPTER 3 – MANUSCRIPT

These expression can be reduced to the familiar forms for angular momentum provided

that

+ =⇒  + 

− =⇒  − 

(†+ )+ (†− )−


|+  − i =⇒ | i = |0 0i
[( + )!]12 [( − )!]12

So we can check

 2 | i = ( + 1)~2 | i

+ | i = [( + 1) − ( − 1)]12 ~ |  + 1i

− | i = [( + 1) − ( − 1)]12 ~ |  − 1i

 | i = ~ | i

B 3.4.1 Spin 1/2 system

We define the spin 1/2 operators ± and  in terms of the creation and annihilation

operators † and  ( =↑ ↓). For  = 12

↑ + ↓ = 2 = 1

161
CHAPTER 3 – MANUSCRIPT

We have two eigenkets for 

|+i = †↑ |0i

|−i = †↓ |0i

The operators are expressed as

+ = ~†↑ ↓

− = ~†↓ ↑

~ +
 = ( ↑ − +
↓ ↓ )
2 ↑

Since [↑  †↑ ] = [↓  †↓ ] = 1, we usually call  and † the boson operators. In the case

of  = 12 2 = († )2 = 0 So they are called hard-core bosons since no two bosons

can be occupied at the same site.

Please check whether the commutation relations still hold if


n o

↑  ↑ = ↑ †↑ + †↑ ↑ = 1

n o
↓  †↓ = 1

n o
↓  †↑ = 0

i.e.,  are fermionic operators.

Quantum statistics can be introduced by making use of commutation and

anticommutation relation of these operators.

162
CHAPTER 3 – MANUSCRIPT

B 3.4.2 Two-spin—1/2 system

We now consider two-spin-1/2 system: 1 and 2  For each  ( = 1 2) we have

|+i = †↑ |0i

|−i = †↓ |0i

There are four possible configurations for combination of two spins

|1+ 2+i = |1+i ⊗ |2+i = †1↑ †2↑ |0i

|1+ 2−i = |1+i ⊗ |2−i = †1↑ †2↓ |0i

|1− 2+i = |1−i ⊗ |2+i = †1↓ †2↑ |0i

|1− 2−i = |1−i ⊗ |2−i = †1↓ †2↓ |0i

We are now in a position to construct eigenstates of the total spin

S = S1 + S2

and its − component

S = S1 + S2 

We first examine the two kets: |1+ 2+i and |1− 2−i.

S |1+ 2+i = (S1 + S2 ) |1+ 2+i = ~ |1+ 2+i

S2 |1+ 2+i = 2~2 |1+ 2+i

163
CHAPTER 3 – MANUSCRIPT

We denote

| = 1  = 1i = |1+ 2+i = †1↑ †2↑ |0i

Similarly,

| = 1  = −1i = |1− 2−i = †1↓ †2↓ |0i

Recall that

− | i = [( + 1) − ( − 1)]12 ~ |  − 1i

1 1
| = 1  = 0i = S− | = 1  = 1i = (†1↑ †2↓ + †1↓ †2↑ ) |0i
212 ~ 212

We have obtained three eigenkets for  2 and  . There should exist another eigenket

since there are four possible configurations for two spins. The fourth eigenket must be

a combination of |1+ 2−i and |1− 2+i

|i =  |1+ 2−i +  |1− 2+i

It is easy to check

 |i = 0

We make use of property of orthogonality of eigenkets

h = 1  = 0 | i = 0

From this equation, we obtain

1
 = − =⇒  = − =
212

164
CHAPTER 3 – MANUSCRIPT

Furthermore,

1
S2 |i = 0 =⇒ |i = | = 0  = 0i = (†1↑ †2↓ − †1↓ †2↑ ) |0i
212

1
| = 0  = 0i = (†1↑ †2↓ − †1↓ †2↑ ) |0i
212

B 3.4.3 Explicit Formula for Rotation Matrices.

Schwinger’s scheme can be used to derive in a very simple way, a closed formula for

rotation matrices. We apply the rotation operator D() to |i. In the Euler angle

rotation the only non-trivial rotation in the second one about y- axis. So

D() = D(  )==0 = − ~

Applying D() on | i, we have

D() |i

(†+ )+ (†− )−


= D() |0i
[( + )!]12 [( − )!]12

[()†+ ()
−1 +
] [()†− ()
−1 −
]
= D() |0i
[( + )!]12 [( − )!]12
1 1
 = (+ − − ) = (†+ − − †− + )
2 2

We have

 |0i = 0

165
CHAPTER 3 – MANUSCRIPT

and

D() |0i = |0i

We come to calculate

 
± () = D()†± D−1 () = − ~  †± + ~ 

 1  
 ~ 
+ () = − − ~  [  +
+ ]
 ~

1 −   †    1
=  ~ −  ~ = − ()
2 2

Similarly
 1
− () = − + ()
 2
2 1 2
2 ± () + ( ) ± () = 0
 2

The solutions are


 
± () = cos ̃1± + sin ̃2±
2 2

The boundary conditions are

± ()|=0 = †± ⇒ ̃1± = †±


¯
 ¯ 1
± ()¯¯ = ± ∓ ()|=0 =⇒ ̃2± = ±†∓
 =0 2

We have
 
± () = cos †± ± sin †∓
2 2

166
CHAPTER 3 – MANUSCRIPT
⎛ ⎞ ⎛ ⎞⎛ ⎞
⎜ †+⎟ −1 ⎜ cos sin ⎟ ⎜ 
2 ⎟

2
†+
D() ⎜

⎟ D () = ⎜
⎠ ⎝
⎟⎜
⎠⎝


†   †
− − sin 2 cos 2 −
Recalling the binomial theorem,

X

!
( + ) =   −
=0
!( − )!

[cos 2 †+ + sin 2 †− ]+ [cos 2 +  + −


− − sin 2 + ]
D() |i = × |0i
[( + )!]12 [( − )!]12
X ()
= 0  () |0 i
0

()
X 0 0 12
0 [( + )!( − )!( +  )!( −  )!]
0  = (−1)−+

( +  − )!!( −  − 0 )!( −  + 0 )

 0  0
×(cos )2−2+− (sin )2−+
2 2

()
Notice that 0  () = h0 | D() |i 

A 3.5 Combination of Angular Momentum and Clebsh-

Gordan Coefficients

One of the important problems in quantum theory is the combination of the angular

momenta associated with two parts of a system (such as the spin and orbital angular

momenta of one electron) to form the angular momentum of the whole system. The

167
CHAPTER 3 – MANUSCRIPT

addition of two vectors 1 and 2 forms a triangle. The third vector  = 1 + 2 . The

maximal value of  is 1 + 2 and the minimal value of  is |1 − 2 |. We start with two

commuting angular momentum operators J1 and J2 : all components of J1 commute

with all components of J2 . The orthonormal eigenkets of J21 and J1 are |1 1 i and

J2 has no effect on them. Similarly, |2 2 i are the eigenkets of J22 and J2  and J1 has

no effect on them. Denote

|1  2  1  2 i = |1  1 i ⊗ |2  2 i

we have

J21 |1  2  1  2 i = 1 (1 + 1)~2 |1  2  1  2 i

J1 |1  2  1  2 i = 1 ~ |1  2  1  2 i

J22 |1  2  1  2 i = 2 (2 + 1)~2 |1  2  1  2 i

J22 |1  2  1  2 i = 2 ~ |1  2  1  2 i 

The infinitesimal rotation operator that affects both subspaces is written as

J1 ·  J2 ·  J1 ⊗ 2 + J2 ⊗ 1
(1 −  ) ⊗ (2 −  ) = 1 ⊗ 2 −  · 
~ ~ ~

where 1 and 2 are two identity matrices, respectively. So we define the total angular

momentum

J = J1 ⊗ 2 + 1 ⊗ J2 ≡ J1 + J2

168
CHAPTER 3 – MANUSCRIPT

It is noted that

J × J = ~J

Since J2  J21  J22 and J are compatible, we denote their simultaneous eigenkets by

|1  2   i ≡ |i

Recall that there are (21 +1) eigenkets |1 1 i and (22 +1) eigenkets |2 2 i  Therefore

there are (21 + 1) × (22 + 1) kets

|1  2  1  2 i ≡ |1  1 i ⊗ |2  2 i (≡ |1  2 i)

These kets form a complete and orthogonal set of basis:

1 2
X X
|1  2 i h1  2 | = 1
1 =−1 2 =−2

The connection between | i and |1  2 i is

X
|i = |1  2 i h1  2 | i
1 2

{h1  2 | i} are the Clebsh-Gordan coefficients. Note that

(J − J1 − J2 ) | i = 0

( − 1 − 2 ) h1  2 | i = 0

It is apparent that h1  2 | i is zero unless  = 1 + 2. The largest value of m

is 1 + 2 and this occurs only when 1 = 1 and 2 = 2  Therefore the largest  is

169
CHAPTER 3 – MANUSCRIPT

1 + 2 , and that there is only such a state

|1 + 2 1 + 2 i = |1  2 i

| = 1 + 2  = 1 + 2 i = |1 = 1  2 = 2 i

The next largest value of m is 1 + 2 − 1 and this occurs twice: when (1) 1 = 1 and

2 = 2 − 1 and (2) 1 = 1 − 1 and 2 = 2  One of the two linear independent

combination of these two states ( = 1 + 2 − 1) must be associated with the new

state with  = 1 + 2 and  = 1 + 2 − 1 Since for each  value there must be values

of m ranging from 1 + 2 to −1 − 2 . The other combination is associated with  =

1 + 2 − 1 By an extension of this argument we can see that each j value, ranging from

1 + 2 to |1 − 2 | by integer steps, appears just once. Each  value is associated with

2 + 1 combinations of the original kets

1 +2
X
(2 + 1) = (21 + 1)(22 + 1)
=|1 −2 |

which is equal to the number of |1  2 i as expected.

B 3.5.1 Clebsch-Gordan coefficients

The Clebsch-Gordan coefficients form a unitary matrix. Furthermore the matrix ele-

ments are taken to be real by convention. An immediate consequence is h |1  2 i =

h1  2 | i  The orthogonality conditions of the matrix are

170
CHAPTER 3 – MANUSCRIPT

(1). h1  2 |01  02 i =  1 01  2 02

X
h1  2 | i h |01  02 i =  1 01  2 02


(2). h | 0  0 i =   0 0

P
1 P
2
h |1  2 i h1  2 | 0  0 i =   0 0
1 =−1 2 =−2

(3). h |1  2 i 6= 0 only if  = 1 + 2

Recursion Relations

We apply ± to the left side of Eq.


2 2
X X
| i = |01  02 i h01  02 | i
01 =−1 02 =−2

2 2
X X
+ |i = (1+ + 2+ ) |01  02 i h01  02 | i
01 =−1 02 =−2

[( + 1) − ( + 1)]12 |  + 1i

P 12
= [1 (1 + 1) − 01 (01 + 1)] |01 + 1 02 i h01  02 | i

P
+ [2 (2 + 1) − 2 (2 + 1)]12 |01  02 + 1i h01  02 | i

Multiplying h1  2 | from the left side

[( + 1) − ( + 1)]12 h1  2 |  + 1i

= [1 (1 + 1) − 1 (1 + 1)]12 h1 − 1 2 | i

+ [2 (2 + 1) − 2 (2 + 1)]12 h1  2 − 1| i

171
CHAPTER 3 – MANUSCRIPT

Similarly,

[( + 1) − ( − 1)]12 h1  2 |  − 1i

= [1 (1 + 1) − 1 (1 − 1)]12 h1 + 1 2 | i

+ [2 (2 + 1) − 2 (2 + 1)]12 h1  2 + 1| i

Construction Procedure

Combination of the two recursion relations and the orthogonality conditions

can leads to all Clebsch-Gordan coefficients . We start with the sector with the largest

value of m :

(1)  = 1 + 2   = 1 + 2

h1 = 1  2 = 2 | = 1 + 2   = 1 + 2 i = 1

(2)  = 1 + 2 − 1

(a) 1 = 1  2 = 2 − 1

(b) 1 = 1 − 1 2 = 2

 = 1 + 2

 = 1 + 2 − 1

| 1 + 2  1 + 2 i = |1  2 i

172
CHAPTER 3 – MANUSCRIPT

[(1 + 2 )(1 + 2 + 1) − ((1 + 2 )(1 + 2 − 1)]12 |1 + 2  1 + 2 − 1i

= [1 (1 + 1) − 1 (1 − 1)]12 |1 − 1 2 i + [2 (2 + 1) − 2 (2 − 1)]12 |1  2 − 1i

µ ¶12
1
h1 − 1 2 |1 + 2 , 1 + 2 − 1i =
1 + 2
µ ¶12
2
h1  2 − 1|1 + 2 , 1 + 2 − 1i =
1 + 2

h1  2 |1 + 2 − 1 1 + 2 − 1i

These two coefficients are determined by the orthogonality conditions h  − 1| − 1  − 1i =

h1 + 2  1 + 2 − 1|1  2 − 1i h1 2 − 1|1 + 2 − 1 1 + 2 − 1i

+ h1 + 2  1 + 2 − 1|1 − 1,2 i h1 − 1 2 |1 + 2 − 1 1 + 2 − 1i

= 0

µ ¶12
2
h1  2 − 1|1 + 2 − 1 1 + 2 − 1i
1 + 2
µ ¶12
1
+ h1 − 1,2 |1 + 2 − 1 1 + 2 − 1i = 0
1 + 2

µ ¶12
1
 =
1 + 2
µ ¶12
2
 =
1 + 2

173
CHAPTER 3 – MANUSCRIPT



⎨ h1 + 2 − 1 1 + 2 − 1|1 + 2 − 1 1 + 2 − 1i = 1


⎩ 2 +  2 = 1

⎪ ³ ´12

⎨ h1 2 − 1|1 + 2 − 1 1 + 2 − 1i = + 1
1 +2
⎪ ³ ´12

⎩ h1 − 1 2 |1 + 2 − 1 1 + 2 − 1i = − 2
1 +2

The convention here is that the first matrix element, which has the form h1   − 1 | i,

is real and positive.

(3)  = 1 + 2 − 2

(a) 1 = 1 2 = 2 − 2

(b) 1 = 1 − 1 2 = 2 − 1

(c) 1 = 1 − 2 2 = 2

 = 1 + 2

 = 1 + 2 − 1

 = 1 + 2 − 2

h1  2 |1 + 2   i and h1  2 |1 + 2 − 1  i are determined by the recursion re-

lations and the coefficients in (2). The coefficients h1  2 |1 + 2 − 2  i are deter-

mined by the orthogonality conditions.

174
CHAPTER 3 – MANUSCRIPT

In this construction procedure, the only difficult part is the use of orthogonally,

which becomes progressively more complicated as the rank of the submatrix increases.

However it needs be employed only once for each submatrix, and it is easier to work

out an example with particular numerical values for 1 and 2 than the general case

just considered.

(a) J− |1 + 2  1 + 2 − 1i
∙ ¸ 12
2 (22 − 1)
h1 2 − 2|1 + 2  1 + 2 − 2i =
(1 + 2 )(21 + 22 − 1)
∙ ¸ 12
41 2
h1 − 12 − 1|1 + 2  1 + 2 − 2i =
(1 + 2 )(21 + 22 − 1)
∙ ¸ 12
1 (21 − 1)
h1 − 22 |1 + 2  1 + 2 − 2i =
(1 + 2 )(21 + 22 − 1)

(b) J− |1 + 2 − 1 1 + 2 − 1i
∙ ¸ 12
1 (22 − 1)
h1 2 − 2|1 + 2 − 1 1 + 2 − 2i =
(1 + 2 )(1 + 2 − 1)

1 − 2
h1 − 12 − 1|1 + 2 − 1 1 + 2 − 2i = 1
[(1 + 2 )(1 + 2 − 1)] 2
∙ ¸ 12
2 (21 − 1)
h1 − 22 |1 + 2 − 1 1 + 2 − 2i = −
(1 + 2 )(1 + 2 − 1)

175
CHAPTER 3 – MANUSCRIPT

A 3.6 Spin Correlation Measurements and Bell’s In-

equality

Anybody who is not shocked by quantum theory has not understand it. – Niels Bohr

B 3.6.1 Spin singlet state



| = 0i = (|+; −i − |−; +i) 2

Suppose we make a measurement on one of spin component A then make

another measurement on B.

If spin 1 is z-up. spin 2 must be z-down, and vice vrese.


| = 0i = (|+; −i − |−; +i) 2

If spin 1 is z-up, we measure spin 2 along x direction. What is the result?

50% x-up and 50% x-down.

To sum up:

• If A measures S  and B measures S  there is a completely randaom correlation

between the two measurements.

• If A measures S  and B measures S  there is a 100% (opposite sign) correlation

between the two measurements.

176
CHAPTER 3 – MANUSCRIPT

• If A makes no measurement, B’s measurements shown random results.

The orthodox quantum mechanical interpretation:

B 3.6.2 Einstein’s Locality Principle and Bell’s inequality

Many physicists have felt uncomfortable with the preceding interpretation of spin-

correlation measure. Their feelings reflect in Einstein’s locality principle: "But on one

supposition we should in my oppinion, absolutely hold fast: The real factual situation

of the system S2 is independent of what is done with the system S1 , which is spatially

separated from the former.", which is also known as Einstein-Podolsky-Rosen paradox.

Consider a large number of spin-1/2 particles. We assign a certain fraction of

them to have the following property

• If S is measured, we obtain a plus sign with certainty.

• If S is measured, we obtain a minus sign with certainty.

A particle satisfying this property is said to belong to type (+ −). Noting

that we are not asserting that we can simultaneously measyre S and S to be + and -.

For a spin-singlet state, the results of the correlation measurement can be

177
CHAPTER 3 – MANUSCRIPT

reproduced if the particle 1 and 2 are matched as follows

Particle 1 Particle 2

(+ −) ↔ (− +)

(+ +) ↔ (− −)

(− +) ↔ (+ −)

(− −) ↔ (+ +)

with equal populations, that is 25% each. Suppose a particular pair belong to type

(+ −) and observer A decides to measure S of particle 1; then he or she neces-

sarily obtains a plus sign regardless of whether B decides to measure S or S . In

this sense that Einstein’s locality principle is incorportaed in this model: A’s results is

predetermined independently of B’s choice as to what to measure.

More complicated situations: we start with three unit vectors a, b, and c

178
CHAPTER 3 – MANUSCRIPT

Population Particle 1 Particle 2

1 (+ + +) (− −  −)

2 (+ + −) (− − +)

3 (+ − +) (− + −)

4 (+ − −) (− + +)

5 (− + +) (+ − +)

6 (− + −) (+ − +)

7 (− − +) (+ + −)

8 (− − −) (+ + +)

(1: a+, 2: b+): 3 + 4

(1: a+, 2: c+): 2 + 4

(1: c+, 2: b+): 3 + 7

 (+; +) ≤  (+; +) +  (+; +)

This is Bell’s inequality, which follows from Einstein’s locality principle.

B 3.6.3 Quantum Mechanics and Bell’s Inequality

In quantum mechanics, assume  the angle between a and b.

1 2 
 (+; +) = sin
2 2

179
CHAPTER 3 – MANUSCRIPT

where the factor 1/2 arises from the probability of initially obtaining S1 · a. From Bell’s

equality,
  
sin2 ≤ sin2 + sin2 
2 2 2

If we take  = 2 = 2 = 2, the inequality is then violated for

0    2

Nature has had the last laugh on EPR. Nearly thirty years after the EPR

paper was published, an experimental test was proposed that could be used to check

whether or not the picture of the world which EPR were hoping to force a return to

is valid or not. It turns out that Nature experimentally invalidates that point of view,

while agreeing with quantum mechanics.

A 3.7 PROBLEM
⎛ ⎞
⎜ 0 − ⎟
1. Find the eigenvalues and eigenvectors of   = ⎜

⎟. Suppose an electron

 0
⎛ ⎞
⎜  ⎟
is in the spin state ⎜ ⎟
⎝ ⎠  If  is measured, what is the probability of the result

~2?

2. Let n be a unit vector in a direction specified by the polar angles ( ). Show

180
CHAPTER 3 – MANUSCRIPT

that the component of the angular momentum in the direction n is

L = sin  cos L + sin  sin L + cos L

1 ¡ ¢
= sin  − L+ +  L− + cos L 
2

If the system is in simultaneously eigenstates of L2 and L belonging to the

eigenvalues ( + 1)~2 and ~ i.e., | i 

(a) what are the possible results of a measurement of L ?

(b) what are the expectation values of L and L2 ?

3. Obtain an explicit expression for

 (φ) = exp(−φn · S~)

in the form of a 2 × 2 matrix when S is the spin operator with  = 12 Let the

unit vector n have the polar angles  and  Show explicitly that your matrix for

 (φ) is unitary and that it is equal to −1 when  = 2

4. Prove a sequence of Euler rotations represented by

(12)(  )

³  ´ µ ¶ ³  ´

= exp −  3 exp −  2 exp −  3
2 2 2
⎛ ⎞
 
⎜ −(+)2 cos 2 −(−)2 sin 2 ⎟
= ⎜

⎟

 
−(−)2 sin 2 (+)2 cos 2

181
CHAPTER 3 – MANUSCRIPT

Because of the group properties of rotations, we expect that this sequence of

operations is equivalent to a single rotation about some axis by an angle . Find



5. Calculate the matrix of Clebsch-Gordan coefficients in the case of 1 = 12 and

2 = 12 According to the matrix, write explicitly the eigenstates of the total

angular momentum and its z component in terms of (1  1 ) and (2  2 )

6. Use the raising or lowering operator to find the following Clebsch-Gordan coeffi-

cients
¯ À ¯ À ¯ À
¯3 1 ¯1 1 ¯1 1
¯  ¯ ¯
¯ 2 2 =  |1 1i ⊗ ¯ 2  − 2 +  |1 0i ⊗ ¯ 2  2 

182
CN Chapter 4

CT Symmetries in Physics

Symmetry is a fundamental attribute of the natural world that enables an investigator

to study particular aspects of physical systems by themselves. For example, the as-

sumption that space is homogeneous, or possesses translational symmetry, leads to the

conclusion that the linear momentum of a closed isolated system does not change as the

system moves. This makes it possible to study separately the motion of the center of

mass and the internal motion of the system. A systematic treatment of the symmetry

properties and the conservation law are useful in solving more complicated problems.

It provides a deeper insight into the structure of physics.

In this chapter we consider the geometrical symmetries that may be associated

with the displacements of a physical system in space and time, with its rotation and

inversion in space, and with the reversal of the sense of progression of time.

183
CHAPTER 4 – MANUSCRIPT

A 4.1 Symmetries and Conservation Laws

B 4.1.1 Symmetry in Classical Physics

An alternative way to describe a classical system is to introduce Lagrangian L:

 L L
− =0
  ̇ 

which is called the Lagrange equation. For a conservative system,

L=T −V

where

V = V(1  2   )

For example, the Lagrangian for a harmonic oscillator is

1 1
L = ̇2 − 2 
2 2

From the Lagrange equation, we obtain

̈ = −

This is the Newtonian equation of motion! When the potential is related to velocity or

generalized velocity, the Lagrangian is constructed in a different way. For instance, the

Lagrangian for a charged particle subjected to an electromagnetic field


L = T −  +  · 

184
CHAPTER 4 – MANUSCRIPT

If the Lagrangian is unchanged under displacement in the general sense,

 →  + 

i.e.,
L
= 0


from the Lagrange equation, it follows that


=0


where that canonical momentum is defined as

L
 = 
 ̇

In other words, the canonical momentum is conserved, i.e. the law of conservation.

B 4.1.2 Symmetry in Quantum Mechanics

In quantum mechanics, we have learned to associate a unitary operator, say U, with an

operation like translation and rotation. It has become customary to call U a symmetry

operator regardless of whether the physical system itself possesses the symmetry to U.

When we say that a system possesses that symmetry to U, which implies that

 †  = 

185
CHAPTER 4 – MANUSCRIPT

For a continuous symmetry, we take an infinitesmal displacement  and expand the

operator,

 =1− 
~

where  is the Hermitian generator of the symmetry operator. Thus,

[ ] = 0

By virtue of the Heisenberg equation of motion, we have


= 0


Hence  is a constant of the motion. In the language of quantum mechanics, we say

that  and  can be diagonalized simultaneously. Suppose |i an eigenstate of .

Then, the state at a later time

| 0  i = ( 0 ) |i

is also an eigenstate of G.

G Proof.

 | 0  i = ( 0 ) |i

= ( 0 ) |i

=  ( 0 ) |i 

186
CHAPTER 4 – MANUSCRIPT

B 4.1.3 Degeneracy

Suppose that

[ ] = 0

for the symmetry operator  and |i is an energy eigenket with eigenvalue  . Then,

 |i is also an eigenket of H with the same energy eigenvalue.

G Proof.

 |i =  |i =   |i 

Suppose |i and  |i represent two different kets. Then these two states with the

same energy are said to be degenerated.

Remark 3 The proof is very trivial, but the concept plays a far more important role

in quantum mechanics.

B 4.1.4 Symmetry and symmetry breaking

If |i is non-degenerated,  |i and |i must represent the same physical state although

they can differ a trivial phase factor, i.e.,

 |i =  |i 

In the case, we say that the state |i also possesses the symmetry. If

 |i 6=  |i 

187
CHAPTER 4 – MANUSCRIPT

we say that the symmetry is spontaneously broken in the state although the Hamil-

tonian possesses the symmetry. The spontaneous symmetry breaking occurs when the

state is degenerated. The spontaneous broken symmetry is one of the most impor-

tant concepts in modern physics, from elementary particle physics to condensed matter

physics.

In 1960 Nambu was the first one to introduce the concept of spontaneous

symmetry breaking to elementary physics. To begin with he first worked on supercon-

ductivity. Then he applied the concept to elementary particles, which becomes one of

the corner stones of the physics of elementary particles, the standard model. Nambu

was awarded Nobel prize in 2008..

EXP Example 1

for a hydrogen atom problem,

1 2
=  +  ()
2

The Hamiltonian is rotational invariant

[() ] = 0

One of eigenstates is denoted by |  i with  . Then

X ()
() |  i = 0 |  0 i
0

188
CHAPTER 4 – MANUSCRIPT

is also an eigenket of  with the same energy  . For each pair of

 and  there are 2 + 1 values of :

 = − − + 1  

Therefore |  i should be at least (2 + 1)-fold degenerated.

EXP Example 2

Two-spin Problem

The Hamiltonian for two spins S1 and S2 is written as

 = S1 · S2 

Denote the eigenstate for H by |  i

£ ¤
 |  i = (S1 + S2 )2 − S21 − S22 |  i
2


= [ ( + 1) − 2( + 1)] |  i
2

The total spin can be

 = 2 2 − 1  0

As the M can be taken to be

 = −  − + 1   

189
CHAPTER 4 – MANUSCRIPT

and the eigenvalue of H is independent of M, so the state |  i

is (2 + 1)-fold degenerated.

The ground state energy:

(1) Antiferromagnetic   0,  = 0

 = −( + 1)

(2) Ferromagnetic   0,  = 2

 = 2 

SU(2) symmetry: When the system is rotated the rotation

operator
(S1 +S2 )·
 =  ~

As

[ S] = 0

we have

 †  = 

For the ground state

(1)   0,  = 0

 | = 0  = 0i = | = 0  = 0i

The state also possesses the SU(2) symmetry.

190
CHAPTER 4 – MANUSCRIPT

(2)  0,  = 2

 | = 2  = 0i 6= | = 2  = 0i

The ground state is (2 + 1)-fold degenerate. The symmetry is

spontaneously broken.

B 4.1.5 Summary: symmetries in physics

Space Translation  →  +  Momentum

Time evolution  →  +  Energy

Rotation  → 0 Angular Momentum

Space Inversion  → − Parity

Time reversal  → − Charge conjugate

Permutation (12) → (21) Quantum Statistics

Gauge Ψ →e Ψ Charge

C Space translation

† (ρ)r (ρ) = r + ρ

P
 (ρ) = exp[− · ρ]
~

† (ρ) (ρ) = 

[P ] = 0

191
CHAPTER 4 – MANUSCRIPT

C Time evolution

† ( ) ( ) =  + 


 ( ) = exp[− ]
~

† ( ) ( ) = 

|()i = exp [−~] |(0)i

C Rotation

0 = 

|i = D() |i


∙ ¸
J·n
D() = exp −
~

D† ()D() = 

[ ] = 0

C Gauge Invariance


Ψ = ~ Ψ


This symmetry is related to the conservation of “charge”, i.e., the number of the par-

ticles. To sovle the Schrodinger equation, we find that there is a degree of freedom to

add a trivial phase factor before the wave function. This originates from the symmetry.

192
CHAPTER 4 – MANUSCRIPT

A 4.2 Discrete Symmetries

In this section we consider two symmetry operators that can be considered to be discrete

B 4.2.1 Parity

Figure 4.1:

The parity operation, as applied to transformation on the coordinate system,

changes a right-handed (RH) system into a left-handed (LH) system. Let us denote the

parity operator Π, which satisfies






⎪ |i → Π |i






⎨ Π† Π = −



⎪ { Π} = 0






⎩ Π† Π = 1

Π† Π = 1−  Π† = Π−1 = Π

193
CHAPTER 4 – MANUSCRIPT

The physical meaning is that under this operation

 = (  ) ⇒  0 = (− − −)

 0 = 

where ⎛ ⎞
⎜ −1 0 0 ⎟
⎜ ⎟
⎜ ⎟
=⎜
⎜ 0 −1 0


⎜ ⎟
⎝ ⎠
0 0 −1
How does an eigenket of the position operator transform under parity?Assume

̂ |i =  |i

Π̂ |i = Π |i = Π |i

−̂Π |i = Π |i

̂ (Π |i) = − (Π |i)

Thus we conclude that

Π |i = − |−i 

Applying the parity operation twice

Π2 ̂ = Π (−̂Π) = ̂Π2

and

̂Π2 |i = Π2 ̂ |i = Π2 |i

194
CHAPTER 4 – MANUSCRIPT

Thus

Π2 = 1

Therefore its eigenvalue can be only +1 or −1.

Example: a simple harmonic oscillator

B 4.2.2 The Momentum Operator

Space translation followed by parity is equivalent to parity followed by translation in

the opposite direction:

ΠT () |i = Π | + i

=  |− − i

= T (−) |−i

= T (−)Π |i

Since

T () = 1 −   + 
~

it follows that

Π = − Π

or

{Π  } = 0

195
CHAPTER 4 – MANUSCRIPT

Example: a free particle

B 4.2.3 The Angular Momentum

Define

L =  × 

then

ΠL = LΠ

For 3 × 3 matrices, we have

() () = () ()

where ⎛ ⎞
⎜ −1 0 0 ⎟
⎜ ⎟
⎜ ⎟
() =⎜
⎜ 0 −1 0


⎜ ⎟
⎝ ⎠
0 0 −1

which implies that parity and rotation operation commute. Therefore

Π() = ()Π

() = 1 −  · } + · · ·

196
CHAPTER 4 – MANUSCRIPT

Π = Π

Π = Π

Π = Π

Example: the hydrogen atom

B 4.2.4 Parity

Let us now look at the parity property of wave function whose state ket is |i :

Ψ() = h|i 

Suppose |i the eigenket of parity

Π |i =  |i  ( = ±1)

Then

h| Π |i =  h |i

h− |i =  h |i

Ψ(−) = Ψ()

The wave function is also parity.

EXP Example 3

197
CHAPTER 4 – MANUSCRIPT

Solution for a square-well potential

~2 2  ()
− = ()
2 2

with || ≤  In the replacement  → −,

~2 2  (−)
− = (−)
2 2

Thus if () is one solution, (−) must be also one solution. The

solutions are

() = ±·}

where 2 = 2}2  We can construct

Odd Parity: () − (−) =  ()

Even Parity:  () + (−) =  ()

Thus the wave functions are sin ~ and cos ~ respec-

tively.

EXP Example 4

Simple harmonic oscillator: The basic Hamiltonian is

2 1
= +  2 2 
2 2

We have

Π† Π = 

198
CHAPTER 4 – MANUSCRIPT

and ⎫ ⎧
⎪ ⎪
Π† xΠ = −x ⎪
⎬ ⎪
⎨ Π† † Π = −†

⎪ ⎪
Π pΠ = −p ⎪
† ⎭ ⎪
⎩ Π† Π = −

The eigenkets are

1 ¡ † ¢
|i =  |0i
!

The ground state wave function is


∙ ¸
1 1 2
h|0i = 12 exp − 2
 0 2 0

So

h|0i = h−|0i

How about h|i?

h|Π|i = (−1) h|i 

h−||i = (−1) h|i

At  = 0, we have h = 0|i = 0 for odd .

EXP Example 5

Double well potential

Let us now look at the parities of energy eigenstates.

199
CHAPTER 4 – MANUSCRIPT

Polar vector: vectors that are odd under parity such as x, p etc

Axial vector or pseudovectors that are even under parity such as J.

Pseudoscalar: scalars that are odd under parity S · x.

Theorem 4 Suppose [ Π] = 0 and |i is an eigenstate of H with eigenvalue 

 |i =  |i

then (a) if |i is nondegenerate, Π |i is also a parity eigenket; (b) if |i is degenerate,

Π |i may not be a parity eigenket.

G Proof. We can construct two states

1
(1 ± Π) |i 
2

These two states are parity eigenstates with eigenvalues ±1

1 1¡ ¢
Π (1 ± Π) |i = Π ± Π2 |i
2 2

1
= ± (1 ± Π) |i
2
1
where the identity Π2 = 1 has been used. Furthermore the states 2
(1 ± Π) |i and |i

must represent the same state otherwise there would be two states with the same energy.

This violating the condition of non-degeneracy.

B 4.2.5 Lattice Translation

200
CHAPTER 4 – MANUSCRIPT

FIGURE (a) Periodic potential in one-dimension with periodicity a. (b) The periodic

potential when the barrier height between two adjacent lattice sites becomes infinite.

(from Sakurai, 1994)

Consider a periodic potential in one dimension:

 ( ± ) =  ()

We introduce a translational operator  () such that

 † () () =  + ;

 † () () () =  ( + )

Thus the Hamiltonian is invariant under the translation

 † () () = 

As  is unitary, we have

 =  

201
CHAPTER 4 – MANUSCRIPT

 and  commute with each other. So  can be diagonalized with  simultaneously. 

is unitary, but not hermitian.  † () 6=  () We expect the eigenvalue to be a complex

number of modulus 1.

Before we determine the energy eigenkets and eigenvalues of   it is instructive

to see a special case of the periodic potential when the barrier height between the two

adjacent lattice sites is made to go to infinite as shown in Figure. Denote |i a state

ket in which the particle is localized at the  site. When  is applied to it

 () |i = | + 1i 

To construct an eigenket for  , we consider a linear combination

X
+∞
|i =  |i 
=−∞

This state is an eigenstate for  

X
+∞ X
+∞
 () |i =   () |i =  | + 1i
=−∞ =−∞

X
+∞
−
=  (+1) | + 1i
=−∞

 () |i = − |i

B 4.2.6 A more realistic example: a 1D system

Assume that the barrier potential is still very large, but finite. The overlap of two states

at nearest neighbor sites is very small, and all other overlaps can be ignored. That is

202
CHAPTER 4 – MANUSCRIPT

h|  |i = 0

h|  | ± 1i = −

and all other elements are equal to zero. That is

 |i = 0 |i −  | + 1i −  | − 1i

X
 = {0 |i h| −  | + 1i h| −  | − 1i h|}

X
 |i = {0 |i h| −  | + 1i h| −  | − 1i h|} |i

= (0 − 2 cos ) |i

The skill used here is the Fourier transform.

Periodic condition: |i = | + i

X

|i =  |i
=0

X

=  | + i
=0

X

= −  (+) | + i
=0

− = 1  = 2 = 2 ( = 0 1 · · ·   − 1)

203
CHAPTER 4 – MANUSCRIPT

A 4.3 Permutation Symmetry and Identical Particles

B 4.3.1 Identical particles

Identical particles cannot be distinguished by means of any inherent properties, since

otherwise they could not be identical in all respects. In classical mechanics the existence

of sharply definable trajectories for individual particles makes it possible in principles to

distinguish between particles that are identical except for their path since each particle

can be followed during the course of an experiment. In quantum mechanics, the finite

size and the spreading of the wave packets that can describe individual particles often

make it impossible to distinguish between identical particles because of their positions,

especially if they interact with each others to an appreciable extent.

C Distinguishable and indistinguishable identical particles

Two identical particles are indistinguishable if the wave functions for the two particles

overlap, and distinguishable if the wave functions do not overlap.

C Permutation Symmetry

Let us consider the Hamiltonian of a system of two identical particles.

1 2 1 2
= 1 +  +  (|1 − 2 |) +  (1 ) +  (2 )
2 2 2

Symmetry: to exchange the positions of two particles: 1 ⇔ 2, the Hamiltonian

204
CHAPTER 4 – MANUSCRIPT

is invariant

 ⇐⇒ 

Define the permutation operator

12 |1 i1 ⊗ |2 i2 = |2 i1 ⊗ |1 i2


12 (1  2 )12 = (2  1 )

Clearly

12 = 21

2
12 = 1

Consider a state ket for two particles:

|i1 ⊗ |0 i2

12 |i1 ⊗ |0 i2 = |0 i1 ⊗ |i2

We can construct two states

1
[(|i1 ⊗ |0 i2 ± |0 i1 ⊗ |i2 )]
2

which are eigenstates of 12 with eigenvalues ±1, respectively.

Our consideration can be extended to a system made up of many identical

particles

 |1  2  · · ·  · · ·    · · ·   i = |1  2  · · ·  · · ·    · · ·   i

205
CHAPTER 4 – MANUSCRIPT

Clearly

2 = 1

just as before, and the allowed eigenvalues of  are ±1 It is important to note that,

in general

[   ] 6= 0

C Connection between spin and statistics

Half-odd-integer spin particles are fermions: electrons, protons, · · ·

Integer spin particles are boson: mesons, 4 He nucleus,· · ·

Although the allowed values of  are ±1, system containing N identical

particles are either totally symmetrical or antisymmetric under the interchange of any

pair.

Boson:

 | i = | i

Fermions

 | i = − | i

It is empirical fact that a mixed symmetry does not occur.

206
CHAPTER 4 – MANUSCRIPT

C The Pauli exclusion principle

An immediate consequence of the electron being a fermion is that the electrons must

satisfy the Pauli exclusion principle, which states that no two electrons can occupy the

same state.

Let us consider a system containing N electrons without interaction

(1 2  · · ·  ) = 0 (1) + 0 (2) + · · · + 0 ()

Assume  () are the wave functions of 0 () with energy 

0 () () =   ()

The product of N one-particle eigen functions are the energy eigen wave function of

total system..

 (1 2 · · ·  ) =  (1) () · · ·  ()

 = ( +  + · · · +  )

As electrons are fermions, we have to express the wave function in an antisymmetric

207
CHAPTER 4 – MANUSCRIPT

form. One way is to express as a determinant of the U’s

 (1 2 · · ·  )
⎛ ⎞
⎜  (1)  (2) · · ·  () ⎟
⎜ ⎟
⎜ ⎟
⎜  (1)  (2) · · ·  () ⎟
⎜    ⎟
= det ⎜
⎜ .
⎟

⎜ .. .. ... .. ⎟
⎜ . . ⎟
⎜ ⎟
⎝ ⎠
 (1)  (2) · · ·  ()

This function is antisymmetric and satisfies the Schrodinger equation with energy

 +  + · · · +  

If two or more of the U’s are the same,

 (1 2 · · ·  ) = 0

which is the so-called Pauli exclusion principle.

To simplify the notation we introduce the creation and annihilation operator

† and  such that

0 † |0i =  † |0i

 |0i = 0

 † + †  = 1

† † + † † = 0

208
CHAPTER 4 – MANUSCRIPT

for fermions.

 † − †  = 1

† † − † † = 0

for bosons. The Shcrodinger equation can be reduced to

 | i = ( +  + · · · +  ) | i

with

| i = † † · · · † |0i 

EXP Example 6

Two-particle problem

A system contains two particles, which is described by the

Hamiltonian

 = 1 + 2

where

1 2 1
1 =  + 1 21
21 1 2

1 2 1
2 = 2 + 2 22
22 2

(1) 1 6= 2 , 1 6= 2

(2) 1 = 2 , 1 = 2

209
CHAPTER 4 – MANUSCRIPT

B 4.3.2 Two electrons

B 4.3.3 The helium atom

A 4.4 Time Reversal

In this section we study a discrete symmetry, called time reversal. The terminology

was first introduced by E. Wigner in 1932.

B 4.4.1 Classical cases

Let us first look at the classical case: a motion of particle subjected to a certain force.

Its trajectory is given by the Newtonian equation of motion,

2 r
 = −∇ ()
2

If r() is the solution of the equation, then r(−) is also the solution of the equation.

In another word, when we make a transformation  → −, the Newtonian equation

of motion keeps unchanged. Of course we should notice the change of the boundary

condition or initial conditions for the problem.

The Maxwell equations and the Lorentz force

F = (E + v × B)

210
CHAPTER 4 – MANUSCRIPT

are invariant under the time reversal provided that

v → −v

j → −j

 → 

B → −B

E → E

1
∇·D = 
0

∇·B = 0

D
∇×H− = 


B
∇×E+ = 0


where D = 0 E + P and H = B0 − M

The Schrodinger equation

Ψ( ) ~2 2
~ = (− ∇ +  )Ψ( )
 2

Ψ( −) is not a solution of the equation because of the appearance of the first order

time derivative and the imaginary sign at the left hand side. However, Ψ∗ ( −) is a

solution.

211
CHAPTER 4 – MANUSCRIPT

Definition: the transformation

¯ E
¯
|i → |̃i =  |i  |i → ¯̃ =  |i

is said to be antiunitary if

D E
̃|̃ = h|i∗ ;

 (1 |i + 2 |i) = ∗1  |i + ∗2  |i 

In this case the operator  is antiunitary operator.

Usually an antiunitary operator can be written as

 = 

where  is a unitary operator and  is the complex conjugator operators.

B 4.4.2 Time reversal operator Θ

Let us denote the time reversal operator by Θ Consider

|i → Θ |i

where Θ |i is the time reversed state. More appropriately, Θ |i should be called the

motion-reversed state. For a momentum eigenstate |pi  Θ |pi should be |−pi up to

a possible phase. Θ is an anti-unitary operator. We can see this properties from the

Schrodinger equation of a time reversal invariant system.

212
CHAPTER 4 – MANUSCRIPT


~ Ψ( ) = Ψ( )


 −1
Θ~ Θ ΘΨ( ) = ΘΘ−1 ΘΨ( )


 −1 
Θ~ Θ = ~ ; ΘΘ−1 = 
 
 
provided that ΘΘ−1 = − and Θ  Θ−1 = (−)

ΘpΘ−1 = −p

ΘxΘ−1 = x

ΘJΘ−1 = −J (= x × p)

Check the invariance of the fundamental commincation relation

£ ¤
Θ x  p Θ−1 = ~ 

From the spherical harmonic  ( ) we have

 ( ) → ( ( ))∗ = (−1) − ( )

Θ | i = (−1) | −i

213
CHAPTER 4 – MANUSCRIPT

B 4.4.3 Time reversal for a spin 1/2 system

For a spin-1/2 system,

Θ =   

where  is the complex conjugate operator. Its inverse matrix is

Θ−1 = −Θ = − 

Consider the eigenstate of S · n with eiegenvalue ~2

| +i = − ~ − ~ |+i

Θ | +i = Θ− ~ − ~ Θ−1 Θ |+i

Because Θ Θ−1 = − and ΘΘ−1 = −

Θ | +i = − ~ − ~ Θ |+i = − ~ − ~ |−i

= | −i

~
 |+i = |+i
2

~
Θ Θ−1 Θ |+i = Θ |+i
2

~
− Θ |+i = Θ |+i
2

Θ |+i =  |−i

214
CHAPTER 4 – MANUSCRIPT

On the other hand,

| −i = − ~ − (+)~ |+i = − ~ − ~ − ~ |+i

Noting that  acting on |+i gives |+i  We have

Θ = − ~  =   

Θ |+i = + |−i

Θ |−i = − |+i

Θ2 = −1

Θ(+ |+i + − |−i) = ∗+ |+i − ∗− |−i

Θ(∗+ |+i − ∗− |−i) = − (+ |+i + − |−i)

In general

Θ = − 

Θ2 = (−1)2

Kramers degeneracy: the ground state for the odd number of electrons in a

time-reversal invariant system has at least double degeneracy.

Time reversal symmetry and energy conservation

215
CHAPTER 4 – MANUSCRIPT

2  
 2 = −2 + 
 

216
CN Chapter 5

CT Approximation Methods for Bound

States

As in the case of classical mechanics, there are relatively few physically interesting prob-

lems in the quantum mechanics which can be solved exactly. Approximation methods

are therefore very important in nearly all the applications of the theory. In this chapter

we shall develop several approximation methods.

Approximation methods can be conveniently divided into two groups, accord-

ing to whether the Hamiltonian of the system is time-independent or time-dependent.

217
CHAPTER 5 – MANUSCRIPT

A 5.1 The Variation Method

The variation method can be used for the approximate determination of the lowest or

ground-state energy level of a system when there is no closely related problem that is

capable of exact solution, so that the perturbation method is inapplicable. The most

famous achievements of the variational method in modern physics are the BCS theory

for superconductivity and Laughlin’s theory for fractional quantum Hall effect.

Why are we so interested in the lowest energy states?

The third law of thermodynamics: "As a system approaches absolute zero, all

processes cease and the entropy of the system approaches a minimum value."

B 5.1.1 Expectation value of the energy


¯ ®
We attempt to estimate the ground state energy by considering a trial ket, ¯0̃  or a

trial wave function. To this end we first prove a theorem of a great of importance. We

define the expectation value ̄ such that


­ ¯ ¯ ®
0̃¯  ¯0̃
̄ = ­ ® 
0̃|0̃

Then we can prove the following

Theorem 5 The expectation value of a trial state ket is always not less than the ground

state energy

̄ ≥ 0 

218
CHAPTER 5 – MANUSCRIPT

G Proof. We can always imagine that any trial state can be expanded as

¯ ® X ­ ®
¯0̃ = |i |0̃

where |i is an exact eigenket of H. The expectation value


­ ¯ ¯ ® P ­ ® ­ ®
0̃¯  ¯0̃  0̃| h||i |0̃
̄ = ­ ® = ­ ®
0̃|0̃ 0̃|0̃
P ­ ®­ ® P ­ ®­ ®
 0̃| |0̃  0̃| |0̃ 0
= ­ ® ≥  ­ ® = 0 
0̃|0̃ 0̃|0̃

In the last step we use the property  ≥ 0 

The variation method does not tell us what kind of trial state ket are to

estimate the ground state energy. Quite often we must appeal to physical intuition. In

practice, we characterize the trial state ket by one or more parameters  can compute

̄ as a function of   Then we minimize the ̄ by setting the derivatives with respect

to the parameters zero


 ̄
= 0


The result of the variation method is an upper limit for the ground state energy of the

system, which is likely to be close if the form of the trial state ket or the trial wave

function resembles that of the eigenfunction. Thus it is important to make use of any

available information such as symmetry analysis or physical intuition in choosing the

trial function.

219
CHAPTER 5 – MANUSCRIPT

B 5.1.2 Particle in a one-dimensional infinite square well

We attempt to estimate the ground state energy of the infinite-well (one-dimensional

box) problem defined by ⎧




⎨ 0 for ||  ;
 =


⎩ +∞ for   

The exact solution are, of course, well known:

1 ³  ´
h|0i = cos
12 2
µ 2 ¶µ 2 ¶
~ 
0 = 
2 42

But suppose we did not know these. Evidently the wave function must vanish at

 = ± Further the ground state should have even parity. The simplest analytic wave

function that satisfies both requirements is

h|0i = 2 − 2 

There is no variational parameter. We can compute ̃ as follows


R  2
 h0|i 2 h|0i
̃ = R
 h0|i h|0i

10
= 0 ≈ 101320 
2

A much better result can be obtained if we use a more general trial function

h|0i =  − || 

220
CHAPTER 5 – MANUSCRIPT

where  is regarded a variational wave parameter. Straightforward algebra gives

( + 1) (2 + 1) ~2
̃ = 
2 − 1 42

which has a minimum at


1 + 612
= ≈ 172
2

This gives
5 + 2612
̃ = 0  ≈ 1002980
2

The state with an odd parity?

h|0i =  ( −  ) 

B 5.1.3 Ground State of Helium Atom

We use the variation method with a simple trial function to obtain an upper limit for

the energy of the ground state of the helium atom. The helium atom consists of a

nucleus of charge +2 surrounded by two electrons, we find that its Hamiltonian is

~2 ¡ 2 ¢
 = − ∇1 + ∇22
2
µ ¶
2 1 1 2
−2 + +
1 2 12

= 1 + 2 + 12

where r1 and r2 are two position vectors of two electrons with respect to the nucleus

2
as origin, and 12 = |r1 − r2 |  If we neglect the interaction energy 12
between the two

221
CHAPTER 5 – MANUSCRIPT

electrons, the ground state wave function of H would be product of two normalized

hydrogenic wave functions

 3 −(0 )(1 +2 )


Ψ (r1  r2 ) =  (5.1)
30

with 0 = ~2 2 and  = 2

Why is the wave function symmetric in the above equation? Spin singlet in

the ground state.

Why  = 2?

We shall use Eq.(5.1) as the trial wave function and permit Z to be the vari-

ation parameter. The kinetic energy of each electron is


Z µ ¶
∗ ~2 2 2  2
1 = r1 r2 Ψ (r1  r2 ) − ∇ Ψ (r1  r2 ) = 
2 1 20

The Coulomb energy of each electron is


Z µ ¶
∗ 22 22 
1 = r1 r2 Ψ (r1  r2 ) − Ψ (r1  r2 ) = − 
1 0

The interaction energy of two electrons is


Z µ ¶
∗ 2 52 
 = r1 r2 Ψ (r1  r2 ) Ψ (r1  r2 ) = − 
12 80

Here we make use of the formula


∞ µ ¶
1 1 X 2
=  (cos ) 1  2
12 1 =0 1

∞ µ ¶
1 1 X 1
=  (cos ) 2  1
12 2 =0 2

222
CHAPTER 5 – MANUSCRIPT

where  is the spherical harmonics and  is the angle of these two vectors.

Totally, the expectation value of the Hamiltonian is

2  2 42  52 
̃ = − +
0 0 80
µ ¶
2 2 27
=  − 
0 8
"µ ¶2 µ ¶2 #
2 27 27
= − − 
0 16 16

The expectation is minimized at

 = 2716 ≈ 169

and
µ ¶2
27 2 2
̃ = − = −285 
16 0 0

The experimental value for a helium atom is 29042 0  so our estimation is about

1.9% higher.

A 5.2 Stationary Perturbation Theory: Nondegener-

ate Case

B 5.2.1 Statement of the Problem

In this section we shall discuss the Rayleigh-Schrodinger perturbation theory, which

analyses the modifications of discrete energy levels and the corresponding eigenfunctions

223
CHAPTER 5 – MANUSCRIPT

of a system when a perturbation is applied.

We consider a time independent Hamiltonian H such that it can be split into

two parts, namely

 = 0 + 

The perturbation theory is based on the assumption that when  = 0, both the exact
¯ ®
eigenvalues  and the exact energy eigenkets ¯(0) of 0 are know
(0)

¯ ® ¯ ®
0 ¯(0) = (0) ¯(0) 

We are required to find approximate eigenkets and eigenvalues for the full Hamiltonian

problem

(0 +  ) |i =  |i 

In general,  is not the full potential operator. We expect that the  is not very

“large” and should not change the eigenvalues and eigenkets of 0 very much.

It is customary to introduce a parameter  such that we solve

(0 +  ) |i =  |i

instead of the original one. We take 0 ≤  ≤ 1 The case of  = 0 corresponds to the

unperturbed problem and the case of  = 1 corresponds to the full strength problem

we want to solve.

224
CHAPTER 5 – MANUSCRIPT

B 5.2.2 The Two-State Problem

Before we embark on a systematic presentation of the basic method, let us see how the

expansion in  might indeed be valid in the exactly soluble two-state problem we have

encountered many times already. Suppose we have a Hamiltonian that can be written

as

 = 0 + 

¯ ®­ ¯ ¯ ®­ ¯
= 1 ¯1(0) 1(0) ¯ + 2 ¯2(0) 2(0) ¯
(0) (0)

¯ ®­ ¯ ¯ ®­ ¯
+12 ¯1(0) 2(0) ¯ + 21 ¯2(0) 1(0) ¯

¯ ®
where  and ¯(0) are the exact eigenvalues and eigenstates for 0 . The index “0”
(0)

stands for the unperturbed problem. We can also express the Hamiltonian by a square

matrix ⎛ ⎞
(0)
⎜ 1 12 ⎟
=⎜

⎟

(0)
21 2
What is the eigenvalues of ?
⎡Ã ! ⎤12
(0) (0) (0) (0) 2
1 + 2 1 − 2
1 = +⎣ + 2 12 21 ⎦ ;
2 2

⎡Ã ! ⎤12
(0) (0) (0) (0) 2
1 + 2 1 − 2
2 = −⎣ + 2 12 21 ⎦ 
2 2

So this problem can be solved exactly. Let us suppose  is small compared with

the relevant energy scale, the difference of the energy eigenvalues of the unperturbed

225
CHAPTER 5 – MANUSCRIPT

problem:
¯ ¯
¯ (0) (0) ¯
 |12 | ¿ ¯1 − 2 ¯

Using the formula


1 1
(1 + )12 = 1 +  − 2 + · · ·  (5.2)
2 8

we obtain the expansion of the energy eigenvalues in terms of 

(0) 2 12 21


1 = 1 + (0) (0)
+··· ;
1 − 2

(0) 2 12 21


2 = 2 − (0) (0)
+ ··· 
1 − 2

As the expansion in Eq.(5.2) holds for   1, the expansion is available in the case of

small  .

B 5.2.3 Formal Development of Perturbation

We are dealing with the perturbation of a bound state. The assumption that  is small

compared with the energy scale suggests that we expand the perturbed eigenvalues and

eigenkets as power series in  . This is most conveniently accomplished in the terms of

the parameter  such that the zero, first, etc., powers of  correspond to the zero, first,

etc., orders of the perturbation calculation.

226
CHAPTER 5 – MANUSCRIPT

The perturbed wave function and energy level are written

¯ ® ¯ ® ¯ ®
|i = ¯(0) +  ¯(1) + 2 ¯(2) + · · · ;

 = (0) + (1) + 2 (2) + · · ·

and are substituted into the wave equation to give

¡¯ ® ¯ ® ¯ ® ¢
(0 +  ) ¯(0) +  ¯(1) + 2 ¯(2) + · · ·

¡ (0) ¢ ¡¯ ® ¯ ® ¯ ® ¢
=  + (1) + 2 (2) + · · · × ¯(0) +  ¯(1) + 2 ¯(2) + · · · 

Since the Equation is supposed to be valid for a continuous range of , we can equate

the coefficients of equal powers of  on the both sides to obtain a series of equations

that represent successively higher orders of the perturbation.

¡ ¢¯ ®
0 − (0) ¯(0) = 0 (5.3)

¡ ¢¯ ® ¯ ®
0 − (0) ¯(1) = ((1) −  ) ¯(0) (5.4)

¡ ¢¯ ® ¯ ® ¯ ®
0 − (0) ¯(2) = ((1) −  ) ¯(1) + (2) ¯(0) (5.5)

¡ ¢¯ ® ¯ ® ¯ ® ¯ ®
0 − (0) ¯(3) = ((1) −  ) ¯(2) + (2) ¯(1) + (3) ¯(0) (5.6)

..
.

Zero-order perturbation

227
CHAPTER 5 – MANUSCRIPT
¯ ®
The Eq. (5.3) means that ¯(0) is one of the unperturbed eigenkets, as
(0)
expected. We can obtain it directly by taking  = 0. The eigenvalue is   Since we
¯ ®
are dealing with the perturbation of a bound state, the state ket ¯(0) is discrete. It

is assumed to be nondegenerate as well, although others of unperturbed eigenkets may

be degenerate or continuously distributed in energy.

First-order perturbation

Multiplying an eigen bra of the unperturbed Hamiltonian from the left hand

on both sides of the Eq.(5.4), we obtain

­ (0) ¯ ¡ ¢¯ ® ­ ¯ ¯ ®
 ¯ 0 − (0) ¯(1) = (0) ¯ ((1) −  ) ¯(0) (5.7)

­ (0) ¯ ¡ ¢¯ ® ­ ¯ ¯ ®
 ¯ 0 − (0) ¯(1) = (0) ¯ ((1) −  ) ¯(0) (5.8)

where  6= . The Eq. (5.7) gives

­ ¯ ¯ ®
(1) = (0) ¯  ¯(0) =  

¯ ®
It is convenient to calculate ¯(1) by expanding it in terms of the unperturbed eigenkets

¯ (1) ® X (1) ¯ (0) ®


¯ =  ¯

where the summation runs over all possible eigenkets of 0 . To simplify the notation,

we define
­ ¯ ¯ ®
 = (0) ¯  ¯(0)

228
CHAPTER 5 – MANUSCRIPT

From the Eq.(5.5),

¡ ¢ X (1) ¯ (0) ® ¯ ®
0 − (0)  ¯ = ((1) −  ) ¯(0)

X ¡ (0) ¢¯ ® ¯ ®
(0) ¯ (0)
(1)
  −   = ((1) −  ) ¯(0)

­ ¯
multiplying (0) ¯ from the left hand side, we obtain


(1)
 = (0) (0)

 − 

Second-order perturbation

The second-order correction to the energy is

­ ¯ ¯ ® X  
(2) = (0) ¯  ¯(1) = (0) (0)

6=  − 

¯ ®
Similarly, we expand ¯(2) in terms of the unperturbed eigenkets

¯ (2) ® X (2) ¯ (0) ®


¯ =  ¯ 

­ ¯
To calculate  , we apply (0) ¯ from the left hand on both side of the Equation and
(2)

obtain

­ (0) ¯ ¡ ¢¯ ® ­ ¯ ¯ ® ­ ¯ ¯ ®
 ¯ 0 − (0) ¯(2) = (0) ¯ ((1) −  ) ¯(1) + (0) ¯ (2) ¯(0) 

That is

1 ©­ (0) ¯ (1) ¯ ® ­ ¯ ¯ ®ª
(2)
 = (0) (0)
 ¯ ( −  ) ¯(1) + (0) ¯ (2) ¯(0) 
 − 

Summary

229
CHAPTER 5 – MANUSCRIPT

In short, the explicit expression for the energy expansion is

X  
 = (0) +  + 2 (0) (0)
+ ··· 
6=  − 

The expansion of the perturbed eigenket is as follows

¯ ®
|i = ¯(0)

X  ¯ (0) ®
+ ¯
(0) (0)
6=  − 
XX  ln ¯ (0) ®
+2 ¯
(0) (0) (0) (0)
6= 6=  −   − 
X   ¯ (0) ®
−2 ¯
(0) (0) (0) (0)
6=  −   − 

+··· 

Remark 6 The state ket |i is not normalized. A normalized state ket should be

1
|i
h|i

A 5.3 Application of the Perturbation Expansion

To illustrate the perturbation method we developed in the last lecture, let us look

at several examples. All examples can be solved analytically. We can compare the

perturbation results with the exact results.

230
CHAPTER 5 – MANUSCRIPT

B 5.3.1 Simple harmonic oscillator

The first two examples concern a simple harmonic oscillator whose unperturbed Hamil-

tonian is the usual one


2 1
0 = + 2 2 
2 2

As we already know the Hamiltonian can be solved in this way


µ ¶
† 1
0 = ~   +
2

where

³  ´12 ³  ´
† = −
2~ 
³  ´12 ³  ´
 = +
2~ 

with the eigenvalues


µ ¶
1
 = ~  + 
2

EXP Example 1

0 = 2  2 2

This problem can be solved exact in this way:

 = 0 + 0

2 1
= + (1 + ) 2 2 
2 2

231
CHAPTER 5 – MANUSCRIPT

The eigenvalues are


µ ¶
12 1
 = ~(1 + )  + 
2

In the perturbation approach, we just concern about the ground

state energy. The first order perturbation is

1 
1 = h0| 0 |0i =  h0|  2 2 |0i = ~
2 4

The second order perturbation is

X |h| 0 |0i|2 |h2| 0 |0i|2 2


2 = = = − ~
6=0
0 −  0 − 2 16

Thus the perturbation correction to the ground state energy is


µ ¶
 2 3
4 = ~ − + 0( ) 
4 16

Compare it with the exact result.

EXP Example 2

0 = 

This problem can also be solved exactly:

2 1
 = 0 + 0 = + 2 2 + 
2 2

2 1 1
= +  2 ( − 0 )2 − 2 20
2 2 2

232
CHAPTER 5 – MANUSCRIPT

where

0 = − 
 2

Making a replacement,  =  − 0 and  =  , we have

~2 2 1 2 2 1
=− 2
+   −  2 20 
2  2 2

The eigenvalues are


µ ¶
1 2
 = ~  + − 
2 22

The first order perturbation is

1 = h0| 0 |0i = 0

The second order perturbation is

X |h| 0 |0i|2
2 =
6=0
0 − 

|h1| 0 |0i|2
=
0 − 1

2
= − 
2 2

B 5.3.2 Atomic hydrogen

The Schrodinger equation for a hydrogen atom is

~2 2 2
(− ∇ − )Φ = Φ
2 

233
CHAPTER 5 – MANUSCRIPT

In the spherical coordinates,


µ ¶ µ ¶
2 1  2  1   1 2
∇ = 2  + 2 sin  + 2 2 
    sin     sin  2

The wave function can be written as

Φ = () ( )

where  ( ) is a spherical harmonic. (See Schiff, P.76 or Sakurai, P. 454). It is the

eigenfunction for orbital angular momentum operators

2  ( ) = ( + 1)~2  ( )

  ( ) = ~ ( )

First several  :


r
1 3
00 = √ ; 10 = cos 
4 4
r
3
1±1 = ∓ sin ±
8

The Schrodinger equation is reduced to


∙ 2µ µ ¶ ¶ ¸
~ 1  2  ( + 1) 2
−  − −  = 
2 2   2 

The solution of this equation is

 2 4  2 2
 = − = −
2~2 2 20 2

234
CHAPTER 5 – MANUSCRIPT

where  = 0 +  + 1 and 0 = ~2 2  The first three radial functions are


µ ¶32

10 = 2−0
0
µ ¶32 µ ¶
  −20
20 = 2− 
20 0
µ ¶32
  −20
21 = 
60 0

Degeneracy

 = 1: the ground state is nondegenerate

 = 2: the first excited state is four-fold degenerate. (1)  = 0 and  = 0; (2)

 = 1 and  = −1 0 +1

EXP Example 3

Gravitational energy shift

Let consider an ordinary hydrogen atom (proton + elec-

tron) whose unperturbed Hamiltonian is

~2 2 2
0 = − ∇ −
2 

where  =   ( +  ) is the reduced mass. Now the proton

and electron interact not only through the electrostatic potential,

but also by means of the gravational interaction. The perturbation

235
CHAPTER 5 – MANUSCRIPT

due to the gravitational force is

 
 = −

where  is the gravitational constant. The ground state wave func-

tion of atomic hydrogen is

Φ = 10 ()00 ( )

where
µ ¶12
1
00 ( ) =
4

To first order in the perturbation theory, the energy shift of the

ground state of atomic hydrogen due to this perturbation is

(1)
1 = h0| 0 |0i

Z
= r |00 ( )10 ()|2  ()

Z  Z 2 Z ∞
=  sin   2 |00 ( )10 ()|2  ()
0 0 0
Z ∞ µ ¶
2 4 −0    
=  3  − = −
0 0  0

The ground state energy is

(0) 2
1 = − 
20

The relative energy shift is


(1)
1 2 
(0)
= 2
≈ 87 × 10−40 
1 

236
CHAPTER 5 – MANUSCRIPT

Needless to say, it is not necessary to calculate higher order correc-

tion in the present case.

B 5.3.3 Quantum well

Exact solution for the square quantum well.

A 5.4 Stationary Perturbation Theory: Degenerate

Case

Until now we have assumed that the perturbed state differs slightly from an unperturbed

bound state. The perturbation method for a nondegenerate case we developed fails

when the unperturbed state are degenerate.

B 5.4.1 Revisited two-state problem

Starting from the two-state Hamiltonian


⎛ ⎞
(0)
⎜ 1 12 ⎟
=⎜

⎟

(0)
21 2

237
CHAPTER 5 – MANUSCRIPT

(0) (0)
we consider the special case of 1 = 2 , that is, the two states are energy degenerate.

In fact, this problem can be solved exactly; the energy eigenvalues are,

(0) 1
1 = 1 + (12 21 ) 2 ; (5.9)

(0) 1
2 = 1 − (12 21 ) 2  (5.10)

However we cannot get this result by using the nondegenerate perturbation theory.

A blind application of the nondegenerate perturbation formula obviously runs into

difficulty because

(0) 2 12 21


1 = 1 + (0) (0)
+··· ;
1 − 2

(0) 2 12 21


2 = 2 − (0) (0)
+ ··· 
1 − 2
(0) (0)
become singular if 12 is nonvanishing and 1 = 2  So we must modify the method

of the nondegenerate case to accommodate such a situation. One way to avoid the

catastrophe is to choose our base kets in such a way that  has no off-diagonal matrix

element. In the two-state case, we choose a new base ket

¯ (0) ® ¯ ® ¯ ®
¯̃ =  ¯1(0) +  ¯2(0)

the transformed perturbation becomes


⎛ ⎞
12
⎜ (12 21 ) 0 ⎟
̃ =  ⎜

⎟

0 −(12 21 )12

238
CHAPTER 5 – MANUSCRIPT

The solution to the coefficients  and  are

1 (12 21 )12


= ;
1 21

2 (12 21 )12


= −
2 12

and the state should be normalized

| |2 + | |2 = 1

The transformation matrix is ⎛ ⎞


⎜ 1 1 ⎟
 =⎜

⎟

2 2
In this way the degeneracy of the perturbed state is removed. We can consider the per-

turbation correction according to the nondegenerate perturbation theory. Note that we

choose the perturbation potential is diagonal when the unperturbed state is degenerate.

The following problem is left as an exercise:


⎛ ⎞
⎜ 11 12 ⎟
 =⎜



21 22

as a perturbation to a two-fold degenerate state.

B 5.4.2 The basic procedure of degenerate perturbation theory

We see from the two-state example that the degeneracy can be removed by choosing a

new set of base kets. Here we present a more detailed approach. Suppose now that there

239
CHAPTER 5 – MANUSCRIPT

are two states, |i and |i, that have the same unperturbed energy. Then we cannot

apply the nondegenerate perturbation formula when  =  unless it happens that

 = 0. We first consider the case in which  6= 0 the initial state is not specified

by its unperturbed energy; the state may be |i or |i or any linear combination of

them. Let us suppose that the perturbation  removes the degeneracy in some order,

so that for finite  there are two states that have different energy. We start from the

perturbation equations in the last lecture

¡ ¢¯ ®
0 − (0) ¯(0) = 0 (5.11)

¡ ¢¯ ® ¯ ®
0 − (0) ¯(1) = ((1) −  ) ¯(0)  (5.12)

Out of infinite number of combinations of the two states we choose the particular pair

which depends on  From the Eq.(5.11), we obtain

¯ ® ¯ ®
|i =  ¯(0) +  ¯(0) ;

(0) = 
(0)
= (0) 

Substitute the combined state into the Eq.(5.12) and take the inner product of this
¯ ® ¯ ®
equation successively with ¯(0) and ¯(0) , we obtain

¡ ¢
 − (1)  +   = 0

¡ ¢
 − (1)  +   = 0

240
CHAPTER 5 – MANUSCRIPT

The two solutions of this equation are

1 1£ ¤12
(1) = ( +  ) ± ( −  )2 + 4 | |2
2 2

These two solutions are equal if and only if

 = 

 = 0

In this case we say the degeneracy is not removed in the first order perturbation and

we have to consider the second order perturbation. On the other hand if either or both
(1)
of two equations are not satisfied, the two valued of  are distinct, and each can be

used to calculate  and  . In this case, we can use the nondegenerate perturbation
(1)
theory to determine the higher order perturbation. The coefficients  are determined

by
(1) (0) (0)
 ( −  ) = −  − ln 

(1) (1)
for  6=   if we assume that  =  = 0

How to solve triplet degenerate problem?

The basic procedure of the degenerate perturbation theory:

(1) Identify degenerate unperturbed eigenkets and construct the perturbation

matrix  , a  ×  matrix if the degeneracy is g-fold.

(2) Diagonalization the perturbation matrix by solving, as usual, the appro-

priate secular equation.

241
CHAPTER 5 – MANUSCRIPT

(3) Identify the roots of secular equation with the first-order energy shifts; the

base kets that diagonalize the  matrix are the correct zero-order kets to which the

perturbed kets approach in the limit  → 0;

(4) For higher orders use the formulas of the corresponding nondegenerate

perturbation theory except in the summations, where we exclude all contributions from

the unperturbed kets in the degenerate subspace.

B 5.4.3 Example: Zeeman Effect

The change in the energy levels of an atom caused by a uniform external magnetic

field is called the Zeeman effect. We now consider the change of first order in the field

strength for a hydrogen atom

A constant magnetic field can be represented by the vector potential

1
A= B×r
2

since B = 5×A and the gauge is ∇·A = 0 The energy related to the vector potential

is
1 
= (p − A)2
2 

~ 2
4 = A · 5+ A2
  2 

 2
= − B·L+  2 2 sin2 
2  8 2

242
CHAPTER 5 – MANUSCRIPT

where L = r × p,  is the angle between r and H, and e is a positive quantity. Since

we wish to work only to the first order in B we can put


 =− B · L
2 

The energy eigenfunctions of the unperturbed hydrogen atom are usually chosen to

be eigenstates of  with the eigenvalues ~ We choose the magnetic field is in the

z-direction, then the first order perturbation is

(0) 
1 = h|  |i = ~
2

So the degeneracy of the 2 + 1 states is removed in first order.

(1). s-wave:  = 0, and  = 0;

(2). p-wave:  = 1, and  = −1, 0, and 1.

We may add the spin magentic-momentum interaction


− ·  = − S·B
 

So the energy caused by a magenetic field is


 = − B · (L + 2S)
2 

B 5.4.4 Spin-Orbit interaction

Assume a central potential:  ()

243
CHAPTER 5 – MANUSCRIPT

A valence electron experience an electric field,

1
E = − ∇ ()

Whenever a moving charge is subjected to an electric field, it "feels" an effective mag-

netic field,
v
  = − ×E

The interaction energy related to the electron spin and the effective magnetic field

 = − · B 
µ ¶
S p 1
= · × ∇ ()
    −
µ ¶
S p 1 r  ()
= · ×
    −  

1 1  ()
= L·S
2 2  

Consider the relativistic quantum correction, the spin-orbit coupling is only half of the

above term.

Weak field:  is dominant.

Strong field: the Zeeman splitting is dominant.

B 5.4.5 Example: First Order Stark Effect in Hydrogen

The change in the energy levels of an atom caused by a uniform external electric field

of strength E is called the Stark effect. The perturbation  is now the extra energy of

244
CHAPTER 5 – MANUSCRIPT

the nucleus and electron in the external field and is ready shown to be

 =  cos 

where the polar axis and E are in the direction of positive  and  is again a positive

quantity. In a hydrogen atom the Coulomb potential is rotationally invariant. The wave

function for any spherically symmetric ponetial energy, when expressed in spherical

harmonics, have even or odd parity according as the azimuthal quantum number  is

even or odd. Since the perturbation is odd parity with respect to inversion, the ground

state has even parity and has no first-order Stark effect. The first excited state ( = 2)

of hydrogen is four fold degenerate; the quantum numbers  and  have the value

(0 0), (1 −1), (1 0), (1 1). Since the perturbation  commutes with the z-component

of angular momentum,

[   ] = 0

we obtain

( −  ) h  |  |0   i = 0

245
CHAPTER 5 – MANUSCRIPT

The nonvanishing matrix elements of  are

1 = h1 0|  |0 0i = h0 0|  |1 0i

Z
=  r∗210 (r) cos 200 ( r)

Z +1 Z ∞
 2  −0
= cos  cos  (2 − )
1640 −1 0 0

= −30 

The first order energy correction is determined by the secular equation


¯ ¯
¯ ¯
¯ −  0 0 ¯
¯ 1 1 ¯
¯ ¯
¯ ¯
¯  0 ¯¯
¯ 1 −1 0
¯ ¯ = 0
¯ ¯
¯ 0 0 −1 0 ¯¯
¯
¯ ¯
¯ ¯
¯ ¯
¯ 0 0 0 −1 ¯

The four roots of the secular equation are






⎪ − |1 |






⎨ 0
1 = 



⎪ 0






⎩ + |1 |

so that half of the four fold degeneracy is removed in the first order.

246
CHAPTER 5 – MANUSCRIPT

A 5.5 The Wentzel-Kramers-Brillouin (WKB) approx-

imation

The Wentzel-Kramers-Brillouin approxiamtion is for the approxiamte treatment of the

Shrodinger wave equation that shows its connection with the Bohr-Sommerfeld quan-

tization rules. It is based on an expansion of the wavefunction in powers of ~, which,

although of a semiconvergent or asymptotic character, is nevertheless also useful for

the approximate solution of quantum-mechanical problems in appropriate cases.

A 5.6 Time-dependent Problem: Interacting Picture

and Two-State Problem

When the Hamiltonian depends on the time, there are no stationary solution of the

Schrödinger equation. Thus our identification of a bound state with a discrete energy

level and stationary eigenket must be modified.

B 5.6.1 Time-dependent Potential and Interacting Picture

We consider a Hamiltonian  such that is can be split into two parts,

 = 0 +  ()

247
CHAPTER 5 – MANUSCRIPT

where 0 does not contain time explicitly. The problem of 0 or  () = 0 is assumed

to be solved completely, i.e.,

0 |i =  |i 

Assume  () = 0 at  = 0. The state ket is given by

X
|i =  (0) |i 

In the system of 0 , the state ket at a later time will evolve into

X
|()i =  (0)− ~ |i 

Our question is how the state ket of  (not 0 ) changes as time goes on.

Now we must work with the time-dependent Schrodinger equation


~ |()i = (0 +  ()) |()i  (5.13)


Our procedure consists in expressing

X
|()i =  ()− ~ |i  (5.14)

Substituting Eq. (5.14) into Eq. (5.13),


" # " #
 X X
~  ()− ~ |i = (0 +  ())  ()− ~ |i 
  

248
CHAPTER 5 – MANUSCRIPT

X∙  ¸ X ∙
 − ~
¸
− ~
 = ~  ()  |i +  () ~  |i

 


X∙  ¸ X
= ~  () − ~ |i +  () − ~ |i

 
" #
X
 = (0 +  ())  ()− ~ |i

X X
=  () − ~ |i +  ()  ()− ~ |i
 

we obtain
X∙  ¸ X
~  () − ~ |i =  ()  ()− ~ |i

 
 X X
~  = h|  |i   ~ = h|  |i  (5.15)
  

where the Bohr frequency is defined as

 − 
  = 
~

This equation is exactly equivalent to the original problem.

Interaction picture

Two kinds of changes have been made in going from Eq. (5.13). First, we

have changed the representation from specified in terms of the coordinates to being

specified in terms of the unperturbed energy eigenvalues. Second, we have changed

from the Schrodinger to the interaction picture.

h|  |i = h| 0 ~  −0 ~ |i

= h|  |i − ~ 

249
CHAPTER 5 – MANUSCRIPT

The equations in the interaction picture


~ | 0  i =  () | 0  i



~  = [  0 ]


B 5.6.2 Time-dependent Two-State Problem

Exact soluble problem of time-dependent potential are rather rare. However, a two

state problem with sinusoidal oscillating potential can be solved exactly.

The Problem is defined by

0 = 1 |1i h1| + 2 |2i h2| (2  1 )

 () =  |1i h2| + − |2i h1| 

In the interaction picture, the Eq. (5.15) can be written as


⎛ ⎞ ⎛ ⎞⎛ ⎞
 ⎜ 1 () ⎟ ⎜ 0 4 ⎟ ⎜ 1 () ⎟
~ ⎝ ⎜ ⎟ ⎜ ⎟⎜ ⎟
 ⎠=⎝ ⎠⎝ ⎠
2 () −4 0 2 ()

The equation is reduced to


~ 1 () = 4 2 () (5.16)



~ 2 () = −4 1 () (5.17)


where 4 =  + 1 − 2 

250
CHAPTER 5 – MANUSCRIPT

How to solve this equation?

From Eq. (5.16) we have


2 () = (4 )−1 ~ 1 ()


Substituting 2 into Eq.(5.17),


∙ ¸
 4 −1 
~ ( ) ~ 1 () = −4 1 ()
 
2  2
− 1 + ∆ 1 = 1
2  ~2

Assume 1 =  exp[] we have

2 − ∆ =  2 ~2 ;

1³ p ´
 = ∆ ± ∆2 +  2 ~2
2

So the general solutions are

 
1 = +  2 (4+ ) + −  2 (4− )

~  ~ 
2 = − (4 +   )+ − 2 (4− ) − (4 −   )− − 2 (4+ )
2 2

 are two coefficients and determined by the initial conditions and


∙ ¸12
4 2
 = 2
+ ( + 12 )2 
~

If initially, at  = 0, only the lower is populated so that

1 (0) = 1 and 2 (0) = 0

251
CHAPTER 5 – MANUSCRIPT

then the probability for being found in each of the two states is given by
µ ¶
2 2 1
|2 ()| = 2 2 2
sin2   
 + ( + 12 ) ~ 4 2

|1 ()|2 = 1 − |2 ()|2

This is Rabi’s formula, after I.I. Rabi.

Spin Magnetic Resonance

The two-state problem has many physical applications: Spin magnetic reso-

nance, maser etc. Consider a spin 1/2 system subjected to a t-independent uniform

magnetic field in the z-direction and, in addition, a t-dependent magnetic field rotating

in the xy plane.

 = 0 ̂ + 1 (̂ cos  + ̂ sin )

with 0 and 1 constant.

We treat the effect of the uniform t-independent field as 0 and the effect of

the rotating field as  ()

~0
0 = − (|+i h+| − |−i h−|)
2 

~1
 () = − [cos  (|+i h−| + |−i h+|) + sin  (− |+i h−| +  |−i h+|)]
2 

~1 £ − ¤
= −  |+i h−| + + |−i h+|
2 

252
CHAPTER 5 – MANUSCRIPT

Figure 5.1:

Figure 5.2:

253
CHAPTER 5 – MANUSCRIPT

Four Nobel Prize winners who took advantage of resonance in two-level sys-

tems

• Rabi (1944): on molecular beams and nuclear magnetic resonance;

• Bloch and Purcell (1952): on B field in atomic nuclei and nuclear magnetic mo-

ments;

• Townes, Basov, and Prochorov (1964): on masers, lasers, and quantum optics;

• Kastler (1966): on optical pumping

A 5.7 Time-dependent Perturbation Problem

B 5.7.1 Perturbation Theory

We now return to Eq. (5.3), replace  by  , and express the  as the power series

in 
2 (2)
 = (0) (1)
 +  +   + · · · 

The substitution yields the set of equations

 (0)
 = 0
 

 (+1) X ()
~  = h|  |i  + ~ 
 

254
CHAPTER 5 – MANUSCRIPT

The zero-order coefficient are constant in time. We shall assume that all except one of
(0)
the  , say  =  are zero, so that the system is initially in a definite unperturbed

energy state. Thus the solutions to the zero order coefficients are

(0)
 = h|i =   

Integration of the first-order equation gives


Z 
(1)

−1
= (~) h|  () |i  ~  (5.18)
−∞

(1)
The integral constant has been chosen in such a way that  vanish at  = 0  To first

order in the perturbation, the transition probability corresponding to the transition

 → , (that is, the probability that the system, initially in the state , be found at

time  in the state  6= ) is therefore

¯ ¯2
 = ¯(1)

¯   6= 

It is also worth noting that for   0 the coefficient  of the state  is given to first

order in the perturbation by

 () ≈ (0) (1)


 () +  ()

Z 
−1
≈ 1 + (~) h|  () |i 
−∞
∙ Z ¸
 
≈ exp − h|  () |i 
~ −∞

So the principal effect of the perturbation on the initial state is to change the phase.

255
CHAPTER 5 – MANUSCRIPT

B 5.7.2 Time-independent perturbation

The result takes particular simple form if the perturbation  is independent of time,

except for being switched on suddenly at a given time.





⎨ 0 for   0
 () =


⎩  for   0

We then obtain


(1)
 = − h|  |i 
~

1 −  ~
(1)
 = h|  |i


− 2~ −  2~


= h|  |i  2~


sin( 2~)
= −2 2~ h|  |i


Hence

Z 
− ~ −1 − ~
  |i = (1 + (~) h|  () |i ) |i
−∞

( + )~


≈  |i

This is in agreement with the result using the stationary perturbation theory. The first

order transition probability from  to  is given by


µ ¶2
(1) 2 sin ( 2~)
 = 4 h|  |i


256
CHAPTER 5 – MANUSCRIPT

It is worth noting that this is valid for nondegenerate case.

In the case of  = 0:

sin ( 2~) 


=
 2~

the probability is
1
(1)
 = h|  |i2 2
~2

B 5.7.3 Harmonic perturbation

Eq. (5.18) takes a particularly simple form if the perturbation  depends harmonically

on the time except for being turned on at one time and off at a latter time. We shall

call these two times 0 and 0 , respectively, and assume that we can write

h|  () |i = h|  |i sin 

where h|  |i is time independent. The first order amplitude at any time  at or after

0
µ ¶
h|  |i ( +)0 − 1 ( −)0 − 1
(1)
 =− −
~   +   − 

The structure of Eq. (??) suggests that the amplitude is appreciable only when the

denominator of one or the other of the two terms is particularly zero. The first term

is important when  ≈ −  and the second term is important when  ≈ +   For

the present, we specialize to a situation in which the initial state  is a discrete bound

257
CHAPTER 5 – MANUSCRIPT

Figure 5.3:

state and the final state  is one of a continuous set of dissociated states. The first order

probability of finding the system in the state |i after the perturbation is removed is

¯ (1) ¯ 2 2
¯ (  0 )¯2 = 4 |h|  |i| × sin [(  − ) 0 2]  (5.19)
~2 (  − )2

B 5.7.4 The Golden Rule

The factor sin2 [(  − ) 0 2]  (  − )2 is plotted in Figure. The height of the

main peak increases in proportional to 20 . Thus if there is a group of states |i that

have energy nearly equal to  + ~, and for which |h|  |i| is roughly independent

of , the probability of finding a system in one or another of these states is proportional

to 0 .

258
CHAPTER 5 – MANUSCRIPT

X ¯ ¯2 Z
¯ ¯2
(1)
 = ¯ (1) ¯
 ⇒  ( ) ¯(1)

¯
 ≈

where () is the number of final states with energies between  and  +  .

the concept of an energy density () of final state is sensible, since we are considering

the case in which the transition is to one or another of a continuous set of dissociated

states.

Z Z
¯ ¯2 ( −  ) | |2
 ( ) ¯(1)

¯ =4  ( ) sin 2
2~ | −  |2

Follow from
1 sin2 
lim = ()
→+∞  2
( − )
sin2 2~ 
→ ( −  )
| −  |2 2~

for a large t (→ +∞).


Z Z
¯ ¯2 
 ( ) ¯(1)

¯ = 4  ( ) | |2 ( −  )
2~

2 ­ ®
= | |2 ( ≈  )
~

The transition probability per unit time is is approximately written as

2
= () |h|  |i|2 
~

This formula, first obtained by P. A. Dirac, was later dubbed by E. Fermi “The Golden

Rule” of perturbation theory.

259
CN Chapter 6

CT Collision Theory

Problems for which the energy eigenvalues are continuously distributed usually arise in

connection with the collision of a particle with a force field. In a collision problem the

energy is specified in advance, the behavior of the wave function at great distance is

found in terms of it. This asymptotic behaviors can then be related to the amount of

scattering of the particle by the force. There are so few systems of physical interest for

which exact solutions can be found that approximation methods play an important part

in applications of the theory. Various methods that are useful in scattering problems

are considered in this chapter.

260
CHAPTER 6 – MANUSCRIPT

Figure 6.1:

A 6.1 Collisions in one- and three-dimensions

B 6.1.1 One-dimensional square potential barriers

We consider first the one-dimensional collision of a particle with the square potential

barrier. In this problem we are interested in a particle that approaches from the region

of negative x and is reflected or transmitted by the carrier. In the corresponding

classical problem, the particle is always reflected if its energy is less than that of the

top of the barrier, and it is always transmitted if the energy is great. We shall see that,

in the quantum problem, both reflection and transmission occur with finite probability

for most energies of the particle.

Asymptotic behaviors

We are interested in representing a particle that approaches from the left with

energy   0 and may be turned back by the potential barrier or penetrate through

261
CHAPTER 6 – MANUSCRIPT

it. Thus the asymptotic behavior in the region where  () = 0 is as follows: for   0

we want the wave function to represent a particle moving to the left (reflected particle)

as well as to the right (incident particle): for   , we want the wave function to

represent only a particle moving to the right (transmitted particle).

The wave function in the region where  () = 0 is

~2  2
−  = 
2 2

Our asymptotic solution are

() =  + −    0

() =   

where
µ ¶12
2
 = ~ =
~2

Normalization

Recall the definition of the possibility current density

~
= [Ψ∗ ∇Ψ − (∇Ψ∗ ) Ψ] 
2

The possibility current densities at   0 and    are

~ ¡ 2 ¢
() = || − ||2    0

~
() = ||2    

262
CHAPTER 6 – MANUSCRIPT

The two currents are independent of the position and should be equal due to the

conversation of the practice.

Scattering coefficients

The character of the solution inside the potential barrier depends on whether

 is greater or less than the potential barrier 0 .

C The case   0

The solution in the potential barrier is

() =   + −  0    

where
µ ¶12
2 ( − 0 )
= 
~2

The continuity of  and  at  = 0 and  =  required by the boundary condition

provides four relations between the five coefficients. At  = 0

+ =  +

( − ) = ( − )

At  = ,

  + − = 

¡ ¢
   + − = 

263
CHAPTER 6 – MANUSCRIPT

We can eliminate  and  and solve for the ratios  and 

 (2 − 2 ) (1 − 2 )
=
 ( + )2 − ( − )2 2
¡ ¢
 4 1 − (−)
=
 ( + )2 − ( − )2 2

The reflection coefficient is


¯ ¯2 ∙ ¸−1
¯ ¯ 4( −  )
¯ ¯ = 1+ 0

¯¯ 02 sin2 

The transmission coefficient is


¯ ¯2 ∙ ¸−1
¯ ¯ 2 2
¯ ¯ = 1 + 0 sin  
¯¯ 4( − 0 )

C The case   0

The reflection and transmission coefficients for 0    0 are obtained by replace 

by  in the wave function in the potential barrier.


µ ¶12
2 (0 − )
= 
~2

The reflection coefficient is


¯ ¯2 ∙ ¸−1
¯ ¯ 4( − )
¯ ¯ = 1+ 0

¯¯ 02 sinh2 

The transmission coefficient is


¯ ¯2 ∙ ¸−1
¯ ¯ 2 2
¯ ¯ = 1 + 0 sinh  
¯¯ 4(0 − )

264
CHAPTER 6 – MANUSCRIPT

Figure 6.2: The transmission coefficient of a square barrier.

B 6.1.2 Datta-Das spin field transistor

In 1990, Datta and Das proposed....

A 6.2 Collision in three dimensions

Scattering cross section

The angular distribution of particles scattered by a fixed center of force or by

other particles is conveniently described in terms of a scattering cross section. Suppose

that we bombard a group of  particles or scattering centers with a parallel flux of 

particles per unit time and count the number of incident particles that emerge per unit

time in a small solid angle 4 0 (?) centered about a direction that has polar angles

0 and 0 with respect to the bombarding direction as polar axis. This number will

265
CHAPTER 6 – MANUSCRIPT

be proportional to ,  and 4 0 provided that the flux is small enough so that there

is no interference between bombarded particles and no appreciable diminution of the

bombarded particles by their recoil out of the target region, and provided also that the

bombarded particles are far enough apart so that each collision process involves only

one of them.

The different scattering cross section  0 (0 0 ): the number of incident parti-

cles that emerge per unit time in 4 0 can be written

  0 (0 0 )4 0 

The total scattering cross section:


Z
0 =  0 (0 0 ) 0 

For a collision of a particle with a fixed scattering center, the definition of the differential

scattering cross section is equally valid in the laboratory and center-of-mass coordinate

systems, since a scattering center that is fixed has an infinite effective mass and so

that the center of mass of the system does not move. For a collision between two

particles of finite mass, however, the differential cross section applies in general only

to the laboratory coordinate system and to the observation of the scattered incident

particle. It does not describe in the observation of the recoil bombarded particle in the

laboratory system, although it is of course possible to obtain a differential cross section

for the recoil particle from 0 (0 0 )

266
CHAPTER 6 – MANUSCRIPT

Figure 6.3:

267
CHAPTER 6 – MANUSCRIPT

Relation between angles in the laboratory and center-of-mass sys-

tems

The relation between the differential cross section and angles in the laboratory

system and in the center-of-mass system can be found by translating the laboratory

system in the direction of the incident particle with sufficient speed to bring the center

of mass to rest. Assume that a particle of mass 1 and initial speed  strikes a particle

of mass 2 that is initially at rest. The center of mass moves with the speed

1
0 = 
1 + 2

00
Thus in the center-of-mass system the speeds of two particles are  00 =  − 0 and − .

If the collision is elastic, they evidently recede from the center of mass after the collision

with the same speed. Thus 0 0 and   are related by

00 cos  +  0 = 1 cos 0

00 sin  = 1 sin 0

 = 0

From these Eqs., we obtain

sin 
tan 0 =
 + cos 

0 1
 = 00
=
 2

268
CHAPTER 6 – MANUSCRIPT

Relation between cross sections

The relation between the cross section in the laboratory and center-of-mass

coordinate systems can be obtained from the definitions, which imply that the same

number of particles are scattered into differential solid angle 0 about 0  0 as are

scattered into  about  

0 (0  0 ) sin 0 0 0 = ( ) sin 

With the help of the relation 0  0 and  , we obtain

12
(1 +  2 + 2 cos )
 0 (0  0 ) = ( )
|1 +  cos |

It should be noted that the total cross section is the same for both laboratory and

center-of-mass systems and also for both the outgoing particles, since the total number

of collision that takes place is independent of the mode of description of the process.
1
Dependence on  = 2

For  = 1,
sin  
tan 0 = = tan
 + cos  2

Thus 0 = 2 and varies from 0 to 2 as  varies from 0 to ; in this case

 0 (0  0 ) = 4 cos 0 (20  0 )

and no particle appear in the backward hemisphere in the laboratory system.

For   1, 0 increase from 0 to  as  increase from 0 to .

269
CHAPTER 6 – MANUSCRIPT

For   1, 0 first increase from 0 to a maximum value cos−1 (−1), which

is less than 2, as  increases from 0 to cos−1 (−1); 0 then decreases to 0 as 

increases further to . In this case 0 (0  0 ) is usually infinite at the maximum value

of 0 , although this singularity gives a finite contribution to the total cross section; no

particles appear beyond the maximum 0 in the laboratory system.

A 6.3 Scattering by Spherically Symmetric Potentials

In this section, we assume that the potential  is a function only of , and we find

the connection between the solution separated in spherical polar coordinate and the

asymptotic form.

Asymptotic behaviors

The differential scattering cross section in the center-of-mass coordinate sys-

tem can be found from


µ 2 ¶
~ 2
− ∇ +   = 
2

where  = 1 2 (1 + 2 ) The scattering is determined by the asymptotic form of

 in the region where  = 0

1
 = [ +  ( ) ] (6.1)

where  = ~ and  is the speed of the incident particle. The probability current

270
CHAPTER 6 – MANUSCRIPT

density J is given by
~ £ † ¡ ¢ ¤
J() = Ψ (∇Ψ) − ∇Ψ† Ψ
2

The gradient operator can be expressed in polar coordinates as

 1  1 
∇= r̂+ θ̂ + φ̂
    sin  

Since the second and third terms are small when  is large, the current for large  is in

the radial direction and


~
 = | ( )|2 2

The number of scattered particles entering the solid angle (or detector) per unit time

is

 =  2 

~
= | ( )|2 

From the definition of the cross section, it follows

 ( ) = |( )|2 

Differential cross section

The asymptotic behavior of the wave function determines the differential

scattering cross section but cannot itself be found without solving the wave equation

271
CHAPTER 6 – MANUSCRIPT

throughout all space. In the spherical polar coordinate,


⎡ ⎤
1 
¡ 
¢
~2 ⎢ 2 
2  ⎥
− ⎣ ⎢ ⎥  = ( −  )
2 ³ ¡ ¢ ´ ⎦
1   2
+ 2 sin  
sin  
+ 2

The radial and angular parts of the solution can be separated by taking

(  ) = () ( )

where

 ( ) =   (cos )

is the spherical harmonic. The problem now possesses symmetry about the polar axis,

so that   , and  are independent of the angle . That is, we just consider the

case of  = 0. In the case the =0 is reduced to the Legendre polynomials. Thus,

the general solution can then be written as a sum of products of radial function and

Legendre polynomials,

X

 = (2 + 1)  () (cos )
=0

X

= (2 + 1) −1  () (cos )
=0

where  is the Legendre polynomial of order , and  satisfies


∙ ¸
2  2 ( + 1)
+  − () −  = 0
2  2

272
CHAPTER 6 – MANUSCRIPT

where
µ ¶12
2
 =
~2

2 ()
() = → 0 as  → +∞
~2
µ ¶
1  2  1 2
 = 
2    2

In the limit of  → +∞ the solution  is one of the form

 ∝ ± →  sin( +  0 )

In some specialized problems, () can be neglected for  greater than some distance

, and  may be small enough so that the  term in not negligible. The general form

for  () has the form

 () =  [cos    () − sin   ()]

 
−→ sin( −  +   )  → ∞ (6.2)
 2

Here  is the spherical Bessel function;


µ ¶12

 () = +12 ()
2

1 +1
→ cos( − )  → ∞
 2


→  →0
(2 + 1)!!

273
CHAPTER 6 – MANUSCRIPT

 is the spherical Neumann function;


µ ¶12
+1 
 () = (−1) −−12 ()
2

1 +1
→ sin( − )  → ∞
 2

(2 − 1)!!
→ −   → 0
+1

To identify the form of  , we require an expansion of  cos  in Legendre polynomials;

X

 cos  = (2 + 1)  () (cos )
=0

Comparison of Eq.(6.1) with the general form  () gives

 + −1 ( )

X

= (2 + 1)  () (cos ) + −1  ( )
=0

X∞

= (2 + 1)  ()−1 sin( − +   ) (cos )
=0
2

 are the coefficient for different , and should take the form

 =  

Thus it gives
1 X

 () = (2 + 1)(2 − 1) (cos )
2 =0

274
CHAPTER 6 – MANUSCRIPT

Thus the differential cross section is

() = | ()|2
¯∞ ¯2
1 ¯X ¯
¯ 2 ¯
= 2 ¯ (2 + 1)( − 1) (cos )¯
 ¯ ¯
=0

Total elastic cross section

The total elastic cross section is given by

Z 
 = 2 () sin 
0

4 X

= (2 + 1) sin2   
2 =0

Here we make use of the orthogonality property of the Legendre polynomials

Z +1
2
 cos  (cos )0 (cos ) = 0 
−1 2 + 1

The total cross section can also be related to (0). it follows from the generating

function for the Legendre polynomials that  (1) = 1 for  = 0 and all , so that

1 X

¡ ¢
 (0) = (2 + 1) 2 − 1 
2 =0

Comparison with the total cross section shows

2 4
= [ (0) −  ∗ (0)] = Im (0)
 

This relation is known as the optical theorem.

275
CHAPTER 6 – MANUSCRIPT

The physical interpretation of the optical theorem is as follows: In order for

scattering to take place, particles must be removed in an amount proportional to 

from the incident beam, so that its intensity is smaller behind the scattering region

( ≈ 0) than in front of it. This can occur only by the interference between two terms

in the asymptotic expression of the wave function.

Phase shifts

The angle   is called the phase shift of the l partial wave, since according to

Eq.(6.2) it is the difference in phase between the asymptotic forms of the actual radial

function  () and the radial function  () in the absence of the scattering potential.

the phase shifts completely determine the scattering, and the scattering cross section

vanishes when each of   is 0 or 180 .

A 6.4 Applications

B 6.4.1 Scattering by a square well

In general, for a given potential, phase shifts are calculated from a numerical solution

of the radial equations. In a few cases an analytical solution is possible, in particular

for scattering by an attractive square well, for which the reduced potential () is



⎨ −0 ( 0)   ;
() =


⎩ 0   

276
CHAPTER 6 – MANUSCRIPT

Inside the well (  ) the radial equation is


∙ ¸
2 2  ( + 1) 2
+ − +   () = 0
2   2

where we set  2 = 2 + 0  The regular solution of this equation is

 () =   ()   

In the exterior region the radial equation is


∙ ¸
2 2  ( + 1) 2
+ − +   () = 0
2   2

The general solution is

 () =  [ () − tan    ()]     (6.3)

The solutions for    and    can be joined smoothly at  =  by requiring that

 () =  ();


¯ ¯
 () ¯¯  () ¯¯
¯ =
 =  ¯=

Eliminating the normalization constants gives

0 () () −  ()0 ()


tan   () = 
0 () () −  ()0 ()

At low energies, the scattering is dominated by the  = 0 wave. Using the formula

sin 
0 () =

cos 
0 () = −

277
CHAPTER 6 – MANUSCRIPT

we have
 tan  −  tan 
tan  0 () = 
 +  tan  tan 

In the low-energy limit,  ¿ 1 tan  ≈ , and


µ ¶
tan 
 0 () = − 1 −


Then the  = 0 partial cross section is

4 2
=0 = sin  0
2
µ ¶2
2 tan 
≈ 4 1 −


B 6.4.2 Scattering by a hard-sphere potential

Another simple, but interesting example is the hard-sphere potential





⎨ +∞   ;
() =


⎩ 0   

Since the scattered particle cannot penetrate into the region   , the wave function in

the exterior region must vanish at  = . Since the scattered particle cannot penetrate

into the region    the wave function in the exterior region, given by Eq.(6.3), must

vanish at  = , from which


 ()
tan   () = 
 ()

278
CHAPTER 6 – MANUSCRIPT

In the low-energy limit,  ¿ 1


∙ ¸
() (2 − 1)!!
tan   () =  −
(2 + 1)!! ()+1

()2+1
= − 
(2 + 1)!!(2 − 1)!!

Hence |tan   ()| quickly decreases as  increases. As a result, the low-energy scattering

is always dominated by the s-wave. Since 0 () = sin  and 0 () = − cos , we

have

 0 = −

So the cross section is

 0 = 42 

At high energies (  0), we find


 = − 
2

so that

4 X

max
2
  ≈ 2
(2 + 1) sin2 ( − )
 =0 2

≈ 22 

It is interesting to compare this result with that obtained from classical mechanics.

279
CHAPTER 6 – MANUSCRIPT

B 6.4.3 Identical Particles and Scattering

Collisions between identical particles are particularly interesting as a direct illustration

of the fundamental differences between classical and quantum mechanics. In the polar

coordinate, the two directions are opposite, ( ) and ( −   + )

The wave functions


 () = · + ()


− () = −· + ( − )

Symmetric or anti-symmetric:

Ψ =  () + − ()

Ψ =  () − − ()

Classical case

The differential cross-section is

( ) = | ( )|2 + |( −   + )|2

Scattering of two identical spinless bosons

The differential cross-section is

( ) = | ( ) + ( −   + )|2

Scattering of two identical spin-1/2 fermions

280
CHAPTER 6 – MANUSCRIPT

For spin 1/2 fermions, the wave functions contains two parts, spin and spatial

part. For spins, two spins can form one spin singlet  = 0 (antisymmetric), and one

spin triplet  = 1 (symmetric). The total wave function should be antisymmetric. Thus

one part of the wave function is symmetric and another part must be antisymmetric.

The differential cross-section is

1
( ) = |( ) +  ( −   + )|2
4

3
+ | ( ) −  ( −   + )|2
4

A 6.5 Applications to interactions with the classical

field

We apply the formalism of time dependent perturbation theory to the interactions of

atomic electrons with the classical radiation field.

B 6.5.1 Absorption and stimulation emission

Consider an electron in an electromagnetic field,

1 
 = (p − A)2 + ()
2 

1 2 
= p + () − A·p
2  

281
CHAPTER 6 – MANUSCRIPT

where the term A2 is omitted and the symmetric gauge is taken, ∇ · A = 0 For a

monochromatic field of a plane wave,


A = 20 ̂ cos( n · x − )

where ̂ · n = 0 We write

 
A = 0 ̂[exp( n · x − ) + exp(− n · x + )
 

The perturbation term is


 = − A·p
 

0 £   ¤
= − ε̂ · p   n·x− + −  n·x+
 

where the first term is responsible for absorption, and the second term is for stimulated

emission.

In the case of absorption,

0  n·x−
† = − (  ε̂ · p)
 

and the transition rate is

2 2 20 ¯¯ 
n·x−
¯2
¯ ( −  − ~)
→ = h|   ε̂ · p |i
~ 2 2

The absorption cross section is

(Energy/unit time)absorbed by the atom (i → n)


  =
Energy flux of the radiation field

282
CHAPTER 6 – MANUSCRIPT

The energy flux of an electromgnatic field is given by the Poynting vector,


= (E × H)
4
1 2
with E = − 1 A and B = ∇ × A 2 
|0 |2

~→
  = 1 2
2 
|0 |2

4 2 ~ 2 ¯¯   n·x−
¯2
¯ ( −  − ~)
= ( ) h|  ε̂ · p |i
2  ~

B 6.5.2 Electric dipole approximation

Assume ~ is of order of the atomic energy levels, so

2 2
~ ∼ '
0  

This leads to
 ~ 
=∼ 2
= 137 
  

Then   ∼ 137 ¿ 1 for light atams. In this case, it becomes a good approxi-

mations
 
  n·x = 1 +  n · x + · · ·


h|   n·x− ε̂ · p |i → ε̂ · h| p |i

Using the Heisenberg equation of motion,

~
[ 0 ] =

283
CHAPTER 6 – MANUSCRIPT


h| p |i = h| [x 0 ] |i =
~

h|   n·x− ε̂ · p |i →  ε̂ · h| x |i

This is called the electric dipole approximation. In this approximation,

  = 4 2   |h| x |i|2 ( −   )

B 6.5.3 Photoelectric effect

The photoelectric effectis the ejection of an electro when the atom is placed in the

radiation field. The basic process is considered to be the transition from an atomic

bound state to a continuum state (  0). Therefore, |i the ket for an atomic state,

say, the ground state of hydrogen atom, while |i is the ket for a continuum state say

a plane wave state |k i 

The basic task in this calculation is the to calculate the number  of the

final state near the energy  and  + , or the density of state


() = lim 
→0 

Consider a cubic box of side L3 . The plane wave has the form,

k ·x
hx|k i =
32

where

 = 2 

284
CHAPTER 6 – MANUSCRIPT

The energy
~2 2 ~2 (2)2 2
= = 2
( + 2 + 2 )
2 2 

 = 2 Ω


= 2 Ω



() = 2
Ω
µ ¶3 µ ¶3
   
= 2
= (2 ~2 )12
2 ~ 2 ~2

So the differential cross section becomes


µ ¶3
4 2 ~ ¯¯  ¯2  
 = 2
hk |   n·x− ε̂ · p |i¯ 
  2 ~2

The initial state is taken to be a ground state of hydrogen-like wave function except

that the Bohr radius 0 is replaced by 0 .


Z
 
n·x−
hk |   ε̂ · p |i = x hk | i   n·x− ε̂ · (−~∇) h |i

Z " µ ¶32 #
−k ·x  n·x− 
= ε̂ · x 32   (−~∇) −0
 0

In this integral, notice that ε̂ · ∇  n·x− = 0
Z " µ ¶32 #

n·x− −~∇−k ·x  n·x− −0 
hk |   ε̂ · p |i = −ε̂ · x  
32 0

Z " µ ¶32 #
~ (ε̂ · k ) −k ·x  n·x− −0 
= x  
32 0

285
CHAPTER 6 – MANUSCRIPT

For th hydrogen atom,

 2 (ε̂ ·  )2  5 1
= 32 
Ω  0 [(0 )2 +  2 ]4
5

where q = k − ()n

A 6.6 Approximate Collision Theory

See Sakurai’s book: section 7.1, 7.2, and 7.3

B 6.6.1 The Lippman-Schwinger Equation

Let us begin with the time-independent formulation of scattering processes. The Hamil-

tonian is written as
2
 = 0 +  = + 
2

0 stands for the kninetic energy. In the collision theory, the source of incident particles

and the detection location of the scattered particles are very far away from the scattering

center. 0 can describe the behaviors of incident and reflected particle very well as the

potential approaches to zero when the distance is much larger than the interaction

distance. We have two Schrodinger equations in the absence and presence of  :

0 |i =  |i

(0 +  ) |Ψi =  |Ψi

286
CHAPTER 6 – MANUSCRIPT

When  → 0 |Ψi → |i  We assume the two equations have the same energy eigen-

value. Combining the two equations

( − 0 ) |Ψi =  |Ψi + ( − 0 ) |i

If ( − 0 )−1 exist, we have

|Ψi = ( − 0 )−1  |Ψi + |i

Unfortunately, the operator ( − 0 )−1 is singular. To avoid the difficulty, we introduce

an infinitesmal complex  ( → 0+ ) such that  →  +  In this way the equation

becomes

¯ (+) ® 1 ¯ ®
¯Ψ = |i +  ¯Ψ(+) (6.4)
 − 0 + 

In the position space, Eq.(6.4) is reduced to

Z
¯ ® 1 ¯ ®
hx ¯Ψ(+) = hx |i + x0 hx| |x0 i hx0 |  ¯Ψ(+)
 − 0 + 
Z
2 ¯ ®
= hx |i + 2 x0 + (x x0 ) hx0 |  ¯Ψ(+)
~

287
CHAPTER 6 – MANUSCRIPT

We evaluate the kernel of integral (x x0 ).

0 ~2 1
+ (x x ) ≡ hx| |x0 i
2  − 0 + 
Z
~2 1
= p hx| pi hp| |pi hp |x0 i
2  − 0 + 
Z
~2 1
= p hx| pi hp| 2
|pi hp |x0 i
2  −  2 + 
Z
~2 1
= p hx| pi 2
hp |x0 i
2  −  2 + 
Z
~2 p exp[p · (x − x0 )~]
=
2 (2~)3  − 2 2 + 

where we have used

hx| pi = (2~)−32 exp[p · x~] (6.5)

hp| xi = (2~)−32 exp[−p · x~]

Take  = ~2  2 2 and  = ~

Z
0 q exp[q · (x − x0 )]
+ (x x ) =
(2)3 2 −  2 + 
Z ∞ 2 Z 2 Z 
  exp[ |x − x0 | cos ]
=  sin 
0 (2)3 0 0 2 −  2 + 
Z +∞
 exp[ |x − x0 |] − exp[− |x − x0 |]
= 
4 2 |x − x0 | 0  2 −  2 + 

1 exp[ | − 0 |]
= −
4 | − 0 |

288
CHAPTER 6 – MANUSCRIPT

Figure 6.4:

In the last step we used the residue theorem:

I

= 2
 − 0
I X
() = 2 −1 = 2 × (sum of enclosed residues).

The two pole points are



 = ± 2 +  = ±( + )

The equation becomes

Z +|x−x0 |
¯ ® 2 0
¯ ®
hx ¯Ψ(+) = hx |i − 2  hx0 |  ¯Ψ(+) 
~ |x − x0 |

Assume that V is a local potential,

hx0 |  |x00 i =  (0 )(x0 −x00 )

Z +|x−x0 |
¯ ® 2 0
¯ ®
hx ¯Ψ(+) = hx |i − 2   (0 ) hx0 ¯Ψ(+) 
~ |x − x0 |

289
CHAPTER 6 – MANUSCRIPT

As the observation point is far away from the scatterer,

|x| =  À |x0 | = 0

|x − x0 | = (2 − 20 cos  + 02 )12

0
= (1 − cos  + )

From this approximation and taking k0 = x |x| we have

0
+|x−x |  −k0 ·r0
≈  
|x − x0 | 

So,
Z
¯ ® 2  ¯ ®
hx ¯Ψ(+) → hx |i − 2 x0−k ·r  (0 ) hx0 ¯Ψ(+) 
0 0

~ 

Comparing with the asymptotic form of the wave function,


∙ ¸
¯ ® 1 
hx ¯Ψ(±) → k•r
+ 0
(k  k)
(2)32 

We have

Z
0 4 2 exp[−k0 · x0 ] ®
 (k  k) = − 2 x0  (x0
) hx0
| Ψ(+)
~ (2)32

4 2 0 ¯¯ (+) ®
= − 2 hk |  Ψ
~

From this result we have the differential cross section

 2
= |(k0  k)|
Ω

290
CHAPTER 6 – MANUSCRIPT

B 6.6.2 The Born Approximation

The formula contains the unknown wave function. We have to introduce the approxi-

mation.

¯ (+) ® 1 ¯ ®
¯Ψ = |i +  ¯Ψ(+)
 − 0 + 
µ ¶
1 1 ¯ (+) ®
= |i +  |i + ¯
 Ψ
 − 0 +   − 0 + 

1
= |i +  |i + 
 − 0 + 
®
Assume the effect of the scattering is not very strong, we replace hx0 | Ψ(+) by hx0 | i

in the integral,
®
hx0 | Ψ(+) → hx0 | i = exp[−k · x0 ]

This is the first-order Born approximation. The amplitude is

Z
0 4 2 1 0 0
 (  ) = − 2 x0 (k−k )•x  (x0 )
~ (2)3

which is the Fourier transform of the potential except for a constant.

For a spherical symmetric potential, assume |k − k0 | =  = 2 sin 2 (|k| =

|k0 | as the energy is conserved.)

Z Z 2 Z 
1 2 2 2
 = −   sin  cos   ()
4 ~2 0 0 0
Z +∞
2
= − 2  () sin 
~ 0

291
CHAPTER 6 – MANUSCRIPT

B 6.6.3 The higher-order Born approximation

Define
¯ ®
 ¯Ψ(+) =  |i

we have

¯ ® 1
 |i =  ¯Ψ(+) =  |i +   |i
 − 0 + 

1
 =  + 
 − 0 + 

1
=  + 
 − 0 + 

1 1
+   + ···
 − 0 +   − 0 + 

The scattering amplitude can be written as

1 2
(k0  k) = − (2)3 hk| |k0 i
4 ~2

B 6.6.4 Optical Theorem



Im  ( = 0) =  
4
R R
where  () =  (k k) and   = 
Ω
Ω = |(k0  k)|2 Ω

G Proof.
1 2
( = 0) = ( ) = − (2)2 hk| |ki
4 ~2

292
CHAPTER 6 – MANUSCRIPT
­ ®
Im hk| |ki = Im k| |Ψ(+)
∙µ ¶ ¸
­ (+) ¯ ­ (+) ¯ 1 ¯ (+) ®
= Im Ψ ¯ − Ψ ¯  ¯Ψ
 − 0 − 

Use the relation,


1 1
= + ( − 0 )
 − 0 −   − 0

£ ­ ¯ ¯ ®¤
Im hk| |ki = Im − Ψ(+) ¯  ( − 0 ) ¯Ψ(+)

­ ¯ ¯ ®
= − Ψ(+) ¯  ( − 0 ) ¯Ψ(+)

Z
­ ¯ ¯ ®
= − k0 Ψ(+) ¯  |k0 i hk0 | ( − 0 ) |k0 i hk0 |  ¯Ψ(+)

Z
­ ¯ ¯ ® ~2 02
= − k0 Ψ(+) ¯  |k0 i hk0 |  ¯Ψ(+) ( − )
2
Z
­ ¯ ¯ ® ~2 02
= − 02 0 Ω0 Ψ(+) ¯  |k0 i hk0 |  ¯Ψ(+) ( − )
2
Z
­ ¯ ¯ ® 2 1
= − Ω0 Ψ(+) ¯  |k0 i hk0 |  ¯Ψ(+)  02 2 0
~ 2
p
At the last step we take 0 = 2~2 .

Z
0 ¯­ ¯ ¯2
Im hk| |ki = − 2 Ω0 ¯ Ψ(+) ¯  |k0 i¯
~

293
CHAPTER 6 – MANUSCRIPT

B 6.6.5 Application: from Yukawa potential to Coloumb poten-

tial

As an illustration of the Born approximation, we consider the Yukawa potential

−
 () = 0

Z
0  sin  1
(k  k) = − () = −0 2
  + 2

where

 = |k0 − k|

Z Z +∞
0
 () sin  = − sin 
 0
Z +∞
0  − −
= −
 0 2
µ ¶
0 1 1
= −
2  −   + 

0 
= 
2 + 2

So the differential cross section is


µ ¶2
20 1
() = 
~ [22 (1 − cos ) + 2 ]

When  = 0, the Yukawa potential is reduced to the Coloumb potential, 0  →  0 2 

We obtain
2 µ ¶2
(2 0 2 ) 1  0 2 1
() = =
~4
16 sin4
4 
2
 16 sin4 
2

294
CHAPTER 6 – MANUSCRIPT

This is the Rutherford scattering cross section. From Rutherford formula we conclude

that the charge-charge interaction in atomic scale obey Coulomb law, i.e., 1. Actually

this formula does not contain the Planck constant, and can be drived from the classical

theory.

295
CN Chapter 7

CT Selected Topics

A 7.1 Relativistic Quantum Mechanics

I shall briefly introduce the Dirac equation. Because of extensive interests in spintronics

and graphene, I shall focus on the spin-orbit coupling (electric manipulation of electron

spin) and graphene (2D massless Dirac particles)

B 7.1.1 The Schrodinger Equation

~ Ψ = Ψ

296
CHAPTER 7 – MANUSCRIPT

2
where  = 2
+.

 →  → ~

 → −~∇

[ ] = ~

The continuity equation:

  + ∇ ·  = 0

where

 = Ψ∗ Ψ

~ ∗
 = − (Ψ ∇Ψ − Ψ∇Ψ∗ ) = Re(Ψ∗ Ψ)
2

In special theory of relativity, Einstein’s relation between energy and mass

tells us that

 2 = 2 4 + 2 2 

In a similar way, we may have the Klein-Gorden Equation

−~2 2 Ψ = (2 4 − 2 ~2 ∇2 )Ψ

For a plane-wave solution,

Ψ = exp[( ·  − )~]

297
CHAPTER 7 – MANUSCRIPT
p
 = ± 2 4 + 2 2

The continuity equation

  + ∇ ·  = 0

where

~
 = [Ψ∗  Ψ − Ψ Ψ∗ ]
22

~ ∗
 = − (Ψ ∇Ψ − Ψ∇Ψ∗ )
2

Difficulties:

1. Negative energy

2. Negative probability

Spectrum of Hydrgen atom

( −  ())2 Ψ = (2 4 − 2 ~2 ∇2 )Ψ

B 7.1.2 Dirac Equation

Dirac wrote a linear equation

 =  ·  + 2 

and required that

 2 = ( ·  + 2 )2

= 2 2 + 2 4

298
CHAPTER 7 – MANUSCRIPT

As a result,

  +   = 2 

  +  = 0

2 = 1

In 2D, we have

 =  

 =  

 = 

In 3D, Dirac realized that  and  should be at least 4×4 matrices. One representation

is
⎛ ⎞
⎜ 0  ⎟
 = ⎜

⎟ =  ⊗ 

 0
⎛ ⎞
⎜  0 ⎟
 = ⎜

⎟ =  ⊗   (=  0 ⊗   )

0 −

In this way we have a linear Dirac equation,

~ Ψ = ( ·  + 2 )Ψ

299
CHAPTER 7 – MANUSCRIPT

Its Hermitian conjugate is

−~ Ψ† = Ψ† (+~∇ · † + 2  † )

we have the continuity equation with

 = Ψ† Ψ

 = Ψ† Ψ

The velocity operator


1
 = [ ] = 
~

The plane wave solution

Ψ( ) =  exp[( ·  − )~]

( ·  + 2 ) = 

Since  and  are 4 × 4 matrices,


⎛ ⎞
⎜ 1 ⎟
⎜ ⎟ ⎛ ⎞
⎜ ⎟
⎜  ⎟
⎜ 2 ⎟ ⎜ 0 ⎟
=⎜

⎟=⎜
⎟ ⎝


⎜  ⎟ 0
⎜ 3 ⎟
⎜ ⎟
⎝ ⎠
4

300
CHAPTER 7 – MANUSCRIPT

The Dirac equation is reduced into

( − 2 )0 −  · 0 = 0

− · 0 + ( + 2 )0 = 0

The solution is
p p
 =  2 2 + 2 4 = ± 2 2 + 2 4

 · 
0 = 
 + 2 0

and ⎛ ⎞
⎜ 0 ⎟
+ = + ⎜



√2 ·
0
 2 +2 4 +2
⎛ ⎞
√2 ·
⎜ 0 ⎟
− = − ⎜

−  2 +2 4 −2 ⎟

0

The rest energy of electron 2 = 051099906 MeV = 051 × 106 The electron energy

in solids is at most about several eV it would be a good approximation in a non-

relativistic limit

⎛ ⎞
⎜ 0 ⎟
+ = + ⎜



0

for the positive energy

301
CHAPTER 7 – MANUSCRIPT

⎛ ⎞
⎜ 0 ⎟
− = − ⎜



0

for the negative energy.

B 7.1.3 Spin of Dirac particles

The orbital angular momentum

L=r×p

[  ] = ~( × )

We need to introduce another operator in define J


⎛ ⎞
⎜  0 ⎟
 =  + ~Σ = r × p+~ ⎜



0 

such that

[ ] = 0

Thus the Dirac theory introduce a spin operator


⎛ ⎞
⎜  0 ⎟
 = ~Σ = ~ ⎜



0 

302
CHAPTER 7 – MANUSCRIPT

B 7.1.4 The Zeeman coupling

Starting with an electron in an electromagnetic field,


[ · ( − ) + 2  +  ]() = ()

⎛ ⎞
⎜ () ⎟
Denote () = ⎜



()


( −  − 2 ) =  · ( − )


( −  + 2 ) =  · ( − )

 1 
( −  − 2 ) =  · ( − ) 2
 · ( − )
  −  +  

In the non-relativistic limit,

 =  + 2  

 1 
( −  ) =  · ( − )  · ( − )
  −  + 22 

1 h  i2
≈  · ( − ) 
2 

303
CHAPTER 7 – MANUSCRIPT
X
( · A) ( · B) =      


X
= (  +    )  


= A · B +  · (A × B)

h  i2  ~
 · ( − ) = ( − )2 +  · (∇ × )
  

The final results


∙ ¸
1  2 ~
  = ( − ) +  + · 
2  2

B 7.1.5 Exact solution in Coloumb potential: Hydrogen atom


£ ¤
 ·  + 2 +  ()  = 

2
 = " #12
 2 2
1+ h √ i2
+12+ (+12)2 − 2 2

where  () = −2  and  is the fine structure constant.

B 7.1.6 Spin-orbit coupling

Higher order approximation:

304
CHAPTER 7 – MANUSCRIPT

 1 
( −  ) =  · ( − ) 2
 · ( − )
  −  + 2 

 1 
( −  ) =  · ( − ) 2
 · ( − )
 2 −  

1   −  
( −  ) = 2
 · ( − )(1 − 2
) · ( − )
2  2 

1 h  i2
=  · ( − ) 
2 

1 h  i h  i
+ 2 2  · ( − ) ( −  )  · ( − ) 
4   

h  i2  ~
 · ( − ) = ( − )2 +  · (∇ × )
  

h  i h  i
 · ( − )   · ( − ) = ~ · (∇ × ) + · · ·
 

Final results
∙ ¸
1  2 ~
  = ( − ) +  + · 
2  2

~
+  · (∇ × p) + · · · · · ·
42 2

Spin-orbit coupling or L S coupling

305
CHAPTER 7 – MANUSCRIPT

In a rotational invariant potential

~
 =  · (∇ × p)
42 2

~ 1  1 1 
= 2 2
 · (r × p) = S·L
4    22 2  

B 7.1.7 Spin-orbital coupling and spintronics

Spin-orbit coupling from the classic point of view

In a 2D plane with ∇ = −ẑ

 = ( × p) = (  −    )

B 7.1.8 Spin transverse force

1
 = [ ]
~

 1
= [ ]
 ~

B 7.1.9 Spin-orbit coupling and Datta-Das field-effect transistor

B 7.1.10 Graphene: 2D massless Dirac quasi-particles

Model for graphene: a monolayer of carbon atoms

306
CHAPTER 7 – MANUSCRIPT

Using Spin-Orbit Coupling: spin FET


V

Figure 7.1:

307
CHAPTER 7 – MANUSCRIPT

Figure 7.2:

308
CHAPTER 7 – MANUSCRIPT

Figure 7.3:

309
CHAPTER 7 – MANUSCRIPT

Figure 7.4:

310
CHAPTER 7 – MANUSCRIPT

Figure 7.5:

311
CHAPTER 7 – MANUSCRIPT

C Graphene and massless Dirac particles

Graphene–a monolayer of carbon atoms packed into a honeycomb lattice, is one of

the most intriguing systems in solid state physics today. The most popular description

of graphene band structure is the tight binding one, first done by Wallace. The band

structure exhibits very unique features: the first two bands do not have a gap between

them, and do not overlap either. In fact they intersect in two inequivalent points,

called Dirac points in the first BZ. resulting in four degenerate modes. The vicinity

of the dirac points is not parabolic (as in most crystals) but in fact conical. Hence,

the group velocity  ≡ ∇k E() is independent of the energy. The Fermi level for

undoped graphene lies exactly at the intersection points, and graphene is a gapless

semiconductor. Since the dispersion curve resembles that of ultra-relativeistic particles,

one can write a relativistic dynamic equation for the excitations. Such equation can be

derived from the tight binding model, and the resulting equations are infact the well

known Dirac equation for massless particles.


 =  3

 being the bond length. We shall took  = 1, then


1 = ( 32 +12)


2 = ( 32 −12)

312
CHAPTER 7 – MANUSCRIPT

Figure 7.6:


Unit cell area is:  = 2 sin(3) = 32.

Fig.: (a) Graphene hexagonal lattice constructed as a superposition of two

triangular lattices A and B, with basis vectors a for lattice A and vectors δ  with

 = 1 2 3 connecting A to B. (b) the green hexagon is a Brillouin zone (BZ) and pink

diamond is the extended BZ for the honeycomb lattice. The reciprocal lattice vectors

are b .

Reciprocal lattice:

̄2 × ̂ √
1 = 2 = 2(1 3 1)

̄1 × ̂ √
2 = 2 = 2(1 3 −1)

313
CHAPTER 7 – MANUSCRIPT

The tight binding Hamiltonian in the nearest neighbors approximation is

X
0 = − †n n+δ  + †n+δ  n  (7.1)
nδ 

 the hopping parameter, n ,n+δ  are annihilation operators corresponding A,B sub-

lattices, and suppressing the spin degree of freedom which is irrelevant for optics. The

above Hamiltonian can be written as


⎛ ⎞⎛ ⎞
X³ ´ ⎜ 0 1 ⎟ ⎜ n ⎟
0 = − †n  †n+δ  ⎜⎝
⎟⎜
⎠⎝

⎠ (7.2)
nδ  1 0 n+δ 

It is possible to diagonalize the Hamiltonian at this point. However, one can

gain some more insight on the excitations in Fourier space. Expanding the operators

( 2 is the number of primitive unit cells, the lattice is  ×  )

1 X k·a
a =  (k) (7.3)
 k

1 X k·(a+δ  )
a+δ  =  (k) (7.4)
 k

and plugging it into Eq. (7.1)

X 1 X X X (k−q)·n −q·δ 
†n+δ  n = 2
  (q)† (k)
nδ 
 k q nδ

X X X
= (k)† (k) −k·δ  ≡ (k)† (k)(k) (7.5)
k δ k

Performing the summation and plugging the results into (7.1) we obtain

X ¡ ¢
0 = Φ† (k)H0 Φ(k) Φ† (k) = (k)†  (k)†  (7.6)
k

314
CHAPTER 7 – MANUSCRIPT
⎛ ⎞

⎜ 0  (k) ⎟
H0 (k) = ⎜


⎠ (7.7)
(k) 0

where
∙ √ √
µ ¶¸
  3 − (2 3)  
(k) = −  + 2 cos (7.8)
2

∙ √
µ ¶¸
  3 −  2 3  
= − 1 + 2 cos  (7.9)
2

Note that diagonalizing H0 , also diagonalize 0 . The energy bands are ob-

tained by the eigenvalues


v Ã√ !
u µ ¶ µ ¶
u     3 
E(k) = ± |(k)| = ±t1 + 4 cos2
  
+ 4 cos cos  (7.10)
2 2 2
√ √
which vanish at six K points ±2(1 3 13), ±2(0 23) and ±2(−1 3 13).

One can show that these points are connected by reciprocal lattice vectors, and only

two are inequivalent, denoted by K+ , K− . One can find the points by requiring  = 0.

For this − 32
must be real, so  = 0. Then   = 23.
√ √
Alternatively, −  3
= −1,  = ±2 3 and  = ±23.

Expanding E(k) around e.g. ±2(0 23) we get


 = − 32( ±  ) (7.11)

Expanding E(k) around these points we obtain



3
E(p) = ±~ |p|   =  (7.12)
2~

315
CHAPTER 7 – MANUSCRIPT

and p measured from the K± points.

One can cast the spectrum in the vicinity of the K± points in the manifestly

relativistic way. For this let us turn the coordinate axes by 90◦ so that  → 0 and

 → − 0 . Then, Eq. 7.11 will look like


 = − 32(−0 ± 0 ) (7.13)

The Hamiltonian in the vicinity of K± then becomes


⎛ ⎞
√ ⎜ 0 ±0 + 0 ⎟
H0 (K± + p) = − 32 ⎜


⎠ (7.14)
±0 − 0 0

=  (±  0 +   0 ) (7.15)

In the graphene, the effective velocity is about c/300 where c is the speed of light

C Solution in a magnetic field and quantum Hall effect

In a magnetic field, we introduce a vector potential  = ( 0) such that

 
 →  −  → ~ − 
 


 →  −  → −~

For (+ + )
⎛ ⎞

⎜ 0 ~ − 
 + ~ ⎟
() =  ⎜



~ −   − ~ 0

316
CHAPTER 7 – MANUSCRIPT

The communtator

   
[~ −  + ~  ~ −  − ~ ] = [−  −~ ] + [~  − ]
   

~
= −2  = 2~2 
2
0

 
 = √ (~ −  + ~ )
2~ 

 
† = √ (~ −  − ~ )
2~ 

⎛ ⎞ ⎛ ⎞
√ ~ ⎜ 0  ⎟ ⎜ 0  ⎟
() = 2 ⎜ ⎟ = 0 ⎜ ⎟
 ⎝ † ⎠ ⎝ ⎠
 0 † 0

with [ † ] = 1 In this way, the stationary Schrodinger equation is reduced to


⎛ ⎞⎛ ⎞ ⎛ ⎞
⎜ 0  ⎟⎜  ⎟ ⎜  ⎟
0 ⎜

⎟⎜ ⎟ =  ⎜ ⎟
⎠⎝ ⎠ ⎝ ⎠
† 0  

0  = 

0 †  = 

(0 † )0  = (0 † ) =  2 

(0 )2 †  =  2 

†  |i =  |i

317
CHAPTER 7 – MANUSCRIPT

 = h|i


 = 0 

0 0 √
 = h||i =  h| − 1i =  h| − 1i
 

For  = 0, =0 = 0.

B 7.1.11 Topological Insulator and Dirac particels

Certain insulators have exotic metallic states surrounding arround the boundary or

surfaces. These states are formed by the strong spin-orbital coupling as a result of

topological effect that the motion of electrons is insensitive to scattering of inpurities.

The electrons in the surface states are well described by an effective Hamiltonoan of

two-component massless Dirac particle.

 =  p ·  =  (   +    )

This system is different from graphene: there are double spin degeneracy and double

valley degeneracy of Dirac cone in graphene. The surface states in topological insulator

is a single Dirac cone.

The topological insulator was predicted theoretically, and confirmed experi-

menatlly. Several families of topological insulator have been discovered. The materials

318
CHAPTER 7 – MANUSCRIPT

Figure 7.7:

may provide new route to generate novel particles and quantum states and to explore

application in spintronic device and quantum computing.

A 7.2 Quantum Statistics

Quantum statistical mechanics is the branch of physics dealing with systems in mixed

states; it is the quantum analogue of classical statistical mechanics. It should be noted

that statistics enters at two levels in quantum statistical mechanics: first, because of

the statistical interpretation of the wave function and second, because of our incomplete

knowledge of dynamical state of the system.

319
CHAPTER 7 – MANUSCRIPT

B 7.2.1 Density Operator and Ensembles

Question: How to describe quantum mechanically an ensemble of physical systems

for which, say   = 60 are characterized by |i and the remaining  = 40 are

characterized by some other ket |i 

Example: a general state for S=1/2 system

|i = + |+i + − |−i

which characterizes a state ket whose spin is pointing some definite direction. We

should not confuse the probability weight of the states |+i and |−i ,  ±  and |± |2 

Ensemble average and density operator.

Pure ensemble: |i

Mixed ensembles:
¯ ®
1 ¯(1)
¯ ®
2 ¯(2)
¯ ®
3 ¯(3)
.. ..
. .

with the normalization condition

X
 = 1

320
CHAPTER 7 – MANUSCRIPT

The ensemble average

X
hi =   h ||  i

X X
=  h |i h| |  i
 
X
=   |h| i|2 


Alternatively,

X
hi =  h| i h |i h || i


X X
= h| { | i h | } |i
 

=  ()

where
X
=   | i h | 

 is the density matrix operator and is independent of representation.

Two properties:

(a).  is Hermitian;

(b).  satisfies the normalization condition.

 () = 1

A pure ensemble:   = 1 for some | i

 = | i h |

321
CHAPTER 7 – MANUSCRIPT

which is just a projection operator.

Several examples of density operator:

(1) A completely polarized beam with S +


⎛ ⎞
µ ¶
⎜ 1 ⎟

 = |+i h+| = ⎝ ⎠ 1 0⎟
0
⎛ ⎞
⎜ 1 0 ⎟
= ⎜



0 0

(2) A completely polarized beam with S +

 = |  +i h  +|

1 1
= (|+i + |−i) (h+| + h−|)
212 212
⎛ ⎞
1 1
⎜ 2 2 ⎟
= ⎜



1 1
2 2

(3) A partially polarized beam with 75 S + and 25 S +

 = 075 |+i h+| + 025 |  +i h  +|


⎛ ⎞
1⎜ 7 1 ⎟
= ⎜ ⎟
8⎝ ⎠
1 1

B 7.2.2 Quantum Statistical Mechanism

Order and Disorder

322
CHAPTER 7 – MANUSCRIPT

The density matrix of a pure ensemble


⎛ ⎞
⎜ 0 ⎟
⎜ ⎟
⎜ ... ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
⎜ ⎟
=⎜
⎜ 1 ⎟

⎜ ⎟
⎜ ⎟
⎜ ... ⎟
⎜ ⎟
⎜ ⎟
⎝ ⎠
0

The density matrix of a completely random ensemble


⎛ ⎞
1
⎜  ⎟
⎜ ⎟
⎜ ⎟
⎜ 1 ⎟
⎜  ⎟
=⎜



⎜ ... ⎟
⎜ ⎟
⎜ ⎟
⎝ ⎠
1

We define a quantity

 = − ( ln )

X
= −  ln 

For a pure ensemble

=0

For a completely random ensemble

 = ln 

323
CHAPTER 7 – MANUSCRIPT

This quantity is related to the entropy in thermodynamics,

 =  ( ln )

where k is the Boltzmann constant.

Basic assumption: Nature tends to maximize  subject to the constraint that

the ensemble average of the Hamiltonian has a certain value.

To minimize :

 = 0

i.e.,

 (ln  + 1) = 0

with the conditions

(a). hi =  = ()

(b).  = 1

we obtain

 [(ln  + 1) +  + ] = 0

 = − −

1
= P −
−


324
CHAPTER 7 – MANUSCRIPT

The partition function:

X
 = −

= (− )

Thus the density matrix operator is

1 −
= 

The parameter  is related to the temperature T as follows

 = 1

Example: Spin-1/2 particle in a magnetic field

Consider a spin-1/2 particle having a magnetic momentum

 = − ~

subjected to a constant magnetic field along the z-axis.

 =    2 = 

The density operator

1 −
 = 

⎛ ⎞
1 −
⎜ 
 0 ⎟
= ⎜



1 +
0 

325
CHAPTER 7 – MANUSCRIPT

where

 = − + 

The average value of the z-component of spin at the temperature T is

~ − − 
h i =
2 − + 

~
= − tanh()
2

B 7.2.3 Quantum Statistics

Consider a system with two identical particles. Each particle has three non-degenerate

states: |1i  |2i  |3i with E1 , E2 , E3 . For classical identical particles: there 9 possible

configurations
|1 1i |1 2i |1 3i

|2 1i |2 2i |2 3i

|3 1i |3 2i |3 3i

The wave function is

Φ = Φ (1)Φ (2)

326
CHAPTER 7 – MANUSCRIPT

For bosons: there are 6 possible configurations

|1 1i  |2 2i  |3 3i

(|1 2i + |2 1i) 212

(|1 3i + |3 1i) 212

(|3 2i + |2 3i) 212

The wave function

Φ = Φ (1)Φ (2) + Φ (1)Φ (2)

For fermions: there are 3 possible configurations

(|1 2i − |2 1i) 212

(|1 3i − |3 1i) 212

(|3 2i − |2 3i) 212

The wave function

Φ = Φ (1)Φ (2) − Φ (1)Φ (2)

B 7.2.4 Systems of non-interaction particles

We shall now discuss the properties of systems of large numbers of non-interacting

objects which are equivalent and possess the same energy levels.

327
CHAPTER 7 – MANUSCRIPT

C Maxwell-Boltzmann Statistics

(Classical identical particles)

If E (j=1, 2, · · · ) denotes an energy level of one of the particles, the total

energy of the system E can be written

X
=  

where the sum runs over all the eigenenergies of a single particle, and where n is the

number of particles with the energy E . The total number of particles is

X
=  

In a Maxwell-Boltzmann system, any number of particles can be in each level, so that

nj can either be zero or any positive integer. If all the energies Ej are different, each

permutation of the particles results in a different wave function, thus each energy level

of the total system is N!-fold degenerate. However if nj1, the interchange between

these particles do not alter the wave function. Thus the number of distinct states

corresponding to a given value of the total energy of the system is

!
 = Q
  !

328
CHAPTER 7 – MANUSCRIPT

The partition function can be expressed as

X X
 =  exp(−   )
{ } 

X ! X
= Q exp(−   )
  !
{ } 

" #2
X
= exp(−  )

The average distribution is

1 X  X
h i =  exp(−   )
 
{ }

 ln 
= −


 −
=  = − −

C Bose-Einstein Statistics (for bosons)

The wave function of many bosons is symmetric and the energy level of the system is

non-degenerate

 = 1 for any 

Here we use the grand canonical ensemble ( → +∞)

1 X  X X
h i =  exp(−   −   )
  
{ }

329
CHAPTER 7 – MANUSCRIPT

where

X X X
 = exp(−   −   )
{ }  

⎛ ⎞
Y X
= ⎝ exp(−  −  )⎠
 { }

Y
= (1 − exp(− − ))−1

In the case,

 ln 
h i = −


1
=
exp( + ) − 1

C Fermi-Dirac Statistics (for fermions)

The wave function of many bosons is antisymmetric, and the energy level of the system

is non-degenerate

  = 1 for  = 1 or 0

= 0 otherwise.

Here we use the grand canonical ensemble ( → +∞)

1 X  X X
h i =  exp(−   −   )
  
{ }

330
CHAPTER 7 – MANUSCRIPT

where

X X X
 =   exp(−   −   )
{ }  

⎛ ⎞
Y X
1
= ⎝ exp(−  −  )⎠
  =0

Y
= (1 + exp(− − ))

In the case,

 ln 
h i = −


1
=
exp( + ) + 1

B 7.2.5 Bose-Einstein Condensation

One of the main features of boson system is that bosons tend to occupy the lowest energy

state at low temperatures. At high temperatures, both distribution laws for bosons and

for fermions become equal to that for the Maxwell-Boltzmann case approximately. The

density of particles on a certain state always tends to be zero. For a boson system, as

1
h i = → +∞
exp( + ) − 1

if both E and  are equal to zero, the density of the particle at E=0 can be nonzero in

the high density limit at low temperatures.

331
CHAPTER 7 – MANUSCRIPT

Consider a non-interacting boson gas in a three-dimensional system. The

spectrum of energy is
~2 2
= 
2

The density of particle is


Z
1
 = 0 + 3 ()
(2)3

where 0 is the density of particles in the lowest energy state.

Z Z
1 3
 () = ()()
(2)3

where
µ ¶32
1 2
() = 2  12
4 ~2

This is the density of state at E.


µ ¶32 Z +∞
1 2  12
 − 0 = 
4 2 ~2 0 + − 1
µ ¶32
1 2
= 12 ()
4 2 ~2

where
Z +∞
12
12 () =  
0 + − 1

Since  ≥ 0 the largest possible values of 12 () occurs when  = 0 and numerically

12 ( = 0) = 2315

332
CHAPTER 7 – MANUSCRIPT

It follows;; that at   0 where


µ ¶23
~2 42 
0 =
2 2315

the function ( − 0 ) becomes less than unit, and the system increasingly condensed

into the lowest energy state.

Bose-Einstein condensation was predicted by Albert Einstein and Satyendra

Nath Bose at 1924 to 25. It was first observed in Helium-4 liquid. Recent years it was

observed in some systems with cooling atoms. It is one of the most active branches in

condensed matter physics. If you are interested in the recent development in this field,

please visit Web site:

http://www.aip.org/pt/webwatch/ww9703.html.

B 7.2.6 Free fermion gas

The free particle Schrodinger equation in three dimensions is


µ ¶
~2 2 2 2
− + + Ψ () =  Ψ ()
2 2  2 2

333
CHAPTER 7 – MANUSCRIPT

If the particle are confined to a cube of edge L and we require the wave function to be

periodic in x, y, z direction with L, i.e.,

Ψ ( +   ) = Ψ ( +   )

Ψ (  +  ) = Ψ (  +  )

Ψ (   + ) = Ψ (   + )

the wave functions are


1
Ψ () = 32
·
(2)
provided that the components of the wave vector k satisfies

2 4 6
 = 0 ± ± ± ···
  

Any component of k is of the form 2n, where n is a positive integer. The energy

spectrum with n is
~2 ¡ 2 ¢
 =  + 2 + 2
2

As two indistinguishable fermions cannot have all their quantum numbers identical,

each energy level (or state) can be occupied by at most one particle. Thus we start

filling the energy level from that with the lowest energy (i.e., k=0 here) until all N

particles are accommodated.

The occupied energy levels may be represented as points inside a sphere in the

k-space. The energy at the surface of the sphere is the Fermi energy; the wave vectors

334
CHAPTER 7 – MANUSCRIPT

at the Fermi surface have a magnitude k such that

~2 2
 =  
2 

The total number of energy levels inside the sphere is

1 4 3
= 3 
(2) 3

Then
¡ ¢13
 = 6 2  

In short, a boson gas condensates into the lowest energy state, and a fermion

gas forms a Fermi sphere in the k-space at absolute zero temperature.

A 7.3 Quantum Hall Effect

The quantum Hall effect is a remarkable phenomenon discovered experimentally in

1980s in which the Hall conductivity of a two-dimensional system of electrons is found

to have plateaus as a function of variables which determine the number of electrons

participating in the effect. The integer quantum Hall effect was observed by von Klitzing

et al in 1980. Simple theory suggests that the Hall conductivity at he plateaus should

be an integral multiple of 2 ~ and the experiments agree with that prediction to within

an accuracy of nearly 0.1 ppm. An application of great importance is to metrology, the

quantum Hall effect promises a method of providing very precise resistance standards

335
CHAPTER 7 – MANUSCRIPT

that are insensitive to the particular sample and the details of its fabrication. With

a more powerful magnetic field and lower temperature, Daniel C. Tsui and Horst L.

Stormer discovered that the plateaus of the conductivity is a fractional multiples of the

basic unit 2 ~ Within a year of their discovery, Rovert B. Laughlin has succeeded in

explaining their result. Through theoretical analysis he showed that the electrons in a

powerful magnetic field can condense to form a kind of quantum fluid related to the

quantum fluids that occur in superconductivity and in liquid helium.

What makes these fluids particularly important for researchers is that events

in a drop of quantum fluid can afford more profound insights into the general inner

structure and dynamics of matter. The Royal Swedish Academy of Sciences awarded

The 1998 Nobel Prize in Physics jointly to Tsui, Stomer and Laughlin “for their dis-

covery of a new form of quantum fluid with fractionally charged excitations.”

B 7.3.1 Hall Effect

The Hall effect occurs when the charge carriers moving through a material experience a

deflection because of an applied magnetic field. This deflection results in a measurable

potential difference across the side of the material which is transverse to the magnetic

field and current direction.

336
CHAPTER 7 – MANUSCRIPT

In Figure, a voltage V drives a current I in the positive X direction. Normal

Ohmic resistance is V/I . A magnetic field tin the positive z direction shifts positive

charge carriers in the negative y direction. This generates a Hall potential ( ) in

the y direction.

Quantitatively, Hall effect indicates that a transverse electric field will be

produced. Charges will accumulate on the transverse edges until a strong enough

electric field is developed to force remaining charges to continue undeflected. The

condition is

= 

The drift velocity is given by



 =  =


337
CHAPTER 7 – MANUSCRIPT

where n is the density of charge carrier. The electric field is

 =  

where l is the thickness of material and d the width. Hence we obtain an expression

for the density of charge carriers in a substance

 1 
= · =
   

Therefore this effect can be used to determine the density of charge carriers in conduc-

tors and semi-conductors, and has become a standard tool in physics laboratories over

all the world. The Hall conductivity

 = 

 
= ∝
 

B 7.3.2 Quantum Hall Effect

In 1980 the German physicist Klaus von Klitzing discovered in a similar experiment that

the Hall resistance does not vary in linear fashion, but ”step-wise” with the strength

of the magnetic field. The steps occur at resistance values that do not depend on

the properties of the material but are given by a combination of fundamental physical

constants divided by an integer. We say that the resistance is quantized. At quantized

Hall resistance values, normal Ohmic resistance disappears and the material becomes

338
CHAPTER 7 – MANUSCRIPT

in a sense superconducting. For his discovery of what is termed the integer quantum

Hall effect von Klitzing was awarded the Nobel Prize in 1985.

In their refined experimental studies of the quantum Hall effect, using among

other things lower temperatures and more powerful magnetic field, Stomer, Tsui and

their co-workers found to their great surprise a new step in the Hall resistance which

was three times higher than von Klitzing’s highest. They subsequently found more and

more new steps, both above and between the integers. All the new step heights can be

expressed with the same constant as earlier but now divided by different fractions. For

this reason the new discovery is named the fractional quantum Hall effect..

339
CHAPTER 7 – MANUSCRIPT

B 7.3.3 Laughlin’s Theory

A year after the discovery of the fractional quantum Hall effect, Laughlin offered a

theoretical explanation. According to his theory the low temperature and the powerful

magnetic field compel the electron gas to condense to form a new type of quantum fluid.

Since electrons are most reluctant to condense (They are what is termed fermions) they

first, in a sense, combine with the ”flux quanta” of magnetic field. Particularly for the

first step ( = 13) discovered by Stomer and Tsui, each of electrons captures three

flux quanta thus forming a kind of composite particle with no objection to condensing.

(They become what is termed bosons). Quantum fluids have earlier occurred at very

340
CHAPTER 7 – MANUSCRIPT

low temperatures in liquid helium and in superconductors. They have certain properties

in common, e.g. superfluidity, but they also show important differences in behaviors.

Apart from its superfluidity which explains the disappearance of Ohmic resistance at the

Hall resistance steps„ the new quantum fluid proposed by Laughlin has many unusual

properties. One of the most remarkable is that if one electron is added the fluid will

be affected (excited) and a number of fractionally charged “quasiparticles” created.

These quasiparticles are not particles in the normal sense but a result of the common

dance of electrons in the quantum fluid. Laughlin was the first to demonstrate that

the quasiparticles have precisely the correct fractional charge to explain the fractional

quantum Hall effect. Subsequent measurements have demonstrated more and more

fractional charged steps in the Hall effect, and Laughlin’s quantum fluid has proved

capable of explaining all the steps experimentally. The new quantum fluid strongly

resists compression; it is said to be incompressible.

Further reading
1. B. Davis, Splitting the electron, New Scientist, 31 Jan, 1998, p36

2. G. P. Collins, Fractionally charged quasiparticles signal their presence with

noise, Physics Today, Nov,1997, p17,

3. S. Kivelson, D. H. Lee and S. C. Zhang, Electrons in flatband, Scientific

American, March,1996, p64.

341
CHAPTER 7 – MANUSCRIPT

B 7.3.4 Charged particle in the presence of a magnetic field

We consider a charged particle moving on a plane in the presence of a perpendicular

uniform magnetic field B. The mass of the particle is M and charge e. The Hamiltonian

is

1 
 = ( + )2
2 

1    
= [(−~ +  )2 + (−~ +  )2 ]
2    

The vector potential A, is such that its curl is equal to B:

 = (∇ × A)

We will work in the isotropic gauge

1
A=− B×r
2

i.e.

1
 = − 
2

1
 = 
2

In the gauge

1  
 = (−~ − )2
2  2

1  
+ (−~ + )2
2  2

342
CHAPTER 7 – MANUSCRIPT

Denote

 
Π = −~ − 
 2

 
Π = −~ + 
 2

The commutator is constant,


~
[Π Π ] = − 

The Hamiltonian is expressed as

1
 = {(Π − Π )(Π + Π ) + [Π  Π ]}
2

1 ~
= [(Π − Π )(Π + Π ) +
2 

 + 1
= ~ [  + ]
 2

where

³  ´12
+ = (Π − Π )
2~
³  ´12    
= (−~ +  − ~ + )
2~  2  2

 1
= 212 (− + )
 ̄ 4

 1
 = 212 (+ + ̄)
 4

where  = ( + )0 and  = ( − )0  The length unit 0 = (~)12 

343
CHAPTER 7 – MANUSCRIPT

Hence the Hamiltonian is reduced to

1
 = ~  (+  + )
2

where

 = 


 and + satisfy [ + ] = 1

Comparing to the simple harmonic oscillator„ the motion of a charged particle

in a uniform magnetic field is equivalent to it. A set of the wave function has the form

 ( ̄) =   −̄4

The energy eigenvalue is


1
 = ~  ( + )
2

The wave functions are also eigenstates of angular momentum operator

 
 = −~( − )
 

 
= ~( − ̄ )
  ̄

with the eigenvalue

  ( ) = ~ ( )

The energy level  is called  Landau level

344
CHAPTER 7 – MANUSCRIPT

The Landau levels have a huge degeneracy which is determined by

  
 = = =
0
~
2  202

To make this degeneracy more apparent, we assume the system has the shape of a

square with dimension L. We introduce another pair of operators

 1
+ = 212 ( − )
 4

 1
 = 212 (− − )
 4

([ + ] = 1)

The two operators commute with H,

[ ] = [+  ] = 0

Also , the operator  annihilates the wave function  ( ), just like the operator 

 ( ) = 0

Define

̄ = exp[20 ]

 = exp[20  † ]

Thus,  is an engenstate of ̄

̄  = 

345
CHAPTER 7 – MANUSCRIPT

A complete set of eigenstates of the  Landau level { }can be constructed ( =

1  )

 ( ̄) =    ( ̄)

= exp[20 ] ( ̄)

From the definitions of  and ̄

̄  =  ̄ exp[20  20 † ]

= exp[2 ] ̄

Therefore the states  ( ̄) have

1
(̄) = ~( + ) ( ̄)
2

̄ (̄) = exp[2 ] ( ̄)

B 7.3.5 Landau Level and Quantum Hall Effect

The degeneracy of the Landau energy levels is  −fold. The ratio of the number of

electrons to 

 =  

is called the filling number of Landau level. When  = 1it means that the first Landau

level is fully filled and when  = 2, the second landau level is also filled. There is a

346
CHAPTER 7 – MANUSCRIPT

simple relation between the filling number and the quantized Hall conductivity. Suppose

the total charge of the sample is

 =  

The current J is then equal to

J =  

 is determined by the magnetic field B and electric field E

 
=
 

So the current density j = J

   
 = = 
 

  2
=  =  
 · ~

~

The Hall conductance is


2
 =  
~

The integer quantum Hall effect occurs at  = , (1,2,...) and the

fractional Hall effect occurs at  = : 1/3, 1/5, 2/3, ...

A 7.4 Quantum Magnetism

Magnetism is inseparable from quantum mechanics, for a strictly classical system in

thermal equilibrium can display no magnetic momentum, even in a magnetic field. The

347
CHAPTER 7 – MANUSCRIPT

magnetic momentum of a free atom has three principal sources: the spin with which

electrons are endowed, their orbital angular momentum about the nucleus; and the

change in the orbital moment induced by an applied magnetic field.

Ordered arrangements of electron spins

348
CHAPTER 7 – MANUSCRIPT

B 7.4.1 Spin Exchange

Ferromagnetism is obtained in solids when the magnetic moments of many electrons

align. Antiferromagnetism and spin density waves describe oscillatory ordering of mag-

netic moments. The classical dipolar interaction between the electron moments (which

is of order 105  ) is for too weak to explain the observed magnetic transition temper-

ature (which are of order 102 − 103 0  in transition metal and rare earth compounds)

The coupling mechanism that gives rise to magnetism derives from the fol-

lowing fundamental properties of electrons:

• The electron’s spin

• The electron’s kinetic energy

• Pauli exclusion principle

• Coulomb repulsion

Before we introduce the physical origin of the magnetic coupling between

electrons in solids, we simply review some standard definitions and basic relations of

second quantization.

For an orthonormal single-particle basis, {| i}

­ ®
 | =   

349
CHAPTER 7 – MANUSCRIPT

The creation operator of state  is † and its Hermitian conjugate is annihilation oper-

ator   Both are defined with respect to the vacuum state |0i such that

| i = † |0i

 |0i = 0

The number operator is defined as

 = † 

For bosons,

h i
  † =  † − †  =  

{   } = 0

For electrons with spin  = 12, we have to introduce a pair of operators, † where

 =↑ ↓ such that

{  † } =    

{   } = 0

350
CHAPTER 7 – MANUSCRIPT

The spin operator can be expressed as

† =  +  = ~†↑ ↓

− =  −  = ~†↓ ↑

~ †
 = ( ↑ − †↓ ↓ )
2 ↑

The commutation relations:

[†  − ] = 2~

[  ± ] = ±~±

B 7.4.2 Two-Site Problem

The Hersenberg spin exchange interaction is written as

 = S1 · S2

To obtain the eigenstates of H, we check the following relations,

(1) [S21  ] = 0but [S1  ] 6= 0

(2) [S22  ] = 0but [S2  ] 6= 0

(3) [S1 + S2  ] = 0

Therefore S21  S22 and S2 = (S1 + S2 )2 and its z-component S are good

quantum numbers, but 1  2 are not. Hence we can denote the simultaneous eigenkets

351
CHAPTER 7 – MANUSCRIPT

of S2  S  S21 and S22 by



|    1  2 i

such that


S21 |    1  2 i


= 1 (1 + 1) |    1  2 i


S22 |    1  2 i


= 2 (2 + 1) |    1  2 i

S1 |  



 1  2 i


=  ( + 1) |    1  2 i

S1 |  



 1  2 i

 
=  |    1  2 i

Fortunately, the state kets are also the eigenkets of H:

 
 |    1  2 i =  |    1  2 i

where

= [ ( + 1) − 1 (1 + 1) − 2 (2 + 1)]
2

352
CHAPTER 7 – MANUSCRIPT

and

 = |1 − 2 |  |1 − 2 | + 1 1 + 2


Since the energy eigenvalues are independent of  can be −   the energy

eigenstates are (2 + 1)−fold degenerated. From the point of view of symmetry,

the degeneracy of the eigenstates originates from the invariance of H under the (2)

symmetry rotation,

 † = 

where

 = exp[−S · n~]

The ground state:

The lowest energy state is determined by the sign of J: when   0 

should be taken to be minimum, otherwise  should be taken to be maximal.

The case of   0 :

 = 1 − 2  (1  2 )

The two spins are antiparallel, which is called antiferromagnetic. The ground state

energy

 = −(1 + 1)2

The case of   0

 = 1 + 2

353
CHAPTER 7 – MANUSCRIPT

The two spins are parallel, which is ferromagnetic

 = − || 1 2

B 7.4.3 Ferromagnetic Exchange (  0)

Ferromagnetic exchange coupling originates from the direct Coulomb interaction and

the Pauli exclusion principle. In the second quantized form, the two-body Coulomb

interaction is given by
Z
1
 = ̃( )Ψ† ()Ψ†0 ()Ψ0 ()Ψ ()
2

The field operator


X
Ψ† () = ∗ ()† 

The interaction can be expressed as


Z
1
 = ̃( )
2

×† ()† () () ()† †0 0 

X
=   −

X
+ 0 (↑ + ↓ )(0 ↑ + 0 ↓ )
0

X
+ 0 † †0 0 0 0 
0

+···

354
CHAPTER 7 – MANUSCRIPT

where
Z
1
0 =  | ()|2 |0 ()|2 ̃( )
2
Z
1
0 = ̃( )†0 ()† ()0 () ()
2

The exchange interaction   acts as a Hersenberg interaction:

X X 1
0 + + 
 0 0 0 0  = −2 0 ( · 0 +  · 0 )
4

The positivity of 0 can be proved as follows:

(1) Complete screening:

 = ( − )

In the case,
Z
1
0 =  | ()|2 |0 ()|2  0
2

(2) Long-ranged Coulomb interaction

2
̃ =
| −  |

Assume  is the plane wave


Z
2
0 ∝  exp[ · ]
||

2
= 4 0
2

This is ferromagnetic!

355
CHAPTER 7 – MANUSCRIPT

B 7.4.4 Antiferromagnetic Exchange

C Two-site (or atom) problem with two electrons

Let’s consider two orthogonal orbitals localized on two atoms labelled by  = 1 2.

Tunnelling between the two atoms (or states) is described by a hopping Hamiltonian

X
  = − +
(1 +
2 + 2 1 )

For simplicity, we consider an on-site interaction

X
 =  ↑ · ↓

To explore the physical origin of anti-ferromagnetic coupling, we consider a special case:

  . In the case, we choose  to be the zero-order(or unperturbed) Hamiltonian

and  + to be the perturbation.

For : there are six possible configurations which are eigenstates of H 

(1)  = 0, |1 ↑ 2 ↑i  |1 ↑ 2 ↓i  |1 ↓ 2 ↑i  |1 ↓ 2 ↓i ;

(2)  = , |1 ↑ 1 ↓i  |2 ↓ 2 ↑i

Denote |i the unperturbed state with energy 0 and |i denote the two state

with E=U (n=1,2). In terms of the c-operators,

|i = + +
1 2 |0i

356
CHAPTER 7 – MANUSCRIPT

and

† †
|1i = 1↑ 1↓ |0i

† †
|2i = 2↑ 2↓ |0i

In the first-order perturbation theory

h | | 0 i = 0

In the second-order perturbation theory

­ ¯ ¯ ® P h |  | i h |  | i
 ¯∆ (2) ¯  =
=12 h | | i − h | | i

1 X ­ ¯¯  ¯¯ ® ­ ¯¯  ¯¯ ®
= −      

|i h| is a projection operator

|1i h1| = 1↑ 1↓ (1 − 2↑ )(1 − 2↓ )(↑↓ −)

|2i h2| = (1 − 1↑ )(1 − 1↓ )2↑ 2↓ (− ↑↓)

357
CHAPTER 7 – MANUSCRIPT
­ ¯ ¯ ® ­ ¯ ¯ 0®
 ¯ ¯ 1 1 ¯ ¯ 

†  † 
= h| (2↑ 1↑ + 2↓ 1↓ + ~)

×1↑ 1↓ (1 − 2↑ )(1 − 2↓ )

†  † 
(2↑ 1↑ + 2↓ 1↓ + ~) |0 i

†  † 
= h| (2↑ 1↑ + 2↓ 1↓ )1↑ 1↓ (1 − 2↑ )(1 − 2↓ )

†  † 
×(1↑ 2↑ + 1↓ 2↓ ) |0 i

†  †  †  † 
= h| (−1↓ 1↑ 2↑ 1↓ − 1↑ 1↓ 2↓ 2↑ ) + 1↑ 2↓

+1↓ 2↑ ) |0 i

1
= h| − 21 · 2 + (1↑ + 1↓ )(2↑ + 2↓ ) |0 i
2

Therefore the effective Hamiltonian ∆ (2) can be written as an isotropic antiferromag-

netic Hersenberg spin exchange form.

42   1
∆ (2) = + (1 · 2 − )
 4

As 42   0, the exchange coupling is antiferromagnetic! The ground state of this

two-site problem is spin singlet, i.e.  = 0. Our discussion on the two-site problem

can be easily generalized to a many-site system. The Hersenberg model is defined on a

358
CHAPTER 7 – MANUSCRIPT

lattice
X
=   · 

where  and  are the lattice sites and usually are of the nearest neighbour pair.  can

be taken any value of half-integer,  = 12 1

One of the most important application of spin superexchange with current

interests is the high temperature superconductivity. the so-called “ − ” model is

extensively discussed over the last decade.

X 42 X   1
 = − †  + ( ·  −   )
  4
hi

which is limited within the Hilbert space excluding double occupancy of electrons on

the same site.

359

También podría gustarte