Вы находитесь на странице: 1из 773

Condensed Matter Physics I

Peter S. Riseborough
February 5, 2015

Contents
1 Introduction 10
1.1 The Born-Oppenheimer Approximation . . . . . . . . . . . . . . 10

2 Crystallography 18
2.0.1 Exercise 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3 Structures 20
3.1 Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.2 Crystalline Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3 The Direct Lattice . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.3.1 Primitive Unit Cells . . . . . . . . . . . . . . . . . . . . . 30
3.3.2 The Wigner-Seitz Unit Cell . . . . . . . . . . . . . . . . . 31
3.4 Symmetry of Crystals . . . . . . . . . . . . . . . . . . . . . . . . 33
3.4.1 Symmetry Groups . . . . . . . . . . . . . . . . . . . . . . 34
3.4.2 Group Multiplication Tables . . . . . . . . . . . . . . . . 35
3.4.3 Point Group Operations . . . . . . . . . . . . . . . . . . . 37
3.4.4 Exercise 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.4.5 Limitations Imposed by Translational Symmetry . . . . . 41
3.4.6 Exercise 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4.7 Point Group Nomenclature . . . . . . . . . . . . . . . . . 44
3.5 Bravais Lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5.1 Exercise 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.5.2 Cubic Bravais Lattices. . . . . . . . . . . . . . . . . . . . 53
3.5.3 Tetragonal Bravais Lattices. . . . . . . . . . . . . . . . . . 58
3.5.4 Orthorhombic Bravais Lattices. . . . . . . . . . . . . . . . 59
3.5.5 Monoclinic Bravais Lattice. . . . . . . . . . . . . . . . . . 60
3.5.6 Triclinic Bravais Lattice. . . . . . . . . . . . . . . . . . . . 61
3.5.7 Trigonal Bravais Lattice. . . . . . . . . . . . . . . . . . . . 63
3.5.8 Hexagonal Bravais Lattice. . . . . . . . . . . . . . . . . . 66
3.5.9 Exercise 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
3.6 Point Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.6.1 Exercise 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
3.7 Space Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

1
3.7.1 Exercise 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
3.8 Crystal Structures with Bases. . . . . . . . . . . . . . . . . . . . 78
3.8.1 Diamond Structure . . . . . . . . . . . . . . . . . . . . . . 78
3.8.2 Exercise 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
3.8.3 Graphite Structure . . . . . . . . . . . . . . . . . . . . . . 79
3.8.4 Exercise 9 . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
3.8.5 Hexagonal Close-Packed Structure . . . . . . . . . . . . . 83
3.8.6 Exercise 10 . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.8.7 Exercise 11 . . . . . . . . . . . . . . . . . . . . . . . . . . 86
3.8.8 Other Close-Packed Structures . . . . . . . . . . . . . . . 86
3.8.9 Sodium Chloride Structure . . . . . . . . . . . . . . . . . 88
3.8.10 Cesium Chloride Structure . . . . . . . . . . . . . . . . . 90
3.8.11 Fluorite Structure . . . . . . . . . . . . . . . . . . . . . . 92
3.8.12 The Copper Three Gold Structure . . . . . . . . . . . . . 93
3.8.13 Rutile Structure . . . . . . . . . . . . . . . . . . . . . . . 94
3.8.14 Exercise 12 . . . . . . . . . . . . . . . . . . . . . . . . . . 95
3.8.15 Zinc Blende Structure . . . . . . . . . . . . . . . . . . . . 95
3.8.16 Zincite Structure . . . . . . . . . . . . . . . . . . . . . . . 96
3.8.17 Exercise 13 . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.8.18 The Perovskite Structure . . . . . . . . . . . . . . . . . . 98
3.8.19 Exercise 14 . . . . . . . . . . . . . . . . . . . . . . . . . . 99
3.9 Lattice Planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
3.9.1 Exercise 15 . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.9.2 Exercise 16 . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.9.3 Exercise 17 . . . . . . . . . . . . . . . . . . . . . . . . . . 104
3.10 Quasi-Crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

4 Structure Determination 112


4.1 X Ray Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4.1.1 The Bragg condition . . . . . . . . . . . . . . . . . . . . . 112
4.1.2 The Laue conditions . . . . . . . . . . . . . . . . . . . . . 114
4.1.3 Equivalence of the Bragg and Laue conditions . . . . . . . 118
4.1.4 The Ewald Construction . . . . . . . . . . . . . . . . . . . 119
4.1.5 X-ray Techniques . . . . . . . . . . . . . . . . . . . . . . . 120
4.1.6 Exercise 18 . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.1.7 The Structure and Form Factors . . . . . . . . . . . . . . 125
4.1.8 Exercise 19 . . . . . . . . . . . . . . . . . . . . . . . . . . 133
4.1.9 Exercise 20 . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.1.10 Exercise 21 . . . . . . . . . . . . . . . . . . . . . . . . . . 136
4.1.11 Exercise 22 . . . . . . . . . . . . . . . . . . . . . . . . . . 136
4.1.12 Exercise 23 . . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.1.13 Exercise 24 . . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.1.14 Exercise 25 . . . . . . . . . . . . . . . . . . . . . . . . . . 140
4.2 Neutron Diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . 141
4.2.1 Exercise 26 . . . . . . . . . . . . . . . . . . . . . . . . . . 143
4.3 Theory of the Differential Scattering Cross-section . . . . . . . . 145

2
4.3.1 Time Dependent Perturbation Theory . . . . . . . . . . . 146
4.3.2 The Fermi Golden Rule . . . . . . . . . . . . . . . . . . . 148
4.3.3 The Elastic Scattering Cross-Section . . . . . . . . . . . . 150
4.3.4 The Condition for Coherent Scattering . . . . . . . . . . . 152
4.3.5 Exercise 27 . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.3.6 Exercise 28 . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.3.7 Exercise 29 . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4.3.8 Anti-Domain Phase Boundaries . . . . . . . . . . . . . . . 156
4.3.9 Exercise 30 . . . . . . . . . . . . . . . . . . . . . . . . . . 157
4.4 Elastic Scattering from Quasi-Crystals . . . . . . . . . . . . . . . 158
4.5 Elastic Scattering from a Fluid . . . . . . . . . . . . . . . . . . . 162

5 The Reciprocal Lattice 164


5.0.1 Exercise 31 . . . . . . . . . . . . . . . . . . . . . . . . . . 165
5.1 The Reciprocal Lattice as a Dual Lattice . . . . . . . . . . . . . . 165
5.1.1 Exercise 32 . . . . . . . . . . . . . . . . . . . . . . . . . . 169
5.2 Examples of Reciprocal Lattices . . . . . . . . . . . . . . . . . . . 169
5.2.1 The Simple Cubic Reciprocal Lattice . . . . . . . . . . . . 169
5.2.2 The Body Centered Cubic Reciprocal Lattice . . . . . . . 169
5.2.3 The Face Centered Cubic Reciprocal Lattice . . . . . . . 170
5.2.4 The Hexagonal Reciprocal Lattice . . . . . . . . . . . . . 171
5.2.5 Exercise 33 . . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.3 The Brillouin Zones . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.3.1 The Simple Cubic Brillouin Zone . . . . . . . . . . . . . . 173
5.3.2 The Body Centered Cubic Brillouin Zone . . . . . . . . . 175
5.3.3 The Face Centered Cubic Brillouin Zone . . . . . . . . . . 177
5.3.4 The Hexagonal Brillouin Zone . . . . . . . . . . . . . . . . 178
5.3.5 The Trigonal Brillouin Zone . . . . . . . . . . . . . . . . . 180

6 Electrons 183

7 Electronic States 183


7.1 Many-Electron Wave Functions . . . . . . . . . . . . . . . . . . . 184
7.1.1 Exercise 34 . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.2 Bloch’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.3 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . 195
7.4 Plane Wave Expansion of Bloch Functions . . . . . . . . . . . . . 198
7.5 The Bloch Wave Vector . . . . . . . . . . . . . . . . . . . . . . . 202
7.6 The Density of States . . . . . . . . . . . . . . . . . . . . . . . . 203
7.6.1 Exercise 35 . . . . . . . . . . . . . . . . . . . . . . . . . . 206
7.7 The Fermi Surface . . . . . . . . . . . . . . . . . . . . . . . . . . 207

3
8 Approximate Models 211
8.1 The Nearly-Free Electron Model . . . . . . . . . . . . . . . . . . 211
8.1.1 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . 212
8.1.2 Non-Degenerate Perturbation Theory . . . . . . . . . . . 213
8.1.3 Degenerate Perturbation Theory . . . . . . . . . . . . . . 215
8.1.4 Empty Lattice Approximation Band Structure . . . . . . 222
8.1.5 Exercise 36 . . . . . . . . . . . . . . . . . . . . . . . . . . 230
8.1.6 Degeneracies of the Bloch States . . . . . . . . . . . . . . 231
8.1.7 Exercise 37 . . . . . . . . . . . . . . . . . . . . . . . . . . 241
8.1.8 Exercise 38 . . . . . . . . . . . . . . . . . . . . . . . . . . 242
8.1.9 Brillouin Zone Boundaries and Fermi Surfaces . . . . . . . 242
8.1.10 The Geometric Structure Factor . . . . . . . . . . . . . . 248
8.1.11 Exercise 39 . . . . . . . . . . . . . . . . . . . . . . . . . . 258
8.1.12 Exercise 40 . . . . . . . . . . . . . . . . . . . . . . . . . . 260
8.1.13 Exercise 41 . . . . . . . . . . . . . . . . . . . . . . . . . . 261
8.1.14 Exercise 42 . . . . . . . . . . . . . . . . . . . . . . . . . . 262
8.1.15 Exercise 43 . . . . . . . . . . . . . . . . . . . . . . . . . . 263
8.2 The Pseudo-Potential Method . . . . . . . . . . . . . . . . . . . . 264
8.2.1 The Pseudo-Potential Theorem . . . . . . . . . . . . . . . 267
8.2.2 The Cancellation Theorem . . . . . . . . . . . . . . . . . 269
8.2.3 The Scattering Approach . . . . . . . . . . . . . . . . . . 272
8.2.4 The Ziman-Lloyd Pseudo-potential . . . . . . . . . . . . . 274
8.2.5 Exercise 44 . . . . . . . . . . . . . . . . . . . . . . . . . . 276
8.2.6 Exercise 45 . . . . . . . . . . . . . . . . . . . . . . . . . . 276
8.2.7 Exercise 46 . . . . . . . . . . . . . . . . . . . . . . . . . . 277
8.3 The Tight-Binding Model . . . . . . . . . . . . . . . . . . . . . . 278
8.3.1 Tight-Binding s Band Metal . . . . . . . . . . . . . . . . . 286
8.3.2 Tight-Binding Bands of Diamond Structured Semicon-
ductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
8.3.3 Exercise 47 . . . . . . . . . . . . . . . . . . . . . . . . . . 293
8.3.4 Exercise 48 . . . . . . . . . . . . . . . . . . . . . . . . . . 294
8.3.5 Exercise 49 . . . . . . . . . . . . . . . . . . . . . . . . . . 296
8.3.6 Exercise 50 . . . . . . . . . . . . . . . . . . . . . . . . . . 296
8.3.7 Exercise 51 . . . . . . . . . . . . . . . . . . . . . . . . . . 297
8.3.8 Exercise 52 . . . . . . . . . . . . . . . . . . . . . . . . . . 297
8.3.9 Wannier Functions . . . . . . . . . . . . . . . . . . . . . . 297
8.3.10 Exercise 53 . . . . . . . . . . . . . . . . . . . . . . . . . . 301
8.3.11 Exercise 54 . . . . . . . . . . . . . . . . . . . . . . . . . . 301
8.3.12 Example of Tight-Binding: Graphene . . . . . . . . . . . 301

9 Electron-Electron Interactions 307


9.1 The Landau Fermi Liquid . . . . . . . . . . . . . . . . . . . . . . 307
9.1.1 The Scattering Rate . . . . . . . . . . . . . . . . . . . . . 309
9.1.2 The Quasi-Particle Energy . . . . . . . . . . . . . . . . . 316
9.1.3 Exercise 55 . . . . . . . . . . . . . . . . . . . . . . . . . . 319
9.2 The Hartree-Fock Approximation . . . . . . . . . . . . . . . . . . 320

4
9.2.1 The Free Electron Gas. . . . . . . . . . . . . . . . . . . . 324
9.2.2 Exercise 56 . . . . . . . . . . . . . . . . . . . . . . . . . . 342
9.3 The Density Functional Method . . . . . . . . . . . . . . . . . . . 344
9.3.1 Hohenberg-Kohn Theorem . . . . . . . . . . . . . . . . . . 345
9.3.2 Functionals and Functional Derivatives . . . . . . . . . . 346
9.3.3 The Variational Principle . . . . . . . . . . . . . . . . . . 350
9.3.4 The Electrostatic Terms . . . . . . . . . . . . . . . . . . . 352
9.3.5 The Kohn-Sham Equations . . . . . . . . . . . . . . . . . 353
9.3.6 The Local Density Approximation . . . . . . . . . . . . . 355
9.4 Static Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
9.4.1 The Thomas-Fermi Approximation . . . . . . . . . . . . . 361
9.4.2 Linear Response Theory . . . . . . . . . . . . . . . . . . . 364
9.4.3 Density Functional Response Function . . . . . . . . . . . 368
9.4.4 Exercise 57 . . . . . . . . . . . . . . . . . . . . . . . . . . 371
9.4.5 Exercise 58 . . . . . . . . . . . . . . . . . . . . . . . . . . 372

10 Stability of Structures 377


10.1 Momentum Space Representation . . . . . . . . . . . . . . . . . . 377
10.2 Real Space Representation . . . . . . . . . . . . . . . . . . . . . . 384

11 Metals 392
11.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 392
11.1.1 The Sommerfeld Expansion . . . . . . . . . . . . . . . . . 393
11.1.2 The Specific Heat Capacity . . . . . . . . . . . . . . . . . 395
11.1.3 Exercise 59 . . . . . . . . . . . . . . . . . . . . . . . . . . 398
11.1.4 Exercise 60 . . . . . . . . . . . . . . . . . . . . . . . . . . 398
11.1.5 Pauli Paramagnetism . . . . . . . . . . . . . . . . . . . . 398
11.1.6 Exercise 61 . . . . . . . . . . . . . . . . . . . . . . . . . . 402
11.1.7 Exercise 62 . . . . . . . . . . . . . . . . . . . . . . . . . . 402
11.1.8 Landau Diamagnetism . . . . . . . . . . . . . . . . . . . . 402
11.1.9 Landau Level Quantization . . . . . . . . . . . . . . . . . 403
11.1.10 The Diamagnetic Susceptibility . . . . . . . . . . . . . . . 405
11.2 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . 408
11.2.1 Electrical Conductivity . . . . . . . . . . . . . . . . . . . 408
11.2.2 Scattering by Static Defects . . . . . . . . . . . . . . . . . 408
11.2.3 Exercise 63 . . . . . . . . . . . . . . . . . . . . . . . . . . 415
11.2.4 The Hall Effect and Magneto-resistance. . . . . . . . . . . 416
11.2.5 Multi-band Models . . . . . . . . . . . . . . . . . . . . . . 424
11.3 Electromagnetic Properties of Metals . . . . . . . . . . . . . . . . 427
11.3.1 The Longitudinal Response . . . . . . . . . . . . . . . . . 430
11.3.2 Electron Scattering Experiments . . . . . . . . . . . . . . 440
11.3.3 Exercise 64 . . . . . . . . . . . . . . . . . . . . . . . . . . 445
11.3.4 Exercise 65 . . . . . . . . . . . . . . . . . . . . . . . . . . 447
11.3.5 The Transverse Response . . . . . . . . . . . . . . . . . . 452
11.3.6 Optical Experiments . . . . . . . . . . . . . . . . . . . . . 457
11.3.7 Kramers-Kronig Relation . . . . . . . . . . . . . . . . . . 459

5
11.3.8 Exercise 66 . . . . . . . . . . . . . . . . . . . . . . . . . . 460
11.3.9 Exercise 67 . . . . . . . . . . . . . . . . . . . . . . . . . . 461
11.3.10 The Drude Conductivity . . . . . . . . . . . . . . . . . . . 462
11.3.11 Exercise 68 . . . . . . . . . . . . . . . . . . . . . . . . . . 466
11.3.12 Exercise 69 . . . . . . . . . . . . . . . . . . . . . . . . . . 468
11.3.13 The Anomalous Skin Effect . . . . . . . . . . . . . . . . . 468
11.3.14 Inter-Band Transitions . . . . . . . . . . . . . . . . . . . . 472
11.4 The Fermi Surface . . . . . . . . . . . . . . . . . . . . . . . . . . 474
11.4.1 Semi-Classical Orbits . . . . . . . . . . . . . . . . . . . . 474
11.4.2 de Haas - van Alphen Oscillations . . . . . . . . . . . . . 478
11.4.3 Exercise 70 . . . . . . . . . . . . . . . . . . . . . . . . . . 481
11.4.4 The Lifshitz-Kosevich Formulae . . . . . . . . . . . . . . . 481
11.4.5 Geometric Resonances . . . . . . . . . . . . . . . . . . . . 487
11.4.6 Cyclotron Resonances . . . . . . . . . . . . . . . . . . . . 489
11.5 The Quantum Hall Effect . . . . . . . . . . . . . . . . . . . . . . 494
11.5.1 The Integer Quantum Hall Effect . . . . . . . . . . . . . . 495
11.5.2 Exercise 71 . . . . . . . . . . . . . . . . . . . . . . . . . . 504
11.5.3 Exercise 72 . . . . . . . . . . . . . . . . . . . . . . . . . . 505
11.5.4 The Fractional Quantum Hall Effect . . . . . . . . . . . . 506
11.5.5 Quasi-Particle Excitations . . . . . . . . . . . . . . . . . . 508
11.5.6 Skyrmions . . . . . . . . . . . . . . . . . . . . . . . . . . . 512
11.5.7 Composite Fermions . . . . . . . . . . . . . . . . . . . . . 520

12 Insulators and Semiconductors 523


12.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 528
12.1.1 Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
12.1.2 Intrinsic Semiconductors . . . . . . . . . . . . . . . . . . . 531
12.1.3 Extrinsic Semiconductors . . . . . . . . . . . . . . . . . . 533
12.1.4 Exercise 73 . . . . . . . . . . . . . . . . . . . . . . . . . . 536
12.2 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . 536
12.3 Optical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 537

13 Phonons 538

14 Harmonic Phonons 538


14.1 Lattice with a Basis . . . . . . . . . . . . . . . . . . . . . . . . . 548
14.2 A Sum Rule for the Dispersion Relations . . . . . . . . . . . . . . 549
14.2.1 Exercise 74 . . . . . . . . . . . . . . . . . . . . . . . . . . 551
14.3 The Nature of the Phonon Modes . . . . . . . . . . . . . . . . . . 552
14.3.1 Exercise 75 . . . . . . . . . . . . . . . . . . . . . . . . . . 554
14.3.2 Exercise 76 . . . . . . . . . . . . . . . . . . . . . . . . . . 555
14.3.3 Exercise 77 . . . . . . . . . . . . . . . . . . . . . . . . . . 556
14.3.4 Exercise 78 . . . . . . . . . . . . . . . . . . . . . . . . . . 556
14.3.5 Exercise 79 . . . . . . . . . . . . . . . . . . . . . . . . . . 556
14.3.6 Exercise 80 . . . . . . . . . . . . . . . . . . . . . . . . . . 557
14.3.7 Exercise 81 . . . . . . . . . . . . . . . . . . . . . . . . . . 558

6
14.4 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
14.4.1 The Specific Heat . . . . . . . . . . . . . . . . . . . . . . 562
14.4.2 The Einstein Model of a Solid . . . . . . . . . . . . . . . . 563
14.4.3 The Debye Model of a Solid . . . . . . . . . . . . . . . . . 564
14.4.4 Exercise 82 . . . . . . . . . . . . . . . . . . . . . . . . . . 566
14.4.5 Exercise 83 . . . . . . . . . . . . . . . . . . . . . . . . . . 567
14.4.6 Exercise 84 . . . . . . . . . . . . . . . . . . . . . . . . . . 567
14.4.7 Exercise 85 . . . . . . . . . . . . . . . . . . . . . . . . . . 568
14.4.8 Lindemann Theory of Melting . . . . . . . . . . . . . . . . 568
14.4.9 Thermal Expansion . . . . . . . . . . . . . . . . . . . . . 572
14.4.10 Thermal Expansion of Metals . . . . . . . . . . . . . . . . 573
14.5 Anharmonicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
14.5.1 Exercise 86 . . . . . . . . . . . . . . . . . . . . . . . . . . 575

15 Phonon Measurements 577


15.1 Inelastic Neutron Scattering . . . . . . . . . . . . . . . . . . . . . 577
15.2 The Scattering Cross-Section . . . . . . . . . . . . . . . . . . . . 578
15.2.1 The Zero-Phonon Scattering Process . . . . . . . . . . . . 582
15.3 The Debye-Waller Factor . . . . . . . . . . . . . . . . . . . . . . 583
15.3.1 The One-Phonon Scattering Processes. . . . . . . . . . . . 585
15.3.2 Multi-Phonon Scattering . . . . . . . . . . . . . . . . . . . 590
15.3.3 Exercise 87 . . . . . . . . . . . . . . . . . . . . . . . . . . 592
15.3.4 Exercise 88 . . . . . . . . . . . . . . . . . . . . . . . . . . 593
15.3.5 Exercise 89 . . . . . . . . . . . . . . . . . . . . . . . . . . 593
15.4 Raman and Brillouin Scattering of Light . . . . . . . . . . . . . . 593

16 Phonons in Metals 599


16.1 Screened Ionic Plasmons . . . . . . . . . . . . . . . . . . . . . . . 600
16.1.1 Kohn Anomalies . . . . . . . . . . . . . . . . . . . . . . . 601
16.2 Dielectric Constant of a Metal . . . . . . . . . . . . . . . . . . . . 602
16.3 The Retarded Electron-Electron Interaction . . . . . . . . . . . . 604
16.4 Phonon Renormalization of Quasi-Particles . . . . . . . . . . . . 605
16.5 Electron-Phonon Interactions . . . . . . . . . . . . . . . . . . . . 608
16.6 Electrical Resistivity due to Phonon Scattering . . . . . . . . . . 609
16.6.1 Umklapp Scattering . . . . . . . . . . . . . . . . . . . . . 614
16.6.2 Phonon Drag . . . . . . . . . . . . . . . . . . . . . . . . . 615

17 Phonons in Semiconductors 616


17.1 Resistivity due to Phonon Scattering . . . . . . . . . . . . . . . . 616
17.2 Polarons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 617
17.3 Indirect Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . 617

7
18 Impurities and Disorder 619
18.1 Scattering by Impurities . . . . . . . . . . . . . . . . . . . . . . . 625
18.1.1 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . 632
18.2 Virtual Bound States . . . . . . . . . . . . . . . . . . . . . . . . . 633
18.2.1 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . . 636
18.3 Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637
18.4 Coherent Potential Approximation . . . . . . . . . . . . . . . . . 638
18.4.1 Exercise . . . . . . . . . . . . . . . . . . . . . . . . . . . 640
18.5 Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 641
18.5.1 Anderson Model of Localization . . . . . . . . . . . . . . . 642
18.5.2 Scaling Theories of Localization . . . . . . . . . . . . . . . 644

19 Magnetic Impurities 647


19.1 Localized Magnetic Impurities in Metals . . . . . . . . . . . . . . 647
19.2 Mean-Field Approximation . . . . . . . . . . . . . . . . . . . . . 647
19.3 The Atomic Limit . . . . . . . . . . . . . . . . . . . . . . . . . . 652
19.4 The Schrieffer-Wolf Transformation . . . . . . . . . . . . . . . . . 656
19.4.1 The Kondo Hamiltonian . . . . . . . . . . . . . . . . . . . 659
19.5 The Resistance Minimum . . . . . . . . . . . . . . . . . . . . . . 660

20 Collective Phenomenon 668

21 Itinerant Magnetism 668


21.1 Stoner Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 668
21.1.1 Exercise 90 . . . . . . . . . . . . . . . . . . . . . . . . . . 671
21.1.2 Exercise 91 . . . . . . . . . . . . . . . . . . . . . . . . . . 671
21.2 Linear Response Theory . . . . . . . . . . . . . . . . . . . . . . . 671
21.3 Magnetic Instabilities . . . . . . . . . . . . . . . . . . . . . . . . 673
21.4 Spin Fluctuations near Ferromagnetic Instabilities . . . . . . . . 677
21.4.1 Ferromagnetic Spin Waves . . . . . . . . . . . . . . . . . . 680
21.5 The Slater-Pauling Curves . . . . . . . . . . . . . . . . . . . . . . 685
21.6 The Heisenberg Model . . . . . . . . . . . . . . . . . . . . . . . . 687

22 Localized Magnetism 688


22.1 Holstein-Primakoff Transformation . . . . . . . . . . . . . . . . . 690
22.2 Spin Rotational Invariance . . . . . . . . . . . . . . . . . . . . . . 694
22.2.1 Exercise 92 . . . . . . . . . . . . . . . . . . . . . . . . . . 697
22.3 Anti-ferromagnetic Spinwaves . . . . . . . . . . . . . . . . . . . . 697
22.3.1 Exercise 93 . . . . . . . . . . . . . . . . . . . . . . . . . . 700

23 Spin Glasses 702


23.1 Mean-Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 706
23.2 The Sherrington-Kirkpatrick Solution. . . . . . . . . . . . . . . . 708

8
24 Magnetic Neutron Scattering 711
24.1 The Inelastic Scattering Cross-Section . . . . . . . . . . . . . . . 711
24.1.1 The Dipole-Dipole Interaction . . . . . . . . . . . . . . . . 711
24.1.2 The Inelastic Scattering Cross-Section . . . . . . . . . . . 711
24.2 Time-Dependent Spin Correlation Functions . . . . . . . . . . . . 715
24.3 The Fluctuation - Dissipation Theorem . . . . . . . . . . . . . . 718
24.4 Magnetic Scattering . . . . . . . . . . . . . . . . . . . . . . . . . 720
24.4.1 Neutron Diffraction . . . . . . . . . . . . . . . . . . . . . 720
24.4.2 Exercise 94 . . . . . . . . . . . . . . . . . . . . . . . . . . 722
24.4.3 Exercise 95 . . . . . . . . . . . . . . . . . . . . . . . . . . 722
24.4.4 Spin Wave Scattering . . . . . . . . . . . . . . . . . . . . 723
24.4.5 Exercise 96 . . . . . . . . . . . . . . . . . . . . . . . . . . 724
24.4.6 Critical Scattering . . . . . . . . . . . . . . . . . . . . . . 724

25 Superconductivity 726
25.1 Experimental Manifestation . . . . . . . . . . . . . . . . . . . . . 727
25.1.1 The London Equations . . . . . . . . . . . . . . . . . . . . 729
25.1.2 Thermodynamics of the Superconducting State . . . . . . 730
25.2 The Cooper Problem . . . . . . . . . . . . . . . . . . . . . . . . . 733
25.3 Pairing Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 738
25.3.1 The Pairing Interaction . . . . . . . . . . . . . . . . . . . 738
25.3.2 The B.C.S. Variational State . . . . . . . . . . . . . . . . 740
25.3.3 The Gap Equation . . . . . . . . . . . . . . . . . . . . . . 742
25.3.4 The Ground State Energy . . . . . . . . . . . . . . . . . . 745
25.4 Quasi-Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 748
25.4.1 Exercise 97 . . . . . . . . . . . . . . . . . . . . . . . . . . 752
25.5 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 752
25.6 Perfect Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . 756
25.7 The Meissner Effect . . . . . . . . . . . . . . . . . . . . . . . . . 758
25.8 Landau-Ginzburg Theory . . . . . . . . . . . . . . . . . . . . . . 760
25.8.1 Extremal Configurations . . . . . . . . . . . . . . . . . . . 764
25.8.2 Characteristic Length Scales . . . . . . . . . . . . . . . . 765
25.8.3 The Surface Energy . . . . . . . . . . . . . . . . . . . . . 768
25.8.4 The Little-Parks Experiment . . . . . . . . . . . . . . . . 770
25.8.5 The Critical Current . . . . . . . . . . . . . . . . . . . . . 772

9
1 Introduction
Condensed Matter Physics is the study of materials in Solid and Liquid Phases.
It encompasses the study of ordered crystalline phases of solids, as well as disor-
dered phases such as the amorphous and glassy phases of solids. Furthermore,
it also includes materials with short-ranged order such as conventional liquids
and liquid crystals which show unconventional order intermediate between those
of a crystalline solid and a liquid. Condensed matter has the quite remarkable
property that, due to the large number of particles involved, the behavior of
the materials may be qualitatively distinct from those of the individual con-
stituents. The behavior of the incredibly large number of particles is governed
by (quantum) statistics which, through the chaotically complicated motion of
the particles, produces new types of order. These emergent phenomena are best
exemplified in phenomenon such as magnetism or superconductivity where the
collective behavior results in transitions to new phases.

In surveying the properties of materials, it is convenient to separate the


properties according to two (usually) disparate time scales. One time scale is
a slow time scale which governs the structural dynamics and the other is a
faster time scale that governs the electronic motion. The large difference be-
tween the time scales is due to the large ratio of the nuclear masses to the
electronic mass, MN /me ∼ 103 . The long-ranged electromagnetic force binds
these two constituents of different mass into electrically neutral material. The
slow moving nuclear masses can be considered to be quasi-static, and are re-
sponsible for defining the structure of matter. In this approximation, the fast
moving electrons equilibrate in the quasi-static potential produced by the nuclei.

1.1 The Born-Oppenheimer Approximation


The difference in the relevant time scales for electronic and nuclear motion
allows one to make the Born-Oppenheimer Approximation1 . In this approxi-
mation, the electronic states are treated as if the nuclei were at rest at fixed
positions. However, when treating the slow motions of the nuclei, the electrons
are considered as adapting instantaneously to the potential of the charged nu-
clei, thereby minimizing the electronic energies. Thus, the nuclei charges are
dressed by a cloud of electrons forming ionic or atomic-like aggregates.

A qualitative estimate of the relative energies of nuclear versus electronic


motion can be obtained by considering metallic hydrogen. The electronic en-
ergies are calculated using only the Bohr model of the hydrogen atom. The
equation of motion for an electron of mass me has the form

Z e2 me v 2
− = − (1)
a2 a
1 M. Born and R. Oppenheimer, Ann. Phys. (Leipzig), 84, 457 (1927).

10
where Z is the nuclear charge and a is the radius of the atomic orbital. The stan-
dard semi-classical quantization condition due to Bohr and Sommerfeld restricts
the angular momentum to integral values of h̄

me v a = n h̄ (2)

These equations can be combined to find the Bohr radius as

h̄2
a = n2 (3)
me Z e2
and also the quantized total electronic energy of the hydrogen atom

Z e2 me v 2
Ee = − +
a 2
Z e2
= −
2a
me Z 2 e4
= − (4)
2 n2 h̄2
which are standard results from atomic physics. Note that the kinetic energy
term and the electrostatic potential term have similar magnitudes.

Now consider the motion of the nuclei. The forces consist of Coulomb forces
between the nuclei and electrons and the quantum mechanical Pauli forces. The
electrostatic repulsions and attractions have similar magnitudes since the inter-
nuclear separations are of the same order as the Bohr radius. In equilibrium,
the sum of the forces vanish identically. Furthermore, if an atom is displaced
from the equilibrium position by a small distance equal to r, the restoring force
is approximately given by the dipole force

Z 2 e2
−α r (5)
a3
where α is a dimensionless constant. Hence, the equation of motion for the
displacement of a nuclei of mass MN is

Z 2 e2 d2 r
−α r = M N (6)
a3 dt2
which shows that the nuclei undergo harmonic oscillations with frequency

Z 2 e2
ω2 = α (7)
MN a3
The semi-classical quantization condition
I
MN dr . v = 2 π n h̄ (8)

11
yields the energy for nuclear motion as

EN = n h̄ ω
 12
me Z 2 e4 1

Z me
= n α2 (9)
h̄2 MN

where we have substituted the expression for the Bohr radius for a in the ex-
pression for ω. Thus, the ratio of the energies of nuclear motion to electronic
motion are given by the factor
  12
EN me
∼ (10)
Ee MN
1
Since the ratio of the mass of electron to the proton mass is 2000 , the nuclear
kinetic energy is negligible when compared to the electronic kinetic energy. A
more rigorous proof of the validity of the Born-Oppenheimer approximation was
given by Migdal2 .

—————————————————————————————————-

Example: Beyond the Born-Oppenheimer Approximation

An example of a correction to the Born-Oppenheimer approximation is given


by (incoherent) inelastic neutron-atom scattering. The scattering occurs only
via the nuclear force between the neutrons and the nucleus. The electrons are
bound to the nucleus and are only excited via an indirect process. We intend to
show that the probability that an electronic transition occurs is governed by he
ratio of the mass of the electron me to the mass of the nucleus MN . First, we
shall consider the elastic-scattering process in the Born-Oppenheimer approxi-
mation and then we shall consider inelastic-scattering.

In the consideration of the elastic-scattering process, the electronic states


may be ignored since the electrons are not excited. Furthermore, consistent with
the Born-Oppenheimer approximation, the mass of the electron is neglected in
comparison with the mass of the nuclei when we consider the kinetics of the
scattering. Therefore, in an incoherent scattering process, a neutron of mass
mn is scattered from an individual nucleus of mass MN . The elastic-scattering
process satisfies conservation of energy and momentum
mn 2 MN mn 2 MN 2
v + V 2i = v + Vf
2 i 2 2 f 2
mn v i + MN V i = mn v f + MN V f (11)

The scattering results in a transfer of energy and momentum between the two
particles. The energy and momentum transferred between the neutron and the
2 A. B. Migdal, Sov. Phys. J.E.T.P. 7, 996 (1958).

12
Nuclear Compton Scattering

pf

θ
mn pi

Pi
MN Pf

Figure 1: The scattering of a neutron of mass mn by an atom with a nucleus of


mass MN .

nucleus is given by

h̄ q = mn ( v i − v f )
mn
h̄ ω = ( v 2i − v 2f )
2
h̄2
= h̄ q . v i − q2 (12)
2 mn
The neutron’s energy and momentum loss can be determined by experiment.
If the nucleus is initially at rest, the magnitude of the final momentum of the
neutron is related to the initial momentum via the scattering angle θ, via
q
2
cos θ + (M N 2
mn ) − sin θ
vf = vi (13)
1 + (M mn )
N

It is seen that, if M
mn
N
 1, the scattering is elastic. The positioning of the
detector selects the neutrons which are scattered through the angle θ. The scat-
tering of the neutrons by the nuclei is characterized by the correlation function
S(q, ω) which embodies the above conservation laws in a delta function factor

h̄2 2
 

S(q, ω) ∝ δ h̄ω − q . pi + q (14)
mn 2mn

13
The correlation function can be directly expressed in terms of the initial prop-
erties of the nuclei via
h̄2
Z  
3 h̄ 2
S(q, ω) = d P i n(P i ) δ h̄ω − q . Pi − q (15)
MN 2MN
where n(P ) is the initial momentum distribution of the nuclei, which is given
by
n(P ) = | φ(P ) |2 (16)
and where φ(P ) is the nuclear wave function in the momentum representation
Z  
1 3 i
φ(P ) = 3 d R exp − P . R χ(R) (17)
( 2 π h̄ ) 2 h̄

in which χ(R) is the nuclear wave function in the real space representation.
Since the scattering potential is represented by a Fermi point-scattering pseudo-
potential,
2 π h̄2
V (rn − R) = b δ(rn − R) (18)
mn
where b is the scattering length, its Fourier transform is given by

2 π h̄2
 
V (q) = b exp i q . R (19)
mn
Therefore, the neutron scattering cross-section can be entirely expressed in
terms of the nuclear density-density correlation function
 2   2
d σ kf mn
= | V (q) |2 S(q, ω) (20)
dΩdω ki 2 π h̄2
From this, we conclude that measurements of the scattering cross-section yields
information about the kinetics of the transition and the matrix-elements of the
interaction potential.

We shall now consider the effect of the neutron scattering from a neutral
atom. Since the interaction between the neutron and the atom is identically
zero in the asymptotic initial and final states, the wave function of the atom
Ψ(r, R) can be analyzed in terms of its center of mass and relative coordinates.
If the atomic Hamiltonian for the asymptotic in and out states is expressed as

P̂ 2 p̂2 Z e2
Ĥ = + − (21)
2 MN 2 me | r−R |
then it decouples in the center of mass and relative coordinates
MN R + me r
RCM =
MN + me
rrel = r − R (22)

14
When expressed in terms of these coordinates, the atomic Hamiltonian decouples
as
h̄2 h̄2 ( MN + me ) 2 Z e2
Ĥ = − ∇2CM − ∇rel − (23)
2 ( M N + me ) 2 M N me | rrel |
Hence, the asymptotic atomic wave functions have the form of products

Ψ(r, R) = φ(rrel ) χ(RCM ) (24)

If we neglect the mass of the electron compared with the mass of the nucleus,
me /MN  1, we expect to recover the Born-Oppenheimer approximation. In
this approximation, the center of mass coordinate reduces to the nuclear coor-
dinate RCM → R. In the (incoherent) inelastic scattering process, the electrons
in the atom are excited, hence the initial and final states are represented as
product states involving the electron and the atomic nucleus

Ψi,n (r, R) = φi (r − R) χn (R)


Ψj,m (r, R) = φj (r − R) χm (R) (25)

In the impulse approximation, the nuclear wave functions are represented by


plane waves  
1
χn (R) = √ exp i k n . R (26)
V
The inelastic scattering cross-section for the neutron beam is represented as
 2   2 X  
d σ kf mn 2
= | < Ψ j,m | V (q) | Ψi,n > | δ h̄ω + E i,n − E j,m
dΩdω ki 2 π h̄2 i,n;j,m
(27)
since h̄ ω and h̄ q are the energy and momentum gained by the atom. The
integration over the vector R yields the condition for the conservation of mo-
mentum. Hence, the scattering cross-section is proportional to the overlap of
the initial and final electronic states
Z
d3 r φ∗j (r − R) φi (r − R) (28)

expressed in terms of the relative coordinates. On using the orthonormality


condition for the electronic energy eigenstates, one obtains a factor of δi,j , so
that the initial and final electronic states are identical. Therefore, in the Born-
Oppenheimer approximation, the scattering is purely elastic.

In going beyond the Born-Oppenheimer approximation, one needs to include


the mass of the electrons, me . The conservation of momentum and energy refers
to the center of mass motion of the atom, where the atomic mass is the total
mass MN + me . The center of mass-coordinate of the atom is defined by
M N R + me r
RCM = (29)
M N + me

15
The center of mass part of the initial and final state atomic wave function,
respectively, contains the factor of
 
M N R + me r
exp i k n . (30)
M N + me

and  
MN R + me r
exp i km . (31)
MN + me
where k n and k m are the initial and final momenta of the atom. We shall express
the electronic position r relative to the nuclear position as r0 = r − R. With
this notation, the integration over the nuclear coordinate still yields conservation
of momentum. However, the electronic part of the scattering matrix elements
now involves the factor which is a function of the momentum q gained by the
nucleus
me r 0
Z  
d3 r0 φ∗j (r0 ) exp − i q . φi (r0 ) (32)
MN + me
The exponential factor allows inelastic transitions to take place. The scattering
amplitude for inelastic transitions is proportional to
  me
me MN
q = q me (33)
M N + me 1 + M N

and also to the “dipole” matrix element


Z
d3 r0 φ∗j (r0 ) r0 φi (r0 ) (34)

Hence, we have shown that the probability amplitude that the electrons are
excited by the nuclear motion is controlled by the ratio
me
(35)
MN
me
The Born-Oppenheimer approximation is valid whenever MN  1.

—————————————————————————————————-

In the first part of the course it is assumed that the Born-Oppenheimer ap-
proximation is valid.

First, the subject of Crystallography shall be discussed, and the characters


of the equilibrium structures of the dressed nuclei in matter will be described.
An important class of such materials are those which posses long-ranged peri-
odic translational order and other symmetries. It shall be shown how these long
range ordered and amorphous structures can be effectively probed by various
elastic scattering experiments in which the wave length of the scattered particles

16
is comparable to the distance between the nuclei.

In the second part, the properties of the Electrons shall be discussed. On


assuming the validity of the Born-Oppenheimer approximation, the nature of
the electronic states that occur in the presence of the potential produced by
the static nuclei shall be discussed. One surprising result of this approach is
that, even though the strength of the ionic potential is quite large (of the or-
der of Rydbergs), in some metals the highest occupied electronic states bear a
close resemblance to the states expected if the ionic potential was very weak or
negligible. In other materials, the potential due to the ionic charges can pro-
duce gaps in the electronic energy spectrum. Using Bloch’s theorem, it shall be
shown how periodic long-ranged order can produce gaps in the electronic spec-
trum. Another surprising result is that in most metals, it appears as though
the electron-electron interactions can be neglected or, more precisely, that the
excitations of the interacting electron system are similar to those of a non-
interacting electron gas, albeit with renormalized masses or magnetic moments.

The thermodynamic properties of electrons in these Bloch states shall be


treated using Fermi-Dirac statistics. Furthermore, the concepts of the Fermi
energy and Fermi surface of metals will be introduced. It shall be shown how
the electronic transport properties of metals are dominated by states with en-
ergies close to the Fermi surface, and how the Fermi surface can be probed.

The third part concerns the motion of the ions or nuclei. In particular, it
will be considered how the fast motion of the electrons dresses or screens the
inter-nuclear potentials. The low-energy excitations of the dressed nuclear or
ionic structure of matter give rise to harmonic-like vibrations. The elementary
excitation of the quantized vibrations are known as Phonons. It shall be shown
how these phonon excitations manifest themselves in experiments, in thermo-
dynamic properties and, how they participate in limiting electrical transport.

The final part of the course concerns some of the more striking examples
of the Collective Phenomenon such as Magnetism and Superconductivity.
These phenomena involve the interactions between the elementary excitations
of the solid which through collective action, spontaneously break the symmetry
of the Hamiltonian. In many cases, the spontaneously broken symmetry is ac-
companied by the formation of a new branch of low-energy excitations.

17
2 Crystallography
Crystallography is the study of the structure of ordered solids, disordered solids
and also liquids. In this section, it shall be assumed that the nuclei are static,
frozen into their average positions. Due to the large nuclear masses and strong
interactions between the nuclei (dressed by their accompanying clouds of elec-
trons), one may assume that the nuclear or ionic motion can be treated classi-
cally. The most notable failure of this assumption occurs with the very lightest
of nuclei, such as He. In the anomalous case of He, where the separation be-
tween ions, d, is of the order of angstroms, the uncertainty of the momentum
is given by h̄d and so the kinetic energy EK for this quantum zero point motion
can be estimated as
h̄2
EK ≈ (36)
2 M d2
The kinetic energy is large since the mass M of the He atom is small. The
magnitude of the kinetic energy of the zero point fluctuations is larger than the
weak van der Waals or London force between the He ions. Thus, the inter-ionic
forces are insufficient to bind the He ion into a solid and the material remains in
a liquid-like state until the lowest attainable temperatures. For these reasons,
He behaves like a quantum fluid. However, for the heavier nuclei, the quantum
nature of the particles only manifest themselves in more subtle ways.

First, the various types of structures and the symmetries that can be found
in Condensed Matter will be described and then the various experimental meth-
ods used to observe these structures will be discussed.

——————————————————————————————————

2.0.1 Exercise 1
Consider the interaction potential between two electrically neutral atoms po-
sitioned at R1 and R2 . The positions of the electrons located on atom 1 are
denoted by the set of ri and the positions of the electrons associated with the
second atom are denoted by the set rj . If the atoms are sufficiently far apart,
the sets of electrons belonging to each atom may be thought of as being distin-
guishable. In this case, the interaction between the two atoms can be expressed
as
Z 2 e2 X e2 X Ze2 X Ze2
Ĥint = + − −
| R1 − R2 | i,j
| ri − rj | i
| r i − R2 | j
| R1 − r j |
(37)
This interaction can be expanded in inverse powers of | R1 − R2 |.

If the eigenstates with energy eigenvalue En of the individual atoms are


(1) (2)
denoted by | Ψn > and | Ψn >, then second-order perturbation theory

18
yields the shift of the ground state energy due to the interaction between the
pair of atoms as

(1) (2) (1) (2)


X | < Ψ(1) Ψ(2) | Ĥint | Ψ(1) (2)
n Ψm > |
2
0 0
∆E = < Ψ0 Ψ0 | Ĥint | Ψ0 Ψ0 > +
n,m
2 E 0 − E n − Em
(38)
Show that the first term is just the classical electrostatic interaction due to the
charge density distributions around each atom. Using the hydrogenic-like 1s
one-electron wave functions
r
κ3
 
φ1s (r) = exp − κ r (39)
π
for the ground state and the 2s and 2p wave functions
r
κ3
   
φ2s (r) = 1 − κr exp − κ r
π
r
κ3
 
φ2p,0 (r) = cos θ κ r exp − κ r
π
r
κ3
   
φ2p,±1 (r) = sin θ exp ± i ϕ κ r exp − κ r

(40)

etc. for the excited states, estimate the sign and magnitude of the energy shift
for atoms with completely filled shells3 .

Estimate the Z dependence of the strengths of the London interaction4 for


the inert gases and compare your results with the values obtained from a part
of the semi-empirical Lennard-Jones potential
  12  6 
a a
Vint (R) = V0 − (41)
R R
where the values of V0 and a, respectively, are given in units of eV and Angstroms
by

element Ne Ar Kr Xe

V0 0.0124 0.0416 0.056 0.080


a 2.74 3.40 3.65 3.98

3 R. Eisenschitz and F. London, Z. für Physik, 60, 491 (1930), J. C. Slater and J. G.

Kirkwood, Phys. Rev. 37, 682 (1931).


4 F. London, Z. für Physik, 63, 245 (1930).

19
——————————————————————————————————

3 Structures
The structure of condensed materials is usually thought about in terms of den-
sity of either electrons or nuclear matter. To the extent that the regions of
non-zero density of the nuclear matter are highly localized in space, with lin-
ear dimensions of 10−15 meters, the nuclei can be discussed in terms of point
objects. The electron density is more extended and varies over length scales of
10−10 meters. The length scale for the electronic density in solids and fluids is
very similar to the length scale over which the electron density varies in isolated
atoms. The similarity of scales occurs as electrons are partially responsible for
the bonding of atoms into a solid. That is, the characteristic atomic length
scale is almost equal to the characteristic separation of the nuclei in condensed
matter. Due to the near equality of these two length scales, the electron den-
sity in solids definitely cannot be represented in terms of a superposition of the
density of well defined atoms. However, the electron density does show a signifi-
cant variation that can be interpreted in terms of the electron density of isolated
atoms, subject to significant modifications when brought together5 . As the elec-
tron density for isolated atoms is usually spherically symmetric, the structure
in the electronic density may, for convenience of discussion, be approximately
represented in terms of a set of spheres of finite radius.

3.1 Fluids
Both liquids and gases are fluids. The microscopic structure of a fluid varies
locally from position to position and in time. The macroscopic characteristics
of fluids are that they are spatially uniform and isotropic, which means that the
average environment of any atom is identical to the average environment of any
other atom.

The density is defined by the function


X
ρ(r) = δ 3 ( r − ri ) (42)
i

in which ri is the instantaneous position of the i-th atom. A measurement of


the density usually results in the time average of the density which corresponds
5 An example of the change in the charge density of the valence electrons of Si, caused by

solid formation, can be found in the contour plot in the article authored by D. R. Hamann,
Phys. Rev. Lett. 42, 622 (1979).

20
Figure 2: Contour plot of the valence charge density of Si (in atomic units).
The positions of the atoms are denoted by dots. [After Hamann (1979).]

to the time averaged positions of the atoms.

In particular, for a fluid, spatial homogeneity ensures that the time averaged
density ρ(r) at position r is equal to the average density at a displaced position
r + R,
ρ(r) = ρ(r + R) (43)
The value of the displacement R is arbitrary, so the average density is inde-
pendent of r and can be expressed as ρ(0). This just means that the average
position of an individual atom is undetermined.

The operations which leave the system unchanged are the symmetry oper-
ations. For a fluid, the symmetry operations consist of the continuous transla-
tions through an arbitrary displacement R, rotations through an arbitrary angle
about an arbitrary axis, and also reflections in arbitrary mirror planes.

The set of symmetry operations form a group called the symmetry group.
For a fluid, the symmetry group is the Euclidean group. Fluids have the largest
possible number of symmetry operators and have the highest possible symmetry.
All other materials are invariant under a smaller number of symmetry opera-
tions.

Nevertheless, fluids do have microscopic short-ranged structure which is ex-


emplified by locating one atom and then examining the positions of the neigh-
boring atoms. The local spatial correlations are expressed by the density -

21
density correlation function which is expressed as an average
C(r, r0 ) = ρ(r) ρ(r0 )

X
= δ 3 ( r − ri ) δ 3 ( r0 − rj )
i,j
(44)
Since fluids are homogeneous, the correlation function is only a function of
the difference of the positions r − r0 . Furthermore, since fluids are isotropic
and invariant under rotations, the correlation function is only a function of the
distance separating the two regions of space | r − r0 |. At sufficiently large
separation distances, the positions of the atoms become uncorrelated, thus,
lim C(r, r0 ) → ρ(r) ρ(r0 )
r−r 0 → ∞

→ ρ(0) ρ(r − r0 ) (45)


That is, at large spatial separations, the density - density correlation function
reduces to the product of the independent average of the density at the origin
and the average density at a position r. From the homogeneity of the fluid, ρ(r)
is identical to the average of ρ(0).

The density - density correlation function contains the correlation between


the same atom, that is, there are terms with i = j. This leads to a contribution
which shows up at short distances,
X X
δ 3 ( r − ri ) δ 3 ( r0 − ri ) = δ 3 ( r − r0 ) δ 3 ( r − ri )
i=j i
3 0
= δ ( r − r ) ρ(r) (46)
which is proportional to the density.

The pair distribution function g(r − r0 ) is defined as the contribution to the


density - density correlation function which excludes the correlation between an
atom and itself,
g( r − r0 ) = C( r − r0 ) − δ 3 ( r − r0 ) ρ(r) (47)
For a system which possesses continuous translational invariance, the pair dis-
tribution function can be evaluated as the temporal and spatial average
X
g( r − r0 ) = δ 3 ( r − ri ) δ 3 ( r0 − rj )
i6=j
Z
1 X
= d3 R δ 3 ( r − ri − R ) δ 3 ( r0 − rj − R )
V
i6=j
1 X 3
= δ ( r − r0 − ri + rj ) (48)
V
i6=j

22
Since the sum over i runs over all of the inter-atomic separations rj − ri for each
fixed value of j, spatial homogeneity demands that the contribution from every
j value is identical. There are N such terms, and this leads to an expression for
the pair distribution function involving an atom at the central site r0 and the
others at sites i in the form
X
g(r) = ρ(0) δ 3 ( r − ri + r0 ) (49)
i

where
N
ρ(0) = (50)
V
As this only depends on the radial distance | r |, this is also called the radial
distribution function g(r). The radial distribution function for liquid Argon
6
is shown in fig(3). For large r, the pair distribution function, like C(r, 0),

Figure 3: The radial distribution function g(r) for Argon at T = 85 K. [After


Yarnell et al. (1973).]
2
approaches ρ(0) , or
2
lim g(r) → ρ(0) (51)
r → ∞

For all fluids, g(r) vanishes as r → 0

lim g(r) → 0 (52)


r → 0
6 J. L. Yarnell, M. J. Katz, R. G. Wenzel, and S. H. Koenig, Phys. Rev. A 7, 2130 (1973).

23
since the presence of an atom at the origin excludes other atoms from residing
at this position. Because correlations in fluids are strongest at short distances,
g(r) usually exhibits a large peak at a radial distance greater than the diameter
of two atoms. The largest peak in g(r) is usually associated with the shell
of atoms nearest to the one at the origin. The integral over the peak in g(r)
approximately yields the number of the atoms in the nearest shell (Nnn ) times
the density
Z R+
4π dr r2 g(r) ≈ Nnn ρ(0) (53)
R−

Liquids are defined as the fluids that have high densities. The liquid phase
is not distinguished from the higher temperature gaseous phase by a change
in symmetry. In the liquid phase, the density is higher and the inter-atomic
forces play a more important role than in the low density gaseous phase. The
interaction forces are responsible for producing the short-ranged correlation in
the density - density function. A model potential that is representative of typical
inter-atomic force between two neutral atoms is the Lennard-Jones potential.
 12  6 
a a
V (r) = 4 V0 − (54)
r r

The potential has a short-ranged repulsion between the atoms caused by the

The Lennard Jones Potential for Neon


0.005

0.0025
V(r) [ eV ]

-0.0025

-0.005
0 1 2 3 4 5 6

r [ Angstroms ]

Figure 4: The radial dependence of the Lennard-Jones potential V (r).

24
overlap of the electronic states, and the long-ranged van der Waals attraction
caused by fluctuation-induced electric polarizations of the atoms. The resulting
1
potential falls to zero at r = a and has a minimum at r = 2 6 a. The potential
at the minimum of the well is given by − V0 . Given the form of the potential,
the atomic positions can be calculated from Newton’s laws to yield a computer
simulated structure of a fluid7 . These molecular dynamics calculations reveal
much more information about the structure than do averages. For example, the
position of the peak in g(r) only gives information about the average nearest
neighbor separation, whereas the molecular dynamics simulation also yields the
statistical variation of these distances.

Another model potential that is often used to describe liquids is the hard
sphere potential which excludes the center of another atom from the region of
radius 2 a centered on the central atom. These mutually impenetrable spheres
are then irregularly packed within some volume such that the resulting structure
contains no cavities large enough to contain another sphere8 . As the repulsion
between atoms dominates the structure of liquids, the Bernal model of random
close packing of hard spheres is responsible for most of the structure of a liq-
uid. This model of randomly close-packed spheres produces a structure which
is irregular but densely packed. The structure does contain small regions in
which the arrangement of atoms has near-perfect hexagonal symmetry. The
packing fraction is defined as the total volume of the hard spheres divided by
the (minimum) volume that contains all the spheres. On randomly packing
spheres, one finds a limiting upper bound on the packing fraction which is given
by 0.638. However, there do exist regular (non-random) close-packed structures
with packing fractions of 0.7405. Bernal also showed that, on averaging over the
random close-packed structure, each sphere was in contact with approximately
8.5 other spheres. The randomly close-packed structure may also be analyzed
in terms of Voronoi polyhedra. A Voronoi polyhedron of a specific sphere in the
structure is constructed by first joining the centers of the sphere to the other
spheres by a set of straight line segments. Next, a bisecting plane is constructed
for each line segment. The set of closest planes which completely enclose the
center of the specific sphere forms the Voronoi polyhedron for that sphere. For
the hard sphere model, the Voronoi polyhedron completely encloses the sphere.
The resulting set of Voronoi polyhedra can then be analyzed in terms of their
volumes, their number of faces and edges. For the randomly packed structure,
the average number of faces was found to be 14.25. A difference between the
average number of faces and the average number of contacts is expected in a
random system, since one can pack more spheres around the central sphere if
the separations between the centers are allowed to vary. The shell of closest
neighboring spheres are defined as the set of spheres which, when joined to the
central sphere by line segments, have bisecting planes that form the faces of the
Voronoi polyhedron. The radial distances between the central sphere and the
7 A. Rahman, J. Chem. Phys. 45, 2585 (1966).
8 J. D. Bernal, Nature, 183, 141 (1959).

25
Figure 5: A random packing of non-overlapping discs in two-dimensions.

spheres in the shell of closest neighbors are found to have values in the range
2 a < r < 2.3 a. The average number of edges per face of the Voronoi poly-
hedra are found to be 5.16.

Random packings of hard spheres can be used to calculate the radial distribu-
tion function g(r). The random close-packing model shows that there are strong
short-ranged correlations between the “closest” atoms. The model also shows
that, in three dimensions, the radial distribution function is zero for r < 2 a.
The radial distribution function then peaks up at a radius slightly greater than
2 a with an intensity which corresponds to the average number of faces of the
Voronoi polyhedra. The short-ranged correlations also show up as other peaks
in the radial distribution function at greater distances which correspond to the
next few shells of neighboring atoms. Due to the larger variation in the radial
separation between the central atom and the atoms in the more distant neighbor
shells, these other peaks are significantly broader than the peak corresponding
to the first shell.

It is noteworthy that the hard sphere model does not yield structures with
packing fractions intermediate between the limiting value of 0.638 (found for
random close packing) and the value of 0.7405 (found for perfectly ordered

26
close-packed structures). One might have expected that there would be a series
of structures with continuously varying packing fractions. The discontinuity in
the densities of the hard sphere structures may be correlated with the signif-
icant but discontinuous change in density that occurs as a liquid freezes and
transforms into a crystalline solid.

27
3.2 Crystalline Solids
A perfect crystal can be partitioned into identical, non-overlapping, structural
units that completely fill the volume of the (infinite) crystal. When the iden-
tical structural units are packed together, they form a periodic structure. If a
point is chosen in one structural unit, the set of equivalent points in the other
units forms a periodic lattice. There are many different ways of partitioning of
the crystal and, therefore, there are many alternate forms of the structural unit.
The structural unit is called the unit cell. The unit cells which have the smallest
possible volume are called primitive unit cells. A unit cell may contain one or
more atoms. The crystal is only specified if the lattice is specified and the types
and positions of all the atoms in the unit cell are specified. The locations and
types of atoms in the unit cell forms the basis of the crystal.

3.3 The Direct Lattice


The set of equivalent points, one taken from each unit cell in a perfect crystal,
form a periodic lattice. The points are called lattice points. Any lattice point
can be reached from any other by a translation R that is a combination of an
integer multiple of three primitive lattice vectors a1 , a2 , a3 ,

R = n1 a1 + n2 a2 + n3 a3 (55)

Here, n1 , n2 and n3 are integers that determine the magnitudes of three com-
ponents of a three-dimensional vector. The set of integers (n1 , n2 , n3 ) can be
used to represent a lattice point in terms of the primitive lattice vectors. The
set of values of the ni run through all positive and negative integers. Any two
lattice vectors R can be combined by addition to produce another lattice vector.
As will be seen later, this is responsible for the set of translations being closed
under addition and, therefore, the translation operations form a group.

Given any lattice, there are many choices for the primitive lattice vectors
a1 , a2 , a3 .

The array of lattice points have arrangements and orientations which are
identical in every respect when viewed from origins centered on different lattice
points. For example, on translating the origin of the unprimed reference frame
through a lattice vector Rm , one obtains a new reference frame (the primed
reference frame). The displacements in the primed reference frame are related
to displacements in the unprimed reference frame via

r0 = r + m1 a1 + m2 a2 + m3 a3 (56)

and the integers labelling the lattice points in the two frames are related via

n0i = ni + mi (57)

28
and as the numbers ni and n0i take on all possible integer values, the set of all
lattice points are identical in the two reference frames.

A crystal structure is composed of a lattice in which a basis of atoms is


attached to each lattice point. That is, a complete specification of a crystal
structure requires specifying the lattice and the distribution of the various atoms
around each lattice point. The basis is specified by giving the number of atoms
and types of the atoms in the basis (j) together with their positions relative to
the lattice points. The position of the j-th atom relative to the lattice point rj ,
is denoted by  
rj = xj a1 + yj a2 + zj a3 (58)

where the set (xi , yi , zi ) may be non-integer numbers.

The choice of lattice and, therefore, the basis, is non-unique for a crystal
structure. An example of this is given by a two-dimensional crystal structure

a"2

a"1

a'2
a2
a1 a'1

Figure 6: A two-dimensional lattice, and some choices for the primitive lattice
vectors.

which can be represented many different ways including the possibilities of a


representation either as a lattice with a one atom basis or as a lattice with a
two atom basis.

29
a2 r2

r1
a1

Figure 7: A two-dimensional lattice, and with a choice of the (non-primitive)


lattice vectors corresponding to a lattice with a two-site basis.

3.3.1 Primitive Unit Cells


The parallelepiped defined by the primitive lattice vectors forms a primitive
unit cell. When repeated, a primitive unit cell will fill all space. The primitive
unit cell is also a unit cell with the minimum volume. Although there are a
number of different ways of choosing the primitive lattice vectors and unit cells,
the number of basis atoms in a primitive cell is unique for each crystal structure.
No basis contains fewer atoms than the basis associated with a primitive unit
cell.

There is always just one lattice point per primitive unit cell.

If the primitive unit cell is a parallelepiped with lattice points at each of the
eight corners, then each corner is shared by eight cells, so that the total number
of lattice points per cell is unity as 8 × 18 = 1.

The volume of the parallelepiped is given in terms of the primitive lattice


vectors via
Vc = | a1 . ( a2 ∧ a3 ) | (59)
The primitive unit cell is a unit cell of minimum volume.

30
A Primitive Unit Cell

a3

a2
a1

Figure 8: A primitive unit cell.

3.3.2 The Wigner-Seitz Unit Cell


An alternate method of constructing a unit cell was proposed by Wigner and
Seitz. The Wigner-Seitz cell has the important property that there are no arbi-
trary choices made in defining the unit cell. The absence of any arbitrary choice
has the consequence that the Wigner-Seitz unit cell always has the same sym-
metry as the lattice. The Wigner-Seitz unit cell is constructed by forming a set
of planes which bisect the lines joining a central lattice point to all other lattice
points. The region of space surrounding the central lattice point, of minimum
volume, which is completely enclosed by a set of the bisecting planes consti-
tutes the Wigner-Seitz cell. Thus, the Wigner-Seitz cell consists of the volume
composed of all the points that are closer to the central lattice site than to any
other lattice site.

The equation of the plane bisecting the vector from the central point to the
i-th lattice point is given by
 
1
r − Ri . Ri = 0 (60)
2

where Ri is the lattice vector. The planes which bound a volume closer to the
origin than any other lattice site form the surface of the Wigner Seitz-cell.

31
A two atom basis

x2 a1
a3 y2 a2

z2 a3 a2
a1

Figure 9: A primitive unit cell, with a two-atom basis. One basis atom is located
at the origin r1 = (0, 0, 0). The second atom is located at r2 = (x2 , y2 , z2 ).

As the definition does not involve any arbitrary choice of primitive lattice
vectors, the Wigner-Seitz cell possesses the full symmetry of the lattice. Fur-
thermore, the Wigner-Seitz cell is space filling since every point in space must
lie closer to one lattice site than any other.

32
Two-dimensional Wigner-Seitz cell

Figure 10: The Wigner-Seitz unit cell of a two-dimensional lattice. The sym-
metry of the Wigner-Seitz cell is independent of the arbitrary choice of the
primitive lattice vectors (a1 , a2 ).

3.4 Symmetry of Crystals


A symmetry operation acts on a crystal producing a new crystal by shifting
the constituent particles to new positions such that the new crystal is identical
in appearance to the original crystal. That is, after the symmetry operation,
the positions of the particles in the new crystal coincide with the positions of
similar particles in the original crystal. Each particle can be assigned a unique
(non-physical) label, irrespective of whether the particles are physically distin-
guishable or not. The uniquely labelled particles are called points. The symme-

33
Wigner-Seitz construction

r r-Ri/2

Ri/2 Ri
O

Figure 11: The plane bisecting the lattice vector Ri used in the construction of
the Wigner-Seitz cell.

try operations can then be described by their actions on the set of points. The
symmetry operations may consist of :

(i) Translation operations which leave no point unchanged.

(ii) Symmetry operations which leave one point unchanged.

(iii) Combinations of the above two types of operations.

3.4.1 Symmetry Groups


A set of symmetry operations form a group if, when the symmetry operations
are combined, the following properties are satisfied :

(I) The combination of any two symmetry operators from the set, say A and
B, defined by A B = C has a product C which is also in the set. That is, the
set of symmetry operations is closed under composition.

(II) The composition of any three elements is associative, which means that
the symmetry operation is independent of whether the first and second operators
are combined before they are combined with the third, or whether the second

34
and third operators are combined before they are combined with the first.

A(BC ) = (AB)C (61)

(III) There exists a symmetry operator which leaves all the atoms in their
original places called the identity operator E. The product of any symmetry
operator arbitrarily chosen from the group with the identity gives back the
arbitrarily chosen operator.

AE = EA = A (62)

(IV) For each operator in the group, there exists a unique inverse operator
such that when the operator is combined with its inverse operator, they produce
the identity.
A A−1 = A−1 A = E (63)

A group of symmetry operators may contain a sub-set of symmetry operators


which also form a group. That is, the group laws are obeyed for all the elements
of the sub-set. This sub-set of elements forms a sub-group of the group, but is
only a sub-group if the elements are combined with the same law of composition
as the group.

The symmetry group of the direct lattice contains at least two sub-groups.
These are the sub-group of translations and the point group of the lattice. Under
a translation which is not the identity, no point remains invariant. The point
group of the lattice consists of the set of symmetry operations under which at
least one point of the lattice is invariant.

3.4.2 Group Multiplication Tables


The properties of a group are concisely represented by the group multiplication
table. The number of elements in the group is called the order of the group, so
the general group with n operators is of order n. The group multiplication table
of a group of order n consists of an n by n array. The group multiplication table
has the convention that if A × B = C, then the operator A which is the first
element of the product is located on the left most column of the table, and the
operator B which is the second element is located in the uppermost row. The
product C is entered in the same row as the element A and the same column as

35
element B.
E . . B .
. . . . .
A . . C .
. . . . .
. . . . .
In general, the symmetry operations do not commute, that is, A × B 6= B × A.
The identity operator is placed as the first element of the series of symmetry
operators, so the first row and first column play a dual role. The first column
and first row play one role as the list of the groups elements. The second role
that they play is as a record of the products found by compounding the elements
with the identity. Every operator appears once, and only once, in each row or
column of the group table. The fact that each operator occurs only once in any
row or in any column, is a consequence of the uniqueness of the inverse.

As an example, consider the point group for a single H2 O molecule. The

The symmetry elements of an H2O molecule

σv'

C2
σv

Figure 12: The symmetry operations of an H2 O molecule. The symmetry ele-


ments are composed of a two-fold axis of rotation, and two mirror planes.

group contains a symmetry operation which is a rotation by π about an axis in


the plane of the molecule that passes through the O atom and bisects the line
joining the two H atoms. This is a two-fold axis, since if the rotation about
the axis is followed by a second rotation about the same axis then the com-
bined symmetry operation is equivalent to the identity. The two-fold rotation

36
is labelled as C2 . In this case, the two-fold axis is the rotation axis of highest
order and thus, is considered to define the vertical direction. In addition to the
two-fold axis, there are two mirror planes. It is conventional to denote a mirror
plane that contains the n-fold axis of rotation (Cn ) with highest n as a vertical
plane. The H2 O molecule is symmetric under reflection in a mirror plane pass-
ing through the two-fold axis in the plane which contains the molecule. That is,
the mirror plane is the plane passing through all three atoms. This is a vertical
mirror symmetry operation and is denoted by σv . The second mirror symmetry
operation is a reflection in another vertical plane passing through the C2 axis
but this time, the mirror plane is perpendicular to the plane of the molecule
and is denoted by σv0 . The symmetry group contains the elements E, C2 , σv ,
σv0 . The group is of order 4. The group table is given by

E C2 σv σv0

C2 E σv0 σv

σv σv0 E C2

σv0 σv C2 E

Since all the operations in this group commute (i.e. A B = B A ), the group
is known as an Abelian group. Inspection of the table immediately shows that
σv × C2 = σv0 .

The symmetry group of a crystal has at least two sub-groups. One sub-group
is the group of translations through the set lattice vectors R. In the case of the
translations, the law of composition is denoted as addition. A general transla-
tion which is not the identity leaves no point unchanged. A second sub-group
is formed by the set of all transformations which leave a particular point of the
crystal untransformed. This sub-group is the point group.

3.4.3 Point Group Operations


The crystallographic point group consists of the symmetry operations that leave
at least one point untransformed. Furthermore, the various symmetry opera-
tions that belong to one point group always leave the same point unchanged.
The possible symmetry elements of the point group are:

Rotations around an axis through angles which are integer multiples of 2nπ .
An n-fold rotation is denoted as Cn . A rotation by 2 πn m can be expressed
as the composition of m successive rotations about the same axis (Cn )m . A
combination of n rotations by 2nπ about the same axis leads to the identity
(Cn )n = E, and n is known as the order of the axis.

37
C3

π/3

Figure 13: A molecule with an n-fold rotation axis Cn (with n = 3).

Reflections which take every point into its mirror image with respect to a
plane known as the mirror plane. Reflections are denoted by σ.

Inversions which take every point r, as measured from an origin, into the point
− r. The inversion operator is denoted by I.

Inversion I

-r

Figure 14: The inversion operator I transforms the point r into the point −r.

Rotation Reflections which are rotations about an axis through integer mul-
tiples of 2nπ followed by reflection in a plane perpendicular to the axis. The

38
n-fold rotation reflections are denoted by Sn . For even n, (Sn )n = E, while
for odd n, (Sn )n = σ.

Roto-Reflection S2

2π/2

Figure 15: A two-fold rotation-reflection axis.

Rotation Inversions which are rotations about an axis through integer mul-
tiples of 2nπ followed by an inversion through an origin. The International
notation for a rotation reflection is n. The rotation inversion and rotation re-
flection operations are related for example, 3 = S6−1 , 4 = S4−1 and 6 = S3−1 .

Since at least one particular point is invariant under all the transformations
of the point group, the rotation axes and mirror planes must all intersect at this
point.

Symmetry operations A and B are defined to be equivalent if the symmetry


group contains an element C such that

A = C −1 B C (64)

The set of elements which are all equivalent to each other form an equivalence
class. It can be seen that equivalent symmetry operations are of the same type,
but may involve different orientations of the translations, rotation axes or mirror
planes. For example, two rotations through the same angle but about different
axes may be equivalent to each other. Likewise, two reflections in different mir-
ror planes may be equivalent to each other.

——————————————————————————————————

39
Roto-Reflection S3

2π/3

Roto-Reflection S4

2π/4

Figure 16: The n-fold rotation-reflection axis, with n = 3 and n = 4.

3.4.4 Exercise 2
Consider a structure with a symmetry group that contains a principal axis of
order n (Cn ) and also contains one vertical mirror plane (σ1 ). Any atom at
point P1 in the structure is related to an identical atom at Q1 by the mirror
symmetry
Q1 = σ1 P1 (65)
Likewise, atoms at points P1 and Q1 , respectively, are related to identical atoms
at Pm and Qm via the repeated actions of Cn ,
Pm = ( Cn )m P1
Qm = ( Cn )m Q1 (66)

40
Show that the symmetry group must contain symmetry operations σm which
are equivalent to σ1 via
σm = ( Cn )m σ1 ( Cn )−m (67)
Identify the symmetry operations σm .

σ1

P1 Q1

C3

P2
Q3
Q2
σ3 P3 σ2

Figure 17: A group with an n-fold rotation axis Cn and one vertical mirror
plane σ1 has a set of equivalent vertical mirror planes σn .

——————————————————————————————————

3.4.5 Limitations Imposed by Translational Symmetry


Although all point group operations are allowable for isolated molecules, certain
point group symmetries are not allowed for periodic lattices. The limitations on
the possible types of point group symmetry operations can be seen by examining
the effect of an n-fold axis in a plane perpendicular to a line through lattice
points A − B . . . C − D, with 1 + m1 lattice points on it. The direction
of the line will be chosen as the direction of the primitive lattice vector a1 , and
the line is assumed to have a length m1 a1 . A clockwise rotation of 2nπ about
the n-fold axis of rotation through point A will generate a new line A − B 0 .
Likewise, a counter clockwise rotation of 2nπ about the n-fold axis of rotation
through point D will generate a new line D − C”. The line constructed through
B 0 − C” is parallel to the initial line A − D. The length of the line B 0 − C”
is equal to m1 a1 − 2 cos 2nπ a1 and must be equal to an integer multiple of
a1 , say m01 . Then
2π m1 − m01
cos = (68)
n 2
Thus, cos 2nπ must be an integer or a half odd integer which is in the set
{±1, 0, ± 12 }. This restriction limits the possible n-fold rotation axis to be of

41
The limitations on rotational symmetry imposed by
periodic translational invariance

m1'
B' C"
θ = 2π/n θ
A B m1 C D

Figure 18: The limitations on rotational symmetry elements imposed by peri-


odicity.

order n = 1 , 2 , 3 , 4 , 6 and allows no others. Thus, a crystalline lattice


can only contain two-fold, three-fold, four-fold or six-fold axes of rotation. How-
ever, there do exist solids that possess five-fold symmetry, such as quasi-crystals.
Quasi-crystals are not crystals as they do not possess periodic translational in-
variance.

——————————————————————————————————

3.4.6 Exercise 3
When point group symmetry operations are combined, new symmetry elements
may arise. These new operations are also point group operations, since they
leave a specific point invariant.

42
P2

θ2/2
φ1

φ3 P3

θ1/2 φ2

P1

Figure 19: The Euler construction shows how a rotation by θ1 about the axes
P1 can be combined with a rotation through angle θ2 about the axis P2 . These
rotations when combined in the correct order are identical to a rotation about
the axis P3 .

Prove that a rotation about an axis Cn1 , followed by a rotation about dif-
ferent axis Cn2 , is identical to a rotation Cn3 about a third axis. The angle
between the Cn1 axis and the Cn2 axis is denoted by ϕ3 . Show that, if there are
no other axes of rotations present, the angle between the axes and the orders of
the axes must satisfy the condition
cos nπ3 + cos nπ1 cos nπ2
cos ϕ3 = (69)
sin nπ1 sin nπ2

Hence, for a periodic crystal, show that the only allowed non-trivial combina-
tions of three rotational axes have orders n1 n2 n3 given by 2 2 2, 2 2 3, 2 2 4,
2 2 6, 2 3 3 and 2 3 4.

43
     
2π 2π 2π
(n1 , n2 , n3 ) θ1 = n1 θ2 = n2 θ3 = n3 ϕ1 ϕ2 ϕ3

π π π
(2,2,2) π π π 2 2 2

2π π π π
(2,2,3) π π 3 2 2 3

π π π π
(2,2,4) π π 2 2 2 4

π π π π
(2,2,6) π π 3 2 2 6

(2,3,3) π 2π
3

3 cos−1 1
3 cos−1 √1
3
cos−1 √1
3
q
(2,3,4) π 2π
3
π
2 cos−1 √1
3
π
4 cos−1 2
3

——————————————————————————————————

3.4.7 Point Group Nomenclature


The point groups are frequently referred to by using either one or the other of
two different notation schemes: the Schoenflies and the International notation.
In the following examples, the groups are first labelled by their Schoenflies des-
ignation and then by their International designation.

In enumerating the point groups, we shall first list the rotational groups and
then list the rotational groups which are adjoined by mirror planes or improper
rotations that do not introduce any further proper rotational symmetry opera-
tions.

The point groups are:

Cn The groups Cn only contain an n-fold rotation axis. The number of


elements in the group Cn is equal to the order of the axis. The group is known
as the cyclic group of order n and it is Abelian. The international symbol des-
ignating this group is n.

Cn,v The groups Cn,v contain the n-fold rotation axis and have vertical mir-
ror planes which contain the axis of rotation. The effect of the n-fold axis, if n

44
The symmetry elements of Cnv

C4

σv
σv'

Figure 20: The group Cn,v has an n-fold rotation axis and equivalent vertical
mirror planes.

is odd, is such that it produces a set of n equivalent mirror planes. This yields
2 n symmetry operations, which are the n rotations and the reflections in the
n mirror planes. If n is even, the effect of repeating Cn only produces n2 equiv-
alent mirror planes. The other n2 rotations merely bring the mirror plane into
coincidence with itself, but with the two surfaces of the mirror interchanged.
A mirror plane is equivalent to its partner mirror plane found by rotating it
through π since by definition, a mirror plane is two-sided. However, for even
n, the effect of the compounded operation Cn σv acting on an arbitrary point
P produces a point P 0 which is identical to the point P 0 produced by reflection
of P in the mirror plane that bisects the angle between two equivalent mirror
planes σv . Thus, the symmetry element given by the product Cn σv is identical
to a mirror reflection in the bisecting (vertical) mirror plane. The effect of Cn
is to transform these bisecting mirror planes into a set of n2 equivalent bisect-
ing mirror planes. Thus, if n is even, there are also 2 n symmetry operations.
These 2 n symmetry operations are the set of n rotations and the two sets of
n
2 reflections. Mirror planes which are not perpendicular to the rotation axis
are recorded as m without any special marking. For even n, the International
symbol for Cn,v is nmm. The two m’s refer to two distinct sets of mirror planes:
one from the original vertical mirror plane and the second m refers to the verti-
cal mirror planes which bisect the first set. For odd n, the international symbol
is just nm, as the group only contain one set of mirror planes and does not

45
contain a set of bisecting mirror planes.

Q1 σv P1
σv'

σv' = C4 σ v C4 Q2

σv

Figure 21: A group with an n-fold rotation axis Cn , where n is even, and one
vertical mirror plane σv has two sets of equivalent vertical mirror planes σv and
σv0 .

Cn,h The groups Cn,h contain the n-fold rotation, and have a horizontal
mirror plane which is perpendicular to the axis of rotation. The group contains
n
2 n elements and, if n is even, the group contains Cn2 . σh = C2 . σh = I
which is the inversion operator. The International notation usually refers to
the group Cn,h as n/m. The diagonal line indicates that the symmetry plane
is perpendicular to the axis of rotation. The only exception is C3h or 6. The
international symbol signifies that C3h is relegated to the group of rotation re-
flections which are, in general, designated by n.

Sn The groups Sn only contain the n-fold rotation - reflection axis. These
groups are obtained by attaching an m0 -fold rotation - reflection axis to the
2
rotation group containing Cm , such that Sm 0 = Cm . That is, the rotation -
reflection axis must coincide with the axis of the rotation group. For even n,
the group contains only n elements since (Sn )n = E. For odd n, (Sn )n = σ,
so the group must contain 2 n elements. The International notation is given
by the equivalent rotation inversion group n. For example, S6 ≡ 3, S4 ≡ 4,
S3 ≡ 6.

Dn The groups Dn contain an n-fold axis of rotation and a two-fold axis


which is perpendicular to the n-fold axis. The effect of the n-fold rotation is
to produce a set of equivalent two-fold axes. If n is odd there are n equivalent
two-fold axes. If n is even, the n-fold rotation produces n2 equivalent two-fold
axes which are two sided. When n is even, the action of a two-fold rotation
followed by an n-fold rotation is equivalent to a new two-fold rotation about an
axis that bisects the original sets of two-fold axes. This can be seen by following

46
Symmetry elements of D3

C3

C2
C2

C2

Figure 22: The dihedral group Dn , for odd n, has an n-fold rotation axis and a
set of n equivalent two-fold rotation axes.

the action of an arbitrary point P with coordinates (x, y, z) under the two-fold
rotation about a horizontal axis, say the x axis. The rotation by π about the x
axis sends z → − z and y → − y. A further rotation of 2nπ about the z axis,
sends the point (x, −y, −z) to the final image point P 0 . Note that the z coordi-
nate of point P 0 is − z. Construct the line joining P and P 0 . The mid-point of
P − P 0 lies on the plane z = 0, and subtends an angle of nπ with the x axis
and, therefore, lies on the bisecting rotation axis. As the bisecting axis passes
through the mid-point of line P − P 0 , this shows that the arbitrary point P
can be sent to P 0 via a π rotation about the bisecting axis. Thus, for even n,
there are n2 bisecting two-fold axes, and the original n2 two-fold axes. In case of
either even or odd n, the group contains 2 n elements consisting of the n-fold
rotations and a total of n two-fold rotations. The International designation for
Dn is either n22 or just n2, depending on whether n is even or odd. These
two designations occur for similar reasons as to why there are two International
designations for Cn,v . For odd n the designation n2 indicates that there is one
n-fold axis and one set of equivalent two-fold axes. For even n, the symbol n22
indicates the existence of an n-fold axes and two inequivalent sets of two-fold
axes.

Dnh The groups Dnh contain all the elements of Dn and also contain a hor-
izontal mirror plane perpendicular to the n-fold axis. The effect of a rotation

47
Symmetry Elements of D4

C4

C2'

C2
C2
C2'

Figure 23: The dihedral group Dn , for even n, has an n-fold rotation axis and
two sets of equivalent two-fold rotation axes.

about a two-fold axis followed by the reflection σh is equivalent to a reflection


about a vertical plane σv passing through the two-fold axis. Since rotating σv
about the Cn axis produces a set of n vertical mirror planes, adding a horizontal
mirror plane to Dn produces n vertical mirror planes σv . The group has 4 n
elements which are formed from the 2 n rotations of Dn , the n reflections in the
vertical mirror planes, and n rotation reflections Cnk σh . For even n, the Inter-
n 2 2
national symbol is m m m which is often abbreviated to n/mmm. The symbol
indicates that the n-fold axis has a perpendicular mirror plane, and also that
the two sets of two-fold axes have perpendicular mirror planes. For odd n, the
International symbol for the group acknowledges the 2n-fold rotation inversion
symmetry and is labelled as 2n2m.

Dnd The groups Dnd contain all the elements of Dn and mirror planes which
contain the n-fold axis and bisect the two-fold axes. The action of the two-fold
rotations generate a total of n vertical reflection planes. There are 4 n elements,
the 2 n rotations of Dn , n mirror reflections σd in the n vertical planes. The re-
maining n elements are rotation reflections about the principle axis of the form
2k+1
S2n where k = 0 , 1 , 2 , . . . , ( n − 1 ). The principle axis is, therefore, a
2n-fold rotation reflection axis. The International symbol is n2m indicating a
n-fold axis, a perpendicular two-fold axis and a vertical mirror plane.

48
T The tetrahedral group corresponds to the group of rotations of the reg-
ular tetrahedron. The elements are comprised of four three-fold rotation axes
passing through one vertex and through the center of the opposite face of the
tetrahedron. The compound action of two of the three-fold rotations yields a ro-

Rotational Symmetries of a Tetrahedron

C2

C3

Figure 24: The rotational symmetry elements of a tetrahedron. The tetrahedron


has four three-fold axes C3 passing through each vertex, and has three two-fold
axes which pass through the mid-points of opposite edges.

tation about a two-fold axis. There are three such two-fold axes passing through
the mid-points of opposite edges of the tetrahedron. The tetrahedral group has
twelve elements. The symmetry operations can also be found in a cube, if the
three four-fold rotation axes present in the cube are discarded. The group has
the International symbol of 23.

Td The group Td corresponds to the tetrahedral group adjoined by a reflec-


tion plane passing through one edge and the mid-point of the opposite edge of
the tetrahedron. The reflection planes bisect a pair of two-fold axes of T . Since
the tetrahedron has six edges, there are six mirror planes σd . For the cube,
these six mirror planes are the diagonal planes which motivates the use of the
subscript d. The mirror planes convert the two-fold axes to produce four-fold
rotation reflection axes S4 . The group Td contains twentyfour elements. The
group Td has the International designation as 43m.

Th The group Th consists of the tetrahedral group adjoined by a mirror

49
A tetrahedron inscribed in a cube

Figure 25: The rotational symmetry operations of a tetrahedron are also found
amongst symmetry elements of a cube. The tetrahedron can be inscribed in a
cube.

plane which bisects the angle between the three-fold axes. For the tetrahedron,
these mirror planes are equivalent to the mirror planes of Td . However, for
the cube, the mirror plane is parallel to opposite faces and is mid-way between
them. There are only three such horizontal planes for the cube. The planes
bisect the angles between the three-fold axes, and, therefore, convert them into
six-fold rotation reflection axes. Since the group contains S6 , it also contains I.
Hence, Th = T × Ci . The group has twentyfour elements. The International
2
designation for the group Th is either m 3 or m3.

O The octahedral group has three mutually perpendicular four-fold axes.


There are four three-fold axes, and six two-fold axes. It has twentyfour ele-
ments. It has an International designation of 432.

Oh On adjoining a mirror plane to the octahedral group, one obtains Oh .


Adding a vertical mirror plane generates three other mirror planes. The effect
of a reflection in the vertical mirror plane followed by a rotation C4 is equivalent
to a reflection in a diagonal mirror plane. There are six of these diagonal mirror
planes. The C3 axes becomes S6 axes, just as in the case of Th . The group con-
4 2
tains fortyeight elements. The International designation is either m 3m or m3m.

50
C4

C3

C2

Figure 26: The rotational symmetry operations of an octahedron.

3.5 Bravais Lattices


There are an infinite number of possible lattices since there are no restrictions
on the primitive lattice vectors. However, only a few special types of lattices are
invariant under point group operations. These special types of lattices are called
Bravais lattices. The Bravais lattices can be categorized by their symmetries.
In three dimensions, there are fourteen Bravais lattices. The fourteen Bravais
lattices are organized according to seven crystal systems. These seven crystal
systems are classified according to the relations between their rotation axes.

The conventional unit cells have lattice vectors a, b, and c, of length a, b


and c, as shown in the figure. The angles between the vectors are denoted by
α, β and γ, such that α is the angle between b and c, etc. That is, α ( 6 b , c ),
β ( 6 a , c ), and γ ( 6 a , b ).

——————————————————————————————————

51
Octahedron inscribed in a cube

Figure 27: The rotational symmetry operations of an octahedron are equivalent


to the rotational symmetry elements of a cube. The octahedron can be inscribed
in a cube.

3.5.1 Exercise 4
Show that the volume of a conventional unit cell, Vc is given by
  12
2 2 2
Vc = a b c 1 + 2 cos α cos β cos γ − cos α − cos β − cos γ (70)

——————————————————————————————————

If the point group contains four three-fold axes C3 or (3), the system is cu-
bic. It is possible to choose three coordinate axes which are orthogonal to each
other and are perpendicular to the faces of a cube that has the four three-fold
axes as the body diagonals.

52
A conventional unit cell

γ b

α
β a

Figure 28: A conventional unit cell is specified by the lengths of the sides (a, b, c)
and the angles (α, β, γ) subtended by the pairs of sides.

3.5.2 Cubic Bravais Lattices.


The cubic Bravais lattices have the highest symmetry. The simple cube (P) has
three four-fold rotation axes and four three-fold axes, along with six two-fold
axes. There are three mirror planes that can be adjoined to the set of rotational
symmetry operations. The three four-fold rotation axes (C4 ) are mutually per-
pendicular and pass through the centers of opposite faces of the cube. Any
rotation which is an integer multiple of 24π will bring the cube into coincidence
with itself. The four three-fold axes (C3 ) pass through pairs of opposite ver-
tices of the cube. A rotation of any multiple of 23π will bring the cube into
coincidence with itself, as can be seen by inspection of the three edges at the
vertex which the rotation axis passes through. The six two-fold axes (C2 ) join
the mid-points of opposite edges. The highest symmetry group when mirror
symmetry is not included is the octahedral group O. The octahedral group O
contains twenty four symmetry operations. On adjoining a mirror plane to the
set of rotations of the octahedral group, one has the highest symmetry point
group which is labelled as Oh or m3m and has fortyeight symmetry elements.

The reason that the cubic group is called the octahedral group is explained
by the following observation. The group of symmetry operations of the cube
is equivalent to the group of symmetry operations on the regular octahedron.
This can be seen by inscribing an octahedron inside a cube, where each vertex

53
Rotational symmetry elements of a cube

C4

C3

C2

Figure 29: The rotational symmetry elements of a cube.

of the octahedron lies on the center of the faces of the cube. Thus, the cubic
point group is called the octahedral group O.

There are three types of cubic Bravais lattices: the simple cubic (P), the
body centered cubic (I) and face centered cubic (F) Bravais lattices.

The primitive lattice vectors for the simple cubic lattice (P) can be taken
as the three orthogonal vectors which form the smallest cube with the lattice
points as vertices. The three primitive lattice vectors have equal length, a, and
are orthogonal. The vertices of the cube can be labelled as (0, 0, 0), (0, 0, 1),
(0, 1, 0), (1, 0, 0), (1, 1, 0), (1, 0, 1), (0, 1, 1) and (1, 1, 1). The primitive cell is the
cube which contains one lattice point and has a volume a3 .

The body centered cubic Bravais lattice (I) has a lattice point at the ver-
tices of the cube and also one at the central point. The central point is located
at a2 (1, 1, 1) when specified in terms of the Cartesian coordinates formed by
the edges of the conventional non-primitive unit cell (which is the cube). The
primitive lattice vectors are given in terms of the Cartesian coordinates by
a1 = a2 (1, 1, −1), a2 = a2 (−1, 1, 1), a3 = a2 (1, −1, 1). These are the three
vector displacements originating at any lattice point which end up at one of
three neighboring body centers. The conventional unit cell contains two lattice
sites and has a volume a3 , where a is the length of the side of the cube. The

54
Conventional unit cell for the b.c.c. lattice

(1/2,1/2,1/2)

(0,0,0)

Figure 30: The conventional unit cell of a b.c.c. lattice.


primitive unit cell is a rhombohedron of edge a 2 3 which contains one lattice
site and has a volume 18 4 a3 . The angles between the primitive lattice vectors
is given by cos γ = − 13 . In the primitive cell, each body center of the conven-
tional unit cell is connected by three primitive lattice vectors to three vertices
of the conventional cell.

The Wigner-Seitz cell for the body centered cubic lattice is a truncated oc-
tahedron. It is made of eight hexagonal planes which are bisectors of the lines
joining the body center to the vertices. These eight planes are truncated by the
planes of the cube which coincide with the bisectors of the lines between the
neighboring body centers. The truncation produces the six square faces of the
body centered cubic Wigner-Seitz cell.

The face centered cubic Bravais lattice (F) consists of the lattice points at
the vertices of the cube and lattice points at the centers of the six faces. The
lattice points at the face centers are located at a2 (1, 1, 0), a2 (1, 0, 1), a2 (0, 1, 1),
a a a
2 (1, 1, 2), 2 (1, 2, 1) and 2 (2, 1, 1). The primitive lattice vectors point from the
vertex centered at (0, 0, 0) to the three closest face centers, a1 = a2 (1, 1, 0),
a2 = a2 (1, 0, 1), a3 = a2 (0, 1, 1). Since each face is shared by two adjacent
non-primitive unit cells there are 4 lattice sites in the conventional non-primitive
cubic unit cell. The primitive unit cell is a rhombohedron with side √a2 . The
edges of the primitive unit cell connect two opposite vertices of the cube via

55
The primitive lattice vectors for a b.c.c. Bravais lattice

a3
a2

a1

Figure 31: The primitive lattice vectors of a b.c.c. Bravais lattice.

the six face centers. The edges of the primitive cell are found by connecting the
vertex to the three neighboring face centers. The volume of the primitive unit
cell is found to be 14 a3 . The angles between the primitive lattice vectors are π3 .

The Wigner-Seitz cell for the face centered cubic lattice is best seen by trans-
lating the conventional unit cell by a2 along one axis. After the translation has
been performed, the unit cell has the appearance of being a cube which has
lattice sites at the body center and at the mid-points of the twelve edges of
the cube. The Wigner-Seitz cell can then be constructed by finding the twelve
planes bisecting the lines from the body center to the mid-points of the edges.
The resulting figure is a rhombic dodecahedron.

——————————————————————————————————

The presence of either one four-fold C4 , (4) or four-fold inversion rotation


axes S4 , (4), makes it possible to choose three vectors such that a = b 6= c,
α = β = γ = π2 and c is parallel to the C4 or S4 axes. This is the tetragonal
system.

56
The conventional unit cell for an f.c.c. lattice

(0,1,1)

(1,0,1)

(0,0,0)
(1,0,1)

Figure 32: The conventional unit cell for an f.c.c. lattice.

The primitive unit cell for a f.c.c. Bravais lattice

a3

a1

a2

Figure 33: The primitive lattice vectors for an f.c.c. lattice.

57
body centered tetragonal unit cell

α = β = γ = π/2

a
a

Figure 34: The body centered tetragonal lattice.

3.5.3 Tetragonal Bravais Lattices.


The tetragonal Bravais lattice can be considered to be formed from the cubic
Bravais lattices by deforming the cube, by stretching it, or contracting it along
one axis. This special axis is denoted as the c axis. Thus, the conventional unit
cell can be constructed, starting with a square base of side a, by constructing
edges of length c 6= a parallel to the normals of the base, from each corner.

The simple tetragonal Bravais lattice has a four-fold rotational axis and two
orthogonal two-fold axes. These symmetry elements generate the group D4 .
On adding a horizontal mirror plane to D4 , one obtains the highest symmetry
tetragonal point group which is D4h or 4/mmm with sixteen symmetry elements.

There are two tetragonal Bravais lattices: the simple tetragonal Bravais lat-
tice (P) and the body centered tetragonal Bravais lattice (I).

The face centered tetragonal lattice is equivalent to the body centered tetrag-
onal lattice. This can be seen by considering a body centered tetragonal lattice
in which the conventional unit cell can be described in terms of a side of length c
perpendicular to the square base of side a and area a2 . Consider the view along
the c axis which is perpendicular
√ to the square base. By taking a new base of
area 2 a2 and sides 2 a which are the diagonals of the original base, one finds
that the body centers can now be positioned as the face centers. That is, the

58
body centered tetragonal is equivalent to the face centered tetragonal unit cell.

The equivalence between the body centered and face centered structures does
not apply to the cubic system. However, the conventional body centered cubic
unit cell is equivalent to a face centered tetragonal unit cell in which the height
along the c-axis has a special relation to the side of the base. The equivalent
face centered tetragonal
√ unit cell has a height of a along the c-axis and the side
of the base is 2 a. Using the converse construction, the face centered cubic
unit cell can be shown to be equivalent to the body centered tetragonal lattice
with a particular length of the c-axis.

——————————————————————————————————

The orthorhombic system has three mutually perpendicular two-fold rota-


tion axes. The existence of the three mutually perpendicular two-fold rotation
axes is compatible with the point groups D2 , C2v and D2h . It is possible to
construct a unit cell α = β = γ = π2 .

3.5.4 Orthorhombic Bravais Lattices.


The conventional orthorhombic unit cell can be considered to be formed by de-
forming the tetragonal unit cell by stretching the base along an axis in the basal
plane. Thus, the base can be viewed as consisting of a rectangle of side a 6= b.
The unit cell has another set of edges which are parallel with the normal to the
base and have lengths c. Thus, the conventional unit cell has edges which are
parallel to three orthogonal unit vectors.

The simple orthorhombic lattice (P) has only two-fold rotation axes. The
two-fold axes are perpendicular, so the rotational group is D2 . The effect of ad-
joining a horizontal mirror plane converts D2 into the orthorhombic point group
with highest symmetry which is D2h or 2/mmm with four symmetry operations.

There are four inequivalent orthorhombic Bravais lattices. These are the sim-
ple orthorhombic lattice (P), the body centered orthorhombic (I), face centered
orthorhombic (F) and a new type of lattice, the c-side centered orthorhombic
lattice (C).

The c-side centered orthorhombic lattice (C) can be constructed from the
tetragonal lattice in the following manner9 . View the square net of side a, which
forms the bases of the tetragonal
√ unit cells, in terms of a non-primitive unit cell
with a square base of side 2 a with sides along the diagonal. This larger
non-primitive unit cell contains one extra lattice site at the center of the base
9 The nomenclature used is that, if the two faces of the unit cell which have the same normal

as the a − b plane are centered, then the structure is designated as C. Likewise, if the faces
parallel to the a − c plane are centered, the structure is designated by B, etc.

59
Primitive monoclinic unit cell

b π/2
γ π/2
a mP

Figure 35: The primitive monoclinic lattice.

and the top. When this base centered tetragonal structure √ is then stretched
along one of its sides (one of the diagonal sides of length 2 a), one obtains the
orthorhombic base-centered lattice.

——————————————————————————————————

The monoclinic lattice system requires a minimum of one two-fold rotation


axis. Due to the conditions imposed by the two-fold rotation symmetry, it is
possible to choose α = β = π2 6= γ. The monoclinic systems are compatible
with the point groups C2 , Cs and C2h .

3.5.5 Monoclinic Bravais Lattice.


The monoclinic Bravais lattice is obtained from the orthorhombic Bravais lat-
tices by distorting the rectangular base perpendicular to the c axis into a par-
allelogram. The base is a parallelogram, and the two basal lattice vectors are
perpendicular to the c axis.

The simple monoclinic lattice (P) has a two-fold axis parallel to the c axis.
The rotational group is C2 . If a horizontal mirror plane is added to C2 , then
one finds that the most symmetric monoclinic point group is C2h or 2/m which
has four elements.

60
Body-centered monoclinic unit cell

γ
mI

Figure 36: The body centered monoclinic lattice.

There are two types of monoclinic Bravais lattices: the simple monoclinic (P)
and the body centered monoclinic Bravais lattice (I). The two types of mono-
clinic Bravais lattices correspond to the two types of tetragonal Bravais lattices.
The four orthorhombic lattices collapse onto two lattices in the tetragonal and
monoclinic systems, as the centered square net is not distinct from a square net.
Likewise, the centered parallelogram is not distinct from a parallelogram.

——————————————————————————————————

The groups C1 and Ci impose no specific restrictions on the lattice. This is


the triclinic lattice system.

3.5.6 Triclinic Bravais Lattice.


The triclinic Bravais lattice is obtained from the monoclinic lattice by tilting
the c axis so that it is no longer orthogonal to the base. There is only the simple
triclinic Bravais lattice (P). The three axes are not orthogonal and the sides are
all different.

Apart from inversion which is required by the periodic translational invari-


ance of the lattice, the triclinic lattice has no special symmetry elements. The

61
Primitive monoclinic unit cell

γ
mC ≡ mP

Body-centered monoclinic unit cell

γ
mF ≡ mI

Body-centered monoclinic unit cell

mC ≡ mI

Figure 37: The base centered monoclinic lattice is equivalent to the primitive
monoclinic lattice. The face centered monoclinic lattice is equivalent to the body
centered monoclinic lattice. The side centered monoclinic lattice is equivalent
to the body centered monoclinic lattice.

62
The Triclinic Lattice

α c
γ a b aP

Figure 38: The primitive triclinic lattice.

point group of highest symmetry is Ci or 1 which has two elements.

——————————————————————————————————

The presence of only one three-fold axes, either C3 (3) or S6 (3), produces
the trigonal system. There are two types of trigonal systems. In one of the
trigonal systems, a primitive unit cell may be chosen with a = b = c and
α = β = γ such that the three-fold axes is along the body diagonal. The
other trigonal system has a = b 6= c and α = β = π2 and γ = 23π . This
latter system is denoted as the hexagonal system.

3.5.7 Trigonal Bravais Lattice.


The trigonal Bravais lattice is a deformation of the cube produced by stretching
it along the body diagonal. The lengths of the sides remain the same and the
three angles between the sides are all identical. There is only one trigonal Bra-
vais lattice. The point symmetry group is D3h or 62m with twelve symmetry
operations.

The body centered cubic and face centered cubic Bravais lattices can be con-
sidered to be special cases of the trigonal lattice. For these cubic systems, the
sides of the primitive unit cells are all equal and the angles are 109.47 degrees

63
The Triclinic Lattice

aC ≡ aP

The Triclinic Lattice

aI ≡ aP

The Triclinic Lattice

aF ≡ aP

Figure 39: The base centered, body centered and face centered triclinic lattice
are equivalent to primitive lattices.

64
The Trigonal unit cell

a=b=c
α=β=γ

Figure 40: The primitive trigonal unit cell.

Conventional Trigonal unit cell

Figure 41: The conventional trigonal unit cell.

65
A

Figure 42: The relationship between the conventional unit cell of the trigonal
Bravais lattice and the primitive unit cell.

for the b.c.c. structure and 60 degrees for the f.c.c. structure.

The trigonal unit cell has two vertices on the extremes of the body diagonal.
The remaining six vertices are arranged on two equilateral triangles. These two
triangles are related by a translation along the body diagonal and a rotation by

6 around the body diagonal. In the trigonal lattice, all the vertices of the unit
cell are arranged on equilateral triangles which form two-dimensional hexagonal
nets. The difference between the trigonal lattice and the hexagonal lattice is
merely due to the different stacking of the hexagonal planes.

——————————————————————————————————

The presence of either a six-fold axes C6 (6) or a rotation - inversion axes


S3 (6), indicates that the system is hexagonal. The hexagonal unit cell has
a = b 6= c and α = β = π2 and γ = 23π .

3.5.8 Hexagonal Bravais Lattice.


The hexagonal Bravais lattice has a unit cell in which the base has sides of
equal length, inclined at an angle of 23π with respect to each other. The c axis
is perpendicular to the base.

The hexagonal system has a point group D6h or 6/mmm which has twenty-

66
The Hexagonal Bravais Lattice

a3

a2
hP a1

Figure 43: The hexagonal Bravais lattice and the primitive lattice vectors.

four symmetry elements.

There is only one hexagonal Bravais lattice. The primitive unit cells are
rhombic prisms which can be stacked to build the hexagonal non-primitive unit
cell. The six-fold rotational symmetry of the hexagonal Bravais lattice is most
evident by inspection of the non-primitive unit cell.

The primitive lattice vector are given in terms of Cartesian coordinates by

a1 = a êx

 
a
a2 = êx + 3 êy
2
a3 = c êz (71)

Possible types of unit cells for a Bravais lattice.

67
Type of unit cell and Symbol Location of Non-origin basis point No. of Lattice Points

Primitive (P) - 1
Body-centered (I) The cell center 2
Side-centered (A) Centers of the A face (1, 0, 0) 2
Side-centered (B) Centers of the B face (0, 1, 0) 2
Side-centered (C) Centers of the C face (0, 0, 1) 2
Face-centered (F) Centers of all faces 4

68
Relation between conventional and primitive lattice vectors.

I
a1 = 12 ( − a + b + c )
a2 = 12 ( a − b + c )
a3 = 12 ( a + b − c )

F
1
a1 = 2 (b + c)
1
a2 = 2 (a + c)
1
a3 = 2 (a + b)

R
a1 = 13 ( 2 a + b + c )
a2 = 13 ( − a + b + c )
a3 = 13 ( − a − 2 b + c )

A
a1 = a
a2 = b
a3 = 12 ( b + c )

B
a1 = a
a2 = b
a3 = 12 ( c + a )

C
a1 = a
1
a2 = 2 ( a + b )
a3 = c

69
In summary the following structures were found:

The Fourteen Bravais Lattices.

Bravais Lattice Lattice Symbol Characteristic Symmetry

Cubic cP, cI, cF a = b = c Four three-fold axes


π
α = β= γ = 2 along body-diagonals

Tetragonal tP, tI a = b 6= c One four-fold axes


π
α = β = γ = 2

Orthorhombic oP, oF, oC, oI a 6= b 6= c Three two-fold axes


π
α = β = γ = 2 mutually perpendicular

Monoclinic mP, mI a 6= b 6= c One two-fold axis


α = β = π2 6= γ

Triclinic aP a 6= b 6= c one-fold axis only


π
α 6= β 6= γ 6= 2 (identity or inversion)

Trigonal hR a = b = c One three-fold axis only


α = β = γ < 2π
3 , 6=
π
2

Hexagonal hP a = b 6= c One six-fold axis only


α = β = π2 , γ = 2π
3

This completes the discussion of the set of fourteen Bravais lattices. In order
to specify crystal structures, it is necessary to associate a basis along with the
underlying Bravais lattice. The addition of a basis can reduce the symmetry of
the crystal from the symmetry of the Bravais lattices. This results in thirty two
point groups, and by adjoining the translations and combined operations, one
finds the two hundred and thirty space groups.

——————————————————————————————————

3.5.9 Exercise 5
Form a table of the number of the n-th nearest neighbors and the distances
to the n-th neighbors for the face centered cubic (f.c.c.), body centered cubic

70
(b.c.c.) and simple cubic (s.c.) lattices, for n = 1, n = 2 and n = 3.

——————————————————————————————————

Distribution of types of unit cells amongst the Crystal Systems.

Type of unit cell and Symbol Crystal System Distribution No. of Lattices

Primitive (P) One in each of the seven crystal systems 7


Body-centered (I) Cubic, Tetragonal, Orthorhombic, Monoclinic 4
Side-centered (A-B-C) Orthorhombic 1
Face-centered (F) Cubic and Orthorhombic 2

Total 14

Having just used symmetry to enumerate all the possible Bravais lattices,
we shall now discuss the possible symmetries of crystals. Due to the addition of
the basis, the point group symmetry of a crystal can be different from the point
group symmetry of the Bravais lattice.

71
3.6 Point Groups
The addition of a basis to a lattice can result in a reduction of the symmetry of
the point group. Here, the point groups are enumerated according to the Bra-
vais lattice types and by the Schoenflies designation followed by the appropriate
(International) symbol.

The cubic system with a basis can have the point symmetry group of either
Oh (m3m), O (43), Td (43m), Th (m3) or T (23).

The tetragonal system can have point group symmetry of D4h (4/mmm),
D4 (42), C4v (4mm), C4h (4/m) or C4 (4).

The orthorhombic system can have point group symmetry of either D2h
(mmm), D2 (222) or C2v (2mm).

The monoclinic system can exist with point group symmetry of either C2h
(2/m), C2 (2) and Cs (m). The group Cs only consists of the identity and a
reflection operation.

The triclinic system only contains C1 (1) and Ci (m). The group Ci only
consists of the identity and the inversion operation.

The trigonal system has the point groups D3h (62m), D3 (32), C3v (3m), S6
(3), or C3 (3).

The hexagonal system has the point groups D6h (6/mmm), D6 (62), C6v
(6mm), C6h (6/m), or C6 (6).

The other four remaining groups are: The groups C3h (6) and D3d (3m)
which are usually included in the hexagonal system; and finally, there are the
groups S4 (4) and D2d (42m) which are included with the tetragonal systems.

This completes the enumeration of the thirtytwo point groups.

——————————————————————————————————

3.6.1 Exercise 6
When point group symmetry operations are adjoined to the translations through
lattice vectors, new symmetry elements arise. These new operations are like the
point group operations but involve other invariant points.

Prove that a rotation about an axis Cn , followed by a translation through a


vector R which is perpendicular to the axis, is equivalent to a rotation Cn about

72
a parallel axis located on the bisecting plane of R at a distance d perpendicular
to the axis. Show that the distance d is given by
1 π
d = |R | cot (72)
2 n

A consequence of this theorem is that, if there is a two-fold rotational axis


C2 which is perpendicular to a primitive lattice vector a, then there must be
another two-fold axis of rotation passing through 12 a.

——————————————————————————————————

3.7 Space Groups


On combining the point group symmetry operations with lattice translations,
one can generate 230 space groups. Often, the space group is composed from
symmetry operations of the point group and symmetry operations that are
translations by the vectors of the direct lattice. These space groups are called
symmorphic groups. Lattices with symmorphic space groups can be constructed
by attaching bases with the various point group symmetries on the various Bra-
vais lattices. For example, a basis which has the symmetry of any one of the five
cubic point groups can be placed on the three cubic Bravais lattices, yielding
a total of fifteen cubic space groups. Likewise, bases with the symmetries of
the seven tetragonal point groups can be placed on the two tetragonal Bravais
lattices, yielding fourteen tetragonal space groups. This process only leads to
61 different space groups. In the other cases, the space groups contain two new
types of symmetry operations that cannot be compounded from translations
by Bravais lattice vectors and operations contained in the point groups. These
groups are non-symmorphic. The new types of symmetry operations occur when
there is a special relation between the basis dimensions and the size of the Bra-
vais lattice. These new symmetry elements include :

Screw Axes. A screw operation is a translation by a vector, not in the


set of Bravais lattice vectors, which is followed by a rotation about the axis
defined by the translation vector. A screw symmetry is denoted by nm , where
n represents the rotations 2nπ , where n = 2 , 3 , 4 , 6 and m represents the
number of translations by lattice vectors which produce one complete rotation
by 2 π. Thus, n screw operations nm , each producing a rotation of 2nπ , result
in a total translation through m lattice spacings. In other words, the subscript
m indicates a translation of m
n lattice spacings produced by one screw operation.

Glide Planes. A glide operation is composed of a translation by a vector,


not in the Bravais lattice, which is followed by a reflection in a plane containing
the translation vector. Glide planes are denoted by either a, b or c (according to

73
21 Screw Axis

21

π
c/2
c

31 Screw Axis

31

2π/3

2π/3
c
c/3

32 Screw Axis

32
2π/3
c

2π/3

2c/3


Figure 44: Screw operations nm , consists of rotations by n and a translation
by m
n lattice spacings.

74
whether the translation is along the a, b or c axis). Also, there are glide planes
which are denoted n or d (the diagonal or diamond glides) that are special cases
involving translations along more than one axis.

——————————————————————————————————

3.7.1 Exercise 7
An axial glide reflection is one in which the translation is parallel to a primitive
unit lattice vector, say c. Prove that in this case, the only allowable axial glides
involve a translation through c/2.

——————————————————————————————————

The hexagonal close-packed lattice structure has both glide and screw types
of non-symmorphic symmetry operations. The hexagonal close-packed struc-
ture can be described by a three-dimensional unit cell which contains a centered
hexagonal base, and which has an identical centered hexagonal top located at
vertical distance c directly above the base. If one considers the base hexagon to
be formed by six equilateral triangles, then there are lattice sites at the vertex
of each triangle. These lattice sites form a triangular net in the basal plane
and there is a similar triangular net in the upper plane. These lattice sites are
designated as the A sites. There is a second net of triangles at a distance 2c
vertically over the base. The centers of the mid-plane equilateral triangles are
located directly over the (central) lattice sites of the base. There are two pos-
sible orientations for these triangles. On choosing any one orientation, the set
of lattice sites on this mid-plane are located such that they lie directly over the
centers of every other equilateral triangle in the base. These mid-plane lattice
sites are designated as the B sites.

Consider a line, parallel to the c axis. The line is equidistant to two neigh-
boring B lattice sites and is equidistant to the two A lattice sites that form
the section of the perimeter of the basal hexagon which is parallel to the line
connecting the above two B lattice sites. Viewed from the c axis, the vertical
line passes through the center of the rectangle formed by the two A and two
B lattice sites. This line is the screw axis. The screw operation 21 consists of
a translation by 2c followed by rotation of π, and brings the A hexagons into
coincidence with the sites of the B hexagons.

The glide planes can also be found by considering the projection of the lattice
along the c axis. A line can be constructed which connects any two of the three
B sites inside the hexagonal unit cell. Then a parallel line can be constructed
which connects a pair of neighboring A sites that forms part of the perimeter
of the hexagonal base. Since this line is on the perimeter of the unit cell, it
is equivalent to the parallel line segment connecting A sites at the opposite

75
Hexagonal close-packed lattice

C3
21-screw axis

Figure 45: The screw-axis in the hexagonal close-packed structure.

Hexagonal close-packed lattice

c-glide σ
l

Figure 46: The glide-plane in the hexagonal close-packed structure.

76
boundary. Consider the pair of parallel lines, one which connects the B sites,
and the other which is the closest line segment that connects the A sites on the
perimeter of the base hexagon. The projection of the glide plane along the c
axis is parallel to and equidistant from the above pair of lines. The glide op-
eration is a translation by 2c along the c axis followed by a reflection in the plane.

There are two different systems of nomenclature for space groups: one is due
to Schoenflies and the other is due to Hermann and Mauguin. The Hermann -
Maugin space group nomenclature consists of a letter P , I , F , R , C which
describes the Bravais lattice type followed by a statement of the essential sym-
metry elements that are present. As an example, the space group P 63 /mmc has
a primitive (P ) hexagonal Bravais lattice with point group symmetry 6/mmm.
Another example is given by the space group P ba2 which represents a primitive
(P ) orthorhombic Bravais lattice and has a point group of mm2 (the a and b
glide planes being simple mirror planes in point group symmetry).

There is some arbitrariness in the distinction between the trigonal and hexag-
onal crystal systems. While it is true that the trigonal point group is a sub-
group of the hexagonal point group, the trigonal lattice cannot be obtained by
infinitesimal distortion of the hexagonal lattice. As a result, the space groups
of the hexagonal and trigonal lattices are listed together.

Distribution of Space Groups.

Crystal System Number of Space Groups

Cubic 36
Tetragonal 68
Orthorhombic 59
Monoclinic 13
Triclinic 2
Hexagonal - Trigonal 52

The International Tables for Crystallography10 includes listings of crystal


structures for each space group.

10 T. Hahn, editor, International Tables for Crystallography, Kluwer publishers, Utrecht,

Holland (1996).

77
3.8 Crystal Structures with Bases.
Crystal structures are specified by giving a basis and a Bravais lattice. The
basis is specified by the positions of and types of atoms in the unit cell.

Sometimes it is also useful to specify the local coordination polyhedron


around each inequivalent site in the lattice. This provides information about
the local environment of the atom which is important for bonding. Small de-
formations in the positions of the atoms can lower the symmetry of a crystal
structure but usually does not affect the connectivity or topology of the atoms.
Therefore, slight deformations of the local environment are often specified by
the same local coordination polyhedra. The local coordination polyhedra have
been enumerated by W. B. Jensen11 , and by Villars and Daams12 .

3.8.1 Diamond Structure


The diamond lattice is formed by the carbon atoms in a diamond crystal. The
structure is cubic, and has the space group F d3m. The underlying Bravais lat-
tice is the face centered cubic lattice and has a two atom basis. In the diamond
structure, both atoms are identical. They are located on the f.c.c. Bravais lat-
tice site (0, 0, 0) and at a second site displaced from the Bravais lattice site by a
distance a( 14 , 14 , 14 ) in terms of the Cartesian coordinates of the conventional unit
cell. There are four lattice points corresponding to the sites of the conventional
f.c.c. unit cell. There are also four interior points which are displaced from the
Bravais lattice points by the basis vector a( 14 , 14 , 14 ). Thus, the diamond struc-
ture consists of two interpenetrating face centered cubic lattices with C atoms
on each lattice site. Diamond possesses a center of inversion located half-way
between the origins of the two interpenetrating f.c.c. lattices. This is a glide-
like inversion operation. The center of inversion is located at a( 18 , 18 , 18 ). When
this is chosen as the origin, the crystal is symmetric under the transformation
r → − r.

Each atom is covalently bonded to four other atoms. The bonds on the two
inequivalent basis sites point in different directions. The four neighboring atoms
form a tetrahedron centered on each atom. The diamond lattice is most stable
for compounds in which the bonds are highly directional. Directional covalent
bonding is often found in the elements of column IV of the periodic table. In
particular, Carbon, Silicon and Germanium can crystallize in the diamond struc-
ture. The great strength of diamond is a consequence of the three-dimensional
network of strong covalent bonds. The diamond structure is relatively open as
the packing fraction is only 0.34.

11 W. B. Jensen, The Structures of Binary Compounds, North Holland publishers (1988).


12 P. Villars and J. L. C. Daams, Journal of Alloys and Compounds, 197, 177 (1993).

78
Diamond Structure

cF

Figure 47: The diamond structure can be considered as being composed of two
interpenetrating f.c.c. lattices.

——————————————————————————————————

3.8.2 Exercise 8
Find the angles between the tetrahedral bonds of diamond.

——————————————————————————————————

3.8.3 Graphite Structure


Graphite is the stable form of Carbon. Graphite has a hexagonal unit cell and
has the space group P 63 /mmc. The primitive lattice vectors may be represented
by

a1 = 3 a êx

3 3
a2 = a êx + a êy
2 2
a3 = c êz (73)

where a is the distance between adjacent atoms in the x − y plane. The atoms

79
Diamond with bonds oriented along the (1,1,1) direction

Figure 48: The diamond structure has atoms on two inequivalent sites which
have bonds that are oriented in different directions.

a2

a1

Figure 49: A vertical projection of the unit cell of graphite. The primitive lattice
vectors a1 , a2 and the vertices of the unit cell in the basal plane are drawn in
green. The positions of the other C atoms in the basal plane are drawn in blue.

are located at [0, 0, z] and [0, 0, 12 + z] where z ≈ 0, where the coordinates


are given in terms of the primitive lattice vectors. Another two atoms are lo-
cated at the positions [ 23 , 23 , z] and [ 31 , 13 , 12 + z], where z ≈ 0. The structure is

80
Crystal Structure of Graphite

Figure 50: The primitive unit cell of the graphite structure.

formed in layers, where each atom is bonded to three other atoms in the plane,
thereby forming a two-dimensional hexagonal network. The central site of the
two-dimensional hexagonal ring is open. The side of the hexagon is of length
a. The stacking sequence of the layers just corresponds to a translation of one
layer by [ 13 , 13 , 12 ] with respect to the other, such that one C atom lies above the
hexagonal hollow in the layer below. The layers are relatively far apart and as
is expected, there is only weak van der Waals bonding between the layers. This
structure explains the cleavage and other characteristic properties of graphite.

Carbon may crystallize into either as a diamond lattice or as graphite, under


different conditions. This is an example of polymorphism which is quite common
among the elements. Diamonds are not forever as they actually are an unstable
form of C under ambient conditions, although the rate of transformation to the
stable form (graphite) is exceedingly slow.

Boron and Nitrogen, which occur on either side of Carbon in the periodic
table, form compounds which have properties that are strikingly similar to Car-
bon. The Boron and Nitrogen atoms can be bonded in either planar structures

81
Crystal Structure of Graphite

ey
ex

Figure 51: The crystal structure of graphite can be considered as being con-
structed by successively stacking open hexagonal layers.

like graphite, or tetrahedral structures, like diamond. The tetrahedral bonded


Boron - Nitrogen materials have extremely high melting points and hardness,
and have great importance in materials engineering.

——————————————————————————————————

3.8.4 Exercise 9
There are two forms of graphite. The most common form is hexagonal graphite
which has a stacking sequence A − B − A − B. The other form of graphite is
rhombohedral graphite. This is based on a trigonal form which has a stacking
sequence A − B − C − A − B − C. Describe the primitive unit cells for the two
forms of graphite. How many atoms are in the primitive unit cells of graphite?

——————————————————————————————————

82
3.8.5 Hexagonal Close-Packed Structure
The hexagonal close-packed structure has hexagonal symmetry and the space
group is P 63 /mmc. It is described as a hexagonal Bravais lattice with a two

The unit cell of the hexagonal close-packed structure

ey

ex
a

Figure 52: The hexagonal close-packed structure can be considered as a hexag-


onal Bravais lattice with a two-atom basis.

atom basis. The two basis atoms are identical and one is positioned at [0, 0, 0]
which is at the vertex of the primitive lattice cell, and the other atom is located
at [ 13 , 13 , 12 ] as expressed in terms of the primitive lattice vectors. (The square
brackets indicate that the direction in the direct lattice are specified with respect
to the primitive lattice vectors.) The primitive lattice vectors are

a1 = a êx

 
a
a2 = êx + 3 êy
2
a3 = c êz (74)

Thus, the hexagonal close-packed structure has a basis of two atoms: one at
r1 = (0, 0, 0) and the other at
 
1 1
r2 = a1 + a2 + a
3 2 3
a a c
= êx + √ êy + êz (75)
2 2 3 2

83
The Hexagonal Bravais Lattice

a3
a2

a1

Figure 53: The hexagonal Bravais lattice.

Since a1 and a2 are inclined at an angle π3 , the structure can be considered to be


formed by two interpenetrating simple hexagonal Bravais lattices. Alternately,
the structure may be viewed as being formed by stacking two-dimensional tri-
angular lattices above one another with a separation between the layers of half
the height of the unit cell. Each atom has twelve nearest neighbors: six within
the hexagonal plane and three in each of the planes above and below the atom.

The name hexagonal close-packed comes from thinking of this structure as


being formed from hard spheres with radii a2 and forming a close-packed hexag-
onal layer. The second layer is formed by stacking a second hexagonal layer of
atoms above the first. However, the centers of the atoms in the second layer
are positioned above the dimples in the first layer. There are two sets of dim-
ples between the atoms so there are two different choices for placing the second
layer of atoms. The third layer is stacked such that the centers of the atoms
are directly above the centers of the atoms of the first layer, and the fourth is
stacked directly over the second layer, etc. Thus, there are two interpenetrating
hexagonal lattices displaced by
1 1 1
a + a + a (76)
3 1 3 2 2 3
or [ 13 , 13 , 12 ].

84
A A
C
B B
A A A
C C
B
A A

Figure 54: The vertical projection of the hexagonal close-packed stacked struc-
ture. The close-packed layer of spheres in the basal plane are centered at the A
sites. The next layer of atoms are centered over the sites B (or equivalently C),
and the third layer is centered over the A sites. Hence, the layers are stacked
in the sequence A − B − A − B etc.

There are a total of twelve nearest neighbor atoms which are distributed as
6 neighbors in the plane, 3 in the plane above, and 3 in the plane below. This
gives a total of twelve nearest neighbor atoms.

On assuming that the atoms are hard spheres with radii r and that√ the
spheres are touching, the lattice constants satisfy a = b = 2 r and c = 4 √23 r.
This yields the hexagonal close-packed structure and has the particular ratio of
the c to the a axis lengths of
r
c 8
= = 1.633 (77)
a 3
This is the ideal c to a ratio. Hexagonal close-packed systems with the ideal
ratio have a packing fraction of 0.74. As atoms are not hard spheres, there is
no reason for this value of the c to a ratio to be found in naturally occurring
crystals, and deviations from the ideal value are found most frequently. Only
He has the ideal c to a ratio.

The most frequently occurring structures are the close-packed structures.


These are the hexagonal close-packed, face centered cubic and body centered
cubic structures, which have packing fractions of 0.74, 0.74 and 0.68, respec-
tively. Both simple and transition metals frequently form in the hexagonal
close-packed structure, or other close-packed structures.

85
——————————————————————————————————

3.8.6 Exercise 10
c
Show that the a ratio for an ideal hexagonal close-packed lattice structure is
  12
c 8
= (78)
a 3

——————————————————————————————————

3.8.7 Exercise 11
N a transforms from b.c.c. to h.c.p. at 23 K via a Martensitic transition. On
assuming that the density remains constant and the h.c.p. structure is ideal,
find the h.c.p. lattice constant a in terms of the b.c.c. value a0 .

——————————————————————————————————

3.8.8 Other Close-Packed Structures


One can form other close-packed structures by altering the sequence of stacking
of the close-packed layers. The hexagonal close-packed can be characterized
by the repeated stacking sequence A - B - A - B etc. That is, the atoms in
the planes above and below the triangular lattice have centers directly over the
dimples and each other, thereby creating a two layer unit cell.

Another stacking sequence is given by A - B - C in which the unit cell


consists of three layers. The A and C layers have the atoms centered on the two
inequivalent sets of triangular dimples of the B layer. This close-packed stacking
corresponds to the face centered cubic lattice. The packing fraction of the face
centered cubic lattice and hexagonal close-packed lattice are identical. The
triangular close-packed nets are the planes perpendicular to the body diagonal
of the conventional f.c.c. unit cell. There are two such planes which pass through
the conventional unit cell and two further planes that each just graze one vertex
of the unit cell. The intercepts of the planes with the conventional (Cartesian)
axes are (1, 0, 0), (0, 1, 0) and (0, 0, 1). The next plane has intercepts (2, 0, 0),
(0, 2, 0) and (0, 0, 2). The sets of planes are known as {1, 1, 1} planes and are
composed of triangular arrays of atoms, where the sides of the triangles have
lengths √a2 . The normal to the planes are in the direction [1, 1, 1] i.e.
 
1
n̂ = √ êx + êy + êz (79)
3

86
Conventional face-centered cubic unit cell

Figure 55: The conventional unit cell of the f.c.c. structure. The diagonal planes
are seen to consist of close-packed layers.

where ê are the orthogonal unit vectors of the conventional cell. The equations
of the planes are  
r − m a êx . n̂ = 0 (80)

where m is an integer that labels the plane by the intercept with the x axis.
The quantity m is related to the perpendicular distance, s, between the plane
and the origin by
a
s = m √ (81)
3
for integer m.

It is convenient to introduce three new orthogonal unit vectors to describe


the positions of the atoms in the planes. The first is n̂ the normal to the planes
 
1
n̂ = √ êx + êy + êz (82)
3
The other vectors ê1 and ê2 are chosen to be vectors in the planes. These form
a new set of Cartesian non-primitive lattice vectors which are defined by
 
1
ê1 = √ êx − êy (83)
2

87
which corresponds to the face diagonal of the conventional unit cell that lies in
the triangular plane and
 
1
ê2 = √ êx + êy − 2 êz (84)
6
which is the “lateral” direction in the triangular plane. The lateral displace-
ments of atoms between one triangular plane, say the plane which passes through
the atom at ( 12 , 12 , 0), and the atoms on the next plane (centered on the origin
(0, 0, 0)) can be written as
 
a a
∆r = êx + êy − n̂ √
2 3
 
a
= êx + êy − 2 êz
6
1 a
= √ √ ê2 (85)
3 2
This can be re-written as √
2 3 a
√ ê2 (86)
3 2 2

as √a2 is the triangular lattice constant and 23 √a2 is the height of the triangle.
Thus, the atoms in consecutive planes are displaced “laterally” by 0, 23 , and 43
and this sequence then repeats. The resulting structure has layers which have
a stacking sequence A − B − C − A − B − C etc.

There are other possible stacking sequences, with longer periodicities. The
earlier lanthanides and late actinides have a stacking sequence A - B - A - C
with four layers per unit cell, however, the Sm structure only repeats itself af-
ter nine layers. The longest known periodicity is 594 layers which is found in
a polytype of SiC. The long-ranged crystallographic order is not due to long-
ranged forces, but is caused by spiral steps caused by dislocations in the growth
nucleus. There is also the possibility of random stacking sequences.

3.8.9 Sodium Chloride Structure


The Sodium Chloride or N aCl structure is cubic. The space group is F m3m.
It has an ordered array of N a and Cl ions located on the sites of a simple cubic
lattice of linear dimension a2 . Each type of ion is surrounded by six ions of the
opposite charge located at a distance a2 away. The twelve next nearest neigh-
bors have like charge and are located at a distance √12 a away along the face
diagonals of the cubic unit cell. There are four units of N aCl in the unit cell.
The structure may be most efficiently visualized as having the N a+ ions located
on the sites of a face centered cubic lattice which has its origin at (0, 0, 0) and
the Cl− ions are located on the sites of an interpenetrating face centered cubic

88
The Sodium Chloride Structure

cF

Figure 56: The Sodium Chloride structure can be considered as an f.c.c. Bravais
lattice with a two-atom basis.

lattice which has its origin at the center of the cubic unit cell ( 12 , 12 , 12 ).

The Sodium Chloride structure is favored by many ionic compounds. In this


structure, the electrostatic interactions are balanced by the short-ranged repul-
sive interactions due to the finite size of the ions. The short-ranged repulsions
are due to the Pauli exclusion principle. The sizes of the ions are important
in determining the stability of this structure. If the ions of opposite charge are
envisaged as just touching, then the ionic radii must satisfy the equality
 
+ −
a = 2 r(N a ) + r(Cl ) (87)

Ions of the same type are closest along the face diagonals, so if they do not
touch, the lattice constant satisfies the inequality
1
√ a > 2 r(Cl− ) (88)
2
Combining the above two equations yields an inequality for the ratio of the ionic
radii of the ions
r(Cl− ) √
+
≤ 1 + 2 (89)
r(N a )
If this inequality is not obeyed, the Pauli forces render the structure unstable.

89
Examples of materials that form in the N aCl structure are the alkali halides
made from the alkaline elements Li, N a, K, Rb or Cs with a halide element
F , Cl Br or I. Alternatively, one can go to the neighboring columns of the
periodic table and combine M g, Ca, Sr or Ba with a chalcogen O, S, Se or T e
to form the N aCl structure.

3.8.10 Cesium Chloride Structure


The ionic compound Cesium Chloride or CsCl has a cubic structure. The space
group is P m3m. The Cs+ ion is located at (0, 0, 0) and the Cl− ion at the body
center of the cube ( 12 , 12 , 12 ). Thus, the CsCl structure resembles a body centered
cubic structure in which one type of atom is at the simple cubic sites and the
other type of atom is at the body center. Each ion is surrounded by eight atoms

Cesium Chloride Structure

cP

Figure 57: The Cesium Chloride structure can be considered as a simple cubic
Bravais lattice with a two-atom basis.

of opposite charge located at a distance 23 a away, which corresponds to half
the length of the body diagonal of the cube. Each atom has six neighbors of
similar charge located a distance a away. The ratio of the ionic radii required
for this structure to be possible is

r(Cl− ) ( 3 + 1)
≤ (90)
r(Cs+ ) 2

90
If the radii ratio is greater than 1.366, but less than 2.42, ionic compounds pre-
fer the N aCl structure.

Examples of compounds that form the CsCl structure are the Cs halides,
T l halides, CuZn (beta brass), CuP d, AgM g and LiHg.

Linus Pauling has produced a set of empirical rules which determine the
coordination numbers in terms of the ionic radii of the ions. If one assumes
that the anion adopts the cubic close-packed structure (f.c.c.), there are three
types of holes between the close-packed spheres and each type of hole has a
different size. It is assumed that the cations fit into one set of holes. The
central void of the conventional f.c.c. unit cell is surrounded by an octahedron
and, therefore, has a coordination number of six. There are also tetrahedral
holes with coordination number four. The tetrahedral holes are located near
the 8 corners of the f.c.c. cube, and the vertices of the tetrahedra are located at
the corner and the three neighboring face centers. Alternatively, the tetrahedral
holes can be seen by considering an octant of the f.c.c. cube. The tetrahedral
hole site is at the center of the octant, and the four vertices of the tetrahedron
are located at four of the octant’s eight corners. There are twelve trigonal
holes which are located near the eight vertices of the conventional unit cell. The
trigonal sites reside in the planes formed by the vertex and any two of the closest
face centers. The radius ratio rule suggests that the structure is determined by
maximizing the coordination numbers while keeping ions of opposite charge in
contact. This procedure seems likely to maximize the electrostatic attraction
energy. By considering the geometry of the holes, one expects that certain
structures will be stable for different values of the radius ratio
r(X − )
rr = (91)
r(R+ )

For the tetragonal sites, by considering the body diagonal of the octant, one
expects that   √
− + 3
2 r(X ) + r(R ) = a (92)
2
and by considering the face diagonal
a
2 r(X − ) < √ (93)
2
Hence, we find the tetragonal hole has the limiting radius ratio of
r
r(X − )

3
> 2 + 1 (94)
r(R+ ) 2

In particular, the radius ratio rules suggest that the range of radii ratios where

91
the various f.c.c. based configurations are stable are given by

Name Coordination No.

6.45 > rr > 4.45 trigonal 3


4.45 > rr > 2.41 tetrahedral 4
2.41 > rr > 1.37 octahedral 6

If the atoms have comparable sizes, then it is necessary to consider more open
structures with higher coordination numbers such as simple cubic. For the sim-
ple cubic structure, the coordination number of the central hole is eight and
the hole size is larger, so 1.37 > rr. Thus, since rr ∼ 1.8 for N a and Cl,
it fits the radius ratio rules as being octahedrally coordinated like in the N aCl
structure. On the other hand, for Cs and Cl where the ions have comparable
sizes, the radius ratio is rr ∼ 1.07 which is compatible with the cubic hole
structure found in CsCl.

3.8.11 Fluorite Structure


Fluorite or CaF2 has a cubic structure. The space group is F m3m. Ionic
compounds of the form RX2 , in which the ratio of the ionic radii r satisfy the
inequality √
r(X − ) ( 3 + 1)
≤ (95)
r(R2+ ) 2
can form the fluorite structure. The unit cell has four Ca2+ ions, one at the
origin and the others are located at the face centers of the cube. The eight F −
ions are interior to the cube. The F − ions are located at the vertices of simple
cubes which are concentric with the unit cells, but the simple cubes have only
half the lattice spacing of the unit cell. Alternatively, the eight F − ions can be
considered to lie on two interpenetrating f.c.c. lattices with origins ( 34 , 14 , 14 ) and
( 34 , 34 , 34 ). Since the two F sites are symmetrically displaced from the Ca site by
one quarter of the f.c.c. body diagonal, there is an inversion symmetry about
each of the Ca sites. Each F anion occupies a site at the center of a tetrahedron
formed by the Ca cations.

Materials such as LiO2 , form an anti-fluorite structure. The anti-fluorite


structure is the same as the fluorite structure except that the positions of the
anions and cations are reversed. The O anions are in the f.c.c. positions and
the Li cations form a simple cubic array.

92
The Fluorite Structure

cF12

Figure 58: The Fluorite structure can be considered as an f.c.c. Bravais lattice
with an interpenetrating simple cubic lattice.

3.8.12 The Copper Three Gold Structure


The Cu3 Au structure is cubic, and has the space group P m3m. The Bravais
lattice corresponds to a primitive cubic structure. There are three Cu atoms
and one Au per unit cell. All the atoms are located on the sites of a face
centered cubic unit cell. The Au atom can be envisaged as being positioned on
the corners of the cube, whereas the three Cu atoms sit on the faces centers of
the cube, thereby forming an octahedron inside the cube. Thus, the basis of the
structure consists of the position of the Au atom

r0 = 0 (96)

and the three Cu atoms are located at


a
r1 = ( êy + êz )
2
a
r2 = ( êx + êz )
2
a
r3 = ( êx + êy ) (97)
2
The Au atoms have twelve Cu nearest neighbors located at a distance √a ,
2
whereas the Cu atoms only have four Au nearest neighbors.

93
Cu3Au Structure

cP4

Figure 59: The Cu3 Au structure is equivalent to a simple cubic structure with
a four atom basis.

Other compounds with the Cu3 Au structure are N i3 Al, T iP t3 and the
metastable compound Al3 Li.

3.8.13 Rutile Structure


The structure possessed by rutile (T iO2 ), by cassiterite (SnO2 ) and by nu-
merous other substances with small cations, is tetragonal. The space group is
P 42 /mnm. The T i4+ ions occupy positions : (0, 0, 0) ; ( 12 , 12 , 12 ) while the O2−
ions occupy the four positions ± (x, x, 0) ; ( 12 ± x, 12 ∓ x, 12 ) where x ≈ 10
3
. Thus,
the titanium atoms occupy the sites of a body centered tetragonal lattice. The
oxygen atoms are located on lines which are oriented along one set of face diago-
nals of the base. Oxygen atoms are also located on horizontal lines through the
body centers, and are orthogonal to the lines in the base. Thus, the titanium
ion is surrounded by six O atoms which form a slightly distorted octahedron.

——————————————————————————————————

94
Rutile Structure

a tP6

Figure 60: A unit cell of the Rutile structure.

3.8.14 Exercise 12
Consider the Rutile structure for T iO2 . What conditions must hold for the O
atoms to form a hexagonal close-packed structure?

——————————————————————————————————

3.8.15 Zinc Blende Structure


The Zinc Blende structure or ZnS structure is cubic. This is also known as the
Sphalerite structure. The space group is F 43m. The Zn2+ ions are positioned
at (0, 0, 0) and at the face centers of the cube. The S 2− are positioned on
the sites of an interpenetrating face centered cubic lattice with origin ( 14 , 14 , 14 ).
There are four units of ZnS in the unit cell. The Zinc Blende structure is similar
to the diamond structure. The main difference between the Zinc Blende and
diamond structures is that the Zinc Blende structure involves two different types
of atoms, while the diamond structure only involves one type of atom. Each
atom in ZnS is surrounded by a regular tetrahedron of atoms of the opposite
type. Unlike diamond, Zinc Blende has no center of inversion, as the diamond
inversion operator interchanges the two different types of atoms. The radius

95
Zinc Blende Structure

cF

Figure 61: The Zinc Blende structure can be considered as an f.c.c. Bravais
lattice with a two-atom basis.

ratio rules suggest that the Zinc Blende structure will be adopted whenever


r 
3
2 + 1 > rr > 1 + 2 (98)
2
The Zinc Blende structure is often found for binary compounds formed from
pairs of elements from either the II - VI columns, III - V columns or the I - VII
columns of the periodic table.

3.8.16 Zincite Structure


Zincite (ZnO) has a hexagonal structure. This structure is also known as the
Wurtzite structure (the hexagonal form of ZnS). The space group is P 63 mc.
The primitive lattice vectors are given by

a1 = a êx
a √
a2 = ( êx + 3 êy )
2
a3 = c êz (99)

The Zn and O atoms occupy the positions [ 0, 0, z ]; [ 23 , 23 , 12 + z ] where


z = 0 for Zn and about z ≈ 83 for O. Since ZnS also is found in this form

96
above 1300 K, it is not surprising that Zincite structure has a local coordina-
tion similar to that of the low-temperature Zinc Blende structure. Each atom
is surrounded by a tetrahedron of atoms of the opposite type. The tetrahedra
form continuous interconnected networks. However, symmetry does not require
that the tetrahedra are regular.

Wurtzite Structure

Figure 62: The Zincite structure can be considered as two interpenetrating


hexagonal close-packed lattices.

The cubic Zinc Blende and the hexagonal Wurtzite structures are closely
related. They merely differ by the stacking sequence of the Zn (S) close-packed
planes. The structure consists of alternate close-packed planes which either
contain only Zn or only S ions. The set of planes form layers consisting of a
pair of planes. Within a layer, the Zn atoms in one plane and the S atoms in
the other plane are bonded by vertical tetrahedral bonds. The remaining three
tetrahedral bonds join the layer with atoms in the successive layers. Due to the
orientation of the inter-layer tetrahedral bonds, successive pairs of planes are
displaced horizontally. Thus, the successive sets of vertical bonds are displaced
horizontally.

In the cubic Zinc Blende sequence, the tetrahedra of the S atom bonds have
the same rotational orientation in each layer so that each S layer is displaced in
the same direction. The net horizontal displacement produced in three vertical
S layers is equal to the periodicity in the direction of the displacement. This

97
can be considered as having a stacking sequence A - B - C which repeats.

In the hexagonal Wurtzite sequence, the tetrahedra of bonds are rotated by


π between successive S layers. Thus, the horizontal displacement that occurs
between one S layer and the next are cancelled by the opposite displacement
that occurs by going to the very next S layer. This stacking sequence is A - B
which repeats.

——————————————————————————————————

3.8.17 Exercise 13
Consider the Wurtzite structure in which one type of atom can be regarded as
occupying some of the tetrahedral interstitial sites. Show that in the ideal case,
the lattice parameters are given by

c 2 6
=
a 3
3
z =
8
Also show that, if the nearest neighbor distances are the same in the Zinc√Blende
and Wurtzite lattices, then the lattice constants are related via az = 2aw .

——————————————————————————————————

3.8.18 The Perovskite Structure


The perovskite structure, as exemplified by BaT iO3 , is cubic at high tempera-
tures but becomes slightly tetragonal on cooling below a ferro-electric transition
temperature. The cubic structure has the space group P m3m. The structure is
composed of the T i atoms positioned on the simple cubic lattice sites (0, 0, 0),
and the Ba atoms positioned at the body center sites ( 21 , 12 , 12 ). The three O
atoms are located at the mid-points of the edges of the cube, i.e. at (0, 0, 21 ),
(0, 12 , 0) and ( 12 , 0, 0). An alternate representation of the unit cell is found by
centering the lattice on the Ba ions and by translating the origin via 12 (1, 1, 1).
In this representation, the T i atoms are located at the body centers, and the O
atoms lie on the face centers. The T iO2 form a set of parallel planes separated
by planes of BaO. Each T i atom is surrounded by an octahedron of O atoms
which have corners that are shared with the octahedron surrounding the neigh-
boring T i atoms.

——————————————————————————————————

98
Perovskite Structure

(1/2,1/2,1/2)

(0,0,1/2)

(0,1/2,0)
(1/2,0,0)

Figure 63: The cubic perovskite structure, as found in BaT iO3 .

3.8.19 Exercise 14
The density of the face centered cubic structure is highest, body centered cubic
is the next largest, followed by simple cubic and then diamond has the lowest
density. This correlates with the coordination numbers. The coordination num-
ber is defined to be the number of nearest neighbors. The coordination numbers
are 12 for the f.c.c. lattice, 8 for b.c.c., 6 for s.c. and 4 for diamond. Assume
that the atoms are hard spheres that just touch. Find the packing fraction or
density of these materials.

——————————————————————————————————

99
3.9 Lattice Planes
A Bravais lattice plane, by definition, passes through three non-collinear Bravais
lattice points. Since these points are connected by combinations of multiples
of the primitive lattice vectors, and due to the periodic translational symmetry
of the lattice, the lattice planes must contain an infinite number of lattice points.

Given one such lattice plane, there exists a family consisting of an infinite
set of parallel lattice planes with the same normal. One such lattice plane must
pass through each Bravais lattice point, since the lattice viewed from any lattice
point is identical to the lattice when viewed from any other lattice point. Thus,
the family of parallel planes contain all the points of the Bravais lattice.

Each member of the set of lattice planes must intersect the axes given by
the primitive lattice vectors a1 , a2 and a3 . The planes need not intersect any
particular axis at a lattice point, however, every lattice point located on any
axis will have one member of the family pass through it. In particular, one plane
must pass through the origin O.

A Family of Lattice Planes

3 a3

(3,2,1)
2 a3

a3
3 a2
2 a2
a2
1/2 a2

O 1/3 a1 a1 2 a1 3 a1

Figure 64: Members of a family of lattice planes. Each plane is uniquely specified
by its intercepts with the axes.

Each plane is uniquely specified by the three intercepts of the plane with the
axes formed by three primitive lattice vectors directed from the origin to the
Bravais lattice points a1 , a2 and a3 . The positions of the intercepts are denoted

100
by the numbers x1 , x2 and x3 , where the x’s are measured in units of the length
of the primitive lattice vectors. That is, the intercepts are x1 a1 , x2 a2 and x3 a3 .

The three points of intersection of one lattice plane with the three primi-
tive axes can be represented as κ [ h11 , 0, 0], κ [0, h12 , 0] and κ [0, 0, h13 ], where κ
is a positive or negative integer, and (h1 , h2 , h3 ) are also positive or negative
integers. The integers (h1 , h2 , h3 ) are chosen such that they have no common
factors. The index κ serves to distinguish between the different members of the
same family of planes. The plane that passes through the origin has κ = 0,
whereas the plane that passes next closest to the origin has κ = 1. The planes
that are at successively further distances from the origin have larger magnitudes
of κ.

The indices (h1 , h2 , h3 ) are found by locating the intercepts of the plane with
the three primitive axes, say x1 a1 , x2 a2 and x3 a3 , inverting the intercepts
1 1 1
x1 , x2 , x3 , and then finding the smallest three integers which have the same
ratio
1 1 1
: : = h1 : h2 : h3 (100)
x1 x2 x3
The set of integers (h1 , h2 , h3 ) are enclosed in round brackets and denote the
Miller indices of the plane. A negative valued integer, such as − h1 , is denoted
by an overbar such as h1 .

The Miller indices label the direction of the normal to the family of planes.
Since the vectors between pairs of intercepts are parallel to the plane, the three
vectors
1 1
a − a
h1 1 h2 2
1 1
a − a
h2 2 h3 3
1 1
a3 − a (101)
h3 h1 1
are parallel to the plane. Any two of these vectors span the plane, so the third
vector is not independent. The normal to the plane is parallel to the vector
product of any two non-collinear vectors in the plane
   
1 1 1 1
n̂ ∝ κ2 a1 − a2 ∧ a2 − a3
h1 h2 h2 h3
2
 
κ
= h3 a1 ∧ a2 + h2 a3 ∧ a1 + h1 a2 ∧ a3
h1 h2 h3
(102)

Thus, the direction of the normal to the plane is given in terms of the compo-
nents hi in the three directions defined by aj ∧ ak . The three vectors have the

101
same directions as the primitive “reciprocal lattice vectors”.

The primitive reciprocal lattice vectors are defined by


a2 ∧ a3
b1 = 2 π (103)
a1 . ( a2 ∧ a3 )

and cyclic permutations of the set (1, 2, 3). These primitive reciprocal lattice
vectors are, in general, not orthogonal. The normal to the plane is then given
by the direction of the reciprocal lattice vector B h

B h = h 1 b 1 + h 2 b2 + h 3 b 3 (104)

where (h1 , h2 , h3 ) are the Miller indices. The length of this reciprocal lattice
vector is defined as
 2
| B h |2 = h 1 b1 + h 2 b 2 + h 3 b3
 2

= (105)
dh

The quantity dh is the minimum distance which separates the two closest mem-
bers of the family of planes. This is seen through the following consideration:

Interplanar spacings

d130

d110 b
a

Figure 65: the spacing between successive members of a family of planes is


related to the length of the reciprocal lattice vector.

102
The equation for the points r on the plane which intercept the primitive lattice
vectors ai at distances xi = hκi is given by
κ
r . Bh = a . Bh
h1 1
 
κ
= a . h1 b 1
h1 1
= 2πκ (106)
The minimum distance s, between the origin and the plane is given by
Bh
s = r.
| Bh |
 
dh
= r . Bh

= κ dh (107)
Thus, it is found that the spacing between successive planes in the family is
given by s = dh , and the planes are equidistant.

Sets of families of planes that are equivalent in a given crystal structure


are denoted by {h, k, l}. For example, in a cubic crystal the families of planes
(1, 0, 0), (0, 1, 0) and (0, 0, 1) are equivalent and are denoted by {1, 0, 0}.

A direction of a vector in the direct lattice is specified by three integers in


square brackets [n1 , n2 , n3 ] and specify a vector
n1 a1 + n2 a2 + n3 a3 (108)
A negative value for a component is also denoted by an overbar. The set of direc-
tions which are equivalent for a crystal structure are denoted by < n1 , n2 , n3 >.

——————————————————————————————————

3.9.1 Exercise 15
Consider the (1, 0, 0) and (0, 0, 1) planes of an f.c.c. lattice with axes described
by the conventional unit cell. What are the indices of the planes when referred
to the primitive axes?

——————————————————————————————————

3.9.2 Exercise 16
The angles α1 ( 6 a2 , a3 ), α2 ( 6 a3 , a1 ) and α3 ( 6 a1 , a2 ) between the
three primitive lattice vectors of the direct lattice, ai , are related to the angles

103
A primitive unit cell

a3

α2
α1
a2
α3 a1

Figure 66: The primitive unit cell and the angles between the primitive lattice
vectors.

βi between the three primitive lattice vectors of the reciprocal lattice, bi . The
angles βi are defined as β1 ( 6 b2 , b3 ), β2 ( 6 b3 , b1 ) and β3 ( 6 b1 , b2 ).
Show that
cos β2 cos β3 − cos β1
cos α1 = (109)
| sin β2 sin β3 |
and also find the inverse relation.

——————————————————————————————————

3.9.3 Exercise 17
Complete the table which shows the values of the inter-planar spacings (dh1 ,h2 ,h3 )
in terms of the Miller indices for the seven primitive Bravais lattices.

——————————————————————————————————

104
System dh1 ,h2 ,h3

 − 12
1 2
Cubic a2 (h1 + h22 + h23 )

  − 12
h21 +h22 h23
Tetragonal a21
+ a23

 − 12
h21 h22 h23
Orthorhombic a21
+ a22
+ a23

 h21 + h22 − 2h1 h2 cos α3  − 12


a2 a2 a1 a2 h23
Monoclinic 1 2
sin2 α3
+ a23

Triclinic ?
 − 21
2 2 2 2 2
1 (h1 +h2 +h3 ) sin α+2(h1 h2 +h2 h3 +h3 h1 )(cos α−cos α)
Trigonal a2 1−3 cos2 α+2 cos3 α

   − 12
4 h21 +h22 +h1 h2 h23
Hexagonal 3 a21
+ a23

105
3.10 Quasi-Crystals
Quasi-crystals have symmetries intermediate between a crystal and a liquid.
Quasi-crystals are usually intermetallic alloys. The quasi-crystal is space filling,
but unlike a regular Bravais lattice, does not have just one unit cell. (They
usually have two types of unit cell.) These different “unit cells” are stacked in a
way such that there is no long-ranged positional order but nevertheless, retain
orientational order. The absence of long-ranged positional order lifts the restric-
tion on the symmetry of the lattice but puts a restriction on the vectors that
describe the “unit cells”. For example, an Al − M n quasi-crystal13 has icosahe-
dral symmetry with two, three and five-fold axes. The structure is made from
blocks consisting of a central M n atom surrounded by twelve Al atoms arranged
at the corners of an icosahedron. This type of icosahedral structure is often the
arrangement of thirteen atoms which has the lowest energy14 . The icosahedra
are stacked together with the same orientation. The voids between the icosahe-
dra are the second structural unit. The five-fold symmetry of the icosahedra is
not allowed for a regular Bravais lattice. The five-fold point group symmetry
imposes a restriction on the lengths of the “lattice vectors” of a quasi-crystal
to have certain irrational ratios. Thus, the reciprocal lattice contains reciprocal
lattice vectors of arbitrary small magnitude which show up as an extremely high
density of reflections in x-ray scattering15 .

A way of obtaining quasi-crystal structures is by projecting a periodic Bra-


vais lattice structure in higher dimensions (six or more) onto three dimensions16 .
To illustrate this, consider a square two-dimensional lattice, with lattice con-
stant a. On any unit cell, construct two parallel lines with slope tan θ passing
through opposite corners. The equations of the lower line is given by

y = x tan θ (110)

and the upper line is determined by

y = a + ( x + a ) tan θ (111)

For rational values of the slope, tan θ = pq , the lattice points cross the line
periodically, with repeat distance q a along the x direction and have periodicity
p a along the y direction. Lines with irrational values of the slope cannot cross
more than one lattice point and, therefore, do not have periodic long-ranged
order. The points (na, ma) contained in the area between the two lines must
satisfy the inequality

1 + ( n + 1 ) tan θ > m > n tan θ (112)


13 D. Shechtman, I. Blech, D. Gratais and J. W. Cahn, Phys. Rev. Lett. 53, 1951 (1984).
14 F. C. Frank, Proc. Roy. Soc. London, 215, 43 (1952).
15 D. Levine and P. Steinhart, Phys. Rev. Lett. 53, 2477 (1984).
16 P. Kramer and R. Neri, Acta. Crystallogr. Sec. A 40, 580 (1984).

106
Construction of a one-dimensional Quasicrystal

θ
y-axis

x-axis

Figure 67: The one-dimensional quasi-crystal is constructed by projecting the


lattice points on a strip of a two-dimensional lattice.

Project the lattice points contained within the strip onto one of the lines. The
distance s along the lower line is given by

s = n a cos θ + m a sin θ (113)

where X
m = m0 Θ(1 + (n + 1) tan θ − m0 ) Θ(m0 − n tan θ) (114)
m0

For irrational values of the slope, the resulting array of points is a quasi-periodic
array. The spacing between adjacent points of the quasi-periodic array is either
given by cos θ or sin θ. The spacings are not distributed periodically but nev-
ertheless, are distributed according to some √ irregular or more complex pattern.
If the slopes of the lines are equal to 12 ( 5 − 1 ), the spacings between the
projected points forms a Fibonacci series. For a Fibonacci series of numbers,
the first term can be chosen in any way but the next term is given by the sum

107
of the preceding two numbers, i.e., Fn+1 = Fn + Fn−1 . Thus, both series 1
, 1 , 2 , 3 , 5 , 8√
, 13 etc. or 3 , 3 , 6 , 9 , 15 etc. are Fibonacci series. The
golden mean 12 ( 5 + 1 ) is the limit of the ratio of the successive terms. In
our example, the sequences of spacings is given by s c s c c s c s c . . .. The first
element of the Fibonacci series is s, the second element is c, the third element
comprises of s c, the next element is c s c, which is followed by s c c s c etc.
If this type of analysis is applied to high-dimensional Bravais lattices, one can
find three-dimensional quasi-crystal structures with five-fold symmetry.

A five-fold symmetry is also found when tiling a two-dimensional plane with


two types of tiles, both having the same length of edge s, but with angles of
π 2 π
5 or 5 . The √
“diameter” to side ratios of these two types of tiles satisfy
d s 5 − 1
s = d0 = 2 . The sides of the tiles are marked and the tiles are ad-
joined so that the markings match17 . The result is a tiling without long-ranged

Figure 68: The two types of Penrose tiles. The tiles are decorated by arrows
which indicate the matching rules.

periodic order although every finite area segment repeats an infinite number of
times in the plane. These types of tilings are known as Penrose tilings. The
Penrose tiling has long-ranged orientational order, as can be seen by decorating
each tile with lines. The lines on the tiles join up to form five families of parallel
lines (Ammann lines). The five sets of lines make an angle of 25π with respect to
each other. The spacing between the successive members of a family of parallel
lines form Fibonacci series.

Most of the theoretical studies of real quasi-crystals have considered projec-


tions of a d-dimensional lattice onto three-dimensions. The d-dimensional hy-
percubic lattice is spanned by d orthogonal unit vectors and contains d(d−1)/2!
different orientations of its two-dimensional areas and d(d − 1)(d − 2)/3! orien-
tations of its three-dimensional volumes etc18 . The projection of a hypercube
onto a three-dimensional space is an isohedral polyhedron whose faces are par-
allelograms. The faces of the polyhedron have d(d − 1)/2! different orientations,
17 M. Gardner, Scientific American, 236, 110 (1977).
18 The number Nk of k-dimensional sub-units of the d-dimensional hypercube is given by
the coefficient of xk in the expansion of (2x + 1)d . That is
 
d
Nk = 2d−k
k

108
Figure 69: A two-dimensional quasi-crystal or Penrose tiling.

Figure 70: If the Penrose tiles are decorated with lines as shown, the Penrose
tiling reveals sets of parallel lines.

and the volume can be partitioned into d(d − 1)(d − 2)/3! rhombohedra. The 24
The generalization of Euler’s formula to d dimensions is given by
d−1
X
( − 1 ) k Nk = 1 − ( − 1 ) d (115)
k=0

109
Figure 71: Ammann lines on a Penrose tiling.

vertices of the four-dimensional hypercube can be projected onto the vertices


of a rhombic dodecahedron. A rhombic dodecahedron has fourteen vertices and
twelve faces which have identical shapes in√ the form of a parallelogram.
√ The
parallelograms have fixed angles tan α21 = 2 and tan α22 = ( 2)−1 . The rhom-
bic dodecahedron can be partitioned into four equivalent rhombohedra. The 26
vertices of the six-dimensional hypercube can be projected onto the vertices of a
triacontahedron. The triacontahedron has the full symmetry of the icosahedral
group. The five-fold symmetry can be regarded as originating from a five-fold
rotation symmetry about a principal axis of the six-dimensional lattice. Under
this rotation, the projections of the remaining five orthogonal unit vectors are
interchanged. Due to the choice of the projection, the sixty four vertices of
the six-dimensional hypercube project onto thirty two vertices of the triaconta-
hedron. Of the thirty two vertices, twelve vertices form a regular icosahedron
and the twenty other vertices form a regular dodecahedron. Five faces meet at
each vertex of the icosahedron and three faces meet at each vertex of the do-
decahedron. Hence, the axes of the five-fold and the three-fold rotations of the
icosahedral point group pass through these vertices. The edges join a vertex of
the icosahedron to a vertex of the dodecahedron (E = 5 × 12 = 3 × 20). Thus,
these thirty two vertices are joined by sixty edges which form thirty identical
faces (V − E + F = 2). The faces are parallelograms which contain two five-fold
vertices and two three-fold vertices (F = 5×12
2 = 3×20
2 ). It can be shown that
the projection of the hypercube from the six-dimensional space results in thirty

110
Triacontahhedron

Figure 72: A triacontahedron.

equivalent parallelograms with fixed angles α1 and α2 , where cos α1 = √15


and α2 = π − α1 . The triacontahedron can be partitioned into twenty
rhombohedra, and these twenty rhombohedra fall into two sets of ten equiv-
alent rhombohedra. That is, the triacontahedron can be partitioned into ten
equivalent prolate rhombohedra and ten equivalent oblate rhombohedra. These
two types of rhombohedra are known as the Ammann rhombohedra, and only
these two types of rhombohedra can be formed from the two-dimensional par-
allelograms. The three-dimensional Penrose tilings are space filling packings of
the Ammann rhombohedra.

111
4 Structure Determination
Structure can be determined by experiments in which beams of particles are
scattered from the material. Elastic scattering experiments are usually preferred
since the underlying material is not dynamically deformed by these processes.
In order that the results be easily interpretable in terms of the structure, it is
necessary that the wave length associated with the beam of particles has the
same order of magnitude as the spacing between atoms in the structure and
secondly, the beam of particles should only interact weakly with the structure.
The first condition allows for a clear resolution of diffraction peaks caused by
the atomic structure. The second condition ensures that the beam is scattered
primarily in the bulk or interior of the material, and not just the surface. It
also allows for an easy interpretation of the data via second order perturbation
theory.

4.1 X Ray Scattering


X-rays are usually used in the determination of the atomic structure of solids.
The strength of the interaction is measured by the deviation of the dielectric
constant from its vacuum value ( 1 ). At energies of about 10 keV, the wave
length of the x-rays λ is ∼ 10−10 m, and at these high energies, the refractive
index is almost unity.

In x-ray diffraction, the x-rays are elastically scattered from the charge den-
sity of the electrons. The formal theory of x-ray scattering shows that the
intensity of the reflected waves is given by the Fourier Transform of the electron
density - density correlation function. For a solid which possesses long-ranged
order, the resulting expression for the intensity can be simplified down to involve
the square of the Fourier transform of the electron density. In order to eluci-
date the role of the Bravais lattice and the coherent nature of x-ray scattering,
the atoms shall first be considered to be point-like objects. Later, the spatial
distribution of the electrons around the nuclei shall be re-introduced.

4.1.1 The Bragg condition


Bragg considered the specular reflection of a beam of x-rays from successive
planes of atoms separated by distances d. If the angle between the beam of x-rays
and the planes (not the normal to the plane) is θ2 , then the difference in optical
path lengths for a beam specularly reflected at the lower of two consecutive
layers is
θ
2 d sin (116)
2

112
Bragg scattering by planes of atoms

θ/2

d
θ/2

d sin θ/2 d sin θ/2

Figure 73: Bragg reflections from successive planes of atoms. The angle θ is the
scattering angle.

In this expression, θ is the scattering angle of the particles in the beam. The
reflected beams superimpose with a phase difference of
d θ
4π sin (117)
λ 2
and constructive interference occurs whenever
θ
n λ = 2 d sin (118)
2
This is Bragg’s law19 . The value of n is called the order of the Bragg reflection.
Since the successive planes are equi-spaced, the scattering for an entire family
of planes is constructive when the scattering from two neighboring planes in the
family is constructive. Since there are a large number of planes in a family, and
since the solid is almost transparent to x-rays, the scattering amplitude from
each member of the family adds coherently, giving rise to a very high intensity
of the scattered beam whenever Bragg’s condition is satisfied.

In the application of Bragg’s law to x-ray scattering, not only must one con-
sider the different coherent scattering conditions from a single family of planes,
19 W. L. Bragg, Proc. Cambridge Phil. Soc. 17, 43 (1913).

W. L. Bragg, Nature, 90, 402 (1913).

113
but one must also consider scattering from all the different families of planes in
the solid. Different families of planes of atoms in a solid have different orien-
tations of their common normal. Since a plane of every family passes through
each lattice point, the spacing between the members of a family of planes may
depend on the orientation of the normal. That is, the minimum spacing between
lattice planes d can vary from family to family. The different Bragg reflections
are usually indexed by the Miller indices (m1 , m2 , m3 ) of the planes that they
are reflected from.

4.1.2 The Laue conditions


Laue’s condition is more general than that of Bragg. The Laue condition20 is
derived by considering scattering from the basis atoms in each of the primitive
unit cells in the solid. The individual cells scatter the x-rays almost isotropi-
cally, however, the scattering in a specific direction will only be coherent at wave
lengths for which the scattered waves from each unit cell add constructively.

The Laue Condition

e
e'

θ'/2 d
θ/2

Figure 74: The Laue condition is equivalent to requiring that the scattering
from any two unit cells interfere constructively.
20 W. Friedrich, P. Knipping and M. von Laue, Proc. Bavarian Acad. Sci. 303 (1912).

114
The wave vector of the incident beam is expressed as k, where

k = ê (119)
λ
and the reflected wave has wave vector k 0
2π 0
k0 = ê (120)
λ
where ê and ê0 are two unit vectors. Let us consider two identical scattering
centers separated by a vector displacement d. Consider two rays in the incident
beam, each of which is scattered from one of the centers. The difference in
optical path lengths of the two x-rays is composed of two non-equal segments
θ
d sin = d . ê
2
θ0
− d sin = d . ê0 (121)
2
where the angle between the wave-front of the incident beam and d is given by
θ π θ
2 , so the angle between d and ê is 2 − 2 . The optical path difference between
the two waves is given by the difference
θ θ0
d sin + d sin = d . ( ê − ê0 ) (122)
2 2
Thus, constructive interference of the scattered waves from two unit cells occurs
whenever  
d. ê − ê0 = mλ (123)

holds for integer m. This condition can be re-expressed in terms of the wave
vectors of the incident and scattered x-rays as
 
0
d. k − k = 2πm (124)

If this condition is fulfilled for the set of vectors d that consists of all the Bravais
lattice vectors R, one finds the Laue condition for coherent scattering
 
0
R. k − k = 2πm (125)

or alternatively    
0
exp i k − k .R = 1 (126)

If this condition is satisfied for all R in a solid with N unit cells, constructive
interference will occur between all pairs of unit cells, giving rise to coherent
scattering. The cross-section will have N 2 such contributions, and the scattered

115
wave will be extremely intense.

If the scattering vector q is defined as

q = k − k0 (127)

the Laue condition is satisfied for the special set of q values Q, which satisfy
 
exp iQ.R = 1 ∀ R (128)

These special q values can be used to obtain the k values for which coherent
scattering will occur. The expression for the momentum transfer is

k0 = k − Q (129)

which can be squared to yield

The Scattering Geometry

k'
q
θ

q = k - k'

Figure 75: The scattering angle θ is determined by the momentum transfer q


and the momentum of the beam of incident particles.

k 02 = k 2 − 2 k . Q + Q2 (130)

This equation may be combined with the condition for elastic scattering | k | =
| k 0 |, to result in a condition on the incident k values for coherent scattering of
the form
Q2 = 2 k . Q (131)

116
k
k-Q/2
O Q/2 Q

Figure 76: The Laue condition is satisfied when k resides on the Bragg plane.

Thus, k will satisfy the Laue condition for coherent scattering when the compo-
nent of k along Q bisects Q. Thus, the projection of k along Q must be equal to
half the length of Q. The incident wave vector k must lie on the plane bisecting
the origin and Q, which is called the Bragg plane.

The Laue condition is satisfied if Q . R = 2 π m for all lattice vectors R.


In particular, if the Laue condition is satisfied, one can choose R to be any one
of three primitive lattice vectors. The three choices of primitive lattice vectors
yields the three equations,

a1 . Q = 2 π m1
a2 . Q = 2 π m2
a3 . Q = 2 π m3
(132)

Since any lattice vector R can be expressed as integer multiples of the primitive
lattice vectors, these three Laue equations are equivalent to the Laue condition.
The three Laue equations have a geometrical interpretation. Namely, Q lies on
a cone around the direction of a1 with projection 2 π m1 . Similarly, Q also
lies on a certain cone around a2 , and also on a cone around a3 . Thus, Q must
lie on the common intersection of the three cones. This is a severe constraint:
the values of k for which this is satisfied can only be found by systematically
sweeping the magnitude of k or by rotating the direction of k which is equivalent
to systematically re-orienting the crystal.

However, once Qi values have been found which satisfy the Laue conditions,

117
other Q values can be found which are integral multiples of the initial Qi ’s.
General considerations show that there are three basis vectors bi which satisfy
the set of equations
ai . bj = 2 π δi,j (133)
The three basis vectors bi can be used to construct the general values of Q which
satisfy the Laue condition. These special values of Q have the form of reciprocal
lattice vectors
X3
Q = mi bi (134)
i=1

where the mi are integers.

4.1.3 Equivalence of the Bragg and Laue conditions


The Laue condition makes it clear that each Q value defines a normal to a set of
lattice planes indexed by the Miller indices (m1 , m2 , m3 ). Since a plane belong-
ing to each family of planes passes through each lattice point, it can be shown
that the Laue condition is equivalent to the Bragg condition.

Let Q = k − k 0 be a scattering wave vector such that Q . R = 2 π m for


all lattice vectors R. Since it can be shown that

Q2 = − 2 k0 . Q
= +2k.Q (135)

and as k and k 0 have the same magnitude, the incident and scattered wave vec-
tors make the same angle θ2 with the Bragg plane.

Due to the elastic scattering condition, one has |Q| = 2 k sin θ2 and, if
the scattering is coherent, the magnitude of Q can be written as |Q| = 2 πd n ,
where n is the order of the reflection and d is the minimum distance between
planes in the family. Combining the elastic and Laue conditions, one has
θ πn
k sin =
2 d
θ
2 d sin = nλ (136)
2
Thus, the Laue diffraction peak associated with the change in k given by
k − k 0 = Q, just corresponds to a Bragg reflection by an effective family
of planes which have Q as their normal.

The order n of the Bragg reflection just corresponds to the magnitude of


| Q | divided by 2dπ , where d is the separation between the closest members of
the family of planes.

118
The Elastic Scattering Condition

k
θ/2
q θ/2
- k'

k = k'

Figure 77: For elastic scattering, the magnitude of the momentum transfer is
given by q = 2 k sin θ2 .

4.1.4 The Ewald Construction


Since the Laue condition is very restrictive, the vectors k which produce co-
herent scattering are relatively few and far between. The Ewald construction21
provides a convenient way of visualizing how the Laue condition may be fulfilled.

The incident wave vector k is centered on the origin O. A sphere of radius


k 0 ( = k ) is constructed which is centered on the tip of k. This is the Ewald
sphere. The scattered wave vectors have the magnitude k 0 and may be repre-
sented by vectors k 0 directed from points on the sphere’s surface towards the
center of the sphere. The vector − k 0 is also represented by a vector centered
on the tip of k with an end on the Ewald sphere. The scattering wave vectors
q = k − k 0 are directed from the origin towards the points on the surface of
the sphere.

The scattering wave vectors Q which are solutions of

Q.R = 2πm (137)

form a lattice of points (the reciprocal lattice) which include Q = 0. When


the lattice of Q points is imposed on the Ewald sphere, a lattice point has to
be centered on the origin. This lattice point corresponds to the un-scattered
beam for which k = k 0 , hence q = 0. The lattice of Q points is indexed by
three integers (m1 , m2 , m3 ) corresponding to the components along three prim-
itive (reciprocal) lattice vectors. When a second point of the lattice of Q points
resides on the surface of the Ewald sphere, say at the tip of − k 0 , it produces
a Bragg reflected beam. In this case, the Laue condition is satisfied and the
21 P. P. Ewald, Z. Krist. 56, 129 (1921).

119
The Ewald Sphere Construction

k'
k
O
-k'
Q

Figure 78: Coherent scattering occurs when a reciprocal lattice vector Q resides
on the surface of the Ewald sphere.

incident beam will be Bragg reflected at this k 0 value. In general, it is expected


that the Ewald sphere will not have a second lattice point on the surface. When
Bragg reflections occur, they are indexed by the integers (m1 , m2 , m3 ) which
describe the family of planes associated with the momentum transfer Q.

4.1.5 X-ray Techniques


There are various techniques which can be used to obtain diffracted beams.

In the Laue Method, a beam of x-rays with a continuum of wave lengths


in the range between λ0 and λ1 is used, and the incident beam has a fixed
direction. Thus, it is only appropriate to use this method for a single crystal,
as a polycrystalline sample would correspond to an average over the relative
orientation with the incident beam. In the Laue method, the continuous wave-
length of the beam broadens the surface of the Ewald sphere into a finite volume
enclosed between two Ewald spheres with the limiting wave lengths. For a large
enough mismatch between the wave length of the interior Ewald sphere λ0 and
the exterior sphere λ1 , it is quite likely that at least one Bragg reflection will
occur. This method provides the simplest method for orienting a single crystal
relative to the direction of the incident beam. If the incident beam is along a
direction of high symmetry of the lattice of Q points, the pattern of reflected

120
Ewald Construction for the Laue method

|k1|
k1
O
k2
|k2|

Figure 79: In the Laue method, coherent scattering occurs whenever a reciprocal
lattice vector is located within the volume enclosed between the Ewald spheres
of radius k2 and k1

beams should exhibit the same symmetry. It should be noted that the x-ray
pattern will always show a center of symmetry, even if the crystal does not have
one. This discovery is due to Friedel22 .

The Rotating Crystal Method uses a monochromatic beam of x rays, and


in the experiment, the relative direction of the incident beam and the crystal is
varied. If one considers the lattice of points Q as being fixed, then the Ewald
sphere rotates around the origin and, for large enough k, will sweep some lattice
points through the surface of the sphere. This experiment produces a set of
Bragg reflected beams that are recorded on a photographic film. In practice,
the crystal is rotated about a crystallographic axis, say a1 , while the incident
beam has a fixed direction perpendicular to a1 . The photographic film is bent
into a cylinder with an axis which is chosen to coincide with the axis of rotation
of the crystal. Since the incident beam is perpendicular to the rotation axis, then
the Bragg reflected beams occur within cones of fixed angle. That is, the b2 and
b3 components of the lattice of Q points form planes which are perpendicular
to a1 . Therefore, under the rotation, these two components of the Q vectors
22 G. Friedel, Comptes Rendus, Acad. Sci. (Paris) 157, 1533 (1913).

121
Ewald Sphere for the Rotating Crystal Method

Ewald Sphere with radius |k|


Spheres with radii |Q|

Figure 80: In the rotating crystal method, Bragg scattering occurs whenever
spheres with radii Q centered on the origin cut through the Ewald sphere.

a1

k'

- Q . a1

Figure 81: In the rotating crystal method, the scattering occurs within cones
around the crystallographic axis a1 .

122
are rotated in the planes. However, the components of Q parallel to a1 remain
invariant and are governed by m1 , since

Q . a1 = 2 π m1 (138)

Furthermore, since k and a1 are perpendicular

k 0 . a1 = − Q . a1 (139)

and the reflected beams produce a series of Bragg spots which exist in rings
wrapped around the photographic film cylinder. Each ring corresponds to a
different value of m1 . Direct observation of the angle between k 0 and a1 allows
the magnitude of a1 to be obtained with ease.

The Debye-Scherrer Method uses a polycrystalline or powdered sample.


Each grain of the sample has a random orientation, therefore, this method is
equivalent to the rotating crystal method in which the sample is rotated over
all possible orientations. Each reciprocal lattice point will generate a sphere of
radius equal to the magnitude of the reciprocal lattice vector. If this spherical
shell of reciprocal lattice vectors intersects with the Ewald sphere, it produces
Bragg reflections. Each lattice vector with length less than 2 k will produce a
cone of Bragg reflections with an angle θ relative to the un-scattered beam. The
magnitude of the reciprocal lattice vector is given by Q = 2 k sin θ2 . Thus,
a measurement of θ will give the lengths of the smallest reciprocal lattice vectors.

These methods can be used to determine the reciprocal lattice vectors and
hence, the Bravais lattice associated with the crystal. In order to completely
determine the crystal structure, one must determine the basis. This can be done
by examining the structure and form factors.

——————————————————————————————————

4.1.6 Exercise 18
The icosahedral molecule C60 molecule has been proposed as having the form
of a truncated icosahedron with twenty six-membered rings and twelve five-
membered rings23 . These molecules can be condensed into solid phases, in
which the C60 molecules are rotationally disordered. Fleming and co-workers
describe the structure of various alkaline metal C60 compounds in their Nature
article, Nature 352, 701, (1991).

In figure (2.a) of their paper, Fleming et al. state that the solid has an f.c.c.
structure. Indicate the axes of the conventional unit cell on their unit cell.

23 H. W. Kroto, J. R. Heath, S. C. O’Brien, R. F. Curl and R. E. Smalley, Nature, 318, 162

(1985).

123
Powder Diffraction

Sample Film

Figure 82: The diffraction of x-rays by a powdered sample gives rise to a pattern
of concentric circles on the photographic film.

Figure 83: Unit cells of crystals of doped C60 . [After Fleming et al. (1991).]

If a powder x -ray diffraction experiment is performed on Rb doped C60 with


x-rays of wavelength λ = 0.9 Å, for the dopings 3, 4 and 6 in the paper, what
are the scattering angles 2 θ for the first five diffraction peaks for the observed
structures?

——————————————————————————————————

124
4.1.7 The Structure and Form Factors
If the lattice has a basis, the scattered wave from each unit cell must be com-
posed from the scattered waves from each atom in the basis. This means that
the scattering from each type of atom in the basis must be determined and
then superimposed to find the scattered wave. The scattering from the electron
density of each atom can be expressed in terms of the form factor24 . The form

28

24 λ = 0.709 Α
Hartree-Fock
20
FFe(q)

16 (110)
(200) (220)
12
(211) (222)
(411)
8 (310)
(321)
(330)
4

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

θ/2) / λ [ Angstroms ]
sin (θ/2)
-1

Figure 84: The experimentally determined and calculated atomic form factors
FF e (q) for F e. The angle θ is defined as the scattering angle. [After Batterman,
Chipman and DeMarco, Phys. Rev. 122, 69 (1961).]

factor for atoms in a solid differ only slightly from the form factors of isolated
atoms25 , and are mainly determined by the atomic charge number Z. Although
there are differences due to the bonding, the form factors are determined by all
the electrons and not just those involved in bonding. The form factor of the
j-th atom in the basis is denoted by Fj (q). It is conventional to use a scale such
that the forward scattering (θ = 0) atomic form factor equals the number of
electrons in the atom. Since the coherent scattering is restricted to scattering
vectors Q that satisfy the Laue condition, the form factor only needs to be
evaluated at these values of Q. The amplitude of the scattered wave from the
atoms in the basis of the unit cell can be expressed in terms of the structure
factor S(Q) which is given by
X  
S(Q) = exp i Q . rj Fj (Q) (140)
j
24 C. G. Darwin, Phil. Mag. 43, 800 (1922).
25 The covalently bonded semiconductors provide notable exceptions to this statement. For
example, in diamond, the electronic charge density has maxima intermediate between the
atomic positions (N. A. W. Holzwarth, S. G. Louie and S. Rabii, Phys. Rev. B 26, 5382
(1982).). This is a manifestation of the covalent bonding, and has a result that the structure
factor does not vanish at Q = 2π a
(2, 2, 2). Hence, this is seen as an extra peak in x-ray
diffraction spectra. (S. Gottlicher and E. Wolfel, Z. Electrochemie, 63, 891 (1959).)

125
Figure 85: The calculated and experimentally determined valence electron
charge densities of graphite. [After Holzwarth et al. (1982).]

This is just the component of Fourier Transform of the electron density from
one unit cell. The intensities of the Bragg peaks are proportional to the factor

| S(Q) |2 (141)

The Q dependence of the intensity can be used to determine the basis of the
crystal. Unfortunately, only the modulus of S(Q) and not its phase, can be
found from diffraction experiments which produce only one scattered Bragg
beam26 . Therefore, indirect methods have to be used to discover the crystal
structure. However, if the crystal is centro-symmetric, then if there is an atom
at the basis point rj then there is another atom of the same type at − rj and
S(Q) is purely real. In this case, the phase problem just simplifies to the ques-
tion as to whether S(Q) is positive or negative.

26 In the case that more than one Bragg is produced in the scattering experiment, say

corresponding to reciprocal lattice vectors Q and Q , then the relative phase of the structure
1 2
factor S(Q ) and S(Q ) can be determined. (L. D. Chapman, D. R. Yoder, and R. Colella
1 2
Phys. Rev. Lett. 46, 1578 (1981), B. Post, Phys. Rev. Lett. 39, 760 (1977)).

126
If the basis of a crystal structure is mono-atomic, the atomic form factor
can be factorized out, and the amplitude of the scattered wave is partially
determined by the geometric structure factor
X  
SG (Q) = exp i Q . rj (142)
j

The geometric structure factor expresses the interference between identical atoms
in the basis. The intensity of the Bragg peak is still determined by the product
of the modulus of the form factor with the modulus of the geometric structure
factor. The vanishing or variation of the Bragg peak intensities due to interfer-
ence can be used to determine the positions of the basis atoms.

An example of the ambiguity imposed by the non-measurability of the phase


of the Structure Factor is given by Friedel’s law for non-centrosymmetric crys-
tals. The structure factor S(Q) is a complex number, and can be written as

S(Q) = A + i B (143)

For each Q that satisfies the Laue condition, there is a vector −Q which corre-
sponds to the negative integers (−m1 , −m2 , −m3 ). The structure factor S(−Q)
is just the complex conjugate of S(Q)

S(−Q) = A − i B (144)

Since the structure factor for both the vectors Q and −Q have the same magni-
tude, the Bragg peaks have the same intensity. Thus, the diffraction pattern has
a center of inversion symmetry, even if the crystal structure does not. Excep-
tions to Friedel’s law only occur if the crystal has anomalous dispersion. This
happens when the x-rays are highly absorbed by the crystal.

Face Centered Cubic Structures.

A structure with a face centered cubic lattice and a one atom basis can also
be represented in terms of a simple cubic lattice with a four atom basis. The
scattering from this structure can be expressed in terms of the Laue condition for
the simple cubic lattice but modulated by the geometric structure factor. The
four atom basis of the non-primitive (conventional) unit cell of a face centered
cubic structure with a monoatomic basis consists of the atomic positions

r1 = 0
 
a
r2 = êx + êy
2
 
a
r3 = êz + êx
2
 
a
r4 = êz + êy (145)
2

127
The Bragg vectors for the conventional simple cubic cell are easily found to be
 

bx = êx
a
 

by = êy
a
 

bz = êz (146)
a

so a general simple cubic Bragg scattering vector is given by


  

Q = m1 êx + m2 êy + m3 êz (147)
a

The geometric structure factor for the conventional f.c.c. unit cell is found to
have four contributions
X  
SG (Q) = exp i Q . rj
j
"  
= 1 + exp + i π ( m1 + m2 ) +
   #
+ exp + i π ( m1 + m3 ) + exp + i π ( m2 + m3 )

(148)

one contribution is provided by each basis atom. When evaluated at the Bragg
vectors, the geometric structure factor adds coherently

SG (Q) = 4 (149)

if the integers (m1 , m2 , m3 ) are either all even or are all odd. The geometric
structure factor interferes destructively

SG (Q) = 0 (150)

if only one integer is different from the other two. That is, if one integer is ei-
ther even or odd, while the other two respectively, are odd or even, then SG (Q)
vanishes. Thus, the f.c.c. lattice has the same pattern of Bragg reflections as
the simple cubic lattice, but has missing Bragg spots. The resulting lattice of
Bragg spots is cubic with twice the dimensions (in q space) but has missing
Bragg spots at the mid-points of the edges and at the face centers. Thus, it is
found that the diffraction pattern has the form of a body centered cubic lattice.

Body Centered Cubic Structures.

128
A structure with a body centered cubic lattice and a one atom basis can also
be viewed as a simple cubic lattice with a two atom basis
r0 = 0
 
a
r1 = êx + êy + êz (151)
2
Then, the geometric structure factor for the conventional b.c.c. unit cell only
contains two terms
 
a
SG (Q) = 1 + exp i Q . ( êx + êy + êz )
2
 
a
= 1 + exp i ( Qx + Qy + Qz ) (152)
2
Each basis atom provides one contribution to the geometric structure factor.
Now the Bragg vectors for the simple cubic structure are just

Q = ( m1 êx + m2 êy + m3 êz ) (153)
a
therefore, at these Q values the geometric structure factor simplifies to
 
SG (Q) = 1 + exp i π ( m1 + m2 + m3 )
 ( m1 + m2 + m3 )
= 1 + − 1

= 2 for ( m1 + m2 + m3 ) even

= 0 for ( m1 + m2 + m3 ) odd
(154)
Thus, the body centered cubic lattice has Bragg spots that form a cubic lattice.
However, the intensity of the odd indexed Bragg spots vanish, leading to a face
centered cubic lattice of Bragg spots.

The Diamond Structure.

The diamond structure is described by an f.c.c. lattice with a two atom


basis. The positions of the basis atoms are given by
r0 = 0
 
a
r1 = êx + êy + êz (155)
4
where the conventional f.c.c. unit cell has sides of length a.

129
From the discussion of scattering from an f.c.c. structure, one finds that the
Q vectors of the Bragg spots can be expressed in terms of the set of primitive
vectors for a b.c.c. lattice X
Q = mi bi (156)
i

The primitive lattice vectors are given by


 

b1 = êy + êz − êx
a
 

b2 = êz + êx − êy
a
 

b3 = êx + êy − êz (157)
a

The geometric structure factor of the diamond lattice relative to the lattice of
Bragg spots of the real space f.c.c. lattice, is given by
 
π
SG (Q) = 1 + exp i ( m1 + m2 + m3 ) (158)
2

From this it is found that the geometric structure factor not only gives rise
to extinctions but also modulates the intensity of the non-zero Bragg spots
according to the rule

SG (Q) = 2 for ( m1 + m2 + m3 ) 2 × even

SG (Q) = 0 for ( m1 + m2 + m3 ) 2 × odd

SG (Q) = 1 ± i for ( m1 + m2 + m3 ) odd


(159)

As the f.c.c. lattice has Bragg spots arranged on a b.c.c. lattice, it is con-
venient to transform the Bragg vectors into the coordinates system used for a
conventional b.c.c. unit cell
"  
4π 1
Q = êx ( m1 + m2 + m3 ) − m1
a 2
 
1
+ êy ( m1 + m2 + m3 ) − m2
2
 #
1
+ êz ( m1 + m2 + m3 ) − m3
2
(160)

130
The rule for the modulation of intensities is expressed directly in terms of the
quantity
X Qi a 1
= ( m1 + m2 + m3 ) (161)
i
4 π 2
Thus, one can describe the system of Bragg spots as residing on a b.c.c. lattice
where the cubic cell has sides with dimensions of 4aπ . The b.c.c. lattice can be
re-interpreted in terms of two interpenetrating simple cubic lattices. Thus, the
Bragg spots with non-equal intensities reside on two interpenetrating simple cu-
bic lattices with dimensions of 4aπ . This scale is twice as large as the reciprocal
lattice spacing of the (simple cubic) lattice constructed from the conventional
unit cell.

One simple cubic lattice contains the origin Q = 0, and the Bragg spots
have integer coefficients for the unit vectors êx , êy and êz . This means that
( m1 + m2 + m3 ) is even for this simple cubic lattice. On dividing by a
factor of 2, the resulting number is odd and even at consecutive lattice points.
When ( m1 + m2 + m3 )/2 is an even integer, S = 2 and the intensities
are finite. However, when ( m1 + m2 + m3 )/2 is odd then S = 0 so the
intensities of the Bragg peaks vanish. Thus, the non-zero intensities on this sim-
ple cubic reciprocal lattice actually forms a face centered cubic reciprocal lattice.

The second interpenetrating simple cubic lattice has Bragg points with half
(odd) integer coefficients for the unit vectors êx , êy and êz . This means that
the sum ( m1 + m2 + m3 ) is odd for this simple cubic lattice. These lattice
points are the body center points of the underlying b.c.c. lattice. The geometric
structure factor is simply SG (Q) = 1 ± i and thus, the Bragg spots on this
simple cubic lattice all have the same intensities.

Extinctions due to Glide Planes and Screw Axes.

The symmetry and metric properties of an X-ray diffraction pattern can be


used to determine the point group symmetry of the Bravais lattice of a sample.
The observation of systematic absences of Bragg spots due to any centering of
unit cells and other symmetry operations with translational components, may
be extremely useful in the identification of the space group of the sample.

Consider an orthorhombic solid with a glide symmetry where the translation


is along the êz axis and the reflection occurs in a plane with a normal along the
êy axis. Thus, if there is an atom at (x, y, z) in units of the lattice parameters,
there is an equivalent atom at (x, y, z + 12 ). The pairs of basis atoms each
contribute a term
   
1
SG (Q) = exp 2 π i ( x m1 + y m2 + z m3 ) + exp 2 π i ( x m1 − y m2 + ( z + ) m3 )
2
(162)

131
Glide Operation
2
ez

(x,-y,z+1/2) x ey
0
-2 -1 0 1 2

(x,y,z)
-1

ex

-2

Figure 86: A crystal is symmetric under the glide operation consists of a trans-
lation by 2c in the êz direction, followed by a reflection in the x − z plane.
Therefore, if an atom is located at (x, y, z), then an identical atom is located at
(x, −y, z + 12 ).

to the geometric structure factor. One can see that for the special case m2 = 0
the structure factor is composed of terms with the form
   
SG (Q) = exp 2 π i ( x m1 + z m3 ) 1 + exp π i m3
  
= exp 2 π i ( x m1 + z m3 ) 1 + ( − 1 )m3

= 0 if m3 is odd
 
= 2 exp 2 π i ( x m1 + z m3 ) if m3 is even

(163)

Thus, reflections of the type (m1 , 0, m3 ) will be missing unless m3 is an even


number.

Similar extinctions occur for screw axes. Consider a two-fold screw axis
parallel to êy . The equivalent positions are (x, y, z) and (x, 12 + y, z). Thus, the

132
Figure 87: The structure of a C60 molecule.

structure factor at m1 = 0 and m3 = 0 has contributions of the form


  
SG (Q) = exp 2 π i y m2 1 + ( − 1 )m2

= 0 if m2 is odd
 
= 2 exp 2 π i y m2 if m2 is even

(164)

Thus, reflections of the type (0, m2 , 0) will be missing unless m2 is an even in-
teger.

——————————————————————————————————

4.1.8 Exercise 19
Experiments on solid Ax C60 show that the C60 molecules are located on a face
centered cubic lattice with lattice spacing a = 14.11 A, and that the (2, 0, 0)
x-ray diffraction peak is very weak when compared to the (1, 1, 1) Bragg peak.

133
Fleming et al. Nature 352, 701 (1991). Calculate the structure factor for these

Figure 88: Data from x-ray diffraction experiments on crystals of doped C60 .
[After Fleming et al. (1991).]

reflections in an approximation which assumes that the electron distribution of


each fullerene molecule is uniformly spread over a spherical shell of radius 3.5 A.

——————————————————————————————————

4.1.9 Exercise 20
The Hendricks-Teller model27 for x-ray diffraction from a disordered system
considers a one-dimensional line of molecules. The probability that a pair of
atoms is separated by a distance a is given by p and the probability that they
are separated by a + da is given by 1 − p. The random system has an infinite
unit cell. Calculate the averaged scattering intensity for this model, and show
27 S. B. Hendricks and E. Teller, J. Chem. Phys. 10, 147 (1942).

134
that
 
p(1 − p) 1 − cos q da
I(q) = N I0
1 − p(1 − p) − p cos qa − (1 − p) cos[ q(a + da) ] + p(1 − p) cos qda
(165)
Due to the random phases introduced by the distribution of distances, the scat-
tering is not fully coherent. However, for small amounts of disorder, the coher-
ence can be almost fully recovered.

——————————————————————————————————

Polyatomic Crystals.

For a polyatomic crystal, the structure factor has both the geometric con-
tribution and the contribution from the atomic form factors of the basis atoms
X  
S(Q) = exp i Q . rj Fj (Q) (166)
j

The atomic form factor Fj (Q) is determined by the internal structure of the
atom that occupies the position rj in the basis.

The atomic form factor is normalized to the electronic charge of the atom.
For a single atom, the form factor is given by
Z  
F (Q) = d3 r ρ(r) exp − i Q . r (167)

where ρ(r) is the atomic electron density. If the charge density is spherically
symmetric, then the form factor can be reduced to a radial integral
Z ∞ Z 1  
F (Q) = 2 π dr r2 ρ(r) d cos θ exp − i Q r cos θ
0 −1
Z ∞
2 sin Q r
= 2π dr r2 ρ(r)
Qr
Z0 ∞
sin Qr
= 4π dr r2 ρ(r) (168)
0 Qr
For forward scattering, Q = 0, the form factor reduces to
Z ∞
F (0) = 4 π dr r2 ρ(r)
0
= Z (169)
where Z is the atomic number. Typically, F (Q) decreases monotonically with
increasing Q, falling off as a power of Q12 for large Q.

——————————————————————————————————

135
4.1.10 Exercise 21
Calculate the x-ray scattering intensities for the close-packed structures formed
by stacking hexagonal layers in the following sequences:

(a) The sequence ABAB... (the h.c.p. sequence).

(b) The sequence ABCABC... (the f.c.c. sequence).

(c) The random sequence in which all the consecutive layers are different,
but given one layer (say A), there is an equal probability that it will be followed
by either one of the two other layers.

——————————————————————————————————

4.1.11 Exercise 22
Find the atomic form factor for the hydrogen atom using the electron density
 
1 2r
ρ(r) = exp − (170)
π a3 a

where a is the Bohr radius.

——————————————————————————————————

Sodium Chloride.

An example of a diatomic crystal with a basis is provided by N aCl. This has


a face centered cubic lattice and has N a+ ions at the positions (0, 0, 0), ( 12 , 12 , 0),
( 12 , 0, 12 ) and (0, 12 , 12 ). The Cl− ions reside at ( 12 , 0, 0), (0, 12 , 0), (0, 0, 12 ) and
( 12 , 12 , 12 ). The structure can be viewed as a simple cubic lattice with a six atom
basis. In this case, we can use the simple cubic representation of the Bragg
vectors Q. Thus, the structure factor is given by
  
S(Q) = FN a (Q) 1 + exp i π ( m1 + m2 )
   
+ exp i π ( m2 + m3 ) + exp i π ( m3 + m1 )
    
+ FCl (Q) exp i π m1 + exp i π m2
   
+ exp i π m3 + exp i π ( m1 + m2 + m3 )

(171)

136
Since exp[ i π m ] = ( − 1 )m , the structure factor can be factorized as
  m1 
S(Q) = FN a (Q) + − 1 FCl (Q)
  (m1 +m2 )  (m2 +m3 )  (m3 +m1 ) 
× 1 + − 1 + − 1 + − 1

(172)

The structure factor is 0 unless the indices are either all odd or all even. This is
characteristic of face centering. The intensities of the Bragg spots with all even
indices and all odd indices are different as the atomic form factors either add or
subtract.

——————————————————————————————————

4.1.12 Exercise 23
Potassium Chloride has the same structure as N aCl. However, unlike N aCl,
K + and Cl− are iso-electronic and so have very similar form factors. Determine
the indices (m1 , m2 , m3 ) of the allowed Bragg reflections.

——————————————————————————————————

4.1.13 Exercise 24
Calculate the structure factor for the Zinc Blende structure. The Zinc Blende
structure is a face centered cubic lattice of side a, with a positively charged ion
at the origin and a negatively charged ion at a4 ( êx + êy + êz ).

——————————————————————————————————

Since the differences between the atomic form factors show up in the exper-
imentally observed structure factor of compounds, it is possible to distinguish
between ordered binary compounds and binary compounds with site disorder.
The order-disorder transition in Cu3 Au has been observed by x-ray scattering.
At high temperatures, the atoms in this material are randomly distributed such
that there is one atom at each site of an f.c.c. lattice. In this disordered state,
the apriori probability that a Au atom will be found on any particular site is
roughly 41 , and the probability that a Cu atom will be found at any particu-
lar site is 34 . However, there is a transition from the disordered phase, which
occurs above a critical temperature of Tc ≈ 660 K, to an ordered phase at
lower temperatures. In the completely disordered phase, the structure factor is
that pertaining to an f.c.c. crystal in which the form factor is replaced by the

137
statistically averaged value
3 1
Fav (Q) = FCu (Q) + FAu (Q) (173)
4 4
Thus, at high temperatures, the structure factor is given by
  
S(Q) = Fav (Q) 1 + exp i π ( m1 + m2 )
   
+ exp i π ( m2 + m3 ) + exp i π ( m3 + m1 )

(174)
Hence, the peaks have intensity of either 16 | Fav (Q) |2 or zero depending on
whether the indices are all even or all odd, or whether they are mixed. In the
ordered phase, the Cu atoms reside on the face center sites and the Au on the
vertices of the cubes. In this ordered phase, “super-lattice” peaks appear in
the spectra for mixed indices. For the completely ordered phase, the structure
factor is given by
  
S(Q) = FAu (Q) + FCu (Q) exp i π ( m1 + m2 )
   
+ exp i π ( m2 + m3 ) + exp i π ( m3 + m1 )

(175)
The “super-lattice” peaks occur for mixed indices. The relative intensity of the
“super-lattice” peaks are approximately given by
 2
I(1, 0, 0) FAu (0) − FCu (0)
∼ (176)
I(2, 0, 0) FAu (0) + 3 FCu (0)
which leads to a relative intensity of about 0.09.

At very low temperatures, CuZn exists as an ordered compound of the CuCl


type. The structure consists of two interpenetrating simple cubic sub-lattices
which have a relative displacement of [ 12 , 12 , 12 ]. The Cu atoms occupy the sites
of one sub-lattice, say the A sub-lattice, and the Zn atoms are located on the
other sub-lattice, say the B sub-lattice. For an infinite solid the A and B sub-
lattices are equivalent thus, the compound may also form with the Cu atoms
on the B sub-lattice and the Zn atoms on the A sub-lattice. Since the x-ray
form factors are FCu (0) = 29 and FZn (0) = 30, the relative intensity of the
“super-lattice” peaks of CuZn, or beta brass, are of the order of 0.0003. Thus,
the super-lattice peaks are difficult to observe in x-ray scattering. However, the
order-disorder transition in CuZn and related compounds is easily observable
by neutron diffraction28 .

28 C. G. Shull and S. Siegel, Phys. Rev. 75, 1008 (1949).

138
Ordered Phase

T < Tc

Figure 89: The ordered phase of CuZn.

Disordered Phase

T > Tc

Figure 90: The disordered phase of CuZn.

At temperatures above the order-disorder transition temperature of CuZn,


Tc ≈ 741 K, the material exists in a disordered phase in which the Cu and Zn
atoms are randomly positioned on the sites of the A and B sub-lattices. At the
transition temperature, a phase transition occurs between the high temperature
disordered phase and the low-temperature ordered phase. The order parameter
for the phase transition is given by the scalar quantity, φ(T ), where

φ(T ) = n(Cu)A − n(Cu)B


= n(Zn)B − n(Zn)A (177)

139
Figure 91: The neutron diffraction pattern of the ordered and disordered phases
of F eCo [After Shull and Siegel (1949).].

where n(Cu)A and n(Cu)B are, respectively, the number of Cu atoms on the A
and B sub-lattices. The second line follows from the fact that an atom of one
type or the other exists at each site. In particular, if the total number of sites
is 2 N , the numbers of Zn atoms at the sites of the A and B sub-lattices are,
respectively, given by
n(Zn)A = N − n(Cu)A
n(Zn)B = N − n(Cu)B (178)
Above the transition temperature, the Cu atoms are equally probable to be
found on the A and B sublattices and so the order parameter is zero, φ = 0.
Below the transition temperature, the order parameter has a non-zero magni-
tude φ0 (T ) which is temperature dependent, and has either a positive or negative
sign depending on whether the Cu atoms spontaneously select to occupy the A
or B sites, φ(T ) = ± φ0 (T ). In the ordered state, the temperature dependence
of the order parameter is given by
φ0 (T ) ∝ ( Tc − T )β (179)
where β ≈ 0.32. As the Hamiltonian is symmetric under interchange of the A
and B sub-lattices, this order-disorder transition provides an example of spon-
taneous symmetry breaking.

——————————————————————————————————

4.1.14 Exercise 25
Express the inelastic x-ray scattering intensity for CuZn in terms of the atomic
form factors FCu (Q), FZn (Q), and the order parameter φ(T ). Assume that the

140
Temperature-dependence of the order parameter

1.2

0.8
φ(T)

0.6

0.4

0.2
Tc
0
0 50 100 150 200 250

T[K]

Figure 92: The ordered parameter of the order-disorder transition of CuZn.

deviations of the site occupancies from the average values at different sites are
un-correlated.

——————————————————————————————————

4.2 Neutron Diffraction


Elastic neutron scattering from the nuclei of a solid involves a change in the
momentum of the neutron from the initial value h̄ k to the final value h̄ k 0 of

q = k − k0 (180)

Conservation of momentum requires that the transferred momentum must be


equal to a momentum component of the interaction potential. This momentum
is ultimately transferred to the solid. Experimentally accessible ranges of q for
neutrons are in the range of 0.01 < q < 30 A, which covers the range that is
useful to determine crystalline structures.

The interaction between the neutron and one nucleus is short-ranged and
can be modelled by a point contact interaction,

2 π h̄2
Ĥint = b δ3 ( r − R ) (181)
mn

141
Figure 93: The neutron diffraction pattern of ZnF e2 O4 (top) and N iF e4 O2
(bottom). [After Hastings and Corliss (1953).]

where b is the scattering amplitude of the order of 10−14 m. The differential


scattering cross-section represents the number of particles per unit time which
are scattered into solid angle dΩ per incident flux. The differential scattering
cross-section for one nucleus is assumed to be isotropic and is given by

= | b |2 (182)
dΩ
Hence, the total cross-section for the one nucleus is given by
Z

σ = dΩ
dΩ
= 4 π | b |2 (183)

where the scattering takes place over all solid angles.

For a crystalline lattice of nuclei, as it shall be shown, the scattering cross-


section is given by
 
dσ X
= exp i q . ( Ri − Rj ) bi b∗j (184)
dΩ i,j

where bi is the scattering amplitude from the i-th nucleus. The value of bi de-
pends on what isotope exists at the lattice site and also on the direction of the
nuclear spin.

In general, the different isotopes are randomly distributed so they must be


averaged over. Thus, bi and bj are independent or uncorrelated if they belong
to different sites, and the average for i 6= j is given by the product of the
averages
bi b∗j = bi b∗j = | b |2 (185)

142
while, if i = j, one has the average of the squared amplitude

bi b∗i = | b |2 (186)

In general, the average of bi b∗j has the form


 
bi b∗j = | b |2 + δi,j | b |2 − | b |2 (187)

The scattering cross-section can be written as the sum of two parts, a coherent
part where i 6= j and an incoherent part which has i = j.

The coherent cross-section is given by


 
dσ X
= exp i q . ( Ri − Rj ) | b |2 (188)
dΩ i,j

For coherent scattering from every nuclei in the solid, the momentum transfer
must satisfy the Laue condition and so q must be equal to Q, where Q satisfies

Q.R = 2πm (189)

for all lattice vectors R and m is any integer. When this condition is satis-
fied, the scattering produces Bragg reflections similar to those observed in x-ray
scattering. When the Bragg scattering condition is satisfied, the coherently
scattered beam has an intensity which is proportional to N 2 .

The incoherent scattering cross-section comes from the terms with i = j


and is given by  
dσ 2
= N | b | − | b |2 (190)
dΩ
The incoherent scattering is proportional to the number of nuclei N and is in-
dependent of the direction of q. It is obvious that the coherent and incoherent
contributions are profoundly different. Only the coherent contribution can be
utilized to determine the crystalline structure. Examples of structural determi-
nation from neutron diffraction patterns are provided by Zinc Ferrite and Nickel
Ferrite29 .

——————————————————————————————————

4.2.1 Exercise 26
Consider a beam of unpolarized neutrons being diffracted from a material which
has nuclei with spin I. The scattering length bJ depends on the total angular
29 J. M. Hastings and L. M. Corliss, Rev. Mod. Phys. 25, 114 (1953).

143
momentum J of the combined neutron and nucleus. The scattering length is b+
for J = I + 12 and b− for J = I − 12 . Show that
 
I + 1 I
b = b+ + b− (191)
2I + 1 2I + 1

and  
I + 1 I
| b |2 = | b+ |2 + | b− |2 (192)
2I + 1 2I + 1

The scattering from iron, F e, is almost entirely coherent since the most abun-
dant isotope has I = 0. On the other hand, the scattering from vanadium, V ,
is almost completely incoherent. Other materials such as those containing Eu,
have nuclei which are strong neutron absorbers and, therefore, neutron diffrac-
tion is not a suitable technique to determine their structure.

——————————————————————————————————

144
4.3 Theory of the Differential Scattering Cross-section

By definition, the differential scattering cross-section dΩ is the ratio of the
number of particles scattered dNscatt (per unit time) into a solid angle dΩ =
sin θ dθ dϕ to the incident flux of particles F (number of particles crossing unit
area per unit time) times the solid angle element

dNscatt = F dΩ (193)
dΩ
Consider a beam of particles collimated to have a momentum k which falls
incident on a crystal. The particles are assumed to interact with either the elec-

Scattering Geometry
Detector


dΩ
k'
Scattered Beam

Incident Beam Scattering angle

k θ

Sample

Figure 94: A collimated beam of particles with momentum k is elastically scat-


tered by the sample. The scattered particles are observed at the detector.

trons or nuclei of the solid. One example is x-ray diffraction in which the beam
of photons interacts elastically with the electron density,. A second example
is neutron diffraction experiments in which the beam of neutrons interacts, via
short-ranged nuclear forces, with the nuclei of the solid.

The interaction Hamiltonian between a particle in the beam and the relevant
particles of the solid can be represented as the sum of single-particle interactions
X
Ĥint = Vj (r − rj ) (194)
j

145
Here, r represents the position of the beam particle and rj is the position of the
j-th particle in the solid.

For x-rays in which the energy of the photon is in the keV range, the photon
energy is much greater than the electronic energy scale. This has the effect
that only certain terms of the interaction Hamiltonian between the x-rays and
the electron need be considered. The non-relativistic form of the interaction
between the electromagnetic field represented by a vector potential A(r, t) and
particles of charge q and mass m is given by
"
q2
  
X q
Hint = − p̂j . A(rj , t) + A(rj , t) . p̂j − A(r j , t) . A(r j , t)
j
2mc 2 m c2
(195)
where rj and p̂j are the position and momentum of the j-th particle. The
first pair of terms involve processes in which a single photon is absorbed or
emitted, whereas the last term involves the interaction of two photons with the
charged particle. To calculate the cross-section for light scattering, one needs to
consider terms of fourth order in the vector potential A(r, t), as both the initial
and final states each involve a photon. In principle, this requires including
the first pair of terms in fourth order as well as the last term in second order.
However, as the fourth order processes involve intermediate states in which
a very high energy photon has either been absorbed or emitted, the energy
denominator involving the intermediate state is large. Thus, these contributions
can safely be ignored and only the last term in the interaction need be considered
explicitly in the calculation of the elastic scattering cross-section. Therefore, in
this approximation,
P the x-rays couple to the density of the charged particles,
ρ(r) = j δ( r − rj ). For electrons, the coupling constant is proportional
to the length scale given by
e2
 2
e h̄
= ∼ 2.82 × 10−15 m (196)
me c2 h̄ c me c
which involves the fine structure constant and the Compton wave length. The
resulting length scale is the so-called classical radius of the electron.

4.3.1 Time Dependent Perturbation Theory


We shall assume that, asymptotically as t → − ∞ before the interaction is
turned on, the scattered particles have the asymptotic form of a momentum
eigenstate with eigenvalue h̄ k and energy E(k)
  12    
1 E(k) t
lim Ψk (r, t) → exp + ik.r exp − i (197)
t → − ∞ V h̄
It is expected that close to the source of the particles, the incident beam can be
described in terms of asymptotically free particles at all times. The choice of

146
normalization corresponds to one particle per volume V . The time-independent
part of the asymptotic initial state will be denoted by | k > in Dirac notation.
The scattered wave at the detector has an asymptotic form of a momentum
eigenstate | k 0 > with momentum eigenvalue h̄ k 0 .

The matrix elements of the interaction potential are given as


Z    
1 X
< k 0 | Ĥint | k > = d3 r exp − i k 0 . r Vj (r − rj ) exp + i k . r
V j V
Z        
1 X
= d3 r0 exp + i k − k 0 . r0 Vj (r0 ) exp + i k − k 0 . rj
V j V

(198)
where we have substituted r0 = r − rj in the second line. The integration
over R0 yields the Fourier transform of the interaction potential between the
scattered particle and the j-th particle of the system
Z  
0
Vj (q) = d r exp + i q . R Vj (r0 )
3 0
V
Z  
3 0 0
≈ d r exp + i q . r Vj (r0 ) (199)

Hence, we find
 
0 1 X
< k | Ĥint |k > = Vj (q) exp + i q . rj (200)
V j

where the sum over j runs over all the particles of the target material.

Given one incident particle in the state | Ψk (t) >, which is initially in an
energy eigenstate | k > before the interaction Ĥint is turned on adiabatically
at t → − ∞, the state of this particle evolves according to the Schrodinger
equation  

i h̄ | Ψk (t) > = Ĥ0 + Ĥint | Ψk (t) > (201)
∂t
As the interaction is weak, the Schrodinger equation can be solved perturba-
tively using the interaction representation. In the interaction representation,
the states are transformed through a unitary operator in a manner such that
 
i
| Ψ̃k (t) > = exp + Ĥ0 t | Ψk (t) > (202)

This unitary transformation would make the eigenstate of the non-interacting
particle time-independent. However, the presence of a non-zero interaction term
leads to the time-dependent equation of motion
∂ ˆ (t) | Ψ̃ (t) >
i h̄ | Ψ̃k (t) > = H̃ int k (203)
∂t

147
where the new interaction operator is time-dependent and is given by
   
ˆ (t) = exp + i Ĥ t Ĥ
H̃ −
i
int 0 int exp Ĥ0 t (204)
h̄ h̄
The equation of motion in the interaction representation can be solved by iter-
ation. The equation is integrated to yield
Z t
i ˆ (t0 ) | Ψ̃ (t0 ) >
| Ψ̃k (t) > = | k > − dt0 H̃ int k (205)
h̄ −∞
On iterating once, it is found that the state is given to first order in the inter-
action by
Z t
i ˆ (t0 ) | k > + . . .
| Ψ̃k (t) > = | k > − dt0 H̃ int (206)
h̄ −∞
This shows that, if wave function is started in an initial state which is an en-
ergy eigenstate of the unperturbed Hamiltonian, the time evolution caused by
the interaction will admix other states into the wave function. In this sense,
the particle described by the wave function may be considered as undergoing
transitions between the unperturbed energy eigenstates.

4.3.2 The Fermi Golden Rule


The rate at which the particle makes a transition from the initial state | k >
to state | k 0 >, due to the effect of Ĥint , is given in second order perturbation
theory by the Fermi Golden rule. The probability that the system has made a
transition at time t is given by the squared modulus of the transition amplitude
< k 0 | Ψk (t) > (207)
However, it is more convenient to calculate the probability based on the matrix
elements evaluated in the interaction representation
< k 0 | Ψ̃k (t) > (208)
These two quantities are equivalent, as they are simply related via
 
i
< k 0 | Ψk (t) > = exp − E(k 0 ) t < k 0 | Ψ̃k (t) > (209)

and the phase factor cancels out in the squared modulus.

To first order in the perturbation, the transition amplitude is given by


Z t
i ˆ (t0 ) | k >
< k 0 | Ψ̃k (t) > = − dt0 < k 0 | H̃ int
h̄ −∞
Z t  
i i
= − dt0 exp ( E(k 0 ) − E(k) − i η ) t0 < k 0 | Ĥint | k >
h̄ −∞ h̄
(210)

148
where E(k) and E(k 0 ) are the unperturbed (non-interacting) energies of the
initial and final states of the beam particles. The factor η corresponds to adia-
batically switching on the interaction at t0 → − ∞. The probability that the
transition has occurred at time t is given by
Z t   2
1 0 i 0 0 0

2 dt exp ( E(k ) − E(k) − i η ) t < k | Ĥ int | k >
h̄ −∞ h̄
(211)

The rate at which the transition occurs is given by the time derivative of the
transition probability
Z t   2
1 ∂ i
P (k → k 0 , t) = 2 0 0 0 0

dt exp ( E(k ) − E(k) − i η ) t < k | Ĥ int | k >
h̄ ∂t
−∞ h̄
(212)

The transition rate is evaluated as


 
2 η t
 exp h̄

0 0 ∂ 2
P (k → k , t) = | < k | Ĥint |k > |
∂t ( E(k 0 ) − E(k) )2 + η 2
 
2 2ηt η
= | < k 0 | Ĥint 2
| k > | exp 0
h̄ h̄ ( E(k ) − E(k) )2 + η 2
(213)

Then, in the limit η → 0, the transition rate becomes time-independent and the
energy-dependent factor reduces to π times an energy conserving delta function
since
η
lim 0 = π δ( E(k 0 ) − E(k) ) (214)
η → 0 ( E(k ) − E(k) )2 + η 2

Hence, we have obtained the Fermi Golden rule



lim P (k → k 0 , t) = | < k 0 | Ĥint | k > |2 δ( E(k 0 ) − E(k) )
η → 0 h̄
(215)

This expression represents the probability per unit time for a transition to oc-
cur from the initial state to a very specific final state with a precisely known k 0
that exactly conserves energy. As the rate contains a Dirac delta function, it is
necessary, for the rate to be mathematically meaningful, to introduce a distri-
bution of final states. Thus, one must sum over all states with k 0 in the solid
angle subtended by the detector, irrespective of the magnitude of k 0 . Thus, the
Dirac delta function is to be replaced by the density of final states with energy
E = E(k) which are travelling in the direction of the detector. That is, for
the particles to be detected, the particles must have been scattered through an

149
angle θ into the solid angle dΩ which is subtended by the detector.

The probability that a particle makes the transition from state k to states
with final momentum in a solid angle dΩ distributed around k 0 , per unit time,
is given by summing over the number of allowed final states
Z ∞
V 2π
P (k → dΩ) = 3
dk 0 k 02 dΩ | < k 0 | Ĥint | k > |2 δ( E − E(k 0 ) )
(2π) 0 h̄

= | < k 0 | Ĥint | k > |2 ρdΩ (E, k 0 ) (216)

where ρdΩ (E, k 0 ) is the density of final scattering states per unit energy range,
defined as
Z ∞
V
ρdΩ (E, k 0 ) = dΩ dk 0 k 02 δ( E − E(k 0 ) ) (217)
( 2 π )3 0

The matrix elements of the interaction operator are to be evaluated with k 0


that have the magnitude of k and are headed in the direction of the detector.
That is, the particles which are detected must be travelling towards the detector
(θ, ϕ) such that their velocities fall within the solid angle dΩ subtended by the
detector to the target.

4.3.3 The Elastic Scattering Cross-Section


The scattering cross-section is defined by
 

dΩ = P (k → dΩ) / F (218)
dΩ
where the incident flux F is the density of particles (which is one per unit
volume, i.e. V1 ) times the velocity. For massive particles, the velocity is just
h̄ k
m . Thus, for particles of mass mn , the flux is given by

h̄ k
F = (219)
mn V
On changing the variable of integration from dk 0 to dE 0 , the density of final
states is evaluated by integrating over the energy conserving delta function
Z ∞
V dk 0 k 02
ρdΩ (E, k 0 ) = dE 0
dΩ δ( E − E(k 0 ) )
( 2 π )3 0 dE 0
V dk 0 k 02
= dΩ (220)
(2π) 3 dE 0
where the magnitude of k 0 is determined by the solution of E = E(k 0 ), hence
k 0 = k. For massive particles, one has the energy momentum relation
h̄2 k 0
dE 0 = dk 0 (221)
mn

150
and so, the density of final states can be written as
V mn k 0
ρdΩ (E, k 0 ) = dΩ (222)
( 2 π )3 h̄2
On inserting the Fermi golden rule expression for P (k → dΩ)

P (k → dΩ) = | < k 0 | Ĥint | k > |2 ρdΩ (E, k 0 ) (223)

the final density of states ρdΩ (E, k) and the flux F into the expression eqn(218)
for the scattering cross-section, one finds that the elastic scattering cross-section
for massive particles such as neutrons is calculated as
 2 Z 2
dσ V mn 3 ∗

= 2
d r Ψk0 (r) Ĥint (r) Ψk (r)
dΩ 2 π h̄
V
 2 X   
mn ∗
= V j (q) V j 0 (q) exp i q . r j − r j 0
2 π h̄2 j,j 0
(224)

where q is the scattering vector

k − k0 = q (225)

The magnitude of the scattering vector is related to the scattering angle θ via
θ
q = 2 k sin (226)
2
On substituting the point contact interaction appropriate for nuclear scattering,
and noting that the Fourier transform of the delta function is q independent,
one finds the expression for the Fourier component of the potential

2 π h̄2
Vj (q) = bj (227)
mn
For crystals with a mono-atomic basis that are in static equilibrium, the posi-
tions of the nuclei rj can be identified with the lattice vectors Rj . Substitution
of Vj (q) in the above expression for the cross-section, yields the formula for the
elastic neutron scattering cross-section
  
dσ X

= bj bj 0 exp i q . Rj − Rj 0 (228)
dΩ 0
j,j

previously discussed.

For massless particles such as photons, the incident flux is just


c
F = (229)
V

151
if the incident vector potential is normalized to yield one photon per volume V .
The appropriately normalized vector potential can be expressed as
r  
4 π h̄
A(r, t) = êα c exp i ( k . r − ω t ) + c.c. (230)
2ωV
With this normalization, the vector potential represents one incident photon
per volume V , with frequency ω and incident polarization êα . The density of
final states (for polarization êβ ) is just

V k 02
ρdΩ (E, k 0 ) = dΩ (231)
( 2 π )3 h̄ c
Thus, it is found that the cross-section for elastic x-ray scattering is simply
given by
2 Z 2
V 2 ω2 e2

dσ 2π 3 ∗

= 2 3 2
d r Ak 0 (r) . ρ̂(r) Ak (r)
dΩ h̄ c ( 2 π c ) 2 me c
V

2  2 X
e2
   


= êα . êβ
S(q) S (q) exp i q . Rj − Rj 0
me c2
j,j 0
(232)

where the structure factor S(q) is the contribution of a unit cell to the Fourier
transform of the electron density. The vectors Ri are the lattice vectors. Thus,
the factors of V and ω cancel, leading to a scattering cross-section that only
depends on the Fourier transform of the electronic density and has a coupling
constant which is the square of the classical radius of the electron
2
e2

re2 = (233)
me c2
From the form of this coupling constant, it can be seen that the scattering of
x-rays from the density of charged nuclei is entirely negligible compared with
the scattering from the electron density.

4.3.4 The Condition for Coherent Scattering


Consider scattering from a crystal which has a mono-atomic basis and has a
finite spatial extent. In this case, the subscript on the atomic potential can be
dropped, and the summation over j and j 0 run over all the lattice sites. For
convenience, it shall be assumed that the crystal has the same shape as the
primitive unit cell but has overall dimensions ( N1 − 1 ) a1 , ( N2 − 1 ) a2
and ( N3 − 1 ) a3 along the various primitive lattice directions. The solid,
therefore, contains a total of N1 N2 N3 primitive unit cells and since the basis
consists of one atom, the solid contains a total of N = N1 N2 N3 atoms.

152
The summation over Rj in the scattering cross-section can be performed by
expressing the general reciprocal lattice vector in terms of the primitive lattice
vectors,
X   X      
exp i q . Rj = exp i n1 q . a1 exp i n2 q . a2 exp i n3 q . a3
j n1 ,n2 ,n3
(234)
The sums over n1 runs from 0 to N1 − 1, and similarly for n2 and n3 . This gives
the products of three factors, each of the form
 
n1 =N
X1 −1   1 − exp i N 1 q . a1
exp i n1 a . a1 =  
n1 =0 1 − exp i q . a1
 
( N1 − 1 )
= exp + i q . a1 ×
2
    
N1 N1
exp i 2 q . a1 − exp − i 2 q . a1
×    
1 1
exp i 2 q . a1 − exp − i 2 q . a1
!
N1
sin 2 q . a1
 
( N1 − 1 )
= exp + i q . a1 1
2 sin 2 q . a1
(235)
This function exhibits the effect of the constructive and destructive interference
between the scattered waves emanating from the various atoms forming the
solid. The numerator of the function falls to zero at
2mπ
q . a1 = (236)
N1
for general integer values of m. The numerator has maximum magnitude at
(2m + 1)π
q . a1 = (237)
N1
The overall q dependence is dominated by the denominator which falls to zero
when q . a1 = 2 m π for integer m. At these special q values, the function has
to be evaluated by l’hopital’s rule and has the limiting value of N1 . This occurs
since, for these q values, the exponential phase factors are all in phase (and
equal to unity) and so the sum over the N1 terms simply yields N1 . Thus, the
scattering cross-section is proportional to the product of the modulus square of
three of these factors
2

= êα . êβ re2 | F (q) |2 ×


dΩ

153
!2 !2 !2
N1 N2 N3
sin 2 q . a1 sin 2 q . a2 sin 2 q . a3
× 1 1 1
sin 2 q . a1 sin 2 q . a2 sin 2 q . a3
(238)
Since for a macroscopic solid the numbers N1 , N2 and N3 are of the order of
107 , the three factors rapidly vary with the magnitudes of q . ai . The maxima

Diffraction Pattern for a finite one-dimensional crystal


500
N=20
400
Ω(q)

300
σ/dΩ

200

100

0
-1.5 -1 -0.5 0 0.5 1 1.5

π
q a /π

Figure 95: The scattering cross-section of a one-dimensional lattice with N unit


cells, as a function of the momentum transfer.

occur when the three conditions


q . a1 = 2 π m1
q . a2 = 2 π m2
q . a3 = 2 π m3
(239)
are satisfied. These special values of q are denoted by Q. In this case, one finds
that the scattering cross-section is simply proportional to
2

∼ êα . êβ re2 | F (Q) |2 N 2

(240)
dΩ
which is just equal to the square of the number of atoms in the solid. The
coherent scattering from an ordered solid should be contrasted with incoherent

154
scattering from the atoms of a gas. Due to the positional disorder in the gas, the
phase factors may be considered to be random. The net scattering intensity for
scattering of a gas of N atoms is then approximately equal to just N times the
scattering intensity for an isolated atom. The coherent scattering from atoms in
a solid possessing long-ranged order is a factor of N 2 larger than the scattering
intensity for an isolated atom. In summary, the condition that there is complete
constructive interference between all the atoms in the solid is given by
 
exp i Q . Ri = 1 ∀ i (241)

The intensity of the scattered beam is exceptionally large at these special values
Q, compared with all other q values. Thus, coherent scattering is the dominant
feature of diffraction from crystalline solids but occurs only infrequently, as it
only occurs when the scattered wave length and scattering angle satisfy the
above stringent condition. These special values of Q are the lattice vectors of
the reciprocal lattice.

——————————————————————————————————

4.3.5 Exercise 27
Consider a sample with N unit cells arranged in M micro-crystals that are
oriented parallel with respect to each other, but their positions are random.
Calculate the width and height of the Bragg peak.

——————————————————————————————————

4.3.6 Exercise 28
At finite temperatures, the atoms of a crystal undergo thermal vibrations. Due
to the vibrations, the intensity of the Bragg peaks are reduced by a Debye-
Waller factor which involves the spectrum of lattice vibrations. However, this
situation can be approximately modelled by assuming that each atom undergoes
a small random displacement δ R from its equilibrium position R. Assume that
the displacements are small compared with the separation between neighbor-
ing atoms, | δ R |  a, and are Gaussian distributed. Also assume that the
displacements of different atoms are entirely uncorrelated δi,R δj,R0 = 0 for
R 6= R0 . Calculate the diffraction peak intensity, and show that the largest
reduction in the intensity occurs for large Q values.

——————————————————————————————————

155
4.3.7 Exercise 29
Evaluate the effect of a significant number of thermally induced vacancies (miss-
ing atoms) in the elastic scattering cross-section from a crystal.

——————————————————————————————————

4.3.8 Anti-Domain Phase Boundaries


The order-disorder transition usually starts at several nucleation centers in a
crystal. For CuZn, the underlying CsCl lattice can be divided into two inter-
penetrating simple cubic sub-lattices: the A and B sub-lattice. In several re-
gions, the nucleation may start with the Cu atoms condensing on the A sub-
lattice, whereas the nucleation may occur in other regions where the Cu atoms
condense on the B sub-lattices. These distinct domains of nucleation grow and
spread through the crystal until they meet and the entire crystal is ordered.
The interfaces of the different domains meet at anti-domain phase boundaries
at which there is a mismatch of the long-ranged ordering of the atoms. Due to
the mismatch, two planes containing similar atoms form the anti-domain phase
boundary30 . The effect of anti-domain phase boundaries is to smear out the
“super-lattice” Bragg peaks. This can be seen by considering the amplitude
of the scattered x-rays as a superposition of the scattering from the various
domains. For simplicity, let us consider the scattering from two domains of
identical shape and size. If the scattering amplitude from one domain is de-
noted by A1 (q) and the scattering from the second domain is denoted by A2 (q)
then, as the scattering amplitudes are additive, one obtains

A(q) = A1 (q) + A2 (q) (242)

where  
A2 (q) = exp i q . δR A1 (q) (243)

δR is the vector displacements of the origins of the two domains. The scattering
amplitude A1 (q) is given by
! N2 qy a
! !
sin N1 2qx a sin 2 sin N3 2qz a
A1 (q) ∝ qy a
sin qx2 a sin 2
sin qz2 a
(244)

For a domain wall in the y − z plane, the displacement between the origins of
the Cu sub-lattices in the two domains is given by
1 a a
δR = ( N1 + ) a êx + êy + êz (245)
2 2 2
30 F. W. Jones and C. Sykes, Proc. Roy. Soc. A, 166, 376 (1938).

156
Hence, for a CsCl-type structure and if q is close to Q, the total scattering
amplitude is given by the expression
  
A(q) ∼ A1 (q) 1 + ( − 1 )m1 +m2 +m3 exp i N1 qz a (246)

The total intensity of the scattered wave is proportional to


 
2 m1 +m2 +m3
I(q) ∝ 2 | A1 (q) | 1 + (−1) cos N1 qx a (247)

Thus, if m1 + m2 + m3 is even, the intensity is modulated by the factor


N 1 qx a
4 cos2 (248)
2
whereas if m1 + m2 + m3 is odd, the intensity is modulated by the factor
N1 qx a
4 sin2 (249)
2
This factor is due to the interference of the scattering from the two domains.
The destructive interference causes an exact cancellation of the intensity at the
exact Bragg wave vector at odd m1 + m2 + m3 . However, for qx slightly off the
Bragg position

qx = m1 + δqx
a
π
δqx ∼ (250)
N1 a
the scattered intensity is finite and large. That is, the single anti-domain phase
boundary between identical domains of identical shapes and sizes produces a
hole in the Bragg peak with odd m1 + m2 + m3 .

For a crystal with a CuCl type structure which contains several anti-domain
phases, one expects there to be three sets of anti-domain phase boundaries, and
one expects that each domain has a different size. On averaging over the distri-
bution of domains, one expects the small oscillations in the scattered intensity
from the single domain S1 (q) to be washed out. Furthermore, one expects that
the intensities of the “super-lattice” peaks to be smeared out in q space.

4.3.9 Exercise 30
Consider the scattering produced by a CuCl type material with anti-domain
walls. For simplicity, only consider the component of the scattering ampli-
tude associated with a single primitive lattice vector. That is, consider a one-
dimensional model. Let p be the probability of not crossing a domain wall on

157
traversing one step a along the primitive lattice vector, and let q be the proba-
bility of crossing a domain wall, where q ∼ N11 . Show that the average scattered
intensity near the “super-lattice” peaks is proportional to the factor
N1
X
| A(qx ) |2 ∝ N1 + ( N1 − m1 ) ( p − q )m1 2 cos m1 qx a (251)
m1 =1

Evaluate the summation. Hence, show that as the number of domain walls in-
creases, the intensities of the “super-lattice” Bragg peaks are diminished and
acquire low amplitude tails.

4.4 Elastic Scattering from Quasi-Crystals


The scattering intensity from three-dimensional quasi-crystals show ten-fold,
six-fold and five-fold symmetric diffraction patterns which can be understood
as arising from a space of six or more dimensions. Icosahedral symmetry can be
found in a six dimensional hyper-cubic lattice. An icosahedron has 20 identical
faces made of equilateral triangles. Five of the faces meet at each of the twelve
vertices of the icosahedron. These vertices are responsible for the five-fold sym-
metry.

Figure 96: Selected electron diffraction patterns from a single grain of an Al −


M n quasi-crystal. [After Gratais et al. (1984).]

The x-ray scattering amplitude A(q) from a one-dimensional quasi-crystal


can be found by a projection from a two-dimensional lattice. The amplitude is

158
a linear superposition from the scattered amplitudes from the sites sn , where

sn = n a cos θ + m0 a sin θ (252)

and where the points (na, m0 a) are restricted to lie in a two-dimensional strip.
The amplitude is given by
X  
A(q) = exp i q sn
n
X  
= exp i q a ( cos θ n + sin θ m0 )
n,m0
X  
= exp i q a ( cos θ n + sin θ m ) Θ(1 + (n + 1) tan θ − m) Θ(m − n tan θ)
n,m
(253)

This can be expressed as an integral over a two-dimensional space


X  
A(q) = exp i q a ( cos θ n + sin θ m ) Θ(1 + (n + 1) tan θ − m) Θ(m − n tan θ)
n,m
Z Z   X
= dx dy exp i q ( cos θ x + sin θ y ) δ(x − na) δ(y − ma) ×
n,m
× Θ(a + (x + a) tan θ − y) Θ(y − x tan θ)

This is a two-dimensional Fourier transform


Z   X
A(q) = d2 r exp i q . r δ(x − na) δ(y − ma) ×
n,m
× Θ(a + (x + a) tan θ − y) Θ(y − x tan θ)
(254)

which is to be evaluated with q restricted to have values on the one-dimensional


line
q = q ( cos θ , sin θ ) (255)
The two-dimensional Fourier transform is recognized as the Fourier transform
of a product Z  
2
A(q) = d r exp i q . r B(r) C(r) (256)

where B(r) is non-zero on the sites of a two-dimensional array


X
B(r) = δ(x − na) δ(y − ma) (257)
m,n

and the function C(r) projects onto a two-dimensional strip

C(r) = Θ(a + (x + a) tan θ − y) Θ(y − x tan θ) (258)

159
The Fourier transform of the product of functions can be evaluated using the
convolution theorem. The result is given by the convolution of the product of
Fourier-transformed functions
d2 q 0
Z
A(q) = B(q − q 0 ) C(q 0 ) (259)
( 2 π )2

The function B(q) is the scattering amplitude from the two-dimensional lattice
X  
B(q) = exp i ( qx n a + qy m a ) (260)
n,m

while the function C(q 0 ) is evaluated as


!
exp[ i qy0 a (1 + tan θ) ] − 1
Z  
0
C(q ) = dx exp i ( + qx0
tan θ ) xqy0
i qy0
!
0 0
exp[ i qy0 a (1 + tan θ) ] − 1
= ( 2 π ) δ( qx + qy tan θ )
i qy0
(261)

The scattering amplitude for the two-dimensional lattice is only non-zero at the
two-dimensional reciprocal lattice vectors q = Q. Thus, the scattering from
the the two-dimensional lattice is represented by the factor
 2 X

B(q) = δ 2 (q − Q) (262)
a
Q

Hence, we find that the amplitude in the two-dimensional space is given by


1 X
A(q) = C(q − Q)
a2
Q

2π X
= δ( qx − Qx + ( qy − Qy ) tan θ ) ×
a2
Q
!
exp[ i ( qy − Qy ) a (1 + tan θ) ] − 1
×
i ( qy − Qy )
(263)

Evaluating this on the line in q space yields the amplitude for scattering from
the one-dimensional quasi-crystal
X
A(q) = 2 π cos θ δ( q a − Qx a cos θ − Qy a sin θ ) ×
Q

160
!
exp[ i ( q sin θ − Qy ) a (1 + tan θ) ] − 1
×
i ( q a sin θ − Qy a )
X
= 2π δ( q a − Qx a cos θ − Qy a sin θ ) ×
Q
!
exp[ i ( Qx sin θ − Qy cos θ ) a (cos θ + sin θ) ] − 1
×
i ( Qx a sin θ − Qy a cos θ )
(264)

This has delta function-like peaks at the wave vectors given by

q a = 2 π ( m1 cos θ + m2 sin θ ) (265)

where m1 and m2 are integers. In contrast to the scattering from a one-


dimensional periodic crystal, the peaks in the scattering cross-section of a one-
dimensional quasi-crystal are indexed by two integers. The intensities of the
peaks are proportional to
 
sin2 π ( m1 sin θ − m2 cos θ ) ( cos θ + sin θ )
| A(q) |2 ∝ (266)
( m1 sin θ − m2 cos θ )2
Thus, the inelastic scattering spectra consists of a dense set of sharp peaks, but
with varying intensities. The intensities are large when the ratios of m2 and m1
are close to the value of tan θ. The quasi-crystal diffraction pattern collapses

0.8

0.6
I(q)/I(0)

0.4

0.2

0
-10 -5 0 5 10
(qa/π)

Figure 97: Calculated Intensities of the Diffraction Peaks for a one-dimenional


quasi-crystal.

into the diffraction pattern of a one-dimenional crystal with a non-trivial basis,

161
in the limit when tan θ is a rational number.

Data from high resolution x-ray diffraction experiments on a three-dimensional


quasi-crystal31 are shown in fig(98).

Figure 98: High resolution x-ray diffraction spectra of quenched (top) and an-
nealed (bottom) Al − M n powder. [After Bancel et al. (1995).]

4.5 Elastic Scattering from a Fluid


The structure of a fluid, as expressed by the pair correlation function, can be
inferred from elastic scattering experiments. The intensity of a beam of particles
scattered from a liquid can be considered as analogous to the scattering from a
solid with an infinite unit cell. First, we shall consider the atoms of the fluid
as static point particles. The amplitude of the beams scattered from each atom
add, giving a total amplitude which is proportional to
X  
S(q) = exp i q . rj
j
Z   X
3
= d r exp iq.r δ 3 (r − rj )
j
(267)

The scattering intensity is given by the square of the scattered amplitude

I(q) ∝ | S(q) |2
X    
= exp + i q . ri exp − i q . rj
i,j

31 P. A. Bancel, P. A. Heiney, P. W. Stephens, A .I. Goldman and P. Horn, Phys. Rev. Lett.

54, 2422 (1995).

162
Z Z   X
= d3 r d3 r0 exp i q . ( r − r0 ) δ 3 (r − ri ) δ 3 (r0 − rj )
i,j
(268)

On considering the long time average of the atomic positions, one obtains
Z Z   X
3 3 0 0
I(q) ∝ d r d r exp i q . ( r − r ) δ 3 (r − ri ) δ 3 (r0 − rj )
i,j
Z Z  
= d3 r d3 r0 exp i q . ( r − r0 ) C(r, r0 )

(269)

The scattering intensity can be expressed in terms of the radial distribution


function g(r), since

C(r − r0 ) = δ 3 (r − r0 ) ρ(0) + g(r − r0 ) (270)

Hence, the
Z Z Z  
3 0 0
I(q) ∝ d r ρ(0) + 3
d r 3
d r exp i q . ( r − r ) g(r − r0 )
Z  
= N + V d3 r exp i q . r g(r)

(271)

However, the integral over g(r) can be split into two parts
Z   Z  
2 2
I(q) ∝ N + V d3 r exp i q . r ρ(0) + V d3 r exp i q . r ( g(r) − ρ(0) )
Z  
2 2
= N + V ( 2 π )3 ρ(0) δ 3 (q) + V d3 r exp i q . r ( g(r) − ρ(0) )

( 2 π )3 3
Z  
2
= N + N2 δ (q) + V d3 r exp i q . r ( g(r) − ρ(0) )
V
Z ∞
( 2 π )3 3 4π 2
= N + N2 δ (q) + V dr r sin q r ( g(r) − ρ(0) )
V q 0
(272)

The first term represents the incoherent scattering. The second term represents
coherent forward scattering. The integral in the last term is convergent and
yields non-trivial information about the structure of the fluid. As an example,
g(r) has been derived from measurements of the intensity of elastically-scattered
neutrons I(q), for liquid argon near its triple point32 .

32 J. L. Yarnell, M. J. Katz, R. G. Wenzel, and S. H. Koenig, Phys. Rev. A, 7, 2130 (1973).

163
Figure 99: A comparison between the theoretical and the experimentally deter-
mined structure factor of liquid argon. [After Yarnell et al. (1975).]

5 The Reciprocal Lattice


The reciprocal lattice vectors play an important role in describing the properties
of a solid that has periodic translational invariance. Any property of the solid,
whether scalar, vector or tensor, should have the same periodic translational
invariance as the potential due to the charged nuclei. This means that, due to
the translational invariance, physical properties only need to be specified in a
finite volume, and this volume can then be periodically continued over all space.
The vectors of the reciprocal lattice play an important and special role in the
Fourier transform of the physical quantity.

The reciprocal lattice vectors have dimensions of inverse distance and are
defined in terms of the direct primitive lattice vectors a1 , a2 and a3 . The
primitive reciprocal lattice vectors b(i) , are defined via the scalar product

ai . b(j) = 2 π δij (273)

where the Kronecker delta function δij has the value 1 if i = j and is zero
if i 6= j. Thus, the primitive reciprocal lattice vectors are orthogonal to two
primitive lattice vectors of the direct lattice.

164
The primitive reciprocal lattice vectors can be constructed via
a2 ∧ a3
b(1) = 2π
a1 . ( a2 ∧ a3 )
a3 ∧ a1
b(2) = 2π
a1 . ( a2 ∧ a3 )
a1 ∧ a2
b(3) = 2π
a1 . ( a2 ∧ a3 )
(274)

where the last two expressions are found from the first by cyclic permutation of
the labels (1, 2, 3). The denominator is just the volume of the primitive unit cell.

The reciprocal lattice consists of the points given by the set of vectors Q
where
Q = m1 b(1) + m2 b(2) + m3 b(3) (275)
and (m1 , m2 , m3 ) are integers. This set of vectors are the reciprocal lattice vec-
tors. The reciprocal lattice vectors denote directions in the reciprocal lattice or
are the normals to a set of planes in the direct lattice. In the latter case, as
it shall be seen, the numbers (m1 , m2 , m3 ) are equivalent to Miller indices and,
hence, are enclosed in round brackets.

——————————————————————————————————

5.0.1 Exercise 31
Find the volume of the primitive unit cell of the reciprocal lattice.

——————————————————————————————————

5.1 The Reciprocal Lattice as a Dual Lattice


The reciprocal lattice vectors can be considered to be the duals of the direct
lattice vectors. This relation can be seen by expressing the primitive lattice
vectors aj in terms of the primitive reciprocal lattice vectors bi , via
1 X
aj = gj,i b(i) (276)
2π i

If the quantity gi,j is identified as the metric, then the above relation defines
the set of b(i) to be basis vectors which are dual to the set of basis vectors ai .
The quantity gi,j is given by the metric, since
1 X
aj . ak = gj,i b(i) . ak (277)
2π i

165
and since
b(i) . ak = 2 π δki (278)
one has
gj,k = aj . ak (279)
Hence, gj,k is defined to be the metric tensor. The metric tensor has the property
that it expresses the length s of a vector r in terms of its components xi along
the basis vectors ai . That is, if one expresses the vector r in terms of the basis
vectors ai and the components xi via
X
r = xi ai (280)
i

then, for a constant metric, the length is given in terms of the covariant com-
ponents33 xi via X
s2 = gi,j xi xj (281)
i,j

The metric tensor, when evaluated in terms of the parameters of the primitive
unit cell, is given by the matrix

a21
 
a1 a2 cos α3 a1 a3 cos α2
( gi,j ) =  a1 a2 cos α3 a22 a2 a3 cos α1  (282)
a1 a3 cos α2 a2 a3 cos α1 a23

The inverse transform is given by


X
b(i) = 2 π g i,k ak (283)
k

where the quantity g i,k is identified as the metric for the dual vectors. Since
1 X
aj = gj,i b(i)
2π i
X X
= gj,i g i,k ak (284)
i k
33 The components xi which occur in the decomposition
X
r = xi a i
i

are usually known as the covariant components of r, whereas if one decomposes the vector r
in terms of the set of basis vectors b(i) via
X
r = xi b(i)
i

then the xi are known as the contra-variant components of r.

166
and as X
aj = δjk ak (285)
k
one infers that X
δjk = gj,i g i,k (286)
i
Hence, the metric tensor (gi,j ) is the inverse of the metric tensor (g i,j ) for the
dual vectors.

The volume of the unit cell, Vc , is given by


Vc2 = det ( gi,j ) (287)
or
 
Vc2 = a21 a22 a23 2 2 2
1 − cos α1 − cos α2 − cos α3 + 2 cos α1 cos α2 cos α3

(288)
The dual metric tensor is given by the inverse of the metric tensor. The dual
metric tensor is expressed as the matrix
a2 a2 (1−cos2 α ) a2 a a (cos α cos α −cos α ) a2 a a (cos α cos α 3 −cos α2 )
 
2 3 1 3 1 2 1 2 3 2 1 3 1
Vc2 Vc2 Vc2
a23 a1 a2 (cos α1 cos α2 −cos α3 ) a21 a23 (1−cos2 α2 ) 2
a1 a2 a3 (cos α2 cos α3 −cos α1 )
g i,j
  
= 
 Vc2 Vc2 Vc2


a22 a1 a3 (cos α1 cos α3 −cos α2 ) a21 a2 a3 (cos α2 cos α3 −cos α1 ) a21 a22 (1−cos2 α3 )
Vc2 Vc2 Vc2
(289)
This dual metric is also defined as
bi . bj = ( 2 π )2 g i,j (290)
From this, one can immediately find that the length of the reciprocal lattice
vectors are given by
a2 a3
b1 = 2 π | sin α1 | (291)
Vc
etc., and the angle β3 between b(1) and b(2) is given by
( cos α1 cos α2 − cos α3 )
cos β3 = (292)
| sin α1 sin α2 |
etc. On using the inverse transformation, the reciprocal lattice vectors are given
in terms of the primitive direct lattice vectors by
a2 a2 a2 (1 − cos2 α1 )
 
(cos α1 cos α2 − cos α3 ) (cos α1 cos α3 − cos α2 )
b(1) = 2π 1 22 3 a1 + a2 + a3
Vc a21 a1 a2 a1 a3
2 2 2 2
 
a a a (cos α1 cos α2 − cos α3 ) (1 − cos α2 ) (cos α2 cos α3 − cos α1 )
b(2) = 2π 1 22 3 a1 + a2 + a3
Vc a1 a2 a22 a2 a3
2 2 2
(1 − cos2 α3 )
 
a a a (cos α1 cos α3 − cos α2 ) (cos α2 cos α3 − cos α1 )
b(3) = 2π 1 22 3 a1 + a2 + a3
Vc a1 a3 a2 a3 a23
(293)

167
These expressions for the b(i) are equivalent to the expressions in terms of the
vector products of the primitive lattice vectors ai , and they also satisfy the
definition of the primitive reciprocal lattice vectors
ai . b(j) = 2 π δij (294)

Any vector of the direct Bravais lattice can be expressed as


R = n1 a1 + n2 a2 + n3 a3 (295)
A reciprocal lattice vector Q can also be written as

Q = m1 b(1) + m2 b(2) + m3 b(3) (296)


where (m1 , m2 , m3 ) are integers. Any vector k in the reciprocal lattice can be
represented as a superposition of the reciprocal lattice vectors
k = µ1 b(1) + µ2 b(2) + µ3 b(3) (297)
where the µi are non-integer. Thus, the scalar product of an arbitrary vector k
in the reciprocal lattice and a Bravais lattice vector R is evaluated as
 
k . R = 2 π µ1 n1 + µ2 n2 + µ3 n3 (298)

If k is a reciprocal lattice vector Q then the set of µi ’s take on integer values


mi , so that the scalar product reduces to
 
Q . R = 2 π m1 n1 + m2 n2 + m3 n3 (299)

As the sum of the products of integers is still an integer ( say M ), the Laue
condition can be expressed as
Q.R = 2πM (300)
for all R. Thus, the Reciprocal Lattice vectors satisfy the Laue condition. This
requirement is equivalent to the condition that the exponential phase factor
given by  
exp iQ.R = 1 (301)

is unity for all Bravais lattice vectors R.

The vectors Q form a Bravais lattice in which the primitive lattice vectors
can be expressed in terms of the vectors b(i) . Also, the reciprocal lattice of a
reciprocal lattice is the original direct lattice.

——————————————————————————————————

168
5.1.1 Exercise 32
Determine the primitive lattice vectors of the lattice that is reciprocal to the re-
ciprocal lattice. How are they related to the vectors of the original direct lattice?

——————————————————————————————————

5.2 Examples of Reciprocal Lattices


Now some examples of reciprocal lattices are examined.

5.2.1 The Simple Cubic Reciprocal Lattice


In terms of Cartesian coordinates, the primitive lattice vectors of the simple
cubic direct lattice are

a1 = a êx
a2 = a êy
a3 = a êz (302)

The primitive reciprocal lattice vectors are determined to be



b(1) = êx
a

b(2) = êy
a

b(3) = êz (303)
a
These are three orthogonal vectors which are oriented parallel to the direct
lattice vectors. The reciprocal lattice of the simple cubic direct lattice is also
simple cubic.

5.2.2 The Body Centered Cubic Reciprocal Lattice


In terms of Cartesian coordinates, the primitive lattice vectors of the body
centered cubic direct lattice are
 
a
a1 = êx + êy − êz
2
 
a
a2 = − êx + êy + êz
2
 
a
a3 = êx − êy + êz (304)
2

169
a3
The volume of the unit cell is Vc = | a1 . ( a2 ∧ a3 ) | = 2 .

The primitive reciprocal lattice vectors are determined to be


 
(1) 2π
b = êx + êy
a
 
(2) 2π
b = êy + êz
a
 
(3) 2π
b = êx + êz (305)
a
The three primitive reciprocal lattice vectors span the three-dimensional recip-
rocal lattice, but have different orientations from the direct lattice vectors. The
reciprocal lattice has cubic symmetry as can be seen by combining the three
primitive reciprocal lattice vectors (adding any two and subtracting the third)
to yield three orthogonal reciprocal lattice vectors of equal magnitude. The
reciprocal lattice of the body centered cubic direct lattice is face centered cubic,
with a conventional cell of side 4aπ .

5.2.3 The Face Centered Cubic Reciprocal Lattice


In terms of Cartesian coordinates, the primitive lattice vectors of the face cen-
tered cubic direct lattice are
 
a
a1 = êx + êy
2
 
a
a2 = êx + êz
2
 
a
a3 = êy + êz (306)
2
The primitive reciprocal lattice vectors are determined to be
 

b(1) = êx + êy − êz
a
 
2 π
b(2) = êx − êy + êz
a
 
2 π
b(3) = − êx + êy + êz (307)
a
These are three non co-planar vectors, but have different orientations from the
direct lattice vectors. The reciprocal lattice has cubic symmetry. This can be
seen by combining pairs of the primitive reciprocal lattice vectors, which yield
three orthogonal reciprocal lattice vectors of equal magnitude. The reciprocal
lattice of the face centered cubic direct lattice is body centered cubic, with a
conventional unit cell of side 4aπ .

170
5.2.4 The Hexagonal Reciprocal Lattice

The Hexagonal Bravais Lattice

a3

a2
a1

Figure 100: The primitive unit cell for the hexagonal Bravais lattice.

The hexagonal lattice has primitive lattice vectors which can be chosen as

 
a
a1 = 3 êx + êy
2

 
a
a2 = − 3 êx + êy
2
a3 = c êz (308)

The volume of the primitive unit cell is



3 2
Vc = a c (309)
2

The primitive reciprocal lattice vectors are


 
(1) 2π 1
b = + √ êx + êy
a 3
 
(2) 2 π 1
b = − √ êx + êy
a 3
(3) 2π
b = êz (310)
c

171
Hexagonal reciprocal lattice

b3

(2π/c)

b1 b2

Figure 101: The primitive unit cell for the reciprocal lattice of the hexagonal
Bravais lattice.

Thus, the reciprocal lattice of the hexagonal lattice is its own reciprocal lattice,
but is rotated about the z axis.

——————————————————————————————————

5.2.5 Exercise 33
A trigonal lattice is defined by three primitive lattice vectors a1 , a2 and a3 , all
of equal length a where the angle α between any pair of these lattice vectors
is a constant. Show that the three vectors a1 = [r, s, t], a2 = [t, r, s] and
a3 = [s, t, r], referenced to an orthonormal basis, are primitive lattice vectors
for a trigonal lattice. Prove that the reciprocal lattice of the trigonal lattice is
another trigonal lattice.

——————————————————————————————————

5.3 The Brillouin Zones


The first Brillouin zone is the Wigner-Seitz cell of the reciprocal lattice. That
is, the first Brillouin zone is a volume of a primitive unit cell in the reciprocal
lattice. This cell is found by first connecting a central reciprocal lattice point
O to all the other reciprocal lattice points via the reciprocal lattice vectors Qi .

172
Secondly, these connecting lines are bisected by planes. The equations for the
set of these planes are given by
 
1
k − Qi . Qi = 0 (311)
2

for each i. The smallest volume around the origin O enclosed by these planes
is the first Brillouin zone. That is, the first Brillouin zone consists of all the
regions of space that can be reached from O without crossing any of the planes.

The regions of the entire reciprocal lattice can be partitioned off into Bril-
louin zones of higher order. The planes defined by eqn(311) form a set of
boundaries for the set of Brillouin zones. The n-th order Brillouin zone consists
of the regions of k space that is accessed from the origin by crossing a minimum
of n − 1 boundaries. Although the n-th order Brillouin zone exists in the form

Brillouin Zones for a Square Lattice

(5) (5) (5)


(4) (4)
(5) (5)
(3) (3)
(5)
(4) (2) (4) (5)

(5) (3) (3) (5)


(4) (4)
(5) (2) (1) (2) (5)
(4) (4)
(5) (3) (3) (5)

(5)
(4) (2) (4) (5)

(5)
(3) (3)
(5)
(4) (4)
(5)
(5) (5)

Figure 102: The higher-order Brillouin zones of a square lattice.

of isolated regions of k space, these regions can be brought together to make


a contiguous volume by translating the isolated regions through appropriately
chosen reciprocal lattice vectors Qi .

5.3.1 The Simple Cubic Brillouin Zone


The first Brillouin zone of the simple cubic direct lattice is a simple cube cen-
tered at the origin O. It is bounded by lattice planes with normals that are
members of the set {1, 0, 0}, and thus are parallel to mirror planes of the direct

173
First Brillouin Zone for a simple cubic lattice

(2π/a)

Figure 103: The first Brillouin zone of the simple cubic lattice.

lattice. The sides of the cubic Brillouin zone are of length 2aπ and the Brillouin
zone has a volume of ( 2aπ )3 which, when given in terms of the volume of the
3
unit cell of the direct lattice, is equal to 8Vπc .

Second Brillouin Zone of the simple cubic lattice

Figure 104: The second Brillouin zone of the simple cubic lattice.

Points of high symmetry are usually given special names. Points interior to
the first Brillouin zone are designated by Greek letters and those on the surface
are designated by Roman letters. The center of the zone (0, 0, 0) is denoted by
Γ, the vertex of the cube 2aπ ( 12 , 12 , 12 ) is called R. The center of the x face lo-

174
cated at 2aπ ( 12 , 0, 0) is called X, and the mid-points of the edges at 2 π
a ( 12 , 12 , 0)
are denoted by M .

Points on high-symmetry lines are also given special designations. The points
between M and X are denoted by Z. The points on the lines between R and
X are denoted by S, the points on the lines between R and M are denoted by
T . The points on high-symmetry lines in the interior have the following desig-
nations: the points between Γ and M are denoted by Σ, the points between Γ
and X by ∆, the points on lines between Γ and R are denoted by Λ.

5.3.2 The Body Centered Cubic Brillouin Zone


The first Brillouin zone for the body centered direct lattice is a rhombic dodec-
ahedron34 . It is bounded by lattice planes that are in the set {1, 1, 0}, and thus
are parallel to planes of reflection symmetry of the direct lattice. The Brillouin
zone can be visualized by capping each square face of a cube with a square
pyramid, and noting that the triangular faces which are joined along one edge
of the cube actually form a single face. The vertices of the cube connect three
edges, whereas four edges intersect at each apex of the square pyramids. The
cell is centered at the origin Γ = (0, 0, 0). The vertices are located either on the
34 The rhombic dodecahedron is the dual of the Archimedean solid which is known as the

cuboctahedron. The Archimedean dual of a solid is formed from an Archimedean solid


by interchanging the numbers of faces with the number of vertices. Therefore, the dual
Archimedean solid has equivalent faces and congruent edges. The dual of an Archimedean
solid can be characterized by listing the number of edges per vertex in cyclic order about a
given face. In this scheme, the rhombic dodecahedron would be labelled by ( 3 . 4 )2 .

Third Brillouin Zone for a s.c. lattice

Figure 105: The third Brillouin zone of the simple cubic lattice.

175
The Brillouin zone of the body-centered cubic lattice
ez

(0,1,0)
(0,0,1)
ey
ex
(1,0,0)

Q=(2π/a)[(m1+m3)ex+(m1+m2)ey+(m2+m3)ez]

Figure 106: Brillouin zone of the b.c.c. lattice.

positive or negative Cartesian axes at H = 2aπ (1, 0, 0) or at diagonal points


P = 2aπ ( 12 , 12 , 12 ). The centers of the faces are denoted by N = 2aπ ( 12 , 12 , 0).

The Brillouin zone of the body-centered cubic lattice

P
N
Γ
H
N

Figure 107: High symmetry points and lines of the b.c.c. Brillouin zone.

176
Points on the high-symmetry lines joining P and H are denoted by F . Other
special points are: G which are on the high-symmetry line between N and H, or
D between P and N . The names of interior points on high-symmetry lines are
Σ which are intermediate between Γ and N , ∆ which are intermediate between
Γ and H, and Λ which are intermediate between Γ and P .

5.3.3 The Face Centered Cubic Brillouin Zone

The Brillouin zone for a face-centered cubic lattice


ez

1/2(1,1,1)
ey

1/2(1,1,0)
ex

Figure 108: Brillouin zone of the f.c.c. lattice.

The Brillouin zone for the face centered cubic lattice has the form of a trun-
cated octahedron. The truncated octahedron is an Archimedean solid as it has
faces which are regular polygons and as it has equivalent vertices and congruent
edges35 . The f.c.c. Brillouin zone is bounded by square faces with normals in
the set {2, 0, 0} and hexagonal faces with normals in the set {1, 1, 1}. Only
the {2, 0, 0} faces are parallel to mirror planes of the point group. The Bril-
louin zone has twenty four vertices located at W = 2aπ (1, 12 , 0). The centers
of the square faces are denoted by X and are located at 2aπ (1, 0, 0). These
squares are connected to eight hexagonal faces with centers at the L points
L = 2π 1 1 1
a ( 2 , 2 , 2 ). The mid-points of the edges joining two hexagonal faces are
2 π 3 3
at a ( 4 , 4 , 0), and are denoted by K. The mid-points of the edges between
35 The are thirteen Archimedean solids which are convex polyhedra. Their faces are com-

posed of regular polygons. Their vertices are equivalent and their edges are congruent. The
Archimedean solids are enumerated by listing the number of edges per face in cyclic order
about a given vertex. The truncated octahedron is labelled by 4 . 62 .

177

the square and hexagonal faces are denoted by U , where U = a ( 14 , 14 , 1).

The Brillouin zone for a face-centered cubic lattice

L
W
U
Γ
K
X

Figure 109: High symmetry points and lines of the f.c.c. Brillouin zone.

The points on the lines between X and U contained on the square faces are
denoted by S while those between X and W are denoted by Z. The points on
the high-symmetry lines between L and W on the hexagonal faces are denoted
by Q. The points on the high-symmetry lines between Γ and K are denoted by
Σ, the points on the lines between Γ and X are denoted by ∆, and the points
on the line running through Γ and L are known as Λ.

5.3.4 The Hexagonal Brillouin Zone


The Brillouin zone for the hexagonal lattice is hexagonal. The upper and lower
faces are hexagons. The hexagonal face centers are at A = 2aπ (0, 0, 2ac ). The
vertices are at the H points, H = 2aπ ( √13 , 13 , 2ac ). The centers of the vertical
rectangular faces are denoted by M and M = 2aπ ( √13 , 0, 0). The mid-points
of the horizontal edges are denoted by L where L = 2aπ ( √13 , 0, 2ac ) and the
mid-points of the vertical edges are denoted by K where K = 2aπ ( √13 , 13 , 0).

The interior high-symmetry points is Γ the zone center. Points on the in-
terior high-symmetry lines are denoted as follows: Σ are points located on the
high-symmetry lines Γ M , ∆ are the points on the lines Γ A, and Λ are the
points on the lines between Γ and K.

178
The hexagonal Brillouin Zone

b3

b1 b2

Figure 110: The Brillouin zone for the hexagonal Bravais lattice.

The Hexagonal Brillouin Zone

L H

Γ
K
M

Figure 111: High symmetry points of the hexagonal Brillouin zone.

Points on the high-symmetry lines on the surface of the hexagonal Brillouin


zone are enumerated as: T the points which are located on the horizontal lines
K M , the points U which reside on the vertical line M L. The high-symmetry
lines on the hexagonal faces are: A L which contain points that are denoted by
R, and the line A H which contains points that are denoted by S.

179
5.3.5 The Trigonal Brillouin Zone
The trigonal unit cell has sides of length a, and the angles between any two
primitive lattice vectors are all equal and denoted by α. If the unit cell is
oriented such that the three primitive lattice vectors subtend the same angle
θ with the z axis, then the primitive lattice vectors can be expressed in the
Cartesian coordinate system as

a1 = a ( sin θ êx + cos θ êz )



1 3
a2 = a(− sin θ êx + sin θ êy + cos θ êz )
2 √2
1 3
a3 = a(− sin θ êx − sin θ êy + cos θ êz ) (312)
2 2
The relation between the angle θ and α is found from the scalar product as
1
cos α = − sin2 θ + cos2 θ (313)
2
which yields r
1 + 2 cos α
cos θ = (314)
3
Since the reciprocal lattice is also trigonal, the primitive reciprocal lattice vec-
tors are given by

b1 = b ( sin Θ êx + cos Θ êz )



1 3
b2 = b(− sin Θ êx + sin Θ êy + cos Θ êz )
2 √2
1 3
b3 = b(− sin Θ êx − sin Θ êy + cos Θ êz ) (315)
2 2
where Θ is related to β, the angle between any two primitive reciprocal lattice
vectors, by a relation analogous to that between θ and α
r
1 + 2 cos β
cos Θ = (316)
3
Since the length of the primitive reciprocal lattice vector is given by

2π a2
b = sin α (317)
Vc
and since
Vc2 = a6 ( 1 − 3 cos2 α + 2 cos3 α ) (318)

180
one finds that b is given by
s
2π 1 + cos α
b = (319)
a ( 1 − cos α ) ( 1 + 2 cos α )

Furthermore, since
( cos2 α − cos α )
cos β = (320)
sin2 α
after simplification, one finds that the angle β is given by
cos α
cos β = − (321)
1 + cos α
The reciprocal lattice vectors can be expressed as the sum of integer multiples
of the primitive reciprocal lattice vectors
 
1
Q = b m1 − ( m2 + m3 ) sin Θ êx
2

3
+b ( m2 − m3 ) sin Θ êy
2
+ b ( m1 + m2 + m3 ) cos Θ êz (322)

The types of planes which bound the Brillouin zone depend on the angle β in

Trigonal Brillouin Zone

(1,1,1)

(1,0,1)
(0,0,1)
(1,0,0)

(0,-1,0)

(-1,-1,0)
(0,-1,-1)

Figure 112: The Brillouin zone of the trigonal lattice.

a non-trivial way. That is, even the number of faces depends on the value of
β. For example, in Bi, one finds that α = 0.497 radians and β = 0.960

181
radians. The Brillouin zone of Bi has fourteen bounding planes. One set of
boundaries correspond to the six planes of the type (1, 0, 0) and (1, 0, 0), which
are located at a distance 2b from the origin. The other boundaries are formed
by six planes of the type (0, 1, 1)√and (0, 1, 1). These planes occur at a distance
from the origin equal to b √12 1 + cos β. The top and bottom boundaries
of the Brillouin zone are given by the planes (1, 1, 1) and (1, 1, 1), which are
√ √
located at a distance from the origin equal to b 23 1 + 2 cos β.

182
6 Electrons
The types of states of single electrons in the potentials produced by the crys-
talline lattice are discussed in the next three chapters. For simplicity, we shall
first implicitly assume that the effect of the Coulomb interactions between elec-
trons can be neglected. The neglect of electron - electron interactions is un-
justified, as can be seen by considering the electrical neutrality of solids. The
condition of electrical neutrality leads to the electron charge density being com-
parable with the charge density due to the lattice of nuclei or ions. Thus, the
strength of the interactions between the electrons is expected to be comparable
to the strength of the potential due to the nuclei. A simple order of magni-
tude estimate, based upon the typical linear dimensions of a unit cell a0 ∼ 2
2
Angstroms, leads to the average value of er ∼ 3 eV for both these interaction
energies. Nevertheless, as a discussion of the effect of pseudo-potentials reveals,
for most metals, the effect of the periodic potential of the lattice may be consid-
ered as small. The small value of the effective potential (or pseudo-potential)
leads to a useful approximation namely, that of nearly-free electrons. The effect
of the finite strength of electron-electron interactions is a more complex issue,
and is not yet fully understood. In principle, density functional theory provides
a method of evaluating the ground state electron density including the effect of
electron-electron interactions. However, the density functional method does not
describe the excited states. The effect of the electron-electron interactions is
that of disturbing the electron density around any excited electron. On assum-
ing that the interactions can be treated as a small perturbation, it can be shown
that most of the effects of electron-electron interactions on the low-energy ex-
cited electrons merely involve the dressing of the single excited electron thereby,
forming a quasi-particle excitation. That is, the effects of the excitation induced
modifications of the surrounding gas of electrons can be absorbed as renormal-
izations of the properties of the single-electron excitation. This feature can lead
to the low-temperature properties of the electronic system being determined by
the gas of quasi-particles, which has the same form as a non-interacting gas of
electrons. Systems where this simplification occurs are known as Landau Fermi
liquids. The effect of electron-electron interactions will be delayed to a later
chapter.

7 Electronic States
In describing electronic states in metals first, the nature of the many-electron
wave function and its decomposition into the sum of anti-symmetric products
of one-electron wave functions shall be described. Then, the general properties
of the one-electron basis wave functions shall be discussed. The one-electron
wave functions, or Bloch functions, are taken to be eigenfunctions of a suitable
non-interacting Hamiltonian in which the potential has the periodicity of the
underlying Bravais lattice.

183
7.1 Many-Electron Wave Functions
The energy of the electrons in a solid can be written as the sum of the kinetic
energies, the ionic potential energy acting on the individual electrons, and the
interaction potential between pairs of electrons. Thus, for a system with Ne
electrons, the Hamiltonian can be written as the sum
i=N
Xe h̄2 e2
 
1 X
Ĥ = − ∇2 + Vions (ri ) + (323)
i=1
2m i 2 | ri − rj |
i6=j

where ri denotes the position of the i-th electron, Vions is the potential due to
the lattice of ions, and the last term is the pair-wise interaction between the
electrons. This Hamiltonian can be separated into two sets of terms,

Ĥ = Ĥ0 + Ĥint (324)

where
i=N
Xe h̄2
 
Ĥ0 = − ∇2 + Vions (ri ) (325)
i=1
2m i
is the sum of one-body Hamiltonians acting on the individual electrons, and the
interaction term is given by the sum of two-body terms

1 X e2
Ĥint = (326)
2 | ri − rj |
i6=j

Since electrons are indistinguishable, the Hamiltonian must be symmetric un-


der all permutations of the indices i labelling the electrons. Also, the modulus
squared wave function must be invariant under all possible permutations of the
electron labels. An arbitrary permutation of the labels can be built up through
sequentially permuting pairs of labels.

The permutation operator P̂i,j is defined as the operator which interchanges


the indices i and j labelling a pair of otherwise indistinguishable particles. Thus,
if

Ψ(r1 , . . . ri , . . . rj , . . . rNe )

is an arbitrary Ne particle wave function, the permutation operator can be


defined as

P̂i,j Ψ(r1 , . . . ri , . . . rj , . . . , rNe ) = Ψ(r1 , . . . rj , . . . ri , . . . rNe )


(327)

184
Since the Hamiltonian is symmetric under interchange of the indices i and j
labelling any two identical particles, the permutation operators commute with
the Hamiltonian
[ P̂i,j , Ĥ ] = 0 (328)
Likewise, the permutation operators must also commute with any physical op-
erator Â
[ P̂i,j , Â ] = 0 (329)
otherwise measurements of the quantity  could lead to the particles being dis-
tinguished.

Since the Hamiltonian commutes with all the permutation operators, one can
find simultaneous eigenstates of the Hamiltonian Ĥ and all the permutation
operators P̂i,j . The energy eigenstates Ψ corresponding to physical states of
indistinguishable particles must satisfy the equations

Ĥ Ψ = E Ψ
P̂i,j Ψ = pi,j Ψ ∀ i, j (330)

where pi,j are the eigenvalues of the permutation operators P̂i,j . As permut-
ing the same pair of particle indices twice always reproduces the initial wave
function, one has
P̂i,j 2 = Iˆ (331)
where Iˆ is the identity operator. Thus, the eigenstates of the permutation
operators satisfy the two equations

P̂i,j 2 Ψ = p2i,j Ψ
= Ψ (332)

Hence, the eigenvalues of the permutation operators must satisfy

pi,j 2 = 1 (333)

or
pi,j = ± 1 (334)
Thus, the Ne particle wave functions have the property that, under any permu-
tation of a single pair of identical particles which are labelled by i and j, the
un-permuted and permuted wave functions are related by

Ψ(r1 , . . . rj , . . . ri , . . . rNe ) = ± Ψ(r1 , . . . ri , . . . rj , . . . rNe )


(335)

The upper sign holds for boson particles and the lower sign holds for fermions.
Also, since pi,j is a constant of motion, the nature of the particles does not
change with respect to time. Electrons are fermions and, thus, the wave func-
tion must always be anti-symmetric with respect to the interchange of any pair

185
of electron labels. Furthermore, the modulus square of the many-electron wave
function must be invariant under all possible permutations of the electron labels.

The energy eigenfunction for the system of Ne electrons can be written as

Ψ(r1 , r2 , . . . rNe ) (336)

The many-electron energy eigenstates Ψ usually cannot be found exactly. How-


ever, they can be expressed in terms of a superposition formed from a com-
plete set of many-electron eigenfunctions Φα1 ,α2 ,...,αNe (r1 , r2 , . . . , rNe ) of the
one-particle Hamiltonian Ĥ0 . The subscript αi represents the complete set of
quantum numbers (including spin) which completely describes the state of a
single fermion state.

Ĥ0 Φα1 ,α2 ,...αNe (r1 , r2 , . . . rNe ) = E0 Φα1 ,α2 ,...αNe (r1 , r2 , . . . rNe )
(337)

This many-electron eigenfunction is interpreted as representing the state in


which the Ne electrons are distributed in the set of single-electron states with
the specific quantum numbers α1 , α2 , . . . αNe . The basis states are orthonormal
and so satisfy the relations
Ne  Z
Y 
d3 r j Φ∗β1 ,β2 ,... (r1 , . . . , rNe ) Φα1 ,α2 ,... (r1 , . . . , rNe ) = δα1 ,β1 δα2 ,β2 . . .
j=1 V
(338)
where we have assumed that the sets of single-electron eigenvalues have been
enumerated and ordered. Since the basis is complete, the exact many-body
eigenstates of the full Hamiltonian Ĥ can be written as a linear superposition
of the complete set of basis functions
X
Ψ(r1 , r2 , . . . , rNe ) = Cα1 ,α2 ,...αNe Φα1 ,α2 ,...,αNe (r1 , r2 , . . . rNe )
α1 ,α2 ,...,αNe
(339)
where the sum over the set of {αi } runs over all possible distributions of the Ne
electrons in the set of all the single-electron states. The coefficients Cα1 ,α2 ,...,αNe
have to be determined. The coefficients represent the probability amplitudes
that electrons occupy the set of single-electron states labelled by α1 , α2 , . . . , αNe .

The set of many-electron basis functions Φα1 ,α2 ,...αNe (r1 , r2 , . . . rNe ) can be
expressed directly in terms of the one-electron wave functions φα (r). First,
note that the non-interacting Hamiltonian Ĥ0 can be decomposed as the sum
of Hamiltonians which only act on the individual electrons
i=N
Xe
Ĥ0 = Ĥi (340)
i=1

186
where the one-particle non-interacting Hamiltonian is given by

h̄2
Ĥi = − ∇2 + Vions (ri ) (341)
2m i
This one-particle Hamiltonian has eigenstates, φβ (ri ), which satisfy the eigen-
value equation
Ĥi φβ (ri ) = Eβ φβ (ri ) (342)
The many-particle non-interacting Hamiltonian Ĥ0 has eigenfunctions which
are the products of Ne one-particle eigenfunctions φβ (r)

χ(r1 , α1 ; r2 , α2 ; . . . rNe , αNe ) = φα1 (r1 ) φα2 (r2 ) . . . φαNe (rNe ) (343)

and the non-interacting energy eigenvalue E0 for the many-particle state is given
as the sum of the one-electron energy eigenvalues Eαi that are occupied by the
electrons
i=N
Xe
E0 = Eαi (344)
i=1

However, the wave functions χ(r1 , α1 ; r2 , α2 ; . . . rNe , αNe ) do not represent phys-
ical wave functions since each of the single-particle states with quantum numbers
α1 , α2 , . . . αNe are occupied by the respective electron labelled by r1 , r2 , . . . rNe
and, hence, the electrons have been labelled. As the electrons are indistinguish-
able, it is impermissible to distinguish them by this type of labelling. Thus,
physical wave functions should contain terms which are related by all the pos-
sible relabelling of the indices of the particles.

Electrons are fermions and, therefore, they have wave functions which are
anti-symmetric under the interchange of any pair of particles. The proper ba-
sis set of the many-electron wave function Φ must correspond to all possible
permutations of the single-particle indices. The proper anti-symmetrized wave
function Φα1 ,α2 ,...,αNe is given by the Slater determinant

φα1 (r1 ) φα1 (r2 ) . . . φα1 (ri ) . . . φα1 (rNe )

φα2 (r1 ) φα2 (r2 ) . . . φα2 (ri ) . . . φα2 (rNe )
.. .. .. ..


1
. . . .
Φα1 ...αNe = √N
φαi (r1 )
e! φαi (r2 ) . . . φαi (ri ) . . . φαi (rNe )
.. .. .. ..

. . . .

φα (r ) φα (r ) . . . φα (r ) . . . φα (r )
Ne 1 Ne 2 Ne i Ne Ne
1
The normalization is √N as there are Ne ! terms in the determinant, corre-
e!
sponding to the Ne ! permutations of the electron labels.

The anti-symmetric wave function has the property that if there are two or
more particles in the same one-particle eigenstate, say αi = αj , then the wave
function vanishes. This can be seen by noting that two rows of the determinant

187
are identical and, hence, the determinant vanishes. As the wave function van-
ishes if two or more electrons occupy the same single-particle eigenstate, there is
no state in which a one-particle eigenstate is occupied by two or more electrons.
The anti-symmetric nature of the fermion wave function directly leads to the
Pauli exclusion principle. The Pauli exclusion principle can be stated as “no
unique single-particle state can be occupied by two or more electrons.” For elec-
trons which have spin one-half, a single-particle state is uniquely specified only
if the spin quantum number is also specified. The single-particle wave function
φα (r) should be supplemented by the spinor χσ . That is, the single-electron
wave function should be replaced by the product

φα (r) → φα (r) χσ (345)

where χσ is a spinor or a normalized two-component column vector. The spin


index σ can be considered to label an eigenstate of a component of an arbitrary
single-electron spin operator, and the label σ should be considered as analogous
to the single-particle eigenvalue α. A complete set of labels for the single-
electron state are given by α and σ. An arbitrary spinor χσ can be decomposed
as the linear superposition of two basis spinors χ±
X
χσ = γ± χ± (346)
±

where the normalization condition is given by


X
| γ± |2 = 1 (347)
±

The two basis spinors χ± are usually denoted by the two component column
vectors  
1
χ+ = (348)
0
corresponding to an eigenstate of the Pauli matrix σz with spin-up and
 
0
χ− = (349)
1

corresponding to the eigenstate with spin-down. Thus, the arbitrary spin state
can be written as  
γ+
χσ = (350)
γ−
In this representation, the two components of an arbitrary spinor, χσ , represent
the internal degree of freedom of the spin and, thus, are analogous to the de-
gree of freedom represented by r in the position representation. The complex
conjugate wave function should be replaced by

φ∗α0 (r) → χTσ0 φ∗α0 (r) (351)

188
which contains χTσ0 which is the complex conjugated transpose of the spinor
states given by the two-dimensional row matrices
χTσ0 = 0 ∗ 0 ∗

γ+ γ− (352)
In the situations where the electron spin has to be explicitly considered, these
replacements lead to the inner product of two one-electron states not only in-
volving the integration of the product φ∗α0 (r) φα (r) over the electron’s position r
but also automatically involves evaluating the matrix elements of the individual
electron’s row spinor state χTσ0 with the column spinor state χσ .

The probability density ρ(r) for finding an electron at position r can be ob-
tained from the matrix elements of the many-electron wave function Ψ(r1 ; r2 , . . . rNe )
with the one-electron density operator ρ̂. The one-electron density operator is
given by a Dirac delta function
i=N
Xe
ρ̂(r) = δ 3 ( r − ri ) (353)
i=1

The density ρ(r) is evaluated as


Z Z Z
ρ(r) = Ne d3 r 1 d3 r 2 . . . d3 rNe δ( r − r1 ) | Ψ(r1 ; r2 , . . . rNe ) |2
V V V
(354)
Thus, the trace over the positions particles can be evaluated by integrating over
all but one of the particles positions
Z Z Z
ρ(r) = Ne d3 r 2 d3 r 3 . . . d3 rNe | Ψ(r; r2 , . . . rNe ) |2 (355)
V V V
The matrix elements of the spin states has also to be taken. The resulting elec-
tron density is normalized to Ne .

The probability density for finding an electron at position r and another


electron at r0 is a correlation function ρ(r, r0 ) which is given by the matrix
elements of the operator
X X
ρ̂(r, r0 ) = δ 3 ( r − ri ) δ 3 ( r0 − rj ) (356)
i j6=i

The resulting expression for the two-particle density is found by integrating over
the positions of all the electrons except two
Z Z Z
ρ2 (r, r0 ) = Ne ( Ne − 1 ) d3 r 3 d3 r 4 . . . d3 rNe | Ψ(r; r0 ; . . . rNe ) |2
V V V
(357)
This two-particle density correlation function is normalized to twice the number
of pairs of electrons, Ne ( Ne − 1 ).

——————————————————————————————————

189
7.1.1 Exercise 34
Evaluate the single-particle density and two-particle density correlation function
for a many-particle basis wave function Φα1 ,α2 ,...αNe given by a single Slater de-
terminant of single-particle wave functions φα (r).

——————————————————————————————————

The properties of the single-electron wave functions, φα (ri ), that are to be


used in forming the many-particle basis functions Φα1 ,α2 ...αNe (r1 , r2 , . . . rNe ) as
Slater determinants, are discussed in the next chapter. In the following, the
electron labels i in the one-electron wave functions are omitted.

7.2 Bloch’s Theorem


Bloch’s theorem describes the properties of the one-electron states φα (r) which
are eigenstates of the one-electron Hamiltonian with a periodic potential. An
electron in the solid experiences a periodic potential that has the periodicity
of the underlying lattice of ions. In particular, the potential is invariant under

Periodic potential of a set of ions

εF
Vions(r)

-4 -3 -2 -1 0 1 2 3 4

r/a

Figure 113: The periodic potential Vions (r) must satisfy the condition
Vions (r − R) = Vions (r), for all Bravais lattice vectors R.

190
translation through any Bravais lattice vector Ri

Vions (r − Ri ) = Vions (r) (358)

General properties of the solution of the Schrodinger equation for a single


electron in a solid can be found by considering the periodicity of Vions (r). If the
electron-electron interactions are neglected, the independent electrons obey the
one-particle Schrodinger equation with the periodic potential,

h̄2
 
Ĥ φα (r) = − ∇2 + Vions (r) φα (r) = Eα φα (r) (359)
2m
For an infinite solid, the physically acceptable solutions of this equation are
known as the Bloch wave functions. The energies of the Bloch states are usually
labelled by two quantum numbers n and k, instead of by α. The one-dimensional
case, where the values of k were restricted to real values, was investigated by
Kramers36 .

Bloch’s theorem applies to the eigenstates of the one-particle Hamiltonian,

h̄2
 
Ĥ = − ∇2 + Vions (r) (360)
2m
in which the potential has the symmetry

Vion (r − Ri ) = Vions (r) (361)

for all lattice vectors Ri in the Bravais lattice. Bloch theorem states that the
eigenfunctions can be found in the form
 
1
φn,k (r) = √ exp i k . r un,k (r) (362)
V
where the function un,k is invariant under the translation through any Bravais
lattice vector
un,k (r − Ri ) = un,k (r) (363)

Bloch’s theorem asserts that the periodic translational symmetry manifests


itself in the transformation of the wave function
 
φn,k (r − Ri ) = exp − i k . Ri φn,k (r) (364)

Thus, a translation of the wave function through a Bravais lattice vector only
shows up through the presence of an exponential factor. Furthermore, if the
36 H. A. Kramers, Physica 2, 483 (1935).

191
wave vector k is real, then the electron density for the Bloch state is identical
for each unit cell in the crystal. This prevents the wave function from diverging
at the boundaries of the solid.

The proof of Bloch’s theorem is based on the consideration of the translation


operator T̂R which, when acting on an arbitrary function f (r), has the effect of
translating it through a Bravais lattice vector R

T̂R f (r) = f (r − R) (365)

This translation operator can be applied to the wave function Ĥ φ(r) which

The Translation Operator TR


1
f(r) f(r-R) TR f(r) = f(r-R)
0.8

0.6

0.4

0.2

r0 r0+R
0
0 1 2 3 0 1 2 3
r r

Figure 114: The effect of the translation operator T̂R on an arbitrary function
f (r).

yields

T̂R Ĥ φ(r) = Ĥ(r − R) φ(r − R)


= Ĥ(r) φ(r − R)
= Ĥ T̂R φ(r) (366)

Thus, the Hamiltonian commutes with the translation operator which produces
a translation through a Bravais lattice vector,

[ Ĥ , T̂R ] = 0 (367)

192
This means that it is possible to find simultaneous eigenstates of both T̂R and Ĥ.

Furthermore, the translation operators corresponding to translations through


different lattice vectors commute. This can be shown by successive translations
T̂R and T̂R0 , which yields

T̂R T̂R0 φ(r) = φ(r − R0 − R)


φ(r − R − R0 ) = T̂R0 T̂R φ(r) (368)

Thus, the translation operators commute

[ T̂R , T̂R0 ] = 0 (369)

This proves that the wave functions can be chosen to be simultaneous eigenstates
of the Hamiltonian and all the translation operators that produce translations
through Bravais lattice vectors. The Bloch functions are chosen such that they
satisfy

Ĥ φ(r) = E φ(r)

T̂R φ(r) = c(R) φ(r) (370)

and, thus, are the simultaneous eigenstates of Ĥ and all the T̂R .

The translation operators can be compounded as

T̂R0 T̂R φ(r) = φ(r − R − R0 )


= T̂R+R0 φ(r) (371)

When two translation operators are successively applied to the simultaneous


eigenfunctions of the translation operators, it may be re-interpreted in terms of
the compound translation

T̂R0 T̂R φ(r) = c(R0 ) c(R) φ(r)


= T̂R+R0 φ(r) = c(R + R0 ) φ(r) (372)

This shows that the products of two eigenvalues of different translation operators
gives the eigenvalue of the compound translation

c(R0 ) c(R) = c(R + R0 ) (373)

Since a general Bravais lattice vector can be expressed as the sum

R = n1 a1 + n2 a2 + n3 a3 (374)

where (n1 , n2 , n3 ) are integers, a general eigenvalue can be decomposed in terms


of products
c(R) = c(a1 )n1 c(a2 )n2 c(a3 )n3 (375)

193
Hence, on introducing the notation
 
c(a1 ) = exp − i 2 π x1
 
c(a2 ) = exp − i 2 π x2
 
c(a3 ) = exp − i 2 π x3 (376)

for the primitive eigenvalues, one can define a vector k via


 
(1) (2) (3)
k = x1 b + x2 b + x3 b (377)

With these definitions, the eigenvalue of the translation operator can be ex-
pressed in terms of the k vector as
 
c(R) = exp − i k . R (378)

Thus, the eigenvalue equation for the translation operator is expressed as

T̂R φ(r) = φ(r − R)


= c(R) φ(r)
 
= exp − i k . R φ(r) (379)

which completes the proof of Bloch’s theorem.

The wave functions which are simultaneous eigenfunctions of the energy


and the periodic translation operators are the Bloch functions37 . The Bloch
functions, φn,k (r), are labelled by the translation quantum number k and a
quantum number n that pertains to the single-particle energy eigenvalue En,k .
It should be noted that Bloch’s theorem does not guarantee that the quantity
k is real. Since k is the quantum number associated with the eigenvalue of the
operator which translates through a Bravais lattice vector

T̂R φn,k (r) = φn,k (r − R)


 
= exp − i k . R φn,k (r) (380)

then it should be clear that as


 
exp iQ.R = 1 (381)

37 F. Bloch, Zeit. für Physik, 52, 555 (1928).

194
the eigenvalue labelled by k is identical to the eigenvalue labelled by k +Q. This
means that the two wave vectors can be identified, i.e., k + Q ≡ k. Thus, the
Bloch wave vector when translated through a reciprocal lattice vector Q leads
to an equivalent wave vector. Furthermore, if the convention

φn,k+Q (r) = φn,k (r) (382)

is adopted, then the eigenvalues must be related through

En,k+Q = En,k (383)

Thus, if k is real, any k value can be restricted to lie within one unit cell of
reciprocal space38 .

φk(r)

Vions(r)

Figure 115: The spatial variation of the real part of a Bloch function (schematic).

7.3 Boundary Conditions


Bloch’s theorem does not ensure that the wave vector k is real. In fact, for
surface states or impurity states, k may become imaginary. However, for bulk
states the wave vector is real, as can be ascertained by applying appropriate
boundary conditions.

Consider a crystalline solid of finite size which has the same shape as the
primitive unit cell of the Bravais lattice but with dimensions L1 = N1 | a1 |,
L2 = N2 | a2 | and L3 = N3 | a3 | along the three primitive axes. The solid
then contains a total number of N = N1 N2 N3 lattice points.

38 L. Brillouin, J. Phys. Radium, 1, 377 (1930).

195
N2 a 2
N1 a 1

N3 a 3
a3

a2 a
1

Figure 116: A hypothetical crystal of the same shape as the primitive unit cell
is constructed by stacking N1 × N2 × N3 unit cells together.

Born-von Karman or periodic boundary conditions39 are imposed on the


wave function

φn,k (r − Ni ai ) = φn,k (r) for i = 1 , 2 or 3 . (384)

The periodic boundary conditions ensure that the electronic states are homoge-
neous bulk states and are unmodified in the vicinity of the surface of the solid.
Therefore, application of Bloch’s theorem yields the condition

φn,k (r) = φn,k (r − Ni ai )


 
= exp − i Ni k . ai φn,k (r) for i = 1 , 2 or 3

(385)

Thus, the periodic boundary conditions are fulfilled if the wave vectors k satisfy
the conditions  
exp − i Ni k . ai = 1 (386)

Since k can be written in terms of the primitive reciprocal lattice vectors, b(i) ,
via
i=3
X
k = xi b(i) (387)
i=1
39 M.Born and Th. von Karman, Zeit. für Physik, 13, 297 (1912), M. Born and Th. von
Karman, Zeit. für Physik, 14, 15 (1913).

196
and as ai . b(j) = 2 π δij , then the periodic boundary conditions require that
 
exp − i 2 π Ni xi = 1 for i = 1 , 2 or 3

(388)

Thus, the components xi must be in the form of ratios


mi
xi = (389)
Ni
where mi are integers. This proves that the general Bloch wave vector k is a
real vector, and the k vectors have the general form
i=3
X mi (i)
k = b (390)
i=1
Ni

Since Ni  1, the k vectors form a dense set of points in reciprocal space.

The properties of a solid can be expressed in terms of summations over the


electronic states. Since each state can be expressed in terms of the discrete
k quantum number, the summation are over a dense set of k vectors. The
summation over a dense set of k vectors can be represented in terms of an
integral over the energy, weighted by the density of states. From the form of k,
the volume of k space per allowed k value is

b(1)
 (2)
b(3)

b
∆3 k = . ∧
N1 N2 N3
 
1 (1)
= b . b(2) ∧ b(3) (391)
N
As the volume of the Brillouin zone is given by
 
b(1) . b(2) ∧ b(3) (392)

the volume of one state is N1 times the volume of the Brillouin zone. This implies
that the number of allowed k values within the Brillouin zone is equal to the
number of unit cells in the crystal. The volume ∆3 k associated with a Bloch
state is given by
1 ( 2 π )3
∆3 k =
N a1 . ( a2 ∧ a3 )
1 ( 2 π )3
= (393)
N Vc
Now, since the volume of the solid V is N times the volume of the cell Vc ,

V = N Vc (394)

197
b2
b1

b3/N3
∆k
3 b3
b2/N2
b1/N1

Figure 117: The volume of k-space associated with a one-electron state, ∆3 k, is


a factor of 1/N smaller than the volume of a primitive unit cell of the reciprocal
lattice.

then the volume of k space associated with each Bloch state is

( 2 π )3
∆3 k = (395)
V
Hence, in the continuum limit, the number of one-electron states (per spin) in
an infinitesimal volume of phase d3 k is given by
V
d3 k (396)
( 2 π )3

7.4 Plane Wave Expansion of Bloch Functions


Any function obeying Born-von Karman boundary conditions can be expanded
as a Fourier series. This implies that the Bloch functions can also be expanded
as  
X 1
φn,k (r) = Cq √ exp i q . r (397)
q V

where the wave vectors q are to be related to k. From Bloch’s theorem, the
Bloch functions can also expressed as
 
1
φn,k (r) = √ exp i k . r un,k (r) (398)
V

198
Since un,k (r) has periodic translational invariance, it only contains reciprocal
lattice vectors Q. The Fourier series expansion of the periodic function is
X  
un,k (r) = un,k (Q) exp i Q . r (399)
Q

and the inverse transform is given by the integral


Z  
1
un,k (Q) = d3 r un,k (r) exp − i Q . r (400)
V V

On comparing the above two forms for the Bloch functions, one has
 
X 1
φn,k (r) = Cq √ exp i q . r
q V
 
X 1
= un,k (Q) √ exp i ( k + Q ) . r
Q
V
(401)

Thus, the allowed q values in the Bloch wave functions are equal to k, modulo a
reciprocal lattice vector. Furthermore, the Cq are equal to the Fourier compo-
nents un,k (Q). Next, it shall be shown how the Cq can be determined directly
from the Schrodinger equation which contains the periodic potential Vions (r).

The Bloch functions can be found by solving the Schrodinger equation where
the Hamiltonian contains the periodic potential Vions (r). The periodic potential
also has a Fourier series expansion
X  
Vions (r) = Vions (Q) exp i Q . r (402)
Q

and the inverse transform is given by the integral


Z  
1
Vions (Q) = d3 r Vions (r) exp − i Q . r (403)
V V

Furthermore, since Vions (r) is real, the Fourier transform of the potential has
the symmetry

Vions (−Q) = Vions (Q) (404)
This follows from taking the complex conjugate of the Fourier series expansion
of Vions (r). A second condition on the Fourier expansion coefficients exists
for crystals which have an inversion symmetry around a suitable origin. The
inversion symmetry implies that the potential is symmetric

Vions (r) = Vions (−r) (405)

199
0
(2,2,2)(4,0,0) (4,2,0)
(2,2,0)
(3,3,1)
(2,0,0) (3,1,1)
-0.2

Vions(Q) a/4πZe2
(1,1,1)
-0.4

-0.6

-0.8

-1
0 1 2 3 4 5 6

(Qa/2π)

Figure 118: The first few Fourier components Vions (Q) for a model potential of
an f.c.c. solid.

and this implies that the Fourier transform of the potential has the property

Vions (Q) = Vions (−Q) = Vions (Q) (406)

The expansion coefficients Cq in the Bloch function are found by substituting


the Fourier series into the energy eigenvalue equation. The kinetic energy term
is evaluated from
p̂2 h̄2
φn,k (r) = − ∇2 φn,k (r)
2m 2m
X h̄2 q 2 1
 
= Cq √ exp i q . r (407)
q
2m V

The potential term in the energy eigenvalue equation has the form of a convo-
lution when expressed in terms of the Fourier Transforms
   
X X
0 1 0 0
Vions (r) φn,k (r) = Vions (Q ) Cq √ exp i q + Q
0 . r (408)
q0 Q0
V

The form of the energy eigenvalue equation is simplified if q 0 is expressed as


q 0 = q − Q0 , so
 
X X
0 1
Vions (r) φn,k (r) = Vions (Q ) Cq−Q0 √ exp iq.r (409)
q Q0
V

200
Then, the energy eigenvalue equation takes the form
 2 2 !
h̄ q
X  X  
0
− E Cq + Vions (Q ) Cq−Q0 exp i q . r = 0 (410)
q
2m 0Q

The wave vectors q are expressed as q = k − Q so that k is always located


within the first Brillouin zone. On equating the coefficients of the plane waves
with zero, one finds the matrix eigenvalue equation
 2
h̄ ( k − Q )2
 X
− E Ck−Q + Vions (Q0 ) Ck−Q−Q0 = 0 (411)
2m 0 Q

The reciprocal lattice vector is transformed as Q0 → Q” = Q0 + Q in the


second term, leading to an infinite set of coupled equations
 2
h̄ ( k − Q )2
 X
− E Ck−Q + Vions (Q” − Q) Ck−Q” = 0
2m
Q”

(412)

Thus, because of the periodicity of the potential, the Bloch functions only con-
tain Fourier components q that are connected to k via reciprocal lattice vectors.
For fixed k, the set of equations couple Ck to all the Ck−Q via the Fourier
component of the potential Vions (Q). In principle, the set of infinite coupled
algebraic equations (412) could be used to find the coefficients Ck−Q and the
eigenvalue En,k . The Bloch function is expressed in terms of the coefficients
Ck−Q as
 
X 1
φn,k (r) = Ck−Q √ exp i ( k − Q ) . r
Q
V
  X  
1
= √ exp + i k . r Ck−Q exp − i Q . r
V Q

(413)

Using this, the Bloch function can be expressed in terms of the periodic function
un,k (r) via  
X
un,k (r) = Ck−Q exp − i Q . r (414)
Q

In order to make this approach tractable, it is necessary to truncate the infinite


set of coupled equations (412) to a finite set. However, if this set of equations are
truncated, it would require approximately 103 to 106 plane wave components
before convergence is attained in three dimensions. Therefore, other methods
are frequently used.

201
7.5 The Bloch Wave Vector
The Bloch wave vector k plays a role similar to that of the momentum of a
free electron. In fact, it reduces to the momentum quantum number in the
limit Vions (r) → 0. However, for a non-zero crystal potential, k is not equal
to the eigenvalue of the electron momentum p̂ = − i h̄ ∇ since it differs
by amounts that are determined by the reciprocal lattice vectors Q and the
coefficients Ck+Q . That is,
 
i h̄
p̂ φn,k (r) = h̄ k φn,k (r) − √ exp ik.r ∇ un,k (r) (415)
V
Thus, h̄ k is known as the crystal momentum.

The crystal momentum can always be chosen to be in the first Brillouin zone
by making the transformation

k = k0 + Q (416)

On substituting this relation into the expression for the Bloch function, one
finds that it can be re-written as
 
1
φn,k (r) = √ exp i k . r un,k (r)
V
    
1
= √ exp i k 0 . r exp i Q . r un,k (r)
V
 
1
= √ exp i k 0 . r ũn,k (r) (417)
V
where  
ũn,k (r) = exp iQ.r un,k (r) (418)

is identified as a periodic function of the type that is used in Bloch’s theorem.


The new function ũn,k (r) transforms like un,k (r) as it has the periodicity of the
Bravais lattice since  
exp iQ.R = 1 (419)

Due to the periodic translational symmetry, the eigenvalue problem can be


reduced to finding a solution for the periodic function un,k (r) in a single cell of
the lattice. The total number of energy eigenfunctions must correspond to the
sum of the numbers of electron states originating from each atom in the crystal,
and there may be many basis atoms in the unit cell. As an isolated atom is
expected to have an infinite number of excited levels, and as the number of
different k points in the Brillouin zone is equal to the number of primitive cells
in the crystal, there must be infinitely many energy eigenfunctions with fixed k.

202
The different one-electron states with fixed k are distinguished by the index n.
The energy En,k is a continuous function of k, forming energy bands. This is
seen by examination of the eigenvalue equation, when the Bloch functions are
expressed as  
1
φn,k (r) = √ exp i k . r un,k (r) (420)
V
This procedure leads to the energy eigenvalue equation
 2  2 

Ĥk un,k (r) = − i∇ + k + Vions (r) un,k (r)
2m
= En,k un,k (r) (421)

Due to the Born-von Karman boundary conditions, each energy band in the
Brillouin zone contains N different states. The different k values are not part
of a continuum but form a discrete dense set of points. The energy eigenvalues
En,k , therefore, although a continuous function of k, only exist at a finite set of
points.

7.6 The Density of States


A physical quantity A may be expressed in terms of the quantities An,k associ-
ated with the individual electrons in each of the occupied Bloch states (n, k) in
the solid. That is, the quantity A is given by
X
A = 2 An,k (422)
n,k

where the sum runs over each level (n, k) that is occupied by an electron. The
factor 2 originates from the spin degeneracy. Since the different k states are
dense and uniformly distributed in the Brillouin zone, the summation may be
represented by an integration. The volume ∆3 k of phase space associated with
a Bloch state is given by
( 2 π )3
∆3 k = (423)
V
The quantity A is expressed as the integral
V X Z
A = 2 d3 k An,k (424)
( 2 π )3 n En,k <EF

where the integration over k runs over the volume of occupied states in the first
Brillouin zone. Thus, for the partially filled bands the integration runs over
a volume of k space enclosed by a surface of constant energy F , and for the
completely filled bands it runs over the entire Brillouin zone.

The integration over k space may be converted into an integral over the
energy , by introducing the one-electron density of states ρ(). The density

203
of states per spin is defined by the sum over Dirac delta functions for each
one-electron state
X
ρ() = δ(  − En,k )
n,k
X Z d3 k
= V δ(  − En,k ) (425)
n
( 2 π )3

If the quantity An,k only depends on (n, k) through En,k , then

An,k = A(En,k ) (426)

so the quantity A can be represented as an integral over the density of states


Z F
A = 2 d ρ() A() (427)
−∞

The density of states ρ() can be calculated by noting that the infinitesimal
R +∆
integral  ρ() d ≈ ρ() ∆ is the number of states in the energy range
between  and  + ∆, or the allowed number of k values between  and  + ∆
in each of the energy bands. Thus, on integrating over an energy range ∆ and
using the definition of the density of states in terms of the Dirac delta function,
one finds
X Z +∆ Z
d3 k
ρ() ∆ ≈ V d δ(  − En,k )
n  ( 2 π )3
X Z  
V 3
= d k Θ( + ∆ − En,k ) − Θ( − En,k )
( 2 π )3 n
(428)

where Θ(x) is the Heaviside step function. Thus, the density of states is ex-
pressed by an integral over a volume of k space enclosed by surfaces of constant
energy  and  + ∆. Furthermore, since ∆ is an infinitesimal quantity, ∆ can
be expressed in terms of the perpendicular distance between the two surfaces of
constant energy.

Let Sn () be the surface En,k =  lying within the primitive cell and let
δk(k) be the perpendicular distance between the surfaces Sn () and Sn ( + ∆)
at point k. Then, as Sn () is a surface of constant  and ∇ En,k is perpendicular
to that surface

 + ∆ ≈  + | ∇ En,k | δk(k)

∆
δk(k) ≈ (429)
| ∇ En,k |

204
k3

Sn(ε+∆ε)
Sn(ε)

d2Sn
δk

Figure 119: The density of states is given in terms of a weighted integration


over a surface in k-space of constant energy, Sn ().

Hence, the density of states can be expressed as an integral over a surface of


constant energy
X Z d2 S 1
ρ() ≈ V 3
(430)
n Sn () ( 2 π ) | ∇ En,k |

This gives an explicit relation between the density of states and the band struc-
ture.

Since En,k is periodic, it is bounded from above and below for each value of
n. This implies that there will be values of k in each Brillouin zone where the
group velocity vanishes,
∇ En,k = 0 (431)
As the band energy En,k is a periodic function of three variables it must have
at least one maximum and one minimum in the Brillouin zone and six saddle
points. At each of these k points, the integrand in the expression for ρ() di-
verges. Other divergences may be expected which originates from k points near
the Brillouin zone boundary, where the dispersion relation is expected to have
zero slope. These divergences give rise to van Hove singularities in the density of
states. L. van Hove provided a general discussion of these types of singularities
using the Morse index theorem40 .

In three dimensions these singularities are integrable. That is, the integra-
tion over the surface area yield a finite value for ρ(). In the three-dimensional
case the divergences show up in the slopes of the density of states ∂ρ() ∂ , and
are the van Hove singularities. The van Hove singularities at the density of
40 L. van Hove, Phys. Rev. 89, 1189 (1953), also see the discussion by H. P. Rosenstock,

Phys. Rev. 97, 290 (1955).

205
Density of states ρ(ε)
Density of States ρ(ε)

Density of states ρ(ε)


Density of states ρ(ε)

ε/ε0 ε/ε0
ε/ε0 ε/ε0

Figure 120: The types of van Hove singularities in the one-electron density of
states of a three-dimensional solid.

states occur at the values of E where ∇ En,k vanishes at some points of the
surface Sn (). Typical van p
Hove singularities occur at the band edges where the
density of states varies as |  | . Although the density of states ρ() at van
Hove singularities does not diverge in three dimensions, the derivatives diverge
and can give rise to anomalies in thermodynamics as can be seen by examining
the Sommerfeld expansion.

In low-dimensional systems, the divergence can show up directly as a diver-


gence in the density of states.

——————————————————————————————————

7.6.1 Exercise 35
The energy dispersion relation at a van Hove singularity has a zero gradient. In
the vicinity of the van Hove singularity, the d-dimensional dispersion relation
can be written as
i=d
X
E k = E0 + E1 αi ki2 a2i (432)
i=1

where the coefficients αi determine whether the extremum is a maximum, min-


imum or saddle point. The coefficients are given by

αi = ± 1 (433)

Characterize the different types of van Hove singularities41 in the density of


states and sketch the energy dependence in the vicinity of the singularity for
41 One theorem states that a periodic function of d variables must have a number of at least
 
d
van Hove singularities with n negative coefficients αi in the d-dimensional primitive
n
 
d
unit cell. The expression is the binomial coefficient
n
d!
 
d
=
n ( d − n )! n!
where d > n.

206
d = 1, 2 and d = 3.

——————————————————————————————————

7.7 The Fermi Surface


The ground state of the electronic system has the lowest possible energy. For
non-interacting electrons, the electrons occupy the lowest possible eigenvalues.
However, the distribution of electrons must satisfy the restriction imposed by
the Pauli exclusion principle, which states that no uniquely specified electron
state can be occupied by more than one electron. This means that a spin de-
generate state cannot be occupied by more than two electrons, one for each
spin value. Thus, the ground state of Ĥ0 is represented by a Slater determinant
wave function in which two electrons are placed in the lowest energy eigenstate,
and, successively, two more are placed in the next lowest states, until all the Ne
electrons have been placed in states. In the following, the convention is adopted
that the electrons which are associated with the states (n, k) have k restricted
to be within the first Brillouin zone.

On distributing all the electrons into the lowest one-electron energy eigen-
states in a manner consistent with the Pauli exclusion principle, one finds two
different types of ground states:

(i) Insulators.

In insulators, a number of bands are completely filled and all other bands
are completely empty. No band is partially filled. In this case, there must exist

1.2

1
µ
Ek [ Rydbergs ]

0.8

0.6

0.4

0.2

-0.2
Γ X W L Γ K X

Figure 121: The calculated electronic dispersion relation for Si. The chemical
potential µ does not cross the energy bands.

an energy interval which separates the lowest unoccupied band state and the

207
highest occupied band state. The density of states must be zero in this energy
interval. The width of the interval, in which ρ() = 0, is the threshold energy
required to excite an electron from an occupied to an unoccupied state. This
energy interval is defined to be the band gap. In an insulator, the chemical po-

Figure 122: The calculated electronic density of states for Si. The chemical
potential µ occurs in the gap in the density of states. [After Chelikowski et al.
(1973).]

tential µ falls in the band gap. An insulating state can only occur if the number
of electrons Ne is equal to an even number times the number of primitive unit
cells N in the direct lattice. This is because each band can be occupied by 2 N
electrons. For example, C being tetravalent when it crystallizes in the diamond
structure is insulating, and has a band gap of over 5 eV. The elements Si and
Ge are also insulating, but have smaller band gaps which are 1.1 eV and 0.67
eV, respectively.

(ii) Metals.

A number of bands may be partially filled. In this case, the highest occu-
pied Bloch states have an energy F which lies within the range of one or more
bands. This case corresponds to a metal, in which the one-electron density of
states at F is non-zero, ρ(F ) 6= 0. Systems with an odd number of electrons
per unit cell should be metallic, such as the simple mono-valent metals like N a
or K. However, systems with two electrons per unit cell can be metallic. For
example, divalent M g is metallic. M g crystallizes in the hexagonal close-packed
system and, hence, has four electrons per unit cell. The small distance between
the atoms is responsible for the large dispersion of the bands which allows the
bands to overlap. The overlapping of the bands leads to divalent M g being
metallic.

For each partially filled band, there will be a surface in the three-dimensional

208
Electron Bands for Al
1.25

1
µ

Ek [ Rydbergs ]
0.75

0.5

0.25

0
f.c.c.

-0.25
Γ Χ W L Γ K W Χ

Figure 123: The electronic dispersion relations for metallic Al. The chemical
potential µ or Fermi energy F cuts across the bands.

Density of States for b.c.c. Li

0.75
ρ(ε) [ States /eV ]

0.5

0.25

εF

0
0 1 2 3 4

[After Ham (1962)] ε [ eV ]

Figure 124: The electronic density of states ρ() for metallic Li. The chemical
potential µ or Fermi energy F occurs in the energy region where the density of
states is finite.

k space which separates the occupied from the unoccupied states. The set of all
such surfaces forms the Fermi surface. The Fermi surface is determined by the
equation
En,k = F (434)
Since En,k is periodic in the reciprocal lattice, the Fermi surface may either be
represented within the full periodic reciprocal lattice or in a single unit cell of
the reciprocal lattice. The Fermi surface is represented in the extended zone
scheme, if the full reciprocal lattice is used. If the Fermi surface is represented
within a single primitive unit cell of the reciprocal lattice, it is represented in a
reduced zone scheme.

209
210
8 Approximate Models
Some of the earlier approaches to electronic structure of solids will be discussed
in this chapter. These methods are not in common use, and are not reliable
methods for calculating electronic structures. These older methods also ne-
glect the effect of electron-electron interactions. By contrast, the most common
method in use today is based on the Density Functional approach of Kohn and
Sham, which is quantitatively reliable and includes the effect of electron-electron
interactions. Nevertheless, the older methods were important in the develop-
ment of the subject and yield important insights into the results of electronic
structure calculations. We shall focus our attention to two different models
which are suitable for different regimes of the energy spectrum. First, we shall
examine the nearly-free electron model that treats continuum states E > 0
and in which the potential of the lattice of ions is introduced as a small pertur-
bation. Then, we shall examine the tight-binding model, which describes how
atomic bound states with E < 0 are perturbed when they are brought together
to form a crystal.

8.1 The Nearly-Free Electron Model


In the nearly-free electron approach to electronic structure calculations, one as-
sumes that the periodic potential due to the lattice is small. This assumption is
not justified, apriori, as the potential is of the order of 10 eV. However, the effect
of the potential can be much smaller than this estimate, and for these cases, the
nearly-free electron model gives results which can be used to phenomenologically
describe metals found in groups I, II, III, and IV of the periodic table. These
materials have an atomic structure which consists of s or p electrons outside a
closed shell configuration.

The nearly-free electron model provides a good description of conduction


electron states due to two principal reasons:-

(i) The region in which the electron - ion interaction is strongest is in the
vicinity of the ion. However, since this region is occupied by the core electrons
and the Pauli principle forbids the conduction electrons to enter this region, the
effective potential is weak.

(ii) In the region of space where the conduction electrons reside, the motion
of other conduction electrons effectively screen the potential.

Since in the nearly-free electron approximation the effective potential is as-


sumed to be small, perturbation theory may be used.

211
8.1.1 Perturbation Theory
The wave function for an electron in a Bloch state with wave vector k is given
by  
X 1
φk (r) = Ck−Q √ exp i ( k − Q ) . r (435)
Q
V

where Q are reciprocal lattice vectors and the coefficients Ck have to be deter-
mined. The coefficients satisfy the set of coupled algebraic equations
 2 
h̄ X
2
( k − Q ) − E Ck−Q + Vions (Q − Q0 ) Ck−Q0 = 0 (436)
2m 0 Q

where the sum runs over all the reciprocal lattice vectors Q0 . For fixed k, there
is an equation for each Q value. The different solutions of this set of equations
for fixed k are labelled by the index n.

If one neglects the potential due to the lattice, one obtains the empty lattice
approximation. This is the result of the zero-th order perturbation theory. To
zero-th order in the perturbing potential Vions , the set of equations reduce to
 
(0)
Ek − Q − E Ck−Q = 0 (437)

where the zero-th order energy eigenvalues are given by

(0) h̄2
Ek − Q = ( k − Q )2 (438)
2m
and the zero-th order energy eigenfunctions are
 
(0) 1
φk (r) = √ exp i ( k − Q ) . r (439)
V
If, for a given k, the energies associated with the set of reciprocal lattice vectors
Q1 , . . . , Qm are degenerate,

(0) (0) (0)


Ek − Q = Ek − Q = . . . = Ek − Q (440)
1 2 m

then the zero-th order approximation for φk (r) can be made of any linear su-
(0)
perposition of the functions φk (r) = √1V exp[ i ( k − Q ) . r ].

The type of perturbation theory that is appropriate depends on whether the


zero-th order eigenvalues are degenerate or not.

212
8.1.2 Non-Degenerate Perturbation Theory
Non-degenerate perturbation theory can be used when the energy separations
(0)
between the level under consideration, Ek − Q , and all other zero-th order
1
eigenvalues are large compared with the magnitude of the potential
(0) (0)
| Ek − Q − Ek − Q |  | Vions (Q1 − Q) | (441)
1

for fixed k and all Q 6= Q1 . This corresponds to the non-degenerate case.

We shall evaluate the one-electron energy eigenvalue to second order in Vions


but first, we need to consider the first order correction to the energy and wave
function of the state under consideration which, to zero-th order, has momentum
k − Q1 . The amplitude corresponding to the plane wave component with this
momentum satisfies the secular equation
 
(0)
X
Ek − Q − E Ck−Q + Vions (Q1 − Q0 ) Ck−Q0 = 0 (442)
1 1
Q0

This shall be used to obtain the energy E and the coefficient Ck−Q to first
1
order in Vions . The term involving the summation is explicitly of the order of
Vions , so the coefficients Ck−Q in this term only need to be calculated to zero-th
order in the Vions . Only one coefficient is non-zero to zero-th order in Vions ,
since
(0)
Ck−Q = 0 ∀ Q 6= Q1 (443)

Thus, to first order in Vions , only one term survives in the summation and the
coefficient Ck−Q satisfies the eigenvalue equation
1

 
(0) (0)
E − Ek − Q Ck−Q = Vions (0) Ck−Q (444)
1 1 1

This equation determines the energy eigenvalue E (1) to first order in Vions . Since
the energy shift is to be calculated to first order in Vions , the coefficient Ck−Q
1
(0)
can be substituted by its zero-th order value Ck−Q . This procedure yields the
1
first order approximation for the energy eigenvalue
(0)
E (1) = Ek−Q + Vions (0) (445)
1

First order perturbation theory only produces a constant shift in the zero-th
order energy eigenvalues which can be absorbed into the definition of the refer-
ence energy. It is also seen from eqn(444) that, to first order, the change in the
coefficient Ck−Q remains undetermined, so we may set
1

(1) (0)
Ck−Q = Ck−Q (446)
1 1

213
(0)
This is seen by substituting the first-order expression for E − Ek−Q into
1
eqn(444). In the following discussion, we shall neglect the effect of the average
potential Vions (0).

The coefficients of the other plane wave components of the Bloch function,
Ck−Q , satisfy
 
(0)
X
Ek − Q − E Ck−Q + Vions (Q − Q0 ) Ck−Q0 = 0 (447)
Q0

(1)
This is used to obtain the coefficients Ck−Q to first order in Vions . Since the
summand is explicitly of first order in Vions , then the coefficients Ck−Q0 need
(0)
only be considered to zero-th order. However, only Ck−Q is non-zero in this
1
order so,
(1) Vions (Q − Q1 ) (0)
Ck−Q = (0)
Ck−Q (448)
E − Ek − Q 1

(1) (0)
The coefficients Ck−Q and Ck−Q completely determine the energy eigenfunc-
1
tion to first order in Vions .

The energy eigenvalue can now be found to second order in Vions using
the wave function that have just been calculated to first order in Vions . On
(1)
substituting the expression for Ck−Q , eqn(448), into the secular equation which
determines Ck−Q , eqn(442), one finds
1


(0)
 X | Vions (Q − Q) |2 (0)
1
E − Ek − Q Ck−Q = (0)
Ck−Q (449)
1 1
Q (E − Ek − Q ) 1

Since both the energy and wave function are unchanged to first order in Vions ,
the lowest order non-zero contribution to the term on the left hand side is found
when Ck−Q is evaluated in zero-th order and E is evaluated to second order.
1
Thus, to second order in Vions , the energy eigenvalue E is given by the solution
of 
(0)
 X | Vions (Q − Q) |2
1
E − Ek − Q = (0)
(450)
1
Q ( E − Ek − Q )
(0)
or, since the eigenvalue E is approximately equal to Ek − Q , the energy eigen-
1
value is given by

(0)
X | Vions (Q1 − Q) |2
E = Ek − Q + (0) (0)
(451)
1
Q ( Ek − Q − Ek − Q )
1

214
This relation shows that weakly perturbed non-degenerate bands repel each
other. For example, if
(0) (0)
E k − Q > Ek − Q (452)
1

then the second order contribution is negative and E is reduced further below
(0)
Ek − Q . On the other hand, if
1

(0) (0)
Ek − Q < Ek − Q (453)
1

then the second-order contribution is positive and E is increased further above


(0)
Ek − Q . Hence, the leading-order effect of the perturbation increases the sep-
1
aration between the energy bands.

8.1.3 Degenerate Perturbation Theory


The most important effect of the potential occurs when a pair of the free electron
eigenvalues are within Vions of each other, but are far from all other eigenvalues.
Under these conditions, the eigenvalues are almost doubly-degenerate and one
can use degenerate perturbation theory to couple these energy levels.

In this case, the set of equations can be truncated to only two non-zero
complex coefficients Ck−Q and Ck−Q . These two coefficients satisfy the pair
1 2
of equations
(0)
( E − Ek − Q ) Ck−Q = Vions (Q2 − Q1 ) Ck−Q (454)
1 1 2

and
(0)
( E − Ek − Q ) Ck−Q = Vions (Q1 − Q2 ) Ck−Q (455)
2 2 1

which can be combined to yield the quadratic equation for E


(0) (0)
( E − Ek − Q ) ( E − Ek − Q ) = | Vions (Q1 − Q2 ) |2 (456)
1 2

The two energy eigenvalues are given by the solutions of the quadratic equation
v
 E (0) (0)  u u E (0) (0)
+ E k − Q − Ek − Q
2
k − Q k − Q
1 2
± 1 2
+ | Vions (Q1 − Q2 ) |2
t
E =
2 2
(457)
Whenever the Bloch wave vector k takes on special values such that unperturbed
bands cross
(0) (0)
Ek − Q = E k − Q (458)
1 2

the energy bands simplify to yield the two energies


(0)
E = Ek − Q ± | Vions (Q1 − Q2 ) | (459)
1

215
Nearly Free Electron Dispersion Relation
3

Ek/E0
1 2 Vions(1,0,0)

-1
0 0.2 0.4 0.6 0.8 1

(kxa/π)

Figure 125: The electronic dispersion relations calculated in the nearly-free


electron approximation. The crystal potential may lift the degeneracy of the
free electron bands at their crossing points.

If the unperturbed bands cross, the non-zero potential produces a splitting of


2 | Vions (Q1 − Q2 ) |. This result is consistent with that previously found by
using non-degenerate perturbation theory.

The avoided crossings of the bands are expected to occur whenever


(0) (0)
Ek − Q ∼ Ek − Q (460)
1 2

This gives rise to a specific condition on the wave vectors. For convenience of
notation, let q = k − Q1 so that this criterion takes the form

(0)
Eq(0) = Eq − Q” (461)

6 0. This requires that vector q lies on


for some reciprocal lattice vector Q” =
the Bragg plane bisecting Q”, as this condition reduces to

Q”2 = 2 q . Q” (462)

The vector q − Q” lies on a second Bragg plane. Thus, the geometric signifi-
cance of the condition for the degeneracy of the unperturbed bands, is that the
electronic states satisfy the condition for Bragg scattering.

The origin of the gaps can be easily understood from consideration of the
wave functions. When q lies on a single Bragg plane, then the energy eigenvalues
are simply given by
E = Eq(0) ± | Vions (Q”) | (463)
The coefficients corresponding to these energies are found from the two coupled
equations. In this case, where the unperturbed bands cross, the coefficients are

216
q Q-q

O Q
Q

Figure 126: If the vector q lies on a Bragg plane, the vector q − Q also lies on
a Bragg plane.

related via  
Cq = ± sign Vions (Q”) Cq−Q” (464)

which produces two standing wave solutions. If Vions (Q”) > 0, then the pair
of standing wave states are the anti-bonding state
Q” . r
 
+ 2 2 2
| φq (r) | ∼ cos
V 2

E+ = Eq(0) + | Vions (Q”) | (465)

and the bonding state


Q” . r
 
2
| φ−
q (r) |
2
∼ sin2
V 2

E− = Eq(0) − | Vions (Q”) | (466)

On the other hand, if Vions (Q”) < 0, then the situation is reversed, and the
anti-bonding state is given by
Q” . r
 
+ 2 2 2
| φq (r) | ∼ sin
V 2

E+ = Eq(0) + | Vions (Q”) | (467)

217
while the bonding state is given by the other form
Q” . r
 
2
| φ−
q (r) |2
∼ cos 2
V 2

E− = Eq(0) − | Vions (Q”) | (468)

In this context, the wave function


r
Q” . r
 
2
φpq (r) ∼ sin (469)
V 2
is called p-like as it vanishes at the lattice points, whereas
r
Q” . r
 
s 2
φq (r) ∼ cos (470)
V 2
is called s-like as it is non-vanishing at the positions of the ions, r = R. The
origin of the gap between the two branches is seen through examination of the
average potential energy of the s and p like wave functions
Z
d3 r Vions (r) | φs,p
q (r) |
2
(471)
V

The s-like electrons congregate at the position of the ions where the potential is
lower, and the p-like electrons congregate between the ions where the potential
is higher. For an attractive interaction Vions (r) < 0, this leads to φsq (r) having
a lower energy than φpq (r), (since Vions (Q”) < 0).

The opening of a gap in the electronic dispersion relation at the Bragg plane
is manifested by structure in the one-electron density of states which is associ-
ated with the van Hove singularities. For example, in the free electron approxi-
mation the surfaces of constant energies are spherical. If the surface of constant
energy 0 corresponding to the momentum k0 first touches the Brillouin zone
boundary at n points, then for energies  ∼ 0 , the gap in the dispersion re-
lation at the Brillouin boundary impedes the progress of the constant energy
surface into the next zone. The effect of the gap on the density of states can be
estimated by only considering the free electron states within the first Brillouin
zone. In this case, the surface of constant energy has an area S() which is given
by
S() = 4 π k 2 − n 2 π k ( k − k0 ) (472)
which is the surface area of the sphere of radius k minus the area of the n
spherical caps. The density of states is proportional to
dk
ρ() ∝ S() (473)
d
Hence, on using
h̄2
d = k dk (474)
m

218
|φQ/2(r)|2 cos(πr/a)2

-4 -3 -2 -1 0 1 2 3 4

Vions(r)

r/a

sin(πr/a)2
|φQ/2(r)|2

-4 -3 -2 -1 0 1 2 3 4

Vions(r)

r/a

Figure 127: The Bloch functions with k on the Bragg planes have the form
of standing waves. The standing waves that have nodes located between the
atoms have lower energies, while the standing waves that have nodes at the
atomic positions have higher energies.

one obtains the approximate expression for the density of states

ρ() 4 π k − n 2 π ( k − k0 )

ρ(0 ) 4 π k0
n n k
= + (1 − )
2 2 k
r0
n n 
= + ( 1− ) (475)
2 2 0

Thus, for energies which fall within the gaps at the Brillouin zone boundaries,
the density of states should be decreased below the free electron density of states.
The states pushed out from the gap region are manifested by an increase in the
density of states just outside the gap. This type of structure in the electronic
density of states is clearly seen above the Fermi energy of Li, as found in the

219
Figure 128: A constant energy surface of free electrons in reciprocal space for a
s.c. solid.

Brillouin Zone Decomposition


Free Electron Density of States
0.15
ρ(ε) [ States (2ma /h ) ]
2 2

0.1

Second Third
Zone Zone
0.05 First Zone

0
0 0.5 1 1.5 2 2.5 3

ε [ units of (h2π2/2ma2) ]

Figure 129: The Brillouin Zone decomposition of the free electron density of
states for a s.c. solid.

calculations of F.S. Ham42 .

The Bragg planes have other significance, as can be inferred from the gradi-
ent of the energy
v
 Eq(0) + E (0)  u (0)
u Eq − E (0) 2
q − Q” q − Q”
E± = ± + | Vions (Q”) |2
t
2 2
(476)
42 F. S. Ham, Phys. Rev. 128, 82 (1962).

220
which is found as
 
(0) (0)
" Eq − Eq #
h̄2 Q” Q” − Q”

∇ q E± = q− ± s
m 2 2 2
(0) (0)
Eq − Eq − Q” + 4 | Vions (Q”) |2
(477)
On the Bragg plane, one has
(0)
Eq(0) = Eq − Q” (478)
therefore, the second term in the expression for the gradient drops out on these
planes. Thus, the gradient of the energy of the mixed bands is given by
h̄2 Q”
 
∇q E± = q − (479)
m 2
Q”
and, as q is on the Bragg plane, the vector q − 2 is parallel to the plane and
so is the gradient. The gradient of the energy is perpendicular to surfaces of
constant energy and so, the constant energy surfaces are usually perpendicular
to the Bragg planes at their points of intersection.

Generally, the vanishing of the normal component of the gradient at the


Brillouin zone boundary is not dependent on the validity of the nearly-free
electron approximation, but is a consequence of symmetry. Consider the case
in which there is a mirror plane symmetry, σ. The mirror plane is assumed to
run through the origin of the Brillouin zone and is parallel to the Brillouin zone
boundary under consideration. Then, the normal component of the gradient is
defined as " #
Ek+δQ − Ek−δQ


Q . ∇k Ek = lim
(480)
δ → 0 2δ
However, since the point k is equivalent to the point k − Q, one has
Ek−δQ = E−Q+k−δQ (481)
and, as there exists a mirror plane σ through the origin and perpendicular to
Q, one also has
E−Q+k−δQ = EQ+σk+δQ (482)
Noting that as k is on the Bragg plane, k ≡ Q + σk, and substituting the
above equality into the definition, one finds that the normal component of the
gradient vanishes at the Brillouin zone boundary


Q . ∇k Ek = 0 (483)

Thus, at the Brillouin zone boundary, either the normal component of the gra-
dient vanishes or the gradient does not exist, i.e. there might be a cusp. The
presence of other types of symmetry can give rise to similar conclusions.

221
Two Bragg Planes and a Mirror Plane

δQ
-Q+k-δ δQ
k-δ

O Q/2

Figure 130: The vanishing of the group velocity at the Brillouin zone boundary
is a consequence of a mirror plane σ parallel to the boundary.

8.1.4 Empty Lattice Approximation Band Structure


Since the nearly-free electron approximation deviates only slightly from the free
electron approximation, the gross features of the band structure can be found
using the empty lattice approximation.

Since the Brillouin zone is a three-dimensional object and is highly sym-


metric, it is only necessary to specify the bands within an irreducible wedge.
Once the bands are specified within the wedge, then by use of symmetry, the
bands are completely known throughout the Brillouin zone. Since it is difficult
to represent the energy dispersion relations in a three-dimensional volume of
reciprocal space, it is customary to specify the dispersion relations on the lines
defining the boundaries of the irreducible wedge. These lines have high symme-
tries.

Consider the case of an f.c.c. Bravais lattice, and consider the bands within
the first Brillouin zone. The high-symmetry points are marked by special letters.

222
The Brillouin zone for a face-centered cubic lattice

L
W
U
Γ
K
X

Figure 131: The high-symmetry points and lines in the Brillouin zone of an
f.c.c. structure.

Γ ≡ (0, 0, 0) ≡ (0, 0, 0)

3 π 2 π
K ≡ 2 a (1, 1, 0) ≡ a ( 34 , 34 , 0)

π 2 π
W ≡ a (2, 1, 0) ≡ a (1, 12 , 0)

2 π 2 π
X ≡ a (1, 0, 0) ≡ a (1, 0, 0)

π 2 π
L ≡ a (1, 1, 1) ≡ a ( 12 , 12 , 12 )

2 π
and, in units of a , these points correspond to

Γ ≡ (0, 0, 0) The zone center

K ≡ ( 34 , 34 , 0) The midpoint of the hexagonal edge

X ≡ (1, 0, 0) The square face center

W ≡ (1, 12 , 0) The vertex

L ≡ ( 12 , 12 , 12 ) The hexagonal face center

The electron bands are usually plotted against k along the directions

223
Γ → X → W → L → Γ → K → X

All of these are high-symmetry lines. The lengths of the linear segments (in
units of 2aπ ) are given by
√ √ √
1 √1 3 3 2 10
1 2 2 2 4 4

In the empty lattice approximation, the single-electron energies can be plotted


along these lines in units of E0 , where

h̄2 4 π2
 
E0 = (484)
2m a2

The components of the reduced wave vectors are defined as the dimensionless
quantities k̃i , where
ki a
k̃i = (485)

It should be noted that two or more Bragg planes may intersect at points lo-
cated on some of these lines.

0
The energies of the various bands can be constructed from the various Ek−Q .
(0)
The lowest energy band that is considered is simply Ek , where QΓ = (0, 0, 0).
This band has the lowest energy, since points in the first Brillouin zone are
closer to the origin than any other reciprocal lattice point Q. The band is
non-degenerate, except if k is located on the Brillouin zone boundary. The
one-electron energy is given by
(0)
Ek
 
= k̃x2 + k̃y2 + k̃z2 (486)
E0

which for Γ → X, reduces to

= k̃x2 for 0 ≤ k̃x ≤ 1 (487)

For the line segment X → W on the Brillouin zone boundary, this band
dispersion relation is evaluated as
1
= 1 + k̃y2 for 0 ≤ k̃y ≤ (488)
2
The band should be doubly-degenerate for most points on the line X W , since
points on this line are equidistant to the origin and the reciprocal lattice point
QX = 4aπ (1, 0, 0). For W → L, this band is described by

(0)
Ek 1 1
= + k̃x2 + ( 1 − k̃x )2 for ≤ k̃x ≤ 1 (489)
E0 4 2

224
The band should be at least doubly-degenerate for points on the line W L,
since the line is on the QL = 4aπ ( 12 , 12 , 12 ) Bragg plane. That is, the line is
symmetrically positioned with respect to the reciprocal lattice points QΓ and
QL . For L → Γ, the dispersion relation reduces to

(0)
Ek 1
= 3 k̃x2 for 0 ≤ k̃x ≤ (490)
E0 2
For Γ → K, the band energy simplifies to
3
= 2 k̃x2 for 0 ≤ k̃x ≤ (491)
4
The last segment is given by the, lower symmetry, interior line K → X, over
which the band energy takes the form
3
= k̃x2 + 9 ( 1 − k̃x )2 for ≤ k̃x ≤ 1 (492)
4

(2) (2)
(4) (3)

(2)
2
(2)
Ek / E0

(1) (2) (1)


(1)
(2)

1 (2)
(2) (1)

(1) (1)

0
Γ Χ W L Γ K Χ

Figure 132: The electronic dispersion relations for an f.c.c. compound in the
empty lattice approximation. The degeneracies of the various branches are
enclosed in parentheses.

The next lowest energy bands can be identified by considering their proxim-
ity to other reciprocal lattice points Q. At any point, this just reduces to finding
the nearest Bragg planes. For our particular path, these are found amongst the
set of reciprocal lattice vectors QX = 4aπ (1, 0, 0), QL = 4aπ ( 12 , 12 , 12 ) and

225
4 π
QL0 = a ( 12 , 12 , − 12 ).
(0) 4 π
The next band to be considered is simply Ek−Q , where QX = a (1, 0, 0).
X
(0)
In the free electron approximation, this band should be degenerate with Ek
along the line segment X W . The dispersion relation for this one-electron band
is given by
(0)
Ek−Q  
2 2 2
X
= ( k̃x − 2 ) + k̃y + k̃z (493)
E0
which, for Γ → X, reduces to

= ( k̃x − 2 )2 for 0 ≤ k̃x ≤ 1 (494)

For X → W , the band dispersion relation reduces to


1
= 1 + k̃y2 for 0 ≤ k̃y ≤ (495)
2
(0)
and, indeed, is degenerate with the band Ek . On the line W → L, the band
(0)
Ek−Q is described by the dispersion relation
X

(0)
Ek−Q 1 1
X
= + ( k̃x − 2 )2 + ( 1 − k̃x )2 for ≤ k̃x ≤ 1 (496)
E0 4 2
Since the line W L is on the Bragg plane with reciprocal lattice vector QL , the
(0) (0)
band energy Ek−Q is higher energy than the band energy Ek−Q . Along the
X L
(0)
line L → Γ, the dispersion relation of Ek−Q is expressed as
X

1
= ( k̃x − 2 )2 + 2 k̃x2 for 0 ≤ k̃x ≤ (497)
2
In the empty lattice approximation, this branch of the band could be expected
to be triply-degenerate since the line is symmetrically placed with respect to the
three equivalent reciprocal lattice points, 4π 4π 4π
a (1, 0, 0), a (0, 1, 0) and a (0, 0, 1).
For Γ → K, this band takes the form
3
= ( k̃x − 2 )2 + k̃x2 for 0 ≤ k̃x ≤ (498)
4
The last segment is given by K → X, for which the band takes the form
3
= ( k̃x − 2 )2 + 9 ( 1 − k̃x )2 for ≤ k̃x ≤ 1 (499)
4
Since, the line is mostly confined within the interior of the Brillouin zone, the
band is generally non-degenerate.

226
(0) 4 π
The next band in the set is Ek−Q , where QL = a ( 12 , 12 , 12 ). In the empty
L
(0)
lattice approximation, this band becomes degenerate with Ek on the Bragg
plane defined by the reciprocal lattice point QL . The one-electron dispersion
relation is given by
(0)
Ek−Q  
L
= ( k̃x − 1 )2 + ( k̃y − 1 )2 + ( k̃z − 1 )2 (500)
E0

which, for Γ → X, is just

= 2 + ( k̃x − 1 )2 for 0 ≤ k̃x ≤ 1 (501)

This branch of the band should be at least four-fold degenerate, since the line
Γ X is symmetrically positioned with respect to the four reciprocal lattice points
4 π 1 1 1
a ( 2 , ± 2 , ± 2 ). For X → W , the branch has a dispersion relation given by

1
= 1 + ( k̃y − 1 )2 for 0 ≤ k̃y ≤ (502)
2
The line W → L is on the Brillouin zone boundary and, therefore, the free
(0) (0)
electron band Ek−Q is degenerate with Ek at these points. On the line W L,
L
the dispersion relation is described by
(0)
Ek−Q 1 1
L
= + ( k̃x − 1 )2 + k̃x2 for ≤ k̃x ≤ 1 (503)
E0 4 2
For L → Γ, this dispersion is given as
1
= 3 ( k̃x − 1 )2 for 0 ≤ k̃x ≤ (504)
2
and one expects the band to be non-degenerate. For Γ → K, this band takes
the form
3
= 1 + 2 ( k̃x − 1 )2 for 0 ≤ k̃x ≤ (505)
4
On the line segments Γ K and K X, the branch of this one-electron band
are generally doubly-degenerate, since the points on the line are symmetrically
positioned with respect to the pair of reciprocal lattice points QL and QL0 . The
last path segment is given by K → X, in which the band has the form of
3
= 1 + ( k̃x − 1 )2 + ( 2 − 3 k̃x )2 for ≤ k̃x ≤ 1 (506)
4

(0)
The last band we consider is Ek−Q , where QL0 = 4aπ ( 12 , 12 , − 12 ). Our
L0
set of lines only meet this Bragg plane of QL0 at the point K. In the empty

227
lattice approximation, the band should be triply-degenerate at the K point.
The dispersion relation is given by
(0)
Ek−Q  
L0 2 2 2
= ( k̃x − 1 ) + ( k̃y − 1 ) + ( k̃z + 1 ) (507)
E0

which, for Γ → X, is just

= 2 + ( k̃x − 1 )2 for 0 ≤ k̃x ≤ 1 (508)

For X → W , the dispersion relation for the band is given by


1
= 1 + ( k̃y − 1 )2 for 0 ≤ k̃y ≤ (509)
2
For W → L, this band is described by
1 1
= + ( k̃x − 1 )2 + ( 2 − k̃x )2 for ≤ k̃x ≤ 1 (510)
4 2
For L → Γ, this dispersion is given as
1
= 2 ( k̃x − 1 )2 + ( k̃x + 1 )2 for 0 ≤ k̃x ≤ (511)
2
This branch may be expected to be triply-degenerate as the line L Γ is symmetri-
cally positioned with respect to the three reciprocal lattice points, 4aπ (− 12 , 12 , 12 ),
4 π 1 1 1 4 π 1 1 1
a ( 2 , − 2 , 2 ) and QL0 = a ( 2 , 2 , − 2 ). The line Γ K is equidistant from
the reciprocal lattice points QL and QL0 and, therefore, this band should be
(0)
degenerate with Ek−Q on this line segment. On the line Γ K, the band takes
L
the form
(0)
Ek−Q 3
L0
= 1 + 2 ( k̃x − 1 )2 for 0 ≤ k̃x ≤ (512)
E0 4
The last line segment is given by K → X, for which the band has the form
3
= 1 + ( k̃x − 1 )2 + ( 2 − 3 k̃x )2 for ≤ k̃x ≤ 1 (513)
4
which is also doubly-degenerate.

It is seen that some branches of these bands are highly degenerate. For a
real solid, the symmetry of the bands is not dictated by the symmetry of the
Brillouin zone (which is determined by the symmetry of the Bravais lattice),
but instead is dictated by the symmetry of the space group of the lattice which
includes the atomic basis. When Vions 6= 0, the degeneracy of the various
branches found in the empty lattice approximation may be lifted. Group theory
can be used to determine whether or not the potential lifts the degeneracy of

228
the branches.

Thus, even in the empty lattice approximation, the method of plotting bands
shows a great deal of structure. The real structure is actually inherent in the
Bragg planes which generally can be associated with an“energy gap” in the dis-
persion relations. The “gap” may or may not extend across the entire Brillouin
zone. A gap only appears in the density of states if the “gap” extends across the
entire Brillouin zone. The nearly-free electron approximation has been worked
out in detail for Al by B. Segall43 and has been compared with the results of
numerical calculations.

Nearly Free Electron Bands for f.c.c. Al

1.25

1
Ek [ Rydbergs ]

0.75

0.5

0.25

0 f.c.c.

-0.25
Γ Χ W L Γ K W Χ

Figure 133: The the nearly-free electron approximation to the dispersion rela-
tions of f.c.c. Al.

For the b.c.c. lattice, the reciprocal lattice vectors are


1 4π
b1 = ( êx + êy )
2 a
1 4π
b2 = ( êx + êz )
2 a
1 4π
b3 = ( êy + êz ) (514)
2 a
The Cartesian coordinates of the high-symmetry points are

43 B. Segall, Phys. Rev. 124, 1797 (1961).

229
The Brillouin zone of the body-centered cubic lattice

P
N
Γ
H
N

Figure 134: The high-symmetry points and lines in the Brillouin zone of a b.c.c.
solid.

Γ ≡ (0, 0, 0)

H ≡ (1, 0, 0)

N ≡ ( 12 , 12 , 0)

P ≡ ( 12 , 12 , 12 )

2 π
in units of a .

——————————————————————————————————

8.1.5 Exercise 36
Derive the lowest energy bands of a b.c.c. lattice in the empty lattice approx-
imation. Plot the dispersion along the high-symmetry directions (Γ → H →
N → P → Γ → N ).

——————————————————————————————————

230
8.1.6 Degeneracies of the Bloch States
The degeneracies of the bands at various points in the Brillouin zone, found in
the empty lattice approximation, can be raised by the crystalline potential. The
character and degeneracies of the bands at symmetry points can be ascertained
by the use of group theory44 .

Given a Bloch function φn,k (r), one can apply a general point group sym-
metry operator Ô(Aj ) to the Bloch function, thereby, transforming it into the
Bloch function corresponding to the wave vector Aj k

Ô(Aj ) φn,k (r) = φn,Aj k (r) (515)

This is proved by considering the combined operation consisting of the point


group operation Ô(Aj ) followed by a translation through a Bravais lattice vector
R. The effect of the combined operation is evaluated as

T̂R Ô(Aj ) φn,k (r) = T̂R φn,k (A−1


j r)
 
= exp − i k . Aj R φn,k (A−1
−1
j r) (516)

where the second line follows from Bloch’s theorem. However, we note that
the scalar product remains invariant if both vectors are transformed. We shall
transform the vectors k and ( A−1
j R ) by Aj . Hence, as

k . ( A−1
j R) = ( Aj k ) . ( Aj A−1
j R)
= ( Aj k ) . R (517)

we find that
 
T̂R Ô(Aj ) φn,k (r) = exp − i ( Aj k ) . R φn,k (A−1
j r)
 
= exp − i ( Aj k ) . R Ô(Aj ) φn,k (r)

(518)

Since the quantity  


exp − i ( Aj k ) . R (519)

is the eigenvalue of the translation operator T̂R , the Bloch wave vector of the
function Ô(Aj ) φn,k (r) is Aj k. As this is an energy eigenfunction, the trans-
formed function is a Bloch function. That is,

φn,Aj k (r) = Ô(Aj ) φn,k (r) (520)


44 L. P. Bouckaert, R. Smoluchowski and E. Wigner, Phys. Rev. 50, 58 (1936).

231
Since the point group symmetry operations commute with the Hamiltonian,

[ Ĥ , Ô(Aj ) ] = 0 (521)

the Bloch states Ô(Aj ) φn,k (r) all have the same energy En,k .

A set of basis functions for a representation of the space group can be con-
structed by repeatedly applying the point group symmetry operators to any
one of the Bloch functions. The same vector k cannot appear in distinct bases
created from a Bloch function in this manner, since the symmetry operations
form a group. This means that two such bases are either identical or have no
wave vector k in common. A basis created from the Bloch function φn,k (r) in
this fashion may be either reducible or irreducible.

An irreducible basis can be constructed by selecting an appropriate subset


of Bloch functions from the above basis set. If one considers the set of wave
vectors Ai k, then certain of these points may be equivalent in that

Ai k = Aj k + Q (522)

where Q is a reciprocal lattice vector. The star of k is the set of all the inequiv-
alent wave vectors Ai k. More precisely, the star of the wave vector k consists
of the set of all mutually inequivalent wave vectors Ai k, where Ai ranges over
all the operations of the point group. Since none of the Bloch wave vectors in
the star are equivalent, the corresponding Bloch functions are all linearly inde-
pendent. Hence, the Bloch functions of the star may be used to construct an
irreducible basis. The star of k contains fewer wave vectors than the order of
the point group, if either k lies on a symmetry line or is on the Brillouin zone
boundary. As an example, consider a crystal with a simple tetragonal Bravais
Lattice and a one-atom basis. The crystal has the D4h point group symmetry.
The stars of the wave vectors Γ, Z, M and A each only contain one wave vector.
However, the star of the wave vector X contains a total of two points, as does
the star of wave vector R. For the case of tetragonal symmetry, the star of
a general k vector contains a total of 2! × (2)3 wave vectors. Since the set of
energies of Bloch states with wave vector k are equal to the set of energies of
Bloch states at each wave vector in the star of k, it is only necessary to find the
electronic dispersion relations in an irreducible wedge of the Brillouin zone.

The group of the k vector consists of all symmetry operations which, when
acting on k, lead to an equivalent point. That is, the symmetry operations of
the group of the k vector satisfy

Aj k = k + Q (523)

where Q is a reciprocal lattice vector. As an example, the groups of the k vectors


for the points Γ, Z, M, A of a crystal with a simple tetragonal lattice coincide
with the D4h point group of the tetragonal lattice itself. The groups of the k

232
Z R
A

Γ X

Figure 135: The tetragonal Brillouin zone and the points of high symmetry.

vectors at X and R are D2h , and D2h is a subgroup of D4h . In general, the
group of the k vector of the Γ point will always coincide with the point group
of the crystal. However, at a general point, the group of the wave vector only
consists of the identity. The group of the k vector has irreducible representa-
tions, and these are called the small representations.

The Bloch functions corresponding to the wave vectors of the star of k can
be symmetrized with respect to the small representations. The symmetrization
can be performed by using the projection method. Although the groups of the
wave vectors in the star may be different, the small representation of any one
can be chosen for the symmetrization process. After the symmetrization, the
resulting basis functions form an irreducible representation of the space group.
Each basis function of the small representation only corresponds to exactly one
wave vector in the star and its equivalent wave vectors. The basis functions
corresponding to the different irreducible representations are orthogonal.

The irreducible representations of the space group constructed from the


Bloch functions are fully determined by the star of the k vector and the small
representation. The basis functions forming the irreducible representation of
the space group constructed from the Bloch state φn,k (r) are eigenstates of Ĥ0
with energy En,k . Barring accidental degeneracies, the degeneracy of this eigen-
value is equal to the dimension of its irreducible representation. As k varies in
the Brillouin zone, the eigenvalue En,k and the corresponding basis functions
vary continuously. The group of the k vector also varies as k varies. Whenever
the dimension of the small representation corresponding to the basis function
φn,k (r) changes, the degeneracy of En,k changes. This may signify that at these

233
points different bands cross or merge together.

Alternatively, at k there are a vast number of bands each corresponding to


a different small representation. The degeneracy of each band is given by the
dimension of the corresponding small representation. If an irreducible represen-
tation of the group of the wave vector k can be decomposed into the irreducible
representations of the group of k 0 , then on varying k to k 0 , the branch will
split into sub-levels. The degeneracies of the sub-levels are determined by the
dimensions of the irreducible representations contained in the decomposition.

——————————————————————————————————

As an example, consider the nearly-free electron bands of Zinc Blende. The


material has tetrahedral point group symmetry, Td . The point group contains
twenty four elements in five equivalence classes. One class consists of the iden-
tity E. There is a class of eight C3 operations, which contain the rotation C3
and the inverse rotation C3−1 about the four axes [1, 1, 1], [1, 1, 1], [1, 1, 1] and
[1, 1, 1]. There is a class consisting of three C2 operations around the [1, 0, 0],
[0, 1, 0] and [0, 0, 1]. There is a class consisting of six S4 operations around the
[1, 0, 0], [0, 1, 0] and [0, 0, 1] axes. Finally, there is a group consisting of six σ op-
erations which are reflections in the six planes (1, 1, 0), (1, 0, 1), (0, 1, 1), (1, 1, 0),
(1, 0, 1) and (0, 1, 1). Therefore, the group has five irreducible representations.
The character table is given by

Td E C2 (3) C3 (8) (S4 )(6) σ(6)

Γ1 1 1 1 1 1
Γ2 1 1 1 -1 -1
Γ3 2 2 -1 0 0
Γ4 3 -1 0 1 -1
Γ5 3 -1 0 -1 1

Let us consider the band structure along the high-symmetry directions [1, 1, 1]
and [1, 0, 0] directions.

At the Γ point the group of the k vector coincides with the point group of
the crystal. Since the free electron approximation for the Bloch wave function
for k = 0 is a constant, it is a basis for the Γ1 representation. Thus, the level
is non-degenerate.

234
Figure 136: A unit cell of the Zinc Blende structure.

At a general point along the eight [1, 1, 1] directions (the Λ high-symmetry


lines), the group of the k vector is C3v and contains six elements in three classes.
These classes are the identity E, a class consisting of the rotation C3 about the
[1, 1, 1] axis and its inverse C3−1 , and the class consisting of three reflections σ in
the three equivalent (1, 1, 0) planes containing the [1, 1, 1] axis. The character
table is given by

C3v E C3 (2) σ(3)

Λ1 1 1 1
Λ2 1 1 -1
Λ3 2 -1 0

Thus, the branches along the Λ high-symmetry lines are either singly or dou-
bly degenerate, when the crystalline potential is introduced. The branch which
(0) 2
emanates from k = 0 with the approximate energy Ek = 2h̄m k 2 belongs to
the Λ1 representation as this is compatible with the Γ1 representation.

At the end point L where k = 2aπ ( 12 , 12 , 12 ), the symmetry operations


are identical to those of Λ. In the free electron approximation, the state at L
is doubly degenerate (ignoring spin) since the wave vectors 2aπ ( 12 , 12 , 12 ) and
− 2aπ ( 12 , 12 , 12 ) differ by a reciprocal lattice vector Q = 2aπ (1, 1, 1). Using
the compatibility relations, one can show that the other state also has Λ1 sym-

235
metry. These two states are accidentally degenerate, since they are not partner
basis functions of a multi-dimensional irreducible representation. Therefore, the
degeneracy may be lifted by the presence of a crystalline potential Vions (Q).

On continuing along the Λ high-symmetry line, one reaches the point k =


2 π
a (1, 1, 1). Since the primitive reciprocal lattice vectors of the f.c.c. lattice
are of the form

b1 = (−1, 1, 1)
a

b2 = (1, −1, 1)
a

b3 = (1, 1, −1) (524)
a
then Q = b1 + b2 + b3 is equal to 2aπ (1, 1, 1). Thus, the point k = 2aπ (1, 1, 1)
is equivalent to the Γ point. The star consists of just one wave vector. At this
point, the eight nearly-free electron bands corresponding to
 
1 2π
φkj (r) = √ exp i ( ± x ± y ±z ) (525)
V a

are degenerate. These eight functions form the basis of an eight-dimensional


representation of the group Td which is reducible. In this representation, a
symmetry transformation A is represented by the 8 × 8 matrices, D(A), which
are constructed according to the prescription

Ô(A) φki (r) = φki (A−1 r)


X
= φkj (r) D(A)j,i (526)
j

The characters of this eight-dimensional representation are given by the trace


of the 8 × 8 matrices and, therefore, the character of an operation is just the
number of wave functions that are unchanged by the transformation.

Class Transformation χ

E x, y, z 8
C2 (3) x, y, z 0
C3 (8) y, z, x 2
S4 (6) x, z, y 0
σ(6) y, x, z 4

236
This eight-dimensional representation, Γ, is reduced into the irreducible repre-
sentations, Γµ , via X
Γ = aµ Γµ (527)
µ

The decomposition can be found from considering the characters. The charac-
ters of a symmetry operation A, χ(A), is decomposed into the characters of the
irreducible representations, χµ (A), via
X
χ(A) = aµ χµ (A) (528)
µ

The multiplicity aµ can be found from the orthogonality relation


X
gi χ(Ai ) χµ (Ai ) = g aµ (529)
i

where the sum over i runs over all the equivalence classes of the group, and gi
is the number of symmetry elements in the i-th equivalence class, and g is the
order of the group. This procedure leads to the decomposition

χ(Ai ) = 2 χΓ1 (Ai ) + 2 χΓ4 (Ai ) (530)

Thus, the eight plane wave basis can be symmetrized into two sets of basis
functions of Γ1 symmetry and two three-dimensional sets of basis functions
of Γ4 symmetry. The symmetrization process is performed by the use of the
projection method. A projector, P̂ µ which projects the functions on to an
irreducible set of basis functions, is constructed from the symmetry operations
Ô(A) and the characters of the operations via

dim(Γµ ) X µ
P̂ µ = χ (A) Ô(A) (531)
g
A

In this, dim(Γµ ) is the dimension of the µ-th irreducible representation, i.e.,


dim(Γµ ) = χµ (E). When the projector acts on an arbitrary combination of
functions with equivalent wave vectors k, φk (r), it produces a basis function,
φµk (r) for the µ-th irreducible representation

P̂ µ φk (r) = φµk (r) (532)

In this way, one can construct the set of symmetrized basis functions:

237
Representation Basis functions

Γ1
cos 2πx
a cos 2πy
a cos 2πz
a

Γ1
sin 2πx
a sin 2πy
a sin 2πz
a

Γ4
cos 2πx
a sin 2πy
a sin 2πz
a
2πy
sin a cos a sin 2πz
2πx
a
sin 2πx
a sin 2πy
a cos 2πz
a

Γ4
sin 2πx
a cos 2πy
a cos 2πz
a
2πy
cos a sin a cos 2πz
2πx
a
cos 2πx
a cos 2πy
a sin 2πz
a

In this basis, all the matrices D(A) representing the symmetry operators A
have the same block diagonal form. The matrices contains two one-dimensional
blocks and two three-dimensional blocks. Thus, these eight levels may be split
by the application of a potential into two non-degenerate levels and two sets of
triply degenerate levels. Therefore, the degeneracies at the eight approximate
(free-electron) bands at Γ are not completely lifted.

Along the X direction (the ∆ high-symmetry line), the wave vectors are of
the form (k, 0, 0) where 0 < k < 2aπ . The group of k is C2v . It has four
elements in four classes: the identity E, a two-fold rotation about the [1, 0, 0]
axis, and the two diagonal mirror planes σd and σd0 . The character table is given
by

238
C2v E C42 σd σd0

∆1 1 1 1 1
∆2 1 1 -1 -1
∆3 1 -1 1 -1
∆4 1 -1 -1 1

Therefore, along this direction, all the irreducible representations are one-
dimensional. The symmetry of the wave function emanating from (0, 0, 0) belong
to ∆1 since this is the only irreducible representation compatible with Γ1 . This
branch continues up to the X point. The point 2aπ (1, 0, 0) is equivalent to the
point − 2aπ (1, 0, 0), as they are related via the Q vector Q = b2 + b3 . At
the X point, the lowest energy level in the nearly-free electron approximation
is doubly degenerate.

The group of the k vector at the X point is D2d and consists of eight el-
ements arranged in five classes. These are the identity E, a two-fold rotation
about the x axis C42 , a class of two elements which are the two-fold rotations
C2 about the y and z axis, and a class containing two S4 operations about the
x axis, and a class of two diagonal reflections σd on the (0, 1, 1) and the (0, 1, 1)
planes. Thus, there are five irreducible representations. The character table is
given by

D2d E C42 (1) C2 (2) S4 (2) σd (2)

X1 1 1 1 1 1
X2 1 1 1 -1 -1
X3 1 1 -1 -1 1
X4 1 1 -1 1 -1
X5 2 -2 0 0 0

At the X point, the wave functions of the two-fold degenerate energy levels, E (0) ,
found in the nearly-free electron approximation belong to the one-dimensional
X1 and X3 irreducible representations. This degeneracy may be raised by the

239
potential.

On continuing along the X direction, one reaches the point QX = 2aπ (2, 0, 0).
The six k points (±2, 0, 0), (0, ±2, 0) and (0, 0, ±2) are all equivalent to the zone
center. The group of the wave vector is Td . The six wave functions
 
1 4π
φQ (r) = √ exp ± i x
X
V a
 
1 4π
φQ (r) = √ exp ± i y
Y
V a
 
1 4π
φQ (r) = √ exp ± i z
Z
V a
(533)
can be used as a basis for a six-dimensional representation. In this representa-
tion, the characters of the symmetry operations are given by:

Class Transformation χ

E x, y, z 6
C2 (3) x, y, z 2
C3 (8) y, z, x 0
S4 (6) x, z, y 0
σ(6) y, x, z 2

This representation is degenerate and can be decomposed via


X
Γ = aµ Γµ (534)
µ

The multiplicities aµ are calculated from the orthogonality relations


X
gi χ(Ai ) χµ (Ai ) = g aµ (535)
i

which leads to the decomposition


Γ = Γ1 + Γ3 + Γ4 (536)
Thus, the six-dimensional representation of the group of the wave vector (200)
is reducible and reduces into a one-dimensional, a two-dimensional and a three-
dimensional irreducible representation. The basis functions can be symmetrized

240
using the projection method. The basis functions for the small representations
are

Representation Basis functions

Γ1
cos 4πx
a + cos 4πy
a + cos 4πz
a

Γ3
cos 4πy
a − cos 4πz
a
2 cos a − cos 4πy
4πx
a − cos 4πz
a

Γ4
sin 4πx
a
sin 4πy
a
sin 4πz
a

(0) h̄2
Hence, the six-fold degenerate free-electron energy level EQ = 2m ( 4aπ )2
X
may have its degeneracy lifted by Vions (Q).

——————————————————————————————————

8.1.7 Exercise 37
Using the symmetrized wave functions at k = ( 2aπ ) (1, 1, 1) in the nearly-free
electron model for Zn Blende
r
8 2πx 2πy 2πz
φΓ1 = cos cos cos
N a3 a a a
r
8 2πx 2πy 2πz
φΓ4 (x) = sin cos cos
N a3 a a a
r
8 2πx 2πy 2πz
φΓ4 (y) = cos sin cos
N a3 a a a
r
8 2πx 2πy 2πz
φΓ4 (z) = 3
cos cos sin
N a a a a
(537)

241
show that the matrix elements of the momentum operator between the Γ1 and
Γ4 basis functions are given by
 2
2 2 2 2 π h̄
| < Γ1 | p̂x | Γ4 (x) > | = | < Γ1 | p̂y | Γ4 (y) > | = | < Γ1 | p̂z | Γ4 (z) > | =
a
(538)
while all other matrix elements are zero.

——————————————————————————————————

8.1.8 Exercise 38
Consider the nearly-free electrons bands in f.c.c. Al, on the the high-symmetry
lines Γ X and X W . Consider the branches of the free electron bands

(0) h̄2
Ek−Q = ( k − Qi )2 (539)
i 2m
where Qi runs over the four reciprocal lattice vectors Qi = 2π a (1, ±1, ±1).
Show that, in the free electron approximation, there is some degeneracy between
the band branches on these line segments. Set up the secular equation for these
four bands on the two line segments, and solve them using Matlab. Assume that
the lattice constant of Al is 4.05 Å, and that the non-zero values of Vions (Q)
are given by

Vions (1, 1, 1) = 0.023 Rydbergs


Vions (2, 0, 0) = 0.043 Rydbergs

while all other matrix elements are zero.

——————————————————————————————————

8.1.9 Brillouin Zone Boundaries and Fermi Surfaces


The Brillouin zone boundaries play an important role in the understanding of
Fermi surfaces. In the empty lattice approximation, the Fermi surface is a
sphere when represented in the extended zone scheme. The nearly-free electron
approximation introduces a distortion to the sphere which is most marked near
the Brillouin zone boundaries.

In general, if the spherical Fermi surface crosses a Bragg plane, then the
sphere may distort. In particular, the constant energy surface generally should
be perpendicular to the Bragg plane at the line where they intersect. Due to
the appearance of the potential Vions (Q) in the expression for the Bloch en-
ergy near the Bragg plane, and also due to the accompanying band splitting,

242
the circles of intersection of the constant energy surfaces (corresponding to EF )
with the Bragg plane do not match up. This is necessary since the distortion of
the Fermi surface must conserve the volume enclosed. This volume is equal to
the volume enclosed by the spherical Fermi surface found in the empty lattice
approximation.

The Fermi surface in the reduced Brillouin zone scheme can be constructed
from the Fermi surface in the extended zone scheme. This is done by translating
the disjoint pieces of the Fermi surface in the higher order zones by reciprocal
lattice vectors, so that the pieces fit back into the first Brillouin zone.

The first Brillouin zone is the Wigner-Seitz unit cell of the reciprocal lattice.
It encloses the set of points that are closer to Q = 0 than they are to any other
reciprocal lattice vector Q 6= 0. This can be restated as, the first Brillouin
zone consists of the volume in the reciprocal lattice which can be accessed from
the origin without crossing a Bragg plane.

The second Brillouin zone is the volume that can be reached from the first
Brillouin zone by crossing only one Bragg plane.

Likewise, the (n + 1)-th Brillouin zone consists of the points, not in the
(n − 1)-th zone, that can be reached from the n-th zone by crossing only one
Bragg plane. Alternatively, the n-th Brillouin zone is the volume that can only
be reached from the origin by crossing a minimum of (n − 1) Bragg planes.

The Fermi surface is constructed by:

(i) Drawing the free electron sphere.

(ii) Distorting the sphere at the Bragg planes.

(iii) For each of the n Brillouin zones, take the portions of the surface in
the n-th zone and translate them by reciprocal lattice vectors so that they lay
within the first Brillouin zone. The resulting surface is the branch of the Fermi
surface assigned to the n-th band in the extended zone scheme.

The Hume-Rothery rules provide a correlation of crystal structure with the


number of electrons per unit cell, or band filling45 . It is an empirical rule
which only applies to alloys of noble metals, such as Cu, Ag and Au, with s-p
elements such as Zn, Al, Si, and Ge. It is assumed that the noble metals have
one electron outside the closed d shell, that Zn has two conduction electrons,
Al has three electrons, and so on. With these assumptions, then the alloys have
an f.c.c. phase for an average number of electrons per atom up to 1.38, while
45 T. B. Massalaski, Binary Alloy Phase Diagrams, eds. J. L. Murray, L. H. Bennett and

H. Baker, Vol. 1, A.S.M., Metals Park, Ohio, (1986).

243
The Fermi surface and Brillouin zone of b.c.c. Na
ez

ey
ex

Figure 137: The Fermi surface of b.c.c. N a is completely enclosed within the
first Brillouin zone.

Figure 138: Cu forms in an f.c.c. structure. The Fermi surface and the Brillouin
zone boundary intersect in circular rings (necks) around the L points. [After D.
Shoenberg, Proc. Roy. Soc. 379, 1 (1983).]

the b.c.c. phase is stable for band-fillings between 1.38 and 1.48. In the b.c.c.
structure, the smallest vectors from the zone center to each face of the Brillouin
zone have the form 12 2aπ (1, 1, 0), whereas for the f.c.c. lattice these vectors
are of the form 12 2aπ (1, 1, 1). Therefore, the radius of the Fermi sphere, kF , at

244
Figure 139: The experimentally observed (T, n) phase diagram of Cu-Zn and
Cu-Ga intermetallic alloys. After W. Hume-Rothery, J. Inst. Metals 90, 42
(1961). The α phase is f.c.c., the β phase is b.c.c., the γ phase has a complex
cubic structure, and a hexagonal -phase occurs near 1.75.

which it first makes contact with the Brillouin zone boundary is given by
√ π
kF = 3 for f.c.c.
a
√ π
kF = 2 for b.c.c. (540)
a
When the Fermi sphere first makes contact with the zone boundary, the occupied
band is depressed by Vions (Q) resulting in an energy lowering which stabilizes
the structure. In the free electron approximation, the number of electrons per

Proposed Density of States for Cu-Zn Alloys

0.4

b.c.c.
0.3
ρ(ε) [ States / eV ]

f.c.c.

0.2

0.1
[After H. Jones (1937)]

0
0 1 2 3 4 5 6 7 8 9
ε [ eV ]

Figure 140: The proposed density of states for Cu − Zn alloys.

245
primitive unit cell, n, is given by
V 4π 3
n = 2 k (541)
N (2π) 3 F
3

where
V a3
= for f.c.c.
N 4
V a3
= for b.c.c. (542)
N 2


Thus, one finds that the critical number n is given by 4 = 1.36 for the f.c.c.

and 23 π = 1.48 for the b.c.c. lattices.

The above explanation of the Hume-Rothery rules as proposed by Jones46


does require further modification. Heine and Weaire47 noted that it is neces-
sary to consider the presence of the Cu d bands. It is also unclear whether the
concept of a sharp Fermi surface remains valid at these high concentration of
impurities, and furthermore, it is unknown what the electron-electron interac-
tion does to the relative stability of these phases48 .

Hume-Rothery observed that the number of valence electrons per unit cell
plays a critical role in the stability of intermetallic alloys. Jones’s hypothe-
sized that the proximity of the Fermi surface to prominent Bragg scattering
planes does play an important role in stabilizing these alloys. Apparently, the
Jones mechanism is also active in intermetallic quasi-crystals. Although quasi-
crystals do not have conventional Bragg planes, they do have prominent peaks
in their diffraction patterns which can be used to construct a pseudo-Brillouin
or Jones zone that takes into account the most important components of the
electronic potential. Friedel49 has noted that Jones’s mechanism is optimal in
quasi-crystals because of the high-multiplicity of Bragg planes due to the icosa-
hedral symmetry, so that the pseudo-zone is almost spherical. Therefore, it is
not surprising that the Fermi energy is found to occur in a dip in the density of
states of many quasi-crystals.

46 H.Jones, Proc. Phys. Soc. A 49, 250 (1937).


47 V.Heine and D. Weaire, Solid State Physics, Volume 22, Academic Press, N.Y., (1970).
48 The above argument is equivalent to assuming that the stable phase is found by minimizing

the band energy given by


Z F
E = 2 d  ρ()
−∞
(543)
and, therefore, double counts the effect of electron-electron interactions.
49 J. Friedel, Phil. Mag. B 65, 1125 (1992).

246
Al–Cu Hume–Rothery alloys 4241

Figure 6. Al p CB distribution curves in pure Al (full line), γ -Al35 Cu65 (starred line) and
i-AlCuFe (line with triangles) alloys as adjusted to the Al 3p intensity at the Fermi level (see
text). For clarity only the Al 3p distribution curve of pure Al is shown.

Figure 7. X-ray diffraction pattern of the δ-Al39 Cu61 phase and its Brillouin–Jones zone.

Figure 141: The X-ray diffraction pattern of δ-Al39 Cu61 . After V. Fournèe et
al., J. Phys. Our
CM. 10,
alloys are 4231 (1998).
HR compounds and consequently a pseudo-gap is expected at EF .
The one we observe is quite small as compared to the pseudo-gap in approximant and
quasicrystalline alloys, where IF goes down to less than 15% in icosahedral phases.
Nevertheless, the FS–BZ interaction should be strong, especially in the case of the γ and
δ phases, where the BZ is constructed with 36 faces corresponding to Bragg peaks that
concentrate almost all the intensity of the x-ray diffraction pattern as shown in figure 7.
The partial Al p DOS in γ -Al4 Cu9 calculated using the TB–LMTO–ASA method and the
corresponding calculated Al Kβ spectra are presented in figure 8. In that case, a direct

4242 V Fournée et al

Figure 8. Experimental (top) and calculated (bottom) Al p DOS curves of the γ -Al4 Cu9 phase
and its Brillouin zone.
Figure 142: The experimentally observed and calculated Al p-density of states
for γ-Al4 Cu 9 . After
comparison with V.
γ -AlFournèe et al., J. Phys. CM. 18, 4231 (1998).
35 Cu65 is possible, as it has the same γ -brass structure. The Fermi
level falls in a pronounced pseudo-gap, confirming the strong FS–BZ interaction. The
width of this pseudogap is about 1 eV. After broadening, the relative intensity at EF is
IF = 40%, just as the experimental value. Taking the number of faces N as a rough
parameter for the sphericity of the BZ, the comparison with the case of the approximant
Al13 Fe4 or the quasicrystalline i-AlCuFe, with N equal to 30 and 42 respectively, shows
that the HR mechanism alone is not sufficient to explain the formation of the pseudogap
observed by SXS in Al–TM–TM0 quasicrystalline phases (TM, TM0 : transition metals). The
wider depletion cannot be the result of a more spherical shape of the PBZ, differences with
the γ phase being too small.
More likely, the effect of the TM should 247 be invoked. Indeed, it has been shown
(Friedel 1992, Trambly de Laissardière et al 1995) that the presence of TM d states of
energies close to the gap, coupled to the sp states, results in an increase of the magnitude
of the potential acting on the Fermi electrons and consequently an increase of the width
of the resulting pseudo-gap. Previous experiments on Al–TM alloys have revealed indeed
the strong Al p–TM d hybridization at EF , the result of which is to repel the electronic
states on both sides of the Fermi level, giving rise to an enhanced pseudo-gap (Belin
et al 1992, 1994a, Belin-Ferré et al 1996). The p–d hybridization is even responsible for
the opening of an almost true gap in Al2 Ru and Ga2 Ru semiconductors (Nguyen Manh
et al 1992, Fournée et al 1997). Note that these two effects, diffraction by Bragg planes
and hybridization with TM d states, should not be considered as distinct effects as the
sublattice of the TM atoms contributes predominantly to the scattering potential VK . Note
8.1.10 The Geometric Structure Factor
The potential Vions (r) is a periodic function and can be defined in terms of the
ionic potentials of the basis atoms Vatom , the lattice vectors R, and the basis
vectors rj , via
X X
Vions (r) = Vatom (r − R − rj ) (544)
R j

The evaluation of the Fourier Transform of the potential can be reduced to an


evaluation of the Fourier Transform in one unit cell of the lattice as
Z   X
1 3
Vions (Q) = d r exp − i Q . r Vatom (r − R − rj )
V V
R,j
Z  
1 3
X
= d r exp − i Q . ( r − R ) Vatom (r − R − rj )
V V
R,j
Z  
1 X
= d3 r 0 exp − i Q . r0 Vatom (r0 − rj )
V V0
R,j

(545)

where we have used the Laue condition


 
exp i Q . R = 1 (546)

and the transformation r0 = r − R. Furthermore, since the Bravais lattice vec-


tors do not explicitly appear in the summand, the sum over R merely produces
a factor of N X
≡ N (547)
R

one has
Z   X
N 3
Vions (Q) = d r exp − iQ.r Vatom (r − rj )
V V j
  Z  
N X 3
= exp + i Q . rj d r” exp − i Q . r” Vatom (r”)
V j V

N
= S(Q) Vatom (Q) (548)
V
where S(Q) is the geometric structure factor associated with the basis and the
other factor is the Fourier transform of the ionic potential
Z  
Vatom (Q) = d3 r exp − i Q . r Vatom (r) (549)
V

Thus, when the geometric structure factor vanishes, the Fourier component of
the lattice potential also vanishes and then the lowest order splitting at the

248
Bragg plane also vanishes. Examples of this are provided by the diamond struc-
ture and also by the hexagonal close-packed lattice.

The vanishing of the form factor of the ionic potential occurs in the diamond
structure phases of Si (and Ge). The nearly-free electron method may be ap-
plied, as the Fourier components of the effective potential from the lattice of ions
(the pseudo-potential) are reasonably small since they are orthogonalized with
the 2p (and 3p) core states. We shall assume that the pseudo-potential method
can be applied and that the pseudo-potential can be approximated by a local
form. This assumption is highly questionable since (as we shall see later) the
cancellation of part of the ionic potential because of orthogonality with the 2p
cores requires that the pseudo-potential should be directional dependent. How-
ever, since the values of the local approximation to the pseudo-potential are not
known apriori, the pseudo-potential model can be viewed as only producing a
fit of the bands and the results should be discarded when they conflict with the
results obtained by rigorous means. The semiconducting materials Si (and Ge)

8 electrons per primitive unit cell


5

(2,0,0)
4
E0k [ arbitrary units ]

3
(1,1,1)
(1/2,1/2,−3/2)
(1,1,0)
2

(1,0,0)
1
(1/2,1/2,1/2)
(0,0,0)
0
L Λ Γ (ka/2π) ∆ X

Figure 143: The free electron dispersion relation for an f.c.c. solid.

form in the diamond structure, and have eight atoms in the conventional cubic
unit cell and have lattice constants of a = 5.43 (and 5.65) Å. Due to the basic
f.c.c. structure, only the conventional unit cell reciprocal lattice vectors of the
form Q = 2π a (m1 , m2 , m3 ) where the mi are either all even or are all odd yield
non-zero values of the Fourier components of Vions . Furthermore, since there
are two identical atoms in the basis, the structure factor is given by
  
π
S(Q) = 1 + exp i ( m1 + m2 + m3 ) (550)
2

249
The structure factor causes the potential to vanish for half of the Q vectors
which correspond to the set of all even mi . That is, the non-zero Fourier com-
ponents correspond either to the set of all odd integers mi or the set of all even
values of mi for which m1 + m2 + m3 is a doubly even integer.

From considering the free electron model where the eight atoms in the con-
ventional f.c.c. unit cell each contribute four electrons to the valence band, the
Fermi wave vector of the empty lattice is expected to be given by

4π kF3 a3
32 = 2 (551)
3 ( 2π )3
Hence, the chemical potential is approximately estimated to be

h̄2 2
µ = k
2m F
  23  2
12 h̄2 2π
=
π 2m a
2  2
h̄ 2π
= 2.4435 (552)
2m a
Therefore, to obtain a reasonable description of the structure near the gap, it is
crucial to include the free electron states corresponding to the set of the eight
reciprocal lattice vectors (±1, ±1, ±1), the set of the six (±2, 0, 0) and (0, 0, 0).
At the Γ point, the approximate free electron states have energies of 0, 3 and 4 in
units of 0.3754 Rydbergs for Si (in units of 0.3455 Rydbergs for Ge). The Fourier
components of the pseudo-potential corresponding to Q = 2π a (2, 0, 0) is ineffec-
tual, since S(Q) = 0 for this value of Q. The non-zero values for the pseudo-
potential are given by V (Q(1,1,1) ) = +0.149 Ry, V (Q(2,2,0) ) = −0.040 Ry and
V (Q(3,1,1) ) = −0.080 Ry for Si (while V (Q(1,1,1) ) = +0.190 Ry, V (Q(2,2,0) ) =
−0.038 Ry and V (Q(3,1,1) ) = −0.035 Ry for Ge). Despite the different size of
the Si and Ge atoms, the pseudo-potentials are almost the same. This simi-
larity occurs since the differences between the atoms mainly occur close to the
cores where the differences in the true potentials are cancelled by the repulsive
contribution from the core wavefunctions. The higher order pseudo-potentials
may be neglected, since the pseudo-potential (like the full potential Vions is ex-
pected to show an overall decrease with Q−2 , for large Q.

The electronic bands of semiconductors with diamond and Zinc-Blende struc-


tures have been calculated by Chelikowsky et al.50 . To lowest order in the
pseudo-potential, a large magnitude gap should open up between the lowest
pair of bands at the L point. However, the lowest two bands at the X point
should remain degenerate for the elemental compounds, due to the vanishing of
the structure factor. For the binary compounds, the Fourier component of the
50 J. R. Chelikowsky, D. J. Chadi and M. L. Cohen, Phys. Rev. B, 8, 2786 (1973).

250
potential does not vanish at the X point, and so the degeneracy of the lowest
pair of bands is lifted. The direct band gap opens up between the next two high-
est sets of bands at the Γ point and these sets of bands do not cross, so the gap
between the valence and conduction bands persists throughout the entire Bril-
louin zone. However, the minimum energy separation between the valence and

1.2

1
µ
Ek [ Rydbergs ]

0.8

0.6

0.4

0.2

-0.2
Γ X W L Γ K X

Figure 144: The nearly-free electron dispersion relation for Si.

GaAs Structure

1.25

0.75
Ek [ Rydbergs ]

0.5

0.25

-0.25

-0.5
L Γ X W L Γ
(ka/2π)

Figure 145: The nearly-free electron dispersion relation for GaAs.

conduction bands occurs between the valence band at the Γ and the conduction
band at the X point. Since the minimum energy to excite an electron between
the valence and conduction bands requires a non-zero change in k, this threshold
energy is known as the indirect gap. The band splitting is mainly caused by the
mixing of the set of the six (2, 0, 0) free electron states with the set of the eight
(1, 1, 1) free electron states. Inspection of the secular equation shows that the
pseudo-potential causes a propensity for the (1, 1, 1) states directed parallel to

251
any one vector connecting the central atom to a neighboring atom in the local
atomic tetrahedra to combine with the linear superposition of (2, 0, 0) states
which has the same orientation. These un-normalized combinations are given

Diamond Structure

(-1/4,-1/4,1/4)

(1/4,1/4,1/4)

(0,0,0)

(-1/4,1/4,-1/4)

(1/4,-1/4,-1/4)

Figure 146: The local atomic tetrahedron in a diamond structure.

by

direction

(1, 1, 1) (2, 0, 0) + (0, 2, 0) + (0, 0, 2)


(−1, −1, 1) (2, 0, 0) + (0, 2, 0) − (0, 0, 2)
(−1, 1, −1) (2, 0, 0) − (0, 2, 0) + (0, 0, 2)
(1, −1, −1) (2, 0, 0) − (0, 2, 0) − (0, 0, 2)

The set of states are the symmetrized components the hybrid atomic (sp3 ) or-
bitals which form a local basis in the tight-binding representation. The appro-
priate basis states can then be combined to describe the uni-directional bonding
and anti-bonding states of the central and each of the four neighboring atoms
of the local atomic tetrahedron. The hybrid states localized on one of the two

252
inequivalent atoms51 , are denoted by

direction

(1, 1, 1) φs + φpx + φpy + φpz


(−1, −1, 1) φs − φpx − φpy + φpz
(−1, 1, −1) φs − φpx + φpy − φpz
(1, −1, −1) φs + φpx − φpy − φpz

and, when properly normalized, the hybrid wave functions on the atom form
an orthonormal set. The gap in the one-electron energy spectra of the com-
pounds occurs between the bonding and anti-bonding states, when the splitting
caused by the non-zero Fourier components of the pseudo-potential is larger
than the dispersion of the free electron bands. This inequality suggests that
the nearly-free electron picture is being pushed to the limits of its applicability.
The diamond form of C is stabilized and has a large gap since the 2p states
have quite similar energies to the 2s states and also since the pseudo-potential
is large, both of which favor the formation of the tetrahedrally bonded (sp3 )
hybrids.

The Brillouin zone boundaries of hexagonal close-packed materials are also


not fully gapped, due to the symmetry of the structure. The unit cell of the
reciprocal lattice of the (direct space) hexagonal closed packed lattice is a hexag-
onal prism. There are two hexagonal planes which have normals pointing along
the positive and negative z axis. These are Bragg planes. The structure factor
vanishes for the Q values which define the Bragg planes that are the hexagonal
top and bottom of the prism. The structure factor can be evaluated as
 
4 2
S(Q) = 1 + exp i π ( m1 + m2 + m3 ) (553)
3 3
which vanishes when m1 = m2 = 0 and m3 = ± 1, corresponding to the
Q value of these Bragg planes. The vanishing of the structure factor at these
particular Bragg planes is a consequence of a glide symmetry. In fact, group
theory shows that the splitting on these planes is rigorously zero in the absence
of spin-orbit coupling52 .

Since the gaps vanish on some faces of the Brillouin zones, it is sometimes
helpful to define a set of zones, called the Jones zones53 , which are enclosed by
51 The two atoms in the primitive unit cell are not equivalent since their local atomic tetra-
hedra have different orientations.
52 C. Herring, Phys. Rev. 52, 361 (1937).
53 H. Jones, Theory of Brillouin Zones and Electronic States in Crystals, North Holland

Publishers, Amsterdam (1960).

253
planes in which gaps do occur. For example, the hexagonal Bravais lattice has
a hexagonal Brillouin zone which is bounded by a set of eight Bragg planes54 .
For the hexagonal close-packed structure, the gaps in the electronic dispersion
relations either vanish or are small on the {001} Bragg planes. One can construct
a Jones zone for the h.c.p. structure which is based on the hexagonal Brillouin
zone but, instead of being bounded by the {001} planes, it is bounded by the
{002} Bragg planes. This Jones zone has twice the volume of the Brillouin zone
and, therefore, contains two states for each unit cell in the crystal. However,
the sets of {101}, {011} and {111} Bragg planes intersect with the zone, so
truncating the above zone with these twelve equivalent planes leads to another
Jones zone with a smaller volume and a more spherical shape. This minimal
volume Jones zone is bounded by a set of twenty Bragg planes55 and contains
a total of  2  4
3 a 3 a
2 − + (554)
4 c 16 c
states per unit cell. The minimal volume Jones zone for the h.c.p. structure is
shown in fig(147).

(002)

(101)
(0-11)
(1-11)

(100) (4π/c)
(0-10) (1-10)

(0-1-1) (10-1)
(1-1-1)

4/3(2π/a)

Figure 147: The minimal volume Jones zone for the hexagonal close-packed
lattice.

The Jones zone concept can be applied to explain the stability of the γ-phases
of Hume-Rothery alloys. The γ-phases have complex cubic structures with
fiftytwo atoms in the unit cell56 . The calculated magnitudes of the structure
54 If the reciprocal lattice vectors are represented by
  
2π (m1 − m2 ) a
Q= √ ê1 + (m1 + m2 ) ê2 + m3 ê3
a 3 c

the hexagonal Brillouin zone is bounded by the sets of {100}, {010}, {110} and {001} Bragg
planes.
55 The minimal volume h.c.p. Jones zone is bounded by twenty Bragg planes which are the

six {100}, {010} {110} planes, the twelve {101}, {011}, {111} planes and the two {002} Bragg
planes
56 A. J. Bradley and J. Thewlis, Proc. Roy. Soc. 112, 678 (1926).

254
Jones zone for Bi

(2,1,2)
(2,2,1)

(1,-1,0) (1,0,-1) (0,1,-1)

(-1,-2,-2)

Figure 148: The Jones Zone for solid Bi.

factor S(Q) at the reciprocal lattice vectors Q are given by:

Q (110) (200) (211) (220) (310) (222) (321) (400) (330) (411)

|S(Q)| 0.32 0 0.32 0 0.32 2.68 1.05 0 8.85 5.63

As seen in fig(149), the structure factors are large at for the (330) and (441)
reciprocal lattice vectors. Therefore, one expects large gaps to open up at the
ASAHI et al. PHYSICAL REVIEW B 71, 16

FIG. 2. Total DOS derived from the FLAPW band


for the Cu5Zn8 ␥ brass.

FIG. 1. Diffraction spectrum taken at Spring-8 synchrotron ra- In the present work, we performed the ab init
Figure 149: X-ray diffraction
diation facility and pattern fromforγ-phase
Rietveld fitting the Cu9AlCu 9 Al4
4 brass. . [After
The inset Asahi et al.
band calculations for the Cu5Zn8 and Cu9Al4 ␥
(2005).] shows its enlargement at low diffraction angles. The Miller indices single out the Fermi surface-Brillouin zone inte
with asterisk indicate diffraction lines observed in the Cu9Al4 with sponsible for the formation of the pseudogap at
space
twelve (330) and thegroup P4̄3m but not
twentyfour in theBragg
(411) Cu5Zn8planes.
brass withThe
spacetwo
groupsets level. As described below, we demonstrate why t
of equiva-
I4̄3m.
lent Bragg planes encloses a region around the origin which is nearly spherical, and Cu 9Al4 ␥ brasses are commonly stabilized
value of 21/ 13 in spite of the entirely different s
past for approximants. The ␥ brass is also known as a struc- centrations. This is, we believe, the first strai
turally complex alloy phase stabilized at specific electron demonstration for the Hume-Rothery electron co
255 below, the unit cell is just
concentrations and, as described rule in a structurally complex alloy phase.
small enough to be handled with the FLAPW band calcula-
tions.
The ␥ brass is characterized by the possession of a com- II. ELECTRONIC STRUCTURE CALCULAT
plex cubic structure containing 52 atoms in the unit cell and A. Atomic structure
formed in many systems at the electron per atom ratio e / a of
21/ 13= 1.615.13–16 Mott and Jones17 discussed why the ␥ The ␥ brass of both Cu5Zn8 and Cu9Al4 contain
brass is stabilized at the specific value of e / a by assuming in the conventional body-centered-cubic unit cell
simultaneous contacts of the free-electron Fermi sphere with constants of 8.84 and 8.675 Å, respectively. Ho
since it is has thirtysix faces. This Jones zone contains ninety electronic states
for each unit cell in the crystal. In an extended zone scheme, the Jones zone
would be fully occupied if an almost spherical Fermi-surface coincided with the
boundaries of this region. In this case, the number of electrons per atom should
be n = 9052 and the structure would be stabilized by the relatively large values
Vions (Q) at the boundaries.

Figure 150: The Jones Zone for the γ-phase of Cu − Zn alloys.

The spin-orbit interaction can lead to the re-occurrence of small gaps in the
bands57 . The spin-orbit interaction is a relativistic effect, which appears as a
low order correction to the non-relativistic limit of the Dirac equation. For a
particle of charge q in the presence of a scalar and vector potential (φ, A), this
process yields the single-particle Hamiltonian in the form
 2
1 q
Ĥ = m c2 + (p − A).σ + qφ
2m c
q h̄3
 
1 q h̄ q
− p4 + σ . ∇ φ ∧ ( p − A ) + ∇2 φ
8 m3 c 2
4m c 2 c 8 m2 c2
57 M. H. Cohen and L. Falicov, Phys. Rev. Lett. 5, 544 (1960).

256
(555)

The first line, apart from the rest energy, coincides with the non-relativistic
Pauli Hamiltonian
 2
1 q
ĤP = (p − A).σ + σ0 q φ (556)
2m c
which, together with the identity
      
σ.a σ.b = σ0 a . b + iσ. a ∧ b (557)

leads to
  2
1 q
ĤP = σ0 − i h̄ ∇ − A + σ0 q φ
2m c
 
h̄ q
− σ. ∇ ∧ A + A ∧ ∇ (558)
2mc
Furthermore, on using
 
∇ ∧ A Ψ(r) = Ψ(r) ∇ ∧ A − A ∧ ∇ Ψ(r) (559)

and on noting that B = ∇ ∧ A, one obtains the non-relativistic Pauli


Hamiltonian including the anomalous Zeeman interaction
  2 
1 q h̄ q
ĤP = σ0 p − A − σ . B + σ0 q φ (560)
2m c 2mc
Thus, all the terms in the first line of equation (555) are found in the non-
relativistic theory whereas the terms in the second line represent interactions,
Ĥrel , which have a relativistic origin. The relativistic terms are given by

q h̄3
 
1 4 q h̄ q
Ĥrel = − p + σ . ∇ φ ∧ ( p − A ) + ∇2 φ
8 m3 c 4 m2 c2 c 8 m2 c2
(561)

The first term which is proportional to p4 represents a relativistic correction


to the kinetic energy. The next term is the spin-orbit interaction which can be
interpreted as being caused by the interaction of the spin with the magnetic field
produced by the electron’s own orbital motion. The last term is the Darwin
term, which is often discussed as an interaction with a classical electron of
finite spatial extent. Thus, the spin-orbit interaction for an electron is truly
a relativistic effect and, unlike the other relativistic corrections, is not very
symmetric. It is given by the pseudo-scalar interaction
q h̄
− σ.(v ∧ E) (562)
4 m c2

257
For systems with an inversion symmetry, the spin-orbit interaction does not
lift the two-fold spin-degeneracy of the Bloch states. This can be seen by con-
sidering time reversal symmetry58 . The Hermitean nature of the Hamiltonian
implies that if the Bloch state with wave vector −k
   
1 χ+
φn,−k (r) = √ exp − i k . r un,−k (r) (563)
V χ−

is an eigenfunction with eigenvalue En,−k , then so is the time reversed state

−χ∗−
   
1
φTn,−k (r) = √ exp i k . r u∗n,−k (r) (564)
V χ∗+

The time reversed state has a spin state which is orthogonal to the original one.
This is a Bloch state with vector k. Since the system has a center of inversion,
then    
1 χ+
φn,−k (−r) = √ exp i k . r un,−k (−r) (565)
V χ−
is also Bloch state φn,k with wave vector k and has the same energy En,−k and
spin state as the original state φn,−k . Hence, for a system with a center of
inversion, the Bloch state with energy En,k must be at least doubly-degenerate
(spin-degenerate). It also follows that for every Bloch state φn,k with spin σ,
there is a state φn,−k with the same spin and the same energy. This is known
as Kramers’ theorem.

Due to its reduced symmetry, the spin-orbit interaction raises the degener-
acy of the bands at high-symmetry points in k space59 , such as those on the
hexagonal faces of the h.c.p. Brillouin zone.

——————————————————————————————————

8.1.11 Exercise 39
The effect of the Bragg planes on the density of states can be calculated from
the nearly-free electron model. For simplicity, consider the effect of one Bragg
plane. The Bloch wave vector k is resolved into components parallel, k k , and
perpendicular, k ⊥ , to the reciprocal lattice vector Q

k = k⊥ + kk (566)
58 The time reversal operator τ̂ can be represented by τ̂ = i σy Ĉ where σy is the Pauli
matrix and Ĉ is the complex conjugation operator.
59 R. J. Elliott, Phys. Rev. 96, 280 (1954).

258
The energy of the two bands can be written as

h̄2 2
Ek,± = k + ∆E± (kk ) (567)
2m ⊥
where
h̄2
  
2 1 2
∆E± (kk ) = kk + Q − 2 kk Q
2m 2
 2   2 ! 12
h̄ 2 2
± Q − 2 kk Q + | V (Q) | (568)
4m

describes the splitting of the two bands. (Note that the band energies are not
periodic in kk . This is a consequence of our artificial assumption that there is
only one Bragg plane.) For each band, the density of state per spin is
Z
V
ρ± () = d3 k δ(  − Ek,± ) (569)
( 2 π )3

Show that the contribution to the density of states from each band is of the
form    
2m V
kmaxk () − kmink () (570)
h̄2 4 π2
where an equation of the form  = ∆E± (kmk ) defines the maximum and min-
imum value of kk .

Show that, if the constant energy surface cuts the Brillouin zone boundary,
i.e.,
(0) (0)
E Q − | V (Q) | ≤  ≤ E Q + | V (Q) | (571)
2 2

then for the lower band, one has


Q
kmaxk () = (572)
2
and if the constant energy does not intersect the Brillouin zone boundary then
r
2m
kmaxk () = + O(|V (Q)|2 ) (573)
h̄2
(0)
where E Q − | V (Q) | ≥  ≥ 0.
2

Show that the density of states for the upper band is given by
  
V m Q (0)
ρ+ () = − kmink () for  ≥ E Q + | V (Q) | (574)
4 π 2 h̄2 2 2

259
∂ρ
Show that the energy derivative of the density of states, ∂ , is singular at
the energies
(0)
 = E Q ± | V (Q) | (575)
2

——————————————————————————————————

8.1.12 Exercise 40
Consider the point W on the Brillouin zone boundary of an f.c.c. crystal. Three
Bragg planes meet at W. The k value at W is
 
2π 1
kW = (1, , 0) (576)
a 2

The three Bragg planes correspond to the reciprocal lattice vectors QX =


2 π 2 π 2 π
a (2, 0, 0), QL = a (1, 1, 1) and QL0 = a (1, 1, 1). The four free electron
energies are

(0) h̄2 2
EΓ,k = k
2m
2
h̄2

(0)
EL,k = k − QL
2m
2
h̄2

(0)
EL0 ,k = k − QL0
2m
2
h̄2

(0)
EX,k = k − QX (577)
2m
(0) h̄2
These four energies are degenerate at W and are equal to EW = 2 m k 2W .

Show that near W , the first order energies are given by the solutions of

E (0) − E

Γ,k V1 V1 V2



(0)
EL,k − E

V1 V2 V1

= 0

(0)

V1 V2 EL0 ,k − E V1



(0)
V2 V1 V1 EX,k − E

260
where V2 = V (Qx ) and V1 = V (QL ) = V (QL0 ), and that at W the roots
are
(0)
E = EW − V 2 doubly degenerate

(0)
E = EW + V 2 ± 2 V 1 singly degenerate

Two Bragg planes meet on the line W U W , where the point U is the
midpoint of the edge of the square face. The point U corresponds to the k value
 
2π 1 1
kU = (1, , ) (578)
a 4 4
Show that for points k which are on the line W U W and are close to U , the
band energies are given by
(0)
E = Ek − V2

V2 1
q
(0)
E = Ek + ± V22 + 8 V12 (579)
2 2
where
(0) h̄2 2
Ek = k (580)
2m
is the free electron energy at k.

——————————————————————————————————

8.1.13 Exercise 41
Consider a nearly-free electron band structure near a Bragg plane. Let
Q
k = + q (581)
2
and resolve q into the components q k and q ⊥ parallel and perpendicular to the
Q
Bragg plane 2. Then, the energy bands are given by
 12
h̄2 2 h̄2 2

(0) (0)
E = EQ + q ± 4 EQ q + | V (Q) |2 (582)
2 2m 2 2m k
It is convenient to express the Fermi energy µ in terms of the energy of the
lower band at the Bragg plane
(0)
µ = EQ − | V (Q) | + ∆ (583)
2

261
Show that when 2 V (Q) > ∆ > 0, then the Fermi surface is only composed
of states in the lower Bloch band. Furthermore, show that the Fermi surface
intersects the Bragg plane in a circle of radius ρ where
r
2m∆
ρ = (584)
h̄2

Show that, if ∆ > 2 | V (Q) |, the Fermi surface cuts the Bragg plane in
two circles of radius ρ1 and ρ2 such that the area between them is
 
2 2 4πm
π ρ1 − ρ2 = | V (Q) | (585)
h̄2
This area is measurable through de Haas - van Alphen experiments.

——————————————————————————————————

8.1.14 Exercise 42
In a weak periodic potential the Bloch states in the vicinity of a Bragg plane
can be approximated in terms of two plane waves.

Let k be a wave vector with polar coordinates (θ, ϕ) in which the z axis
is taken to be the direction Q of the reciprocal lattice vector that defines the
Bragg plane.
 2
h̄2 Q
(i) If E < 2 m 2 , show that to order V (Q)2 the surface of energy E
is given by r  
2mE
k(θ, ϕ) = 1 + δ(θ) (586)
h̄2
where
| V (Q) |2
m E
δ(θ) = √ (587)
h̄2 Q2 − 2 h̄ Q cos θ 2mE

(ii) Show that | V (Q) |2 results in a shift of the Fermi energy given by

∆µ = µ − µ0 (588)

where
1 | V (Q) |2
 
2 kF Q + 2 kF
∆µ = − ln (589)
8 µ0 Q Q − 2 kF

262
——————————————————————————————————

8.1.15 Exercise 43
Consider an energy E which lies within the gap between the upper and lower
bands at point k on the Bragg plane which is defined by the reciprocal lattice
vector Q. Let
Q
k = + q (590)
2

(i) Find an expression for the imaginary part of k for E within the gap.

(ii) Show that for E at the center of the gap, the imaginary part of k satisfies
2 s 
2 2
Q2

Q2

2m
=m k = − ± + 2 V (Q)
(591)
2 2 h̄
Q
Thus, on solving for k given E, there is a range of =m k when <e k = 2.

Complex wave vectors are important for the theory of Zener tunnelling be-
tween two bands, caused by strong electric fields. Complex wave vectors also
occur in the description of states that are localized near surfaces.

——————————————————————————————————

263
8.2 The Pseudo-Potential Method
The failure of the nearly-free electron model is primarily due to the large values
of the potentials, V (Q), calculated from first principles, and the small values of
the experimentally observed splittings between the bands. Due to the large value
of the lattice potential, if the wave functions are expanded terms of plane waves
then very many plane waves (of the order of 106 ) are needed to obtain conver-
gence. Furthermore, band structure calculations with the exact lattice potential
are expected to reproduce the entire set of wave functions ranging from the core
wave functions located within the ions, up to the valence and/or conduction
wave functions. Since the core electrons are very localized and almost atomic, a
large number of plane waves are needed for an accurate calculation of the core
wave functions. Large numbers of plane waves are also needed to calculate the
valence band wave functions. The need for a large number of Fourier compo-

Vatom(r) = - 14 e2/r Si: 1s2 2s2 2p6 3s2 3p2

core 1s
2s
2p
3s
3p

Figure 151: The spatial separation of the core and the valence electrons in a Si
atom (schematic).

nents to calculate the valence band wave functions can be understood by the
consideration of the fact that the conduction or valence band states have to be
orthogonal to the wave functions of the core electrons. Thus, the conduction
electrons should have wave functions that exhibit rapid oscillations in the vicin-
ity of the ion cores. Historically, there have been many methods which were
used to avoid the need to use many plane waves. The methods used range from
orthogonalized plane waves, augmented plane waves and pseudo-potentials. All
these methods have some common features, namely the feature of producing
wave functions that require fewer plane wave components in the expansion and,
thereby, increase the rate of convergence, and concomitantly diminish the effect
of the ionic potential. The pseudo-potential method provides a first principles
way of explicitly finding a smaller effective potential.

The electrons in the valence band move in a periodic potential Vions (r) pro-
vided by the ions. The ionic potential already includes a partial screening of

264
the nuclear potential by the ion core electrons.

The valence band Bloch functions φvk,n (r) undergo many oscillations in the
region of the core as they must be orthogonal to the core electron wave functions
φck,α (r). In the Dirac notation, the orthogonality condition is expressed as

< φvk,n | φck,α > = 0 (592)

The valence band Bloch function can be expressed in terms of a smooth function
v
ψk,n (r) (593)

that doesn’t contain the oscillations that orthogonalize the Bloch state, | φvk,n >,
with the core wave states. The smooth function is known as the pseudo-wave
function. The pseudo-wave function is related to the valence band Bloch func-
tion by the definition
X
| φvk,n > = | ψk,n
v
> − | φck,α > < φck,α | ψk,n
v
> (594)
α

This definition automatically ensures the othornomality of the core states with
the valence band states without placing any restriction on the form of the
pseudo-wave function. The basic idea behind pseudo-potential theory is that
the smooth pseudo-wave function represents the electronic wave function in the
region between the cores, and may be expressed in terms of only a few plane
wave components60 .

Since the Bloch state, | φvk,n >, satisfies the one-particle Schrodinger equa-
tion
Ĥ | φvk,n > = Ek,n v
| φvk,n > (595)
one finds that the smooth function satisfies
X
v
Ĥ | ψk,n > − Eαc | φck,α > < φck,α | ψk,n
v
> =
α
 X 
v v
= Ek,n | ψk,n >− | φck,α > < φck,α | ψk,n
v
>
α
(596)

This equation can be re-arranged to yield an eigenvalue equation for the (un-
known) smooth function, which has the same energy eigenvalues as the exact
eigenfunction. The rearranged equation has the form
 
v v v v
Ĥ + V̂ (Ek,n ) | ψk,n > = Ek,n | ψk,n > (597)

60 J. C. Phillips and L. Kleinman, Phys. Rev. 116, 287 (1959).

265
where
X  
v v c
V̂ (Ek,n ) = Ek,n − Ek,α | φck,α > < φck,α | (598)
α
is a non-local and energy-dependent contribution to the potential. The impor-
tant point is that this potential may be regarded as being positive and, therefore,
counteracts the effect of the large negative potential due to the ions. This can
be seen by taking the expectation value of the energy dependent potential in
any arbitrary state | Ψ >
X  
v v c
< Ψ | V̂ (Ek,n )|Ψ > = Ek,n − Ek,α | < Ψ | φck,α > |2 (599)
α
v
and as the valence electrons have a higher energy than the core electrons, Ek,n >
c
Ek,α , one finds
v
< Ψ | V̂ (Ek,n )|Ψ > ≥ 0 (600)
Thus, the potential operator is effectively positive as it increases the expectation
value of the energy for an arbitrary state.

The operator V̂ is non-local. This can be seen by considering the action of V̂


on an arbitrary wave function Ψ(r). The operator has the effect of transforming
the state through
X   Z
v
V̂ (Ek,n ) Ψ(r) = v
Ek,n c
− Ek,α d3 r0 φ∗c 0 0 c
k,α (r ) Ψ(r ) φk,α (r) (601)
α V

Thus, the operator when acting on the wave function at position r changes the
position to r0 .

If the original one-particle Schrodinger equation for φvk,n (r) has the form

h̄2
 
− ∇ + Vions (r) φvk,n (r) = Ek,n
2 v
φvk,n (r) (602)
2m
v
then the Schrodinger equation for the smooth function ψk,n (r) has the form

h̄2
 
− ∇2 + Vions (r) + V̂ (Ek,nv v
) ψk,n (r) = Ek,nv v
ψk,n (r) (603)
2m
The Schrodinger equation for the smooth wave function has exactly the same
energy eigenvalues as the original potential. The pseudo-potential is defined as
v
V̂pseudo = Vions (r) + V̂ (Ek,n ) (604)
and, as has been shown, the effect of the pseudo potential is much weaker than
v
that of Vions (r). Also as the eigenstate ψk,n (r) is a smooth function it can be
expanded in terms of a few planes waves
 
v
X 1
ψk,n (r) = Ck−Q √ exp i ( k − Q ) . r (605)
Q
V

266
Thus, the pseudo-potential may be treated as a weak perturbation and gives
results very similar to those of the nearly-free electron model. Since its energy

0.75
1s
0.5

3s
r Rnl(r)

0.25

0
0 0.5 1 1.5 2 2.5 3 3.5 4

80%
-0.25

2s Rc RWS
-0.5

r/a

Figure 152: The schematic spatial variation of the ns atomic wave functions
for Al. X-ray scattering reveals that the core electrons are localized within
Rc ≈ 1.1 atomic units and the linear dimension of the unit cell is given by the
Wigner-Seitz radius RW S ≈ 3 a.u. It is seen that roughly 80 % of the atomic
3s electrons would reside outside the unit cell, and hence can be described as
nearly-free electrons.

dependence is weak, E v can be set to zero in the pseudo-potential.

There are many different forms that a pseudo-potential can take61 .

8.2.1 The Pseudo-Potential Theorem


If the valence band Bloch functions | φvk,n > satisfy the energy eigenvalue
equation
v
Ĥ | φvk,n > = Ek,n | φvk,n > (606)
then one can find a pseudo-wave function | ψk > which satisfies the energy
eigenvalue equation
v
Ĥef f | ψk > = Ek,n | ψk > (607)
with the same eigenvalue as the valence states but in which the effective Hamil-
tonian Ĥef f has the form
X
Ĥef f = Ĥ + | φck,α > < Fα | (608)
α
61 B. J. Austin, V. Heine and L. J. Sham, Phys. Rev. 127, 276 (1962).

267
and the pseudo-wave function is given by
X
| ψk > = | φvk,n > + aα | φck,α > (609)
α

Proof

On assuming that the conduction band states and the valence states form a
complete orthonormal set, one can expand the pseudo-wave function as
X X
| ψk > = an0 | φvk,n0 > + aα0 | φck,α0 > (610)
n0 α0

where an0 and aα0 are coefficients that are to be determined. On substituting
this form in the energy eigenvalue equation
v
Ĥef f | ψk > = Ek,n | ψk > (611)

one finds
X X
an0 Ĥ | φvk,n0 > + aα0 Ĥ | φck,α0 >
n0 α0
X X
+ an0 | φck,α > < Fα | φvk,n0 > + aα0 | φck,α > < Fα | φck,α0 >
α,n0 α,α0
 X X 
v
= Ek,n an0 | φvk,n0 > + aα0 | φck,α0 > (612)
n0 α0

On taking the matrix elements of the above equation with the valence band
state < φvk,n00 | and using the orthogonality relations between the valence and
core Bloch functions, one finds the condition
X  
v v
Ek,n00 − Ek,n an0 δn00 ,n0 = 0 (613)
n0

where the sum over n0 can be trivially performed. The above condition is auto-
matically satisfied independent of an if n00 = n. On the other hand, for n00 6= n,
one has an00 = 0. Hence, the expansion of the pseudo-wave function with the
v
energy eigenvalue Ek,n only contains the valence band Bloch function with band
index n.

On taking the matrix elements of eqn(612) with the core state < φck,α00 |,
using the orthogonality relations and the condition an0 = δn0 ,n , one finds
X  X
c v
Ek,α 00 − Ek,n aα0 δα00 ,α0 = − < Fα00 | φvk,n > − aα00 < Fα00 | φck,α >
α0 α
(614)

268
or
 X 
c v
Ek,α00 − Ek,n + < Fα00 | φck,α > aα00 = − < Fα00 | φvk,n > (615)
α

which determines a in terms of | Fα00 > or vice versa. Either the choice of
α00
Fα is arbitrary or the choice of aα is arbitrary.

There are many common choices made for Fα . Previously, we made the
choice
c
| Fα > = ( E − Ek,α ) | φck,α > (616)
Austin, Heine and Sham made an alternate choice:
| Fα > = V̂ | φk,α > (617)
which corresponds to the pseudo-potential
 X 
V̂pseudo = V̂ ˆ
I − c c
| φk,α > < φk,α | (618)
α

which, if the cores wave functions were complete within the core region of space,
would result in the effective potential being zero within this region.

The non-local pseudo-potential can be approximated by a local potential. In


this approximation, the pseudo-potential is almost zero within the core. This is
a result of the so-called cancellation theorem62 .

8.2.2 The Cancellation Theorem


The Austin, Heine and Sham potential is optimal in the sense that the cancella-
tion is most complete as the kinetic energy is as close as possible to the energy
eigenvalue. This can be seen by minimizing the functional
Z  2 
3 h̄ 2 2
Λ[ψ] = d r | ∇ ψ | − Ek,v | ψ | (619)
2m
w.r.t to the coefficients aα . For a pseudo-wave function ψ, the above functional
is evaluated as
 X 
c
Λ[ψ] = − < ψ | V̂ + | φk,α > < Fα | | ψ >
α

= − < ψ | V̂pseudo | ψ > (620)


Hence, if we minimize Λ[ψ] the cancellation is optimal. We shall make a variation
w.r.t. the coefficients aα , so
X
δψ = δaα φk,α (621)
α
62 M. H. Cohen and V. Heine, Phys. Rev. 122, 1821 (1961).

269
Hence, the first-order variation in Λ[ψ], δΛ(1) is given by
X X
δΛ(1) = − δa∗α < φck,α | V̂pseudo | ψ > − δaα < ψ | V̂pseudo | φck,α >
α α
(622)
Since the general form of the operator V̂pseudo is not a Hermitean, the extremal
condition δΛ(1) = 0 for arbitrary δaα yields

< ψ | V̂ | φck,α > + < ψ | Fα > = 0 (623)

which is satisfied if
| Fα > = − V̂ | φck,α > (624)
Thus, the requirement that the cancellation is most complete reduces the pseudo-
potential to the Austin, Heine and Sham pseudo-potential.

The cancellation theorem can be understood through classical considera-


tions. Classically, the gain in kinetic energy of a conduction electron as it enters
the core region is equal to the potential energy. As the oscillations in φck,α (r)
give rise to the kinetic energy of the electron in the core region, one expects the
pseudo-potential to cancel in the core region. Therefore, the pseudo-potential
follows the ionic core potential for distances larger than the ionic core radius
Rc , at which point the attractive potential almost shuts off. The empty core
Empty Core Approximation Pseudopotential
Vps(r)

-Z e2/r

Rc r

Figure 153: The empty core approximation for the atomic pseudo-potential.

approximation to the atomic pseudo-potential63 is given by

Z e2
Vpseudo (r) = − for r > Rc
r

Vpseudo (r) = 0 for r < Rc


(625)
63 N. W. Ashcroft, Phys. Lett. 23, 48 (1966).

270
Basically, this is a reflection of the fact that the valence electrons do not probe
the region of the cores as this region is already occupied by the core electrons
and the Pauli exclusion principle forbids the overlap of states.

The Fourier transform of the atomic pseudo-potential is a smooth function


of the wave vector q.

4 π Z e2
Vpseudo (q) = − cos q Rc (626)
q2
Only the values of Vpseudo (q) at the reciprocal lattice vectors Q are physically
important, and Vpseudo (Q) are small at most of Q. When one includes the effect
of the screening electron clouds, the pseudo-potential is replaced by the screened
pseudo-potential
Z e2
 
Vpseudo (r) = − exp − kT F r for r > Rc
r

Vpseudo (r) = 0 for r < Rc (627)

The Fourier transform of the screened pseudo-potential is given by


4 π Z e2
Vpseudo (q) = − cos q Rc (628)
q 2 + kT2 F
which is weakened with respect to the original potential. The pseudo-potential

Screened Pseudopotential in Momentum Space


0.2

0
q0
Vps(q)/EF

-0.2

-0.4

-0.6

-0.8
0 0.5 1 1.5 2
q/2kF

Figure 154: The q-dependence of the screened atomic pseudo-potential.

for the crystalline solid is equal to the Fourier transform of the atomic pseudo-
potential multiplied by the structure factor times the inverse of the volume of a
unit cell. For a lattice with a one atom basis, this yields the empirical potential
N 4 π Z e2
Vions (q) = − cos q Rc (629)
V q 2 + kT2 F

271
The weakening of the potential occurs mainly in the q region around the q0 where
the potential changes sign. That is, the pseudo-potential method is most efficient
for systems where the most important reciprocal lattice vectors have magnitudes
similar to q0 . For example, in Al, one finds that the pseudo-potential at the
(2,0,0) reciprocal lattice vector is given by Vions (2, 0, 0) ≈ + 0.5 eV, whereas the
Fourier transform of the bare lattice potential is of opposite sign and is an order
of magnitude larger at the same reciprocal lattice vector, Vions (2, 0, 0) ≈ − 5
eV. The periodic oscillations found in the Ashcroft empty core pseudo-potential
are spurious, since they are due to the sharp cut-off. Nevertheless, changes in
sign are also found in other pseudo-potentials64 . In these cases, the oscillations
are rapidly damped so that only a few changes in sign occur. For example, on
treating the atomic 1s wave functions of Li as hydrogenic-like, then one finds
an unscreened pseudo-potential of the form

4 π Z e2 F ∗ (k + Q) F (k + Q0 )
Vions (Q, Q0 ) = − + ( E − E 1s ) (630)
Vc (Q − Q0 )2 Vc

where Vc is the volume of a unit cell of the Bravais lattice and


r
κ3 8πκ
F (q) = (631)
π ( q 2 + κ2 )2

in which κ represents the inverse radial dimension of the 1s electron core. The
non-locality of the pseudo-potential shows up as an extra dependence on k.
This extra k dependence can be ignored for s states, since the core electrons
only occupy a tiny fraction of the unit cell volume (about 6 % for Al). That
means that the product κ a is sufficiently large (κ a ≈ 14 for Li), so that there
is only a small variation in F (q) when q is varied.

8.2.3 The Scattering Approach


The pseudo-potential is a potential that gives the same eigenvalues as the full
potential Vions (r), for the valence electron states. The pseudo-potential may be
obtained from scattering theory.

Consider a single ionic scattering center with a spherically symmetric poten-


tial V (r) which is zero for r > R. Then for r > R, the radial wave function
has the asymptotic form
 
Rl (r, E) = Cl jl (kr) − tan δl ηl (kr) (632)

where
h̄2 k 2
E = (633)
2m
64 A. E. O. Animalu and V. Heine, Phil. Mag. 12, 1249 (1965).

272
0.2

(1,1,0) (2,0,0)
Li

Vions(q) [ Rydbergs ]
0.1

(2,1,1)
0 (2,2,0)

-0.1 εF = 0.3492 Ryd


kF a = (6π2)1/3
-0.2

−2/3 ε F q0 q1
-0.3
0 0.5 1 1.5 2 2.5
q/2kF

Figure 155: The q dependence of the pseudo-potential for Li. The reciprocal
lattice vectors Q for the b.c.c. structure are denoted by the red triangles.

and jj (x) and ηl (x) are the spherical Bessel and Neumann functions. The coef-
ficients Cl and the phase shifts δl (E) are obtained by matching the asymptotic
form to the solution at some large distance r = R. The exact logarithmic
derivative of Rl (r, E) at r = R can be defined as
Rl0 (R, E)
Ll (E) = (634)
Rl (R, E)
The matching condition of the logarithmic derivative of the asymptotic form
with the logarithmic derivative of the wave function at r = R leads to the
equation
jl (kR) Ll (E) − k jl0 (kR)
tan δl (E) = (635)
ηl (kR) Ll (E) − k ηl0 (kR)
The phase shifts δl (E) determine the scattering amplitude f (θ, E) for a particle
of energy E to be scattered through an angle θ. Partial wave analysis yields the
relation
   
1 X
f (θ, E) = ( 2 l + 1 ) exp 2 i δl − 1 Pl (cos θ) (636)
2ik
l

The scattering amplitude only depends on the phase shift modulo π. The phase
shift can always be restricted to the range − π2 to + π2 by defining

δ l = nl π + ∆ l (637)

where nl is an integer chosen such that


π
| ∆l | < (638)
2

273
The value of nl denotes the number of the oscillations in the radial wave func-
tion Rl (r, E). The (truncated) phase shifts ∆l produce the same scattering
amplitude as the original phase shift δl (E).

The atomic pseudo-potential is defined as any potential in which the com-


plete phase shifts are the truncated phase shifts ∆l and, thus, gives rise to the
same scattering amplitude, but does not produce any bound states (according
to Levinson’s theorem). The pseudo-radial wave functions R̃l (r, E) have no
nodes and, thus, have no rapid oscillations. Therefore, the pseudo-radial wave
function can be represented in terms of a finite superposition of plane waves of
long wave length. The pseudo-potential actually only depends on the function
Ll (E). From the knowledge of logarithmic derivative, Ll (E), one can construct
the pseudo-potential. One method has been proposed by Ziman and Lloyd.

8.2.4 The Ziman-Lloyd Pseudo-potential


Ziman and Lloyd independently65 proposed a pseudo-potential which is local
in r and is zero everywhere except on the surface of a shell of radius R. The
potential operator, V̂ ZL , is written as
X
V̂ ZL = Bl (E) δ( r − R ) P̂l (639)
l

where P̂l projects onto the states with angular momentum l. Inside the sphere
the potential is zero and so the radial wave function is just proportional to
jl (kr), since the Neumann function is excluded due to the boundary condition
at r = 0. The amplitude Bl (E) is chosen so as to give the proper asymptotic
properties of the wave function of the true potential V , for r > R.

The pseudo-radial wave functions satisfy the radial Schrodinger equation,


given by

h̄2 1 ∂
   2 
2 ∂ h̄ l ( l + 1 ) ZL
− r R̃ l + + V (r) R̃l (r) = E R̃l (r)
2 m r2 ∂r ∂r 2 m r2
(640)
The derivative of the pseudo-radial wave function is found by integrating the
Radial Schrodinger equation over the shell at r = R
R+
h̄2 ∂
− R̃l (r) + Bl (E) R̃l (R) = 0 (641)
2 m ∂r R−

The pseudo-wave function is matched with the true wave function at the radius
r = R+ . The matching condition determines the function Bl (E) in the pseudo-
potential in terms of the logarithmic derivative of the true wave function, Ll (E).
65 J. M. Ziman, Proc. Phys. Soc. (London) 86, 337 (1965), P. Lloyd, Proc. Phys. Soc.

(London), 86, 825 (1965).

274
Thus, the coefficient Bl (E) is related to Ll (E) via

jl0 (kR) 2m
Ll (E) − k = Bl (E) (642)
jl (kR) h̄2
Therefore, the Bl (E), for different l, are determined in terms of the exact value
of logarithmic derivatives. The projection operator is simply given as
X
P̂l = | l, m > < l, m | (643)
m

which also gives rise to the non-locality of the pseudo-potential operator. The
pseudo-potential for the solid can be constructed as a superposition of the
pseudo-potentials of the ions.

It should be noted that the pseudo-potential only cancels for states of an-
gular momentum l if there are core states with angular momentum l otherwise,
the electrons experience the full potential. Thus, in C the 2s electron ex-

0.4
l=2 l=0
n=3

0.2
r R3,l(r)

0
0 4 8 12 16 20

-0.2
l=1

-0.4
r/a0

Figure 156: The schematic spatial variation of the atomic wave functions and
pseudo-wave functions, for various values of the angular momentum. The pseudo
wave functions are depicted by the dashed lines and are nodeless.

perience the cancelled pseudo-potential but the 2p electrons interact with the
full potential. The 2p electrons are relatively tightly bound compared with the
2s. Thus, the s → p promotion energy is lower than in the other group IV
elements Si, Ge, Sn and P b. This allows C to easily form the tetrahedrally
directed sp3 valence bonds and, therefore, is partly responsible for its ability
to form the diamond structure. Similarly, in the 3d transition metals, the 3d
electrons are tightly bound compared with the 4d or 5d electrons in the second
and third series. Thus, the 3d electrons form tightly bound narrow bands, and
pseudo-potential theory is inappropriate.

275
3

l=0

0
l=1

Vl(r)
-3

- Z e2/r

-6
0 2 4 6 8 10 12
r/a0

Figure 157: The schematic spatial-dependence of the (norm-conserving) atomic


pseudo-potential, for various values of the angular momentum l.

In summary, the pseudo-potentials can be created from first principles and


then, if the pseudo-potential is weak enough, the nearly-free electron model can
be used to obtain the results for the valence bands of real solids.

——————————————————————————————————

8.2.5 Exercise 44
An electron outside a hydrogen atom with a 1s core state is treated by the
pseudo-potential method. Calculate the Bloch wave function for an electron
which has a pseudo-wave function that can be approximated by a single plane
wave. Discuss whether this function is appropriate to represent a 2s wave func-
tion. Evaluate the magnitude of the pseudo-potential, for low-energy electron
states.

——————————————————————————————————

8.2.6 Exercise 45
Show that the non-local pseudo-potential displayed in eqn.(598) is non-unique.
In particular, show that the magnitude of contribution from each core state
can be changed by an arbitrary factor and yet the energy eigenvalue remains
unchanged.

——————————————————————————————————

276
8.2.7 Exercise 46
Consider the pseudo-potential for a hydrogen-like atom for the 3s and 3p states.
Fourier transform the energy eigenvalue equation. Show that the contributions
to the pseudo-potential from the 1s and 2s core states do not reduce the mag-
nitude of the effective potential on the 3p states. Also show that the non-local
nature of the pseudo-potential from the 2p core states is important for the re-
duction of the magnitude of the effective potential on the 3p states.

——————————————————————————————————

277
8.3 The Tight-Binding Model
The tight-binding method is appropriate to the situation in which the electron
density in a solid can be considered to be mainly a superposition of the densi-
ties of the individual atoms66 . However, the tight-binding method does produce
slight corrections to the atomic densities. It should be a good approximation
for the inner core orbitals where the ratio of the radius of the atomic orbit to
the inter-atomic separation is small.

Consider a lattice with a mono-atomic basis. The Hamiltonian for a single


ion centered at 0 is Ĥ0 and has eigenstates | φm > defined by the eigenvalue
equation
Ĥ0 | φm > = Em | φm > (644)
The periodic potential of the ions can be written as the sum of the potential
from the ion at site 0, V0 , and the potential due to all other ions in the crystalline
lattice ∆V
Vions = V0 + ∆V (645)
Thus, the Hamiltonian is written as the sum of a single ion Hamiltonian and

Periodic Potential due to a lattice of ions


1

0
Vions(r)

-1

-2
-3 -2 -1 0 1 2 3
r/a

Figure 158: The potential periodic potential Vions (r).

the potential due to the rest of the ions

Ĥ = Ĥ0 + ∆V (646)

In the tight-binding method it is convenient to define Wannier functions, φen ,


as a transform of the Bloch functions
 
1 X
φk,n (r) = √ exp i k . R φen ( r − R ) (647)
N R
66 J. C. Slater and G. F. Koster, Phys. Rev. 94, 1498 (1954).

278
1

0
∆V(r)

-1

-2
-4 -3 -2 -1 0 1 2 3 4
r/a

Figure 159: The potential at the origin due to all other atoms (excluding the
atom at the origin), ∆V (r).

In the above expression, the Wannier functions are centered around the differ-
ent lattice points R. The Wannier states are almost localized states and are
composed of a linear superposition of the atomic bound states
X
| φen > = Cn,m | φm > (648)
m

The band structure is found from the energy eigenvalue equation for the Bloch
wave functions
Ĥ | φk,n > = Ek,n | φk,n > (649)
or  
Ĥ0 + ∆V | φk,n > = Ek,n | φk,n > (650)

This energy eigenvalue equation is projected onto the atomic wave function
| φm > bound to the origin, O, which leads to
 
< φm | Ĥ | φk,n > = < φm | Ĥ0 + ∆V | φk,n >

= Ek,n < φm | φk,n > (651)

However, the state | φm > is an eigenstate of the atomic Hamiltonian Ĥ0 and
so the overlap is given by

< φm | Ĥ0 | φk,n > = Em < φm | φk,n >


(652)

279
On substituting this relation into the matrix elements of the eigenvalue equation,
the equation reduces to

( Ek,n − Em ) < φm | φk,n > = < φm | ∆V | φk,n >


(653)

The Bloch wave function can be expressed in terms of the Wannier functions,
and then the Wannier functions are expressed in terms of the atomic wave
functions via
 
1 X
φk,n (r) = √ exp i k . R φen ( r − R )
N R
 
1 X
= √ Cn,m0 exp i k . R φm0 ( r − R )
N R,m0
(654)

The overlap of the Bloch functions and the atomic wave function is expressed
as the sum of the overlap of atomic wave functions at the same site and the
overlaps of atomic wave functions centered at different sites
√ X
N < φm | φk,n > = δm,m0 Cn,m0 +
m0
X X   Z
+ Cn,m0 exp ik.R d3 r φ∗m (r) φm0 (r − R)
m0 R6=0

(655)

Substituting this into the energy eigenvalue equation, one obtains the equation
X  
Ek,n − Em δm,m0 Cn,m0 +
m0
  X   Z
+ Ek,n − Em C
n,m0 exp ik.R d3 r φ∗m (r) φm0 (r − R)
m0 ,R6=0

X Z
= Cn,m0 d3 r φ∗m (r) ∆V (r) φm0 (r) +
m0
X   Z
+ C
n,m0 exp ik.R d3 r φ∗m (r) ∆V (r) φm0 (r − R)
m0 ,R6=0

(656)

The first term on the left side involves the overlap of two atomic wave function
both bound to the site 0. These atomic wave functions are part of an orthonor-
mal set of eigenfunctions67 . The second term on the left hand side involves the
67 The eigenfunctions corresponding to the continuum of scattering states have to be dis-

carded for reasons which will become obvious.

280
overlap of bound atomic wave functions centered at site 0 and at site R, and
may be expected to be exponentially smaller than the first term.
Z
3 ∗

1  d r φm (r) φm0 (r − R) (657)

The two terms on the right both involve the potential ∆V and the atomic wave
function φm (r) located at site 0. The first term on the right hand site involves
the effect of the potential due to the other ions on the central atom. This
term represents the effect of the crystalline electric field on the atomic levels.
The remaining term represents the delocalization of the electrons. The magni-
tudes of the coefficients Cn,m that appear in the expansion of the Wannier state
crucially depend on the ratios of the overlap integrals to the energy difference
(Ek,n − Em ). The dependence of Cn,m upon (Ek,n − Em ), is similar to
the way that the coefficients Ck+Q that occur in the plane wave expansion of
(0)
a Bloch state depend on the energy difference (Ek,n − Ek+Q ). Generally, the
dependence of Cn,m upon (Ek,n − Em ) allows one to approximate the Wannier
functions by retaining only a finite number of atomic wave functions in their
expansions. That is, the expansion of the Wannier function is truncated by only
Li Be B C N O F Ne
0
H
He B C N O F Ne
-5
Li 1s
E [ Rydbergs ]

-10 Be 1s

2s

-15
B 2p

-20
C

-25
1 2 3 4 5 6 7 8

Figure 160: The one-electron energies for the first and second row elements of
the periodic table, as calculated in the Hartree-Fock approximation.

considering atomic wave functions that have energies close to the energy of the
Bloch state.

The set of equations can be solved approximately by considering the spatial


dependence.

If one assumes that the potential ∆V is non-zero only in the range where
φm (r) is negligibly small, both terms on the right hand side will be approxi-
mately zero. Thus, in a first order and very crude approximation, it is found
that Ek,n = Em .

281
On keeping the two-center and three-center integrals in which R is limited
to a few sites which are close to O, and to atomic states with a few energies
close to Em , the set of equations truncate into a finite set. This set of equations
can be solved to yield the Bloch state energies and the Bloch wave functions.

In general, the band widths are linearly related to the overlap matrix ele-
ments, γi,j , where
Z
γi,j (R) = − d3 r φ∗i (r) ∆V (r) φj (r − R) (658)

in which φj are atomic wave functions and R represent atomic positions relative
to the central atom 0. Tight-binding overlap integrals have been tabulated for
an extensive number of elemental solids by Papaconstantopoulos68 . The band
widths increase with the increase in the ratio of the spatial extent of φi (r) to
the typical separation R. Thus, bands with large binding energies which tend
Electronic radius [ units of Bohr radius ]

3.5

3 1s

2.5 1s

2 2s

1.5 2p

0.5

0
1 2 3 4 5 6 7 8

Figure 161: The radii of the one-electron wafefunctions for the first and second
row elements of the periodic table, as calculated in the Hartree-Fock approxi-
mation.

to have wave functions with small spatial extents form narrow bands while the
higher energy bands have broader band widths.

The overlap integrals are conventionally expressed in terms of the angular


momentum quantum numbers (l, m) of the atomic wave functions that are quan-
tized along the axis joining the atoms. The matrix elements are non-negligible
only if the z-component of the angular momentum satisfies a selection rule. The
non-zero overlap matrix elements are then characterized by m. In analogy to
the atomic wave functions, the type of bonding is labelled by the Greek letters
σ, π and δ respectively, corresponding to m = 0, m = ± 1 and m = ± 2.
68 Handbook of the Band Structure of Elemental Solids, D. A. Papaconstantopoulos, Plenum

Press, NY (1989).

282
The overlap integrals corresponding to ssσ and ppπ bonds are negative, as the
lobes of the wave function with the same sign overlap the negative crystal field
potential. The ppσ bonds are positive at large to intermediate separations as
lobes of opposite sign overlap the negative potential, but become negative at
small values of R where the overlap of lobes with the same sign start to domi-
nate. The spσ overlap is an odd function of R and vanishes for zero separation
R = 0 as the different atomic wave functions are orthogonal. The sign of the
spσ overlap depends on the ordering of the s and p orbitals along the axis. The
spσ bond is positive if lobes of different sign overlap and is negative if lobes of
the same sign overlap.

The Helmholtz-Wolfsberg approximation69 consists of replacing the value of


the potential ∆V by a constant. The magnitude of the potential is factorized out
of the integral. Therefore, the overlap integrals merely depend on the displaced
atomic wave functions, i and j. The overlap integrals are then written as

γi,j (R) = ∆V ti,j (R) (659)

where Z
ti,j (R) = − d3 r φ∗i (r) φj (r − R) (660)

The overlap between hydrogen-like 1s wave functions


r
κ3
 
φ1s (r) = exp − κ r (661)
π

can be evaluated from the Fourier transformed wave function


r
κ3 8πκ
φ1s (q) = (662)
π ( q 2 + κ2 )2

The overlap of two wave functions, with a relative displacement R, can be


evaluated via the convolution theorem
d3 q
Z Z  
d3 r φ∗1s (r) φ1s (r − R) = φ1s (−q) φ1s (q) exp i q . R (663)
( 2 π )3

with the result that


   
1 2 2
t1s,1s,σ = − 1 + κR + κ R exp − κR (664)
3

On using the hydrogenic-like 2s and 2p wave functions,


r
κ3
   
φ2s (r) = 1 − κr exp − κ r
π
69 M. Wolfsberg and L. Helmholtz, J. Chem. Phys. 20, 837, (1952).

283
r
κ3
 
φ2p,0 (r) = cos θ κ r exp − κ r
π
r
κ3
   
φ2p,±1 (r) = sin θ exp ± i ϕ κ r exp − κ r

(665)
one finds that the Fourier transform of the 2s and 2p wave functions are given
by

32 π κ ( q 2 − κ2 ) 0
φ2s (q) κ3 = Y0 (θq , ϕq )
( κ2 + q 2 )3
r
κ5 64 π κ q
φ2p,0 (q) = i Y 0 (θq , ϕq )
3 ( κ2 + q 2 )3 1
r
κ5 64 π κ q
φ2p,±1 (q) = − i Y ±1 (θq , ϕq ) (666)
3 ( κ2 + q 2 )3 1
where the dependence on the direction of q is expressed through the factors
Ylm (θq , ϕq ). The functions Ylm (θ, ϕ) are the spherical harmonics. On using the
d d d
Polar plot of Ψ(r,θ,ϕ) m=0 m=+1 m=+2
m=-1 m=-2
S p p
m=0 m = +1 +
m = -1
- +
+
+ - + - - + +
-
+ -
+

ez θ
Polar plot of Ψ(r,θ,ϕ)

Figure 162: A polar plot exhibiting the angular dependence of the spherical
harmonics with quantum numbers l and m.

convolution theorem, the approximate overlap integrals are evaluated as


   
1 2 2 1 4 4
t2s,2s,σ = − 1 + κ R + κ R + κ R exp − κ R
3 15
   
1 3 3
t2s,2p,σ = κ R 1 + κR exp − κ R
15
   
1 2 2 2 3 3 1 4 4
t2p,2p,σ = − 1 + κ R + κ R − κ R − κ R exp − κ R
5 15 15
   
2 2 2 1 3 3
t2p,2p,π = − 1 + κ R + κ R + κ R exp − κ R
5 15
(667)

284
where κ determines the spatial extent of the wave function and R is the inter-
atomic separation. Typically for a material such as C, the relative strength of

Tight-binding Overlap Integrals


1

0.5 tspσ

tppσ
t(R)

tppπ
-0.5 tssσ

-1
0 2 4 6 8 10

κR
Figure 163: The dependence of the tight-binding overlap integrals on the inter-
atomic separation R.

the bonds are given by the ratios at the radius R where the bonding saturates.
Typical values of the relative strengths are given by

t2s,2s,σ : t2s,2p,σ : t2p,2p,σ : t2p,2p,π = − 1 : 1 : 0.75 : − 0.49 (668)

The structure of tight-binding d bands can be found by expressing the Bloch


functions in terms of five atomic d wave functions that correspond to the different
eigenvalues of the z component of the orbital angular momentum mz = ± 2,
mz = ± 1 and mz = 0. If mz is quantized along the axis between two
atoms, the tight-binding overlap integrals between these sets of states are de-
noted, respectively, by td,d,δ , td,d,π and td,d,σ . The matrix elements for arbitrary
orientations are tabulated in the article of Slater and Koster70 . Representative
ratios of the strengths of the td,d,δ , td,d,π and td,d,σ bonds are given by

td,d,δ : td,d,π : td,d,σ = − 6 : 4 : − 1 (669)

In general, the tight-binding bands obtained by considering d bands alone are


highly inaccurate. Usually, a broad s band crosses the narrow set of d bands.
This degeneracy is lifted as the d and s bands hybridize strongly71 .

70 J. C. Slater and G. F. Koster, Phys. Rev. 94, 1498 (1954).


71 V. Heine, Phys. Rev. 153, 673 (1967).

285
The Bloch functions are constructed out of localized atomic levels with equal
amplitude, but only involves the phase exp[ i k . R ]. Thus, the electrons are
equally likely to be found in any atomic cell of the crystal. Also, <e φk,n
shows that the atomic structure is modulated by the sinusoidal variation of
exp[ i k . R ]. Since the mean velocity is given by
1
v(k) = ∇Ek 6= 0 (670)

then the electrons have a non-zero velocity and will be able to move through-
out the crystal. The non-zero velocity is due to the coherent tunnelling of the
electron between the atoms.

For a lattice with a basis, the Bloch wave function is given


1 X X
φk (r) = √ exp[ i k . R ] Cj,m φm (r − rj − R) (671)
N R j,m

where rj are the positions of the basis atoms and Cj,m are the amplitudes of
the orbitals on the j-th basis atom. The equation for the Bloch function has a
structure in which the basis atoms in each unit cell can be viewed as forming
molecules72 . The molecular wave functions in each lattice cell are then com-
bined via the tight-binding method.

8.3.1 Tight-Binding s Band Metal


For a simple s-band metal the Wannier state | φen > can be approximated by
the atomic s wave function. As this s wave function is non-degenerate, one has

| φe1 > ≈ | φs > (672)

as Cs ≈ 1. All other coefficients are set to zero, corresponding to the assump-


tion that the energy of the s band, Es , is well separated from the energies of
the other bands. This is probably a good assumption for the 1s band which is
often regarded as forming part of the core of the ions. The energy eigenvalue
equation truncates to
  X  Z 
Es,k − Es 1 + exp i k . R d3 r φ∗s (r) φs (r − R)
R6=0
X  Z
= < φs | ∆V̂ | φs > + exp ik.R d3 r φ∗s (r) ∆V (r) φs (r − R)
R6=0

(673)
72 By using symmetry adapted wave functions, the tight-binding secular equation may be

put in block diagonal form. Hence, an N × N secular equation might be reducible to a set of
smaller secular equations.

286
The overlap between the atomic wave functions on different sites is defined to
be a function α(R) through
Z
d3 r φ∗s (r) φs (r − R) = α(R) (674)

The matrix elements of the atomic functions centered at 0 with the tail of the
potential, ∆V , is defined to be β where

< φs | ∆V̂ | φs > = − β (675)

and the matrix elements of the atomic functions centered at 0 and R with the
tail of the potential is defined to be γ(R) through
Z
d3 r φ∗s (r) ∆V (r) φs (r − R) = − γ(R) (676)

The dispersion relation can be expressed in terms of these three functions via
 
P
β + R6=0 γ(R) exp i k . R
!
Es,k = Es −   (677)
P
1 + R6=0 α(R) exp i k . R

Since γ(R) = γ(−R) and α(R) = α(−R) the dispersion relation E1,k is an
even periodic function of k. For bonding only to the nearest neighbors, the sums
over R are truncated to run only over the nearest neighbors.

For the f.c.c. structure the dispersion relation becomes


!
β + γ(k)
Es,k = Es − (678)
1 + α(k)

where
 
kx a ky a kx a kz a ky a kz a
γ(k) = 4 γ cos cos + cos cos + cos cos
2 2 2 2 2 2
(679)
and
 
kx a ky a kx a kz a ky a kz a
α(k) = 4 α cos cos + cos cos + cos cos
2 2 2 2 2 2
(680)
Usually α is neglected as it is small. The tight-binding bands are off-set from
Es by an energy β due to the tail of the potential of all other atoms at O,

β = − < φs | ∆V̂ | φs > (681)

287
The band width is governed by the overlap of the central atom’s wave function
with the nearest neighbor atomic wave function. This overlap, γ, is evaluated
from Z
γ = − d3 r φ∗s (r) ∆V (r) φs (r − Rnn ) (682)

The band width for the f.c.c. lattice is 12 γ.

For small | k | a one can expand the dispersion relation in powers of k

E1,k = Es − β − 12 γ + γ k 2 a2 (683)

which is independent of the direction of k near k = 0. Thus, the constant


energy surfaces are spherical around k = 0.

The gradient of the energy has a component perpendicular to the square


face of the Brillouin zone (the face containing the X point) that is given by
 
∂Ek kx a ky a kz a
= 2 a γ sin cos + cos (684)
∂kx 2 2 2

Thus, if E1,k is plotted along any line in k space which is perpendicular to the
square face, it crosses with zero slope.

The points on the hexagonal face satisfy the equation


 
3π 3 2π
kx + ky + kz = = (685)
a 2 a

Since there is no plane of symmetry parallel to the hexagonal face, the energy
plotted along any line perpendicular to the hexagonal face is not required to
cross with zero slope,
 
kx a ky a kz a
∇ E1,k . ê ∝ sin cos + cos
2 2 2
 
ky a kx a kz a
+ sin cos + cos
2 2 2
 
kz a kx a ky a
+ sin cos + cos
2 2 2
(686)

This only vanishes along the lines joining L ( 12 , 12 , 12 ) to the vertices W (1, 21 , 0).

For degenerate levels such as p or d levels, the tight-binding method leads


to a N × N secular equation where N is the orbital degeneracy.

288
For heavy elements, spin-orbit coupling should be included. In this case, the
potential ∆V should have a spin-dependent contribution. The spin-orbit cou-
pling breaks the spin-degeneracy and increases the size of the secular equation73
by a factor of 2.

8.3.2 Tight-Binding Bands of Diamond Structured Semiconductors


Tight-binding calculations for semiconducting materials with the diamond struc-
ture, such as C, Si or Ge, require a minimum basis set which consists of one
s state and three p states for the atom at each of the two lattice sites. These
eight states are required in order to provide a reasonable description of the
valence bands. More states need to be included in the basis set, in order to
yield a reasonable description of the lowest energy conduction bands. We shall
use the Hückle approximation, in which the non-zero overlap matrix elements
between atomic wave functions centered on different atomic sites are neglected.
Furthermore, the non-zero tight-binding hopping matrix elements will be evalu-
ated in the Helmholtz-Wolfsberg approximation. The Bloch wave functions are
expressed in terms of a sum of primitive Bloch functions based on the atomic
orbitals with atomic quantum numbers β (φ0β ) centered on the two lattice sites.
Hence,
 
1 X
φk,α (r) = √ Cαβ,j exp i k . ( R + rj ) φ0β (r − R − rj ) (687)
N R,j,β

where the coefficients Cαβ,j are to be determined. On substituting the above


ansatz for the wave function into the energy eigenvalue equation, and on taking
the matrix elements with φ0β (r), one finds that the dominant overlap integrals
are either on-site or occur with the atomic states on the atoms of the surrounding
tetrahedron. If the central atom has coordinates (0, 0, 0), then the neighboring
atoms which are tetrahedrally coordinated with it are located at the set of
equivalent sites
a
r1 = (1, 1, 1)
4
a
r2 = (1, 1, 1)
4
a
r3 = (1, 1, 1)
4
a
r4 = (1, 1, 1) (688)
4
These four sites are equivalent since they are related to the each other by a com-
bination of the f.c.c. primitive lattice vectors, a(0, 21 , 12 ), a( 12 , 0, 12 ), a(0, 12 , 12 ).
For the other lattice site, say the site at a4 (1, 1, 1), the separations to the tetra-
hedrally coordinated neighboring atoms are given by the four equivalent vectors
a
r5 = (1, 1, 1)
4
73 J. Friedel, P. Lenghart and G. Leman, J. Phys. Chem. Solids, 25, 781 (1964).

289
a
r6 = (1, 1, 1)
4
a
r7 = (1, 1, 1)
4
a
r8 = (1, 1, 1) (689)
4
Thus, the vectors parallel to the bonds with the atoms of the surrounding tetra-
hedron have opposite orientations for the two lattice sites. This is expected since
the two sites are connected via a glide-like inversion symmetry about the point
a
8 (1, 1, 1). Thus, the phase factors in the matrix form of the energy eigenvalue
equation involve combinations of the four phase factors
 
a
exp i ( + kx + ky + kz )
4
 
a
exp i ( − kx − ky + kz )
4
 
a
exp i ( + kx − ky − kz )
4
 
a
exp i ( − kx + ky − kz ) (690)
4
or their complex conjugates.

On considering the matrix element of the eigenvalue equation with the s


wave function located at site (0, 0, 0) and on utilizing the tight-binding approx-
imations, one obtains the linear equation
0 = ( Es − Ek,α ) Cαs,0 − 4 γss,σ gs (k) Cαs,1
 
4 γsp,σ
− √ gx (k) Cαpx ,1 + gy (k) Cαpy ,1 + gz (k) Cαpz ,1 (691)
3
which for the s − s overlaps, involves the combination of the phase factors
    
1 a a
gs (k) = exp i ( kx + ky + kz ) + exp i ( − kx − ky + kz )
4 4 4
   
a a
+ exp i ( kx − ky − kz ) + exp i ( − kx + ky − kz )
4 4
kx a ky a kz a kx a ky a kz a
= cos cos cos − i sin sin sin (692)
4 4 4 4 4 4
The s − p hopping matrix elements are found by resolving the p orbitals in new
coordinate systems where the z 0 -axes are parallel to the ri . In this case, only
the σ matrix elements of the s and the p wave functions are non-zero, since the
z 0 component of the angular momenta is conserved. For example, for the px
orbital, one finds that the s − p overlap is given by
    
1 a 1 a
γsp,σ √ exp i ( kx + ky + kz ) − √ exp i ( − kx − ky + kz )
3 4 3 4

290
   
1 a 1 a
+ √ exp i ( kx − ky − kz ) − √ exp i ( − kx + ky − kz )
3 4 3 4
(693)

The negative signs in this expression occur since the overlap of the central s
wave function with the px wave functions located at the positions (xj , yj , zj )
and (xj , yj , zj ) have opposite signs. The above expression is written as
4 γsp,σ
√ gx (k) (694)
3
where the relevant combination of phase factors is given by
kx a ky a kz a kx a ky a kz a
gx (k) = − cos sin sin + i sin cos cos (695)
4 4 4 4 4 4
Likewise, the sum of the other combinations of phase factors for the s − py and
s − pz hopping are given by the analogous expressions
kx a ky a kz a kx a ky a kz a
gy (k) = − sin cos sin + i cos sin cos (696)
4 4 4 4 4 4
and
kx a ky a kz a kx a ky a kz a
gz (k) = − sin sin cos + i cos cos sin (697)
4 4 4 4 4 4

The matrix element of the eigenvalue equation with the other s wave function
φ0s (r − r1 ) is given by

0 = ( Es − Ek,α ) Cαs,1 − 4 γss,σ gs∗ (k) Cαs,0


 
4 γsp,σ ∗ px ,0 ∗ py ,0 ∗ pz ,0
+ √ gx (k) Cα + gy (k) Cα + gz (k) Cα (698)
3
The sign of the term proportional to γsp,σ is opposite to that of eqn.(691), since
the bonds with the atoms in the surrounding tetrahedra have different orienta-
tions.

The overlap integrals between the p wave functions on the two atoms can
be evaluated in much the same way. For the overlap between the two p wave
functions located on atoms separated by the vector a4 (1, 1, 1), one must express
the px wave functions along a new z 0 axis (êz0 ) parallel to the line joining the
atoms √13 (1, 1, 1) and two other mutually orthogonal directions (say, √16 (1, 2, 1)
and √12 (1, 0, 1)). Thus, since the px wave function has the form

x f (r) = ( r . êx ) f (r)


(699)

291
and as
1 1 1
êx = √ êx0 + √ êy0 + √ êz0 (700)
6 2 3
one can express the px wave function in terms of the p wave functions which are
quantized in the new coordinate system
1 1 1
φ0px (r) = √ φ0px0 (r) + √ φ0py0 (r) + √ φ0pz0 (r) (701)
6 2 3
Likewise,
2 1
φ0py (r) = − √ φ0px0 (r) + √ φ0pz0 (r) (702)
6 3
and
1 1 1
φ0pz (r) = √ φ0px0 (r) − √ φ0py0 (r) + √ φ0pz0 (r) (703)
6 2 3
The px − px overlap integral for the a4 (1, 1, 1) bond is then found to be given by
1 1 1 1
(+ ) γpp,π + γpp,σ = ( 2 γpp,π + γpp,σ ) (704)
6 2 3 3
while the corresponding px − py and px − pz overlap integrals are both given by
1
( γpp,σ − γpp,π ) (705)
3
On evaluating the matrix element of the px wave function at the origin with the
energy eigenvalue equation, one obtains the linear equation
4 4
+ √ γsp,σ gx (k) Cαs,1 + ( Ep − Ek,α ) Cαpx ,0 − ( γpp,σ + 2 γpp,π ) gs (k) Cαpx ,1
3 3
4 4
− ( γpp,σ − γpp,π ) gz (k) Cαpy ,1 − ( γpp,σ − γpp,π ) gx (k) Cαpz ,1 = 0 (706)
3 3
Similar equations are found for the matrix element of the energy eigenvalue
equation with the py and pz wave functions. On repeating this procedure for
the other atom located at a4 (1, 1, 1), one obtains a closed set of coupled linear
equations. The energy eigenvalues Ek,α are then given by the solutions of the
secular equation. The 8 × 8 secular matrix is given by
4γsp,σ 4γsp,σ
Es − Ek,α −4γss,σ gs 0 0 0 − √ gx − √ gy
3 3
∗ 4γsp,σ ∗ 4γsp,σ ∗ 4γsp,σ ∗
−4γss,σ gs Es − Ek,α √ gx √ gy √ gz 0 0
3 3 3
4γsp,σ 4(γpp,σ +2γpp,π ) 4(γpp,σ −γpp,π )
0 √ gx Ep − Ek,α 0 0 − gs − gz
3 3 3
4γsp,σ 4(γpp,σ −γpp,π ) 4(γpp,σ +2γpp,π )
0 √ gy 0 Ep − Ek,α 0 − gz − gs
3 3 3
4γsp,σ 4(γpp,σ −γpp,π ) 4(γpp,σ −γpp,π )
0 √ gz 0 0 Ep − Ek,α − gy − gx
3 3 3
4γsp,σ ∗ 4(γpp,σ +2γpp,π ) ∗ 4(γpp,σ −γpp,π ) ∗ 4(γpp,σ −γpp,π ) ∗
− √ gx 0 − gs − gz − gy Ep − Ek,α 0
3 3 3 3
4γsp,σ ∗ 4(γpp,σ −γpp,π ) ∗ 4(γpp,σ +2γpp,π ) ∗ 4(γpp,σ −γpp,π ) ∗
− √ gy 0 − gz − gs − gx 0 Ep − Ek,α
3 3 3 3
4γsp,σ ∗ 4(γpp,σ −γpp,π ) ∗ 4(γpp,σ −γpp,π ) ∗ 4(γpp,σ +2γpp,π ) ∗
− √ gz 0 − gy − gx − gs 0 0
3 3 3 3

The tight-binding parameters for C, Si and Ge have been tabulated by Chadi


and Cohen74 , and are given in Rydbergs in the following table:

74 D. J. Chadi and M. L. Cohen, Physica Status Solidi, B 68, 405 (1975).

292
Ep − Es γss,σ γsp,σ γpp,σ γpp,π

C 0.544 0.279 − 0.326 − 0.360 0.097

Si 0.529 0.149 − 0.187 − 0.334 0.080

Ge 0.618 0.125 − 0.169 − 0.299 0.077

0.2
Ek [ Rydbergs ]

-0.2

-0.4

-0.6
Γ X W L Γ K X

Figure 164: The tight-binding approximation for the valence bands of crystalline
Si.
The lowest four bands are completely occupied. At the Γ point these consist of
the bonding s band which has the energy Es − 4 | γss,σ |, followed by three-fold
degenerate bonding bands with energies Ep − 43 | ( γpp,σ + 2 γpp,π ) |. The
anti-bonding bonds are unoccupied. The anti-bonding s band has the energy
Es + 4 | γss,σ |, whereas the three-fold degenerate anti-bonding p bands have
the energy eigenvalues Ep − 43 | ( γpp,σ + 2 γpp,π ) |.

——————————————————————————————————

8.3.3 Exercise 47
Consider two p orbitals, one located at the origin and another at the point
R (cos Θx , cos Θy , cos Θz ), where R is the separation between the two ions and

293
Figure 165: The (occupied) valence bands of diamond as determined by Angle
Resolved Photoemission Spectroscopy experiments [From Jiménez et al., Phys.
Rev. B, 56, 7215 (1997).].

the cos θ are the direction cosines of the displacements. The overlap parameters
for the orbitals φi (r) and φj (r) are defined by
Z
γi,j (R) = − d3 r φ∗i (r) ∆V (r) φj (r − R) (707)

Show that the overlap parameters are given by


 
2 2
γx,x = ∆V tppσ cos Θx + tppπ sin Θx
 
γx,y = ∆V tppσ − tppπ cos Θx cos Θy (708)

Thus, the tight-binding parameters not only depend on the distance, R, but
also depend on the direction.

——————————————————————————————————

8.3.4 Exercise 48
Consider the p bands in a cubic crystal, which have the p wave functions

φpx (r) = x f (r)


φpy (r) = y f (r)
φpz (r) = z f (r) (709)

294
where f (r) is a spherically symmetric function. The energies of the three p
bands are found from the secular equation
 

Ek − Ep δi,j + βi,j + γi,j (k) = 0 (710)

and  
X
γi,j (k) = exp ik.R γi,j (R) (711)
R

and Z
γi,j (R) = − d3 r φ∗i (r) ∆V (r) φj (r − R) (712)

and
βi,j = γi,j (0) (713)

Show that, using cubic symmetry,

βx,x = βy,y = βz,z = β (714)

and all other overlap matrix elements are zero

βx,y = βy,z = βx,z = 0 (715)

Assuming that only the nearest neighbor overlaps γi,j (R) are non-zero, show
that for a simple cubic lattice γi,j (k) are diagonal in i and j. Hence, the px , py
and pz wave functions generate three independent bands

Ex,k = Ep − β − 2 γppσ cos kx a − 2 γppπ ( cos ky a + cos kz a )


Ey,k = Ep − β − 2 γppσ cos ky a − 2 γppπ ( cos kx a + cos kz a )
Ez,k = Ep − β − 2 γppσ cos kz a − 2 γppπ ( cos kx a + cos ky a )
(716)

The relative values of these parameters can be estimated from first princi-
ples calculations of bulk silicon, where the ratios were found to be given by
tppσ : tppπ = 3.98 : − 1 .

——————————————————————————————————

295
8.3.5 Exercise 49
Consider the p bands in a face-centered cubic lattice with nearest neighbor
hopping γi,j (R). Show that the system is described by a 3 × 3 secular equation
which is expressed in terms of four integrals

E − Ek0 + Mx0 − Mz1 − My1




1 0 0 1

0 = − M z E − E k + M y − M x

(717)
1 1 0 0
− My − Mx E − E k + Mz

where the functions Mi0 and Mi1 are given by

ky a kz a
Mx0 = 4 γ0 cos cos
2 2
ky a kz a
Mx1 = 4 γ1 sin sin (718)
2 2
and cyclic permutations. The energy Ek0 is given by
 
ky a kz a kx a kz a kx a ky a
Ek0 = Ep − β − 4 γ2 cos cos + cos cos + cos cos
2 2 2 2 2 2
(719)
Evaluate the integrals γn in terms of the overlap of atomic wave functions by
using the Helmholtz-Wolfsberg approximation. Also show that the three energy
bands are degenerate at the Γ point, and that when k is directed along the cube
axis (Γ X) or the cube diagonal (Γ L), two bands are degenerate.

——————————————————————————————————

8.3.6 Exercise 50
The parent compound of the doped high temperature superconductors is La2 CuO4
which has the Perovskite structure. In this structure, the CuO2 atoms form
planes. Each Cu atom is surrounded by an octahedra of O atoms of which four
atoms are in the plane. The in-plane Cu − O bonds can serve to define the
x and y axes. The O atoms that have the Cu − O bonds parallel to the x axis
are denoted as Ox , whereas the other O atoms are denoted by Oy . In this coor-
dinate system, the appropriate basis orbitals are the Cu dx2 −y2 orbitals, while
the only Ox states which mix with the Cu states are the px states and the only
Oy states that mix with the Cu are the py states.

Using the tight-binding form of the Bloch wave function


  
1 X d p a p a
φk = √ exp i k . R a φx2 −y2 (r) + bx φx (r− êx ) + by φy (r− êy )
N R 2 2
(720)

296
find the energy bands for the CuO2 planes.

——————————————————————————————————

8.3.7 Exercise 51
Evaluate the tight-binding density of states for the s states of a simple hyper-
cubic lattice in d = 1, d = 2, d = 3, d = 4, in which only the nearest
neighbor hopping matrix elements t are retained. Calculate the form of the
density of states when d → ∞.

——————————————————————————————————

8.3.8 Exercise 52
Consider the tight-binding density of states for s states on a tetragonal lattice
where the overlap in the c direction is t0 and the overlap in either the a or b
direction is t. Assume that t  t0 . Examine the form of the Fermi surface
when the band is nearly half-filled. Evaluate the density of states.

——————————————————————————————————

8.3.9 Wannier Functions


Bloch functions are eigenstates of the set of operators T̂R which produce transla-
tions through the discrete lattice vectors R. The eigenvalues of T̂R are exp[ − i k . R ].
The set of periodic translations can be viewed as being produced by a generator,
ˆ , via

 
T̂R = exp − i R . ℘ ˆ (721)

The operator ℘ ˆ is the crystal momentum operator and has eigenvalues k, and
its eigenvectors are the Bloch functions


ˆ φn,k (r) = k φn,k (r) (722)

The eigenvalues of ℘ ˆ are only defined in a bounded region of k space, since


k ≡ k + Q. One expects that there is an operator < ˆ which is canonically con-
jugate to ℘ˆ . Like the relation between canonically conjugate operators L̂z and
ϕ̂ in quantum mechanics, the eigenvalues of < ˆ may be expected to be discrete
since k is only defined in a bounded region of space. The eigenstates of < ˆ are
the Wannier functions.

297
Consider the position r to have a fixed value, and n to be fixed. That is,
we shall consider the Bloch functions to be functions of k. The Bloch functions
can be written as
 
1 X
φk,n (r) = √ exp i k . R fn (r, R) (723)
N R

The Bloch function φk for fixed r is periodic in k, with periodicity given by the
primitive reciprocal lattice vectors Q. Clearly
 
1 X
φk+Q,n (r) = √ exp i ( k + Q ) . R fn (r, R)
N R
 
1 X
= √ exp i k . R fn (r, R)
N R
= φk,n (724)

since Q and R satisfy the Laue condition. Thus, the Bloch functions are periodic
functions in k space. The Fourier coefficients, fn (r, R), that appear in the k
space Fourier expansion can be found from the inversion formulae
 
1 X 0
fn (r, R) = √ exp − i k . R φk0 ,n (r) (725)
N k0

where the sum over k 0 is restricted to run over the volume of the first Brillouin
zone, Ωc .

The simultaneous transformations r → r − R0 and R → R − R0 leave


the function fn (r, R) unchanged

fn (r, R) = fn (r − R0 , R − R0 ) (726)

This is proved by considering the effect of the transformation r → r − R0 on


the definition of the functions fn (r, R)
 
1 X
φk,n (r) = √ exp i k . R fn (r, R0 )
0
(727)
N R0

Applying the transformation on the Bloch function yields


 
1 X
φk,n (r − R0 ) = √ exp i k . R0 fn (r − R0 , R0 ) (728)
N R0

and then, on transforming the sum over R0 as R0 = R − R0 , one has


 
1 X
φk,n (r − R0 ) = √ exp i k . (R − R0 ) fn (r − R0 , R − R0 ) (729)
N R

298
On comparing the above expression with the result of Bloch’s theorem
   
1 X
φk,n (r − R0 ) = exp − i k . R0 √ exp i k . R fn (r, R) (730)
N R

one recovers the symmetry relation

fn (r, R) = fn (r − R0 , R − R0 ) (731)

Using the above symmetry of f (r, R) under a translation R0 , and on choosing


R0 = R, one finds

fn (r, R) = fn (r − R, 0) = φen (r − R) (732)

which shows that the function only depends on the difference r − R. Hence, it
has been shown that the Bloch function can be expressed as
 
1 X
φk,n (r) = √ exp i k . R φen (r − R) (733)
N R

where φen (r) are the Wannier functions75 . The Wannier functions at different
sites are orthogonal. Thus, as they are linearly related to the Bloch wave func-
tions φk,n (r), the set of Wannier functions form a complete orthogonal set.

The Wannier functions are given in terms of the Bloch functions via the
inverse transform
 
1 X
φn (r − R) = √
e exp − i k . R φk,n (r) (734)
N k

where the sum over k is restricted over the volume of the first Brillouin zone,
Ωc . The Wannier functions defined in this way are not unique76 . The Wannier
ˆ with eigenvalues of R 77 . The
functions are eigenstates of the operator <,
75 G. Wannier, Phys. Rev. 52, 191 (1947).
76 Sincethe Bloch functions are only defined up to an arbitrary phase factor, if φn,k (r) is a
Bloch function and ϕ(k) ≡ ϕ(k +Q) + 2π M is an arbitrary phase, then exp[ i ϕ(k) ] φn,k (r)
is also a Bloch function. This means that there is an arbitrariness in the definition of the
Wannier functions. The Wannier functions could equally well be defined as
   
en (r − R) = √1
X
φ exp − ik.R exp i ϕ(k) φk,n (r)
N
k

The phase factor is usually chosen such that the Wannier function φ en (r − R) is maximally
localized around site R.
ˆ = − i ∇ . This
77 This is most easily seen if one works in a representation where <
k
is discussed in detail by G. Weinreich in Solids, Elementary Theory for Advanced Students,
John Wiley & Sons Inc. New York, (1965).

299
Wannier functions are localized around the site R, as can be seen by substituting
the expression for the Bloch functions in the above equation
r
d3 k
Z  
V
φn (r − R) =
e exp + i k . ( r − R ) un,k (r) (735)
N Ωc ( 2 π )3

The phase factor in the integral over d3 k has the effect of localizing the Wannier
function around r = R, as at this r value the phase of the integral is stationary.
The integral is easy to evaluate for free electrons for which un (r) = 1. The
Wannier functions appropriate to free electrons in an orthorhombic lattice are
given by
  sin [ π y ]  
sin [ πaxx ] sin [ πazz ]
 
1 ay
φn (r) = √
e
π x π y π z (736)
ax ay az ax ay az

which have amplitudes that decay algebraically outside the unit cell. This alge-

Free Electron Wannier Functions

1.2

0.8
√a
φ(x)/√

0.4

0
-4 -2 0 2 4

-0.4

x/a

Figure 166: The Wannier functions for free electrons.

braic decay is found only for bands with infinite width. Bands that have allowed
energies that are separated by forbidden ranges of E of finite width have Wan-
nier functions that decay exponentially. Furthermore, the rate of exponential
decay is dependent on the band width78 .

——————————————————————————————————

78 W. Kohn, Phys. Rev. 115, 809 (1959), E. I. Blount, Solid State Physics, Vol 13, Acad.

Press, (1962).

300
8.3.10 Exercise 53
Prove that the Wannier functions centered on different lattice sites are orthog-
onal Z
d3 r φe∗n0 (r − R0 ) φen (r − R) ∝ δn0 ,n δR0 ,R (737)

Also show that the Wannier functions are normalized to unity


Z
d3 r | φen (r) |2 = 1 (738)

——————————————————————————————————

8.3.11 Exercise 54
Given that the Bloch functions form a complete set, show that the Wannier
functions also form a complete set.

——————————————————————————————————

8.3.12 Example of Tight-Binding: Graphene


Carbon in its stable form of graphite is a semimetal. The structure of graphite
consists of well separated stacks of two-dimensional layers, as the interatomic
spacing in the layers is a = 1.42 Å, while the interlayer spacing is given by
c
2 = 3.35 Å. As a consequence of theGraphite
Two-dimensional two-dimensional
Structure form of the structure,

a1 = 3/2 a ex + √3/2 a ey

a1

a2

a2 = 3/2 a ex - √3/2 a ey

Figure 167: The crystal structure of a sheet of graphene.

301
the electronic states can almost be described by a two-dimensional tight-binding
model79 . A single layer of graphite is also known as Graphene. In the two-
Brillouin Zone of two-dimensional Graphite

ky 2π/3a(0,2/√3)
b1 = (2π/3a) ( ex + √3 ey )

b2 = (2π/3a) ( ex - √3 ey )
P 2π/3a(1,1/√3)

kx
2π/3a(1,0)
Γ Q

Figure 168: The first Brillouin Zone of a graphene sheet.

dimensional model, the pz wave functions are out of plane and decouple from
the s, px and py orbitals. The s, px and py orbitals mix and form bonding and
anti-bonding σ orbitals which are separated by a very large band gap. Since
there are two carbon atoms in the unit cell of a single sheet of graphene, the
three bonding levels are completely occupied with six electrons. The remaining
two electrons are to be distributed amongst the pz -states. The pz -orbitals form
the bonding and anti-bonding π bands. Their dispersion relations are given by
the solution of the secular equation

Ep − Ek,α − γpp,π gs (k)
0 = (739)
− γpp,π gs∗ (k)

Ep − Ek,α

where
    √
kx a 3 ky a
gs (k) = exp − i kx a + exp + i 2 cos (740)
2 2

Due to the vanishing of gs (k), the bonding and anti-bonding π bands just touch
at the single point P ≡ 2π √1
3a (1, 3 , 0). Since the remaining two electrons com-
pletely fill the anti-bonding π band, the Fermi surface consists of just one point,
namely P . Hence, the two-dimensional model describes graphite as a zero band
gap semiconductor. However, if the small inter-layer hopping is reintroduced,
the bands energies show a small modulation as kz is varied. This slight mod-
ulation of the energy on the lines H - P - H modulates both the bonding
and anti-bonding π bands. As a result, the density of states from the bond-
ing and anti-bonding π bands slightly overlap and the material is semimetallic.
The Fermi surface of the three-dimensional material consists of small (equal
volume) needle shaped hole pockets and electron pockets centered on the six
vertical edges of the hexagonal Brillouin zone located at 2π √1
3a (1, 3 , κ3 ). The

79 F. Bassani and G. Pastori Parravicini, Il Nuovo Cimento B 50, 95, (1967).

302
Two-dimensional Graphite

0.1

-0.1
π

Ek [ Hartrees ]
-0.3
EF
-0.5 π

-0.7

-0.9

-1.1
P Γ Q P
(3ka/2π)

π Density
Figure 169: The dispersion relation
of States for the p-bands
of two-dimensional of graphene.
Graphite

12
ρ(ε) [ States / Hartree ]

10

2 εF

0
-0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0

ε [ Hartrees ]

Figure 170: The density of states for the π-bands of graphene.

occupied portions of the electronic bands of graphite and graphitic layers sepa-
rated by Li or Rb (intercalated graphite) have been measured in angle resolved
photoemission experiments80 and are found to be in reasonable agreement with
the two-dimensional model.

Carbon Nanotubes

The semi-metallic nature of the single layer of graphite is an important


ingredient in the description of carbon nanotubes. These carbon nanotubes
were an unexpected by-product of studies of C60 molecules81 . It was found that
it was possible to grow carbon tubules, which may have walls composed of one
or more layers of graphene. A single-walled nanotube may be viewed a being
constructed from a single graphene sheet, which is cut by two parallel lines,
80 I.T. McGovern, W. Eberhardt, E.W. Plummer and J.E. Fischer, Physica B, 99, 415 (1980).
81 S. Ijima, Nature, 354, 56, (1991).

303
Figure 171: Typical forms of single-walled Carbon Nanotubes. The indices of
the chiral vector cp are denoted by (p1 , p2 )

both normal to the chiral vector

cp = p1 a1 + p2 a2 (741)

for integer values of (p1 , p2 ). The lattice points on the line passing through
the origin must be identified with lattice points connected by cp . It should be
noted that the two sub-lattices must remain distinct after this construction has
taken place. (Both these conditions must be satisfied, for the local geometry
to be unique at any site on the line.) The resulting structures consists of a
graphite tubules or cylinders, which may or may not have a chiral character82 .
The fundamental translation vector τ is given by the vector originating from O,
normal to cp which ends at the first lattice point it meets. This translational
vector is given by
( p1 + 2 p2 ) ( 2 p1 + p 2 )
τ = a1 − a2 (742)
d d
where d is the greatest common divisor of (p1 + 2p2 ) and (2p1 + p2 ).

The properties of a single nanotubule are governed by cp , that is by the


chiral indices (p1 , p2 ). The unit cell for the graphite tubule consists has primitive
lattice vectors cp and τ . The radius of the nanotubule is given by

| cp | 3a
q
r = = p21 + p22 + p1 p2 (743)
2π 2π
82 R. Saito, M. Fujita, G. Dresselhaus and M.S. Dresselhaus, Phys. Rev. B 46, 1804 (1992).

304
τ

Ο θ
cp

Figure 172: The representation of a single-walled carbon nano-tube in terms of


a sheet of graphene. The chiral vector cp is depicted by a red dashed line. A
line parallel to the tube axis is depicted by the dark blue dashed lines. The
chiral angle is denoted by θ.

The length of the unit cell of the nanotube is given by


3a
q
|τ | = p21 + p22 + p1 p2 (744)
d
Since the area of a single unit cell of graphene is

27 2
a (745)
2
the number of unit cells of the honeycomb lattice in the unit cell of the nan-
otubule is given by
 
2
N = p21 + p22 + p1 p2 (746)
d

The zig-zag axis is defined as the direction (0, 1). The chiral angle θ is defined
as the angle between cp and the zig-zag axis. The chiral angle governs the screw
symmetry of the graphene structure along the tubule, and is given by

3 p1
tan θ = (747)
2 p2 + p1
The nanotubues which have mirror planes are considered to be non-chiral.

The finite radius of the tubule causes the crystalline momentum along the
c-direction (perpendicular to the tubule’s axis) to be quantized. Each quantized

305
k value forms its own one-dimensional band, which is separated from the next
one-dimensional band by an energy gap due to the finite size. The allowed
momentum values are found to lie on parallel lines which slice through the
two-dimensional Brillouin zone. This is seen by applying periodic boundary
conditions across the “circumference” of the tubule

k . cp = 2 π nt (748)

where nt is an integer. Hence, the quantization condition is



3 3
2 π nt = ( p1 + p2 ) kx a + ( p1 − p2 ) ky a (749)
2 2
Therefore, the allowed values of k fall onto a set of discrete lines. This quanti-
zation condition leads to three types of nanotubes. In one set of states, there is
a line of quantized crystal momenta (k1 , k2 ) which passes through the P point
where the two bands are degenerate. This occurs when

3 n t = 2 p1 + p2 (750)

which is satisfied if
p1 − p2 = 3 j (751)
for integer j, since
n t = p1 − j (752)
Thus, if p1 − p2 = 3 j the system is metallic. Since there is a one-dimensional
band of states running through the P point and since the energy of the P point
is equal to the Fermi energy, the system acts like a one-dimensional metal83 .
Due to the reduced dimensionality of phase space, the density of states at the
Fermi energy is finite. There are another two types of states where

p1 − p2 = 3 j + 1 (753)

or
p1 − p 2 = 3 j + 2 (754)
In these other two types of materials, the crystal momentum corresponding
to the P point is absent and since there is always a finite energy gap between
the occupied and unoccupied states, these other two types of carbon nanotubes
are semiconducting. Hence, nanotubes have a broad range of properties which
sensitively depend on the tube’s structure.

83 The curvature of the tube may introduce tiny gaps in the density of states at the Fermi

energy.

306
Semiconducting Nanotube Density of States Metallic Nanotube Density of States
2 2

ρ(Ε) [arbitrary units]


ρ(E) [arbitrary units]
1
1

0
0 -3 -2 -1 0 1 2 3
-3 -2 -1 0 1 2 3 E / γpπ
E / γpπ

Figure 173: Typical forms of the density of states for semiconducting and metal-
lic nanotubes.

9 Electron-Electron Interactions
In the last chapter, the effects of interactions between electrons were neglected
when calculating the energies of single-electron excitations and the single-electron
wave functions. The neglect of electron-electron interactions is certainly not jus-
tifiable from considerations of the strength of the effect of the Coulomb interac-
tions due to the potential of the lattice of nuclei relative to the electron-electron
interactions. However, due to the Pauli exclusion principle, the lowest energy
excitations of an interacting electron gas can be put into a one to one correspon-
dence with the excitations of a non-interacting gas of fermions. The effects of
electron-electron interactions are weak for low-energy excitations and this leads
to the concept of treating the interacting electron system as a Landau Fermi
Liquid.

9.1 The Landau Fermi Liquid


The Pauli exclusion principle plays an important role in reducing the effect of
electron-electron interactions. An important result of this blocking principle is
that the low-energy excitations of an electron gas behave very similarly to those
of a non-interacting electron gas. This allows one to consider the low-energy
excitations as quasi-particles, which have a one to one correspondence with the
excitations of a non-interacting electron gas. This is the basis of the Landau
theory of Fermi liquids.

An important step in deriving the Landau theory was proved by J. M. Lut-


tinger84 , who showed that electrons have scattering rates that vanish as their
energies approach the Fermi energy, to all orders in the electron-electron inter-
action. This can already be be seen from the lowest order calculation of the
lifetime of an electron in a Bloch state due to electron-electron interactions. Al-
84 J. M. Luttinger, Phys. Rev. 121, 942 (1961), J. M. Luttinger, Phys. Rev. 119, 1153,
(1960).

307
though, a rigorous derivation of Fermi Liquid theory85 must consider processes
of all orders in the electron-electron interaction, we shall only consider the low-
est order processes. Consider the lowest order process, in which an electron,
initially in a state k above the Fermi surface, is scattered to a state k − q.
In this scattering process, a second electron is excited from an initial state k 0
below the Fermi surface to a state k 0 + q above the Fermi surface. This process
conserves momentum and will conserve energy if

h̄2 k 2 h̄2 ( k − q )2 h̄2 ( k 0 + q )2 h̄2 k 02


− = − (755)
2m 2m 2m 2m
or
( k − k0 ) . q = q2 (756)
0
For fixed k and k , the allowed values of the momentum transfer, q, are on the
surface of a sphere of diameter | k − k 0 | centered at the point ( k − k 0 ) / 2.
Thus, the momentum transfer, q, ranges from 0 to k − k 0 . Therefore, con-
servation of energy and momentum ensures that both of the electrons possible
final state momenta k − q and k 0 + q are located on the surface of sphere of
radius | ( k − k 0 ) | / 2, centered at the point ( k + k 0 ) / 2. This spherical
surface also passes through both the points k and k 0 , which correspond to the
electrons initial state momenta. It is seen that, if the electrons are initially
travelling parallel, so k ∼ k 0 , then the radius of the sphere is quite small.
Thus, for glancing collisions, the phase space of final states is quite small. By

k+q k'-q

k'

Figure 174: The range of allowed final states for the scattering of an electron in
state k with an electron of momentum k 0 inside the Fermi sphere. The allowed
states are constrained to lie in a ring on the spherical surface. For glancing
collisions, there is a small number of allowed final states.

comparison, the radius of the sphere can be larger for head-on collisions, where
85 P. Nozières and J. M. Luttinger, Phys. Rev. 127, 1423 (1962), J. M. Luttinger and P.

Nozières, Phys. Rev. 127, 1431 (1962).

308
k ∼ − k 0 . This result shows that there is a larger range of final states available

k'-q
k+q

k'

Figure 175: The range of allowed final states for the scattering of an electron in
state k with an electron of momentum k 0 inside the Fermi sphere. The allowed
states are constrained to lie in a ring on the spherical surface. For head on
collisions, there is a large number of allowed final states.

for head-on collisions.

Since electrons are fermions and the Fermi sphere is filled, the Pauli exclusion
principle further restricts the phase space of the final states. That is, both k − q
and k 0 + q must be above the Fermi surface. Hence, one has the additional
restrictions that
| k − q | ≥ kF (757)
and
| k 0 + q | ≥ kF (758)
Thus, only a segment of the surface of the sphere represents allowed final states
of possible electron scattering processes. The allowed segment is in the form of
a circular strip. The thickness of the strip becomes small as k approaches kF . In
the limit | k | → kF , this segment tends to a circle in the plane of intersection
of the sphere and the Fermi surface, unless of course k = − k 0 . The net result
is that the phase space available for the scattering process vanishes as k → kF ,
and the scattering rate vanishes86 .

9.1.1 The Scattering Rate


In considering scattering of an electron above the Fermi surface with electrons
below the Fermi surface, one must consider spin-dependent effects. The scat-
tering rate can be estimated from scattering processes in which the an electron
86 J. J. Quinn and R. A. Ferrell, Phys. Rev. 112, 812 (1958).

309
of spin σ is only scatters from electrons which have anti-parallel spins, since
the scattering rate from pairs of electrons with parallel spins is diminished due
to the exchange process. In fact, for short-ranged interactions, the scattering
rate for electrons with parallel spins vanishes identically. Hence, we shall only
consider the scattering of of electrons with anti-parallel spins.

We shall consider the finite temperatures scattering rate of an electron ini-


tially in the state (k, σ). At finite temperatures, there is a finite probability
that single-electron states with energies below µ are unoccupied, and also that
the single-electron states with energies above µ are occupied. The rate for scat-
tering the electron out of the state can be estimated by using the Fermi Golden
rule expression
2 X
4 π e2

1 2π 1 X
= fk0 ,−σ (1 − fk0 −q,−σ ) (1 − fk+q,σ )
τk h̄ V 2 q q 2 + kT2 F 0 k

× δ( Ek + Ek0 − Ek+q − Ek0 −q ) (759)

where the wave-vector kT F is the inverse Thomas-Fermi screening length, which


is given by  2 
2 2 m e2
 
kT F
= (760)
kF2 π h̄2 kF
The expression fk0 ,−σ in the scattering rate represents the Fermi function

1
f (Ek0 ) = (761)
exp[ β (Ek0 − µ) ] + 1

which represents the probability that the one-electron state (k 0 , −σ) is occupied.
The factors (1 − fk0 −q,−σ ) and (1 − fk+q,σ ) represent the probabilities that the
one-electron states (k + q, σ) and (k 0 − q, −σ) are unoccupied. These latter
two factors ensure that the final states are consistent with the Pauli exclusion
principle. The final factor in eqn(759) expresses the condition of conservation
of energy.

We shall evaluate the sum over the Bloch states k 0 . On denoting the energy
loss of the scattered electron by h̄ ω, via

h̄ ω = Ek − Ek+q (762)

then the electron-electron scattering rate becomes


2
4 π e2

1 2π X
= (1 − fk+q,σ )
τk V2 q q 2 + kT2 F
Z ∞
× dω IT (ω, q) δ( Ek − Ek+q − h̄ ω ) (763)
−∞

310
where the energies are independent of σ and the k 0 dependent factors have been
expressed as the integral
Z
V
IT (ω, q) = d3 k 0 fk0 (1 − fk0 −q ) δ( Ek0 − Ek0 −q + h̄ ω ) (764)
( 2 π )3
The integral can be evaluated in the limit of zero temperature, first by expressing
it in the form
q2
Z
V m mω
IT =0 (ω, q) = 3 2 d3 k 0 fk0 (1−fk0 −q ) δ( k 0 . q − + ) (765)
( 2 π ) h̄ 2 h̄
and then by noting that the Fermi functions reduce to step functions. The
values of k 0 at which the delta function is non-zero lie on the plane

q2 mω
k0 . q = − (766)
2 h̄
Therefore, the integral over k 0 is proportional to the area of the plane which is
interior to the Fermi sphere centered on the origin and is exterior to the sphere
of radius kF centered on q.

k'-q
q > 2 kF q

k'
ω/!
q . k' = q2/2 - mω

Figure 176: The range of integration of k 0 for IT =0 (ω, q) when q > 2kF . The
point k 0 must lie inside the Fermi sphere and must lie outside the sphere centered
on q, if the process is to obey the Pauli principle. Energy conservation constrains
k 0 to the plane surface.

For q > 2 kF , the spheres do not overlap. As a result the integral is


evaluated as the area of the plane enclosed by the Fermi sphere
  2 
V m 2 q mω
IT =0 (ω, q) = kF − −
( 8 π 2 ) h̄2 q 2 h̄ q
q mω
for kF > − > − kF (767)
2 h̄ q

311
and is zero otherwise, since in these other cases the surface does not intersect
with the Fermi sphere.

q < 2 kF
q k'-q

ω/!
q . k' = q2/2 - mω ! k'

Figure 177: The range of integration of k 0 for IT =0 (ω, q) for q < 2kF . En-
ergy conservation constrains the allowed values of k 0 to the surface. The Pauli
exclusion principle further restricts the points k 0 to an annulus in the plane.

For q < 2 kF , the two spheres overlap. If q − kF > 2q − mh̄ ωq > − kF


then, as before, the results is
  2 
V m 2 q mω
IT =0 (ω, q) = kF − −
( 8 π 2 ) h̄2 q 2 h̄ q
q mω
for q − kF > − > − kF (768)
2 h̄ q
since the plane does not intersect the region where the spheres overlap. However,
if 2q > 2q − mh̄ ωq > q − kF , the integral yields
 2  2 
V m q mω q mω
IT =0 (ω, q) = + − −
( 8 π 2 ) h̄2 q 2 h̄ q 2 h̄ q
q q mω
for > − > q − kF (769)
2 2 h̄ q
which is the area of the annulus bounded by the two spheres.

From the analysis of the zero temperature form, we note that for small
(positive) values of ω and q, one has

V m2 ω
IT =0 (ω, q) = (770)
4 π 2 h̄3 q
Furthermore, the function is zero for negative values of ω. This is due to the
Pauli exclusion principle and conservation of energy. The electron in state

312
(k 0 , −σ) involved in the T = 0 scattering process is initially inside the Fermi
sphere and since the only available final states are outside the Fermi sphere, the
electron must be excited in the scattering process. Hence, the electron in state
k must lose energy. We also note that the range of q for which the process takes
place is restricted by the condition
r r
2 2mω 2mω
kF + + kF > q > kF2 + − kF (771)
h̄ h̄

At finite temperatures, the integration can be performed by using the prop-


erty of the delta function, so
Z
V
IT (ω, q) = d3 k 0 f (Ek0 ) (1 − f (Ek0 + h̄ω)) δ( Ek0 − Ek0 −q + h̄ ω )
( 2 π )3
(772)
We shall express the product of Fermi functions as
  
f (Ek ) (1 − f (Ek + h̄ω)) = 1 + N (h̄ω)
0 0 f (Ek ) − f (Ek + h̄ω)
0 0 (773)

where N (x) is the Bose-Einstein distribution function


1
N (x) = (774)
exp[ βx ] − 1
Hence, we have found that
 Z  
V 3 0
IT (ω, q) = 1 + N (h̄ω) d k δ( Ek − Ek −q + h̄ ω ) f (Ek0 ) − f (Ek0 +h̄ω)
0 0
( 2 π )3
(775)
or equivalently
  Z
V
IT (ω, q) = 1 + N (h̄ω) d3 k 0 δ( Ek0 − Ek0 −q + h̄ ω ) ×
( 2 π )3
 
× fk0 (1 − fk −q ) − fk −q (1 − fk )
0 0 0

where we have again used the properties of the delta function and have intro-
duced a pair of terms that cancel. It should be noted that the first term in the
big round parenthesis has a form similar to the form IT (ω, q) given in eqn(764).
The second term is identified as having the form similar to IT (−ω, q). This
shows that our function satisfies the condition
  
IT (ω, q) = 1 + N (h̄ω) IT (ω, q) − IT (−ω, q) (776)

required from the consideration of the principle of detailed balance.

313
As we shall see later, in the evaluation of the scattering rate, it is justifiable
to make the approximation
  
IT (ω, q) ≈ 1 + N (h̄ω) IT =0 (ω, q) − IT =0 (−ω, q) (777)

We shall formally extend the definition of IT =0 (ω, q) to the negative axis, making
it an antisymmetric function. That is, we shall define

I˜T =0 (ω, q) = IT =0 (ω, q) for ω > 0 (778)

and
I˜T =0 (ω, q) = − IT =0 (−ω, q) for ω < 0 (779)
The scattering rate can then be expressed as
2 Z ∞
4 π e2
  
1 2π X
= dω 1 + N (h̄ω)
τk V2 q q 2 + kT2 F −∞

× (1 − f (Ek − h̄ω)) δ( Ek − Ek+q − h̄ ω ) I˜T =0 (ω, q)

where
V m2 ω
I˜T =0 (ω, q) ∼ (780)
4 π 2 h̄3 q
The integration over q can be evaluated by choosing the direction of k as the
polar axis. The integration over the direction of q can be performed by noting
that the integrand is independent of the azimuthal angle ϕ, and only depends
on cos θ through the argument of the delta function. We shall change variables
from cos θ to x, where x is defined by

h̄2
x = k q cos θ (781)
m
The angular integration is evaluated as
Z  
2πm q mω
dΩ δ( Ek − Ek+q − h̄ ω ) = 2 Θ k − 2 + h̄ q
(782)
h̄ k q

where Θ(x) is the Heaviside step function. Therefore,


Z ∞ 2 Z ∞
4 π e2
  
1 1 m
= dq q dω 1 + N (h̄ω)
τk ( 2 π ) h̄2 k V 0 q 2 + kT2 F −∞
 
q mω
I˜T =0 (ω, q)

× (1 − f (Ek − h̄ω)) Θ k − +
2 h̄ q
The integration over h̄ω is cut off at Ek − µ by the Fermi function factor, and at
− kB T by the Bose-Einstein factor. Therefore, the scattering rate is dominated
by the low-frequency processes, such that

EkF  | h̄ ω | (783)

314
The q integration is also dominated by the region

2 kF > q > 0 (784)

On using the approximate expression for I˜T =0 (ω, q), the scattering rate becomes
Z 2kF 2 Z ∞
m3 4 π e2
  
1 1 ω
= dq dω
τk ( 2 π )3 h̄5 k 0 q 2 + kT2 F −∞ 1 − exp[−βh̄ω]
 
1
× (785)
1 + exp[β(h̄ω − Ek + µ]

The integral over ω can be performed exactly using the identity


Z ∞
x 1 y2 + π2
dx =
−∞ (1 − exp[ − x ] ) ( 1 + exp[ x − y ] ) 2 ( 1 + exp[ − y ] )
(786)
Therefore,
Z 2kF 2
m3 4 π e2 ( π kB T )2 + ( Ek − µ )2

1 1
= dq
τk ( 2 π )3 2 h̄7 k 0 q 2 + kT2 F ( 1 + exp[−β(Ek − µ] )

We note that if we had used the bare Coulomb interaction the integral over q
would diverge, so the scattering rate would have diverged. The divergence has
been suppressed by the inclusion of screening. On performing the integral over
q, the scattering rate is evaluated as

kF ( e2 kF )2 ( π kB T )2 + ( Ek − µ )2
 
1 1 2kF
= 2 F
τk ( 8 π h̄ ) k ( h̄ kF kF T )3 kT F ( 1 + exp[−β(Ek − µ] )
2 m

where  
1 −1 x
F (x) = tan x + (787)
2 1 + x2
On introducing the electronic plasma frequency ωp via,

e2 kF3
ωp2 = (788)
3 π2 m
(ωp ∼ 1015 sec−1 ) one finds that, in the limit T → 0, the scattering rate can be
expressed as
  12 2
π2

1 3 Ek − µ
≈ ωp (789)
τk 128 π µ
Thus, the quasi-particle scattering rate vanishes as Ek → µ at zero temper-
ature. At finite temperatures, the quasi-particle scattering rate at the Fermi
energy87 varies as ( kB T )2 . In conclusion, we have indicated why the quasi-
particle concept may remain valid in the limit Ek → µ and T → 0.
87 E. Abrahams, Phys. Rev. 95, 834 (1954).

315
9.1.2 The Quasi-Particle Energy
The quasi-particle excitation energy Ek is affected by the interaction with the
other electrons in the system. The manner in which this change in energy
occurs can be exhibited using perturbation theory. For convenience, we shall
assume that the interaction between the electrons is a highly screened point
contact interaction. To second order in the perturbation, the approximate en-
ergy eigenvalue of the state where an electron is added to the Bloch state k is
given by
X Y Y
Ek+ = Ek0 + Ek0n + < k k n | Ĥint | k kn >
|kn |<kF |kn |<kF |kn |<kF
2
Q Q
|kn |<kF k n | Ĥint | k − q k m + q
< k
|kn |<kF ,n6=m k n >

X X
+
q
Ek0 + Ek0 0
− Ek−q − Ek0 +q
|km |<kF m m
2

< k Q Q
|kn |<kF k n | Ĥint | k, k m0 − q k m + q |kn |<kF ,n6=m,m0 k n >

X X
+
q
Ek0 + Ek0 − Ek0 −q − Ek0 +q
|km |,|km0 |<kF m0 m m0 m

(790)

To second order in the interaction, the ground state energy is given by


X Y Y
Egs = Ek0n + < k n | Ĥint | kn >
|kn |<kF |kn |<kF |kn |<kF
2
Q Q
X X < |kn |<kF k n | Ĥint | k m0 − q k m + q |kn |<kF ,n6=m,m0 k n >


+
Ek0 + Ek0 − Ek0 −q − Ek0 +q
q m,m0 m0 m m0 m

(791)

The excitation energy for adding an electron to state k is defined by the energy
difference

Ekexc = Ek+ − Egs (792)

To this order, the excitation energy is expressed in terms of two-particle states


as
X
Ekexc = Ek0 + < k k n | Ĥint | k k n >
|kn |<kF
2

< k k m | Ĥint | k − q k + q >
X X m
+
Ek0 + Ek0 0
− Ek−q − Ek0 +q
|k−q|>kF |km |<kF m m

316
2

< k + q k m | Ĥint | k k m + q >
X X
− 0
Ek+q + Ek0 − Ek0 − Ek0 +q
|k+q|<kF |km |<kF m m

(793)
The terms first order in the interaction represent the interaction of the particle
with the average density due to the other electrons. The last two terms are
second order terms. The first of this pair represents the scattering of a pair of
electrons in the initial states k above the Fermi energy and k m below the Fermi
energy. This pair of particles is scattered into the final states k − q and k m + q
which are both above the Fermi energy. The last term represents a subtraction,
as this represents a scattering process that is allowed in the ground state but
which is forbidden in the state where an extra electron is added to k. This
particular scattering process is forbidden since it violates the Pauli exclusion
principle. That is, the process whereby an electron is scattered from the ground
state to the state k is forbidden when the state k is already occupied by an elec-
tron. The k independent terms in the excitation energies are usually absorbed
into a shift of the Fermi energy.

The above expression represents the excitation energy for a state corre-
sponding to the non-interacting state in which one electron is added to the
system. Since, in the limit of large systems, the energy eigenvalues form a
quasi-continuum and since the interactions are turned on adiabatically, the
quasi-particle does not correspond to a single exact eigenstate but instead cor-
responds to a linear superposition of states with almost degenerate energies.
However, the corresponding many-body energy eigenstate consists of a linear
superposition of the state with an electron added to the ground state and states
where the added electron is scattered and dressed by many electron-hole pairs.
The quasi-particle weight Z −1 (k) is defined as the fraction of the initial bare
electron contained in this state. To lowest order in Ĥint , the quasi-particle
weight or wave function renormalization is calculated as
2

< k k m | Ĥint | k − q k m + q >
X X
Z(k) = 1 + 0 0 0 0 2
( Ek + Ek − Ek−q − Ek +q )
|k−q|>kF |km |<kF m m

2

< k + q k | Ĥint | k k + q >
X X m m
+ 0
( Ek+q + Ek0 − Ek0 − Ek0 +q )2
|k+q|<kF |km |<kF m m

(794)
which is greater than unity. Thus, the fraction of the bare electron in the quasi-
particle state is always less than unity88 . This conclusion remains valid to all
88 The quasi-particle is actually described as a wave packet constructed from the above exact

energy eigenstates.

317
orders of perturbation theory, if the Fermi Liquid phase is stable. When |k|
crosses kF , the quasi-particle changes from a quasi-particle to a quasi-hole. At
zero temperature due to the vanishing of the quasi-particle scattering rate, the
distribution of the number of bare particles has a discontinuity at the Fermi
energy of Z(k)−1 . This discontinuity is small compared with the discontinuity
for non-interacting electrons which is completely represented by the Fermi func-
tion. Thus, the concept of a Fermi surface remains well defined for interacting
electron systems.

The quasi-particle weight has the effect that the excitation energy for a single
quasi-particle is given by the expression
Ekexc
Eqp (k) = (795)
Z(k)
In addition to the shift in the excitation energy, the quasi-particle excitation
energy is reduced by Z(k) and these two effects combine to yield a flattening of
the quasi-particle’s dispersion relation. The reduced dispersion is interpreted in
terms of an increase in the effective mass of the quasi-particle. The density of
single-electron excitations has a quasi-particle contribution which is given by
X  
ρ(E) ≈ Z(k)−1 δ E − Eqp (k) (796)
k

where E is the excitation energy relative to the Fermi energy. Due to quasi-
particle weight factor, the single-electron density of states is narrowed and peaks
up near the Fermi energy. As the quasi-particles obey Fermi-Dirac statistics,
the quasi-particles can give rise to an enhancement of the coefficient of the linear
T term in the low-temperature electronic specific heat.

Despite the apparent simplicity of the Fermi Liquid picture, it is exceed-


ingly difficult to quantitatively derive the Fermi Liquid description appropriate
to a specific microscopic Hamiltonian. Since the perturbation due to electron-
electron interaction is long-ranged, there are divergent terms in the perturba-
tion expansion. The divergent terms first appear in the expansion when taken
to second order. The divergent terms can actually be re-summed to yield finite
results. The re-summations are made possible by the fact that the long-ranged
Coulomb interaction between a pair of electrons in a metal is screened by the
other electrons. The screening processes involves the Coulomb interaction to
infinite order. By taking into account the screening of the long-ranged Coulomb
interaction, the divergent terms can be summed to infinite order leading to
finite results. That is, the divergence associated with any term can be elim-
inated by combining it with a subset of other divergent terms. However, the
re-summation of all the terms in the perturbation expansion presents a serious
challenge and so approximations have been developed. These approximations
involve the summation of infinite subsets of the terms that appear in the pertur-
bation expansion. One such approximation is the Hartree-Fock approximation.

318
The Hartree-Fock approximation is self-consistent first-order perturbation the-
ory in that it just consists of the first order terms in the perturbation expansion.
However, in these terms, all the wave functions are calculated self-consistently
by taking the first-order processes into account.

——————————————————————————————————

9.1.3 Exercise 55
Using a perturbation expansion, find the energy of a free electron gas to first-
order in the electron-electron interaction.

——————————————————————————————————

319
9.2 The Hartree-Fock Approximation
The Hartree-Fock approximation consists of approximating the many-electron
wave function by a single Slater determinant. The resulting approximate wave
function has the same form as for independent or non-interacting electrons.
This should be contrasted with the exact wave function which is expected to be
composed of a linear superposition of Slater determinants. The Hartree-Fock
approximation, therefore, involves finding the best one-electron basis functions
that takes the average effect of electron-electron interactions into account89 .

The Hartree-Fock approximation can be expressed in terms of the Rayleigh-


Ritz variational principle90 , in which the many-particle wave function is written
as a single Slater determinant91 . The Hamiltonian operator is expressed as

X  p̂2 
1 X e2
i
Ĥ = + Vions (ri ) + (797)
2m 2 | ri − rj |
i i 6= j

The expectation value of the Hamiltonian in a state described by a single Slater


determinant Φ of a complete set of, as yet, unspecified single-electron wave
functions φα,σ (r) is written as
i=N
Ye  Z 
Ĥ = d3 ri Φ∗α1 , . . . αNe (r 1 , . . . rNe ) Ĥ Φα1 , . . . αNe (r 1 , . . . rNe )
i=1
(798)

The expectation value of the energy is evaluated as

h̄2
X Z  
3 ∗ 2
E = d r φα (r) − ∇ + Vions (r) φα (r)
α
2m
e2
Z Z
1 X
+ d3 r d3 r0 φ∗α (r) φ∗β (r0 ) φβ (r0 ) φα (r)
2 | r − r0 |
α,β

e2
Z Z
1 X
− 3
d r d3 r0 φ∗α (r) φ∗β (r0 ) φα (r0 ) φβ (r)
2 | r − r0 |
α,β
(799)

where the sums over α and β run over all the single-particle quantum numbers
labelling the Slater determinant Φ. The first term just represents the sum of
one-particle energies of the electrons. The second term represents the interaction
energy between an electron and the average charge density of all the electrons.
The last term is the exchange term; it arises due to the Coulomb interaction
and the anti-symmetry of the many-electron wave function. The spin indices
89 D. R. Hartree, Proc. Cambridge Philos. Soc. 24, 89 (1928).
90 V. A. Fock, Z. Physik, 61, 126, (1930).
91 J. C. Slater, Phys. Rev. 35, 210 (1930).

320
have been suppressed in the expression for the energy. The quantum number α
needs to be supplemented by the spin quantum number σ to uniquely specify the
state which will be written as either φα (r) = ψα (r) χσ or φ∗α (r) = χTσ ψα∗ (r).
Therefore, in the matrix elements there are not only an integrations over r, but
also the matrix elements of the spin states have to be evaluated.

The single-electron wave functions are to be chosen such that they minimize
the energy, subject to the constraint that they remain normalized to unity.
Hence, subject to this condition, the single-electron wave functions are chosen
such that the first order variation of the energy is identically equal to zero.
The minimization is performed by using the Lagrange method of undetermined
multipliers. First, one forms the functional Ω which is the average value of the
Hamiltonian minus the Ne constraints that ensure that the one-electron wave
functions are normalized to unity. The functional Ω is given by
i=N
Ye  Z 
Ω = d ri3
Φ∗α1 , . . . αNe (r 1 , . . . rNe ) Ĥ Φα1 , . . . αNe (r 1 , . . . rNe )
i=1
i=N
Xe  Z 
− λαi d3 ri φ∗αi (ri ) φαi (ri ) − 1 (800)
i=1

where the λα are the undetermined multipliers. Since φα is an arbitrary complex


function, the real and imaginary parts are independent. Instead of working with
the real and imaginary parts, we shall consider the function φα and its complex
conjugate φ∗α as being independent. The second step of the Lagrange method
consists of considering the effect of varying the set of φ∗α . The deviation of the
variational functions φ∗α (r) from the extremal function, φ∗HF,α (r), are denoted
by δφ∗α , i.e.,
φ∗α (r) = φ∗HF,α (r) + δφ∗α (r) (801)
To first order in the deviation δφ∗α (r), the expectation value of the functional Ω
changes to first order in δφ∗α by an amount δΩ. The change δΩ is evaluated as

h̄2
X Z  
3 ∗ 2
δΩ = d r δφα (r) − ∇ + Vions (r) − λα φHF,α (r)
α
2m
XZ e2
Z
+ 3
d r d3 r0 δφ∗α (r) φ∗HF,β (r0 ) φHF,β (r0 ) φHF,α (r)
| r − r0 |
α,β
XZ e2
Z
− d3 r d3 r0 φ∗HF,β (r) δφ∗α (r0 ) φHF,β (r0 ) φHF,α (r)
| r − r0 |
α,β
(802)

The expression for δΩ must vanish identically for any of the independent and ar-
bitrary variations δφ∗α (r), if the Hartree-Fock wave functions φHF,α (r) minimize
the average energy. In order for this to be true, the coefficient of δφ∗α (r) must
vanish identically for each value of α. After interchanging the variables r and

321
r0 in the last term, one finds that the normalized Hartree-Fock wave functions
must satisfy the set of equations

h̄2
 
2
0 = − ∇ + Vions (r) − λα φHF,α (r)
2m
e2
X Z  
+ d3 r0 φ∗HF,β (r0 ) φHF,β (r 0
) φHF,α (r)
| r − r0 |
β

e2
X Z  
− d3 r0 φ∗HF,β (r0 ) φHF,α (r 0
) φHF,β (r)
| r − r0 |
β
(803)

in order to minimize the energy. To simplify further analysis, we shall explicitly


display the spin-dependence by writing

φHF,α (r) = ψα (r) χσ


φHF,β (r) = ψβ (r) χσ0 (804)

This notation recognizes that the spatial component of the wave function, ψα (r),
may depend on all the quantum numbers represented by α, including the spin
quantum number, as in the un-restricted Hartree-Fock approximation. The
Hartree-Fock equations are re-written as

h̄2
 
0 = − ∇2 + Vions (r) − λα ψα (r) χσ
2m
e2
X Z  
3 0 T ∗ 0 0
+ d r χσ0 ψβ (r ) ψβ (r ) χσ 0 ψα (r) χσ
| r − r0 |
β

e2
X Z  
3 0 T ∗ 0 0
− d r χσ0 ψβ (r ) ψα (r ) χσ ψβ (r) χσ0
| r − r0 |
β
(805)

In the inner product, the integrations over the position r0 of the spatial com-
ponent of the wave function is combined with the matrix elements of the spin
wave functions. The spin matrix elements are given by

χTσ0 χσ = δσ0 ,σ (806)

Since the Coulomb interaction is spin-independent, that last term contains a


Kronecker delta function that is non-vanishing only when σ = σ 0 . The set of
Hartree-Fock equations are eigenvalue equations for a non-local linear operator

h̄2
 
0 = − ∇2 + Vions (r) − λα ψα (r)
2m
e2
X Z  
+ d3 r0 ψβ∗ (r0 ) ψ β (r 0
) ψα (r)
| r − r0 |
β

322
e2
X Z  
− δσ0 ,σ d3 r 0 ψβ∗ (r0 ) ψβ (r) ψα (r0 )
| r − r0 |
β
(807)

There is one such equation for each value of α. In solving the above equations
for ψα (r), one should consider the functions ψβ (r) as known quantities. In
this case, the eigenvalue equations are linear in the eigenfunctions, ψα , and the
undetermined multipliers, λα , are the eigenvalues. The term proportional to
X Z e2 | ψβ (r0 ) |2
Vdirect (r) = d3 r 0 (808)
| r − r0 |
β

represents a contribution to the potential


P from the average electrostatic potential
due to the electron density ρ(r0 ) = β |ψ β (r 0 2
)| from all the electrons in the
system. That is, this potential even includes the contribution from an electron
in the state α. This potential is independent of the spin states of the electrons,
and is called the direct interaction. The last term in the Hartree-Fock equation
(808) is non-local, as it relates the unknown eigenfunction ψα (r) to the weighted
average of the unknown eigenfunction at other points in space, ψα (r0 ). The non-
local potential represented by
X e2
σ
Vexch (r, r0 ) = − δσ,σ0 ψβ∗ (r0 ) ψβ (r) (809)
| r − r0 |
β

is called the exchange interaction. Since the Coulomb interaction is spin inde-
pendent, the matrix elements in the non-local exchange potential are non-zero
only if the spin of state α is identical to the spin of state β. If the spins are anti-
parallel, the exchange term is zero. Thus, the exchange term is spin-dependent.
The terms with β = α are spurious, since they represent the interaction
of an electron with itself. However, their contributions to the direct and ex-
change terms cancel exactly. Therefore, there are no self-interaction terms in
the Hartree-Fock approximation. This cancellation of the self-interaction has
the effect that the linear potential operator is the same for all the single-electron
wave functions.

With this notation, the Hartree-Fock equations can be written as

h̄2
  Z
− ∇2 + Vion (r) + Vdirect (r) ψα (r) + d3 r0 Vexch
σ
(r, r0 ) ψα (r0 ) = λα ψα (r)
2m
(810)
This set of equations can be solved iteratively. Using approximations for the
direct and exchange potentials, one can solve the equations to find a set of
wave functions which are approximations for the ψα (r). These approximate
wave functions are then used to construct new approximations for the direct
and exchange potentials. The procedure is repeated until self-consistency is

323
achieved. Roothaan92 has shown how, by introducing an appropriate set of
basis functions θm (r), one can express the Hartree-Fock wave functions ψα (r)
as
M
X
α
ψα (r) = Cm θm (r) (811)
m=1

and, hence, the Hartree-Fock equations can be reduced to a set of M coupled


simultaneous equations. The resulting set of non-linear simultaneous equations
are known as the Roothaan equations.

k'
k'

Hint
Hint
k-k'

k
k

Figure 178: Contributions to the Hartree-Fock interaction energy for a uniform


system. The process shown in (a) is the direct interaction or Hartree term. The
process in (b) is the exchange or Fock term.

The Hartree-Fock approximation can be solved exactly for the free electron
gas in which the potential of the lattice of ions is replaced by a constant value.
This (unrealistic) case of a uniform potential is of special importance, since the
solution is often used as a starting point to discussing the electronic structure
of a non-uniform electron gas. Specifically, the most common method of de-
termining electronic structure, the local density functional method, utilizes the
expression for the ground state energy of the uniform electron gas.

9.2.1 The Free Electron Gas.


The Hamiltonian for the free electron gas is invariant under all translations and,
as long as the translational symmetry is not spontaneously broken, the Hartree-
Fock eigenstates should be simultaneous eigenstates of the momentum operator.
Thus, the Hartree-Fock equations for a uniform potential Vions = V0 should
have the eigenfunctions
 
1
φα,σ = ψk,σ (r) χσ = √ exp i k . r χσ (812)
V
92 C. C. J. Roothaan, Rev. Mod. Phys. 23, 68 (1951).

324
where V is the volume of the crystal. That this is true can be seen by substi-
tuting the wave functions into the Hartree-Fock eigenvalue equations.

The charge density due to the electrons is a constant, and this combines with
the uniform charge density from the background gas of ions. Due to charge
neutrality, the resulting net direct Coulomb potential from the total charge
density vanishes
Vions (r) + Vdirect (r) = 0 (813)

In order to evaluate the exchange potential, one has to perform the sum
over values of k 0 , σ 0 . The sum over k 0 , σ 0 only runs over the occupied states.
We shall assume that the Hartree-Fock state does not spontaneously break the
spin rotational symmetry and lead to magnetism. Likewise, we shall also as-
sume that the Hartree-Fock solution does not break translational invariance.
Magnetic solutions which also break translation invariance have been found by
Overhauser93 . However, it is unclear whether theses inhomogeneous phases are
stable when screening is taken into account. Kohn and Nettel94 have pointed out
that the absolute stability of the broken-symmetry Hartree-Fock states requires
the presence of significant anisotropy (or low dimensions). In the non-magnetic
translationally invariant case, the Hartree-Fock states are spin degenerate, and
the one-particle states are filled according to the magnitude of the kinetic en-
ergy. All the one-particle states labelled by (k, σ), where k is contained inside
a sphere of radius kF , are filled with electrons. The spin-dependent exchange
term is evaluated as
e2
 
1 X
0
Vexch (r, r0 ) = − δ σ,σ 0 exp i k . ( r − r 0
)
V | r − r0 | 0 0
|k |≤kF , σ

(814)
The exchange potential also has translational invariance, and so it is possible
that plane waves are eigenfunctions of the Hartree-Fock equations. The ex-
change potential is evaluated as the integral
Z 2π Z π
1
Vexch (r, r0 ) = − dϕ dθ sin θ
( 2 π )3 0 0
Z kF
e2
 
0 02 0 0
× dk k exp i k . ( r − r )
0 | r − r0 |
1
= − 2 π e2
( 2 π )3
   
Z kF exp + i k 0 | r − r0 | − exp − i k 0 | r − r0 | !
× dk 0 k 0
0 i | r − r0 |2
93 A. W. Overhauser, Phys. Rev. Lett. 4, 462 (1960), Phys. Rev. 128, 1437 (1962).
94 W. Kohn and S. J. Nettel, Phys. Rev. Lett. 5, 8 (1960).

325
(815)
The integration over k 0 can be performed with the aid of an identity obtained
by differentiating the expression
Z 1
sin α
dx cos α x = (816)
0 α
with respect to α. That is,
Z 1  
sin α cos α
dx x sin α x = − (817)
0 α2 α
The resulting expression for the exchange potential is
!
0 e2 kF4 sin kF | r − r0 | cos kF | r − r0 |
Vexch (r, r ) = − 0

2 π2 ( kF | r − r | )4 ( kF | r − r0 | )3
(818)
The long-ranged oscillatory behavior of the exchange potential is due to the
sharp cut off of the integration at kF . This cut off occurs as the Fermi wave
vector kF is the largest wave vector associated with the occupied one-electron
states.

0.01
Vexc(r) [ units of e2 kF4/(2π2) ]

-0.01

-0.02

-0.03

-0.04
0 2 4 6 8 10 12 14 16
kF r

Figure 179: The radial dependence of the exchange potential for the free electron
gas. This plot emphasizes the oscillations due to the Fermi cut off and does not
show the 1/r divergence at the origin.

The contribution of the exchange potential to the energy eigenvalue λk can


be found from
Z   Z  
1
d3 r0 Vexch (r, r0 ) ψk (r0 ) = √ exp i k . r d3 r0 exp i k . ( r0 − r ) Vexch (r, r0 )
V
(819)

326
Thus, the contribution of the eigenvalue stemming from exchange potential is
just the Fourier transform of the exchange term, Vexch (k),
!
e2 kF4
Z  
3 sin kF R cos kF R
Vexch (k) = − d R exp i k . R −
2 π2 ( kF R )4 ( kF R )3
(820)
which can be evaluated directly. An alternate method involves using the con-
volution theorem, in which case the expression

e2
Z
V 3 0
Vexch (k) = − d r
( 2 π )3 | r − r0 |
Z  
3 0 ∗ 0 0
× d k ψk0 (r ) ψk0 (r) exp i k . ( r − r )
|k0 |≤kF
(821)

can be used. The plane wave nature of the eigenfunctions can be utilized to
write the expression for Vexch as

e2
Z
V 3 0
Vexch (k) = − d r ×
( 2 π )3 | r − r0 |
Z  
3 0 0 2 0 0
× d k | ψk0 (r ) | exp i ( k − k ) . ( r − r )
|k0 |≤kF
(822)

The electron density, per spin, arising from state k is just | ψk0 (r0 ) |2 = V1
for | k 0 | ≤ kF . Since this is independent of r0 , the exchange contribution to
the eigenvalue involves the Fourier Transform of the Coulomb potential. The
Fourier transform of the exchange potential is found as

e2
Z Z  
1 3 0 3 0 0 0
= − d r d k exp i ( k − k ) . ( r − r )
( 2 π )3 |k0 |≤kF | r − r0 |
(823)

Hence, the expression for the exchange contribution to the eigenvalue λk is given
by

4 π e2
Z
1
Vexch (k) = − d3 k 0
(2π) 3
|k0 |≤kF | k − k 0 |2
kF
e2 | k + k0 |
Z
= − dk 0 k 0 ln (824)
πk 0 | k − k0 |

The integral can be evaluated as

2 e2
 
k
Vexch (k) = − kF F (825)
π kF

327
where
1 1 − x2 |1 + x|
F (x) = + ln (826)
2 4x |1 − x|
At k = 0, the function F (0) is unity. At k = kF , the function falls to the
value F (1) = 12 and has a logarithmic singularity in the slope. This singularity
in the slope is due to the long-ranged nature of the Coulomb interaction ( 4k2π ).
The function F (x) falls to zero in the limit limx → ∞ F (x) → 0. Thus, the
eigenvalue λk is given by

h̄2 k 2 2 e2
 
k
λk = − kF F (827)
2m π kF

0.5 rs=3
λk [ Rydbergs ]

0
rs=4

-0.5

-1
0 0.5 1 1.5
k / kF

Figure 180: The Hartree-Fock expression for the dispersion relation λk for free
electrons. The curves corresponding to the different rs values correspond to
different values of the electronic density.

The total energy of the electron system is given by the sum of the kinetic
energy and the exchange energy of the occupied states
X h̄2 k 2
EHF = 2
2m
k

e2
X Z Z
− d3 r d3 r0 ψk∗ (r) ψk∗0 (r0 ) ψk (r0 ) ψk0 (r)
| r − r0 |
k,k0
X  h̄2 k 2 
= + λk (828)
2m
k

where the summations are restricted to the values of k and k 0 which are within

328
the Fermi sphere. The Hartree-Fock energy can be re-expressed as
X h̄2 k 2
EHF = 2
2m
k ≤ kF

2 e2 kF2 − k 2
   
X 1 | kF + k |
− kF + ln
π 2 4 k kF | kF − k |
k ≤ kF
(829)

The summations over k can be evaluated by transforming them into integrals


over the volume of the Fermi sphere

h̄2 k 2
Z
4πV
EHF = 2 dk k 2
( 2 π )3 k ≤ kF 2m
2 e2 kF2 − k 2
   
| kF + k |
Z
4πV 2 1
− kF dk k + ln
π ( 2 π )3 k ≤ kF 2 4 k kF | kF − k |
V h̄2 kF5 V e2 4
 
1 1
= 2
− kF + (830)
π 10 m π3 3 6

Ne
The number of electrons, per spin, 2 is given by

Ne V 4π 3
= k (831)
2 8 π3 3 F
Using this, the Hartree-Fock approximation for the cohesive energy of the free
electron gas can be expressed as

3 h̄2 kF2 3 e2
 
EHF = Ne − kF (832)
5 2m 4 π

An alternative expression is given by introducing a characteristic dimension, or


radius rs , such that there exists one electron in a sphere of radius rs a0 , where
2
a0 is the Bohr radius (a0 = mh̄ e2 ). Then, the uniform electron density, ρ, is
given by the equivalent expressions
1 4π 3 3
= a r
ρ 3 0 s
3 π2
= (833)
kF3

Thus, the magnitude of the Fermi wave vector kF is given by


  13
9π 1
kF = (834)
4 rs a0

329
and so the electronic energy is expressed as
2 1
3 h̄2 3 e2
 
EHF 9π 3 1 9π 3 1
= −
Ne 10 m a20 4 rs2 4 π a0 4 rs
2
  13    13 
e 9π 3 9π 1 3 1
= −
2 a0 4 5 4 rs2 2 π rs
2.2099 0.9163
= − Rydbergs (835)
rs2 rs
2
where 1 Rydberg = 2ea0 . The Hartree-Fock energy has a minimum at the rs
value given by rs ∼ 4.8 which corresponds to a value of the cohesive energy
which is about 0.1 Rydbergs. Typical materials have spatially varying densities,
hence, the local value of rs also varies. For a hydrogen-like atom, the ground
state density is given by

Z3
 
2Z r
ρ(r) = exp − (836)
π a30 a0

Therefore, typical values of rs correspond to the density at the nuclear position


1
( 34 ) 3
rs =
Z
0.9086
= (837)
Z
or to the density at the first Bohr radius r = Z a0
1 2
( 34 ) 3 e 3
rs =
Z
1.7696
= (838)
Z
Since for metals the density of electrons corresponds to rs values in the range
of 2 to 5, the exchange term is of similar magnitude to the kinetic energy term.
The Hartree-Fock approximation indicates that the cohesive energy is largest
for low density metals, i.e., those with rs ∼ 5.

In the particular case of the free electron gas where the potential due to the
ions is uniform, the Hartree-Fock approximation coincides with second order
perturbation theory. If an infinite number of higher order terms are included in
the perturbation series95 , one obtains the expression for the energy per electron
 
E 2.2099 0.9163
= − + 0.06218 ln rs − 0.094 + O(rs ) (839)
Ne rs2 rs
95 M. Gell-Mann and K. A. Brueckner, Phys. Rev. 106, 347, (1957), W. J. Carr and A. A.

Maradudin, Phys. Rev. A 133, 371 (1964).

330
Hartree-Fock Total Energy
0.08

E(rs)/Ne [ units of e /(2a0) ]


2
0

-0.08

-0.16
0 2 4 6 8
rs

Figure 181: The Hartree-Fock approximation for cohesive energy of the free
electron gas.

2
in units of 2ea0 . The energy is a form of an expansion in rs , valid for rs < 1.
Thus, the Hartree-Fock result can be thought of as an approximation which
reproduces the high density limit (small rs limit) correctly. The other terms in
the expression are due to electron correlations. A completely different behavior
is expected to occur in the low density limit. In reducing the density from the
high density metallic limit to the low density limit, the system is expected to
undergo a transition to a Wigner crystal phase96 . In a Wigner crystal, the
electrons are expected to localize in a b.c.c. structure. The total energy is
expected to be dominated by the electrostatic interaction and the energies of
the vibrations of the electronic lattice97 . The energy of the Wigner crystalline
phase is given by
e2
 
E 1.792 2.65 0.73
= − + 3 − + . . . (840)
Ne 2 a0 rs rs2 rs2

for rs  1.

The electronic wave functions described by a Slater determinant are not


devoid of correlations. The correlations are a result of the Pauli exclusion prin-
ciple. The two-particle density-density correlation function for a single Slater
determinant can be written as
X 1
ρ2 (r, r0 ) = | φα (r) φβ (r0 ) − φβ (r) φα (r0 ) |2
2
α,β
X X X
= | φα (r) |2 | φβ (r0 ) |2 − φ∗α (r) φβ (r) φ∗β (r0 ) φα (r0 )
α β α,β

96 E. P. Wigner, Phys. Rev. 46, 1002 (1934).


97 W. J. Carr, R. A. Coldwell-Horsfall, and A. E. Fein, Phys. Rev. 124, 747 (1961).

331
(841)
On making the spin dependence explicit, by writing
φα (r) = ψα (r) χσ
φβ (r) = ψβ (r) χσ0 (842)
one finds that the two-particle density-density correlation function is given by
X X X
ρ2 (r, r0 ) = | ψα (r) |2 | ψβ (r0 ) |2 − δσ,σ0 ψα∗ (r) ψα (r0 ) ψβ∗ (r0 ) ψβ (r)
α β α,β
X
0
= ρ(r) ρ(r ) − Gσ (r , r) Gσ (r, r0 )
0
(843)
σ

where Gσ is given by a sum over the single-particle states labelled by α which


have the spin quantum number σ
X
Gσ (r, r0 ) = ψα∗ (r0 ) ψα (r) (844)
α

The last term in the two-particle density-density correlation function is the ex-
change term. The exchange term originates from pairs of electrons with parallel
spins. In the Hartree-Fock approximation for the free electron gas, the exchange
contribution to the two-particle density-density correlation function ρ2 (r, r0 ) is
expressed in terms of the factors
X
Gσ (r, r0 ) = ψk∗0 (r0 ) ψk0 (r)
|k0 | < kF
 
1 X
= exp i k0 . ( r − r0 )
V
|k0 | < kF

kF3 sin kF | r − r0 | cos kF | r − r0 |


 
= −
2 π2 ( kF | r − r0 | )3 ( kF | r − r0 | )2
(845)
where the summation is over the Fermi sphere. The density-density correlation
function shows a hole in the density of parallel spin electron around the electron
and vanishes as | r − r0 | → 0, as expected from the Pauli exclusion principle.
The exchange hole describes the exclusion of just one electron. The exchange
potential has a similar form to the density-density correlation function and can
be thought of arising from a deficiency in the density of parallel spin electrons
around an electron at r. The Hartree-Fock approximation is deficient in that it
does not include a similar correlation hole between electrons with anti-parallel
spins98 .

98 The effect of the Coulomb interaction between pairs of electrons (with anti-parallel spins)

is also expected to produce a correlation hole. Recent experiments in which a single photon
produces the emission of two electrons, via a process which necessarily involves inter-electronic
interactions, allows for a mapping of the exchange-correlation hole. [F. O. Schuhman et al.,
Physical Review B 73, 041404 (2006).]

332
1.2

0.8

2
ρ2(r)/ρ
ρ
0.6

0.4

0.2

0
0 1 2 3 4 5
kF r

Figure 182: The radial dependence of the two-particle density-density correla-


tion function for the free electron gas, as calculated in the Hartree-Fock approx-
imation. The Pauli-exclusion principle causes the density-density correlation
function at r = 0 to fall to half of the asymptotic value.

In the Hartree-Fock approximation, the energies of the excited states are


given by Koopmans’ theorem99 . That is, the energy for adding or removing an
electron from the system is given by the eigenvalue λk , if the other one-electron
states in the many-particle Slater determinant are not changed or that the other
electrons in the ground state are not re-arranged. Thus, in the Hartree-Fock

Fermi-level
30
ρqp(E) [States/Rydberg]

25
non-interacting

20

15
Hartree-Fock
10

0
-0.5 0 0.5 1 1.5
E [Rydbergs]

Figure 183: The Hartree-Fock approximation for the quasi-particle density of


states for a free electron gas.

99 T. A. Koopmans, Physica 1, 104 (1933).

333
approximation, the quasi-particles energies are given by
Eqp (k) = λk − µ (846)
The zero of the quasi-particle energy is chosen to be the renormalized Fermi
energy. The quasi-particle density of states, per spin, is given by
X
ρqp (E) = δ( E − Eqp (k) )
k
Z kF
V
= dk k 2 δ( E − Eqp (k) )
2 π2 0
 −1
V dEqp (k)
k2

= (847)
2 π2 dk
k(E)

where k(E) is the value of k that satisfies the equation


Eqp (k) = E (848)
The Fermi energy is defined by
Eqp (kF ) = 0 (849)
which fixes µ in terms of kF . At the Fermi energy, the quasi-particle density of
states is zero since
dEqp (k) h̄2 k e2 kF
= −
dk m π k
e2 ( kF2 + k 2 ) | kF + k |
+ ln (850)
π 2 k2 | kF − k |
which diverges logarithmically at k(E = 0) = kF . Thus, the Hartree-Fock ap-
proximation for the free electron gas is of limited utility in discussing properties
of real metals. This is caused by the divergent slope of the one-electron eigen-
values near the Fermi surface. This spurious divergence is due to the neglect of
screening and results in the one-electron density of states falling to zero just at
the Fermi energy.

——————————————————————————————————

Broken Symmetry: The Spiral Spin Density Wave State

The spiral spin density wave state is an example of a state with combined
broken spin rotational invariance and translational invariance. In the Hartree-
Fock approximation, it is represented by a single Slater determinant Φ composed
of one-electron states of the form
uk vk
φk,+ (r) = √ exp[ i ( k − Q/2 ) . r ] χ↑ + √ exp[ i ( k + Q/2 ) . r ] χ↓
V V
vk∗ u∗k
φk,− (r) = √ exp[ i ( k − Q/2 ) . r ] χ↑ − √ exp[ i ( k + Q/2 ) . r ] χ↓
V V
(851)

334
where uk and vk are complex coefficients that have yet to be determined. It
should be noted that k serves to label the one-electron wave functions but does
not represent the momentum of the electron. The above one-electron states
form an orthonormal basis set if

| uk |2 + | vk |2 = 1 (852)

For uk ≡ 1 and vk ≡ 0, the broken-symmetry Hartree-Fock state reduces


to the Hartree-Fock approximation to the free electron gas. The spin density is
defined as

− h̄ X ∗
S (r) = φk,± (r) −

σ φk,± (r) fk,± (853)
2
k,±

where →−
σ is the vector Pauli spin matrix and where fk,± are the occupation
numbers of the broken-symmetry one-electron states. The broken-symmetry
state may have a spiral spin density, since the components are evaluated as
 
h̄ X
S z (r) = | uk |2 − | vk |2 ( fk,+ − fk,− ) (854)
2V
k

and
 
h̄ X
S x (r) = uk vk∗ exp[ − i Q . r ] + vk u∗k exp[ + i Q . r ] ( fk,+ − fk,− )
2V
k
 
h̄ X ∗ ∗
S y (r) = i uk vk exp[ − i Q . r ] − vk uk exp[ + i Q . r ] ( fk,+ − fk,− )
2V
k

(855)

The z-component of the spin density, just like the charge density, is uniform for
T
this state. We shall define a complex number with amplitude MQ and phase ϕ
via
h̄ X
MQ exp[ i ϕ ] = vk u∗k ( fk,+ − fk,− ) (856)
V
k

so that the components of the spin density can be simply expressed as

S x (r) = MQ cos( Q . r + ϕ )
y
S (r) = MQ sin( Q . r + ϕ ) (857)

If the amplitude MQ is non-zero, the static magnetic spin density spirals in the
x − y plane with wave vector Q. Hence, the state has a broken spin-rotational
invariance and translational invariance if the product vk u∗k 6= 0.

The coefficients uk and vk are to be determined by minimizing the energy.


The energy of the broken-symmetry Hartree-Fock state is given by minimizing
the expectation value
< Φ | Ĥ | Φ > (858)

335
with respect to variations of the complex coefficients uk and vk subject to the
normalization conditions given by eqn(852). Instead of considering the real and
imaginary parts of the coefficients uk and vk as independent variables, we shall
consider the pair of coefficients uk and vk and their complex conjugates u∗k and
vk∗ as independent. The minimization is performed by using Lagrange’s method
of undetermined multipliers, which minimizes the functional
X  
Ω[uk , u∗k , vk , vk∗ ] = < Φ | Ĥ | Φ > − fk,± λk,± | uk |2 + | vk |2 − 1
k,±
(859)
where λk,± are the undetermined multipliers. The functional is evaluated as
X  
0 2 0 2
Ω = fk,+ Ek−Q/2 | uk | + Ek+Q/2 | vk |
k
X  
0
+ fk,− Ek−Q/2 | vk |2 + Ek+Q/2
0
| uk |2
k
2
4 π e2
 
1 X ∗ ∗

− ( fk,+ fk0 ,+ + fk,− fk0 ,− ) 0 2
uk uk0 + vk vk0
2 V |k − k |
k,k0
2
4 π e2
 
1 X
− ( fk,+ fk0 ,− + fk,− fk0 ,+ ) 0 2
uk vk0 − vk uk0
2 V |k − k |
k,k0
X  
− fk,± λk,± | uk |2 + | vk |2 − 1 (860)
k,±

The first two terms represent the kinetic energies of the independent electrons.
The energy due to the direct Coulomb interaction is assumed to cancel with
the corresponding interaction energy due to the positive uniform charge back
ground. The third and fourth terms represent the exchange part of the pairwise
Coulomb interactions between the electrons. The last term represents originates
from the normalization condition and vanishes if the one-electron wave functions
are properly normalized. The Lagrange undetermined multipliers λk,± are to
be identified with the quasi-particle energies. The occupation numbers fk,± are
one or zero, depending on whether or not the corresponding one-electron wave
functions φk,± (r) are included in the Slater determinant. Therefore, the fk,±
can be treated as independent quantities.

On requiring that Ω be an extrema with respect to a variation of u∗k , one


obtains
4 π e2
 X   
0 2 2
0 = Ek−Q/2 − λk,+ − fk ,+ | uk | + fk ,− | vk |
0 0 0 0 uk
V | k − k 0 |2
k0
X  4 π e2
 

− ( f 0
k ,+ − f 0
k ,− ) u k 0 v 0 vk (861)
0
V | k − k 0 |2 k
k

336
Similarly, varying Ω with respect to vk∗ and demanding that the first-order vari-
ation is zero, yields

4 π e2
 X   
0 2 2
0 = Ek+Q/2 − λk,+ − f 0
k .+ | vk 0 | + f 0
k ,− | u k 0 | vk
V | k − k 0 |2
k0
X  4 π e2
 

− ( fk ,+ − fk ,− ) vk uk0
0 0 0 uk (862)
0
V | k − k 0 |2
k

On defining the real k-dependent quantities


X  4 π e2

∆0 (k) = ( fk0 ,+ + fk0 ,− ) (863)
0
V | k − k 0 |2
k

and
X  4 π e2

∆± (k) = ( fk0 ,+ − fk0 ,− ) ( | uk0 |2 − | vk0 |2 ) (864)
V | k − k 0 |2
k0

which represent the first-order shifts of the bands and the complex quantity
X  4 π e2

∆(k) = ( fk0 ,+ − fk0 ,− ) uk0 vk∗0 (865)
0
V | k − k 0 |2
k

which represents the mixing of the bands, the equations can be re-written as
 
0 1 1
0 = Ek−Q/2 − ∆0 (k) − ∆± (k) − λk,+ uk − ∆(k) vk
2 2
 
1 1
0 = 0
Ek+Q/2 − ∆0 (k) + ∆± (k) − λk,+ vk − ∆∗ (k) uk
2 2
(866)

The above set of equations can be solved for the eigenvalues λk,+ , yielding
s
 E0 0  E0 0
− Ek+Q/2 − ∆± (k) 2
k−Q/2 + Ek+Q/2 − ∆0 (k)

k−Q/2
λk,+ = ± + | ∆(k) |2
2 2
(867)
We shall choose the solution with the lower sign, since we have assumed that
the states with weight fk,+ are occupied in the ground state. The undeter-
mined multipliers λk,− can be found by an analogous procedure, however, in
this case, the solution corresponding to the upper sign should be chosen. The
kz -dependence of the eigenvalues λk,± is sketched in fig(184). It is seen that
the dispersion relation consists of two branches which asymptotically approach
0 0
Ek−Q/2 and Ek+Q/2 , and are separated by a gap of 2 ∆(0) at k = 0. Since the
interaction is a property of the system as a whole, it seems a reasonable guess

337
1.5

λk-
λk

0.5

λk+
0
-1.5 -1 -0.5 0 0.5 1 1.5

-0.5
kz/Q

Figure 184: The form of the quasi-particle dispersion relations in a spiral spin
density wave state in label space.

that the states (k, ±) should be filled in the order of increasing quasi-particle
energies λk,± . The functions | uk |2 and | vk |2 are determined as
0 0
1
 Ek−Q/2 − Ek+Q/2 − ∆± (k) 
| vk |2 = 1 + q
2 0
( Ek−Q/2 0
− Ek+Q/2 − ∆± (k) )2 + 4 | ∆(k) |2
0 0
1
 Ek−Q/2 − Ek+Q/2 − ∆± (k) 
| uk |2 = 1 − q
2 0
( Ek−Q/2 0
− Ek+Q/2 − ∆± (k) )2 + 4 | ∆(k) |2
(868)

These functions, respectively, represent the probability density for finding spin-
down or spin-up electrons in the states of the lower quasi-particle branch. Hence,
the states in the lower branch are primarily spin-up when the label k ≈ Q/2 and
are primarily spin-down when k ≈ −Q/2. The roles of uk and vk are reversed
for the upper branch of quasi-particles. In momentum-space, the mixed spin-
character of the states at the label value kz = 0 corresponds to up-spin electrons
at the back of the Fermi-surface being mixed with down-spin electrons at the
front of the Fermi-surface. Therefore, the spiral spin density wave state roughly
corresponds to a state which is described in our k-label space as having an up-
spin Fermi-sphere centered on k = Q/2 and a down-spin Fermi sphere centered
on k = −Q/2. However, the “Fermi-surfaces” are ellipsoids of revolution that
are expected to merge and then finally a new sheet will re-emerge as the wave
vector Q is reduced from Q > 2kF to 2kF < Q. The formation of the new sheet

338
1.2

| vk | 2 | uk |2
0.8
2
| uk |

0.6

0.4

0.2

0
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
kz/Q
Figure 185: The probability density for finding spin-up |uk |2 and spin-down
|vk |2 electrons in the lower branch of quasi-particles in the spiral spin density
wave state.

corresponds to the filling of the upper band. The product uk vk∗ is evaluated as

1.5
kx/kF

0.5

kz/kF
0
-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5 2 2.5

-0.5

-1

-1.5

Figure 186: The ky = 0 section across the “Fermi surfaces” of spiral density wave
states, for ordering waves vector Q = 2.2kF (blue), Q = 2kF and Q = 1.8kF
(red).

339
∆(k)
uk vk∗ = q (869)
0 0
( Ek−Q/2 − Ek+Q/2 − ∆± (k) )2 + 4 | ∆(k) |2

On substituting this product into the definition of the gap ∆(k), we find the
gap equation
X  4 π e2 ( fk0 ,+ − fk0 ,− ) ∆(k 0 )

∆(k) =
V | k − k 0 |2
q
k0 ( Ek00 −Q/2 − Ek00 +Q/2 − ∆± (k 0 ) )2 + 4 | ∆(k 0 ) |2
(870)
The spiral magnetization only occurs if the gap equation has a non-zero solution
for ∆(k). It should be noted that the absolute value of the phase ϕ of the
gaps ∆(k) drop out of the gap equation and so the phase can be considered
as being arbitrary. In fact, the absolute phase is spontaneously chosen by the
system when the system condenses into the spiral spin density wave state. The
absolute phase is expected to be determined by the interaction between the
spiral magnetization and the impurities and surfaces in the solid. The relative
shift of the quasi-particle bands is found to be given by
0 0 0
k ,+ − fk0 ,− ) ( Ek0 −Q/2 − Ek0 +Q/2 − ∆± (k ) )
(f 0
X  4 π e2
∆± (k) = −
V | k − k 0 |2
q
k0 ( Ek00 −Q/2 − Ek00 +Q/2 − ∆± (k 0 ) )2 + 4 | ∆(k 0 ) |2
(871)
The set of equations (870), (871) and (863) are to be solved self-consistently for
arbitrary distributions of fk,± which conserve the total number of electrons.

It might be expected that for non-zero values of ∆(k), the solution with
lowest energy corresponds to the spiral state with Q = 2kF since, in this case,
the electrons only occupy states in the lower sub-band which have quasi-particle
energies lower than the free-electron energies due to the opening of the gap. This
possible optimal value of Q could be considered quite reasonable since 2kF is
the only momentum scale in the system. On the other hand, the system can also
be viewed as having a static spiral magnetization which, through the exchange
interaction, simultaneously scatters and flips the spins of the electrons. That
is, through the electronic interactions, the electrons produce an effective spin-
dependent potential V (r, −→
σ ) of the form
 

− 1 ∗ + −
V (r, σ ) = V (Q) exp[ − i Q . r ] σ + V (Q) exp[ + i Q . r ] σ
2
 
x y
= | V (Q) | cos( Q . r + ϕ ) σ + sin( Q . r + ϕ ) σ


∝ S (r) . −

σ (872)

Since the Hartree-Fock approximation is self-consistent, the spin-dependent po-


tential must yield the spiral magnetization. This leads one to expect that the
broken-symmetry state will be more stable with Q ≈ 0, at least for small values

340
of the spiral magnetization, since the appropriate response function is largest
when Q = 0. The reason why the response function is largest for Q ≈ 0 is sim-
ply because the number of electrons available for elastic scattering across the
Fermi-surface involving small momentum transfers Q = 2 kF cos θ, where the
polar angle is θ ≈ π2 , is proportional to the area 2πkF2 sin θ ∆θ. The available
area for small Q is large compared to the corresponding area for large momen-
tum transfers, for which θ ≈ 0. The number of electrons available for scattering
through large momenta is proportional to the area πkF2 sin θ ∆θ and, therefore,
vanishes as ∆θ2 . That is, the response function at large momentum transfers
is small since it rapidly cuts off for momentum transfers equal to the diameter
of the Fermi-sphere Q = 2kF . This reasoning leads one to expect that, if the
symmetry breaking is sufficiently weak, the ferromagnetic state should be stable.

The effective spin-dependent potential only has components at ±Q, and


this is responsible for the unusual form of the quasi-particle dispersion rela-
tion. In momentum-space, one has up-spin electrons being spin-flip scattered
through momentum +Q. The effective potential with momentum component

kF
Q

Figure 187: A three-dimensional depiction of the momentum-space Fermi-


surface of a spiral density wave state with wave vector Q. The Fermi-surface is
gapped in annular regions centered on ±Q/2.

Q results in the dispersion relation for spin-up electrons being flattened at the
value −Q/2. Likewise, since the spin-down electrons are spin-flip scattered
backward through −Q, the dispersion relation for spin-down electrons is flat-
tened at +Q/2 in momentum-space. Hence, for a spiral spin density wave state,
the quasi-particle mass enhancements at the Fermi-surface are anisotropic and
spin-dependent. Furthermore, the spiral spin density wave state should have
small annular regions on the Fermi-surface which are gapped.

——————————————————————————————————

341
9.2.2 Exercise 56
Show, using perturbation theory, that the second-order correction to the energy
of a free electron gas from the interaction between electrons with anti-parallel
spins is given by
2
4 π e2

m X 1
∆E (2) = − (873)
2
h̄ k,k0 ,q 2
q V q . ( k − k0 + q )

where k < kF , k 0 < kF , | k + q | > kF and | k 0 − q | > kF . Since this


integral is dominated by the region q → 0, the dominant values of k and k 0
are given by k ∼ kF and k 0 ∼ kF . Show that the contribution to ∆E (2) is
proportional to

d3 q
Z Z
dq
= 4π
q3 q
= 4 π ln q (874)

and, thus, diverges for q → 0.

——————————————————————————————————

The second-order contribution to the energy of the electron gas can be ex-
pressed as
2
4 π e2 fk fk0 ( 1 − fk−q ) ( 1 − fk0 +q )
X 
∆E (2) =
V q2 Ek0 + Ek00 − Ek=q
0 − Ek00 +q
k,σ;k0 ,σ 0 ;q

4 π e2 4 π e2 fk fk0 ( 1 − fk−q ) ( 1 − fk0 +q )


X   

V q2 V |k − k 0 − q|2 Ek0 + Ek00 − Ek−q
0 − Ek00 +q
k,k0 ,σ;q

(875)

where fk is one or zero, depending on whether k < kF or not. The first term is
the direct Coulomb interaction and the second term only acts between electrons
with the same spin and represents the exchange interaction. The second order
exchange contribution can be evaluated analytically100 and expressed in units
of Rydbergs per electron as

e2 e2
 
1 3
ln 2 − ζ(3) ∼ 0.0484 (876)
2 a0 3 2 π2 2 a0

where ζ(n) is Riemann’s zeta-function. The above result for the second-order
exchange energy is independent of rs .

100 L. Onsager, L. Mittag and M.J. Stephen, Ann. Physik 18, 71 (1966).

342
Simple second-order perturbation theory does not work for the free electron
gas. In fact, the energy corrections from each higher order in the perturbation
series diverges. None the less, perturbation theory can be applied by using more
elaborate techniques which take into account the screening of the Coulomb in-
teraction.

343
9.3 The Density Functional Method
The density functional method provides an exact method for calculating the elec-
tron density and ground-state energy for interacting electrons in the presence of
a crystalline potential. As such, it can be used to determine the stability of var-
ious lattice structures. It can also be used to determine ground-state properties
or static properties of the electronic systems, such as those provided by elas-
tic scattering experiments. It is based on the Hohenberg and Kohn Theorem101 .

The Hohenberg-Kohn theorem considers the form of a many-electron Hamil-


tonian in which the form of the Coulomb interaction term between pairs of elec-
trons, Vint (r, r0 ), is known, but in which the one-particle potential due to the
ion cores is considered to be an external potential. Thus, the external potential
Vions (r) varies between one crystal structure and the next. The Hamiltonian
is written as the sum of the interaction energies of the pairs of electrons, the
kinetic energies of all the electrons and the external potentials acting on each
electron.

The Hohenberg-Kohn theorem starts by noting that every non-degenerate


many-body ground state wave function is associated with a unique external po-
tential. The theorem proves that a second map exists between the ground state
electron density and the external potential. This establishes a unique mapping
between the many-body ground-state wave function and the ground-state elec-
tron density. Therefore, the expectation value of any ground-state property can
be expressed in terms of a unique functional of the electron density. Having
established this, the ground state properties and electron density can then be
determined by using the Rayleigh-Ritz variational principle for the ground state
energy in which the electron density is the function to be varied.

This leads to a knowledge that, if one can construct the energy functional as
the sum of a unique energy functional and a simple part due to the external po-
tential due to the lattice, then one can find the ground state energy and electron
density. The unique energy functional is not known, however, it is customary
to make the Local Density Approximation (LDA). In this approximation, an
un-testable assumption is made about the electron-electron interactions in a
non-uniform electron gas. The method also generates eigenvalues which are in-
terpreted in terms of the energies of independent Bloch electrons. The energy
dispersion relations generated this way do show marked similarities with the
experimentally determined bands of simple metals.

101 P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).

344
The basis for density functional theory is provided by a theorem proved by
Hohenberg and Kohn102 .

9.3.1 Hohenberg-Kohn Theorem


The Hohenberg-Kohn theorem first assures us that the electron density in a
solid, ρ(r), uniquely specifies the electrostatic interaction potential between the
Ne electrons and the ionic lattice. Since there is a one to one mapping between
Vions (r) and the ground state many-body wave function Ψ, then the density
ρ(r) can be used as the basic variable. As a consequence of this theorem and
the Rayleigh-Ritz principle, the energy can be expressed as the sum of a unique
functional of the electron density and a simple functional describing the po-
tential energy of the electrons arising from their electrostatic interaction with
the ionic lattice. This establishes a variational principle which can be used to
calculate the electron density, ρ(r), and the total energy of the electronic system.

First, it shall be assumed that Vions (r) is not uniquely specified if ρ(r) is
given. That is, it is assumed that there exists at least two potentials V and V 0
which give rise to the same ground state electron density. These potentials are
related to the exact ground state many-particle wave functions via the energy
eigenvalue equations,

Ĥ Ψ(r1 , . . . rNe ) = E Ψ(r1 , . . . rNe ) (877)

and
Ĥ 0 Ψ0 (r1 , . . . rNe ) = E 0 Ψ0 (r1 , . . . rNe ) (878)
From the Rayleigh-Ritz variational principle, one finds that the primed wave
function Ψ0 (r1 , . . . rNe ) provides an upper bound to the ground state energy
of the unprimed Hamiltonian Ĥ,
i=N
Ye  Z 
E = d3 ri Ψ∗ (r1 , . . . rNe ) Ĥ Ψ(r1 , . . . rNe )
i=1
i=N
Ye  Z 
E < d3 ri Ψ0∗ (r1 , . . . rNe ) Ĥ Ψ0 (r1 . . . rNe ) (879)
i=1

However, as the primed and unprimed Hamiltonian are related through

Ĥ = Ĥ 0 + V̂ − V̂ 0 (880)

and as Ψ0 is the ground state of Ĥ 0 with energy eigenvalue E 0 , the energies


satisfy an inequality
Z
0
E < E + d3 r ρ(r) ( V (r) − V 0 (r) ) (881)

102 P. Hohenberg and W. Kohn, Phys. Rev. 136, B 864 (1964).

345
The assumption that the ground state densities of the primed and unprimed
Hamiltonian are equal has been used. By virtue of similar reasoning, it can also
be shown that the energies also satisfy the inequality
Z
0
E < E + d3 r ρ(r) ( V 0 (r) − V (r) ) (882)

where the prime and unprimed quantities are interchanged. Adding these two
inequalities leads to an inconsistency
E + E0 < E + E0 (883)
Therefore, the assumption that the same ground state density can be found for
two different potentials is false. Furthermore, the potentials can, at most, only
differ by a constant V 0 (r) − V (r). Thus, the ground state electron density ρ(r)
must correspond to a unique V (r). Since there is a unique one to one mapping
between Ψ and V (r), there is a unique one to one mapping between Ψ and ρ(r).
This means that the electron density, ρ(r), can be taken to be the principal
variable. It also implies that the Rayleigh-Ritz principle can be re-stated as a
variational principle for an energy functional E[ρ] which is expressed in terms
of the electron density.

9.3.2 Functionals and Functional Derivatives


As a mathematical prelude, we shall define functionals and functional deriva-
tives.

A functional is a generalization of a function. A function f (r) can be defined


as a mapping which maps each point in space, r, to a number. The value of
the number depends on the position of the point. The functional is similar in
that it maps a scalar function onto a number. The value of the functional, F [ρ],
depends upon the function ρ(r), i.e., the values of the function ρ at each point
in space. Functionals are usually expressed in terms of integrals over space,
usually as multiple integrals.

A simple example of a functional is given by the number of electrons Ne [ρ],


which is a functional of the density. The number of electrons is given by
Z
Ne [ρ] = d3 r ρ(r) (884)

It is a functional as different densities may correspond to different number of


particles i.e., N a has a different density than Li and they have different numbers
of electrons.

The classical Coulomb energy is a more interesting functional. The Coulomb


energy is defined as the pair-wise sum of interactions
e2 ρ(r) ρ(r0 )
Z Z
ECoul [ρ] = 3
d r d3 r 0 (885)
2 | r − r0 |

346
This yields a number which is the value of the energy, and this number depends
on the electron density at all points of space.

Given a functional F [ρ], one can define a functional derivative. The defini-
tion of the functional derivative is similar to the definition of a derivative of a
function. However, instead of defining the derivative in terms of the difference
of the function at two nearby points, one defines the functional derivative in
terms of the difference of the functional for two functions that are close. For
example, an arbitrary family of functions, ρ0 (r), can be defined in terms of a
fixed function ρ(r) and an arbitrary deviation δρ(r) via

ρ0 (r) = ρ(r) + λ δρ(r) (886)

The scale factor λ varies from unity to zero continuously. When λ = 1, this
relation defines the shape of the deviation δρ(r). If λ is changed continuously
to zero, the differences between the function ρ0 and the fixed function ρ van-
ish. The shape of the deviation λ δρ(r) is arbitrary and does not change, only
the magnitude of the deviation is changing. The functional derivative can be
expressed in terms of the limit of the difference of the functional evaluated for
these two functions. If one assumes that one may Taylor expand the functional
in powers of λ, one has
1 2 2
F [ρ0 ] = F [ρ] + λ δ 1 F [ρ, δρ] + λ δ F [ρ, δρ] + . . . (887)
2
since the differences now depend on two functions ρ and δρ. If one defines the
terms of first order in λ to have the form
Z
δF [ρ]
δ 1 F [ρ, δρ] = d3 r δρ(r) (888)
δρ(r)

then the quantity


δF [ρ]
(889)
δρ(r)
is independent of the shape of the deviation, and is defined to be the first order
functional derivative. Sometimes a functional may depend on the higher order
derivatives of ρ i.e., Z
F [ρ] = d3 r f (r, ρ, ∇ρ) (890)

In this case, one can define a functional derivative in terms of the partial deriva-
tives,
∂f
(891)
∂ρ
and the vector quantity
∂f
(892)
∂∇ρ

347
etc., where the functions ρ and ∇ρ etc. are treated as independent variables.
This yields the first order variation as
Z  
∂f ∂f
δ 1 F [ρ, δρ] = d3 r δρ + ∇δρ . (893)
∂ρ ∂∇ρ

If the functions ρ satisfy appropriate conditions at the boundaries of the region


of integration, the equation can be integrated by parts to eliminate the term
Z
∂f
d3 r ( ∇δρ ) . (894)
∂∇ρ
In this case, the first order functional derivative is evaluated as

δF [ρ] ∂f ∂f
= − ∇. (895)
δρ(r) ∂ρ ∂∇ρ

The extension to functionals containing higher order derivatives is quite straight-


forward.

An alternative method of evaluating functional derivatives is based on the


observation that the functional derivative is independent of the variation δρ.
Since δρ is arbitrary, one may choose δρ to have any particular form. The
particular variation of the form of a Dirac delta function proves to be a useful
choice
δρ(r) = δ 3 (r − r1 ) (896)
since, for this particular choice, the value of δF 1 [ρ, δ 3 (r − r1 )] is given by

δF [ρ]
δ 1 F [ρ, δ 3 (r − r1 )] = (897)
δρ(r1 )

An example of the first order functional derivative is given by the functional


derivative of the Coulomb energy

e2 e2 ρ(r0 )
Z Z
δECoul [ρ] ρ(r)
= 3
d r + d3 r 0
δρ(r1 ) 2 | r − r1 | 2 | r0 − r1 |
Z
ρ(r)
= e2 d3 r (898)
| r − r1 |

In obtaining the second line, we have relabelled the variable of integration. The
first order functional derivative of the mono-nomial functional
Z
Fn [ρ] = d3 r ρ(r)n (899)

is simply evaluated as
δFn [ρ]
= n ρ(r1 )n−1 (900)
δρ(r1 )

348
The delta function method also proves useful for evaluating functional deriva-
tives of higher orders.

The first order functional derivative is often encountered in variational prin-


ciples. In a variational principle, there exists a function ρ(r) which yields an
extremal value of the functional. That is, if the functional is changed by an
arbitrary small variation λδρ away from the extremal function, the functional
does not change. On regarding the functional F [ρ + λδρ] as a function of λ, the
extremal condition is equivalent to


F [ρ + λδρ] = 0 (901)
∂λ λ=0

since the value of the functional does not change to order λ, as λ approaches
zero. This equation is satisfied for an arbitrary shape δρ(r), if the functional
derivative is identically zero
δF [ρ]
= 0 (902)
δρ(r)
for all r. The extremal function ρ(r) must satisfy this extremal condition for all
r. Often, the extremal condition provides an integro-differential equation that
can be used to uniquely determine ρ(r). The above condition only guarantees
that for this particular ρ the functional F [ρ] is extremal. In order that the
functional F [ρ] is minimized, we require that

δ 2 F [ρ, δρ, δρ0 ] > 0 (903)

for every δρ.

The second order functional derivative is defined via


δ 2 F [ρ]
Z Z
0
2
δ F [ρ, δρ, δρ ] = 3
d r d3 r0 δρ(r) δρ0 (r0 ) (904)
δρ(r) δρ(r0 )
On using the choice
δρ(r) = δ 3 (r − r1 ) (905)
for the deviation centered at r in the first derivative and the choice

δρ0 (r0 ) = δ 3 (r0 − r2 ) (906)

when differentiating the second time, one obtains

δ 2 F [ρ]
δ 2 F [ρ, δρ, δρ0 ] = (907)
δρ(r1 ) δρ(r2 )
As an example, the second order functional derivative of the Coulomb energy is
found to be
δ 2 ECoul [ρ] e2
= (908)
δρ(r1 ) δρ(r2 ) | r1 − r2 |

349
A second example is provided by the functional derivative of the mono-nomial
Z
Fn [ρ] = d3 r ρ(r)n (909)

for real φ. For this functional, the second order functional derivative has the
form
δ 2 Fn [ρ]
= δ 3 (r1 − r2 ) n ( n − 1 ) ρ(r1 )n−2 (910)
δρ(r1 ) δρ(r2 )
etc.

9.3.3 The Variational Principle


Hohenberg and Kohn defined an energy functional of the electron density
Z
E[ρ] = F [ρ] + d3 r Vions (r) ρ(r) (911)

in which the energy functional F [ρ] depends on the kinetic energy T̂ given by
Ne
h̄2 X
T̂ = − ∇2 (912)
2 m i=1 i

and the electron-electron interaction energy, V̂int , given by

1 X e2
V̂int = (913)
2 | ri − rj |
i6=j

The functional F [ρ] can be evaluated as


i=N
Ye  Z   
F [ρ] = d3 ri Ψ∗ (r1 , . . . rNe ) T̂ + V̂int Ψ(r1 , . . . rNe )
i=1
(914)

The functional F [ρ] is a universal functional of ρ, as the functional F [Ψ] is a


universal functional of Ψ. Furthermore, as will be shown, the energy of the
electronic system E is given by the minimum value of the functional E[ρ] where
ρ(r) is the correct ground state density of an Ne electron system associated with
the lattice potential Vions (r). In fact, E is the minimum value of E[ρ] evaluated
for the set of functions, ρ(r), which correspond to the Ne -electron ground state
densities of arbitrary potentials. Such densities are known as V -representable
densities. Not all densities are V -representable.

350
Let ρ0 (r) 6= ρ(r) be an arbitrary density associated with some many-body
wave function Ψ0 6= Ψ, that is not the ground state of our system. The ground
state energy is defined by
i=N
Ye  Z 
E = d ri Ψ∗ (r1 , . . . rNe ) Ĥ Ψ(r1 , . . . rNe ) = E[ρ]
3

i=1
(915)

The Rayleigh-Ritz variational principle asserts that the expectation values of


the Hamiltonian satisfies the inequality
i=N
Ye  Z 
d3 ri Ψ∗ (r1 , . . . rNe ) Ĥ Ψ(r1 , . . . rNe )
i=1
i=N
Ye  Z 
≤ d3 ri Ψ0∗ (r1 , . . . rNe ) Ĥ Ψ0 (r1 , . . . rNe ) (916)
i=1

and so
E[ρ] ≤ E[ρ0 ] (917)
This establishes the minimum principle for the energy functional

δE[ρ]
= 0 (918)
δρ(r)

subject to the constraint that the total number of electrons are fixed
Z
d3 r ρ(r) = Ne (919)

The condition that ρ is V -representable may be replaced by a less stringent


condition of N representable103 , which only requires

ρ(r) > 0
Z
d3 r ρ(r) = Ne
Z
1
d3 r | ∇ ρ 2 (r) |2 < ∞ (920)

Having established the existence of the variational function, the precise form of
the functional E[ρ] remains to be determined.

103 M. Levy, Proc. Nat. Acad. Sci. 76, 6062 (1979).

351
9.3.4 The Electrostatic Terms
Hohenberg and Kohn suggest that one should separate the long-ranged classical
Coulomb energy of the electrons from the functional F [ρ]. This term represents
the average Coulomb interaction between the electrons in the system and, there-
fore, represents the Hartree terms. That is, the energy functional representing
the kinetic and electron-electron interaction energies is written as

e2 ρ(r) ρ(r0 )
Z Z
F [ρ] = 3
d r d3 r 0 + G[ρ] (921)
2 | r − r0 |

The total energy functional is given by

e2 ρ(r) ρ(r0 )
Z Z Z
E[ρ] = d3 r Vions (r) ρ(r) + d3 r d3 r 0 + G[ρ]
2 | r − r0 |
(922)

The electrostatic potential φes (r) is given by the sum of the potential due to
the lattice of ions and the electron-electron interaction
ρ(r0 )
Z
− | e | φes (r) = Vions (r) + e2 d3 r 0 (923)
| r − r0 |

This potential may be obtained directly from Poisson’s equation from the density
of the ions and electrons
 
− ∇2 φes (r) = 4 π | e | Z ρions (r) − ρ(r) (924)

where | e | is the magnitude of the charge on the electron. The electrostatic


potential determines the chemical potential through the variational procedure.
The energy functional is minimized w.r.t variations of ρ(r) subject to the con-
straint that the density is normalized to Ne . This is performed by using La-
grange’s method of undetermined multipliers. The method consists of construct-
ing the functional Ω[ρ] as
 Z 
Ω[ρ] = E[ρ] − µ d3 r ρ(r) − Ne (925)

Then on writing ρ0 (r) = ρ(r) + λ δρ(r) and Taylor expanding in λ, one has
Z
δΩ[ρ]
Ω[ρ0 ] = Ω[ρ] + λ d3 r δρ(r)
δρ(r)
λ2 δ 2 Ω[ρ]
Z Z
+ d3 r d3 r0 δρ(r) δρ(r0 ) + . . . (926)
2 δρ(r) δρ(r0 )

The extremal condition


δΩ[ρ]
= 0 (927)
δρ(r)

352
becomes
δE[ρ]
= µ (928)
δρ(r)
The first order functional derivative of E is evaluated from the Taylor expansion
by retaining the terms of first order in λ. The first order term in E, δE 1 , is
evaluated as
Z
δE 1 = d3 r δρ(r) Vions (r)

e2 ρ(r0 )
Z Z  
3 3 0 0 ρ(r)
+ d r d r δρ(r) + δρ(r )
2 | r − r0 | | r − r0 |
Z
δG[ρ]
+ d3 r δρ(r) (929)
δρ(r)

On interchanging the variables of integration r and r0 in the second part of the


Coulomb term and combining it with the first, one obtains

ρ(r0 )
Z  Z 
δG[ρ]
δE 1 = d3 r δρ(r) Vions (r) + e2 d3 r 0 +
| r − r0 | δρ(r)
(930)

Since the first two terms are identified with the electrostatic potential, the func-
tional derivative is given by

ρ(r0 )
Z
δE[ρ] δG[ρ]
= Vions (r) + e2 d3 r 0 0
+
δρ(r) |r − r | δρ(r)
δG[ρ]
= − | e | φes (r) + (931)
δρ(r)

Hence, Ω is minimized if ρ satisfies the equation

δG[ρ]
− | e | φes (r) + = µ (932)
δρ(r)

For large Ne , µ is equal to the chemical potential given by


∂E
µ = (933)
∂Ne

9.3.5 The Kohn-Sham Equations


The Kohn-Sham equations104 provide a formal correspondence between the
many-body problem and an effective (non-interacting) one-body problem. This
104 W. Kohn, L. J. Sham, Phys. Rev. 140, A 1133 (1965).

353
allows the kinetic energy term in the energy functional to be determined.

The kinetic energy functional T [ρ] can be defined via


i=N
Ye  Z 
T [ρ] = d3 ri Ψ∗ (r1 , . . . rNe ) T̂ Ψ(r1 , . . . rNe ) (934)
i=1

so the non-electrostatic contribution to the energy functional may be written as


the sum
G[ρ] = T [ρ] + Exc [ρ] (935)
which defines the exchange and correlation functional Exc [ρ]. The variational
principle for the density functional gives

δExc [ρ] δT [ρ]


− | e | φes (r) + + = µ (936)
δρ(r) δρ(r)

Thus, the quantity


δExc [ρ]
− | e | φes (r) + (937)
δρ(r)
plays the role of an effective potential, Vef f [ρ, r], which not only depends on r,
but is also a functional of ρ. The effective potential is given by

δExc [ρ]
Vef f [ρ, r] = − | e | φes (r) + (938)
δρ(r)

Thus, minimizing the energy functional entails solving the equation

δT [ρ]
Vef f [ρ, r] + = µ (939)
δρ(r)

Formally, this is equivalent to solving for the ground state of a (non-interacting)


problem with the energy functional given by
Z
Es [ρ] = T [ρ] + d3 r Vs (r) ρ(r) (940)

in which the electron-electron interaction terms are absent. The variational


procedure leads to
δT [ρ]
Vs (r) + = µ (941)
δρ(r)
Since the particles are non-interacting, this equation is solved by exactly finding
the single-particle wave functions φs,α which make up the single Slater determi-
nant that represents the non-interacting ground state. The set of single-particle
wave functions are given as the solutions of the eigenvalue equation

h̄2
 
2
− ∇ + Vs (r) φs,α (r) = Es,α φs,α (r) (942)
2m

354
and then the electron density is given by
i=N
Xe
ρ(r) = | φs,αi (r) |2 (943)
i=1

By analogy, one can find the solution of the effective one-body eigenvalue equa-
tion
h̄2
 
2
− ∇ + Vef f [ρ, r] φef f,α (r) = λef f,α φef f,α (r) (944)
2m
and the electron density is given by
i=N
Xe
ρ(r) = | φef f,αi (r) |2 (945)
i=1

The value of the kinetic energy functional for this effective one-body problem
can be found from the eigenvalues λef f,αi by
XNe Z
T [ρ] = λef f,αi − d3 r Vef f [ρ, r] ρ(r) (946)
i=1

Thus, one also has to minimize the sum of the effective one-body eigenvalues
Ne
X
λef f,αi (947)
i=1

This shows that the Kohn-Sham equations provide a method of obtaining the
kinetic energy functional and also minimizes the energy functional. Although
Kohn-Sham eigenvalues λef f,α are often used to describe electron excitation en-
ergies, they have no physical meaning. In general, the method only provides the
ground state energy and ground state electron density. However, there is a den-
sity functional analogue of Koopmans’ theorem: the eigenvalue of the highest
occupied effective single-particle level is the Fermi energy. All the non-trivial
information about the many-body ground state is contained in the exchange and
correlation function. This is usually approximated in an uncontrolled fashion
by using the local density functional approximation.

9.3.6 The Local Density Approximation


In the Kohn-Sham equations, the remaining unknown function is the exchange
and correlation functional Exc [ρ]. This contains the information about the
many-body interactions. The local density approximation is motivated by an
assumption namely, that this functional can be represented as an integral over
all space of a function of ρ. This assumes that the functional has no non-local
terms. It may be thought that the first correction terms to the local density
function could be expressed by expanding it in powers of the gradient105 . This
105 S. K. Ma and K. A. Brueckner, Phys. Rev. 165, 18 (1968), V. Sahni, J. Greunebaum

and J. P. Perdew, Phys. Rev. B 26, 4371 (1982).

355
type of expansion would be justifiable if the density ρ(r) was slowly varying in
space. The first few terms of the gradient expansion of the exchange-correlation
energy would be
Z  
3 2
Exc [ρ] = d r Exc0 (ρ(r)) + Exc2 (ρ(r)) | ∇ ρ(r) | + . . . (948)

where the coefficients Exc0 (ρ(r)) and Exc2 (ρ(r)) are ordinary functions of the
density. The local density approximation neglects the gradient terms and uses
the same form of the exchange-correlation function Exc0 (ρ(r)) as it pertains to
the free electron gas. In the free electron gas, the electron density ρ is inde-
pendent of r. However, in the local density approximation, the uniform density
appearing in the expression for Exc for the uniform electron gas is replaced by
a similar expression but which depend upon the local electron density.

The exchange and correlation terms from the local density approximation
are taken from the free electron gas. The energy of the free electron gas is
written as
3 h̄2 kF2 3 e2 m e4
 
1
E = Ne − kF − 2 ( 1 − ln 2 ) ln kF + O(1) (949)
5 2m 4 π π h̄2
where the first term is due to the kinetic energy, and the second term is the ex-
change energy. The final term is the leading term in the high density expansion
of the electron correlation energy, as evaluated by Gell-Mann and Brueckner106 .
For the free electron gas, the electrostatic interaction energy between the elec-
trons and the smeared out lattice of ions cancels identically with the Hartree
term. To obtain the exchange-correlation energy, the kinetic energy term is
omitted to find
3 e2 m e4
 
Exc = Ne − kF − 0.0311 ln kF + O(1) (950)
4 π h̄2
Combining this together with the two relations
  13
kF = 3 π2 ρ (951)

and
Ne = V ρ (952)
the exchange-correlation term can be expressed as
 13
3 e2 m e4
  
Exc = V ρ − 3 π2 ρ − 0.0104 ln ρ + O(1) (953)
4 π h̄2
106 M. Gell-Mann and K. A. Brueckner, Phys. Rev. 106, 364 (1957).

356
The exchange-correlation energy in the local density approximation is simply
given by
 13
3 e2 m e4
Z   
1
3 2
Exc [ρ] = d r ρ(r) − 3π ρ(r) − 0.0104 2 ln ρ(r) + O(1)
3
4 π h̄
(954)

Since the effective potential is given by the sum


δG[ρ]
Vef f [ρ, r] = − e φ(r) + (955)
δρ(r)
the local density approximation for the exchange and correlation energy func-
tional contributes a term to the potential in the Kohn-Sham equations of
 13
e2 m e4

2
Vxc [ρ] = − 3 π ρ(r) − 0.0104 ln ρ(r) + O(1) (956)
π h̄2
which adds to the electrostatic potential. The first term comes from the ex-
change interaction, and has the form that was originally proposed by J. C.
Slater but has a different coefficient107 . The higher order terms come from the
correlation energy. In practice, the form of the exchange-correlation energy that
is used as an input to the local density approximation is a form which interpo-
lates between the high-density limit and the low-density limit. One interpolation
formulae which has been used is
 13
3 e2 m e4
Z   
3 2 1 1
Exc [ρ] = d r ρ(r) − 3π ρ(r) −
3 ( 1 − ln 2 ) 2 ln ρ(r) + O(1)
4 π 6 π2 h̄
(957)

which is based on the work of Nozières and Pines108 . As the density is reduced,
the electrons are expected to undergo a phase transition and form a Wigner
crystal. Since the energy is expected to be a non-analytic function at the phase
transition, the interpolation is of doubtful utility. It seems more appropriate to
use the results of Monte Carlo calculations109 for the correlation energy of the
homogenous electron gas.

The local density functional approximation has been used to successfully de-
scribe many different materials, and fails miserably for some others. Attempts
to justify this expression based on the gradient expansion have failed. Basi-
cally, the electron density varies too rapidly for the gradient expansion to be
useful. However, despite the rapid varying charge densities, a generalized gra-
dient expansion110 , in which appropriately chosen cut-off’s are introduced, has
107 J.C. Slater, Phys. Rev. 81, 385 (1951).
108 P.Nozières and D. Pines, Phys. Rev. 111, 442 (1958).
109 D. M. Ceperley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980).
110 D. C. Langreth and M. J. Mehl, Phys. Rev. B 28, 1809 (1983), J. P. Perdew and Y.

Wang, Phys. Rev. B, 33, 8800 (1986).

357
Interpolated Total Energy
0.1

E(rs)/Ne [ Rydbergs ]
0

-0.1

-0.2

-0.3
0 2 4 6 8

rs

Figure 188: The density dependence of the total energy for a uniform electron
gas, as suggested by Nozières and Pines (1958).

produced some encouraging results.

358
9.4 Static Screening
The response of an electronic system to a static or time-independent external
potential is quite remarkable in a metal. In a metal, the long-ranged part of the
static external potential is completely screened out by the electron response.
The screening is characterized by the dielectric constant. Classically, the total
electrostatic potential φes (r) is related to the charge density through Poisson’s
equation. In the absence of the external potential, Poisson’s equation is written
as  
2
− ∇ φes (r) = 4 π | e | Z ρions (r) − ρ(r) (958)

For a free electron gas, the charge density for the electrons exactly cancels the
contributions from the smeared out charges of the ions. The corresponding
potential is constant, and the reference value φes (r) may be set to be zero.
It is expected that a positive external charge with density ρext (r) will induce
a change in the electronic density ρind (r). The external charge produces an
external potential which is defined by the Poisson equation

− ∇2 φext (r) = 4 π | e | ρext (r) (959)

The total potential φes (r) satisfies Poisson’s equation


 
− ∇2 φes (r) = 4 π | e | ρext (r) − ρind (r) (960)

where ρext is assumed to be a positive charge, and the induced electron density
ρind is associated with the negatively charged electrons. The external potential
is related to the total potential via the dielectric constant through the non-local
relation Z
φext (r) = d3 r0 ε(r, r0 ) φes (r0 ) (961)
V
In a spatially homogeneous system, the dielectric constant is translationally
invariant and, therefore, only depends upon the difference r − r0 . In this case,
the linear response relation is expressed as a convolution
Z
φext (r) = d3 r0 ε(r − r0 ) φes (r0 ) (962)
V

This non-local relation, which is valid for homogeneous systems, is simpler after
it has been Fourier transformed. The Fourier transform of φext (r) is defined by
Z  
1 3
φext (q) = d r φext (r) exp − i q . r (963)
V V

and the Fourier transform of the dielectric constant is defined by


Z  
ε(q) = d3 r ε(r) exp − i q . r (964)
V

359
Hence, the Fourier transform of the convolution is just the product of the re-
spective Fourier transforms. Thus, the relation becomes

φext (q) = ε(q) φes (q) (965)

Hence, the total potential is reduced by the dielectric constant


φext (q)
φes (q) = (966)
ε(q)

This relation is sometimes used as the definition of the dielectric constant.

The Fourier Transforms of Poisson’s equations, given in eqn.(959) and eqn.(960),


yield
q 2 φext (q) = 4 π | e | ρext (q) (967)
and  
2
q φes (q) = 4 π | e | ρext (q) − ρind (q) (968)

On using the first equation to eliminate ρext (q) in the second, one obtains

q 2 φes (q) = q 2 φext (q) − 4 π | e | ρind (q) (969)

Taking the induced charge density term to the other side of the equation pro-
duces
q 2 φext (q) = q 2 φes (q) + 4 π | e | ρind (q) (970)
The definition of the dielectric constant, ε(q), can be used to eliminate φext (q),
thereby yielding the relation

4 π | e | ρind (q)
ε(q) = 1 + (971)
q2 φes (q)

The total scalar potential can be expressed as a potential energy term δV (q)
acting on the electrons
δV (q) = − | e | φes (q) (972)
The response function χ(q) is defined as the ratio of the induced density to the
potential, δV (q), which produces it

ρind (q)
χ(q) = (973)
δV (q)

Therefore, one finds that the dielectric constant is related to the response func-
tion via
4 π | e |2
ε(q) = 1 − χ(q) (974)
q2
Thus, the dielectric constant is related to the response of the charge density
to the total potential. This response function can be calculated via different

360
techniques. However, in making approximations, it is imperative that only the
response to the total potential is approximated and not the response to the
external potential. In a metal, it is all the electrons that take part in screening
an external charge. If each electron were to react independently to screen the
external charge, the external charge density would be over-screened by a factor
of Ne as each electron by itself could neutralize a charge of | e |. The simplest
approximate theory of the system’s response to the total field is given by the
Thomas-Fermi approximation. The Thomas-Fermi theory pre-dates linear re-
sponse theory and density functional theory. A more accurate approximation
for weak potentials is based on linear response theory.

The above derivation has the following drawbacks: First, the use of Poisson’s
equation only treats the classical direct Coulomb interactions between aggre-
gates of electrons, neglecting the effect of the exchange interactions. Second,
the assumption of spatial homogeneity neglects the effect of Umklapp interac-
tions in a solid. The neglect of Umklapp interactions produced simple algebraic
equations relating φ(q) and ρind (q). The inclusion of Umklapp scattering pro-
duces an infinite set of coupled equations which has no known analytic solution.

9.4.1 The Thomas-Fermi Approximation


The Thomas-Fermi approximation111 is based on the assumption that the po-
tential is slowly varying. Hence, the energy of a Bloch state is given by

h̄2 k 2
− | e | φes (r) (975)
2m
The momentum of the highest occupied energy is r dependent and is defined to
be kF (r). Then kF (r) is given by

h̄2 kF2 (r)


− | e | φes (r) = µ (976)
2m
The above equation shows that kF (r) is slowly varying if φes (r) is slowly vary-
ing. Since by assumption kF (r) is slowly varying, the electron density ρ(r) at
position r can be expressed in terms of the local Fermi wave vector

1 4 π kF3 (r)
ρ(r) = 2 (977)
8 π3 3
On expressing the Fermi wave vector in terms of the chemical potential and the
electrostatic potential, the total density becomes
  32   32
1 2m
ρ(r) = µ + | e | φes (r) (978)
3 π2 h̄2
111 L. H. Thomas, Proc. Cambridge Philos. Soc. 23, 542 (1927), E. Fermi, Zeit. für Physik,

48, 73 (1928).

361
E
2 2
µ = h kF(r) /2m − e φes(r)

E0 - e φes(r)

0 r

Figure 189: The schematic spatial variation of the one-electron energies, as


envisaged in the Thomas-Fermi approximation.

The induced density is then given in terms of the electrostatic potential via
 3 "   32   32 #
1 2m 2
ρind (r) = µ + | e | φes (r) − µ (979)
3 π2 h̄2
The above equation forms the basis of the Thomas-Fermi Theory. On assuming
that φes (r) is small compared with µ, the equation can be linearized to yield
 
∂ρ0
ρind (r) = | e | φes (r) (980)
∂µ
Thus, the Thomas-Fermi response function is given by
 
∂ρ0
χT F = −
∂µ
 3
1 2m 2 1
= − µ2
2 π2 h̄2
m kF
= − 2 2 (981)
π h̄
This leads to the Thomas-Fermi approximation for the dielectric constant
4 π e2
 
∂ρ0
ε(q) = 1 +
q2 ∂µ
2
k
= 1 + T2F (982)
q
The Thomas-Fermi wave vector is given in terms of the Fermi wave vector by
4 m e2
kT2 F = kF
π h̄2
4 kF
= (983)
π a0

362
and by the alternate expressions
r
kT F 4
=
kF π kF a0
  13
16 1
= 2
rs2

1
= 0.8145 rs2 (984)

Thus, kT F is of the order of kF in a metal, and depends on the density of


mobile electrons available to perform screening. This means that the external
potential or charge is screened over distances of the order of kT−1F ∼ 1 Angstrom.

The spatial dependence of the screened potential can be most clearly seen
through the application of the Thomas-Fermi approximation to the screening
of a point charge Z e in a metal. The charged particle is located at the origin.
From the Fourier transform of Poisson’s equation, the external potential is given
by
4πZ |e|
φext (q) = (985)
q2
The total potential is given by

φext (q)
φes (q) =
ε(q)
4πZ |e|
= (986)
q 2 + kT2 F

which no longer shows the divergence associated with a long-ranged interaction


when q → 0. On performing the inverse Fourier transform, thereby transform-
ing the potential back into direct space, one has
 
Z |e|
φes (r) = exp − kT F r (987)
r

Thus, the charged impurity is exponentially screened over a distance kT−1F . The
induced charge density is given by

Z e2
   
∂ρ0
ρind (r) = exp − kT F r
r ∂µ
 2   
Z kT F
= exp − kT F r (988)
r 4π

On integrating this over all space, one finds that the screening in a metal is
perfect in that the total number of electrons in the induced density is equal to
Z.

363
φes(r)
Z e/r

Z e/r exp[-kTFr]
0 r
kTF-1

Figure 190: A comparison of the spatial dependence of the Thomas-Fermi


screened and the unscreened electrostatic potential.

The Thomas-Fermi approximation is deficient. It over-screens negative point


charges in the free electron gas. That is, the Thomas-Fermi approximation
produces an unphysical negative total electron density at the position of the
point charge. Likewise, for isolated atoms, it can be shown that the Thomas-
Fermi approximation breaks down as it predicts that the electron density at
the nuclear position is infinite (L. D. Landau and E. M. Lifshitz, Quantum
Mechanics), i.e.,
3
lim ρ(r) ∼ r− 2 (989)
r → 0

The Thomas-Fermi approximation cannot describe negative ions. That is, in


the Thomas-Fermi approximation, the number of electrons must always be less
than the nuclear charge. Furthermore, the Thomas-Fermi method also precludes
the binding of neutral atoms into molecules112 . The Thomas-Fermi method is
deficient as it assumes that the potential is slowly varying in space compared
to the distance over which the electrons adjust to the potential. Therefore, the
Thomas-Fermi method assumes that a local approximation for the kinetic en-
ergy is valid. This is not the case for most simple metals, where the potential
due to the ions varies over distances of the order of Angstroms. However, the
Thomas-Fermi theory is valid113 in the limit Z → ∞, and provides a useful
description of the bulk of the atom for finite Z.

9.4.2 Linear Response Theory


Linear response theory describes the response of a system to a weak perturbing
potential. In such cases, the response is assumed to be approximately linear
in the perturbation, so first-order perturbation theory may be used. The effect
112 E. Teller, Rev. Mod. Phys. 34, 627 (1962), see also E. H. Lieb, Rev. Mod. Phys. 48,
553 (1976).
113 E. H. Lieb and B. Simon, J. Chem. Phys. 61, 735 (1977).

364
1.5

ρ(r)/ρ
0.5

-0.5
0
r/a0

Figure 191: The spatial dependence of the total electron density near a negative
point charge, as calculated in the Thomas-Fermi approximation. The electron
density is unphysical near the origin, since it becomes negative.

of a perturbing potential δV (r) on the electronic system is considered. The


effect of this one-body potential on the single-particle Bloch functions φn,k (r),
is examined via perturbation theory. To first order in the perturbation, the
one-electron eigenfunctions are altered. The one-electron eigenfunctions are no
longer Bloch functions, but are approximated by the expressions
X Mn0 ,k0 ;n,k
ψn,k (r) = φn,k (r) + φn0 ,k0 (r) (990)
En,k − En0 ,k0
n0 ,k0 6=n,k

which include the first-order corrections from the perturbation series expan-
sion. In the above equation, Mn0 ,k0 ;n,k is the matrix element of the perturbing
potential between two Bloch functions,
Z
Mn0 ,k0 ;n,k = d3 r0 φ∗n0 ,k0 (r0 ) δV (r0 ) φn,k (r0 ) (991)

The induced change in the electron density, to first order in δV (r), is found as
" #
X X
∗ Mn0 ,k0 ;n,k
ρind (r) = φn,k (r) φn0 ,k0 (r) + c.c. (992)
0 0
En,k − En0 ,k0
n,k,σ n ,k 6=n,k

where the summation over n, k runs over all the occupied states and c.c. denotes
the complex conjugated term. Thus, the induced charge density at position r
is not just related locally to the perturbation at the same point, but instead is
related to δV (r0 ) applied at all the points r0 . The induced charge density ρind (r)
is expressed as the response to the perturbation δV (r0 ) through the non-local
relation Z
ρind (r) = d3 r0 χ(r, r0 ) δV (r0 ) (993)

365
where the response function χ(r, r0 ) is given by the expression

φ∗n0 ,k0 (r0 ) φn0 ,k0 (r)


" #
X X
0 ∗ 0
χ(r, r ) = φn,k (r) φn,k (r ) + c.c. (994)
0 0
En,k − En0 ,k0
n,k,σ n ,k 6=n,k

In the above expression for χ(r, r0 ), the summation over n, k, σ runs over all
one-electron states that were occupied before the perturbation was turned on.
Due to the Pauli exclusion principle, the summation over n0 , k 0 is restricted to
the unoccupied states. The expression for the response is expected to be modi-
fied by the presence of electron-electron interactions.

The expression for the non-interacting response can easily be evaluated for
free electrons. First, the variables k 0 and k are interchanged in the complex
conjugate term, and then, due to a cancellation between the two terms, the
range of one integration in each term is extended over all momentum space.
Once again, the variables k 0 and k are interchanged in the second term, to yield
"
d3 k d3 k 0
Z Z  
0 4m 0 0
χ(r, r ) = exp i ( k − k ) . ( r − r )
h̄2 |k|≤kF ( 2 π )3 ( 2 π )3
#  
1
+ c.c.
k 2 − k 02
(995)

where the integration over k 0 now runs over all space. As the Hamiltonian
possesses translational invariance, the response function only depends on the
vector R = r − r0 . Thus, for the homogeneous electron gas, the real space
linear response relation is in the form of a convolution. The integrations over the
directions of k and k 0 can be evaluated by standard means. The integration can
be transformed so that the integration over the magnitude of k 0 extends over
the range (∞, −∞). The resulting integral is evaluated by means of contour
integration, and yields
kF
sin 2 k | r − r0 |
Z
0 2m 2
χ(r, r ) = − 2 dk k
h̄ ( 2 π )3 0 | r − r0 |2
(996)

The resulting expression is

cos 2 kF | r − r0 | sin 2 kF | r − r0 |
 
2m 1 4
χ(r, r0 ) = k F −
h̄2 π 3 ( 2 kF | r − r0 | )3 ( 2 kF | r − r0 | )4
(997)

This is the response to a delta function perturbation at the origin. This delta
function perturbation requires the electron gas to adjust at very short wave

366
lengths. Instead of having the exponential decay as predicted by the Thomas-
Fermi approximation, the response only decays algebraically, with characteristic
oscillations determined by the wave vector 2 kF due to the sharp cut off at the
Fermi surface. That is, 2 kF is the largest wave vector available for a zero-energy
density fluctuation in which an electron is excited from just below to just above
the Fermi surface. The oscillations in the density that occur in response to a
potential are known as Friedel oscillations114 .

It is more convenient to consider the Fourier transform of the response func-


tion Z  
χ(q) = d3 r χ(r) exp − iq.r (998)
V

The response function χ(q) is evaluated from


 
2m 1 X 1 1
χ(q) = 2 +
h̄2 V k<k k 2 − ( k + q )2 k2 − ( k − q )2
F

(999)

where the summation over k runs over the occupied states within the Fermi
sphere. The summation can be replaced by an integration
Z kF Z +1 
2m 1 2 1
χ(q) = − 2 2 2
dk k d cos θ 2
h̄ 4 π 0 −1 q + 2 k q cos θ

1
+ 2
q − 2 k q cos θ
Z kF
2m 1 |q + 2k|
= −2 2 2
dk k ln (1000)
h̄ q 4 π 0 |q − 2k|

The response is given explicitly by

4 kF2 − q 2
  
m kF 1 | 2 kF + q |
χ(q) = − + ln (1001)
h̄2 π 2 2 8 q kF | 2 kF − q |

This is the Lindhard function115 for the free electron gas. The Lindhard function
reduces to the value of the corresponding Thomas-Fermi response function at
q = 0, which is
k2
χT F = − T F 2 (1002)
4πe
Thus, for very slowly varying potentials, the response of the free electron gas
is identical to the response function found using the Thomas-Fermi approxima-
tion. The magnitude of the Lindhard function drops with increasing q, falling
to half the q = 0 value at q = 2 kF . At this point, the slope has a weak
114 J. Friedel, Phil. Mag. 43, 153 (1952), Nuovo Cimento, Suppl. 7, 287 (1958).
115 J. Lindhard, Kgl. Danske Videnskab. Selskab. Mat.-Fys. Medd. 28, 8 (1954).

367
1.2

0.8

χTF
χ(q)/χ
0.6

0.4

0.2

0
0 0.5 1 1.5 2
q/2kF

Figure 192: The dependence of the Lindhard function on q/2kF .

logarithmic singularity. The electron gas is ineffective in screening the applied


potential for q ≥ 2 kF as 2 kF corresponds to the largest wave vector at which
electrons on the spherical Fermi surface can readjust.

The electron density induced by a point external charge of strength Z can


be written as
 
(q) − 1
Z
ρind (r) = Z d3 q exp[ i q . r ] (1003)
(q)

where we have used the relation


4 π e2
(q) = 1 − χ(q) (1004)
q2
to simplify the numerator. For short distances, the induced electron density
shows an exponential decay similar to that found in the Thomas-Fermi approx-
imation. For large distances, the electron density exhibits Friedel oscillations
and varies as
4 Z kF3 kT2 F kF2 cos 2kF r
ρind (r) ∼ (1005)
π ( kT2 F + 2 ( 2 kF )2 )2 kF3 r3

This expression was first derived by Langer and Vosko116 .

9.4.3 Density Functional Response Function


The change in the electron density ρind (r) due to an external potential, φext (r),
in which electron-electron interactions are included can be obtained from density
116 J. S. Langer and S. H. Vosko, J. Phys. Chem. Solids, 12, 196 (1960).

368
0.2

ρind(r)/(ZkF )
3 0.15

0.1
rs=4

0.05

rs=3
0
0 1 2 3 4

kF r

Figure 193: The short-distance variation of the electron density induced to


screen a point charge of strength Z. The values of rs = 4 and rs = 3 have been
used in these calculations. [After Langer and Vosko (1960).]

0.004

rs=3
rs=4
ρind(r)/(ZkF )

0.002
3

0
2 3 4 5 6 7 8 9 10

-0.002

kF r

Figure 194: The Friedel oscillations in the induced electron density around a
point charge of strength Z. [After Langer and Vosko (1960).]

369
functional theory. The relation between the induced density and the external
potential is given in terms of the screened response function by
Z
ρind (r) = − d3 r0 χs (r − r0 ) | e | φext (r0 ) (1006)

Density functional theory yields an effective potential which contains the effect
of the electron-electron interactions
" #
| e |2 δ 2 Exc
Z
3 0 0
| e | φef f (r) = | e | φext (r) − d r ρind (r ) +
| r − r0 | δρ(r) δρ(r0 )
(1007)
The relation between the induced electron density and the effective potential is
given by Z
ρind (r) = − d3 r0 χ0 (r − r0 ) | e | φef f (r0 ) (1008)

where χ0 (r − r0 ) is the Lindhard response function for non-interacting electrons.

The response function, including the effects of the electron-electron interac-


tions, can be found by Fourier transforming the above set of equations. Thus,
the full response function is given by

ρind (q) = − χs (q) | e | φext (q) (1009)

and the non-interacting response function is given by

ρind (q) = − χ0 (q) | e | φef f (q) (1010)

The relationship between the effective and external potential is given by


 
4π|e| π|e|
φef f (q) = φext (q) + − + Γxc (q) ρind (q) (1011)
q2 kT2 F
This equation can be solved for χs (q) in terms of the non-interacting response
function χ0 (q).

χ0 (q)
χs (q) =   (1012)
4 π π
1 − |e |2 q2 − 2
kT
Γxc (q) χ0 (q)
F

The dielectric constant ε(q) is given by

1 φes (q)
=
ε(q) φext (q)
1 4 π | e | ρind (q)
= 1 −
ε(q) q2 φext (q)
1 4 π e2
= 1 + χs (q) (1013)
ε(q) q2

370
The exchange contribution to Γxc (q) is given in the limit q → 0 by
  2  4 
5 q 73 q
Γxc (q) = 1 + + + ... (1014)
9 2 kF 225 2 kF

It is noted that if the effect of the exchange-correlation terms to the screening


could be dropped, then the dielectric constant is approximated by

1 4 π e2 χ0 (q)
≈ 1 + (1015)
ε(q) q2 ε(q)

which is consistent with the result for free-electrons using the Lindhard approx-
imation for the response to the total field, and treating the total scalar potential
classically via Poisson’s equation. In obtaining this approximate result, it was
necessary to calculate the response of the system to the external potential by
including processes, to all orders in e2 , in which the electron gas is polarized.
That is, the electron gas is polarized by the external potential and then the re-
sulting polarization and the external potential are screened by the electron gas,
ad infinitum. This infinite regression is necessary for the external charge to be
completely screened at large distances, and is a consequence of the long-ranged
2
nature of the Coulomb interaction limq → 0 4 πq2 e → ∞. This re-emphasizes
the importance of only making approximations in the response to the total po-
tential χ and not in the response to the external potential χs .

The response of the electronic system to an applied potential can be used


to examine the stability of a structure. The electronic energy change due to
the perturbation consists of the potential energy of interaction between the
ions and the electron gas, as well as the change induced into the energy of
electron-electron repulsions. All of these energies can be expressed in terms of
the induced charge density.

——————————————————————————————————

9.4.4 Exercise 57
Calculate the Lindhard function for a free electron gas with the dispersion re-
(0) h̄2 k2
lation Ek = 2 m in d = 1, d = 2 and d = 3 dimensions, at zero
temperature.

——————————————————————————————————

371
9.4.5 Exercise 58
Consider the Lindhard function for a tight-binding non-degenerate s band on a
hyper-cubic lattice with the dispersion relation
i=d
X
Ek = E0 − 2 t cos ki a (1016)
i=1

Show that the response function at the corner of the Brillouin zone q =
π
a (1, 1, 1, ., ., .) diverges as the number of electrons in the band approaches one
per site.

——————————————————————————————————

Worked Example.

We have seen that, in the Hartree-Fock approximation, the density of states


of the uniform electron gas is pathological. We have asserted that these patholo-
gies are related to the neglect of screening. In fact, we have shown that simple
perturbation theory for the unform electron gas is unworkable due to divergences
caused by the long-ranged coulomb interaction. Here we shall show that, if it
is assumed that the electrons interact via a Thomas-Fermi screened coulomb
interaction, the Hartree-Fock approximation is no longer pathological.

The Hartree-Fock quasi-particle energy for a screened coulomb gas is given


by
h̄2 k 2
λk = + Vexch (k) (1017)
2m
where the exchange energy is related to the screened coulomb interaction
XZ e2
Vexch (k) = − d3 r 0 exp[ − kT F | r − r0 | ] ψk∗0 (r0 ) ψk0 (r) exp[ i k . ( r0 − r ) ]
0
| r − r0 |
k
(1018)
where the states k 0 are occupied by electrons. Since the single-particle wave
functions ψk0 (r) are of the form of plane waves, the exchange interaction has
the form of a Fourier transform. The Fourier transform is evaluated as
e2
Z Z
1 3 0
Vexch (k) = − d k d3 r 0
(2π) 3
k0 <kF | r − r0 |
× exp[ − kT F | r − r0 | ] exp[ i ( k − k 0 ) . ( r0 − r ) ]
4 π e2
Z
1
= − d3 k 0
(2π) 3
k0 <kF ( k − k 0 )2 + kT2 F
e2 ( k − k 0 )2 + kT2 F
Z
= dk 0 k 0 ln (1019)
2 π k k0 <kF ( k + k 0 )2 + kT2 F

372
0

-0.1

Vexch(k)
-0.2

-0.3
0 1 2 3 4

k/kF

Figure 195: The exchange energy of the screened electron gas as a function of
k/kF .

The result is
e2 kF ( kT2 F + kF2 − k 2 ) (k + kF )2 + kT2 F
  
Vexch (k) = − ln
2π 2 k kF (k − kF )2 + kT2 F
  
kT F (k − kF ) (k + kF )
+2 tan−1 − tan−1 + 2
kF kT F kT F

This expression is a negative function which monotonically increases towards


zero as k increases. Furthermore, the derivative of the Vexch (k) and the deriva-
tive of the quasi-particle energy λk do not diverge logarithmically as k → kF .
Hence, in the screened Hartree-Fock approximation, the quasi-particle density
of states does not exhibit pathological behavior.

——————————————————————————————————

Worked Example: The Spiral Spin Density Wave State

We shall examine a broken-symmetry Hartree-Fock solution for the screened


free electron gas. We shall replace the Coulomb interaction by a highly-screened
Thomas-Fermi interaction of arbitrary strength

4 π e2 4 π e2
   
→ (1020)
| k − k 0 |2 kT2 F

which is separable. For a separable interaction, the band shifts ∆0 (k), ∆± (k)
and the gap become independent of k. The relative shift ∆± is found to be zero
and the overall shift ∆0 can be absorbed in the definition of the zero of energy.
For a fixed value of the gap, the conservation of electron number is ensured by

373
determining µ from the equation
 Z kzmax 
3 X ±
kF3 = dkz k⊥ (kz )2 (1021)
4 ± kzmin

where
s 2
h̄2 ± h̄2 2 h̄2
k⊥ (kz )2 = µ − k ± kz2 Q2 + | ∆ |2 (1022)
2m 2m z 2m

and the maximum and minimum values of kz are determined from the solutions
±
of k⊥ (kz ) = 0 as either
s
2 2 2
h̄ h̄ Q h̄2 Q2
(kz )2 = µ + ± µ + | ∆ |2 (1023)
2m 8m 2m

or zero, depending on the value of µ [see fig(186)]. The integrals can be easily
performed and expressed in terms of elementary functions. Likewise, the gap
equation either has a trivial solution

|∆| = 0 (1024)

or it has a non-trivial solution found from the equation


Z kzmax ±
4 π e2 (kz )2
   X  
1 k⊥
1 = ± dk z
kT2 F 8 π2
q
2
± kzmin ( 2h̄m )2 Q2 kz2 + | ∆ |2
(1025)
which can also be expressed in terms of the same elementary functions. The
right-hand side of the above equation is seen to be a decreasing function of
2
| ∆ |2 , therefore, one expects that the Coulomb interaction 4kπ2 e must exceed
TF
a critical value if there is to be a non-trivial solution (| ∆ | = 6 0) of the gap
equation. It is instructive to examine the ∆ → 0 limit of the above equation
which reduces to
m e2 kF 4 kF2 − Q2
     
Q + 2 kF
1 = 1 + ln
(1026)
π h̄2 kT2 F 4 Q kF Q − 2 kF

This expression confirms that there exists a minimum value of the Coulomb
interaction which has to be exceeded if the system is to have a finite amplitude
of the spiral magnetization. The above expression also shows that the para-
magnetic free electron gas has a preference to be unstable with respect to a
ferromagnetic state since the critical value of the interaction required for the
system to respond with a magnetic spiral of wave vector Q is minimum when
Q = 0. This conclusion is expected to change when the ionic potential is taken
into account or if the material is highly anisotropic.

374
3

2 d=1

χ(q)/χ(0)
d=2
d=3

0
0 1 2 3 4

q/kF

Figure 196: The normalized response function for a free electron gas as a func-
tion of momentum transfer q, for different dimensionalities.

For non-zero values of ∆, the determination of the optimal magnitude of Q


is more intricate. One way to determine the relative stability of the symmetry-
broken states is to compare the values of their total energies E(Q) (per unit
volume). The total energy can be expressed as
 2 
X kT F
E(Q) = λk,± fk,± + | ∆ |2 (1027)
4 π e2
k,±

From this expression, it can be shown that the value of | ∆ | which minimizes
E(Q) corresponds to the solution of the gap equation. It can also be seen that
whenever the gap equation has a non-trivial solution, the state with broken sym-
metry has lower energy than the paramagnetic metal. Furthermore, it can be
seen that the condensation energy is proportional to | ∆ |2 . For computational
purposes, it is more convenient to evaluate the total energy E(Q) from
X Z µ  2 
kT F
E(Q) = dE E ρ±qp (E) + 2
| ∆ |2 (1028)
± −∞ 4 π e

in which the ρ±qp (E) are the quasi-particle density of states (per unit volume) for
each sub-band. It should be noted that the quasi-particle density of states, just
like the values of ∆ and µ, implicity depend on Q. The quasi-particle density
of states (per unit volume) for each sub-band is evaluated as
 
± 2m max min
ρqp (E) = kz (E) − kz (E) (1029)
4 π 2 h̄2
where the non-zero maximum and minimum values of kz for fixed E are found
from
s
h̄2 h̄ 2
Q2
h̄2 Q2
kz (E)2 = E + ± E + | ∆ |2 (1030)
2m 8m 2m

375
The form of the total quasi-particle density of states is sketched in fig(197). The

0.15

ρqp(E) [ arbitrary units ] 0.1

2∆
0.05

0
0 0.5 1 1.5
8mE/h2Q2
Figure 197: The total quasi-particle density of states for a spiral spin density
wave state.

quasi-particle density of states is extremely close to the result for a paramagnetic


metal since the transition only alters states in a small energy interval with a
width of the order of ∆.

376
10 Stability of Structures
In this chapter, the structural stability of a metal is discussed. The total energy
of the metal will be expressed in terms of the energy for a uniform electron gas,
and the interaction with the periodic structure will be treated as a perturbation.

10.1 Momentum Space Representation


In the uniform electron gas, the electro-static energy between pairs of electrons
and also between the particles forming the background positive charge exactly
cancels with the interaction between the electrons and the positive charges.
When the periodic potential is introduced as a perturbation, the change in the
total energy can be expressed in terms of the change in the one-electron eigen-
values. However, the inclusion of the Coulomb interaction between the lattice
and the electrons will also require that the contributions from electron-electron
and ion-ion interaction be explicitly reconsidered in the calculation of the total
energy.

The energy of a one-electron Bloch state, calculated to second in the poten-


tial due to the ionic lattice, can be expressed in terms of the one-electron energy
eigenvalues for a free electron gas as

h̄2 k 2 2 m X | Vions (k 0 , k) |2
En,k = + Vions (k, k) + (1031)
2m h̄2 k0 6=k k 2 − k 02

The zero-th order and first order terms in this energy are independent of the
lattice structure of the ionic potential. This can be seen by examining the matrix
elements
Z  
0 1 3 0
Vions (k , k) = d r Vions (r) exp i ( k − k ) . r (1032)
V

which, when k = k 0 , is just the average potential. The sum of the energies
of all the occupied Bloch states, (k, σ), contributes to the total energy of the
solid. The first order contribution from Vions (k, k), like the kinetic energy of
the free electron gas, does not depend on the structure. These terms combine
to produce a volume-dependent contribution to the solid’s total energy.

The other volume-dependent contribution to the total energy of the solid


originates from the electron-electron interactions and the ion-ion interactions.
It is convenient to combine these terms with the energy of the zero-th order
electron-ion interaction, due to the exact cancellation for the uniform electron
gas. This combination is the total electrostatic interaction energy. It can be
evaluated in the approximation that the Coulomb interactions between different

377
Wigner-Seitz cells are totally screened117 . This means that the ion-ion inter-
actions need not be considered explicitly. The electrostatic contribution to the
energy is then written as

e2 ρ(r) ρ(r0 )
Z Z Z
Ees = d3 r Vions (r) ρ(r) + d3 r d3 r 0 (1033)
2 | r − r0 |

To lowest order in the structure, the electrostatic contribution to the total


energy can be evaluated by considering the Wigner-Seitz unit cell to be spherical
with radius RW S . The electron density is given in terms of the Wigner-Seitz
radius118 by
3Z
ρ = 3 (1034)
4 π RW S
In the case of a uniform electron density, the electron-electron repulsion term is
evaluated as
3 Z 2 e2
Z
Ees = d3 r Vions (r) ρ(r) + (1035)
5 RW S
For the free-electron approximation for the kinetic energy to be reasonable, the
electrostatic contribution from the ions should be calculated using the pseudo-
potential. We shall use the Ashcroft empty core approximation for the ionic
pseudo-potential. Inside the Wigner-Seitz cell, the pseudo-potential reduces to
that of an isolated atom
Z e2
Vatom (r) = − for r ≥ Rc
r
= 0 for r ≤ Rc (1036)

where Rc is the radius of the ionic core. Hence, for a structureless metal, the
electrostatic terms can be expressed as
" 2 #
3 Z 2 e2 3 Z 2 e2

Rc
Ees = − 1 − + (1037)
2 RW S RW S 5 RW S

The potential terms inversely proportional to the Wigner-Seitz radius can be


9 Z 2 e2 9
combined as − 10 RW S . The coefficient α = 10 is the Madelung constant for
a solid composed of spherical unit cells. In general, the Madelung constant will
depend slightly on the structure of the lattice.

117 E. Wigner and F. Seitz, Phys. Rev. 43, 804 (1933), E. Wigner and F. Seitz, Phys. Rev.

45, 509 (1934).


118 The Wigner-Seitz cell radius is related to the electron density parameter r via
s
1
RW S = Z 3 r s a 0
.

378
For a solid with structure, the electrostatic energy can be expressed as the
sum
E = EM + Ec (1038)
where EM is the Madelung energy and Ec is the core energy. The Madelung
energy is the electrostatic energy due to point charges immersed in a neutralizing
uniform distribution of electrons. The Madelung energy is given by
Z 2 e2
EM = − α (1039)
RW S
where α is the structure-dependent Madelung constant. The Madelung con-
stants of various structures are evaluated as

Structure α

b.c.c. 0.89593
f.c.c. 0.89587
h.c.p. 0.89584
simple hexagonal 0.88732
simple cubic 0.88006

The Madelung energy is seen to increase as the symmetry is lowered. The


remaining contribution to the electrostatic energy is defined to be the core
energy. The core energy is given by
2
3 Z 2 e2

Rc
Ec = (1040)
2 RW S RW S
and, as it is the electrostatic energy associated with the spherical pseudo-
potential core, it is not dependent on the solid’s structure.

The largest structural-dependent contribution to the energy originates from


the second order terms of the Bloch energies in the electron-ion interaction
(2) 2 m X | Vions (k 0 , k) |2
En,k = (1041)
h̄2 k0 6=k k 2 − k 02

On summing over all the occupied Bloch states ( | k | < kF ) and both spin
values σ, one obtains a contribution E2 to the total energy of
2m X X | Vions (k 0 , k) |2
E2 = (1042)
h̄2 |k|<kF ,σ k0 6=k
k 2 − k 02

379
In the free electron basis, the matrix elements of the electron-ion interaction,
Vions (k 0 , k), only depends on the momentum difference q = k 0 − k.
Z  
0 1 3 0
Vions (k , k) = d r Vions (r) exp i(k − k ).r (1043)
V

The potential due to the lattice can be written as the sum of the individual
potentials from the atoms. The basis position of the j-th atom in the unit cell
is denoted by rj and the Bravais lattice vector is denoted by Ri . Thus, the
potential for the lattice of ions is given by
X
Vions (r) = Vj (r − Ri − rj ) (1044)
i,j

The matrix elements are then given by


Z  
1 3
X
Vions (q) = d r Vj (r − Ri − rj ) exp − i q . r
V i,j
Z    
1 X
= d3 r exp − i q . Ri Vj (r − Ri − rj ) exp − i q . ( r − Ri )
V i,j
(1045)

This can be expressed as


  X  
1 X
Vions (q) = exp − i q . Ri exp − i q . rj Vj (q)
V i j
(1046)

where Vj (q) is related to the Fourier transform of the potential from the j-th
atom of the basis
Z  
Vj (q) = d3 r Vj (r) exp − i q . r (1047)

For simplicity, a crystal with a mono-atomic basis is considered. The matrix


elements are only non-zero when q is a reciprocal lattice vector Q. The matrix
can be expressed in terms of the structure factor S(Q), via

N
Vions (Q) = S(Q) V0 (Q) (1048)
V
The structure dependence of the total electronic energy is contained in the
second order contribution

N2 X 2 m X | S(Q) |2 | V0 (Q) |2
E2 = (1049)
V2
k<kF ,σ
h̄2 Q6=0 k 2 − ( k + Q )2

380
where the sum over k, σ runs over the occupied states (k < kF ), and the term
with Q = 0 is omitted. On interchanging the order of the summations over k
and Q, one finds that the second order term can be expressed in terms of the
Lindhard function χ(q),

N2 X X fk
E2 = 2
| S(Q) |2 | V0 (Q) |2
V Ek − Ek+Q
Q6=0 k,σ

1 N2 X X fk − fk+Q
= 2
| S(Q) |2 | V0 (Q) |2
2 V Ek − Ek+Q
Q6=0 k,σ

1 N2 X
= | S(Q) |2 | V0 (Q) |2 χ(Q) (1050)
2 V
Q6=0

The summation over q is limited to the set of non-zero reciprocal lattice vectors
Q (including their multiplicity). For convenience sake, the first few non-zero
values of the structure factor and the associated reciprocal lattice vectors for an
ideal hexagonal close-packed crystal are tabulated below.

Q (100) (002) (101) (102) (110) (103) (200) (112) (004)

 
Q a
2 π 1.1545 1.225 1.307 1.683 2 2.170 2.309 2.345 2.449

|S(Q)|2 1 4 3 1 4 3 1 4 4

From this we see that, the second-order energy E2 depends on the lattice struc-
ture through the number of equivalent reciprocal lattice vectors, the structure
factors | S(Q) |2 , and on the electron density through the factors χ(q), and the
nature of the ions through V0 (q). The latter is often expressed in terms of the
Thomas-Fermi screened pseudo-potential
cos q Rc
V0 (q) = − 4 π Z e2 (1051)
q2 + kT2 F

where Rc is the radius of the ionic core. The potential has a node at q0 Rc = π2 .
The structural part of the electronic energy depends sensitively on the position
of the node q0 with respect to the smallest reciprocal lattice vectors Q. Recip-
rocal lattice vectors close to a node q0 contribute little to the cohesive energy.
The system may lower its structural energy, if Q moves away from q0 without
causing a change in the volume-dependent contribution to the energy. Recip-
rocal lattice vectors greater than 2 kF contribute little to the cohesive energy,

381
Z 4πe2/q2 Z

χ(q) k+q
Z 4πe2/q2 4πe2/q2 Z

2 2 χ(q)
Z 4πe /q 4πe2/q2 χ(q) 4πe2/q2 Z

Figure 198: A diagramatic depiction of the first few terms in the (R.P.A.) ex-
pression for the screened Coulomb interaction between two charges Z emmersed
in an electron gas. The (red) dashed lines represent the instantaneous Coulomb
interaction between the charges 4π q 2 and the (blue) bubbles represent the polar-
ization induced in the electron gas χ(q). After summing the infinite number of
2 2
terms in this series, the resulting effective interaction is 4 qπ2 Zε(q)e

since the response of the electron gas is negligible.

In addition to these terms, there is a structural contribution arising from


the electron-electron interactions which comes from the induced change in the
electron density
1 X ∗ 4 π e2 V
E2 es = − ρind (q) ρind (q) (1052)
2 q q2

This term occurs since the effect of electron-electron interactions have been
double counted. On noting that the ionic potential only has non-zero Fourier
components at q = Q, and that
ρind (Q) = χ(Q) Vions (Q) (1053)
one can combine the energy contribution represented by eqn(1052) with the
2
contribution from the Bloch energies. The factor 4 πq2 e χ(q) is related to the
dielectric constant ε(q) through

4 π e2
ε(q) = 1 − χ(q) (1054)
q2
The two second order terms can be combined to yield the dominant contribution
to the structural energy
1 N2 X
Estructural = | S(Q) |2 | V0 (Q) |2 χ(Q) ε(Q) (1055)
2 V
Q6=0

382
Since both pseudo-potential terms V0 (Q) include screening, the explicit factor
of ε(q) cancels with one factor of ε(q) in the denominators. Thus, the structural
energy is only screened by one factor of the dielectric constant. The magni-
tude of the structural energy is quite small. The maximum magnitude of the
2
pseudo-potential is ZRec which may be as small as 12 eV. The magnitude of χ
is given by the inverse of the Fermi energy which is typically 5 eV. Thus, the
structural energy is of the order of milli-Rydbergs. Since the structure factor
vanishes unless q = Q, the structural energy depends on the screened potential
only at the reciprocal lattice vectors. Note that the pseudo-potential contains
a node at the wave vector q0 = 2 πRc . The structural energy is composed of
negative contributions, but the contributions from the reciprocal lattice vectors
which are close to the node, contribute little to the stability of the structure. In
fact, reciprocal lattice vectors at the node would correspond to the special case
in which the band gaps at the appropriate Brillouin zone boundaries are zero.
Usually, the opening of a band gap at a Brillouin zone boundary in a conduction
band can result in an increased stability of the structure. The electronic states
below the “band gap” are depressed and, if occupied, result in a lowering of the
solid’s energy. However, the states above the “band gap,” if empty, are raised
but don’t contribute to the solid’s energy.

Al is f.c.c. and the reciprocal lattice vectors (1, 1, 1) and (2, 0, 0) are both
larger than q0 . On moving down the column of the periodic table from Al to
Ga and then In, the ratios of Q/q0 are reduced.

Al Ga In

Q(1, 1, 1)/q0 1.04 0.94 0.93

Q(2, 0, 0)/q0 1.20 1.09 1.08

As the magnitude Q of the six equivalent (2, 0, 0) reciprocal lattice vectors


approaches q0 in Ga, there is a loss in structural stability and the series un-
dergoes a transition from the f.c.c. to a tetragonal structure119 . When this
transition occurs, the set of equivalent f.c.c. reciprocal lattice vectors that have
Q
q0 ∼ 1 are split. In the tetragonal structure, as the structure is sheared, the
reciprocal lattice vectors undergo different changes. Some values of qQ0 move to
higher values while others move to lower values. This type of transformation
119 V. Heine and D. L. Weaire, Solid State Physics, 24, 1 (1970).

383
leaves the atomic volume unchanged, but as all the “band gaps” V (Q) increase,
the transition lowers the energy of the structure. This structural transition oc-
curs when the lowering of the electronic energy outweighs the increase in the
Madelung energy.

On proceeding further down the group III column, from In to T l, the re-
ciprocal lattice vectors pass through the nodes of the pseudo-potential. As a
consequence of this progression, In has a less distorted structure and T l has a
close-packed structure.

10.2 Real Space Representation


The dominant electronic structural energy is given by a sum over all Q of the
form
1 N2 X
Estructural = | S(Q) |2 | V0 (Q) |2 χ(Q) ε(Q) (1056)
2 V
Q6=0

where S(Q) is the structure factor evaluated at a reciprocal lattice vector. This
can be written as a sum over all vectors q, by using the Laue identity
X X  
2 0
N δq,Q = exp i q . ( R − R ) (1057)
Q R6=R0

where R and R0 are Bravais lattice vectors. The structural energy then takes
the form
 
1 X X 0
Estructural = exp i q . ( R − R ) | S(q) |2 θ(q) (1058)
2V 0
q6=0 R6=R

where θ(q) is defined to be

1
θ(q) = | V0 (q) |2 χ(q) ε(q) (1059)
V
It should be noted that in this approximation, θ(q) is independent of the direc-
tion of q. The product of the structure factors can be written as
X  
2
| S(q) | = exp i q . ( ri − rj ) (1060)
i6=j

Thus, on denoting the position of the atoms by Rj = R + rj , one has


 
1 X X
Estructural = exp i q . ( Ri − Rj ) θ(q) (1061)
2
i6=j q6=0

384
(1,1,1) (2,0,0) (2,2,0) (3,1,1)(2,2,2)
0
(2,1,1)(2,2,0)
Li (2,0,0)
θ(q) [Rydbergs] -0.02

-0.04 (1,1,0)
fcc
-0.06
bcc

-0.08

-0.1
0 0.5 1 1.5 2 2.5
q/2kF
Figure 199: The variation of θ(q) with q, as calculated for Li.

The Fourier transform of θ(q) is defined as


X  
θ(Ri,j ) = θ(q) exp i q . Ri,j (1062)
q

where the vector Ri,j denotes the relative position of the two atoms. On chang-
ing the sum over q into an integration, θ(Ri,j ) is evaluated as
Z  
V
θ(Ri,j ) = d3 q θ(q) exp i q . Ri,j
( 2 π )3
Z ∞ Z 1  
2πV 2
= dq q d cos θ θ(q) exp i q Ri,j cos θ
( 2 π )3 0 −1
   
Z ∞ exp i q Ri,j − exp − i q Ri,j !
V
= dq q 2 θ(q)
( 2 π )2 0 i q Ri,j
Z ∞  
V 2 sin q Ri,j
= dq q θ(q) (1063)
2 π2 0 q Ri,j

Thus, the electronic contribution to the structural energy has the real space
representation
1 X
Estructural = θ(Ri,j ) (1064)
2
i6=j

385
The Madelung energy, which is the sum over the interaction energies of the ions
1 X Z 2 e2
EM adelung = (1065)
2 | Ri,j |
i6=j

should also be added to the structural energy. Thus, the total structural energy
can be expressed in terms of the sum of pair-potentials Θ(R), where
Z 2 e2
Θ(R) = + θ(R) (1066)
R
The pair-potential represents the interaction between a pair of bare ions in the
solid plus the effect of the screening clouds. The pair-potential does not de-
scribe the volume dependence of the energy of the solid, but only the structure-
dependent contribution to the energy. The pair-potential can be expressed as
Z ∞
Z 2 e2
 
V 2 sin q R
Θ(R) = + dq q θ(q) (1067)
R ( 2 π2 ) 0 qR
The first and second term can be combined to yield the interaction between an
ion and a screened ion. This can be seen by expressing the potential in terms
of a dimensionless function Ṽ (q) defined by
4 π Z e2
V0 (q) = Ṽ (q) (1068)
q 2 ε(q)
Thus, the interaction can be expressed as
" Z ∞ #
Z 2 e2 4 π e2 χ(q)
 
2 sin q R
Θ(R) = 1 + dq | Ṽ (q) |2
R π 0 q q2 ε(q)
" #

Z 2 e2
 
1 − ε(q)
Z
2 sin q R 2
= 1 + dq | Ṽ (q) |
R π 0 q ε(q)
" #

Z 2 e2
Z  
2 sin q R 2
= 1 − dq | Ṽ (q) |
R π 0 q
" #

Z 2 e2
Z  
2 sin q R 1
+ dq | Ṽ (q) |2
R π 0 q ε(q)
(1069)

The integral in the first term of the last line can be evaluated with the calculus
of residues, and is evaluated in terms of the pole at q = 0. Since Ṽ (0) = 1
and as R > Rc , the integral is equal to unity. Therefore, the first term cancels
identically. Hence, the interaction energy between a bare ion and a screened ion
is given by the expression
" Z ∞ #
Z 2 e2
 
2 sin q R 1 2
Θ(R) = dq | Ṽ (q) | (1070)
R π 0 q ε(q)

386
The long-ranged nature of the Coulomb interaction between the bare ions has
been completely eliminated due to the screening120 . The repulsive ion-ion in-
teraction is reduced to a repulsive core of linear dimension 2 ( Rc + kT−1F ).
The branch cut at q = 2 kF of the logarithmic term in the dielectric constant
leads to Friedel oscillations in the potential at asymptotically large distances R
cos 2 kF R
Θ(R) = A (1071)
( 2 kF R )3

where
A ∝ cos2 2 kF Rc (1072)
However, at intermediate distances, the pair-potential can be approximately
expressed as the sum of three (damped) oscillatory terms121
3
Z 2 e2 X
   
Θ(R) = Bn cos αn 2 kF R + φn exp − βn 2 kF R (1073)
R n=1

where the phase shift depends on the ionic core radius Rc and the electron
density (through rs ). This form is obtained as a result of approximating the
Lindhard function χ(q) by a ratio of polynomials (Padé approximation). The in-
tegration over q can be performed via contour integration. The pairs of complex
poles in the integrand produces terms which have damped oscillatory dependen-
cies on R. The fit parameters for N a are given by:

Na

n 1 2 3

αn 0.291 0.715 0.958


βn 0.897 0.641 0.271
Bn 1.961 0.806 0.023
φn
π 1.706 1.250 1.005

while for M g the interaction is specified by

120 The detailed form of the pair potential is very sensitive to the approximation used in the

calculation of the dielectric function.


121 D. G. Pettifor and M. A. Ward, Solid. State. Commun. 49, 291 (1984).

387
Mg

n 1 2 3

αn 0.224 0.664 0.958


βn 0.834 0.675 0.277
Bn 5.204 1.313 0.033
φn
π 1.599 0.932 0.499

and for Al, one has

Al

n 1 2 3

αn 0.156 0.644 0.958


βn 0.793 0.698 0.279
Bn 7.964 1.275 0.030
φn
π 1.559 0.832 0.431

The contributions to the pair-potential are arranged in order of increasing range


i.e., they are arranged in order of decreasing βn . The Z dependence of the
phase shifts determine the position of the minima of the pair-potential. This
pair-potential, although it only has a magnitude of about 10−2 eV, dominates
the structural energy.

388
0.002

0.001 Al

Θ(R)/Z2e2kF
0 Mg

-0.001

-0.002 Na

-0.003
1 2 3 4 5 6
2 kF R / π

Figure 200: The calculated pair potentials for N a, M g and Al. [After Pettifor
and Ward (1984).]

neighbor shell number 1 2 3 4 5

b.c.c.

number of neighbors 8 6 12 24 8

3
√ √
11

neighbor distance 2 1 2 2 3

f.c.c.

number of neighbors 12 6 24 12 24

2

6
√ √
10
neighbor distance 2 1 2 2 2

h.c.p.

number of neighbors 12

6 2 18

12

2 √2 6 √11
neighbor distance 2 1 3 2 6

389
4
hcp
hcp
fcc

3 Al bcc

Z fcc

2 Mg

bcc
hcp
fcc Na
1
0 0.5 1 1.5 2
φ3/π

Figure 201: The schematic (Z, φ3 ) phase diagram.

The energy difference between the f.c.c. and h.c.p. structures are determined
by the third, fourth and fifth nearest neighbors, as the number and positions
of the nearest and next nearest neighbors are the same. Hence, the relative
stability of this pair of structures is determined by the reasonably long distance
behavior of the pair-potential. The form of the pair-potential can be used to
describe the relative stability122 of the h.c.p. and f.c.c structures of N a, M g
and Al. At ambient pressure, N a123 and M g are predicted to be h.c.p. and
Al is predicted to have an f.c.c. structure. The f.c.c. form of M g is unstable
due to a repulsive contribution from the pair-potentials between the (12) fourth
nearest neighbor pairs. The h.c.p. form of Al is unstable due to a repulsive
contribution from the pair-potentials between the (12) fifth nearest neighbor
pairs. This trend is understood as almost entirely being due to the long-ranged
component of the pair-potential. Basically, as the value of Z increases, when
going across the column from N a to Al, the phase shift of the long-ranged in-
teraction decreases. This means that the oscillations in the pair-potential move
out to larger distances. This causes the changes in the pair-potential at the
positions of the fourth or fifth nearest neighbors.

Under pressure, these materials are predicted to transform to a b.c.c. phase.


The phase shift of the long-ranged component decreases monotonically with in-
creasing Rrsc , which corresponds to increasing pressure. The change in the phase
shifts moves the oscillations in the pair-potentials to distances larger than the
neighbor distances. This shows that as the pressure is increased, one may ex-
pect the energy differences between the h.c.p. and f.c.c. phases to oscillate. The
energy differences between the b.c.c. and close-packed phases originate from the
combined (14) first and second nearest neighbors in b.c.c. and the (12) nearest
122 A. K. McMahan and J. R. Moriarty, Phys. Rev. B 27, 3235 (1983).
123 N a has the close-packed Samarium structure below T = 5 K, and above 5 K N a has a
b.c.c. structure.

390
neighbors of the close-packed structures. The separations of the neighbors in
1
the b.c.c. structure should be scaled by a factor of 2− 3 to yield the same elec-
tron density as the close-packed structures. After this scaling, it is found that
the nearest neighbor distances in the close-packed structures are intermediate
between the nearest neighbor and the next nearest neighbor distances of the
b.c.c. structure. On decreasing the phase shift, one may expect to see the b.c.c.
phase become unstable to a close-packed phase when the (8) nearest neighbors
experience the hard core repulsive potential. On further decreasing the phase
shift, the (12) neighbors of the close-packed phase will experience the same hard
core potential at which point, the b.c.c. becomes stable again. This region of
stability of the b.c.c. structure will remain until the (8) next nearest neighbors
are compressed to distances where the pair-potential has the form of a hard core
repulsion.

These and similar considerations illuminate the origins of the stability of


different structures124 , which are hard to extract from other methods, as the
structural energy typically amounts to only 1% of the cohesive energy of a solid.
In general, the cohesive energy of the solid will also involve three and four-atom
interactions etc., in addition to the pair-potential. Even though these multi-
atom terms contain additional powers of the small factor Vions /EF , they still
can produce significant contributions to the nearest-neighbor interactions125 .
More specifically, the three-body terms do seem to be important in describing
the stability of the more open structures found in materials such as the group
IV elements Si, Ge and Sn. To obtain a more accurate description of structural
stability, it is necessary to utilize density functional calculations.

124 J. Hafner, From Hamiltonians to Phase Diagrams, Springer Verlag, Berlin (1987).
125 E. G. Brovman and Yu. M. Kagan, Sov. Phys. J.E.T.P. 25, 365 (1967).

391
11 Metals
In a metal with Ne electrons, the state with minimum energy has the Ne low-
est one-electron energy eigenvalue states filled with one electron per state (per
spin) in accordance with the Pauli exclusion principle. In a metal, the highest
occupied and the lowest unoccupied state have energies which only differ by an
infinitesimal amount. This energy is called the Fermi energy, F . Thus, the
one-electron states have occupation numbers distributed according to the law
f () = 1 if  < F

f () = 0 if  > F (1074)


The number of electrons in a solid Ne is dictated by charge neutrality to be equal
to the number of nuclear charges N Z. At finite temperatures, the electron
occupation numbers are statistically distributed according to the Fermi-Dirac
distribution function
1
f () =   (1075)
1 + exp β (  − µ )

where β −1 = kB T is the inverse temperature. The Fermi-Dirac distribution


represents the probability that a state with energy  is occupied. Due to the
Pauli exclusion principle, the distribution also represents the average occupation
of the level with energy . The value of the chemical potential coincides with
the Fermi energy at zero temperature µ(0) = F . Since the solid remains
charge neutral at finite temperatures, the chemical potential is determined by
the condition that the solid contains Ne electrons. For a solid with a density of
states given by ρ(), per spin, the total number of electron is given by
Z +∞
2 d ρ() f () = Ne (1076)
−∞

which is an implicit equation for µ. The factor of two represents the number of
different spin polarizations of the electron.

11.1 Thermodynamics
Due to the Pauli exclusion principle, the density of states at the Fermi energy
can often be inferred from measurements of the thermodynamic properties of a
metal. As the characteristic energy scale for the electronic properties is of the
order of eV, and room temperature is of the order of 25 meV, the thermody-
namic properties can usually be evaluated in the asymptotic low-temperature
expansion first investigated by Sommerfeld126 . The low-temperature Sommer-
feld expansion of the electronic specific heat, for non-interacting electrons, shall
126 A. Sommerfeld, Zeit. für Physik, 47, 1 (1928).

392
be examined.

11.1.1 The Sommerfeld Expansion


The total energy of the solid can be expressed as an integral
Z +∞
E = 2 d ρ()  f () (1077)
−∞

Integrals of this type can be evaluated by expressing them in terms of the zero
temperature limit of the distribution and small deviations about this limit.
Z µ
E = 2 d ρ() 
−∞
Z µ  
+2 d ρ()  f () − 1
−∞
Z +∞
+2 d ρ()  f () (1078)
µ

The variable of integration in the terms involving the Fermi function is changed
from  to the dimensionless variable x defined by

 = µ + kB T x (1079)

The Fermi function becomes


1
f (µ + kB T x) = (1080)
1 + exp x
Thus, the integral becomes
Z µ
E = 2 d ρ() 
−∞
Z 0  
+ 2 kB T dx ρ(µ + kB T x) ( µ + kB T x ) f (µ + kB T x) − 1
−∞
Z +∞
+ 2 kB T dx ρ(µ + kB T x) ( µ + kB T x ) f (µ + kB T x)
0
(1081)

The integral over the negative range of x is re-expressed in terms of the new
variable y where
y = −x (1082)
Thus, the energy is expressed as
Z µ
E = 2 d ρ() 
−∞

393
Z ∞  
+ 2 kB T dy ρ(µ − kB T y) ( µ − kB T y ) f (µ − kB T y) − 1
0
Z +∞
+ 2 kB T dx ρ(µ + kB T x) ( µ + kB T x ) f (µ + kB T x)
0
(1083)
However, the Fermi function satisfies the relation
1 − f ( µ − kB T y ) = f ( µ + kB T y ) (1084)
or equivalently
1 1
1 − = (1085)
1 + exp[ − y ] 1 + exp[ y ]
On setting y back to x, one finds
Z µ
E = 2 d ρ() 
−∞
Z +∞
1
+ 2 kB T dx ×
0 1 + exp x
" #
× ρ(µ + kB T x) ( µ + kB T x ) − ρ(µ − kB T x) ( µ − kB T x )

(1086)
The terms within the square brackets can be Taylor expanded in powers of
kB T x, and the integration over x can be performed. Due to the presence of
the Fermi function, the integrals converge. One then has an expansion which is
effectively expressed in powers of kB T / µ. Thus, the energy is expressed as
Z µ
E = 2 d ρ() 
−∞

(  " #)
Z ∞ (2n+1)
X x(2n+1) ( kB T )2n+1 ∂
+ 4 kB T dx µ ρ(µ)
0 n=0
1 + exp x (2n + 1)! ∂µ
(1087)
The integrals over x are evaluated as
Z ∞ Z ∞ ∞
xn X  
dx = dx xn ( − 1 )l+1 exp − l x
0 1 + exp x 0 l=1

X ( − 1 )l+1
= n! (1088)
ln+1
l=1

which are finite for n ≥ 1. Furthermore, the summation can be expressed in


terms of the Riemann ζ functions defined by

X 1
ζ(m) = (1089)
lm
l=1

394
Using this, one finds that

( − 1 )l+1
 2n+1 
X 2 − 1
= ζ(2n + 2) (1090)
l2n+2 22n+1
l=1

The Riemann zeta functions have special values


π2
ζ(2) =
6
π4
ζ(4) = (1091)
90
Thus, the Sommerfeld expansion for the total electronic energy only involves
even powers of T 2 , that is,
Z µ
E = 2 d ρ() 
−∞
∞  2n+1
(  " #)
 (2n+1)
2
X 2 −1 2n ∂
+ 4 ( kB T ) ζ(2n + 2) ( kB T ) µ ρ(µ)
n=0
22n+1 ∂µ
(1092)
The coefficients may be evaluated in terms of the Riemann ζ functions.

Although the expansion contains an explicit temperature dependence, there


is an implicit temperature dependence in the chemical potential µ. This tem-
perature dependence can be found from the equation
Z +∞
Ne = 2 d ρ() f () (1093)
−∞

which also can be expanded in powers of T 2 as


Z µ
Ne = 2 d ρ()
−∞
∞  2n+1
(  " #)
 (2n+1)
X 2 − 1 ∂
+ 4 ( kB T )2 2n+1
ζ(2n + 2) ( kB T )2n ρ(µ)
n=0
2 ∂µ
(1094)
Since Ne is temperature independent, in principle, the series expansion can be
inverted to yield µ in powers of T .

11.1.2 The Specific Heat Capacity


The electronic contribution of the heat capacity, for non-interacting electrons,
can be expressed as
 
∂S
CNe (T ) = T
∂T Ne

395
 
∂E
= (1095)
∂T Ne

as the solid remains electrically neutral. Using the Sommerfeld expansion of the
energy, the specific heat can can be expressed as the sum of the specific heat at
constant µ and a term depending on the temperature derivative of µ at constant
Ne .

(  )
 2n+1  (2n+1) 
2
X 2 −1 2n ∂
CNe = 4 kB T (n + 1) ζ(2n + 2) ( kB T ) µ ρ(µ)
n=0
22n ∂µ
  "
∂µ
+ 2 µ ρ(µ)
∂T Ne
∞  2n+1
(  ) #
 (2n+2) 
2 2
X 2 −1 2n ∂
+ 4 kB T ζ(2n + 2) ( kB T ) µ ρ(µ)
n=0
22n+1 ∂µ
(1096)

In the above expression, µ is to be expanded in powers of T about its zero tem-


perature value µ = F . The temperature derivative of the chemical potential
can be evaluated from the temperature derivative of the equation for the fixed
number of electrons Ne ,

(  )
 2n+1  (2n+1)
2
X 2 −1 2n ∂
0 = 4 kB T (n + 1) ζ(2n + 2) ( kB T ) ρ(µ)
n=0
22n ∂µ
  "
∂µ
+ 2 ρ(µ)
∂T Ne
∞  2n+1
(  )#
 (2n+2)
2 2
X 2 −1 2n ∂
+ 4 kB T ζ(2n + 2) ( kB T ) ρ(µ)
n=0
22n+1 ∂µ
(1097)

This equation yields the temperature dependence of µ which can be substituted


back into the expression for the temperature dependence of CN . This yields the
leading term in the low-temperature expansion for the electronic-specific heat
of non-interacting electrons as
2 4
CN = kB T 4 ζ(2) ρ(µ) + O(kB T 3)
2 2 π2 4
= kB T ρ(µ) + O(kB T 3) (1098)
3
The coefficient of the linear term is proportional to the density of states, per
spin, at the Fermi energy. The result is understood by noting that the Pauli
exclusion principle prevents electrons from being thermally excited, unless they
are within kB T of the Fermi energy. There are ρ(µ) kB T such electrons, and
each electron contributes kB to the specific heat. Thus, the low-temperature

396
2
specific heat is of the order of kB T ρ(µ). The inclusion of electron-electron
interaction changes this result, and in a Fermi liquid, may increase the coefficient
of T . The low-temperature specific heat is enhanced, due to the enhancement of
the quasi-particle masses. This can be demonstrated by a simplified calculation
in which the quasi-particle weight is assumed to be independent of k. Since the
quasi-particle width in the vicinity of the Fermi energy is negligible, one has
the relationship between the quasi-particle density of states and the density of
states for non-interacting electrons given by
X  
ρqp (E) = δ Z(k) E − Ek + µ
k

( Ek − µ )
 
X 1
= δ E −
Z(k) Zk
k
 
X 1
= δ E − Eqp (k) (1099)
Z(k)
k

Also, the quasi-particle density of states at the Fermi energy is un-renormalized


as
X  
ρqp (0) = δ µ − Ek
k

= ρ(µ) (1100)

The γ term in the low-temperature specific heat is calculated from the quasi-
particle entropy S defined in terms of the quasi-particle occupation numbers
nqp
k by

X  
S = − kB nqp
k ln nqp
k + (1 − nqp
k ) ln( 1 − nqp
k )
σ,k
Z ∞  
= − 2 kB dE Z ρqp (E) f (E) ln f (E) + ( 1 − f (E) ) ln( 1 − f (E) )
−∞
(1101)

Thus, in this approximation, the coefficient of the linear T term is given by


CN
γ = lim
T
T → 0

∂S
=
∂T Ne
2 2 π2
= kB Z ρ(µ) (1102)
3
In the more general case, the specific heat coefficient is enhanced through a
k weighted average of the quasi-particle mass enhancement Zk . For materials

397
like CeCu6 , CeCu2 Si2 , CeAl3 and U Be13 , the value of the γ coefficients are
extremely large127 , of the order of 1 J / mole of f ion / K2 , which is 1000
times larger than Cu. The quasi-particle mass enhancements are inferred by
comparison to LDA electronic density of states calculations128 and are about 10
to 30. The enhancement is assumed to be due to the strong electron-electron
interactions, which the LDA fails to take into account.

——————————————————————————————————

11.1.3 Exercise 59
Calculate the next to leading-order term in the low-temperature electronic-
specific heat.

——————————————————————————————————

11.1.4 Exercise 60
CeN iSn is thought to be a zero-gap semiconductor with a V shaped density of
states. The density of states near the Fermi level is approximated by

ρ() = α0  for  > 0


ρ() = − α1  for  < 0 (1103)

where α0 and α1 are positive numbers. Find the leading temperature depen-
dence of the low-temperature specific-heat.

——————————————————————————————————

11.1.5 Pauli Paramagnetism


In the absence of spin-orbit scattering effects, the susceptibility of a metal can
be decomposed into two contributions; the susceptibility due to the spins of the
electrons, and the susceptibility due to the electrons orbital motion. The spin
susceptibility for non-interacting electrons gives rise to the Pauli-paramagnetic
susceptibility which is positive, and is temperature independent at sufficiently
low temperatures. The susceptibility due to the orbital motion has a negative
sign and, therefore, yields the Landau-diamagnetic susceptibility.

127 F. Steglich, J. Aaarts, C. D. Bredl, W. Lieke, D. Meschede, W. Franz and H. Schafer,

Phys. Rev. Lett., 43, 1982 (1979), H. R. Ott, H. Rudigier, Z. Fisk and J. L. Smith, Phys.
Rev. Lett., 50, 1595 (1983), G. R. Stewart, Z. Fisk, J. O. Willis and J. L. Smith, Phys. Rev.
Lett. 52, 679 (1984), G. R. Stewart, Z. Fisk and M. S. Wire, Phys. Rev. B, 30, 482 (1984).
128 C. S. Wang, M. R. Norman, R. C. Albers, A. M. Boring, W. E. Pickett, H. Krakauer,

N. E. Christiensen, Phys. Rev. B, 35, 7260 (1987).

398
The magnetization due to the electronic spins can be calculated from
 
∂Ω
Mz = − (1104)
∂Hz
where the grand canonical potential is given by

X  !
Ω = − kB T ln 1 + exp − β ( Eα − µ ) (1105)
α

and where the sum over α runs over the quantum numbers of the single-particle
states including the spin. The applied magnetic field Hz couples to the quantum
number corresponding to the z component of the spin of the electron, σ, via the
Zeeman energy
g|e|
ĤZeeman = − Hz Sz
2 me c
= − µB Hz σ (1106)

where the spin angular momentum is given by S = h̄2 σ and the gyromagnetic
ratio g = 2 originates from the Dirac or Pauli equation. The quantity µB is
the Bohr magneton and is given in terms of the electron’s charge and mass by
| e | h̄
µB = (1107)
2 me c
The energy of a particle can then be written as

Eσ,k = Ek − µB σ Hz (1108)

where σ is the eigenvalue of the Pauli spin matrix σ̂z . The density of states, per
spin, in the absence of the field is defined as
X
ρ() = δ(  − Ek ) (1109)
k

Thus, in the presence of a field, one has the spin-dependent density of states

ρσ () = ρ( + µB σHz ) (1110)

The grand canonical potential can be expressed as an integral over the density
of states
Z ∞ X  !
Ω = − kB T d ρ( + µB σHz ) ln 1 + exp − β (  − µ )
−∞ σ
Z ∞ X  !
= − kB T d0 ρ(0 ) ln 1 + exp − β ( 0 − µB σHz − µ )
−∞ σ
(1111)

399
where the variable of integration has been changed in the last line. The summa-
tion over σ runs over the values ± 1. The spin contribution to the magnetization
induced by the applied field is given by
Z ∞ X 1
Mz = µB d0 σ ρ(0 )  
−∞ 0
σ 1 + exp β (  − µB σHz − µ )
Z ∞ X
= µB d0 σ ρ(0 ) f ( 0 − µB σHz )
−∞ σ
 
= µB Ne (σ = 1) − Ne (σ = − 1) (1112)

The magnetization due to the spins is just proportional to the number of up-spin
electrons minus the down-spin electrons. The spin susceptibility is given by
∂M z
 
χzz
p (T, H z ) = (1113)
∂Hz
and is given by
Z ∞ X ∂
χp (T, Hz ) = − µ2B d σ 2 ρ() f (  − µB σHz )
−∞ σ
∂
(1114)

It is usual to measure the susceptibility at zero field. Since the derivative of the
Fermi function is peaked around the chemical potential, only electrons within
kB T of the Fermi energy contribute to the Pauli-susceptibility. At sufficiently
low temperatures, one may use the approximation

− f = δ(  − F ) (1115)
∂
so that the zero temperature value of the Pauli-susceptibility is evaluated as

χp (0) = 2 µ2B ρ(F ) (1116)

which is inversely proportional to the free electron mass, and is also proportional
to the density of states at the Fermi energy. The finite temperature susceptibility
can be evaluated by integration by parts, to obtain
Z ∞
2 ∂
χp (T ) = 2 µB d f () ρ() (1117)
−∞ ∂
The zero field spin susceptibility can then be obtained via the Sommerfeld ex-
pansion
∞  2n+1
" (  )#
 (2n+2)
2 2 2
X 2 −1 2n ∂
χp (T ) = 2 µB ρ(µ) + 2 kB T ζ(2n+2) (kB T ) ρ(µ)
n=0
22n+1 ∂µ
(1118)

400
Thus, the spin susceptibility has the form of a power series in T 2 . The temper-
ature dependence of the chemical potential can be found from the equation for
Ne . The leading change in the chemical potential ∆µ due to T is given by
∂ρ(F )
!
2
2 2 π ∂F 4 4
∆µ = − kB T + O( kB T ) (1119)
6 ρ(F )

The temperature dependence of the chemical potential depends on the logarith-


mic derivative of the density of states, such that it moves away from the region
of high density of states to keep the number of electrons fixed. This leads to
the leading temperature dependence of the Pauli susceptibility being given by

( ∂ρ( F)
" #
π 2 2 2 ∂ 2 ρ(F ) )2
 
2 ∂F 4 4
χp (T ) = 2 µB ρ(F ) + k T − + O( kB T )
6 B ∂2F ρ(F )
(1120)
The temperature dependence gives information about the derivatives of the den-
sity of states.

The coefficient γ of the linear T term in the low-temperature specific-heat


and the zero temperature susceptibility are proportional to the density of states
at the Fermi energy. The susceptibility and specific heat can be used to define
the dimensionless ratio
T χp (T ) χp (0)
lim = (1121)
T → 0 C(T ) γ

This ratio is known as the Sommerfeld ratio. For free electrons, this ratio has
the value
T χp (T ) 3 µ2
lim = 2 B2 (1122)
T → 0 C(T ) π kB
The effect of electron-electron interactions can change this ratio, as they may
affect the susceptibility in a different manner than the specific heat. The Stoner
model, discussed in the chapter on magnetism, shows that the effect of electron-
electron interactions can produce a large enhancement of the paramagnetic sus-
ceptibility for electron systems close to a ferromagnetic instability. Thus, near
a ferromagnetic instability, the Sommerfeld ratio is expected to be large. For
example, in P d the ratio is found to be almost a factor of ten larger than pre-
dicted for non-interacting electrons129 . However, for heavy fermion materials
where both C(T )/T and χp (0) are highly enhanced, the value of the Sommerfeld
ratio is very close to that of non-interacting electrons130 .

——————————————————————————————————

129 D .N. Budworth, F. E. Hoare and J. Preston, Proc. Roy. Soc. (London) A 257, 250

(1960), W. E. Gardner and J. Penfold, Phil. Mag. 11, 549 (1965).


130 Z. Fisk, H. R. Ott and J. L. Smith, J. Less-Common Metals, 133, 99 (1987)

401
11.1.6 Exercise 61
Determine the field dependence of the low-temperature Pauli susceptibility.

——————————————————————————————————

11.1.7 Exercise 62
Determine the high temperature form of the Pauli susceptibility.

——————————————————————————————————

11.1.8 Landau Diamagnetism


Free electrons in a magnetic field aligned along the z axis have quantized energies
given by
h̄2 kz2
 
1
Ekz ,n = + n + h̄ ωc (1123)
2m 2
where
| e | Hz
ωc = (1124)
mc
is the cyclotron frequency and n is a positive integer. For a cubic environment
of linear dimension L, the value of kz is given by

kz = nz (1125)
L
The Landau levels have their orbits in the x − y plane quantized and have a
level spacing of h̄ ωc . Each Landau level is highly degenerate. The degeneracy
D, or number of electrons with a given n and kz , can be found as the ratio of
the area of the sample divided by the area enclosed by the classical orbit

L2
D = (1126)
2 π rc2

where rc is the radius of the classical orbit. This radius can be obtained by
equating the field energy with the zero point energy of the Landau level
m 2 2 1
ω c rc = h̄ ωc (1127)
2 2
Thus, the degeneracy is given by

L2
D = m ωc
2 π h̄
| e | L2
D = Hz (1128)
hc

402
Since Hz ∼ 1 kG, a typical value of the degeneracy is of the order of
D ∼ 1010 . These levels can be treated semi-classically as there are an enormous
number of Landau levels in an energy interval. The number of occupied Landau
levels is given by the Fermi energy µ divided by h̄ ωc ,
µ µ
= | e | h̄
(1129)
h̄ ωc Hz
m c

The numerical constant has the value


| e | h̄
∼ 1.16 × 10−8 eV / G (1130)
mc
so, with µ ∼ 1 eV and Hz ∼ 104 G, one finds that the number of occupied
Landau levels is approximately given by
µ
∼ 104 (1131)
h̄ ωc

11.1.9 Landau Level Quantization


The Hamiltonian of a free electron in a magnetic field is given by
 2
|e|
Ĥ = p̂ + A /(2m) (1132)
c
Using the gauge A = (0, Hz x, 0) appropriate for a field along the z axis, then
the Schrodinger equation takes the form

h̄2 e2 Hz2 2
 
2 | e | Hz ∂φ
− ∇ φ − i x + x φ = Eφ (1133)
2m mc ∂y 2 m c2
This can be solved by the substitution
 
φ(r) = f (x) exp i ( ky y + kz z ) (1134)

so that f (x) satisfies


 2  " 2
h̄2 h̄2 kz2
 
∂ f | e | Hz 1
− + h̄ ky + x − E− f (x) = 0
2 m ∂x2 c 2m 2m
(1135)
which is recognized as the equation for the harmonic oscillator with energy
eigenvalue
h̄2 kz2
 
1
E − = n + h̄ ωc (1136)
2m 2

403
where
e2 Hz2
ωc2 = (1137)
m2 c2
That is, the motion in the plane perpendicular to the field, Hz , is quantized
into Landau levels131 . The energy spacing between the levels is given by h̄ ωc ,
where
| e | Hz
ωc = (1138)
mc
and the orbit is centered around the position
h̄ ky c
x0 = − (1139)
| e | Hz
The momentum dependence of the position x0 has a classical analogy. The
center of the classical orbit is determined by its initial velocity vy via

vy = ωc x0 (1140)

so the center of the quantum orbit is determined by py . The energy of the


Landau orbit is given by

h̄2 kz2
 
1
Ekz ,n = + n + h̄ ωc (1141)
2m 2
The degeneracy of the n-th level must correspond to the number of kx , ky values
for Hz = 0 that collapse onto the Landau levels as Hz is increased.

The degeneracy can be enumerated in the case of periodic boundary condi-


tions,
φ(x, y, z) = φ(x, Ly − y, z) (1142)
The periodic boundary conditions imply that
 
exp i ky Ly = 1 (1143)

or

ky = ny (1144)
Ly
The x dependent factor of the wave function f (x) is centered at x0 where
h̄ ky c
x0 = − (1145)
| e | Hz
For a sample of width Lx , one must have Lx > x0 > 0, so one has the equality
| e | Hz Lx
> − ky > 0 (1146)
h̄ c
131 L. D. Landau, Z. Physik, 64, 629 (1930).

404
The degeneracy, D, is the number of quantized ky values that satisfy this in-
equality. The degeneracy is found to be

| e | Hz Lx 2 π
D = /
h̄ c Ly
| e | Hz Lx Ly
= (1147)
2 π h̄ c
independent of n. Thus, the degeneracy D of every Landau is given by

| e | Hz Lx Ly
D = (1148)
2 π h̄ c
The degeneracy can be expressed in terms of the amplitude of the oscillations
in the x direction, which is defined as the length scale that determines the
exponential fall off of the ground state wave function
r

rc = (1149)
m ωc
The degeneracy of the Landau levels can also be expressed as
Lx Ly
D = (1150)
2 π rc2

as previously found from classical considerations. The quantization of the or-


bital motion, in the presence of a periodic potential, has been considered by
Rauh132 and by Harper133 . These authors have shown that the periodic poten-
tial causes the Landau levels to be broadened or split.

11.1.10 The Diamagnetic Susceptibility


The diamagnetic susceptibility is determined from the field dependence of the
grand canonical potential, Ω,
Z ∞ X h̄2 kz2 h̄2 kz2
  
D Lz 1 1
Ω = 2 dkz h̄ ωc ( n + ) + − µ Θ µ −h̄ ωc ( n + ) −
2 π h̄ −∞ n
2 2m 2 2m
(1151)
On integrating over kz , one finds
µ
 12 −1
h̄ωc  32
8 D Lz

2m X2  1
Ω = − µ − h̄ ωc ( n + ) (1152)
3 2π h̄2 n=0
2
132 A. Rauh, Phys. Stat. Solidi, B 65, K131 (1974), A. Rauh, Phys. Stat. Solidi, B 69, K9

(1975).
133 P. G. Harper, Ph.D. Thesis, University of Birmingham (1954), P. G. Harper, Proc. Phys.

Soc. London, A 68, 874 (1955).

405
or
µ
 12 −1
h̄ωc  32
2 V

2m X2  1
Ω = − m ωc µ − h̄ ωc ( n + ) (1153)
3 ( π 2 h̄ ) h̄2 n=0
2

The summation over n can be performed using the Euler-MacLaurin formula


n=N Z N
X 1 1
F (n) = dx F (x) + ( F (0) + F (N ) ) + ( F 0 (N ) − F 0 (0) ) + . . .
n=0 0 2 12
(1154)
This produces the leading-order field dependence of the grand canonical poten-
tial, given by
 1 " #
2 V 2m 2 2 5 1 2 2 1
Ω = − m µ −
2 h̄ ωc µ + . . .
2
3 ( π 2 h̄2 ) h̄2 5 16
(1155)
The diamagnetic susceptibility is given by the second derivative with respect to
the applied field
 2 
∂ Ω
χd = −
∂Hz2
 1
V m 2m 2 1
= − µ2B µ2 (1156)
3 π 2 h̄2 h̄2
where we have expressed the orbital magnetic moment in terms of the (orbital)
Bohr magneton
| e | h̄
µB = (1157)
2mc
The diamagnetic susceptibility χd can be compared with the Pauli paramagnetic
susceptibility χp . For free electrons, the Pauli susceptibility is given by
χp = 2 µ2B ρ(µ)
V m kF
= 2 µ2B
2 π 2 h̄2
 1
2 V m 2m 2 1
= µB 2 2 µ2 (1158)
π h̄ h̄2
Hence, the spin and orbital susceptibilities are related via
1
χd = − χp (1159)
3
Thus, the Landau diamagnetic susceptibility is negative and has a magnitude
which, for free electrons, is just one third of the Pauli paramagnetic susceptibil-
ity134 . The diamagnetism results from the quantized orbital angular momentum
134 L. D. Landau, Z. für Phys. 64, 629 (1930).

406
of the electrons. The value of µB in the diamagnetic susceptibility is given by
the band mass m∗ , whereas the factor of µB in the Pauli susceptibility is defined
in terms of the mass of the electron in vacuum me . In systems such as Bismuth,
in which the band mass is smaller than the free electron mass, the diamagnetic
susceptibility is larger by a factor of
 2
χd 1 me
= − (1160)
χp 3 m∗

and the diamagnetic susceptibility can be larger than the Pauli susceptibility.
The susceptibility of Bismuth is negative.

In the presence of spin-orbit coupling, the orbital angular momenta are cou-
pled with the spin angular momenta. As a result, the components of the total
susceptibility are coupled. The manner in which the total angular momentum
couples to the field is described by the g factor135 .

135 Y. Yafet, Solid State Phys. 14, 1 (1963).

407
11.2 Transport Properties
11.2.1 Electrical Conductivity
The electrical conductivity of a normal metal is considered. The application
of an electromagnetic field will produce an acceleration of the electrons in the
metal. This implies that the distribution of the electrons in phase space will
become time-dependent, and in particular the Fermi surface will be subject to
a time-dependent distortion. However, the phenomenon of electrical transport
in metals is usually a steady state process, in that the electric current density j
produced by a static electric field E is time independent and obeys Ohm’s law

j = σE (1161)

where σ is the electrical conductivity. This steady state is established by scatter-


ing processes that dynamically balances the time-dependent changes produced
by the electric field. That is, once the steady state has been established, the
acceleration of the electrons produced by the electric field is balanced by scat-
tering processes that are responsible for equilibration.

Since Ohm’s law holds almost universally, without requiring any noticeable
non-linear terms in E to describe the current density, it is safe to assume that
the current density can be calculated by only considering the first order terms in
the electro-magnetic field. The validity of this assumption can be related to the
smallness of the ratio of λ | eµ | E where λ is the mean free path, E the strength
of the applied field and µ the Fermi energy. This has the consequence that the
Fermi surface in the steady state where the field is present is only weakly per-
turbed from the Fermi surface with zero field. A number of different approaches
to the calculation of the electrical conductivity will be described. For simplic-
ity, only the zero temperature limit of the conductivity shall be calculated. The
dominant scattering process for the conductivity in this temperature range is
scattering by static impurities.

11.2.2 Scattering by Static Defects


The electrical conductivity will be calculated in which the scattering is due to
a small concentration of randomly distributed impurities. The potential due
to the distribution of impurities located at positions rj , each with a potential
Vimp (r) is given by X
V (r) = Vimp (r − rj ) (1162)
j

This produces elastic scattering of electrons between Bloch states of different


wave vectors. The transition rate in which an electron is scattered from the
state with Bloch wave vector k to a state with Bloch wave vector k 0 is denoted
1
by τ (k→k 0) . If the strength of the scattering potential is weak enough, the

408
transition rate can be calculated from Fermi’s golden rule as
2
1 2 π
< k | V | k > δ( E(k) − E(k 0 ) )
0

0 =
τ (k → k ) h̄
  2
2π 1 X
exp i (k − k ) . (ri − rj ) Vimp (k − k ) δ( E(k) − E(k 0 ) )
0 0

= 2

h̄ V i,j
(1163)

where the delta function expresses the restriction imposed by energy conserva-
tion in the elastic impurity scattering processes. As usual, the presence of the
delta function requires that the transition probability is calculated by integrat-
ing over the momentum of the final state. As the positions of the impurities
are distributed randomly, the scattering rate shall be configurational averaged.
The configurational average of any function is obtained by integrating over the
positions of the impurities
Y  1 Z 
3
F = d rj F ({rj }) (1164)
j
V

The configurational average of the scattering rate is evaluated as


2
1 2 π 1 X
Vimp (k − k ) δ( E(k) − E(k 0 ) ) (1165)
0

0 == 2
τ (k → k ) h̄ V j

where only the term with i = j survives. The conductivity can be calculated
from the steady state distribution function of the electrons, in which the scat-
tering rate dynamically balances the effects of the electric field. This is found,
in the quasi-classical approximation, from the Boltzmann equation.

The Boltzmann Equation.

The distribution of electrons in phase space at time t, f (k, r, t), is deter-


mined by the Boltzmann equation. The Boltzmann equation can be found be
examining the increase in an infinitesimally region of phase space that occurs
during a time interval dt. The number of electrons in the infinitesimal volume
d3 k d3 r located at the point k, r at time t is

f (k, r, t) d3 k d3 r (1166)

The increase in the number of electrons in this volume that occurs in time
interval dt is given by
 

f (k, r, t + dt) − f (k, r, t) d3 k d3 r = f (k, r, t) d3 k d3 r dt + O(dt2 )
∂t
(1167)

409
This increase can be attributed to changes caused by the regular or deterministic
motion of the electrons in the applied field, and partly due to the irregular
motion caused by the scattering. The appropriate time scale for the changes in
the distribution function due to the applied fields is assumed to be much longer
than the time interval in which the collisions occur. The deterministic motion
of the electrons trajectories in phase space results in a change in the number
of electrons in the volume d3 k d3 r. The increase due to these slow time scale
motions is equal to the number of electrons entering the six-dimensional volume
through its surfaces in the time interval dt minus the number of electrons leaving
the volume. This is given by
    
∆ f (k, r, t) d3 k d3 r = − dt ∇ . ṙ f (k, r, t) + ∇k . k̇ f (k, r, t)

(1168)
and the slow rates of change in position and momentum of the electrons is
determined via
h̄ k
ṙ =
m
|e|E
h̄ k̇ = −
m
(1169)
Hence, the deterministic changes are found as
    
h̄ k |e|E
∆ f (k, r, t) d3 k d3 r = − ∇ . f (k, r, t) − ∇k . f (k, r, t) d3 k d3 r dt
m m h̄
(1170)
This involves the sum of two terms, one coming from the change of the electrons
momentum and the other from the change in the electrons position. The two
gradients in this expression can be evaluated, each gradient yields two terms.
One term of each pair involves a gradient of the distribution function, while the
other only involves the distribution function itself. One term, originating from
the change in the electrons position involves the spatial variation of the velocity.
From Hamilton’s equations of motion it can be shown that the coefficient of this
term is equal to the second derivative of the Hamiltonian,
∇ . ṙ = ∇ . ∇p H (1171)

while the similar term originating from the change in particles momentum is
just equal to the negative of the second derivative
∇p . ṗ = − ∇ . ∇p H (1172)

Since the Hamiltonian is ana analytic function these terms are equal magnitude
and of opposite sign. Thus, these terms cancel yielding only
∆ f (k, r, t) d3 k d3 r =

410
    
h̄ k |e|E
= − .∇ f (k, r, t) − . ∇k f (k, r, t) d3 k d3 r dt
m m h̄
(1173)

The remaining contribution to the change in number of electrons per unit time
occurs from the rapid irregular motion caused by the impurity scattering. The
net increase is due to the excess in scattering of electrons from occupied states
at (k 0 , r) into an unoccupied state (k, r) over the rate of scattering out of state
(k, r) into the unoccupied states at (k 0 , r). The restriction imposed by the Pauli
exclusion principle, is that the state to which the electron is scattered into
should be unoccupied in the initial state. This restriction is incorporated by
introducing the probability that a state (k, r) is unoccupied, through the factor
( 1 − f (k, r, t) ).
"
X 1
∆ f (k, r, t) d3 k d3 r = 0 f (k 0 , r, t) ( 1 − f (k, r, t) )
0
τ (k → k)
k
#
1 0
− f (k, r, t) ( 1 − f (k , r, t) ) d3 k d3 r dt (1174)
τ (k → k 0 )

On equating these three terms, cancelling common factors of d3 k d3 r dt one


obtains the Boltzmann equation
      
∂ h̄ k |e|E
f (k, r, t) = − ∇ . f (k, r, t) − ∇k . f (k, r, t) + I f (k, r, t)
∂t m m h̄
(1175)
 
where the functional I f is the collision integral and is given by
  "
X 1
I f (k, r, t) = 0 f (k 0 , r, t) ( 1 − f (k, r, t) )
τ (k → k)
k0
#
1 0
− f (k, r, t) ( 1 − f (k , r, t) )
τ (k → k 0 )
(1176)

Thus, the Boltzmann equation can be written as the equality of a total derivative
obtained from the regular motion and the collision integral which represents the
scattering processes  
d
f (k, r, t) = I f (k, r, t) (1177)
dt
Since
1 1
= (1178)
τ (k → k 0 ) τ (k 0 → k)

411
the collision integral can be simplified to yield
  "  #
X 1
I f (k, r, t) = f (k 0 , r, t) − f (k, r, t) (1179)
0
τ (k → k 0 )
k

Due to time reversal invariance of the scattering rates and conservation of en-
ergy, the collision integral vanishes in the equilibrium state. In equilibrium,
the distribution function is time independent and uniform in space. The dis-
tribution function, therefore, only depends on k in a non-trivial manner, and
can be written in terms of the Fermi function f0 (k). Thus, in this case, the
distributions are related via
1
f (k, r, t) = f0 (k) (1180)
V
The equilibrium distribution function f0 (k) is only a function of the energy
E(k). The condition of conservation of energy which occurs implicity in the
scattering rate requires f0 (k) = f0 (k 0 ). Hence, in equilibrium the collision
integral vanishes.

In the steady state produced by the application of an electric field, the


electron density will be time independent and uniform throughout the metal,
and so the temporal and spatial dependence of f (k, r, t) can still be neglected.
In this case, the distribution function in momentum space is still related to the
distribution function in phase space via
1
f (k, r, t) = f (k) (1181)
V
where f (k) is the non-equilibrium distribution describing the steady state.

The Solution of the Boltzmann equation.

Since the electron distribution in the steady state conduction of electrons is


close to equilibrium one may look for solutions, for f (k) close to the equilibrium
Fermi-Dirac distribution function. Thus, solutions of the form can be sought

∂f0 (k)
f (k) = f0 (k) + Φ(k) (1182)
∂E(k)

where Φ is an unknown function, with dimensions of energy. It is to be shown


that Φ(k) is determined by the electric field and small compared with the Fermi
energy µ. The above ansatz for the non-equilibrium distribution function is
motivated by the notion that the term proportional to Φ occurs from a Taylor
expansion of the steady state distribution function. In other words, the varia-
tion of Φ with k occurs from the distortion of the Fermi surface in the steady
state.

412
If the above ansatz for the steady state distribution is substituted into the
Boltzmann equation one obtains
   
|e|E ∂f0 (E(k))
− ∇k . f (k, r, t) = I Φ(k)
m h̄ ∂E(k)
(1183)

This shows that the energy Φ has a leading term which is proportional to the
first power of the electric field. However, in order to obtain a current that
satisfies Ohm’s law, only the terms in Φ terms linear in E need to be calculated.
Therefore, the Boltzmann equation can be linearized by dropping the term that
involves the electric field and Φ, since this is second order in the effect of the
field. The linearized Boltzmann equation can be solved by noticing that the
collision integral is equal to the source term which is proportional to the scalar
product ( k . E ). Hence, it is reasonable to assume that Φ(k) has a similar
form
Φ(k) = A(E(k)) ( k . E ) (1184)
where A(E) is an unknown function of the energy, or other constants of motion.
Due to conservation of energy, the unknown coefficient can be factored out of
the collision integral, as can be the factor of ∂f ∂E since both are only functions
0

of the energy. It remains to evaluate an integral of the form


Z  
d3 k 0 δ( E(k) − E(k 0 ) ) | Vimp (k − k 0 ) |2 ( k 0 − k ) . E (1185)

The integration over k 0 can be performed by first integrating over the magnitude
of q = k − k 0 . On using the property of the energy conserving delta function,
2m
δ( E(k) − E(k 0 ) ) = δ( q 2 − 2 k q cos θ0 ) (1186)
h̄2
this sets the magnitude of q = 2 k cos θ0 , where the direction of k was chosen
as the polar axis. For simplicity it shall be assumed that the impurity potential
is short-ranged, so that the dependence of V (q) on q is relatively unimportant.
The integration over the factor of q . E can easily be evaluated, and the result
can be shown to be proportional to just ( k . E ) . That is, on expressing the
scalar product as

q . E = ( q sin θ0 cos φ0 Ex + q sin θ0 sin φ0 Ey + q cos θ0 Ez ) (1187)

on integrating over the azimuthal angle φ0

dΩ0 = dθ0 sin θ0 dφ0 (1188)

the terms proportional to Ex and Ey vanish. The integration over the polar
angle θ0 produces a factor
Z 1
2m
8 π 2 k2 d cos θ0 cos3 θ0 | V ( 2 k cos θ0 ) |2 Ez (1189)
h̄ −1

413
This yields the result
Z 1
2m
= 4π k(k.E)2 d cos θ0 cos3 θ0 | V ( 2 k cos θ0 ) |2 (1190)
h̄2 −1

which can be expressed as an integral over the scattering angle θ = π − 2 θ0


Z 1
2m θ
= 2π 2 k(k.E) d cos θ ( 1 − cos θ ) | V ( 2 k sin ) |2 (1191)
h̄ −1 2

On identifying the non-equilibrium part of the distribution function with

∂f0 (k) ∂f0 (k)


Φ(k) = ( k . E ) A(E) (1192)
∂E(k) ∂E(k)

yields the solution for the non-equilibrium contribution of the distribution func-
tion as  
|e|
Φ(k) = + τtr (k) E . ∇k E(k) (1193)

Thus, Φ is proportional to the energy change of the electron produced by the
electric field in the interval between scattering events. In the above expression,
the term
 
1 2π X 0 0 2
= c δ( E(k) − E(k )) | V (k − k ) | 1 − cos θ (1194)
τtr (k) h̄ 0
k

is identified as the transport scattering rate, in which c is the concentration of


impurities. The transport scattering rate has the form of the rate for scattering
out of the state k but has an extra factor of ( 1 − cos θ ). In the quantum for-
mulation of transport this factor appears as a vertex correction. Basically, the
electrical current is related to the momentum of the electrons in the direction
of the applied field. Forward scattering processes do not result in a reduction
of the momentum and, therefore, leave the current unaffected. The transport
scattering rate involves a factor of ( 1 − cos θ ) where θ is the scattering angle.
This factor represents the relative importance of large angle scattering in the
reduction of the total current.

The Current Density.

The current density can be obtained directly from the expression.

1 X 1 ∂E(k)
j = −2|e| f (k)
V h̄ ∂k
k
 
1 X 1 ∂E(k) ∂f0 (k)
= −2|e| f0 (k) + Φ(k)
V h̄ ∂k ∂E(k)
k
(1195)

414
where the factor of 2 represents the sum over the electron spins. On viewing
the electron distribution function as the first two terms in a Taylor expansion,
the electron distribution function can be described by an occupied Fermi vol-
ume which has been displaced from the equilibrium position in the direction of
the applied field. The displacement of the Fermi volume produces the average
current in the direction of the field. The first term represents the current that
is expected to flow in the equilibrium state. This term is zero, as can be seen
by using the symmetry of the energy E(k) = E(−k) in the Fermi function.
Due to the presence of the velocity vector h̄1 ∇k E(k), it can be seen that the
current produced by an electron of momentum k identically cancels with the
current produced by an electron of momentum −k.

Thus, the non-zero component of the current originates from the non-equilibrium
part of the distribution function. This can only be evaluated once the Bloch
energies are given. The current is given by

e2 X
  
∂f0 (k)
j = −2 2 τ (k)tr ∇k E(k) ∇k E(k) . E (1196)
h̄ ∂E(k)
k

On recognizing the zero temperature property of the Fermi function


 
∂f0 (k)
− = δ( E(k) − µ ) (1197)
∂E(k)

it is seen that the electrical current is carried by electrons in a narrow energy


shell around the Fermi surface. On using the symmetry properties of the inte-
gral, one finds that only the diagonal component of the conductivity tensor is
non-zero and is given by

2 δα,β e2 X
 
2 ∂f0 (k)
σα,β = − τ (k)tr | ∇k E(k) | (1198)
3 h̄2 k ∂E(k)

For free electron bands, the conductivity tensor is evaluated as

ρ e2 τtr
σα,β = δα,β (1199)
m
where ρ is the density of electrons, m is the mass of the electrons and τtr is the
Fermi surface average of the transport scattering rate τtr (k).

——————————————————————————————————

11.2.3 Exercise 63
Determine the conductivity tensor σα,β (q, ω) which relates the Fourier compo-
nent of a current density jα (q, ω) to a time and spatially varying applied electric

415
field with a Fourier amplitude Eβ (q, ω) via Ohm’s law
X
jα (q, ω) = σα,β (q, ω) Eβ (q, ω) (1200)
β

Assume that ω is negligibly small compared with the Fermi energy so that the
scattering rate can be evaluated on the Fermi surface.

The above result should show that in the zero frequency limit ω → 0
the q = 0 conductivity is purely real and given by the standard expression
2
σα,β (0, 0) = δα,β ρ em τtr , and decreases for increasing ω. The frequency width
of the Drude peak is given by the scattering rate τ1tr .

——————————————————————————————————

11.2.4 The Hall Effect and Magneto-resistance.


The Hall effect occurs when an electrical current is flowing in a sample and a
magnetic field is applied in a direction transverse to the direction of the cur-
rent density. Consider a sample in the form of a rectangular prism, with axes
parallel to the axes of a Cartesian coordinate system. The magnetic field is ap-
plied along the z direction and a current flows along the y direction. The Hall
effect concerns the appearance of a voltage (the Hall voltage) across a sample
in the x direction. The Hall voltage appears in order to balance the Lorentz
force produced by the motion of the charged particles in the magnetic field.
The initial current flow in the x direction sets up a net charge imbalance across
the sample in accordance with the continuity equation. The build up of static
charge produces the Hall voltage. In the steady state, the Hall voltage balances
the Lorentz force opposing the further build up of static charge. The sign of the
Hall voltage is an indicator of the sign of the current carrying particles.

The Hall Electric field is given by

E x êx = + vy Bz êy ∧ êz (1201)

The Hall voltage VH is related to the electric field and the width of the sample
dx via

VH = − E x dx = − vy Bz dx
jy Bz dx
= −
ρq
(1202)

Hence, measurement of the Hall voltage VH and jy , together with the magnitude
of the applied field Bz , determines the carrier density ρ and the charge q. This

416
is embodied in the definition of the Hall constant, RH
Ey
RH = (1203)
jx Hz
which for semi-classical free carriers of charge q and density ρ is evaluated as
1
RH = (1204)
ρq

In other geometries, one notices that the current will flow in a direction other
than parallel to the applied field. The conductivity tensor will not be diagonal,
as will the resistivity tensor. The dependence of the resistivity on the magnetic
field is known as magneto-resistance. The phenomenon of transport in a mag-
netic field can be quite generally addressed from knowledge of the conductivity
tensor in an applied magnetic field. This can be calculated using the Boltzmann
equation approach.

The Boltzmann Equation.

The Boltzmann equation for the steady state distribution f (p), in the pres-
ence of static electric and magnetic fields, can be expressed as
   
− | e | E + v ∧ B . ∇p f (p) = I f (p) (1205)

Since only a solution for f (p) is sought which contain terms linear in the electric
field E, the equation can be linearized by making the substitution f (p) → f0 (p)
but only in the term explicitly proportional to E.
 
− | e | E . ∇p f0 (p) − | e | ( v ∧ B ) . ∇p f (p) = I f (p) (1206)

The substitution of the zero-field equilibrium distribution function f0 (p) in the


first term (without any magnetic field corrections) is consistent with the equi-
librium in the presence of a static magnet field. This can be seen by examining
the limit E = 0, where the Boltzmann equation reduces to
 
− | e | ( v ∧ B ) . ∇p f (p) = I f (p) (1207)

which has the solution f (p) = f0 (p) since in this case the collision integral
vanishes and the remaining term is also zero as
∂f0 (p)
( v ∧ B ) . ∇p f0 (p) = ( v ∧ B ) . ∇p E(p)
∂E(p)
∂f0 (p)
= (v ∧ B).v = 0
∂E(p)
(1208)

417
since due to the vector identity

(A ∧ B).A = 0 (1209)

the scalar product vanishes. This is just a consequence of the fact that a mag-
netic field does not change the particles energy. The observation can be used
to simplify the Boltzmann equation as, has been seen, the magnetic force term
only acts on the deviation from equilibrium.

The ansatz for the steady state distribution function is

∂f0 (p)
f (p) = f0 (p) + Φ(p)
∂E(p)
∂f0 (p)
= f0 (p) + v . C (1210)
∂E(p)

where C is an unknown vector function. It shall be shown that the vector C


is independent of p. In this case, the collision integral simplifies to the case
that was previously considered. Namely, the collision integral reduces to the
transport scattering rate times the non-equilibrium part of the steady state
distribution function. On cancelling the common factor involving the derivative
of the Fermi function, and using
1
∇p . v = (1211)
m∗
one finds
|e| 1
|e|E.v + (v ∧ B).C = (v.C ) (1212)
m∗ c τtr
The solution of this equation is independent of v, hence C is a constant vector.
This can be seen explicitly by substituting the identity

(v ∧ B).C = (B ∧ C ).v (1213)

back into the Boltzmann equation. The resulting equation can be solved for all
v if C satisfies the algebraic vector equation
 
|e| 1
|e|E + B ∧ C = C (1214)
m∗ c τtr

To solve the above algebraic equation it is convenient to change variables


 
|e|
ωc = B (1215)
m∗ c

This shall be solved by finding the components parallel and transverse to B.

418
If the scalar product of the algebraic equation is formed with ω c and on
recognizing that ω c . ( ω c ∧ C ) = 0, one finds that the component of C
parallel to the magnetic field is given by

ω c . C = | e | τtr ω c . E (1216)

The transverse component of C can be obtained by taking the vector product


of the algebraic equation with ω. This results in the equation
1
ωc ∧ ( | e | E ) + ωc ∧ ( ωc ∧ C ) = ω ∧ C (1217)
τtr c
but
ω c ∧ ( ω c ∧ C ) = ω c ( ω c . C ) − ωc2 C (1218)
so one recovers the relation between the transverse component and C from
1
ω c ∧ ( | e | E ) + ω c ( ω c . C ) − ωc2 C = ω ∧ C (1219)
τtr c
by eliminating the longitudinal component. The resulting relation is found as
1
ω c ∧ ( | e | E ) + ω c ( | e | τtr ω c . E ) − ωc2 C = ω ∧ C (1220)
τtr c

The transverse component can be substituted back into the original algebraic
equation to find the complete expression for C.
 
1 1
| e | ωc ∧ E + τ ωc ( ωc . E ) − ωc2 C = 2 C − | e | E (1221)
τ τ

Therefore, C is given by the constant vector


   
2 2 2
1 + ωc τ C = τ | e | E + τ ( ωc . E ) ωc + τ ( ωc ∧ E ) (1222)

which only depends upon E and B but not on the momentum p. This leads to
the explicit expression for the non-equilibrium distribution function of

f (k) = f0 (k)
 
τ |e| 2 ∂f0 (k)
+ 2 2
v . E + τ ( v . ω c ) ( ω c . E ) + τ ( v ∧ ω c ) . E
1 + ωc τ ∂E(k)
(1223)

The deviation from equilibrium can be interpreted in terms of an anisotropic


displacement of the Fermi function involving the work done by the electric field

419
on the electron in the time interval between scattering events.

The Conductivity Tensor.

The average value of the current density is given by


2 X
j = −|e| v(k) f (k) (1224)
V
k

where f (k) is the steady state distribution function, and the factor of 2 rep-
resents the summation over the electrons spin. On substituting for the steady
state distribution function, and noting that because of the symmetry k → − k
in the equilibrium distribution function, no current flows in the absence of the
electric field. The current density j is linear in the magnitude of the electric
field E, and is given by

d3 k
Z  
2 τ ∂f0
j = 2|e | v − ×
1 + ωc2 τ 2 ( 2 π )3 ∂E
" #
× ( v . E ) + τ 2 ( v . ωc ) ( ωc . E ) + ( v ∧ ωc ) . E

(1225)

Thus, the conductivity tensor is recovered in dyadic form as


d3 k
Z  
τ ∂f0
e = 2 | e2 |
σ − ×
1 + ωc2 τ 2 ( 2 π )3 ∂E
" #
× v v + τ 2 ( v . ωc ) v ωc + τ v ( v ∧ ωc )

(1226)

Furthermore, if E(k) is assumed to be spherically symmetric, one finds that the


components of the tensor can be expressed as
∂f0 v 2
Z    
2 τ 2
σα,β = e dE ρ(E) − δα,β + τ ωα ωβ ± ( 1 −δα,β ) τ ωγ
∂E 3 1 + ω2 τ 2
(1227)
where in the off-diagonal term the convention is introduced such that γ is chosen
such that (α, β, γ) corresponds to a permutation of (x, y, z). Since the density
3
of states per unit volume, ρ(E), is proportional to E 2 , the conductivity tensor
can be evaluated, by integration by parts, to yield
ρ e2
 
τ 2
σα,β = δα,β + τ ωα ωβ ± ( 1 − δα,β ) τ ωγ (1228)
m 1 + ω2 τ 2

where the ± sign is taken to be positive when (α, β, γ) are an odd permutation of
(x, y, z) and is negative when (α, β, γ) are an even permutation of (x, y, z). Thus,

420
if the field is applied along the z direction it is found the diagonal components
of the conductivity tensor are given by
ρ e2 τ
σx,x = σy,y =
m 1 + ωc2 τ 2
ρ e2
σz,z = τ (1229)
m
The non-zero off-diagonal terms are found as
ρ e2 ωc τ 2
σy,x = − σx,y = (1230)
m 1 + ωc2 τ 2
Thus, for the diagonal component of the conductivity tensor are anisotropic.
The component parallel to the field is constant while the other two components
decrease like ωc−2 in high fields. The off-diagonal components are zero at zero-
field, but increase linearly with the field for small ωc but then decreases like ωc−1
at high fields.

A useful representation of the conductivity is through the Hall angle. For


example, if one applies the magnetic field along the z-direction and then an
electric field along the x-direction Ex 6= 0, then the current will have an x and
y component that can be characterized by a complex number z
z = Jx + i Jy
 
1 − i ωc τ
= σ 0 Ex
1 + ωc2 τ 2
1
= σ0 (1231)
1 + i ωc τ
This complex number z lies on a semi-circle of radius σ20 Ex centered on the
point ( σ20 Ex , 0), as
 
σ0 σ0 1 − i ωc τ
z − Ex = Ex (1232)
2 2 1 + i ωc τ
and the modulus is just given by
σ0 σ0
|z − Ex | = Ex (1233)
2 2
Thus, the number z lies on a semi-circle of radius σ20 Ex passing through the
origin. The Hall angle ΨH is defined as the angle between the line subtended
from the point z to the origin and the Jx axis. Thus
Jy
tan ΨH = (1234)
Jx
and from the Boltzmann equation analysis of the magneto-conductivity
ΨH = tan−1 ωc τ (1235)

421
Thus, from knowledge of σ0 and E one can find z and, thence, J.

The resistivity tensor ρi,j is obtained from the conductivity tensor σi,j by
inverting the relation X
Ji = σi,j Ej (1236)
j

to obtain X
Ei = ρi,j Jj (1237)
j

The resistivity tensor is found as


 
ρ0 ρ0 ω c τ 0
ρe =  − ρ0 ωc τ ρ0 0 
0 0 ρ0

Thus, for the free electron model the diagonal part of the resistivity tensor is
completely unaffected by the field. There is neither a longitudinal or transverse
magneto-resistance.

However, as Hz increases, the transverse component of the electric field Ey


increases. This is the Hall field. The Hall field is given by
J x Hz
Ey = ωc τ ρ0 Jx = (1238)
ρ|e|

Thus, the Hall resistivity is


Ey
ρyx =
Jx
Hz
= (1239)
ρ|e|

Thus, the Hall constant RH is given by


EY 1
RH = = (1240)
Hz J x ρ|e|

The magneto-resistivity is usually classified as being longitudinal or trans-


verse. The longitudinal magneto resistance is the change in the resistivity tensor
ρz,z due to the application of a magnetic field along the z direction. The trans-
verse magneto-resistance is given by the change in ρx,x or ρy,y due to a field
Hz . The longitudinal magneto-resistance is usually due to the dependence of the
scattering rate on the magnetic field, whereas the transverse magneto-resistance
is due to the action of the Lorentz force.

The general features of the magneto-resistance are:-

422
(i) for low fields, Hz such that ωc τ < 1 then

∆ρx,x = ρx,x (Hz ) − ρx,x (0) ∝ Hz2 (1241)

(ii) There is an electric field Ey transverse to Jx and Hz , which has a mag-


nitude proportional to Hz .

(iii) For large fields ρx,x (Hz ) may either continue to increase with Hz2 or
saturate.

(iv) For a set of samples all which have different residual resistivities ρzz (T =
0, Hz = 0), then the transverse magneto resistance usually satisfies Koehler’s
rule  
∆ρx,x (Hz ) Hz
= F (1242)
ρx,x (T = 0, Hz = 0) ρx,x (0, 0)
Basically, Koehler’s law expresses the fact that ρ(Hz ) only depends on Hz
through the combination ωc τ and that ∆ρx,x and ρzz (T = 0, Hz = 0) are
both proportional to τ −1 .

The standard form of the relationship between E and J is expressed as a


vector equation

E = ρ0 J + a ( J ∧ H ) + b H 2 J
+ c ( J . H ) H + d Te J (1243)

where Te is a tensor which only has diagonal components that, when referred to
the crystalline axes, are (Hx2 , Hy2 , Hz2 ). That is Te is the matrix

Hx2
 
0 0
 0 Hy2 0 
0 0 Hz2
(1244)

The five unknown quantities may be determined by five experiments.

(1) When J and H are parallel to the x axis there is the longitudinal magneto-
resistance given by

ρx,x = ρ0 + ( b + c + d ) H 2 (1245)

(2) When J k x , H k y then

ρx,x = ρ0 + b H 2 (1246)

423
which is the transverse magneto-resistance.
(3) With J k (1, 1, 0) i.e.
J
J = √ (1, 1, 0) (1247)
2
and
H
H = √ (1, 1, 0) (1248)
2
there is a different longitudinal magneto-resistance
d
∆ρ = ( b + c + ) H2 (1249)
2

(4) When H = H (0, 0, 1), then a second transverse magneto-resistance is


found as
∆ ρ = b H2 (1250)
H
but when H = √
2
(1, −1, 0) then the magneto-resistance is found as

d
∆ρ = (b + ) H2 (1251)
2

(5) The constant a makes no contribution to the magneto-resistance, but is


found from the Hall effect. If H is transverse to J then the Hall effect is only
determined by a alone, and is isotropic.

The magneto-resistance is usually negative except for cases where the scat-
tering is of magnetic origin, such as disorder with spin - orbit coupling or from
Kondo scattering by magnetic impurities in metals136 .

11.2.5 Multi-band Models


The transverse magneto-resistance for a multi-band model is non-trivial, unlike
the one band free electron model. The resistance can be obtained from the
current field diagram, in which the currents originating from the various sheets
of the Fermi surface are considered separately.

For example, a two band model with positive and negative charge carriers
produces two components of the current J+ and J− by virtue of their responses
σ+ , σ− in response to the electric field Ex . On assuming that the carriers have
the same Hall angles ΨH , then the total current is found as
 
σ+ + σ−
Jx = Ex ( 1 + cos 2 ΨH ) (1252)
2
136 J. Kondo, Prog. Theor. Phys. 32, 37 (1964).

424
and  
σ+ − σ −
Jy = Ex sin 2 ΨH (1253)
2
Thus,

σx,x = σ0 cos2 ΨH
 
ρ+ − ρ−
σy,x = σ0 sin ΨH cos ΨH
ρ+ + ρ−
(1254)

Since the conductivity tensor is anisotropic and given by


 
σx,x σx,y 0
σ̃ =  − σy,x σx,x 0 
0 0 σz,z

then the transverse magneto-resistivity can be found from


σx,x σz,z
ρx,x = 2 2
σz,z ( σx,x + σx,y )
σx,x
= 2 2
( σx,x + σx,y )
1 cos2 ΨH
= 2
σ0

ρ+ − ρ−
( cos4 ΨH + ρ+ + ρ− sin2 ΨH cos2 ΨH )
1 1
= 2
σ0

ρ+ − ρ−
( cos2 ΨH + ρ+ + ρ− sin2 ΨH )

1 sec2 ΨH
= 2 (1255)
σ0

ρ+ − ρ−
1 + ρ+ + ρ− ωc2 τ 2

since tan Ψh = ωc τ .

Furthermore, as

sec2 ΨH = 1 + tan2 ΨH
= 1 + ωc2 τ 2 (1256)

then
1 ( 1 + ωc2 τ 2 )
ρx,x = 2 (1257)
σ0

ρ+ − ρ−
1 + ρ+ + ρ− ωc2 τ 2

This saturates if | ρ+ − ρ− | > 0 and increases indefinitely for a compensated


metal ρ+ = ρ− . Basically, the positive magneto-resistance occurs because the

425
Lorentz force produces a transverse component of the current in each sheet of
the Fermi surface. The Lorentz force, the acting on these transverse currents
then produces a shift of the Fermi surface opposite to the shift produced by the
electric field.

A similar analysis can be performed on the Hall coefficient


E⊥
RH = (1258)
H J
The value of E⊥ is the component of the field perpendicular to the current.
This is found from the angle θ between J and E
Jx
cos θ =
J
Jy
sin θ = (1259)
J
Thus
E sin θ E Jy
RH = = 2
H J H J 
ρ+ − ρ− tan ΨH
1 ρ+ + ρ− cos2 ΨH
= 2
σ0 H

ρ+ − ρ−
1 + ρ+ + ρ− tan2 ΨH
 
ρ+ − ρ−
ρ+ + ρ− ( 1 + ωc2 τ 2 )
ωc τ
= 2
σ0 H

ρ+ − ρ−
1 + ρ+ + ρ− tan2 ΨH
 2
ρ+ − ρ−
ρ+ + ρ− ( 1 + ωc2 τ 2 )
1
= 2
| e | ( ρ+ − ρ− )

ρ+ − ρ−
1 + ρ+ + ρ− tan2 ΨH

(1260)

The Hall coefficient saturates to


1
RH → (1261)
| e | ( ρ+ − ρ− )

for large magnetic fields.

426
11.3 Electromagnetic Properties of Metals
Maxwell’s equations relate the electromagnetic field to charges and current
sources ρ(r; t) and j(r; t). Maxwell’s equations can be formulated as

∇ . E(r; t) = 4 π ρe (r; t)
∇ . B(r; t) = 0
4π 1 ∂E(r; t)
∇ ∧ B(r; t) = j(r; t) +
c c ∂t
1 ∂B(r; t)
∇ ∧ E(r; t) = − (1262)
c ∂t
where E and B represent the microscopic electric and magnetic fields, and ρe and
j are the microscopic charge and current densities. There are eight equations
for the six unknown quantities. The six unknown quantities are the components
of E and B.

The sourceless equations have a formal solution in terms of a scalar potential


φ and a vector potential A which are related to the electric field E and magnetic
field B via
1 ∂A
E = −∇φ −
c ∂t
B = ∇ ∧ A (1263)

The solutions for the potentials are not unique, as the gauge transformations

A → A0 = A + ∇ Λ (1264)

and
1 ∂Λ
φ → φ0 = φ − (1265)
c ∂t
where Λ is an arbitrary scalar function, yield new scalar and vector potentials,
(A0 , φ0 ), that produce the same physical E and B fields as the original potentials
(A, φ). The four quantities φ and A satisfy the four source equations
 
2 1 ∂
∇ φ + ∇.A = − 4 π ρe (1266)
c ∂t

and
1 ∂2A
 
1 ∂φ 4π
∇2 A − − ∇ ∇.A + = − j (1267)
c2 ∂t2 c ∂t c

These equations are usually simplified by choosing a gauge condition. The gauge
conditions which are usually chosen are either the Coulomb Gauge

∇.A = 0 (1268)

427
or the Lorentz Gauge
1 ∂φ
∇.A + = 0 (1269)
c ∂t
The Lorentz gauge has the advantage that it is explicitly covariant under Lorentz
transformations. The Coulomb gauge, also known as the transverse gauge or
radiation gauge, is quite convenient for non-relativistic problems in that it sep-
arates the effects of radiation from electrostatics.

The space and time Fourier transform of the charge density is defined as
Z Z  
1
ρe (q, ω) = d3 r dt exp − i ( q . r − ω t ) ρe (r; t) (1270)
V

On Fourier transforming the source equations with respect to space and time,
one has
ω
− q 2 φ(q, ω) + q . A(q, ω) = − 4 π ρe (q, ω)
c
ω2
  
2 ω 4π
− q + 2 A(q, ω) + q q . A(q, ω) − φ(q, ω) = − j(q, ω)
c c c
(1271)

In the wave-vector and frequency domain, the Coulomb gauge condition is ex-
pressed as
q . A(q, ω) = 0 (1272)
which shows that the vector potential is transverse to the direction of q. In the
transverse gauge, the equation for the vector potential reduces to

ω2
   
2 ω 4π
− q + 2 A(q, ω) − q φ(q, ω) = − j(q, ω)
c c c
(1273)

The first term is transverse and the second term is longitudinal. Thus, the
current can also be divided into a longitudinal term
 
j L (q, ω) = q̂ q̂ . j(q, ω) (1274)

and a transverse term


 
j T (q, ω) = j(q, ω) − q̂ q̂ . j(q, ω) (1275)

Thus, the second non-trivial Maxwell equation separates into the transverse
equation

ω2
 
2 4π
− q + 2 A(q, ω) = − j (q, ω) (1276)
c c T

428
and the longitudinal equation
ω 4π
−q φ(q, ω) = − j (q, ω) (1277)
c c L
In the Coulomb gauge, the other non-trivial Maxwell equation relates the charge
density to the scalar potential via
− q 2 φ(q, ω) = − 4 π ρe (q, ω) (1278)
This is just Poisson’s equation, and it has the solution

φ(q, ω) = 2 ρe (q, ω) (1279)
q
which is equivalent to Coulomb’s law. When Fourier transformed with respect
to space and time, Poisson’s equation yields an instantaneous relation between
the charge density and the scalar potential in the form of Coulomb’s law. Al-
though this is an instantaneous relation, the signals transmitted by the electro-
magnetic field still travel with speed c and are also causal. This is because, in
the Coulomb gauge, the retardation effects are contained in the vector potential.

Poisson’s equation actually has the same content as the longitudinal equa-
tion. This can be seen by examining the continuity equation which expresses
conservation of charge
∂ρe
+ ∇.j = 0 (1280)
∂t
The continuity equation can be Fourier transformed to yield
− ω ρe (q, ω) + q . j(q, ω) = 0 (1281)
This shows that the fluctuations in the charge density are related to the longi-
tudinal current. On solving the continuity condition, one finds that the longi-
tudinal current is given by
ω
j L (q, ω) = q̂ ρe (q, ω) (1282)
q
On substituting the above expression for the longitudinal current into the lon-
gitudinal equation, one finds
ω 4π ω
−q φ(q, ω) = − q̂ ρe (q, ω) (1283)
c c q
On cancelling the factors of ω/c and q, one obtains Poisson’s equation. This
proves that the longitudinal equation is equivalent to Poisson’s equation. We
have also found that the longitudinal current can be expressed in the forms
ω
j L (q, ω) = q̂ ρe (q, ω)
q

= φ(q, ω) (1284)

so the longitudinal current can be viewed as being produced either by the charge
density or by the scalar potential.

429
11.3.1 The Longitudinal Response
The currents and charge densities are usually broken down into the external
contributions and the induced contributions via
j(q, ω) = j ind (q, ω) + j ext (q, ω)
ρe (q, ω) = ρe ind (q, ω) + ρe ext (q, ω) (1285)
The external scalar potential is defined in terms of the external charge density
via Poisson’s equation
− q 2 φext (q, ω) = − 4 π ρe ext (q, ω) (1286)
The frequency and wave vector-dependent dielectric constant for a homogeneous
medium, ε(q, ω), is defined by the ratio

φext (q, ω)
ε(q, ω) = (1287)
φ(q, ω)
The dielectric constant describes the screening of the external potential by lon-
gitudinal or charge density fluctuations. The dielectric constant is related to
the longitudinal conductivity. This can be seen by reducing the relation

j L (q, ω) = φ(q, ω) (1288)

into an expression for the induced component of the longitudinal current

 
j L (q, ω)ind = φ(q, ω) − φext (q, ω) (1289)

Hence, on using the definition of the frequency-dependent dielectric constant to
eliminate φext (q, ω), one obtains

 
j L (q, ω)ind = 1 − ε(q, ω) φ(q, ω) (1290)

The total scalar potential φ(q, ω) can be related to the longitudinal electric
field, E L (q, ω), since the electric field can be written as the sum of the time
dependence of the vector potential and the gradient of the scalar potential

E(q, ω) = A(q, ω) − i q φ(q, ω) (1291)
c
If the longitudinal part of the electric field is identified as
E L (q, ω) = − i q φ(q, ω) (1292)
then one obtains the relation between the longitudinal current and the longitu-
dinal electric field in teh form
 

j L (q, ω)ind = 1 − ε(q, ω) E L (q, ω) (1293)

430
Hence, as the longitudinal conductivity σL is defined by the relation

j L (q, ω)ind = σL (q, ω) E L (q, ω)


(1294)

one finds that the conductivity and the dielectric constant are related through
 

σL (q, ω) = 1 − ε(q, ω) (1295)

The frequency-dependent dielectric constant can be expressed in terms of the


response of the charge density due to the potential

φext (q, ω)
ε(q, ω) =
φ(q, ω)
φ(q, ω) − φind (q, ω)
ε(q, ω) =
φ(q, ω)
4 π ρe ind (q, ω)
ε(q, ω) = 1 − (1296)
q2 φ(q, ω)

The charge density is related to the electron density via a factor of the electron’s
charge
ρe ind (q, ω) = − | e | ρind (q, ω) (1297)
and the scalar potential acting on the electrons produces the potential δV (q, ω)
where
δV (q, ω) = − | e | φ(q, ω) (1298)
Thus, the frequency-dependent dielectric constant may be written137 as

4 π e2 ρind (q, ω)
ε(q, ω) = 1 −
q2 δV (q, ω)
4 π e2
= 1 − χ(q, ω) (1299)
q2
where we have used the definition of the frequency-dependent response function
χ(q, ω). The frequency-dependent response function is defined by

ρind (q, ω)
χ(q, ω) = (1300)
δV (q, ω)

The real space and time form of the linear response relation can be found by
re-writing this relation as

ρind (q, ω) = χ(q, ω) δV (q, ω) (1301)


137 H. Ehrenreich and M. H. Cohen, Phys. Rev. 115, 786 (1959).

431
and then performing the inverse Fourier transform. The real space and time
form of the linear response relation is in the form of a convolution
Z Z ∞
ρind (r, t) = d3 r 0 dt0 χ(r − r0 ; t − t0 ) δV (r0 , t0 ) (1302)
−∞

The dependence of the response function on r − r0 is a direct consequence of


our assumption that space is homogeneous. As the response function relates
the cause and effect in a linear fashion, the response function can be calculated
perturbatively. The induced electron density is found, in real space and time, by
treating the time-dependent potential as a perturbation. The resulting causal,
non-local relation is then Fourier transformed with respect to space and time.
This procedure is a generalization of our previous treatment of static screening.

The expectation value of the electron density operator ρ̂(r) at time t, is cal-
culated in a state that has evolved from the ground states due to the interaction.
The electron density operator is given by
X  
3
ρ̂(r) = δ r − ri (1303)
i

and the time-dependent perturbation is


Z
Ĥint (t) = d3 r0 ρ̂(r0 , t) δV (r0 , t) (1304)

The expectation value of the electron density is to be evaluated in the interaction


representation. The expectation value of the density is given by
ρ(r, t) = < Ψint (t) | ρ̂int (r, t) | Ψint (t) > (1305)
where the state and operators are expressed in the interaction representation. In
the interaction representation, the operators evolve with respect to time under
the influence of the unperturbed Hamiltonian Ĥ0 , and are given by
   
it it
ρ̂int (r, t) = exp + Ĥ0 ρ̂(r) exp − Ĥ0 (1306)
h̄ h̄
In the interaction representation, the state evolves under the influence of the
interaction Ĥint (t). To first order in the perturbation, the ground state is given
by
" Z t #
i 0 0
| Ψint (t) > = 1 − dt Ĥint (t ) + . . . | Ψ0 > (1307)
h̄ −∞

where | Ψ0 > is the initial ground state eigenfunction of Ĥ0 . The induced
electron density is defined as
ρind (r, t) = < Ψint (t) | ρ̂int (r, t) | Ψint (t) > − < Ψ0 | ρ̂int (r, t) | Ψ0 >
(1308)

432
The second term is time independent, as it is the expectation value in the ground
state of the time independent Hamiltonian Ĥ0 . On substituting the expression
for the perturbed wave function, one finds a linear relationship between the
induced density and the perturbing potential
Z t  
i 0 0
ρind (r, t) = − dt < Ψ0 | ρ̂int (r, t) , Hint (t ) | Ψ0 >
h̄ −∞
Z t Z  
i
= − dt0 d3 r0 < Ψ0 | ρ̂int (r, t) , ρ̂int (r0 , t0 ) | Ψ0 > δV (r0 , t0 )
h̄ −∞
Z +∞ Z
0
= dt d3 r0 χ(r, r0 ; t − t0 ) δV (r0 , t0 ) (1309)
−∞

This is a causal relation in which the response function is identified as


 
0 0 i
χ(r, r ; t − t ) = − < Ψ0 | ρ̂int (r, t) , ρ̂int (r , t ) | Ψ0 > Θ( t − t0 )
0 0

(1310)
where Θ(t) is the Heaviside step function. Thus, the response function is a
two time correlation function which involves the ground state expectation value
of the commutator of the density operators at different positions, and differ-
ent times. Due to the time homogeneity of the ground state, the correlation
function only depends on the difference of the two times. For a spatially ho-
mogeneous system, the correlation function only depends on the difference r−r0 .

The expression can be evaluated by using the completeness relation


X
| Ψn > < Ψn | = Iˆ (1311)
n

On inserting a complete set of states between the density operators, one obtains
"
0 0 i X
χ(r, r ; t − t ) = − < Ψ0 | ρ̂int (r, t) | Ψn > < Ψn | ρ̂int (r0 , t0 ) | Ψ0 >
h̄ n
#
− < Ψ0 | ρ̂int (r0 , t0 ) | Ψn > < Ψn | ρ̂int (r, t) | Ψ0 > Θ( t − t0 )

(1312)

On expressing the time dependence of the operators in terms of the eigenvalues


of the unperturbed Hamiltonian, Ĥ0 , the response function reduces to
 
i X i
= − exp + (t − t )(E0 − En ) < Ψ0 | ρ̂(r) | Ψn > < Ψn | ρ̂(r0 ) | Ψ0 >
0
h̄ n h̄
 
i X i
+ exp − (t − t )(E0 − En ) < Ψ0 | ρ̂(r0 ) | Ψn > < Ψn | ρ̂(r) | Ψ0 >
0
h̄ n h̄
(1313)

433
for t − t0 > 0, and is zero otherwise. In the above expression for the response
function, the density operators are no longer time-dependent.

Up to this point, our analysis has been completely general. To illustrate the
structure of the response function, we shall now make the assumption that the
electrons are non-interacting. The ground state | Ψ0 > and the excited states
| Ψn > can be represented by single Slater determinants composed of the set of
one-electron energy eigenfunctions {φαj (rj ); j ∈ 1, 2, . . . Ne } and {φβj (rj ); j ∈
1, 2, . . . Ne }, respectively. The matrix elements of the one-electron operator
ρ̂(r) are non-zero only if the set of quantum numbers {αj ; j ∈ 1, 2, . . . Ne } and
{βj ; j ∈ 1, 2, . . . Ne } only differ by at most one element, say the i-th value. Thus,
we may permute the indices in the set βj until one has

αi 6= βi (1314)

and
αj = βj ∀ j 6= i (1315)
In this case, the matrix elements < Ψ0 | ρ̂(r) | Ψn > are trivially evaluated
as
Z
< Ψ0 | ρ̂(r) | Ψn > = d3 ri φ∗αi (ri ) δ 3 ( r − ri ) φβi (ri )

= φ∗αi (r) φβi (r) (1316)

The matrix element is only non-zero if the spin state of α is identical to the spin
state of β, so the spin quantum number is conserved. In the above expression,
the single-electron state αi is occupied in the initial state | Ψ0 > and unoccu-
pied in the final state | Ψn > while the single-electron state βi is unoccupied in
the initial state | Ψ0 > and occupied in the final state | Ψn > . All the other
single-electron quantum numbers in | Ψ0 > and | Ψn > are unchanged, i.e.,
αj = βj for ∀ j 6= i. Furthermore, the Pauli exclusion principle requires that
βi 6= βj . This shows that the final states of the non-interacting many-electron
system are obtained by exciting a single electron from the state αi to the state
βi . For non-interacting electrons, the excitation energy En − E0 is simply
given by the difference in the single-electron energy eigenvalues

En − E0 = Eβi − Eαi (1317)

Thus, the response function is simply given by


 
0 i X i
χ(r, r ; t) = − exp + t ( Eα − Eβ ) φ∗α (r) φβ (r) φ∗β (r0 ) φα (r0 )
h̄ h̄
α,β
 
i X i
+ exp − t ( Eα − Eβ ) φα (r) φ∗β (r) φβ (r0 ) φ∗α (r0 )
h̄ h̄
α,β
(1318)

434
for t > 0. The sum over α is restricted to run over the single-particle quantum
numbers that are occupied in the ground state, and the sum over β runs over the
quantum numbers that are unoccupied in the ground state. The spin quantum
number is conserved, that is, σα = σβ .

On evaluating the response function for free electrons, summing over spin
states and using the Bloch state energy eigenvalues, one finds
   
2i X X it i 0 0
= − exp + (Ek − Ek0 ) exp − (k − k ) . (r − r )
h̄ V 2 h̄ h̄
|k|<kF |k0 |>kF
   
2i X X it i 0 0
+ exp − (E k − E k 0 ) exp + (k − k ) . (r − r )
h̄ V 2 0
h̄ h̄
|k|<kF |k |>kF

(1319)

for t > 0. Since the free electron gas is homogeneous, the response function
only depends on the distance between the perturbation and the response r − r0 .
On Fourier transforming the response function with respect to space and time,
one obtains χ(q, ω) as
Z +∞ Z  
χ(q, ω) = dt d3 r exp − i(q.r − ωt) χ(r; t) (1320)
−∞

Since the response function contains the Heaviside step function Θ(t), the inte-
gral over t can be evaluated in the interval ∞ > t ≥ 0. The integral over t
converges faster if ω is analytically continued into the upper-half complex plane
to ω → z = ω + i η. The factor of exp [ − η t ] damps out the oscillations
in the integrand as t → ∞. Thus, in the (q, ω) domain, one finds that the
response function is complex and is given by the expression
" #
2 X 1
χ(q, ω + iη) =
V h̄ ω + i η + Ek − Ek+q
|k|<kF |k+q|>kF
" #
2 X 1

V h̄ ω + i η + Ek − Ek+q
|k|>kF |k+q|<kF

(1321)

The restrictions on the summation over k can be simplified. To see this, we


shall introduce a function fk which behaves like the T → 0 limit of the Fermi
function. The function is defined by

fk = 1 for Ek < EF (1322)

and
fk = 0 for Ek > EF (1323)

435
15

2kFq+q2
10

2mω/ kF2h
Ek+q-Ek
5

-2kFq+q2
0
0 1 2 3 4
2kFq-q2
Ek-Ek-q
-2kFq-q2
-5
q/kF

Figure 202: The region of (ω, q) space available for single-particle excitations,
subject to the constraint that the initial state is inside the Fermi sphere. The
region where the first delta function of eqn(1325) is non-zero is bounded by the
two blue lines.

The response function can then be written as the sum over all k as
" #
2 X fk ( 1 − fk+q )
χ(q, ω + iη) =
V h̄ ω + i η + Ek − Ek+q
k
" #
2 X fk+q ( 1 − fk )

V h̄ ω + i η + Ek − Ek+q
k
" #
2 X fk − fk+q
= (1324)
V h̄ ω + i η + Ek − Ek+q
k

In the last line, it is seen that the factors which explicitly enforce the Pauli-
exclusion principle cancel. For ω just above the real axis, i.e in the limit η → 0,
the imaginary part of the response function is found as
 
2π X
=m χ(q, ω + iη) = − δ h̄ ω + Ek − Ek+q
V
|k|<kF
 
2π X
+ δ h̄ ω + Ek − Ek+q (1325)
V
|k+q|<kF

From this analysis, one can see that for positive ω, the imaginary part of χ(q, ω)
is non-zero in the region of (ω, q) phase space, where
h̄ h̄
( − 2 kF q + q 2 ) < ω < ( + 2 kF q + q 2 ) (1326)
2m 2m

436
It is only in this region that the argument of the first delta function in =m χ(q, ω)
has a solution
h̄ h̄
k.q = ω − q2 (1327)
m 2m
with k < kF . These conditions divide (q, ω) space into non-overlapping regions.

For completeness, the complete expressions for the real and imaginary parts
of the Lindhard dielectric function at finite frequencies are given138 . The real
part is given by
" (
kT2 F (2 m ω − h̄ q 2 )2
  
kF
<e ε(q, ω) = 1 + 1 + 1 − ×
2 q2 2q 4 h̄2 q 2 kF2

2 m ω − 2 h̄ q k − h̄ q 2
F
× ln

2 m ω + 2 h̄ q kF − h̄ q 2

 )#
(2 m ω + h̄ q 2 )2 2 m ω + 2 h̄ q kF + h̄ q 2

+ 1 − ln
4 h̄2 q 2 kF2 2 m ω − 2 h̄ q kF + h̄ q 2

(1328)
and the imaginary part is given by
π kT2 F m ω
 
=m ε(q, ω + iη) = 2 m ω < 2 h̄ q kF − h̄ q 2
2 q 2 h̄ q kF
(1329)
" #
π kT2 F kF (2 m ω − h̄ q 2 )2
 
=m ε(q, ω + iη) = 1 −
4 q2 q 4 h̄2 q 2 kF2
2 h̄ q kF − h̄ q 2 < 2 m ω < 2 h̄ q kF + h̄ q 2 (1330)
and
 
=m ε(q, ω + iη) = 0 2 h̄ q kF + h̄ q 2 < 2 m ω (1331)

The real part is an even function of ω and the imaginary part is an odd function
of ω. For ω = 0, the response function reduces to the real static response
function calculated previously. For | ω | > 2 h̄m ( 2 kF q + q 2 ), the imaginary
part of the function vanishes, since the denominator never vanishes for any k
value in the range of integration. In this region of q and ω, there are no poles,
therefore, the real part of the response function χ(q, ω) can be expanded in
powers of q 2 . To the order of q 4 , one finds
" 2 #
kF3 q2

3 h̄ kF q
<e χ(q, ω) = + 1 + + ... (1332)
3 π2 m ω2 5 mω
138 J. Lindhard, Kgl. Danske Videnskab. Selskab. Mat.-Fys. Medd. 28, 8 (1954).

437
80

q/kF = 0.3
40
ε(q,ω)

0
-4 -3 -2 -1 0 1 2 3 4

-40

-80
2mω/hkF2

Figure 203: The frequency dependence of the real (blue) and imaginary (red)
parts of the (R.P.A.) dielectric constant for a q value such that q < 2kF .

1.5

q/kF = 2.5
1
ε(q,ω)

0.5

0
-20 -10 0 10 20

-0.5
2mω/hkF2
Figure 204: The frequency dependence of the real (blue) and imaginary (red)
parts of the (R.P.A.) dielectric constant for a q value such that q > 2kF .

438
Thus, for high frequencies such that ω  q h̄ mkF the dielectric constant can
be approximated by
" 2 #
4 π ρ e2

3 h̄ kF q
ε(q, ω) = 1 − 1 + + ...
m ω2 5 mω
" 2 #
ωp2

3 h̄ kF q
= 1 − 2 1 + + ... (1333)
ω 5 mω

where the expression for the electron density ρ

kF3
ρ = 2 (1334)
6 π2
has been used, and the plasmon frequency ωp has been defined via

4 π ρ e2
ωp2 = (1335)
m
Thus, the dielectric constant has zeros at the frequencies ω = ωp (q), where
"  2 #
2 2 3 h̄ kF q
ωp (q) = ωp 1 + + ... (1336)
5 m ωp (q)

The finite value of the frequency ωp (0), is a direct consequence of the long-
ranged nature of the Coulomb interaction. If the external potential is zero,
φext (q, ωp (q)) = 0, and the total potential is non-zero φ(q, ωp (q)) 6= 0, then the
real and imaginary parts of the dielectric constant must vanish, ε(q, ωp (q)) = 0,
as

ε(q, ωp (q)) φ(q, ωp (q)) = φext (q, ωp (q))


ε(q, ωp (q)) φ(q, ωp (q)) = 0 (1337)

In this case, when the total potential inside the solid φ(q, ωp (q)) is non-zero,
the induced density and current fluctuations must be finite. These longitudinal
collective charge oscillations excitations are plasmons. A typical energy range
for the plasmon energy, h̄ ωp , in metals, ranges from the low values of 3.72 eV
found in K, 5.71 eV found in N a, to values as high as 15.8 eV found in Al. The
dielectric materials Si, Ge etc. also have plasmon energies of the order 16 eV.

One may inquire as to the nature of the excitations at larger q values, such
ω
that the phase velocity of the plasmons qp becomes smaller than the Fermi
h̄ kF
velocity vF = m . At a critical value of q the denominator of the response
function may vanish so the response function acquires a sizeable imaginary part.
The plasmon excitations merge with a continuum of particle hole excitations
which have excitation energies given by

h̄ ω(q, k) = Ek+q − Ek (1338)

439
15

2kFq+q2
10
Single-particle
excitations

2mω/ kF2h
5
Damped
2mωp(q)/kF2h Plasmons
-2kFq+q2
0
0 1 2 3 4
2kFq-q2
2
-2kFq-q
-5
q/kF

Figure 205: The phase space of collective (plasmon) excitations and single-
particle excitations. The branch of plasmon excitations becomes damped as it
merges with the continuum of electron-hole excitations.

2
for k < kF . The edges of the continuum stretch from 2h̄m ( 2 kF q + q 2 ) to
h̄2 2
2 m ( − 2 kF q + q ). When the plasmon merges into the continuum, it un-
dergoes significant broadening. This sort of damping is called Landau damping.
Landau damping can also be viewed classically, in terms of electrons surf-riding
the waves in the potential field. Imagine that a wave with phase velocity ωq is
propagating through an electron gas, and consider the electrons with velocity
almost parallel and close to the phase velocity of the wave. In the frame of
reference travelling with the wave, the electron is at rest and experiences an es-
sentially time independent electric field. The electric field continuously transfers
energy from the wave to the electrons that have the same velocity. If there is
a slight mismatch in the velocities, electrons with lower velocity than the wave
draw energy from the wave and accelerate, whereas electrons that are moving
faster lose energy and slow down. This has the consequence that the rate of
energy loss of the wave is proportional to the derivative of the electron velocity
distribution, evaluated at the wave’s phase velocity.

11.3.2 Electron Scattering Experiments


The longitudinal excitations of the electrons in a metal can be probed by the
scattering of a beam of charged particles or fast electrons. The coupling takes
place via the Coulomb interaction
X e2
Ĥint = (1339)
i
| r − ri |

440
50

40
q/kF = 1.3

Im[ ω/ε(q,ω) ]
30

20

10

0
3 3.5 4 4.5 5
2mω/hkF2

Figure 206: The Landau damped plasmon pole in the inverse of the (R.P.A.)
dielectric constant.

e
Detector
θ k'

e
Thin Film

Figure 207: A beam of high-energy electrons scatters from a thin metal film.

where ri labels the positions of the electrons in the plasma and r is the posi-
tion of the incoming high energy electron. If the incident beam is composed
of electrons which have high energies, the beam electrons can be considered to
be as classical and are, therefore, distinguishable. This ignores the possibility
of exchange interactions with the electrons in the metal. Analysis of the Mott
scattering formula for electrons also shows that the neglect of the exchange
scattering is an excellent approximation for scattering through small angle scat-
tering. Therefore, we shall consider the charged particles in the beam as being
distinguishable from the electrons in the solid.

The rate at which a charged particle is scattered inelastically from state k

441
with energy E(k) to state k 0 with energy E(k 0 ) is given by
2  
1 2 π 0
0
= < k Ψn | Ĥint | k Ψ0 > δ En + E(k ) − E0 − E(k)

τ (k → k 0 ) h̄
(1340)
where | Ψ0 > and E0 are the ground state wave function and ground state
energy of the solid. The final state wave function and energy is given by | Ψn >
and En . The momentum and energy loss of the charged particle are defined to
be
h̄ q = h̄ k − h̄ k 0
h̄ ω = E(k) − E(k 0 ) (1341)
On performing the integral over the position of the fast charged particle, one
has
2 2
2 π 4 π e2

1
< Ψn |
X
0 = 2
exp[ i q . r i ] | Ψ0 > ×
τ (k → k ) h̄ q V
i

 
δ h̄ ω + E0 − En

(1342)
The energy conserving delta function can be replaced by an integral over time
by using the identity
  Z ∞  
dt t
δ h̄ ω + E0 − En = exp[ i ω t ] exp i ( E0 − En )
−∞ 2 π h̄ h̄
(1343)
The energy eigenvalues in the exponential time evolution factors can be replaced
by the general time evolution operators involving the unperturbed Hamiltonian
operator Ĥ0 ,
2 Z ∞
2 π 4 π e2

1 dt X
0 = 2 2
exp[ i ω t ] < Ψn | exp[ i q . ri ] | Ψ0 > ×
τ (k → k ) h̄ q V −∞ 2 π i
X Ĥ0 t Ĥ0 t
< Ψ0 | exp[ i ] exp[ − i q . ri ] exp[ − i ] | Ψn >
i
h̄ h̄
(1344)
The factor involving ri can be expressed as the Fourier transform of the electron
density operator
1 X 1
Z X  
exp[ − i q . ri ] = d3 r exp[ − i q . r ] δ 3 r − ri
V i V i
Z
1
= d3 r exp[ − i q . r ] ρ̂(r)
V
= ρ̂q (1345)

442
Thus, on combining the above expressions, the inelastic scattering rate is found
as
2 Z ∞
2 π 4 π e2

1 dt
0 = 2 2
exp[ i ω t ] < Ψn | ρ̂−q | Ψ0 > ×
τ (k → k ) h̄ q −∞ 2 π
Ĥ0 t Ĥ0 t
< Ψ0 | exp[ i ] ρ̂q exp[ − i ] | Ψn >
h̄ h̄
2 Z ∞
4 π e2

2π dt
= exp[ i ω t ] < Ψ0 | ρ̂q (t) | Ψn > < Ψn | ρ̂−q (0) | Ψ0 >
h̄2 q2 −∞ 2π
(1346)

where the density operator is evaluated in the interaction representation. If the


final state of the solid | Ψn > is not measured, there is a distribution of possible
final states of the solid. If only the final state of the charged particle is measured,
and the final state of the solid is not measured, the index n corresponding to
the different possible final states must be summed over

1 X 2 π  4 π e2 2 Z ∞ dt
= exp[ i ω t ] ×
τ (k → k 0 ) n
h̄2 q2 −∞ 2 π

< Ψ0 | ρ̂q (t) | Ψn > < Ψn | ρ̂−q (0) | Ψ0 >


(1347)

The sum over the final states can be evaluated using the completeness relation
which leads to the result
2 Z ∞
2 π 4 π e2

1 dt
= exp[ i ω t ] < Ψ0 | ρ̂q (t) ρ̂−q (0) | Ψ0 >
τ (k → k 0 ) h̄2 q2 −∞ 2 π
(1348)

The factor of q −4 shows the scattering process is dominated by small momentum


transfers. The density-density correlation function, S(q, ω), is defined via
Z ∞
2 dt
S(q, ω) = V exp[ i ω t ] < Ψ0 | ρ̂q (t) ρ̂−q (0) | Ψ0 > (1349)
−∞ 2 π h̄
On substituting this relation into the scattering rate, we obtain the result
2
2 π 4 π e2

1
= S(q, ω) (1350)
τ (k → k 0 ) h̄ q2 V
Thus, it is seen that the long wavelength electron density fluctuations are mainly
responsible for scattering the incident charged particle. For non-interacting
electrons, S(q, ω) is evaluated as
X  
S(q, ω) = 2 δ h̄ ω + Ek − Ek+q (1351)
|k|<kF |k+q|>kF

443
where the summation over k is over the filled Fermi sphere, subject to the re-
striction that the final state be allowed by the Pauli exclusion principle.

The inelastic scattering cross-section can be evaluated in terms of the scat-


tering rate, and is found to be given by the expression
2
d2 σ k0 2 mq e2

= h̄ S(q, ω) (1352)
dΩ dω k h̄2 q 2
where mq is the mass of the charged particle. Thus, in the Born approximation,
the scattering cross-section is directly related to the density-density correlation
function. This type of correlation function was first introduced by van Hove in
the context of neutron scattering139 . Since the Coulomb interaction is relatively
strong, it is frequently necessary to incorporate the effects of multiple-scattering
in order to obtain agreement with experimental data.

The Fluctuation-Dissipation Theorem140 relates the spectrum of electron


density fluctuations to the imaginary part of the dielectric constant. At finite
temperatures, this relation has the form
q2 V
   
1 1
S(q, ω) = =m (1353)
4 π 2 e2 exp[ − β h̄ ω ] − 1 ε(q, ω + iη)
The relation between S(q, ω) and the inverse dielectric constant can be seen
through the following classical argument. The power, per unit volume, dissi-
pated by the electromagnetic field of the charged particle is given by
1 ∂D
P (r, t) = E. (1354)
4π ∂t
For a negatively charged particle travelling with velocity v(t), the displacement
field D(r, t) is the experimentally controllable quantity and is given by the
expression " #
|e|
D(r, t) = − ∇ − (1355)
| r − r(t) |
On Fourier transforming D(r, t) with respect to space and time, one finds
D(q, ω). However, D(q, ω) is related to the Fourier transform of the electric
field E(q, ω) via a factor of the dielectric constant
D(q, ω)
E(q, ω) = (1356)
ε(q, ω + iη)
On Fourier transforming the expression for the power density, P (r, t), with
respect to r and t, one finds P (q, ω) to be given by
" # 2
ω 1
P (q, ω) = − =m D(q, ω)
8π ε(q, ω + iη)
139 L. van Hove, Phys. Rev. 95, 249 (1954).
140 H. B. Callen and T. A. Welton, Phys. Rev. 83, 34 (1951).

444
" # 2
ω =m ε(q, ω + iη)
= D(q, ω)
8π | ε(q, ω + iη) |2
" # 2
ω =m ε(q, ω + iη)
= D(q, ω)
8π ( <e ε(q, ω) )2 + ( =m ε(q, ω + iη) )2

(1357)

This result implies that the zero in the real part of ε(q, ω) should show up as a
delta function peak in the power loss.

——————————————————————————————————

11.3.3 Exercise 64
Use linear response theory to express the change in the electron density induced
by an external charge. Hence, express the inverse of the dielectric constant in
terms of the exact eigenstates and energy eigenvalues of the interacting many-
electron system. Use the resulting expression to find the T = 0 form of the
fluctuation-dissipation theorem141 .

——————————————————————————————————

Solution 64

In the Coulomb gauge, the Fourier transform of the external charge density
ρext (r, t) is related to the external potential via Poisson’s theorem

− q 2 φext (q, ω) = − 4 π | e | ρext (q, ω) (1358)

The total field is related to the external charge density and the induced charge
density via

− q 2 φ(q, ω) = − 4 π | e | ρext (q, ω) − 4 π | e | ρind (q, ω) (1359)

The dielectric constant is defined as


1 φ(q, ω)
= (1360)
ε(q, ω) φext (q, ω)

which can be expressed as

1 ρind (q, ω)
= 1 + (1361)
ε(q, ω) ρext (q, ω)
141 P. Noziêres and D. Pines, Nuovo Cimento, 9, 470 (1958).

445
Hence, the linear response relation can be expressed as
 
1
ρind (q, ω) = − 1 ρext (q, ω)
ε(q, ω)
q2
 
1
= − 1 φext (q, ω)
ε(q, ω) 4π|e|
q2
 
1
= 1 − δV (q, ω) (1362)
ε(q, ω) 4 π e2
However, the interaction operator is given by
Z
Ĥint (r, t) = d3 r ρ̂(r) δV (r, t) (1363)

where the electron density is given by


X
ρ̂(r) = δ 3 (r − ri ) (1364)
i

The induced electron density is evaluated using linear perturbation theory, in


the interaction representation. In the absence of the perturbation, the ground
state is denoted by | Ψ0 > . The perturbation is turned on adiabatically, and
the ground state evolves to the state | Φ0 (t) > which, to first order in the
interaction, is given by
i t
 Z 
| Φ0 (t) > = 1 − dt0 Ĥint (t0 ) | Ψ0 > (1365)
h̄ −∞
The induced electron density ρind (r, t) is then given by the expectation value of
the commutator
Z t
i
ρind (r, t) = − dt0 < Ψ0 | [ ρ̂(r, t) , Ĥint (t0 ) ] | Ψ0 >
h̄ −∞
Z t Z
i 0
= − dt d3 r0 < Ψ0 | [ ρ̂(r, t) , ρ̂(r0 , t0 ) ] | Ψ0 > δV (r0 , t0 )
h̄ −∞
Z ∞ Z
0
= dt d3 r0 χ(r − r0 , t − t0 ) δV (r0 , t0 ) (1366)
−∞

The material is assumed to be homogeneous, therefore, the response function


is only a function of the spatial separation r − r0 . Furthermore, since Ĥ0 is
independent of time, the response function is only a function of t − t0 . On
Fourier transforming this equation with respect to space and time, one finds
ρind (q, ω) = χ(q, ω) δV (q, ω) (1367)
where
X  | < Ψ0 | ρ̂q | Ψn > |2
χ(q, ω) = V
n
h̄ ω + i η + En − E0
| < Ψ0 | ρ̂q | Ψn > |2

+ (1368)
− h̄ ω − i η + En − E0

446
The imaginary part of the response function is found as
X  
2
=m χ(q, ω) = − π V | < Ψ0 | ρ̂q | Ψn > | δ( h̄ ω + En − E0 ) − δ( h̄ ω − En + E0 )
n
(1369)
Thus, the zero temperature limit of the fluctuation-dissipation theorem has the
form
4 π 2 e2 V X
 
1 2
=m = | < Ψ0 | ρ̂ q | Ψn > | δ(h̄ω + E n − E 0 ) − δ(h̄ω − E n + E 0 )
ε(q, ω) q2 n
4 π 2 e2
 
= S(q, −ω) − S(q, ω) (1370)
q2 V
The first term is only non-zero if 0 > ω, and the second term is only non-zero
in the range ω > 0.

——————————————————————————————————

11.3.4 Exercise 65
Show, using classical electromagnetic theory, that the power loss spectrum of a
particle with charge e moving with velocity v due to plasmons, can be expressed
as
2 e2
 
ω q0 v
P (ω) = − =m ln (1371)
πv ε(ω + iη) ω
Assume that the dielectric constant is independent of q, for q < q0 where h̄ q0
is the maximum momentum transfer.

——————————————————————————————————

Solution 65

The average power P dissipated by the charged particle can be expressed as


the limit τ → ∞
Z ∞  
1 t
P = dt P (t) exp −
τ 0 τ
Z ∞ Z  
1 t
= dt d3 r P (r, t) exp −
τ 0 τ
Z ∞ Z  
1 ∂ t
= dt d3 r E(r, t) D(r, t) exp −
4πτ 0 ∂t τ
(1372)

where we have inserted an exponential convergence factor. The convergence


factor will be absorbed in the displacement and electric fields. The Fourier

447
transform is expressed as
Z Z ∞  
1
D(q, ω) = d3 r dt D(r, t) exp − i ( q . r − ω t ) (1373)
V −∞

and the inverse Fourier transformation is given by


Z Z  
V 3 dω
D(r, t) = d q D(q, ω) exp + i ( q . r − ω t ) (1374)
( 2 π )3 2π

On inserting the expressions for the inverse Fourier transforms into the expres-
sion for the average power loss, one finds
Z ∞
i V2 d3 q
Z

P = − E(−q, −ω) ω D(q, ω)
2 τ ( 2 π ) −∞ 2 π ( 2 π )3
Z ∞
i V2 d3 q D(q, ω)
Z

= ω D∗ (q, ω)
2 τ ( 2 π ) −∞ 2 π ( 2 π ) ε(q, ω + 2τi )
3
Z ∞
i V2 d3 q
Z
dω ω
= | D(q, ω) |2(1375)
2 τ ( 2 π ) −∞ 2 π ( 2 π ) ε(q, ω + 2τi )
3

in the limit τ → ∞. The Fourier transform of the displacement field is given


by
Z ∞  
|e|
Z
1 t
D(q, ω) = d3 r dt exp − i ( q . r − ω t ) − ∇
V 0 2 τ | r − vt|
Z ∞
4πiq|e|
 
t
= − dt exp i ( ω − q . v ) t −
V q2 0 2 τ
4πiq|e| 2τ
= − (1376)
V q2 1 − i(ω − q.v)2τ

Hence,
∞ d3 q
4 e2
Z Z
iω 1
P = dω i 1
2 τ ( 2 π )3 −∞
2
q ε(q, ω + 2τ ) 4 τ 2
+ ( ω − q . v )2
(1377)
which in the limit τ → ∞ reduces to
Z ∞
2 e2 d3 q
Z

P = dω δ( ω − q . v ) (1378)
( 2 π )2 −∞ q 2 ε(q, ω + iη)

This yields the expression


Z ∞
2 e2
Z  
dq iω
P = dω θ( ω + q v ) − θ( ω − q v )
( 2 π ) −∞ q v ε(q, ω + iη)
(1379)

448
Hence, on assuming that the dielectric constant is independent of q, from q = 0
to an upper cut off q = q0 , one obtains the result
Z ∞
e2 iω q0 v
P = − dω ln (1380)
π v −∞ ε(ω + iη) ω
The power loss spectrum, P (ω), is defined in terms of an integral over positive
frequencies Z ∞
P = dω P (ω) (1381)
0
On using the symmetry properties of the dielectric constant under the transfor-
mation ω → − ω, one finds that the contribution from the real part of the
inverse dielectric constant vanishes. Hence, one obtains the final result for P (ω)
2 e2
 
ω q0 v
P (ω) = − =m ln (1382)
πv ε(ω + iη) ω

——————————————————————————————————

In a typical experiment, monochromatic electron beams with energies E(k)


of the order of keV fall incident on thin films, and the energies of the scattered
electrons, E(k 0 ) are analyzed142 . Experimentally, it is found that the fast elec-
tron loses energy in almost exact multiples of h̄ ωp . That is, the energy loss
spectrum shows peaks separated by energies which are multiples of h̄ ωp . The
above analysis predicts a single pole near ω = ωp . The discrepancy is caused
by the use of the Born Approximation which neglects the effects of multiple
scattering. The experiments are usually analyzed by fitting the intensities of
the peaks to a Poisson distribution
 n  
1 L L
In = exp − I0 (1383)
n! λ λ
where In is the intensity of the n-th plasmon peak, L is the sample thickness
and λ is the mean free path. The mean free path is then compared with the
theoretically derived inelastic scattering cross-section.

The mean free path can be estimated from the scattering cross-section. On
expressing the density-density correlation function in terms of the imaginary
part of the inverse of the dielectric constant, one finds
2 " #
d2 σ k0

mq e 1
= − h̄ V =m (1384)
dΩ dω k π h̄2 q ε(q, ω + iη)

For the frequency range


2 m ω > 2 h̄ q kF + q 2 (1385)
142 L. Marton, J. A. Simpson, H. A. Fowler and N. Swanson, Phys. Rev. 126, 182 (1962).

449
Figure 208: The energy loss spectrum, in units of h̄ωp , for electrons of energy
20 keV falling on an aluminum foil of thickness 2500 Å. [After Marton et al.
(1962).]

the imaginary part of χ(q, ω + iη) vanishes as η → 0. Therefore, in this limit,


the imaginary part of the dielectric constant also vanishes
=m ε(q, ω + iη) → κ η (1386)
for some value finite of κ. Hence, in this limit, one has
" #
1
=m = − π δ( <e ε(q, ω) ) (1387)
ε(q, ω + iη)

Furthermore, in this region of (ω, q) space, one has the approximate expression

ωp2 (q)
ε(q, ω) = 1 − (1388)
ω2
where the plasmon dispersion relation is given by
 
2 2 3 2 2
ωp (q) = ωp + q vF + . . . (1389)
5
and the plasmon frequency by
4 π ρ e2
ωp2 = (1390)
m
Thus, one finds the single (plasmon) pole approximation for the inverse dielectric
constant
" #  2
ω − ωp2 (q)

1
=m = −πδ
ε(q, ω + iη) ω2

450
ω2
= −π δ( ω − ωp (q) )
ω + ωp (q)
π
= − ωp (q) δ( ω − ωp (q) ) (1391)
2
for positive ω. The plasmon contribution to the differential scattering cross-
section, is found by integrating over the energy loss ω, and is given by
 2 2
dσ k 0 2 m2q ρ V e
= (1392)
dΩ k m h̄ ωp h̄ q
Thus, the scattering cross-section is directly proportional to the number of elec-
trons in the solid. In deriving the differential scattering cross-section, we have
neglected the q dependence of the plasmon dispersion relation. The total plas-
mon scattering cross-section is found by integrating the differential cross-section
over the scattering angle θ. We note that energy and momentum conservation
leads to the two conditions

q2 = k 2 + k 02 − 2 k k 0 cos θ
θ
q2 = ( k − k 0 )2 + 4 k k 0 sin2 (1393)
2
and
2 mq
( k − k0 ) ( k + k0 ) = ωp

2 mq ω p
( k − k0 ) = (1394)
h̄ ( k + k 0 )
For small scattering angles, θ  1, these can be combined to yield

h̄2 ωp2
 
2 2 2 θ
q ≈ k 4 sin +
2 4 E(k)2
 
≈ k 2 θ2 + θ02 (1395)

Hence, one has

dσ k 0 2 m2q ρ V e4

dΩ k m h̄ ωp h̄2 k 2 ( θ2 + θ02 )
mq e4 ρ V
≈ (1396)
m h̄ ωp E(k) ( θ2 + θ02 )
On integrating over the solid angle dΩ, but restricting the range of θ from zero
to a maximum momentum transfer given by 2 kF ∼ θm k, one finds the total
cross-section, σ, for plasmon scattering, is given by
e4 ρ V mq θm
σ ≈ 2π ln (1397)
h̄ ωp E(k) m θ0

451
The mean free path, λ, is then found by noting that a trajectory of cross-
sectional area σ covers a volume λ σ between consecutive collisions. This must
equal V , the volume of the solid. This leads to the mean free path being given
by
e4 ρ mq θm
λ−1 ≈ 2 π ln (1398)
h̄ ωp E(k) m θ0
Thus, the mean free path depends linearly on the kinetic energy of the incident
electron. This value has been found to track the mean free path obtained by
fitting the measured intensities of the multi-plasmon peaks.

11.3.5 The Transverse Response


In the Coulomb or radiation gauge, the vector potential describes the transverse
electromagnetic field. It satisfies the equation

ω2
 

− q2 + 2 A(q, ω) = − j (q, ω) (1399)
c c T

The situation in which there are no transverse external currents impressed on


the system, j T ext (q, ω) = 0, is considered. Thus, one obtains the microscopic
equation
ω2
 
2 4π
− q + 2 A(q, ω) = − j (q, ω) (1400)
c c T ind
Ohm’s law can be expressed in the form

jT ind
(q, ω) = σT (q, ω) ET (q, ω) (1401)

where σT is the transverse conductivity, and the total transverse electric field is
given by
ω
ET (q, ω) = i A(q, ω) (1402)
c
This leads to
ω2
 
4πω
− q2 + 2 + i σ T (q, ω) A(q, ω) = 0 (1403)
c c2

The transverse dielectric function is identified in terms of the optical conduc-


tivity
4πi
εT (q, ω) = 1 + σT (q, ω) (1404)
ω
The photon dispersion relation can be re-written as
 2
ω
εT (q, ω) = q 2 (1405)
c

If εT (q, ω) > 0, then there are undamped transverse electromagnetic waves.

452
ω > ωp
ε0 = 1 ε(ω) > 0

-3 -2 -1 0 1 2
z

z>0

Figure 209: The spatial dependence of a transverse electromagnetic field falling


incident on a metal, for ω > ωp . The electromagnetic radiation is transmitted
through the metal.

ε0 = 1 ω < ωp
ε(ω) < 0

-3 -2 -1 0 z 1 2

z>0

Figure 210: The spatial dependence of a transverse electromagnetic field falling


incident on a metal, for ω < ωp . The electromagnetic radiation is attenuated
by the metal.

453
√(ωp2 + c2q2)

cq

ω
ωp

Figure 211: The energy dispersion for transverse electromagnetic radiation in a


metal. The dispersion relation starts off at the plasmon frequency ωp at q = 0
and then merges with the dispersion relation of light.

Otherwise, q would be complex which implies that E T (r; t) is attenuated as it


enters into the sample. In other words, if =m ε(q, ω) = 0 and <e ε(q, ω) > 0,
the material is transparent to transverse electromagnetic waves. The dispersion
relation is given by
 2
cq
εT (q, ω) = (1406)
ω
Thus, the transverse excitations have a completely different character to the
longitudinal excitations, specially at large q. As q → 0, one expects that
σL (q, ω) → σT (q, ω) since electrons cannot differentiate between transverse
and longitudinal waves in this limit. In this limit, the conductivity may be
modelled by the complex Drude expression

ρ e2 τ 1
σ(0, ω) = (1407)
m 1 − iωτ
which leads to the dielectric constant being given by

4 π e2 ρ 1
ε(0, ω) = 1 − (1408)
m ω 2 1 + ωi τ

This approximate equality between the longitudinal and transverse dielectric


constants implies that the plasmon frequency also sets the frequency scale for
the interaction of photons with a metal. The dispersion relation for transverse
radiation becomes

ε(0, ω) ω 2 = c2 q 2 (1409)

454
which is given by
ω 2 − ωp2 = c2 q 2 (1410)
Thus, for ω < ωp , the wave will be reflected from a metal. For a typical metal,
where ρ ∼ 1022 electrons/cm3 , a typical plasmon frequency is 1015 sec−1 . This
typical frequency corresponds to a typical wave length of light in vacuum of
λp ∼ 10−7 m. Incident light with longer wave lengths will be reflected from
the metal. Hence, as εL (0, ω) = εT (0, ω), optical experiments that measure
εT (q, ω) produce similar information to energy loss experiments that determine
εL (q, ω).

The transverse conductivity may be evaluated directly by linear response


theory. The vector potential couples to the electrons via the interaction
"  #
|e| X |e| 2
Ĥint = p̂i . A(ri , t) + A(ri , t) . p̂i + A(ri , t)
2mc i c
(1411)
The interaction contains a paramagnetic contribution that involves a coupling to
the momentum density and a diamagnetic contribution that involves a coupling
with the density of the charged electrons. The transverse current density j(r, t)
is the mechanical current density e v, and is given by the sum of a paramagnetic
current j p and a diamagnetic current j d

j(r, t) = j p (r, t) + j d (r, t) (1412)

where the paramagnetic current is given by the symmetric operator


 
|e| X
j p (r) = − δ 3 (r − ri ) p̂i + p̂i δ 3 (r − ri ) (1413)
2m i

and the diamagnetic current is given by


| e |2 X 3
j d (r) = − δ (r − ri ) A(ri ) (1414)
mc i

To linear order in the vector potential, the interaction can be written as


Z
1 1
Ĥint = − d3 r j p (r) . A(r) (1415)
c
The induced paramagnetic current density is then found from linear response
theory in which the ground state is evaluated to first order in the perturbing
1
interaction Ĥint . The components of the induced paramagnetic current are
given as a causal convolution of a paramagnetic current - paramagnetic current
tensor correlation function, and the components of the total vector potential.
X Z +∞ Z
1 α,β
0
α
jind p (r, t) = dt d3 r0 Rj,j (r, r0 , t − t0 ) Aβ (r0 , t0 ) (1416)
−∞ c
β

455
where the paramagnetic response function is given by the ground state expec-
tation value
 
α,β 0 0 i 0 α β 0 0
Rj,j (r, r , t − t ) = + Θ( t − t ) < Ψ0 | jp (r, t) , jp (r , t ) | Ψ0 >

(1417)

This is known as the Kubo formula for the conductivity143 . The structure of the
Kubo formula for the response R is similar to that of the longitudinal response
function χ. They both involve the expectation value of a retarded two time
commutator. However, the Lindhard function involves the commutator of the
density operator and the Kubo formula involves the commutator of the current
operator.

On Fourier transforming the non-local relation between j p and A with re-


spect to space and time, one has
X 1 α,β
α
jind p (q, ω) = R (q, ω) Aβ (q, ω) (1418)
c j,j
β

Hence, to linear order in the vector potential, the total transverse current is
given by
" #
2
X 1 | e |
α
jind (q, ω) = Rα,β (q, ω) − δα,β ρ0 Aβ (q, ω) (1419)
c j,j mc
β

where it is assumed that the electron density is uniform and is given by ρ0 . The
transverse conductivity is then found with the aid of the relation between the
transverse electric field and the vector potential
ω
ET (q, ω) = i A(q, ω) (1420)
c
as " #
1 | e |2
σTα,β (q, ω) = α,β
Rj,j (q, ω) − δα,β ρ0 (1421)
iω m
The conductivity should be evaluated using a microscopic theory. The conduc-
tivity has a real part and an imaginary part that are connected by causality.
The conductivity determines the material’s properties and how transverse elec-
tromagnetic radiation or light interacts with the electrons in the metal.

The energy loss due to a longitudinal field is related to the inverse of the
dielectric constant, but the energy loss of a transverse field is related to the
143 R. Kubo, J. Phys. Soc. Jpn. 12, 570 (1957).

456
conductivity or the imaginary part of the dielectric constant. This can be seen
from the expression for the time-averaged dissipated power density
ω
P (q, ω) = =m ε(q, ω + iη) | E T (q, ω) |2

ω3
= =m ε(q, ω + iη) | A(q, ω) |2 (1422)
4 π c2
As the imaginary part of the dielectric constant is related to the real part of the
conductivity,  

=m ε(q, ω + iη) = <e σ(q, ω) (1423)
ω
the absorption of light measures the conductivity.

11.3.6 Optical Experiments


The optical conductivity can be measured in optical absorption and reflection
experiments. The wave vector of light in the medium is given by the complex
number
 1
ω 4 π i σ(ω) 2
q = 1 +
c ω
 
ω
= n + iκ (1424)
c
and this has the effect that intensity of light is exponentially attenuated as it
passes through the material
   
nz κωz
E(r, t) = E 0 exp i ω ( − t ) exp − (1425)
c c
Experiments measure the absorption coefficient η which is the fraction of light
absorbed in passing through unit thickness of the material

<e j . E κω
η = = 2 (1426)
n | E |2 c
Another experimental method measures the reflectance of polarized light. This
involves measuring the ratio of the reflected intensity of light to the incident
intensity, and gives rise to the real reflection coefficient. At oblique incidence,
with angle of incidence θ, one distinguishes between s and p polarized light. The
s polarized light has the polarization perpendicular to the plane of incidence and
the p polarized light has polarization parallel to the plane of incidence. The
reflectances of the s and p polarized light are given in terms of the complex
refractive index ñ = n + i κ, via the Fresnel formulas
cos θ − ( ñ2 − sin2 θ ) 21

Rs (θ) = 1
(1427)
cos θ + ( ñ2 − sin2 θ ) 2

457
s-polarized

k1
E1 ε(ω) > 0

ε0 = 1 E2
k2
θ θ2
θ

z>0
k0 E0

Figure 212: The geometry for s-polarized light falling incident on a metal.

p-polarized

k1
E1
ε(ω) > 0
ε0 = 1
k2
θ θ2 E2
θ

z>0
k0 E0

Figure 213: The geometry for p-polarized light falling incident on a metal.

458
Im z
C
z = ω+iη

Re z

Figure 214: The contour of integration C about the simple pole at z = ω + iη.

and
ñ cos θ − ( ñ2 − sin2 θ ) 21
2

Rp (θ) = 1
(1428)
ñ2 cos θ + ( ñ2 − sin2 θ ) 2

The complex refractive index can then be inferred from measurements of Rs (θ)
and Rp (θ). However, it is usual to infer the real part from the imaginary part
via the Kramers-Kronig relation144 .

11.3.7 Kramers-Kronig Relation


Causality requires that the frequency be continued in the upper-half complex
plane ω + i η in the response functions. This has the consequence that the
response function is analytic in the upper-half complex plane. Also, it is re-
quired that the integrand vanishes over a semi-circular contour at infinity in the
upper-half complex plane. With these restrictions, one can consider the Cauchy
integral
ε(q, z) − 1
Z
1
( ε(q, ω + iη) − 1 ) = dz (1429)
2πi c z − ω − iη
where the contour is taken around the point z = ω + iη. If ε(q, z) does not
have a pole at z = 0, the contour of integration can be deformed to the real
axis and an infinite semi-circular contour in the upper-half complex plane. In
this case, one finds
Z +∞
1 ε(q, z) − 1
ε(q, ω + iη) − 1 = Pr dz (1430)
πi −∞ z − ω
in which the contribution of the small semi-circle around the pole at z = ω + i η
has cancelled with half of the left-hand side. On writing the dielectric constant
as the sum of the real and imaginary parts
ε(q, z) = <e ε(q, z) + i =m ε(q, z) (1431)
144 H. A. Kramers, Nature, 117, 775 (1926), R. de L. Kronig, J. Opt. Soc. Am. 12, 547

(1926).

459
Im z
C'

z = R eiθ

z = ω+iη

Re z

Figure 215: The contour of integration C is deformed to C 0 which includes the


real axis and the semi-circular contour at infinity.

and on equating the real terms, one finds


Z +∞ =m ε(q, z)
1
<e ε(q, ω + iη) − 1 = Pr dz (1432)
π −∞ z − ω
The imaginary terms are related via
Z +∞ <e ε(q, z) − 1
1
=m ε(q, ω + iη) = − Pr dz (1433)
π −∞ z − ω
These relations can be recast in the form
Z ∞ z =m ε(q, z)
2
<e ε(q, ω + iη) − 1 = Pr dz (1434)
π 0 z2 − ω2
and Z ∞ <e ε(q, z)

=m ε(q, ω + iη) = − Pr dz (1435)
π 0 z2 − ω2
These are the Kramers-Kronig relations145 . They can be used to analyze exper-
imental data or as consistency checks.

——————————————————————————————————

11.3.8 Exercise 66
Derive the form of the Kramers-Kronig relation for the imaginary part of the
dielectric constant ε(q, ω)

4 π σ(q, 0)
Z ∞ <e ε(q, z)

=m ε(q, ω) = − Pr dz (1436)
ω π 0 z2 − ω2
145 H. A. Kramers, Nature, 117, 775 (1926), R. de L. Kronig, J. Opt. Soc. Am. 12, 547

(1926).

460
valid for a material which has a finite d.c. conductivity σ(q, 0).

——————————————————————————————————

Another sum rule, the optical sum rule, is stated as


Z ∞  
π 2
dω ω =m εT (q, ω + iη) = ω (1437)
0 2 p
The optical sum rule can be derived by exact methods. However, it can also be
proved by noting that at high frequencies
ωp2
lim εT (q, ω) = 1 − (1438)
ω → ∞ ω2
On expressing the imaginary part of the dielectric constant in terms of the real
part, one can verify the sum rule using contour integration. A more usual form
of the optical sum rule is stated in terms of a sum rule for the conductivity
Z ∞
π ρ e2
dω σ(0, ω) = (1439)
0 2 m
where ρ is the electron density. Kramers-Kronig relations and sum rules can be
established for a variety of response functions146 . Since the inverse dielectric
constant is the longitudinal response function, 1/ε(q, ω) − 1 also satisfies a
Kramers-Kronig relation.

——————————————————————————————————

11.3.9 Exercise 67
The n-th moment of the imaginary part of the dielectric constant is defined by
Mn Z ∞
Mn = dω ω n =m ε(q, ω + iη) (1440)
0
Show that M1 is given by
e2 ρ
M1 = 2 π 2 (1441)
m
and that M−1 is given by
 
π
M−1 = ε(q, 0) − 1 (1442)
2

——————————————————————————————————

146 P. C. Martin, Phys. Rev. 161, 143 (1967).

461
11.3.10 The Drude Conductivity
Metals have a large conductivity, and as a result, electromagnetic fields only
penetrate a small distance into the metal before the energy of the field is ab-
sorbed and dissipated as Joule heating. For low frequencies, or slowly spatially
varying fields, the penetration depth δ can be calculated from Maxwell’s equa-
tions using the frequency-dependent Drude electrical conductivity. The Drude
conductivity is calculated by assuming that the photon has a long wavelength,
therefore, q ≈ 0. On assuming that the medium is homogeneous and isotropic,
one finds that the conductivity tensor is diagonal
e2 ρ τ 1
σ α,β (ω) = δ α,β (1443)
m 1 − iωτ

The Drude formula for the conductivity can be obtained directly from Kubo
formulae, for the case of non-interacting electrons. On expressing the Kubo
formulae in terms of the completes set of exact eigenstates of the many-particle
Hamiltonian Ĥ0
Ĥ0 | Ψn > = En | Ψn > (1444)
for t > 0, one finds
 
α,β 0 i X α β 0 t
R (r, r , t) = < Ψ0 | jp (r) | Ψn > < Ψn | j (r ) | Ψ0 > exp + i (E0 − En )
h̄ n h̄
 
i X t
− < Ψ0 | jpβ (r0 ) | Ψn > < Ψn | j α (r) | Ψ0 > exp − i (E0 − En )
h̄ n h̄
(1445)
On Fourier transforming the Kubo formula with respect to time, one obtains
X < Ψ0 | jpα (r) | Ψn > < Ψn | j β (r0 ) | Ψ0 >
Rα,β (r, r0 , ω) = −
n
h̄ ω + i η + E0 − En
X < Ψ0 | jpβ (r0 ) | Ψn > < Ψn | j α (r) | Ψ0 >
+
n
h̄ ω + i η + En − E0
(1446)
where the convergence factor η is to be assigned a physical meaning. This
expression is to be evaluated for non-interacting electrons in which case, the
states | Ψn > can be taken to be Slater determinants. The matrix elements
of the current density operators can be expressed in terms of the one-electron
wave functions φγ (r) and φγ 0 (r) via
 
α | e | h̄ ∗ α α ∗
< Ψn | j (r) | Ψ0 > = − φγ 0 (r) ∇ φγ (r) − ∇ φγ 0 (r) φγ (r)
2im
 
| e | h̄ ∗ α
= − =m φγ 0 (r) ∇ φγ (r) (1447)
m

462
where the electron in the state labelled by the one-electron quantum number γ
is excited to the state with quantum number γ 0 in the final state. The energy
difference between the initial and final states is given by the energy difference
between the initial and final energies of the excited electron

En − E 0 = E γ 0 − Eγ (1448)

Thus, one has


   
X =m φ∗γ (r) ∇α φγ 0 (r) =m φ∗γ 0 (r0 ) ∇β φγ (r0 )
e2 h̄2
Rα,β (r, r0 , ω) = −
m2 h̄ ω + i η + Eγ − Eγ 0
γ,γ 0
   
∗ 0 β 0 ∗ α
=m φγ (r ) ∇ φγ (r ) =m φγ 0 (r) ∇ φγ (r)
0
e2 h̄2 X
+
m2 0
h̄ ω + i η + Eγ 0 − Eγ
γ,γ
(1449)

where Eγ < µ and Eγ 0 > µ.

The conductivity response function will be evaluated for free-electrons. On


inserting the single-electron wave functions, the response function is found as
 
i 0 0
exp − (k − k ) . (r − r )
e2 h̄2 X h̄
Rα,β (r, r0 , ω) = − 2 2
(k α + k 0α ) (k β + k 0β )
4m V h̄ ω + i η + Ek − Ek0
k,σ;k0
 
i 0 0
exp − h̄ (k − k ) . (r − r )
e2 h̄2 X
α 0α β 0β
+ (k + k ) (k + k )
4 m2 V 2 0
h̄ ω + i η − Ek + Ek0
k,σ;k

(1450)

where k α and k β denote the α and β components of the vector k. The summation
over k, σ runs over the occupied states k < kF whereas the sum over k 0 runs
over the unoccupied states with the spin σ but with k 0 > kF . The initial
and final state spin quantum numbers are identical. On Fourier transforming
with respect to the space variable, and re-arranging the summation index in the
second term, one finds

e2 h̄2 X f (Ek ) − f (Ek+q )


Rα,β (q, ω) = − 2
(2k α + q α ) (2k β + q β )
4m V h̄ ω + i η + Ek − Ek+q
k,σ

(1451)

In this expression, the effect of the Pauli-exclusion principle is automatically


accounted for.

463
Due to the large magnitude of c, for fixed ω, this expression can be evaluated
to leading-order in q. In this case, as space is isotropic, the response function
is also isotropic. That is, the response function is diagonal in the indices α and
β and the diagonal components have equal magnitudes. Hence, the diagonal
components can be evaluated from the relation
3
1 X β,β
Rα,α (q, ω) = R (q, ω) (1452)
3
β=1

The response function can be expressed as


2 e2 X f (Ek ) − f (Ek+q )
Rα,β (q, ω) = − δ α,β Ek
3mV h̄ ω + i η + Ek − Ek+q
k,σ

(1453)
On Taylor expanding the Fermi function f (Ek+q ) in powers of (Ek+q − Ek ),
one has
2 e2 X ( Ek − Ek+q )
 
∂f
Rα,β (q, ω) = − δ α,β Ek
3mV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

2 e2 X
   
α,β ∂f h̄ ω + i η
= −δ Ek 1 −
3mV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

(1454)
The first term can be evaluated through integration by parts
  Z ∞  
2 X ∂f 2 X ∂f
− Ek = − dE E ρ(E)
3 ∂Ek 3 σ 0 ∂E
σ,k
Z ∞  
2 X ∂
= dE f (E) E ρ(E)
3 σ 0 ∂E
(1455)
since
√ the boundary terms vanish. Furthermore, since for free electrons ρ(E) ∝
E, this term is evaluated as
2 X

∂f
 X Z ∞
− Ek = dE f (E) ρ(E)
3 ∂Ek σ 0
σ,k

= Ne (1456)
as the factor of 32 cancels with the factor of 32 from the derivative. Due to this
simplification, the response function is given by
ρ0 e2 2 e2 X
 
∂f h̄ ω + i η
Rα,α (q, ω) = + Ek
m 3mV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

(1457)

464
On substituting this expression into the conductivity, one finds that the first
term cancels with the diamagnetic current. This cancellation is responsible for
prohibiting current to flow occurring in a metal as a response to an applied
magnetic field. In other words, a normal metal does not superconduct due to
the cancellation of the diamagnetic current.

The conductivity is simply given by

2 e2
 
X ∂f h̄
σ α,β (q, ω) = δ α,β Ek
3imV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

(1458)

The derivative of the Fermi function is only non-zero in the vicinity of the Fermi
energy. In the limit, T → 0, the derivative may be expressed as
 
∂f
− = δ(  − F ) (1459)
∂

The appearance of the derivative of the Fermi function in the expression for the
conductivity is a consequence of the Pauli-exclusion principle. Only electrons
close to the Fermi energy can absorb relatively small amounts of energy and
be excited to unoccupied single-electron states and, hence, carry current. A
phenomenological relaxation time τ can be defined via

= η (1460)
τ
The relaxation time τ can be thought of as the lifetime of the current-carrying
hole or the current-carrying excited electron. This lifetime must be caused
by a scattering mechanism. In more rigorous treatments of the conductivity,
the scattering rate is calculated using microscopic descriptions of the scattering
interaction. When expressed in terms of the relaxation rate, the conductivity
becomes
 
∂f
Ek ∂Ek
e2 τ X 2
σ α,β (q, ω) = − δ α,β (1461)
mV 3 1 − i τ ( ω − q . v(k) )
k,σ

In the limit, q → 0, one recovers the Drude approximation for the conductivity

ρ e2 τ 1
σ α,β (ω) = δ α,β (1462)
m 1 − iτ ω
where ρ is the electron density. The Drude conductivity is purely real in the
d.c. limit, and is given by

ρ e2 τ
σ α,β (0) = δ α,β (1463)
m

465
and at finite frequencies has a real part that decays like ω −2

ρ e2 τ 1
<e σ α,β (ω) = δ α,β (1464)
m 1 + ω2 τ 2
Thus, the Drude conductivity has a peak at zero frequency and the width of
the peak is determined by the relaxation time. The integrated strength of the
low-energy Drude peak in the conductivity is given by
Z ∞
π ρ e2
 
dω <e σ (ω) = δ α,β
α,β
(1465)
0 2m
Hence, the intensity of the Drude peak provides a measure of the number of con-
duction electrons in a system of non-interacting electrons. This sum rule, like
the total number of electrons, is unchanged by electron-electron interactions.
Electron-electron interactions do renormalize the mass in the expression for the
conductivity. However, the interactions also renormalizes the lifetime, τ , by the
same factor, Z, which is the wave function renormalization. In this way, the d.c.
limit of the conductivity essentially remains un-renormalized. However, the fre-
quency width of the Drude peak is narrowed by the factor Z. In heavy fermion
materials, such as CeAl3 147 , the experimentally width of the determined Drude
peak has been found to decrease dramatically at low temperatures as the heavy
quasi-particles form. The mass enhancement determined from the width at low
temperatures is consistent with the large quasi-particle mass determined from
the low temperature limit of the electronic specific heat.

——————————————————————————————————

11.3.11 Exercise 68
Show that the E field of microwaves, with low frequency ω, satisfy the equation
4 π i σ(ω) ω
− ∇2 E(r, ω) = E(r, ω) (1466)
c2
where σ(ω) is the diagonal component of the conductivity tensor. Solve this
equation for the electric field and, therefore, calculate the classical skin depth
δ. The classical skin depth is defined as the distance δ that an electric field
penetrates into a metal before being attenuated.

——————————————————————————————————

The analysis of Exercise 68 is only valid if the electric field vary slowly over
distances of the order of a mean free path λ. The analysis is only valid for low
frequencies and dirty metals. However, for good metals, E(r, ω) varies rapidly
147 A. M. Awasthi, L. Degiorgi, G. Grüner, Y. Dalichaouch and M. B. Maple, Phys. Rev. B

48, 10692 (1993).

466
Figure 216: The frequency and temperature dependence of the Drude peak in
U P t3 as measured by S. Donovan, A. Schwartz, and G. Grüner, Phys. Rev.
Lett., 79, 1401 (1997).

467
in space. This regime corresponds to the anomalous skin effect148 . Since the
electrons do not respond to the field instantaneously and locally, the retarded
and non-local response function ought to be used. In this case, one should solve
Maxwell’s equations by solving for the Fourier components of the fields E(q, ω)
and B(q, ω), and by using an expression for the conductivity tensor in which
both the wave vector and frequency dependence are kept149 . This procedure is
crucial for the discussion of the anomalous skin effect.

——————————————————————————————————

11.3.12 Exercise 69
The conductivity tensor can be expressed as an integral over the Fermi surface,

e2 d2 S τ v α (k) v β (k)
Z
σ α,β (q, ω) = 3
(1467)
4π | h̄ v(k) | 1 − i τ ( ω − q . v(k) )

Consider a clean material with a sufficiently long mean free path λ, such that
qz λ  1 for fixed qz . Show that the transverse conductivity σ x,x (qz êz , ω) is
given by the approximate expression150
3π σ0
σ x,x (qz êz , 0) = (1468)
4 | qz | λ

——————————————————————————————————

11.3.13 The Anomalous Skin Effect


For clean materials with large mean free-paths λ, the penetration of an electric
field into a metal is governed by the anomalous skin effect. In the low frequency
limit, the component of the electric field parallel to the surface Ex (z, ω) is
governed by
∂ 2 Ex ω2 4πiω
2
+ 2 Ex = − jx (1469)
∂z c c2
where the surface of the material is the z = 0 plane. We shall assume that
electrons are specularly reflected from the surface. This boundary condition
can be understood by imagining that the surface of the metal demarcates the
boundary between two identical solids. One solid represents an extension of
the actual material generated by mirror symmetry. The condition of specular
reflection amounts to assuming that the electrons and fields in the mirror image
148 A. B. Pippard, Proc. Roy. Soc. A 191, 385 (1947), A. B. Pippard, Proc. Roy. Soc. A,

224, 273 (1954).


149 D. C. Mattis and G. Dresselhaus, Phys. Rev. 111, 403 (1958).
150 G. E. H. Reuter and E. H. Sondheimer, Proc. Roy. Soc. A 195, 336 (1948).

468
z=0
σk'
k'

σk k
e e
z=0

k'
z>0

k e

Figure 217: Imposing specular boundary conditions on electrons in a solid (lower


frame) can be re-interpreted in terms of two identical crystals (upper frame)
related by mirror symmetry σ.

solid behave in the same way as in the actual solid. This leads to the boundary
condition for the electric field being given by
   
∂Ex ∂Ex
= − (1470)
∂z z=− ∂z z=
Therefore, on subsuming the boundary condition in the equation of motion for
the field, one has
∂ 2 Ex ω2
 
4πiω ∂Ex
+ E x = − jx + 2 δ(z) (1471)
∂z 2 c2 c2 ∂z z=0
Hence, on Fourier transforming with respect to z and using Ohm’s law
jx (qz , ω) = σ x,x (qz , ω) Ex (qz , ω) (1472)
one obtains the solution
 
∂Ex

2 ∂z


z=0
Ex (qz , ω) = ω2 4 π i ω
(1473)
c2 − qz2 + c2 σ x,x (qz , ω)
In the limit of extremely long mean free path λ → ∞, the conductivity
simplifies to151
3 π σ(0, 0)
σ x,x (qz , 0) = (1474)
4 | qz | λ
The spatial dependence of the electric field is given by the inverse Fourier
151 G. E. H. Reuter and E. H. Sondheimer, Proc. Roy. Soc. A 195, 336 (1948).

469
1

0.8

σxx(0,0)
0.6

σxx(qz,0)/σ
0.4

0.2

0
0 2 4 6 8

Figure 218: The q-dependence of the low-frequency transverse conductivity


σ xx (qz , ω).

transform,
 
  Z ∞ exp − i qz z
∂Ex dqz
Ex (z, ω) = ω2 3 π2 i ω
(1475)
∂z z=0 −∞ π c2 − qz2 + c2 λ | qz | σ(0, 0)

On defining δ via
 13
3 π 2 ω σ(0, 0)

−1
δ = (1476)
c2 λ
one has
 
 Z ∞ | q
 z | exp − i q z z
∂Ex dqz
Ex (z, ω) = 2
∂z
z=0 −∞ π ωc2 | qz | − | qz3 | + i δ −3
Z ∞
x cos( x zδ )
 
∂Ex dx
= −2iδ
∂z z=0 0 π 1 + i x3 − i ω2c2δ2 x
(1477)
In the last line, we have introduced a dimensionless variable x = qz δ, multiplied
the top and bottom by a factor of iδ 3 , and folded the region of integration to the
positive x axis. For low frequencies, the decay of the electric field is governed
by δ, the anomalous skin depth.

At the surface, the value of the field is given by


  Z ∞
∂Ex dx x
Ex (0, ω) = − 2 i δ ω2 δ2
∂z
z=0 0 π 1 + ix − i
3
c2 x

470
0.8

ω)
E(z,ω
Im E(z)
0.6
Re E(z)

0.4

0.2

0
-8 -6 -4 -2 0
δ2
z/δ 4 6 8

-0.2

Figure 219: The spatial-dependence of the transverse electric field as calculated


with the short wave length approximation for the conductivity σ xx (qz , ω), as a
function of zδ .

 
2δ ∂Ex i
≈ − (1 + √ ) (1478)
3 ∂z
z=0 3
Far from the surface, the field has an exponential decay
 2  
δ z
Ex (z, ω) ∼ exp − (1479)
z δ
which decays over a distance δ.

This result for the anomalous skin depth δ was first obtained by Pippard152 ,
up to a numerical factor. Pippard noted that only the fraction of the electrons
δ
λ close to the surface may participate in the screening process. That is, only
the electrons moving parallel to the surface are strongly affected by the electric
field. The electrons that remain within the penetration depth δ before being
scattered, subtend an angle of
δ
dθ ≈ (1480)
λ
The number of the electrons which are capable of responding to the field is
proportional to the solid angle dΩ,
dΩ = 2 π sin θ dθ ∼ 2 π dθ (1481)
π
since θ ≈ 2. Hence, the effective electron density, ρef f , is given by
δ
ρef f ≈ ρ (1482)
λ
152 A. B. Pippard, Proc. Roy. Soc. A, 191, 385 (1947), A. B. Pippard, Proc. Roy. Soc. A,

224, 273 (1954).

471
δ z>0

λ
δθ

Figure 220: The geometry determining the effective number of electrons which
are able to screen the external field.

where ρ is the uniform electron density. This implies that conductivity parallel
to the surface should be reduced by the factor λδ . Thus, on applying the analysis
of the classical skin effect, one recovers the relation
4πω δ
δ −2 ∼ σ(0) (1483)
c2 λ
Hence, one has Pippard’s relation
  13
−1 4 π ω σ(0)
δ ∼ (1484)
c2 λ

which only differs by a numerical factor from the previously given expression
for the skin depth.

11.3.14 Inter-Band Transitions


The absorption of a photon of wave vector q, may cause an electron to make a
transition between the initially occupied state with Bloch wave vector k to a final
state with wave vector k + q. Since the wave vector of light q is small compared
with kF , for a given ω, the final state must be in a different band and must be
empty. These are inter-band transitions. Materials with small inter-band gaps
can have large dielectric constants. The inter-band contribution to the dielectric
constant can be obtained by neglecting q, thereby producing vertical transition
between the different Bloch bands. The imaginary part of the dielectric constant

472
due to inter-band transitions can be written as
2 Z 2
d3 k
  
e
=m ε(ω+iη) = 8 π 2 h̄2

êα . M k δ( Ec,k − Ev,k − h̄ ω )
mω ( 2 π )3
(1485)
The matrix elements for the inter-band transitions are given by
Z    
êα . M k = êα . d3 r exp − i k . r uv,k (r) ∇ exp + i k . r uc,k (r)

(1486)

where un,k are the periodic functions of r in the Bloch functions. The sum over
k can be transformed into an integral
2

2 Z êα . M
k
d2 S
  
e
=m ε(ω+iη) = 8 π 2 h̄2
mω ( 2 π )3
∇ ( Ec,k − Ev,k )

Ec −Ev =h̄ω
(1487)
where d2 S represents an element of the surface in k space defined by the equation
h̄ ω = Ec,k − Ev,k . The quantity
2

êα . M k
d2 S
Z
J(ω) = (1488)
( 2 π )3

k ( E c,k − E v,k )

Ec −Ev =h̄ω

is known as the joint density of states. The joint density of states varies rapidly
with respect to ω at the critical points, which are defined by


∇k ( Ec,k − Ev,k ) = 0 (1489)
Ec −Ev =h̄ω

The inter-band transitions produce a broad continuum in the absorption spec-


trum, and only the van Hove singularities may be uniquely identified in the
spectrum. The analytic behavior of the dielectric constant near a singularity
may be obtained by Taylor expanding about the critical point.

When other processes such as electron-phonon scattering are considered,


second order time-dependent perturbation theory describes indirect transitions.
In this case, a phonon may be absorbed or emitted by the lattice while the
photon is being absorbed. The emission or absorption of the phonon introduces
a change of momentum q. Conservation of momentum leads to the momenta
of the initial and final state of the electron being related via q = k 0 − k.
Since the energy of the phonon is usually negligible compared with the energy

473
of the photon, the energy of the absorbed photon is approximately given by the
difference of the electron’s initial and final state energies

h̄ ω ≈ Ec,k − Ev,k0 (1490)

Thus, since q varies continuously, the indirect inter-band transitions have a con-
tinuous spectrum. The threshold energy for the inter-band transition is close to
the minimum value of Ec,k − Ev,k0 , for all possible values of k and k 0 . If the
minimum value of Ec,k − Ev,k0 occurs for k = k 0 the band gap is known as
a direct band gap, whereas if k 6= k 0 the band gap is called an indirect band gap.

11.4 The Fermi Surface


The Fermi surface determines most of the thermodynamic, transport and opti-
cal properties of a solid. The geometry of the Fermi surface can be determined
experimentally, through a variety of techniques. The most powerful of these
techniques is the measurement of de Haas - van Alphen oscillations. The de
Haas - van Alphen effect is manifested as an oscillatory behavior of the magne-
tization153 . The magnetization is periodic in the inverse of the applied magnetic
1 1
field H . Onsager154 pointed out that the period in H is given by the expression
 
1 |e|
∆ = 2π Ae (1491)
H h̄ c

where Ae is the extremal cross-sectional area of the Fermi surface in the plane
perpendicular to the direction of the applied field H . By observing the period of
oscillations for the different directions of the applied field, one can measure the
extremal areas for each direction. This information can then be used to recon-
struct the Fermi surface of three-dimensional metals155 . First, some properties
of the electron orbits in an applied field will be examined, then the experimental
methods used in the determination of the Fermi surface will be described.

11.4.1 Semi-Classical Orbits


In the classical approximation, the Hamilton equations of motion are given by
the pair of equations
1
ṙ = v(k) = ∇ Ek (1492)
h̄ k
and
|e|
h̄ k̇ = − v(k) ∧ H (1493)
c
153 W. J. de Haas and P. M. van Alphen, Proc. Amsterdam, Acad. 33, 1106 (1930).
154 L. Onsager, Phil. Mag. 45, 1006 (1952).
155 D. Schoenberg, Proc. Roy. Soc. A 170, 341 (1939).

474
From these, one finds that k changes in a manner such that it remains on the
constant energy surfaces. This is found by observing that, from the above equa-
tion, k̇ is perpendicular to ∇k Ek . Also since k̇ is perpendicular to H, the k
space orbits are a section of the constant energy surfaces with normal along
the z-axis. That is the orbits traverse the constant energy surfaces, but kz is a
constant.

The real space orbits are perpendicular to the k space orbits. To show this,
we shall first prove that k̇ is perpendicular to ṙ. On taking the vector product
of the equation of motion with the vector H, one finds
 
|e|
h̄ H ∧ k̇ = − H ∧ v(k) ∧ H (1494)
c
The component of the velocity perpendicular to the applied field is given by
 
ṙ⊥ = ṙ − Ĥ ṙ . Ĥ
 
= Ĥ ∧ ṙ ∧ Ĥ (1495)

where Ĥ is the unit vector in the direction of H. Hence,


c h̄
ṙ⊥ = − Ĥ ∧ k̇ (1496)
| e | Hz
Thus, on integrating this with respect to time, one finds that the displacement
∆r⊥ is given in terms of the displacement in k space through
c h̄
∆r⊥ = − Ĥ ∧ ∆k (1497)
| e | Hz
Thus, the real space orbit is perpendicular to the k space orbit and is scaled by
a factor of ec Hh̄z .

The period T at which the orbit is traversed is given by the integral over
one orbit I
dk
T = (1498)

The rate of change of k is given by the Lorentz Force

| e |
k̇ = 2 ∇k E ∧ H
h̄ c

|e|
= 2 Hz ∇k E⊥
(1499)
h̄ c
where ∇k E⊥ is the component of the gradient perpendicular to H, i.e., the
projection in the plane of the orbit. Thus,
h̄2 c 1
I
dk
T = (1500)
∇ k E⊥ | e | H z

475
If semi-classical quantization considerations are applied, then the energy of
the orbits become quantized as do the orbits themselves. The area enclosed by
the orbits are related to the energy, and so the areas are also expected to be
quantized. This shall be shown by two methods, in the first the quantization
condition is imposed through the energy - time uncertainty relation, and the
second method will utilize the Bohr-Sommerfeld quantization condition.

Quantization Using Energy - Time Uncertainty.

The relationship between the energy and the areas enclosed by the k space
orbits can be found from consideration of two classical orbits, one with energy
E and another with energy E + ∆E where both orbits are in the same kz
plane. Then, let ∆k be the minimum distance between these two orbits. The
value of ∆k is related to ∆E via


∆E = ∇k E⊥ ∆k
(1501)

This relation can be substituted into the expression for the period to yield
h̄2 c
I
1
T = ∆k dk (1502)
| e | Hz ∆E
However, the area between the two successive orbits in reciprocal space is given
by the integral I
∆A = ∆k dk (1503)

Thus, the period can be expressed as


h̄2 c ∆A
T = (1504)
| e | Hz ∆E
The orbits can be quantized through the energy uncertainty relation

En+1 − En =
T
| e | Hz ∆E
= (1505)
h̄ c ∆A

Furthermore, as the ratio of the energy difference to the difference in areas of


consecutive orbits is defined as
∆E En+1 − En
= (1506)
∆A An+1 − An
one can cancel a factor of ∆E to find that the area enclosed between consecutive
Landau orbits is quantized
2π|e|
An+1 − An = Hz (1507)
h̄ c

476
This difference equation can be solved to yield the area (in reciprocal space)
enclosed by the n-th Landau orbital as
 
2π|e|
An = n + λ Hz (1508)
h̄ c

where λ is a constant, independent of n. Thus, the area of a Landau orbit in


k space is related to n and the applied field Hz , through the Onsager equation156 .

Bohr-Sommerfeld Quantization.

An alternate derivation of the Onsager equation follows from the Bohr-


Sommerfeld quantization condition
I  
1
p . dr = 2 π h̄ n + (1509)
2

The mechanical momentum is given by


|e|
p = h̄ k − A (1510)
c
The integral is evaluated over an orbit in the x − y plane perpendicular to H.
The orbit is obtained from the equation of motion with the Lorentz Force Law
|e|
h̄ k̇ = − ṙ ∧ H (1511)
c
The equation of motion can be integrated with respect to time, to yield
|e|
h̄ k = − r ∧ H (1512)
c
Thus, the Bohr-Sommerfeld quantization condition reduces to
I    
|e| 1
− r ∧ H + A . dr = 2 π h̄ n +
c 2
 
|e|
I I
= H . dr ∧ r − dr . A (1513)
c

However, the integral I


dr ∧ r = 2 Ar êz (1514)

is just twice the area enclosed by the real space orbit, and the integral of the
vector potential around the loop is given by
I
dr . A = Φ (1515)

156 L. Onsager, Phil. Mag. 43, 1006, (1952).

477
where Φ = Ar Hz is the flux enclosed by the orbit. Hence, the magnitude of
the area of the orbit, Ar , in real space is quantized and is given by
 
1 2 π h̄ c
Ar = n + (1516)
2 | e | Hz

Since the real space and momentum space orbitals are related via
h̄ c
∆r = ∆k (1517)
| e | Hz

one can scale the areas of the real and momentum space orbits. Thus, one
recovers the Onsager formulae for the area of the momentum space orbit
 
1 2π|e|
An = n + Hz (1518)
2 h̄ c

This quantization condition does not always hold. Some electron orbits do not
form closed curves in the Brillouin zone. These open orbits can be understood
by noting that the Bloch state energy is periodic in k space, and if an orbit
crosses the boundary of the first Brillouin zone, it will continue into the next
zone. Sometimes the sections of the orbits in the Brillouin zones can be com-
bined, by translating them through reciprocal lattice vectors, to form closed
orbits. Trajectories that extends across an entire Brillouin zone can not be used
to construct a closed orbit, and are called open orbits157 Another cause for the
failure of the quantization condition is caused when the field is larger than the
energy gaps between successive bands. This gives rise to the phenomenon of
magnetic breakdown. In this case, the levels are broadened and the cyclotron
orbits tunnel between Bloch states in the different bands, ignoring the band
gaps158 . Magnetic breakdown is likely to occur when h̄ ωc µ  Eg2 , where Eg
is the gap between the bands.

11.4.2 de Haas - van Alphen Oscillations


Given a solid with Hz = 0, surfaces of constant energy do not intersect when
plotted in k space. The consecutive constant energy surfaces, corresponding to
the different allowed values of energy, completely fill momentum space. The
states on the surfaces which have energy less than µ will be occupied, and
those with energy greater than µ are empty. On applying a magnetic field, Hz ,
the momentum perpendicular to the field is no longer a constant of motion,
but kz is constant. However, as time evolves, an orbit never leaves its surface
of constant energy. The magnetic field quantizes the orbits. In momentum
space, the allowed orbits form a nested set of discrete Landau tubes. Orbits in
157 I. M. Lifshitz and M. I. Kaganov, Sov. Phys. Uspekhi, 2, 831 (1959).
158 A. B. Pippard, Proc. Roy. Soc. A 270, 1 (1962).

478
the regions between the tubes are forbidden. For a general Fermi surface of a
three-dimensional crystal, the intersection of the constant energy surfaces with
a plane of fixed kz need not be circular, so that the Landau tubes need not have
cylindrical cross-sections. However, for free electrons, the zero field constant
energy surfaces are spherical and the Landau tubes are cylindrical. The radius
of the tubes is determined by the energy of the x-y motion, while the height
is determined by the component of the kinetic energy due to motion in the z-
direction. For free electrons, an occupied orbit is specified by kz and n. The
energy of the free-electron orbit is given by

1 h̄2 kz2
En,kz = ( n + ) h̄ ωc + (1519)
2 2m
where
| e | Hz
ωc = (1520)
mc
The orbit maps out a circle of area
 
2π|e| 1
An = Hz n + (1521)
h̄ c 2

so the orbits will consist of concentric circles. On varying kz but holding n fixed,
the consecutive orbits will map out a tube in k space. The occupied portions
of k space will lie on portions of a series of tubes. These portions will be con-
tained in a volume similar to the volume of the Fermi surface, when Hz = 0.
The bounding volume must reduce to the volume enclosed by the Fermi surface
when the field is decreased. For fields of the order of H ∼ 1 kG, the Fermi
surface cuts about 103 such tubes, so the quasi-classical approximation can be
expected to be valid.

As the field increases, the cross-sectional area enclosed by the tubes also
increases, as does the number of electrons held by the tubes. The extremal tube
may cross the zero field Fermi surface (H = 0) at which point the electrons
in the tube will be entirely transferred into the tubes with lower n values. The
changing structure gives rise to a loss of tubes from the occupied Fermi volume
when the field changes by amounts ∆H. Thus, if at some value of kz , the
occupied Landau tube with the largest area has the largest value n given by the
extremal area of the Fermi surface A(kF )

2π|e| 1
Hz ( n + ) ∼ π kF2 (kz ) = A(kF ) (1522)
h̄ c 2
then, on changing Hz to Hz + ∆H the tube becomes unoccupied so the largest
Landau tube changes from n to n − 1. This occurs when
1 1
(n − ) ( Hz + ∆H ) = ( n + ) Hz
2 2
(1523)

479
Thus, the extremal orbit crosses the Fermi surface when Hz is increased by

n ∆H ≈ Hz (1524)
∆H
which can be used to eliminate n and relate Hz to the momentum space area
of the extremal orbit.
Hz 2 π | e |
A(kF ) = π kF2 (kz ) = Hz (1525)
∆H h̄ c
Thus, n decreases by unity at fields given by

∆H 2π|e| 1
− 2
= − (1526)
H h̄ c A(kz )
 
1
In other words, ∆n changes by − 1 with increasing ∆ Hz . The non-
monotonic variation of the occupancy of the extremal orbits or tubes gives rise
to oscillations in the Free energy as Hz is varied. This can also be seen from
examination of the density of states, per spin polarization, for free electrons
Z ∞
X dkz h̄2 kz2
ρ() = D Lz δ(  − n h̄ ωc − )
n −∞ 2 π 2m
Lx Ly Lz X Z ∞ h̄2 kz2
= m ω c dk z δ(  − n h̄ ω c − )
4 π 2 h̄ n 0 2m
Lx Ly Lz 2 X θ(  − n h̄ ωc )
= 2 2 m ω c p
4 π h̄ n 2 m (  − n h̄ ωc )
(1527)

where D is the degeneracy of a Landau orbital. The degeneracy is given by the


ratio of the cross-section of the crystal to the real space area enclosed between
the Landau orbits
Lx Ly
D =
∆Ar
Lx Ly
= m ωc (1528)
2 π h̄
which increases with increasing field. The density of states has equally spaced
square root singularities determined by the energies of the Landau levels, but
yet still roughly follows the zero field density of states. On changing the field
the spacing between the singularities increases. This means that, as the field is
increased, successive singularities may cross the Fermi energy, and give rise to
oscillations in physical properties.

Physical properties are expressible as averages which are weighted by the


product of the Fermi function and the density of states. For zero spin-orbit

480
coupling, the average of A is given by
X Z +∞
A = d f () ρ(  − µB σ Hz ) Aσ () (1529)
σ −∞

in which the electronic density of states is spin split by the Zeeman field. This
splitting is comparable with the effect of h̄ ωc . Increasing the field will produce
regular oscillations in the integrand which will show up in A. Due to the thermal
smearing manifested by the Fermi function, the oscillations of A as a function
of H1z can only be seen at sufficiently low temperatures such that
kB T  h̄ ωc (1530)
If this condition is not satisfied, the Fermi function becomes broad and washes
out the peaks in the integrand near µ. As
| e | h̄
∼ 1.34 × 10−4 k/G (1531)
m c kB
it is found that, for a typical field of H = 10 kG, the oscillations will only be
appreciable below T ∼ 2 K.

——————————————————————————————————

11.4.3 Exercise 70
A non-uniformity of the magnetic field in a de Haas - van Alphen experiment
may cause the oscillations in M z to be washed out. Calculate the field derivative
of the electron energy
∂En,k
(1532)
∂H
for an extremal orbit. Determine the maximum allowed variation of Hz that is
allowable for the oscillations to still be observed. Show that it is given by δH,
where
δH 2π|e|
< (1533)
Hz2 h̄ c A
and A is the area of the extremal orbit.

——————————————————————————————————

11.4.4 The Lifshitz-Kosevich Formulae


The de Haas - van Alphen Oscillations in the magnetization M can be found
from the grand canonical potential Ω
X   
Ω = − kB T ln 1 + exp − β ( Eα − µ ) (1534)
α

481
where the sum over α runs over all the one-electron states.

The Lifshitz-Kosevich formulae159 describes the oscillatory parts of M . This


shall be examined in the T → 0 limit. In the limit T → 0, one has
X  
lim Ω = Eα − µ Θ( µ − Eα ) (1535)
T → 0
α

where Θ(x) is the Heaviside step function. Also, the total number of electrons
is given by X
Ne = Θ( µ − Eα ) (1536)
α

The dispersion relation for free electrons in an applied field is given by

h̄2 kz2
 
1
Eα = + n + h̄ ωc − µB Hz σ (1537)
2m 2
so
n=∞ Z ∞  2 2   
| e | Hz V X X h̄ kz 1
Ω = dkz + n + h̄ ωc − µB Hz σ − µ
4 π 2 c h̄ σ n=0 −∞ 2m 2
h̄2 kz2
 
1
×Θ µ − − (n + ) h̄ ωc + µB Hz σ
2m 2
(1538)

For fixed n, the step function has the effect that the kz integration is limited to
the range of kz values, kz (σ, n) > kz > − kz (σ, n) , where

h̄2 kz (σ, n)2


 
1
= µ − n + h̄ ωc + µB Hz σ (1539)
2m 2

The kz integration can be performed yielding


  12 X n=∞   ! 32
4 | e | Hz V 2m X 1
Ω = − µ − n + h̄ ωc + µB Hz σ
3 4 π 2 c h̄ h̄2 σ n=0
2
 
1
×Θ µ − (n + ) h̄ ωc + µB Hz σ
2
(1540)

Thus, the summation over n only runs over a finite range of values, where n
runs from 0 to n+ , where n+ denotes the integer part of
µ + µB Hz σ 1
n+ = − (1541)
h̄ ωc 2
159 I. M. Lifshitz and A. M. Kosevich, Sov. Phys. J.E.T.P. 2, 636 (1956).

482
Hence,
 12 X n=n ! 23
4 | e | Hz V

2m X+ 
1

Ω = − µ − n + h̄ ωc + µB Hz σ
3 4 π 2 c h̄ h̄2 σ n=0
2
(1542)

The thermodynamic potential shows oscillatory behavior as H increases, since


when h̄ µωc changes by an integer the upper limit of the summation over n also
changes by an integer.

In order to make the oscillatory nature of the summation more explicit, a


periodic function β(x) is introduced. The periodic function is defined as
n=+∞  
X 1
β(x) = δ x − (n + ) (1543)
n=−∞
2

The summation over n in the thermodynamic potential can be expressed in


terms of an integral over β(x) via

 12 X Z ! 32
 x+
4 | e | Hz V 2m
Ω = − dx β(x) µ − x h̄ ωc + µB Hz σ
3 4 π 2 c h̄ h̄2 σ 0

(1544)

where the upper limit of integration is given by


µ mc
x+ = + µB σ (1545)
h̄ ωc | e | h̄

However, as
| e | h̄
µB = (1546)
2mc
the upper limit of integration becomes
µ σ
x+ = + (1547)
h̄ ωc 2
in which the mass of the electron has cancelled in the second term. In general,
the spin splitting term will depend on the ratio of the mass of the electron to
the band mass.

On Fourier analyzing β(x), one has



X 1
β(x) = 1 + 2 cos 2πp ( x − ) (1548)
p=1
2

483
which on substituting into the expression for Ω yields the expression
 1 "   52
4 | e | Hz V 2m 2 X 2
Ω = − µ + µB σ Hz
3 4 π 2 c h̄ h̄2 σ
5 h̄ ωc
∞ Z
! 23 #
x+
X 1
+2 dx cos 2πp ( x − ) µ − x h̄ ωc + µB Hz σ
p=1 0 2
(1549)

The first term is non-oscillatory. The term containing the summation produces
the oscillatory terms. The second term can be evaluated by integration by parts
Z x+  3
1 µ µB Hz σ 2
Ip = dx cos 2πp ( x − ) − x +
0 2 h̄ ωc h̄ ωc
Z x+    23
dx d 1
= sin 2πp ( x − ) x+ − x
0 2 π p dx 2
Z x+   12
3 dx 1
= sin 2πp ( x − ) x+ − x
2 0 2πp 2
(1550)

since the boundary term vanishes. Integrating by parts once again, leads to
Z x+    21
3 d 1
Ip = − dx cos 2πp ( x − ) x+ − x
8 π 2 p2 0 dx 2
" Z x+   − 12 #
3 1 1 1
= x+2 cos πp − dx cos 2πp ( x − ) x+ − x
8 π 2 p2 2 0 2
(1551)

The integral is performed by changing variables from x to u, where u is defined


as
u2
 
2 p x+ − x = (1552)
2
so that the infinitesimal quantity dx is given by
1
dx = − u du (1553)
2 p
In terms of the new variable, u, the integration becomes
" Z u0  #
3 1 1 π
Ip = x+2 cos πp − √ du cos ( u20 − u2 − 2 p )
8 π 2 p2 4p 0 2
(1554)

484
The cosine term can be decomposed as

π u2 π u2 π u2
cos ( + φ ) = cos cos φ − sin sin φ (1555)
2 2 2
so one has the integrals
x
π u2
Z
C(x) = du cos
0 2
x
π u2
Z
S(x) = du sin (1556)
0 2
which for large x have the limits
1
C(∞) = S(∞) = (1557)
2
Thus, the integral Ip is evaluated as
"
3 1
Ip = 2 2
x+2 cos πp +
8π p
 #
1 π 2 π 2
− √ C(u0 ) cos ( u0 − 2 p ) + S(u0 ) sin ( u0 − 2 p )
4p 2 2
" #
3 1 1 π 2 1
∼ x+ cos πp − √
2
cos ( u0 − 2 p − )
8 π 2 p2 8p 2 2
(1558)

Thus, the oscillatory part of the grand canonical potential ∆Ω is


  23  1
| e | Hz V 2m 2
∆Ω ∼ h̄ ωc ×
4 π 4 c h̄ h̄2
∞    
X X 2 µ σ−1 π
× 5 cos 2πp + −
σ p=1 ( 2 p )2 h̄ ωc 2 4
(1559)

This depends on the ratio of the extremal cross-section of the zero field Fermi
surface, AF = π kF2 , and the difference in areas of the Landau orbits in
momentum space, ∆A = 2π |h̄ec| Hz ,
  32 1

| e | Hz V 2m 2
∆Ω ∼ h̄ ωc ×
4 π 4 c h̄ h̄2

c h̄ π kF2
   
X X 2 σ 1 π
× 5 cos 2πp + − −
σ p=1 ( 2 p )2 2 π | e | Hz 2 2 4

485
  32  1
| e | Hz V 2m 2
∆Ω ∼ h̄ ωc ×
4 π 4 c h̄ h̄2
∞    
X X 2 AF σ−1 π
× 5 cos 2πp + −
σ p=1 ( 2 p )2 ∆A 2 4
(1560)

Thus, oscillations in the grand canonical potential occur when the number of
Landau orbits inside the extremal cross-sectional area change. The oscillations
occur in the magnetization Mz as it is related to the grand canonical potential
Ω via  
∂Ω
Mz = − (1561)
∂Hz
Thus, the magnetization also has oscillations that are periodic in H1z . Fur-
thermore, for a free electron gas the extremal area of the Fermi surface is just
AF = π kF2 , so the period of oscillations is proportional to the extremal cross-
sectional area of the zero field Fermi sphere. In addition to the fundamental
oscillations, there are also higher harmonics which can be observed in experi-
ments. For the more general situation, where the Fermi surface is non-spherical
different extremal cross-sections will be observed when the magnetic field is
applied in different directions. Measurements of the de Haas - van Alphen os-
cillations can be used to map out the Fermi surface.

The Lifshitz-Kosevich formulae160 , valid at finite temperatures, yields the


contribution from one extremal area of the Fermi surface to the oscillatory part
of the grand canonical potential as
 
π p
  32 exp − ωc τ
| e | Hz kB T V X X 1
∆Ω = 1 3 2
2 π h̄ c ( 2 π )2 σ p p
2 sinh 2π h̄p ωkcB T
   
AF σ m∗ π
× cos πp cos 2πp + −
∆A 2 me 4
(1562)
1
For a Fermi surface of general shape, the factor ( 2 π ) 2 in the denominator is
to be identified as the magnitude of the second derivative of the cross-sectional
area of the Fermi surface evaluated at the extremal area. In the case pertinent
to free electrons, the cross-sectional area is given by A(kz ) = π ( kF2 − kz2 ),
therefore,
1
∂ A(kz ) − 2
2
1

∂k 2
= ( 2 π )− 2 (1563)
z
160 I. M. Lifshitz and A. M. Kosevich, Sov. Phys.-JETP 2, 636 (1956).

486
On performing the sum over the spin polarizations, one obtains the result
 
π p
 3 exp − ωc τ
| e | Hz 2 2 kB T V X 1
∆Ω = 1 3 2
2 π h̄ c ( 2 π )2 p p
2 sinh 2π h̄p ωkcB T
     
m∗ AF π
× cos πp cos πp cos 2πp −
me ∆A 4
(1564)

The spin splitting factor depends on the ratio of the band mass of the electron to
the electron mass in vacuum. The splitting between the up-spin and down-spin
bands has modified the relative phase of the higher harmonics in the oscillations.
The interference between the oscillations from the up-spin and down-spin bands
may even result in the suppression of the fundamental oscillations. For systems
which are on the verge of ferromagnetism, the spin splitting factor should be en-
hanced by including the effective field on the spins due to the interactions with
the other electrons. This formula also includes the exponential damping of the
oscillations due to T through the thermal smearing of the Fermi surface and also
has an exponential damping term depending on the rate for elastic scattering
by impurities τ1 . Both these effects reduce the amplitude of the de Haas - van
Alphen oscillations161 . The oscillations can only be seen at low temperatures
T < 1 K and for samples of high purity, as indicated by small residual resis-
tances. The oscillations are only seen in materials where the zero temperature
limit of the resistivity ρ(0) is less than 1 µΩ cm. The term involving the lifetime
comes from the width of the quasi-particle spectrum, and should also be accom-
panied with the change in quasi-particle energy due to interactions. Therefore,
the increase in the quasi-particle mass can also be extracted from the amplitude
of the de Haas - van Alphen oscillations. However, the amplitude of the heavier
mass bands are small compared with the light quasi-particle bands. In the heavy
fermion materials such as CeCu6 and U P t3 quasi-particle masses of about 200
free electron masses have been observed in de Haas - van Alphen experiments162 .

11.4.5 Geometric Resonances


There are many other probes of the Fermi surface, these include the attenuation
of sound waves163 . Consider sound waves in a crystal propagating perpendicular
to the direction of the applied magnetic field and having a transverse polariza-
tion that is also perpendicular to Hz . The motion of the ions is accompanied
by an electric field of the same frequency, wave vector and polarization. The
electrons interact with the sound wave through the electric field. If the wave
161 R. B. Dingle, Proc. Roy. Soc. A 211, 257 (1952).
162 S.Chapman, M. Hunt, P. Meeson, P. H. P. Reinders, M. Springford and M. Norman, J.
Phys. CM. 2, 8123 (1990), L. Taillefer, R. Newbury, G. G. Lonzarich, Z. Fisk and J. L. Smith,
J. Mag. Mag. Mat., 63 & 64, 372 (1987).
163 H. W. Morse and J. D. Gavenda, Phys. Rev. Lett. 2, 250 (1959).

487
length of the mean free path is sufficiently long, i.e. ωc τ  1, the attenuation
of the sound waves can be used to determine the Fermi surface.

The electrons follow real space orbits which have projections in the plane
perpendicular to Hz , which are just cross-sections of the constant energy surfaces
in momentum space, but are scaled by | eh̄| cHz and rotated by π2 . As velocities
of the ions are much smaller than the electrons’ velocities, the electric field may
be considered to be static. If the phonon wave vector q is comparable to the
radius of the real space orbit or, more precisely, the diameter of the orbit in
the direction of q, then the electric field can significantly perturb the electron’s
motion. This strongly depends on the mismatch between q −1 and the diameter
of the orbit164 . When the radius of rc the orbit is such that
λ
2 rc = (1565)
2
then the electron may be accelerated tangentially by the electric field at both
extremities of the orbit. The coupling is coherent over the electron’s orbit and
the coupling is strong. When the condition

2 rc = λ (1566)

is satisfied, the electron is sequentially accelerated and decelerated by the field.


The coupling is out of phase on the different segments of the electron’s orbit so
that the resulting coupling is weak. In general the condition for strong coupling
is that of constructive interference
1
2 rc = ( n + )λ (1567)
2
and weak coupling occurs when the interference is destructive

2 rc = n λ (1568)

The period differs slightly from the asymptotic large n variation just described.
Assume that the projection of the electron’s trajectory on the plane perpendic-
ular to the applied field Hz is circular. The energy transfer between the electron
and the electronic wave in one orbit is given by
Z 2π
ωc
∆E = dt E(r, t) . v(t) (1569)
0

The electric field is assumed to be polarized along the y direction, and q is


directed along the x direction. Since the orbit in momentum space is rotated
by π2 with respect to the real space orbit, one has

vy (t) = v sin ωc t (1570)


164 A. B. Pippard, Phil. Mag. 2, 1147 (1957).

488
and
x(t) = rc sin ωc t (1571)
Thus, the energy transferred in one period is evaluated as
Z ω2π  
c
∆E = Ey v dt exp i q rc sin ωc t − ωq t sin ωc t
0 

∼ Ey v J1 ( q rc ) (1572)
ωc
since, in the limit that the phonon can be considered static, ωq → 0, so only the
term with the n = 1 Bessel function need be considered. Hence, the absorption
of transverse sound waves exhibits geometric resonances. The resonances occur
for phonon wave lengths which match the maxima of the Bessel function J1 (x).

Only electrons near the Fermi surface can absorb energy from the sound
wave. This is caused by the Pauli exclusion principle. The Pauli exclusion
principle forbids electrons in states far below the Fermi energy to undergo low-
energy excitations, since the states with slightly higher energy are completely
occupied. The strength of the geometric resonances are determined by the dis-
tribution of the z component of the electron velocity on the Fermi surface. The
electrons with the extremal diameter on the Fermi surface are more numerous
and, therefore, play a dominant role in the attenuation process. Thus, the sound
wave may display an approximately periodic variation in λ where the asymptotic
period is determined by  
1 1
∆ = (1573)
λ 2 rc
By studying the attenuation of sound with different wave vectors, q, and differ-
ent magnetic fields H one can map out the Fermi surface165 .

11.4.6 Cyclotron Resonances


This experimental method of probing the Fermi surface requires the application
of an microwave electric field at the surface of a metal. The field is attenuated
as it penetrates into the metal, and is only appreciable with a skin depth δ from
the surface. Since the field does not penetrate the bulk, electrons can only pick
up energy from the field when they are within the skin depth from the surface.

In one experimental geometry, a static (d.c.) magnetic field is applied par-


allel to the surface, say in the x direction, so that the electrons undergo spiral
orbits in real space. The x component of the electron’s velocity, vx , remains
constant, but the electrons undergo circular motion in the y − z plane. It is
only necessary to consider the electrons that travel in spirals that are close and
parallel to the surface, as it is only these electrons that couple to the microwave
165 M. H. Cohen, M. J. Harrison and W. A. Harrison, Phys. Rev. 117, 937 (1960).

489
field. The size of the orbit and the electron’s mean free path λ should be much
larger than the skin depth δ. This holds true when the cyclotron frequency ωc
is large, and for microwave frequencies ω for which the anomalous skin depth
phenomenon occurs. The condition of a long mean free path and large cy-
clotron frequency is necessary for the electrons to undergo well defined spirals,
so ωc τ  1.

The electrons pick up energy from the field only if they are within δ of the
surface. The electrons in the spiral orbits only experience the electric field each
time they enter the surface region. They enter the skin depth periodically, with
period

TH = (1574)
ωc
which is the period of the cyclotron motion. In general, the period is given by
the expression
h̄2 c
 
∂A
TH = (1575)
| e | Hx ∂E
where A(E) is the area of the surface of constant energy in reciprocal space.

The electron will experience an E field with the same phase, if the applied
field has completed an integral number of oscillations during each cyclotron
period. If the frequency of the a.c. field is ω, then the period of the a.c. field
TE is given by

TE = (1576)
ω
For the a.c. field to act coherently on the electron, the periods must be related
via

TH = n TE = n (1577)
ω
for some integer n. Hence, this condition requires that the frequency of the
cyclotron orbit match with the frequency of the a.c. electric field
ω = n ωc (1578)
so that the a.c. field resonates with the electronic motion in the uniform field.
The resonance condition can be written as
1 n
= 2 π | E | h̄2 c ω   (1579)
Hx ∂A
∂E

The factor
h̄2
 
∂A
mc = (1580)
2π ∂E
is known as the cyclotron mass. For free electrons, A(E) = π k 2 and E =
h̄2 k2
2 m , so the cyclotron mass coincides with the electron’s band mass

mc = m (1581)

490
1
If the microwave absorption is plotted versus Hx a series of uniformly spaced
resonance peaks should be found.

The calculation of the absorption is the simplest in the case when the wave
length of the electromagnetic field λ is much larger than the cyclotron orbit and
ωc τ  1.

Consider the geometry in which the surface has its normal in the z direction
and the d.c. magnetic field is applied parallel to the surface in the x direction
H = êx Hx (1582)
The a.c. electric field is applied in the y direction, so the oscillatory dependence
is given by  
E = êy Ey exp i(qz − ωt) (1583)

for | z |  δ. The electron in its orbit experiences a rapidly alternating electric


field. For most values of z the contributions cancel. The cancellation only fails
at the extremal values of z where the z component of the velocity falls to zero.
In this case, the electron velocity is entirely parallel to the planes of constant z.

To the zero-th order approximation, the z component of the electron’s posi-


tion can be expressed as
v
z(t) = z0 + sin ωc t (1584)
ωc
The total change in momentum of the electron due to the oscillating field, in
one period, can be calculated in the semi-classical approximation. The change
of momentum and, hence, the current will be in the direction of the a.c. field.
The impulse imparted to the electron is given by the integral of the electric field
evaluated at the electron’s position
Z t+TH   
v
h̄ δky = | e | dt0 Ey exp i q z0 + q sin ωc t0 − ω t0 (1585)
t ωc
In general, the integral can be evaluated with the aid of the identity
  n=∞
X  
exp i x sin θ = Jn (x) exp i n θ (1586)
n=−∞

where Jn (x) is the Bessel function of order n. When the frequencies precisely
satisfy the resonance condition, the integral is given by the n-th order Bessel
function. However, for large qωcv and with an arbitrary frequency mismatch, the
integral can be performed by the method of steepest descents. This yields to
the asymptotic form of the impulse given by
  12   
2π π
h̄ δky ∼ − | e | Ey exp i q z0 − ω t − (1587)
q v ωc 4

491
Due to the variation of the phase of the field experienced by the electron as it
performs its orbit, only a fraction
  12
2 π ωc
(1588)
qv

of the orbit contributes to the integral. The energy gain of the electron in one
traversal is given by
  12   
2πv π
v h̄ δky = − | e | Ey exp i q z0 − ωt − (1589)
q ωc 4

The previous traversal caused a similar displacement, but with t → t − 2ωcπ .


However, only the fraction  

exp − (1590)
ωc τ
of electrons survive traversing one cyclotron orbit without scattering. Thus,
the contribution to the average increase in the energy of the electron from the
previous orbit is given by
 

∼ v h̄ δky exp − (1 + iωτ ) (1591)
ωc τ

As the contributions from all the previous orbits all have similar forms, one can
obtain the average energy displacement experienced by an electron between one
scattering and the next
  12   
2πv π
v h̄ ∆ky = − | e | Ey exp i q z0 − ω t − F (1592)
q ωc 4

The factor F reflects the sum of the probabilities that the electron survive n
orbits without scattering. The value of F is given by the sum of the contributions
from each orbit
∞  
X 2πn
F = exp − (1 + iωτ )
n=0
ωc τ

1
=   (1593)
2 π
1 − exp − ωc τ (1 + iωτ )

When the contributions from the previous orbits survive and are in phase, there
is resonant absorption of the electromagnetic radiation. The resonance is deter-
mined by the denominator of F , and is experimentally manifested by resonances
in the surface impedance166 .

166 M. Ia. Azbel’ and E. A. Kaner, Sov. Phys. J.E.T.P. 5, 730 (1957).

492
To calculate the current at a depth z0 , one must examine the orbits on the
Fermi surface. The orbits circulate around the Fermi surface in sections that
are perpendicular to the d.c. field. Thus, the orbits are in the y − z plane. The
portions of the Fermi surface orbits which contribute most to the current are
those in which the electrons are moving parallel to the surface. These portions
of the orbits are those where kz = 0, and form the effective zone. An electron
on the effective zone has kz = 0 and, thus, is moving at the extreme of its orbit.
Due to the effect of the a.c. field, this orbit has been displaced by a distance
∆ky from the orbit performed by the electron in the absence of the a.c. field.

The total current from the orbits in a section of width dkx , around kx , can
be calculated by considering the contribution of the electrons around the orbit.
Due to the phase differences around the orbits only those within a distance
  12
kx 2 qπ vωc of the effective zone contribute. These orbits are displaced from
their equilibrium positions by an amount ∆ky . Only the displacements from
equilibrium contribute to the current. The contribution to the current density
is
 1  
2 2 π ωc 2 π
δJy = − | e | dk x k x exp i ∆ky v
8 π3 qv 4
e2
 
1
= dk k
x x F Ey (1594)
2 π2 h̄ q
where the phases of π4 cancel. The integration over dkx can be converted into an
integral over the effective zone via the polar angle ϕ. If the effective mass and
cyclotron frequency are constant over the Fermi surface, then F is also constant.
The conductivity σ is proportional to F . When ωc τ  n, the conductivity
exhibits resonances at n ωc = ω.

The surface impedance is defined as


4 π i ω Ex
Z(ω) =   (1595)
∂Ex
∂x

This can be crudely estimated from the theory of the (normal) skin depth, as
 1
2πω 2
Z(ω) ∼ ( 1 − i )
σ
1
∼ F−2 (1596)
A more rigorous calculation167 requires the use of the theory of the anomalous
skin depth, and yields the result
√ 2 1
Z(ω) ∼ ( 1 − i 3 ) ω 3 F − 3 (1597)
167 L.E. Hartmann and J.M. Luttinger, Phys. Rev. 151, 430 (1966).

493
However, the conclusion remains that the surface impedance shows oscillations
with varying field strength Hx . It should be obvious from the above discussion
that the oscillations provide information on ωωc , and that the dominant contri-
bution occurs from the extremal parts of the line of intersection of the Fermi
surface with the plane kz = 0. Thus, the cyclotron resonance can be used to
study points on the Fermi surface168 .

The other experimental geometry involves the static magnetic field being
applied along the normal to the surface. The cyclotron orbits are in the plane
parallel to the surface. Since the a.c. electric field is parallel to the orbit, the
resonance condition involves the Doppler shift169 .

11.5 The Quantum Hall Effect


The Quantum Hall Effect is found in quasi two-dimensional electron systems.
Experimentally, electrons can be confined to a two-dimensional sheet in a metal
oxide semiconductor field effect transistor. The application of a strong elec-
tric field perpendicular to the surface may pull down the conduction band at
the surface of the semiconductor. The energy eigenvalue equation can be sep-
arated into parts describing motion parallel and perpendicular to the surface.
The wave function is factorized into two parts. One part describes the motion
parallel to the surface which can be approximated by a two-dimensional Bloch
function which corresponds to an eigenvalue in a continuum. The other fac-
tor describes one-dimensional motion of electrons which are confined near the
surface by the applied field and corresponds to discrete energy levels. If the
states which correspond to the lowest discrete energy level are occupied, the
electrons are confined in the vicinity of the surface and the system corresponds
to a two-dimensional electron gas. If the energy levels are lower than the Fermi
energy in the metal, these states are occupied by electrons which tunnel from the
metal, across the insulating oxide barrier into the semiconductor. After equi-
librium has been established, the electrons at the surface of the semiconductor
form a two-dimensional electron gas. The Quantum Hall Effect occurs in the
two-dimensional electron gas when a magnetic field is applied that is so strong
that the mixing of the Landau levels by disorder or electron-electron interac-
tions is negligible. The Integer Quantum Hall Effect occurs when the disorder is
stronger than the electron-electron interactions. The Fractional Quantum Hall
effect occurs when the effect of electron-electron interactions are greater than
the effects of disorder.

168 T. W. Moore and F. W. Spong, Phys. Rev. 125, 846 (1962).


169 P. B. Miller and R. R. Haering, Phys. Rev. 128, 126 (1962).

494
11.5.1 The Integer Quantum Hall Effect
The Integer Quantum Hall Effect can be understood entirely within the frame-
work of non-interacting electrons. The calculation of the Hall coefficient can
be performed using the Kubo formula. However, in applying the Kubo for-
mulae, one must recognize that the vector potential has two components: an
a.c. component responsible for producing the applied electric field and a second
component which produces the static magnetic field. The usual derivation only
takes the a.c. component of the vector potential into account as a perturbation.
The d.c. component of the vector potential must be added to the paramagnetic
current operator, yielding the electron velocity operator appropriate for the sit-
uation where the weakly perturbing electric field is zero.

The application of a static magnetic field Bz perpendicular to the surface


will quantize the motion of the electrons parallel to the surface. The motion
parallel to the surface is quantized into Landau orbits, and the energy eigenvalue
equation reduces to the energy eigenvalue equation of a one-dimensional simple
harmonic oscillator. Due to the confinement in the direction perpendicular to
the surface, kz will not be a good quantum number, and the perpendicular
component of the energy will form highly degenerate discrete levels n . For
large enough fields only the lowest level 0 will be occupied. On choosing a
particular asymmetric gauge for the vector potential,

A(r) = + êy Bz x (1598)

the Hamiltonian for the two-dimensional motion in the x − y plane is given by


2
p̂2x m ωc2

Ĥ = + x̂ − X̂ (1599)
2m 2
where the cyclotron frequency ωc is given by
| e | Bz
ωc = (1600)
mc
The operator X̂ is given in terms of the y-component of the momentum operator
p̂y . Since the Hamiltonian is independent of y, both py and X can be taken as
constants of motion. The momentum component p̂x is canonically conjugate to
the x component of the particle’s position relative to the center of the orbit
p̂y c
x̂ − X̂ = x + (1601)
| e | Bz
The energy eigenvalues of the shifted Harmonic oscillator are given by
1
Eν,0 = E0 + h̄ ωc ( ν + ) (1602)
2
where ν is the quantum number for the Landau levels. The Landau levels are
independent of ky and, therefore, are degenerate. Due to the periodic boundary

495
conditions, the values of ky are quantized by

2 π ny
ky = (1603)
Ly

for a surface of length Ly . Since the centroid of the wave function is constrained
by
Lx > X > 0 (1604)
the possible value of ky or ny are restricted by

2 π h̄ c ky
Lx > > 0 (1605)
| e | Bz

Hence, the total number of degenerate states is given by

Bz | e |
D = Lx Ly
2 π h̄ c
Φ|e|
= (1606)
2 π h̄ c
where Φ is the total magnetic flux passing through the sample. The fundamental
flux quantum Φ0 is defined as the quantity
2 π h̄ c
Φ0 = (1607)
|e|

Therefore, the degeneracy is equal to the number of flux quanta. The density
of states can be approximately expressed as a discrete set of delta functions
 
| e | Bz Lx Ly X 1
ρ() = δ  − E0 − h̄ ωc ( ν + )
hc ν
2
 
m ωc Lx Ly X 1
= δ  − E0 − h̄ ωc ( ν + )
2 π h̄ ν
2
(1608)

where each delta function corresponds to a Landau level. The weight associated
with each delta function corresponds to the degeneracy of each Landau level.
On defining the cyclotron radius rc by
s
h̄ c
rc = (1609)
| e | Bz

one finds that the relative position operator can be expressed as


 
rc †
x̂ − X̂ = √ aky + aky (1610)
2

496
and from the Heisenberg equation of motion for x̂, one finds that the x-component
of the velocity is given by
 
1
v̂x = x̂ , Ĥ
i h̄
 
rc ω c †
= i √ aky − aky (1611)
2
as p̂y has been diagonalized. The y-component of the velocity is found from the
Heisenberg equation of motion for ŷ and is given by
 
1
v̂y = ŷ , Ĥ
i h̄
= ωc ( x̂ − X̂ ) (1612)

= − | ie h̄| Bc z . On substituting for the x-


 
since the commutator ŷ , X̂
component of the displacement from the center of the orbit, one finds
rc ω c
v̂y = √ ( a†ky + aky ) (1613)
2
Hence, v̂x and v̂y do not commute. Furthermore, as the velocity operators are
not diagonal in the quantized Landau level indices, the Landau orbitals do not
carry a net current.

The Kubo formula can be expressed in terms of the electron velocity op-
erators, which includes the diamagnetic current contributions from the static
magnetic field. The Kubo formula for the conductivity tensor, per unit area,
appropriate for non-interacting electrons has the form
"
i e2 1 X
σα,β (ω) = < ν, ky | v̂α | ν 0 , ky > < ν 0 , ky | v̂β | ν, ky > ×
ω + i η Lx Ly
ν,ν 0 ,ky
#
f (Eν ) − f (Eν 0 ) X f (Eν )
× − δα,β
h̄ ω + i η + h̄ ωc ( ν − ν 0 ) m
ν,ky

(1614)

since ky is a constant of motion. On evaluating the Kubo formula170 , one finds


that the diagonal components are zero

<e σx,x (0) = 0 (1615)

but, nevertheless, the off-diagonal components are finite and quantized


e2
<e σx,y (0) = − n (1616)
2 π h̄
170 P. Streda and L. Smrcka, Phys. Stat. Sol., B, 70, 537 (1975), D. J. Thouless, J. Phys. C,

14, 3475 (1981), Q. Niu and D. J. Thouless, Phys. Rev. B, 35, 2188 (1987).

497
Daniel C. Tsui: 2D electron gas in intense magnetic fields

neto-transport coefficients ␳ xx and ␳ xy of a 2DEG FIG. 2. ␳ xx and ␳ xy of a relatively low-mobility 2DEG in


Ga1⫺x As at 0.35 K in moderately lowFigure B. The 221:
in- GaAs/Al
The x Ga1⫺x
Integer As. The plateaus
Quantum in ␳ xy are
Hall Effect in aquantized
GaAs − in GaAlAs
the hetero-
e measurement geometry. The magnetic field B is natural conductance unit e 2 /h with integer quantum numbers
junction at T = 66 mK [data taken by H.P. Wei]. The plateau indices are (from
r to the plane of the 2DEG and to the i⫽1,2, . . . . Data taken by H. P. Wei.
thecurrent
right)I. 1, 2, 3, 4, 6, 8... The odd-indexed plateaus can only be resolved at
V and V H are measured along and perpendicular
high fields. [After
vely, ␳ xx ⫽(V/L)/(I/W) is the resistance across a ferentD.C. Tsui, Rev.
physical Mod.
regimes. Phys.
The first71,
is 892
the (1999).]
disorder-
endent of the square size, and ␳ xy ⫽V H /I is the dominant regime, when the sample is dirty with low 2D
ce independent of the sample width. Data taken electron mobility (e.g., ␮ ⬍105 cm2/V sec in the case of
dar. where (n − 1)GaAs). is the quantum
The strikingnumber forinthe
features thehighest occupiedthe
data constitute Landau orbital.
That is, the off-diagonal
integral quantum conductivity
Hall effectis(IQHE)
quantized(vonand the etquantum
Klitzing al., number
gy, separating the filled from the corresponds
empty lev- to1980),the total
whichnumber, n, of occupied
is understood in terms Landau levels. of
of the physics Since the diag-
ommensuration energy. I assumedonal he meant:
component independent electrons and
of the conductivity their localization
vanishes, the current in flow
the pres-
is dissipationless.
␸ extreme quantum limit, at commensuration ence of random impurities in the semiconductors.
As the field is changed, the peaks in the density of states associated The with the
n ␸ ⫽1/i, some interaction energyLandau
might be-levels sweep through the Fermi level. The Hall resistivityinshould, there-
fractional quantum Hall effect (FQHE) is observed
nant to drive the 2D system into fore,someshow
new a sethigh-mobility
of steps assamples in the field
the applied regime where the electron-
is increased. This phenomenon is
e. I was not brave enough to ask him: ‘‘What electron interaction dominates. It manifests the many-
the integer quantum Hall effect.
n?’’ But I felt affirmed that I should continue body interaction physics of the 2DEG in the intense B
ate on the extreme quantum limit. field. Furthermore, even in the cleanest samples, the
with the advent of molecular beam epitaxy Experimentally,
FQHEitseriesis found that the
terminates intosteps in σx,y (0)
an insulator arehigh-B
in the not discontinuous,
and the invention of modulationbut instead
doping to showlimit.a This
finiteinsulator
slope inisthe transition
believed to be region. Furthermore,
an electron crystal the diag-
ghly perfect 2D electron systemsonal component
(Stormer of the
pinned byresistivity
defects to isthenot zero in the region
semiconductor. where
The third re-the transition
it soon became quite clear to Art Gossard,successive
between gime isplateaus ␮ and high-B-field
this high-occur, but shows limit,spikeswhere
there.disorder
This phenomenon
mer, and me that where we wanted is to go to
associated and interaction
with impurity play equally important
scattering. The effectroles and need tois to broaden
of impurities
w many-body interaction physics should
the setbeofa delta
be function
treated on an equal
peaks in thefooting.
density of states into a set of Gaussians.
bility 2DEG sample placed in a most intense
This allows the transition between the plateaus to be continuous. In fact, if
ld.
all the states QUANTUM
contributed to the
PHASE conductivity,
TRANSITIONS the steps of the staircase would
IN IQHE
be smeared out into a straight line, just like in the Drude theory for three-
SIONAL MAGNETO-TRANSPORT dimensional metals. However,
Quantization of the
the states in the tails
Hall resistance of natural
in the each Gaussian
con- are local-
ized171 , as theductance
deviationunit
frome 2 /h
theisideal
currently
Landau understood in terms
level energy of
indicates that these
esence of a perpendicular magneticstates the the existence
field, experience of an energy
a larger impurity gap, separating
potential than average. the excited
The large potential
s of a two-dimensional electron collapse, as
acts to localize states from the
the states withground state,inand
energies the localized
Gaussianstates inside
tail and so these states
Landau quantization of its cyclotron orbits, the gap. In the IQHE case, where the quantum numbers
do not contribute to the Hall conductivity. In fact, in two-dimensions with zero
e Landau levels separated by the cyclotron are integers identified with the number of completely
field, one can filled
ntum. Scattering broadens the Landau levels
showLandau
that alllevels,
the states are localized
the energy gap is theinLandau
an infinite
gap ofsample. How-
se to 2D magneto-transport described byLevine,
171 H. the S.a B.
cyclotron energy
Libby and A. M.quantum. ThePhys.
M. Pruisken, accurate
Rev. quantization
Lett. 51, 1915 (1983).
ura theory (Ando and Uemura, 1979). Fig- was shown by Laughlin, using a gedanken experiment,
example showing the quantum oscillations in to be a consequence of charge quantization. He showed
al resistivity ␳ xx , reflecting the broadened that the experiment in effect measures the charge car-
el structure of the 2DEG, and the Hall resis- 498 The localized states arise
ried by the excited electron.
well known from the Drude model. However, from disorder in the 2D system, and the data, as shown
DEG is taken to the extreme condition of in Fig. 2, show the localization-delocalization phase tran-
low T, much more striking features appear, sitions. In other words, for B in the plateau regions, the
interplay of disorder and electron-electron states at E F are localized, and in between, delocalized.
in the system. More specifically, different As T is decreased, the range of B for the existence of
nomena are observed in three distinctly dif- delocalized states decreases and the transition regions
1.5

ρ(ε) 0.5

0
-3 -1 1 3
ε

Figure 222: The schematic density of states for a two-dimensional electron gas
in a high-magnetic field. Due to the effects of disorder, the states in the tails
of the Landau levels are localized (shaded). Although the region of occurrence
of localized states appears similar to that found in Anderson Localization in
two-dimensions, the effect of the magnetic field breaks time-reversal invariance
and, therefore, yields important differences.

ever, the samples are finite and have edges. The edges have extended states that
carry current. The edge states can be understood by analogy with the classi-
cal motion where there are skipping orbits. In a classical skipping orbit, the
cyclotron orbits are reflected at the edges. The classical skipping orbits would
produce oppositely directed currents at pairs of edges. Quantum mechanically,
the bulk states do not contribute to the current since the velocity operator for
a particle with charge q, is given by
p̂y q
v̂y = − Ay
me me c
= ωc ( x̂ − X̂ ) (1617)
and the probability density for the shifted harmonic oscillator is symmetrically
peaked about X. Since the current carried by individual particles in the bulk
states are given by integrals which are almost anti-symmetric
Z Lx
jky ,ν = q ωc dx | φν ( x − X ) |2 ( x − X ) (1618)
0

the bulk contributions to the current vanish. On the other hand for X close
to the boundary, say at X = 0, then the wave function must vanish at the
boundary for a hard core potential and so the current is given by
Z Lx
jky ,ν = q ωc dx | φν ( x ) |2 x (1619)
0

499
V
Br

Φ

Figure 223: The geometry of Laughlin’s gedanken experiment which identifies


the quantum of Hall conductance with the electric charge. A magnetic flux Φ
threads through the axis of the cylinder. On increasing the magnetic flux by
the flux quantum Φ0 , there is a net transfer of one electron from one end of the
cylinder to the other.

Since the wave function is cut off at x = 0, the integral is positive and the
edge state carries current. The other edge state carries an oppositely directed
current. The presence of the confining potential also lifts the degeneracy of the
states in the Landau levels, by increasing the energy of the states close to the
boundary. As the wave function of the odd order excited state Landau levels
of the homogeneous system vanish at x = X, one finds that the energy of the
Landau levels with hard wall confining potentials increases from ( ν + 12 ) h̄ ωc
to ( 2 ν + 32 ) h̄ ωc as X → Lx . The increase in the Hall resistivity only occurs
when the Fermi level sweeps through the itinerant or delocalized portions of the
density of states.

As the above calculations completely neglects the effect of impurities and


localization, Laughlin172 proposed a gauge theoretic argument which overcomes
these shortcomings. Laughlin envisaged an experiment in which the two-dimensional
sample is in the form of the surface of a hollow cylinder of radius R and length
Lz . The axis of the cylinder is taken to be along the z-direction. A uniform
magnetic field Br is arranged to flow through the sample in a radial direction.
Locally, this field is perpendicular to the plane in which the electrons are con-
fined. A second field is arranged to thread through the cylinder parallel to its
axis, but is entirely contained inside the hollow and falls to zero at r = R.
This field does not affect the motion of the electrons directly since it is zero
inside the sample. However, the associated flux threading through the cylinder
172 R. B. Laughlin, Phys. Rev. B, 23, 5632 (1981).

500
Φ does lead to a finite vector potential AΦ which satisfies

∇ ∧ AΦ = 0 (1620)

Therefore, the vector potential can be written in the form

AΦ = ∇ Λ (1621)

This vector potential does cause the wave functions of the charged particles to
acquire an Aharonov-Bohm phase factor of
 
q
exp − i Λ (1622)
h̄ c

For the vector potential


Φ
AΦ = êϕ (1623)
2πr
one finds that
Φ
Λ = ϕ (1624)

and, hence, the phase factor is given by
 

exp − i ϕ (1625)
2 π h̄ c

Thus, on traversing a singly connected path around the cylinders axis, the
Aharonov-Bohm flux changes the extended state wave function by a factor of
 

exp − i 2π (1626)
2 π h̄ c

Since the fundamental flux quantum Φ0 is defined by


2 π h̄ c
Φ0 = (1627)
|e|

the Aharonov-Bohm factor for electrons can be written as


 
Φ
exp i 2π (1628)
Φ0

Thus, if Φ is an integer multiple of Φ0 , i.e.,

Φ = µ Φ0 (1629)

the extended wave functions are single-valued.

The presence of the perpendicular field B r within the sample quantizes the
motion into Landau levels. These states may either be localized or may be

501
extended throughout the sample. The vector potential at position z is purely
tangential and is given by
 
Φ
A = Br z + êϕ (1630)
2πR
If the vector potential is absorbed into the wave function by gauge transforma-
tion, involving the factor
   
Φ ϕ
exp i z R Br + (1631)
2π Φ0
The flux quantization condition is found by requiring that the wave function to
be single-valued. Hence,
 
2 π z R Br + Φ = µ Φ0 (1632)

must be an integer multiple of Φ0 . If Φ is adiabatically increased by ∆Φ = Φ0 ,


then the extended states in a Landau level must be translated along the z-axis
by amounts ∆z given by
Φ0
∆z = − (1633)
2 π R Br
as the phase of an extended wave function must be single valued. If the flux
through the entire surface is equal to ΦS , then ∆z is given by
Φ0
∆z = − Lz (1634)
ΦS
The phase of the localized states can shift by arbitrary amounts. The presence
of the gap forbids excitation of electrons to states in the higher Landau levels. In
the pure system, the maximum number of electronic states, M , in a Landau level
is given by the total number of the integer valued quantum numbers µ = m
which satisfy eqn(1632) with Lz > z > 0,

2 π R Lz Br = M Φ0 (1635)

Hence, a fully occupied Landau level contains a number


ΦS
M =
Φ0
Lz
= (1636)
| ∆z |
of electrons. Furthermore, for the pure system, the adiabatic increase of Φ by Φ0
results in the transfer of electrons between neighboring extended states. Hence,
in the dirty system, all the delocalized electrons in a Landau level are translated
along the z-direction by one spacing, skipping over the localized states. The net
result is that one electron, per Landau level, is translated across the entire

502
length of the sample. In the absence of an applied electric field, the initial and
final states have the same energy. Thus, by gauge invariance, increasing the
Aharonov-Bohm flux by Φ0 maps the system back on itself. However, if there
is an electric field Ez across the length of the cylinder, this process requires an
energy change of

∆E = − q Ez ( − Lz )
= − | e | Ez Lz (1637)

per Landau level. The current wrapping around the cylinder Iϕ , from all the
electrons in a single Landau level, is given by
∂E
Iϕ = − c (1638)
∂Φ
which leads to a current density

c | e | Ez Lz
jϕ =
Lz Φ0
2
e
= Ez (1639)
2 π h̄
Hence, on summing over all the n occupied Landau levels, one has

e2
jϕ = n Ez (1640)
2 π h̄
In this expression n is the number of completely occupied Landau levels with
extended states, and the Fermi energy is assumed to lie in an energy range
where the bulk states are localized. The Hall conductivity is given by

jy e2
σH = = n (1641)
Ez 2 π h̄
Hence, as long as there are Landau levels with extended states, there is an in-
teger quantum Hall effect.

The integer quantum Hall effect was measured experimentally by von Kl-
itzing173 in 1980. The steps can only be discerned in very clean samples. At
much higher fields, where only the lowest Landau level should be occupied, Gos-
sard, Stormer, and Tsui discovered a similar type of effect174 which is known as
the fractional quantum Hall effect. This phenomenon involves the effect of the
Coulomb repulsion between electrons in the Landau levels. Laughlin showed175
that the energy of the interacting electron states can be minimized by allow-
ing the electrons to form a ground state with a different symmetry from the
173 K. von Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494 (1980).
174 D. C. Tsui, H. L. Stormer and A. C. Gossard, Phys. Rev. Lett. 48, 1559 (1982).
175 R. B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983).

503
886 Horst L. Stormer: The fractional quantum Hall effect

There are many fascinating open questions


with the ␯ ⫽1/2 state, such as: how does the
with energy for CFs? and what is the microsco
ture of the particles? Also, how does the ele
(which we were neglecting throughout this le
fect CF formation? A beautiful picture of com
mions being tiny dipoles is emerging. While o
vortices is placed directly on the electron (P
ciple), the position of the second vortex is a bit
from exact center, rendering the object an elec
in the 2D plane. There is great promise for f
covery and future theoretical insight.

All those other FQHE states

Bose condensation of CBs consisting of elec


an odd number of flux quanta rationalizes th
ance of the FQHE at the primary fraction
Landau-level filling factor ␯ ⫽i⫾1/q with quan
resistances R H ⫽h/( v e 2 ) and deep minima in
comitant magnetoresistance R. However, a mu
FIG. 18. The FQHE as it appears today in ultrahigh-mobility other FQHE states have been discovered over
Figure 224: The Fractional Quantum
modulation-doped GaAs/AlGaAs Hall Effect
2DESs. Many in fractions
an ultra-high
are mobility
Figure 18 shows one of the best of today’s exp
modulation-dopedvisible.
GaAs The − most
GaAlAsprominent2-d electron ␯ ⫽p/(2p⫾1),
sequence, gas. [After H.L.con-Stormer, on a specimen with a multimillion cm2/
traces Rev.
Mod. Phys. 71,verges
874 toward
(1999).] ␯ ⫽1/2 and is discussed in the text. bility. What is the origin of these other states?
posite fermion model offers an extraordinarily
being fermions they are prevented from condensing into ture. We shall discuss it for the sequence of
the lowest state
bulk. Since Laughlin’s energyhas state.a Instead, it can not befractions
they fill up successively
lower symmetry, adiabati- 2/5, 3/7, 4/9, 5/11, . . . and 2/3, 3/5, 4/
cally continuedthe
to sequence
states of of thelowest-lying
ideal Landau energy states,
levels, until a maxi-
without (i.e., ␯ ⫽p/(2p⫾1),
states crossing the p⫽2,3,4 . . . ) around ␯ ⫽
mum is Laughlin’s
reached andgauge all CFs have been accommodated. At half filling the electron system has be
Fermi-energy. Hence, argument does not forbid the possibility
formed into CFs consisting of electrons which
The process is equivalent
of non-integer quantization of the Hall conductance.to the filling of states by elec-
trons at B⫽0. Hence, from the point of view of CFs, the magnetic flux quanta. All of the external mag
␯ ⫽1/2 state appears equivalent to the case for electrons has been incorporated into the particles and t
—————————————————————————————————— in an apparently field-free 2D plane. Since the
at B⫽0. In spite of the huge external magnetic field at
mions, the system of CFs at ␯ ⫽1/2 resembles a
half filling of the Landau level, CFs are moving in a
electrons of the same density at B⫽0. What h
similar fashion to electrons moving in zero field. This
the magnetic field deviates from B⫽0? For
11.5.2 Exercise has been
71 directly observed in experiment. Flux quantum their motion becomes quantized into electro
attachment has transformed these earlier electrons and orbits. They fill up their electron-Landau level
Evaluate the Kubo formula for the real part of the
they are propagating along straight trajectories in a high diagonal and off-diagonal
ter the energy gaps, and exhibit the w
quantum Hall conductivities
magnetic field, where by first taking
normal the limit
electrons would → 0onand IQHE.
ω orbit then tak- CFs around ␯ ⫽1/2 follow the same
→ tight
ing the limit  very circles.
0. Also The mass
estimate theof effects
a CF, usually consideredscattering
of introducing due field deviates from exactly ␯ ⫽1/
the magnetic
to be a property
to random impurities. The effect of theofparticle, is unrelated
the scattering to the can
lifetime massbe introduced
tion of CFs becomes quantized into CF-Land
of the underlying
by including imaginary parts of electron. Instead, of
the energies thethe
mass depends on
occupied They fill up their CF-Landau levels, encou
and unoccupied
the magnetic
single-particle states of the field
form and ± i onlyh̄ on the magnetic field. In
. Choose the signs to ensureenergy
that gaps,
the and exhibit an IQHE. However,
fact, it is a mass of purely2 τmany-particle origin, arising
wave functions for the excited states (with an electron - hole pair) willandecay IQHE of electrons, but an IQHE of CFs. T
to
solely from interactions, rather than being a property of of CFs arises exactly at ␯ ⫽p/(2p⫾1), whic
the ground state after a time τ . Compare your result
any individual particle. It is another one of these baffling with the conductivities
positions of the FQHE features. In fact, the
obtained for theimplications
three-dimensional Drude model.
of e-e interactions in high magnetic fields. features in the magnetoresistance R of th
The absence of condensation and the lack of an energy around ␯ ⫽1/2 closely resemble the oscillatin
gap prevents the ␯ ⫽1/2 state from showing a quantized
—————————————————————————————————— in R around B⫽0 and, once they have been sh
Hall resistance. Instead the Hall line is featureless, just B⫽0 to ␯ ⫽1/2, they coincide with their positi
as it is for electrons around B⫽0 (see Fig. 18). very remarkable in several ways.
The difference between ␯ ⫽1/3 and ␯ ⫽1/2 is striking. CFs ‘‘survive’’ the additional (effective) mag
One is a Bose-condensed504 many-particle state showing a (away from ␯ ⫽1/2), and the orbits of these
quantized Hall effect and giving rise to fractionally particles mimic the orbits of electrons in the
charged particles. The other is a Fermi sea, in spite of magnetic field in the vicinity of B⫽0. The C
the existence of a huge external field, and its particles ‘‘good’’ particles. In this way, a complex elect
have a mass that arises from interactions. One flux quan- particle problem at some rational fractional fil
tum per electron makes all the difference. has been reduced to a single-particle problem

Rev. Mod. Phys., Vol. 71, No. 4, July 1999


11.5.3 Exercise 72
Using the definition of the Hamiltonian for non-interacting electrons
 2
1 q
Ĥ = p̂ − A (1642)
2m c

in the presence of an Aharonov-Bohm phase Φ, show that the total current I


flowing around the cylinder enclosing the flux is given by the derivative
∂E
I = −c (1643)
∂Φ

——————————————————————————————————

Solution

The total charge, per unit time, crossing a line parallel to the cylinder’s axis
is given by the expectation value
 
q X q
I = < Ψ | p̂ϕj − Aϕ (rj ) | Ψ > (1644)
me 2 π R L j c

where A is the vector potential due to the field in the sample. It has been
assumed that the charges are distributed uniformly on the sheet. Therefore, the
number of electrons crossing the line, in the time interval dt, is just ρ vϕ dt.
The Hamiltonian can be written as
X 1   2 
2 q qΦ
Ĥ = p̂z j + pˆϕ j − Aϕ (rj ) − (1645)
j
2 me c 2πRc

where the term involving Φ represents the change in the kinetic energy due to
the Aharonov-Bohm field. On taking the expectation value and then taking the
derivative with respect to Φ, one obtains
 
∂E q X q
= < Ψ | p̂j ϕ − Aϕ (rj ) | Ψ >
∂Φ 2 π R me c j c
1
= − I (1646)
c

——————————————————————————————————

505
11.5.4 The Fractional Quantum Hall Effect
Consider a particle of mass me , confined to move in the x − y plane with a
uniform magnetic field in the z direction. Using the circularly symmetric gauge,
the energy eigenstates are also eigenstates of angular momentum. The single-
particle wave functions describing the states in the lowest Landau level with
angular momentum m are all degenerate. The wave functions of these states
can be written as
m
x2j + yj2
  
1 xj + i yj
φm (rj ) = p exp − (1647)
2 π m! rc2 2 rc 4 rc2

where the length rc is given by


s
h̄ c
rc = (1648)
| e | Bz

The probability density for finding a particle at position r has a peak which
form a circle around the origin. The radius of the circle depends on m, so that
for large m p
r ∼ rc 2(m + 1) (1649)

The many-particle ground state wave function corresponding to the com-


pletely filled lowest Landau level is constructed as a Slater determinant from
the states of different m. The spins of the electrons are assumed to be fully
polarized by the applied field. The Ne particle wave function is
Ne
Y  x2k + yk2
 Y  
Ψ ∼ ( xi − xj ) + i ( yi − yj ) exp −
i>j
4 rc2
k=1
(1650)

This satisfies the Pauli exclusion principle as the wave function vanishes linearly
as ri → rj . The linear vanishing is a signature that the a pair of particles are
in a state of relative angular momentum m = 1, together with contributions
from states of higher angular momentum. This can be seen by expressing the
prefactor as a van der Monde determinant

z12 . . . z1Ne −1
0
z11

z1
z22 . . . z2Ne −1

Y   z20 z21
zi − zj = . (1651)

..
..

i>j .

z0
Ne z1 Ne z2Ne . . . z Ne −1
Ne

where zj = xj + i yj . Hence, the Ne electrons occupy the zero-th Landau


levels single-particle states with all the angular momentum quantum numbers

506
in the range between m = 0 and m = Ne − 1. The wave function is an
eigenfunction of the total angular momentum. The total angular momentum
about the origin is Mz = Ne ( N2e − 1 ) h̄. Since all the one-electron states in
the lowest Landau level are occupied, the many-particle state corresponds to a
uniform particle density of
D 1
ρ = = (1652)
Lx Ly 2 π rc2
particles per unit area. Hence, this many-particle state corresponds to the com-
pletely filled lowest Landau level.

For larger Bz the lowest Landau level is only partially filled. The wave
function which minimizes the interactions between pairs of particles is given by
the Laughlin trial wave function176
p YNe
Y  x2 + y 2
 
Ψp ∼ ( xi − xj ) + i ( yi − yj ) exp − k 2 k (1653)
i>j
4 rc
k=1

for odd integers p. The power of p in the prefactor has the effect of minimizing
the interactions between particles, since the square of the wave function vanishes
like a power law with power 2p instead of quadratically. This is a consequence
of the pairs of particles being in states with relative angular momentum p. The
Laughlin state is also an eigenstate of total angular momentum with eigenvalue
Mz = Ne ( N2e − 1 ) p h̄. Since the trial state contains one particle for every p
quantum states in the lowest Landau level, the many-particle state corresponds
to a uniform particle density of
1
ρ = (1654)
2 π rc2 p
particles per unit area. The filling factor, ν, is defined as
Ne
ν =
D
Φ0
= Ne
Φ
ρ Φ0
= (1655)
Bz
The Laughlin state corresponds to a state with the fractional filling, ν = p1 , of
the lowest Landau level. The energy of the Laughlin ground state is determined
by the Coulomb interaction energy. Since the Coulomb potential is a central
potential, it conserves the relative angular momentum. In the Laughlin state,
the energy is evaluated as177
  2
Eg 0.78213 0.211 0.012 e
= − √ 1 − 0.74 + (1656)
Ne p p p1.7 ε rc
176 R. B. Laughlin, Phys. Rev. Lett. 50, 1395 (1983).
177 D. Levesque, J. J. Weis and A. H. MacDonald, Phys. Rev. B, 30, 1056 (1984).

507
The energy per particle, for small p, is lower than any other candidate state by
an amount determined by the Coulomb interaction between states with angular
momentum m < p.

11.5.5 Quasi-Particle Excitations


Quasi-particles can either be thermally excited, or can be excited by coupling the
system to an external perturbation. One method of introducing quasi-particles
into the fractional quantum Hall state is by slightly changing the applied mag-
netic field. The quasi-particle excitations of the Laughlin state, like the quasi-
particle excitations of a completely filled Landau level, can be obtained from
considering the effect of adding a number of flux quanta, Φ, passing through the
center of the system. Although the magnetic field generating the extra flux does
not act on the electrons in the sample, it does add an Aharonov-Bohm phase to
the system. First, we shall consider the completely filled nr = 0 Landau level.

The single-particle wave function φ(r, ϕ) experiences a vector potential of


the form  
Bz r Φ
A = + êϕ (1657)
2 2πr
where Bz is the uniform field and Φ is the Aharonov-Bohm flux. Since the
single-particle energy eigenstates satisfy
2
h̄2 1 ∂
    
∂ 1 i h̄ ∂ q Bz r Φ
− r + − − ( + ) −E φ(r, ϕ) = 0
2 me r ∂r ∂r 2 me r ∂ϕ c 2 2πr
(1658)
Then, with the ansatz
 
1
φ(r, ϕ) = √ exp + iµϕ R(r) (1659)

one finds that the radial wave function is given by the solution of
2
h̄2 1 ∂ h̄2 q Bz r 2
    
∂ qΦ
− r + µ− − −E R(r) = 0
2 me r ∂r ∂r 2 me r2 2 π h̄ c 2 h̄ c
(1660)
Hence, the solutions for the lowest Landau level are of the form

r2
     

φ(r, ϕ) ∼ exp i ϕ exp i ν ϕ rν exp − (1661)
2 π h̄ c 4 rc2

where

ν = µ − (1662)
2 π h̄ c
Since the wave function is single valued, µ must be an integer, say m. The
maximum in the probability density is determined by ν and rc . On increasing

508
ν the particles move away from the origin. On subtracting one flux quantum,
Φ0 , through the center of the loop, where
2 π h̄ c
Φ0 = (1663)
|e|
then the degenerate eigenfunctions transform into themselves, m → m + 1. If
the Landau level had been completely filled, then one particle has been pushed
to the edge of the system and a hole has been created in the m = 0 orbit. This
is the quasi-hole excitation in the filled Landau level.

The wave function of the Laughlin state with an excited quasi-hole is given
by similar considerations. The insertion of a flux quanta produces an extra
Aharonov-Bohm phase. The requirement that the wave function is single valued
restricts µ to be integer, m. The Laughlin state in which the flux is decreased by
one flux quantum Φ0 has m shifted by m → m + 1 which creates a quasi-hole
at the origin. The many-particle wave function with a quasi-hole at the origin
is given by the expression
Ne
p Y
Y x2k + yk2
Y  
+
Ψp ∼ ( xi + i yi ) ( xi − xj ) + i ( yi − yj ) exp −
i i>j
4 rc2
k=1
(1664)
The wave function of the Laughlin ground state with a quasi-hole present at r0
is given by
Ne  Ne
p Y
x2k + yk2
Y Y  
Ψ+p ∼ xi − x0 + i ( y i −y 0 ) ( xi − xj ) + i ( y i − y j ) exp −
i=1 i>j
4 rc2
k=1
(1665)
where one flux quanta has also been removed from r0 . This state has angular
momentum of
Ne ( Ne − 1 )
Mz = p h̄ + Ne h̄ (1666)
2
as there is now an extra zero at point r0 . Due to the zero, the charge density
of this state is depleted around r0 . The charge deficiency is smaller than that
around the position of any electron by a factor of p1 . Hence, the quasi-hole has
the fractional charge − pq . The above argument demonstrating the existence
of fractionally charged quasi-particles can be re-phrased in a classical language.
On inserting a microscopic solenoid through the two-dimensional electron gas,
and adiabatically increasing the flux through the metal by an amount Φ0 , one
introduces a small electric field. From Faraday’s law of induction,
1 ∂B
∇ ∧ E = − (1667)
c ∂t
one finds, after using Stoke’s theorem, that the electric field encircles the solenoid
and the field at a radius r is given by
1 ∂Φ(t)
2 π r Eϕ (t) = − (1668)
c ∂t

509
Φ(t)

jr

Figure 225: The geometry of Laughlin’s gedanken experiment which identifies


the fractional charge on the quasi-hole excitations.

Since the two-dimension electron gas exhibits the fractional quantum Hall effect,
this field induces a radial current
jr = σr,ϕ Eϕ (1669)
where the transverse conductivity is given by
e2
σr,ϕ = − ν (1670)

and ν is the filling factor. This equation is integrated with respect to time. One
finds that, when the flux is increased by Φ0 , a charge Q is transferred into the
region
Z 0
Q = 2πr dt jr (t)
−∞
Z 0
e2 1 ∂Φ
= ν dt
h̄ c −∞ ∂t
e2 1
= ν Φ0
h̄ c
= ν|e| (1671)
Hence, the quasi-particle excited by the non-uniform flux has a fractional charge.
Alternatively, one may notice that by adding p quasi-holes at the same point
and then on adding an electron there, one just obtains the wave function for the
Laughlin state with one more electron. Hence, p quasi-holes are neutralized by
an extra electron. The operator creating a quasi-hole can be written just as
Ne 
Y 
Sp ∼ xi − x0 + i ( yi − y0 ) (1672)
i=1

510
since it just adds zeros to the wave function.

Creating a quasi-particle is a little more complicated, as adding a flux quan-


tum results in the transformation m → m − 1. Hence, the circles contract to
the origin, but the state with m = 0 is already filled in the initial state. This
must be lifted to the next Landau level, nr = 1. An operator Sp† which adds
a flux quantum at r0 , creating a quasi-particle, can be written just as
Ne  
Y ∂ ∂ x0 − i y0
Sp† ∼ − i − (1673)
i=1
∂xi ∂yi rc2

where this operator only acts on the polynomial part of the wave function and
not the exponential part. It reduces the angular momentum of each single-
particle state by one unit of h̄ and sends the particle at r0 into the higher
Landau levels. This activation process ensures that the quasi-particle excita-
tion spectra has a gap. The threshold energies quasi-hole and quasi-particle
excitations are determined by the Coulomb interaction, and for p = 3 have been
2 2
evaluated178 as 0.073 erc and 0.026 erc .

Since each quasi-particle of charge pq is attached to one flux quantum, the


statistics are neither fermionic nor bosonic179 . Two quasi-particles can be ex-
changed by a process whereby one quasi-particle is rotated by π in a semi-
circle centered on the other fixed quasi-particle and then the two particles are
translated in the same direction along the diameter. This process results in
an interchange of the electrons between their initial position eigenstates. The
phase of the wave function changes in this permutation. The rotation of the
quasi-particle through π around a flux tube produces an Aharonov-Bohm phase
of
q Φ0 π
π = (1674)
p 2 π c h̄ p
since the quasi-particle has a fractional charge. Thus, the quasi-particles have
fractional statistics.

These types of fractional or anyon statistics are only possible in two or less
dimensions. If the permutation of two particles produces a phase difference of πp ,
then the time reversed permutation process must yield a phase change of − πp .
In two-dimensions, the permutation process and the time reversed process are
distinguishable. However, if the two-dimensional process is embedded in three
dimensions, the processes are no longer distinct. On flipping over the plane in
which the particles are contained in, the interchange becomes equivalent to the
time reversed interchange. Hence,
   
π π
exp + i = exp − i (1675)
p p
178 F. D. M. Haldane and E.H. Rezayi, Phys. Rev. Lett. 54, 237 (1985).
179 B.I. Halperin, Phys. Rev. Lett. 52, 1583 (1984).

511
which restricts the value of p so that p = 1. Thus, in three dimensions, the
quasi-particles have to obey Fermi-Dirac statistics.

11.5.6 Skyrmions
Although we have considered the effect of extremely high magnetic fields, the
electronic spin system is not completely polarized, as we have been assuming.
The reasons for the relatively weak coupling between the electronic spin and the
magnetic field, compared with the coupling of the orbital motion to the field is
mainly due to the small band mass of the electron. The strength of the orbital
coupling to the field is determined by the quantity
q
(1676)
2 m∗ c
where m∗ is the band mass, which in GaAs has the value of m∗ ≈ 0.067 me
where me is the free electron mass. The strength of the spin coupling to the
field is governed by the electron mass and is given by
gq
(1677)
2 me c
where g is the gyro-magnetic ratio, which for free electrons would be g ≈ 2.03.
The strength of the anomalous Zeeman interaction in GaAs is further reduced by
the strong spin-orbit coupling by reducing g to the value given by the Lande gL
factor, gL ≈ 0.44. Since the importance of the anomalous Zeeman interaction
has been reduced by a factor of about 60, the spin directions are, therefore,
determined via the exchange parts of the Coulomb interaction. The magnitude
of the Coulomb interaction is given by the screened interaction

q2
(1678)
 rc
where  ∼ 12 and rc is the magnetic length. For fields of order B = 10 Tesla,
the Coulomb interaction is the same order of magnitude as h̄ ωc . Therefore,
the optimal polarization of the ground state is determined by the ratio of the
orbital part of the Zeeman interaction and the Coulomb exchange and partially
spin-polarized states may occur.

The spin magnetization M contributes to the effective magnetic field accord-


ing to
B ef f = B + 4 π M (1679)
Hence, if we consider excitations in the spin system, the magnetization will be
spatially varying, and so will the effective field. A variation in the effective field
will result in a local change of the filling factor. The system will respond to
the change of the filling factor by transferring charge. Thus, spin and charge
excitations are coupled. The lowest energy coupled spin-charge excitations are

512
skyrmions, not the Laughlin quasi-particles.

Consider an electron with spin S moving in the exchange field of the other
fixed electrons. The spin degree of freedom is governed by the effective Zeeman
Hamiltonian
g µB
ĤZ = − B ef f (r) . S (1680)

In the lowest energy state, the electron aligns its spin with the static effective
magnetic field. If the electron is moved around a closed contour, the spin will
remain aligned with the local magnetic field all along the contour. However,
the spin wave function does not return to its initial value but instead acquires a
phase, the Berry phase. The Berry phase is related to the solid angle enclosed
by the spin’s trajectory, as mapped onto the unit sphere in spin space. The
solid angle Ω traced out by the spin when completing the contour is given by
I
Ω = dϕ ( 1 − cos θ ) (1681)

where the spin direction is specified by the polar coordinates (θ, ϕ). After the
contour is traversed the spin wave function acquires an extra phase is S Ωh̄ .

The Berry phase can be illustrated by considering a spin one-half in a mag-


netic field of constant magnitude oriented along the direction (θ, ϕ). In this
case, the Zeeman Hamiltonian is given by

ĤZ = − µB ( B . σ ) (1682)

which can be expressed as


 
cos θ sin θ exp[− i ϕ ]
ĤZ = − µB B (1683)
sin θ exp[ + i ϕ ] − cos θ

For fixed (θ, ϕ), the Zeeman Hamiltonian ĤZ has an eigenstate χ+ given by

cos θ2
 
χ+ = (1684)
sin θ2 exp[ + i ϕ ]

which corresponds to the eigenvalue

E0 = − µB B (1685)

Thus, in this state the spin is aligned parallel to the applied field. For a static
field, one has the time-dependent wave function given by

cos θ2
   
µB B
χ+ (t) = exp + i t (1686)
sin θ2 exp[ + i ϕ ] h̄

where the time dependence is entirely contained in the exponential phase factor.

513
If the direction of the field (θ(t), ϕ(t)) is changed very slowly, one expects the
spin will adiabatically follow the field direction. That is, if the field is rotated
sufficiently slowly, one does not expect the spin to make a transition to the
state with energy E = + µB B where the spin is aligned anti-parallel to the
field. However, the wave function may acquire a phase which is different from
the time and energy dependent phase factor expected for a static field. This
extra phase is the Berry phase δ, and can be calculated from the Schrödinger
equation
    
∂ α(t) cos θ(t) sin θ(t) exp[− i ϕ(t) ] α(t)
i h̄ = − µB B
∂t β(t) sin θ(t) exp[ + i ϕ(t) ] − cos θ(t) β(t)
(1687)
We shall assume that the wave function has the adiabatic form
!
cos θ(t)
    
α(t) 2
µB B
= exp + i t − δ(t)
β(t) sin θ(t)
2 exp[ + i ϕ(t) ] h̄
(1688)
which instantaneously follows the direction of the field but is also modified by
the inclusion of the Berry phase δ(t). On substituting this ansatz into the
Schrodinger equation, one finds that the non-adiabatic terms satisfy
!
cos θ(t)
 
∂δ ∂ϕ 0
− 2 +
∂t sin θ(t)
2 exp[ + i ϕ(t) ] ∂t sin θ(t)
2 exp[ + i ϕ(t) ]
!
i ∂θ − sin θ(t)
2
=
2 ∂t cos θ(t)
2 exp[ + i ϕ(t) ]
(1689)

The above equation is projected onto the adiabatic state by multiplying it by


the row matrix  
cos θ(t)
2 sin θ(t)
2 exp[ − i ϕ(t) ] (1690)

On performing the projection, one finds that the derivative of θ w.r.t. t cancels
and that the equation simplifies to
∂δ ∂ϕ θ
− + sin2 = 0 (1691)
∂t ∂t 2
Hence, the Berry phase is given by integrating w.r.t. to t,
Z t
∂ϕ θ(t0 )
δ(t) = dt0 0 sin2
∂t 2
Z0
θ
= dϕ sin2
2
Z
1
= dϕ ( 1 − cos θ ) (1692)
2

514
On completing one orbit in spin space, the extra phase is given by

δ = (1693)
2
as was claimed.

Thus, an inhomogeneous effective field on the spin introduces an extra phase


of Ω2 in the wave function of the electron which is dragged around a contour.
This extra phase has the same effect as if the contour contains an additional
contribution to the magnetic flux given by
Ω Φ0
∆Φ = (1694)
2 2π
since encircling a flux quanta Φ0 produces a phase change of 2 π. Furthermore,
as the effective flux enclosed in the region is increased by ∆Φ, and the filling
fraction ν is constant, the number of electrons in the region must change by an
amount ∆Ne . Since the number of electrons is related to the flux by the filling
factor
ν Φ = N e Φ0 (1695)
the change in the number of electrons is given by

ν ∆Φ = ∆Ne Φ0 (1696)

Hence, the contour encloses an extra charge

∆Q = q ∆Ne
∆Φ
= νq
Φ0

= νq (1697)

Thus, the extra charge is determined by the Berry phase and the filling fraction
ν, and also the spin and charge excitations are coupled.

Due to the coupling of spin and charge, a localized spin-flip excitation of a


fully polarized ground state introduces a non-uniform charge density. Consider
a skyrmion excitation in the fully filled lowest Landau level180 . The ground
state wave function is written in second quantized form as
Y †
| Ψ0 > = am,↑ | 0 > (1698)
m

The creation of a charged spin-flip excitation at the origin requires adding a


down spin electron in the state m = 0. However, to allow for the spin excitation
to have a finite spatial extent and the charge density to re-adjust, the wave
180 D. H. Lee and C. L. Kane, Phys. Rev. Lett. 64, 1313 (1990).

515
function needs to be able to reduce the charge density and net spin at the origin
by redistributing them on neighboring shells. The excited state must also be an
eigenstate of total angular momentum Jz . Thus, to a first approximation the
excited state wave function can be written as
  Ne
v0 + u0 a†1,↓ a0,↑ a†0,↓ a†m,↑ | 0 >
Y
| Ψ+ > ≈
m=0
  Ne
v0 a†0↑ + u0 a†1,↓ a†0,↓ a†m,↑ | 0 >
Y
≈ (1699)
m6=0

where v0 and u0 are variational parameters which, since the wave function is
normalized, must satisfy

| u0 |2 + | v0 |2 = 1 (1700)

Iterating this process leads to the skyrmion wave function181


Ne  
vm a†m↑ + um a†m+1,↓ a†0,↓ | 0 >
Y
| Ψ+ > = (1701)
m=0

One expects that as m → Ne one recovers the fully polarized state, so


| vm | → 1 and | um | → 0. This is a variational wave function for the
excited state, and the parameters vm and um are to be determined by minimiz-
ing the expectation value of Ĥ.

The Hamiltonian can be approximated by


1 X
Vm,m0 a†m,↑ a†m0 +1,↓ am0 ,↑ am+1,↓ (1702)
X
Ĥ = m,σ a†m,σ am,σ +
m,σ
2! 0
m,m

which is a simplified version of the skyrmion Hamiltonian. On using the relations

< Ψ+ | a†m,↑ am,↑ | Ψ+ > = | vm |2


< Ψ+ | a†m,↑ am+1,↓ | Ψ+ > ∗
= vm um
< Ψ+ | a†m+1,↓ am,↑ | Ψ+ > = u∗m vm
< Ψ+ | a†m+1,↓ am+1,↓ | Ψ+ > = | um |2 (1703)

the expectation value of the Hamiltonian is found as


X  
2 2
< Ψ+ | Ĥ | Ψ+ > = m,↑ | vm | + m+1,↓ | um |
m
1 X ∗
+ Vm,m0 vm um u∗m0 vm0 (1704)
2! 0
m,m

181 H. A. Fertig, L. Brey, R. Cote and A. H. MacDonald, Phys. Rev. B, 50, 11018 (1994).

516
The energy of this excited state is to be minimized w.r.t. um and vm subject to
the constraint
| vm |2 + | um |2 = 1 (1705)
The minimization is performed using Lagrange’s method of undetermined mul-
tipliers, λm . The minimization results in the set of equations

∗ 1 ∗ X ∗
( m,↑ − λm ) vm + u Vm,m0 um0 vm 0 = 0 (1706)
2 m 0
m

and
1 ∗ X
( m+1,↓ − λm ) u∗m + v Vm,m0 vm0 u∗m0 = 0 (1707)
2 m 0
m

These sets of equations can be solved to yield the undetermined multipliers


  s 2
m↑ + m+1,↓ m↑ − m+1,↓
λm = ± + | ∆m |2 (1708)
2 2

where we have defined the parameter


1 X ∗
∆m = Vm,m0 vm 0 um0 (1709)
2! 0m

which we expect will decrease with increasing m. The factors | vm |2 and | um |2


are then found as
 
2 1 m,↑ − m+1,↓
| vm | = 1 ∓ p
2 ( m↑ − m+1,↓ )2 + 4 | ∆m |2
 
1 m,↑ − m+1,↓
| um |2 = 1 ± p (1710)
2 ( m↑ − m+1,↓ )2 + 4 | ∆m |2

Far from the center of the spin-flip excitation one expects that the ground state
will be recovered, so the spins will be polarized parallel to the field. Hence, we
shall use the upper signs. The equations (1709) and (1710) have to be solved
self-consistently for ∆m . We shall use a real solution for the gap. These are
combined to yield the “gap” equation
1 X ∆m0
∆m = Vm,m0 p (1711)
2 ( m0 ↑ − m0 +1,↓ )2 + 4 | ∆m0 |2
m0

The solution uniquely determines the wave function up to an undetermined


phase. From the gap equation, we note that as the Zeeman splitting increases,
the magnitude of ∆m decreases. The energy of the skyrmion is reduced by the
non-zero value for ∆m , since spreading out the charge reduces the Coulomb
interaction. However, the cost in Zeeman energy for flipping the spins in a large

517
Page 1 of 1

Figure 226: The spin distribution in a skyrmion excitation. The spin distri-
bution is radially symmetric, with an overturned spin at the center. The fully
polarized state is recovered at r → ∞.

region of space limits the size of the skyrmions.

Once the wave function has been determined, one can examine the spin
distribution. The direction of the spin at the point (r, ϕ) is defined as the
direction along which the spin density operator is maximum. The spin density
operator, projected along a unit vector η̂ in an arbitrary direction (θ0 , ϕ0 ) in
second quantized form is given by
( η̂ . σ̂(r) ) = Ψ† (r) ( η̂ . σ ) Ψ(r) (1712)
However, the field creation and annihilation operators are given by
X
Ψ† (r) = a†m,α φ∗m (r) χ†α
m,α
X
Ψ(r) = am0 ,β φm0 (r) χβ (1713)
m0 ,β

Hence, one findshttp://phys.thu.edu.tw/~mfyang/qhe/skyrmion.gif


the component of the spin density operator in the 14.10.2008 form
X X
( η̂ . σ̂(r) ) = φ∗m (r) ( η̂ . σ )α,β φm0 (r) a†m,α am0 ,β
m,m0 α,β
 
a†m,↑ am0 ,↓ exp[ − i ϕ0 ] + a†m,↓ am0 ,↑ exp[ + i ϕ0 ]
X
= φ∗m (r) φm0 (r) sin θ0
m,m0
 
a†m,↑ a†m,↓
X
+ φ∗m (r) φ
m0 (r) cos θ 0
a
m0 ,↑ − a
m0 ,↓ (1714)
m,m0

On taking the expectation value of the spin density operator in the skyrmion
state, one finds
X
< Ψ+ | ( η̂ . σ̂(r) ) | Ψ+ > = sin θ0 exp[ − i ϕ0 ] φ∗m (r) φm+1 (r) um vm

518
X
+ sin θ0 exp[ + i ϕ0 ] φ∗m+1 (r) φm (r) vm u∗m
m
X  
+ cos θ0 | φm (r) |2 | vm |2 − | φm+1 (r) |2 | um |2
m
(1715)

The wave functions can be expressed in planar polar coordinates as


 
1
φm (r, ϕ) = √ exp + i m ϕ Rm (r) (1716)

where Rm (r) is real. Hence, we have
X
< Ψ+ | ( η̂ . σ̂(r) ) | Ψ+ > = sin θ0 exp[ − i ( ϕ0 − ϕ ) ] Rm (r) Rm+1 (r) um vm

m
X
+ sin θ0 exp[ + i ( ϕ0 − ϕ ) ] Rm+1 (r) Rm (r) vm u∗m
m
X  
+ cos θ0 2
Rm (r) | vm |2 − Rm+1
2
(r) | um |2
m
(1717)

On maximizing w.r.t ϕ0 , one finds


X X
exp[ − 2 i ( ϕ0 − ϕ ) ] ∗
Rm (r) Rm+1 (r) um vm = Rm+1 (r) Rm (r) vm u∗m
m m
(1718)
Hence, ϕ0 = ϕ, that is, the in-plane component of the spin is directed radially
outwards. On substituting this relation into the wave function, one finds that
the spin density along this direction simplifies to
X
< Ψ+ | ( η̂ . σ̂(r) ) | Ψ+ > = sin θ0 Rm+1 (r) Rm (r) ( vm u∗m + vm

um )
m
X  
+ cos θ0 | vm |2 Rm
2
(r) − | um |2 Rm+1
2
(r)
m
(1719)

The out of plane component of the spin is determined by θ0 . This is found from
maximizing w.r.t. θ0 , and leads to
( v u∗ + vm ∗
P
um ) Rm+1 (r) Rm (r)
0
tan θ =  m m
m  (1720)
2 2
P
0 | vm0 |2 R 0 (r) − | um0 |2 R 0 (r)
m m m +1

At large distances r from the origin, the wave functions are dominated by a
range of m values around the value given by

r2 = 2 m rc2 (1721)

519
r
In this case, one has Rm+1 (r) ∼ 2 √m rc
Rm (r), hence, the out of plane angle
at the distance r from the origin is governed by

θ0

r um
tan ∼ √ (1722)
2 2 m rc vm

Since, the ratio decreases with increasing m, the spin direction varies from
θ0 = π at the origin to θ0 = 0 as r → ∞. The texture can be expressed
empirically as
θ0 λ
tan = (1723)
2 r
where λ expresses the size of the skyrmion. In fact the size of the skyrmion is
determined by the m variation of um , or more explicitly on the ratio
 
m+1,↓ − m,↑
(1724)
∆m

The size of the skyrmion decreases as the magnitude of the Zeeman splitting
increases. This reflects the fact that the energy required to flip the spins in a
region of large spatial extent becomes prohibitively costly as the Zeeman inter-
action is increased.

Skyrmions can also be created in the Laughlin state. The skyrmions have
lower energy than the Laughlin quasi-particles for all values of the gyromagnetic
ratio g. The energy difference is largest for g → 0. However, as g → ∞
the region over which the spin is varying is reduced, and the energy approaches
that of the Laughlin quasi-particle. In fact, in this limit, the skyrmion becomes
identical to the Laughlin quasi-particle.

Skyrmion excitations have been found in N.M.R. experiments182 . The Knight


shift is proportional to the spin polarization of the electron gas. The electron gas
can be fully spin polarized if ν < 1 but can only be partially spin polarized for
ν > 1. In this latter case, the partial polarization is limited by ( 2 − ν )/ν, for
non-interacting electrons. The spin polarization is reduced by having a number
of skyrmions present. Skyrmions can be introduced into the system by chang-
ing the filling factor from an integer or fractional value. The dependence of the
Knight shift on the filling factor in a range close to ν = 1 follows the theo-
retical results and indicates that each skyrmion contains about 3.6 flipped spins.

11.5.7 Composite Fermions


The Laughlin wave function describes states with filling fractions p1 , where p
is odd. However, the sequence of fillings at which the fractional quantum Hall
182 S. E. Barrett, G. Dabbagh, L. N. Pfieffer, K. N. West, and R. Tycko, Phys. Rev. Lett,

74, 5112 (1995).

520
Figure 227: A cartoon of Composite Fermions in which each electron is bound
to an even number of flux quanta.

effect is observed is given by the expressions


n
ν = (1725)
2nr ± 1
and
n
ν = 1 − (1726)
2nr ± 1
where n and r are integers. These two filling fractions are related by approxi-
mate electron-hole symmetry, in which occupations of only the lowest Landau
level are considered.

These other states can be expressed in terms of composite fermions183 . The


general Laughlin state can be written as
Ne
2r+1 Y
Y  x2 + y 2
 
Ψ2r+1 ∼ ( xi − xj ) + i ( yi − yj ) exp − k 2 k
i>j
4 rc
k=1
Y  2r
= ( xi − xj ) + i ( yi − yj ) ×
i>j
Ne
Y  x2 + y 2
 Y  
× ( xi − xj ) + i ( yi − yj ) exp − k 2 k
i>j
4 rc
k=1
(1727)

This can be thought of as taking the state in which the lowest Landau level has
the filling ν = 1 and attaching an even number 2 r flux quanta to each particle.
Then, the statistics of the composite particle (composed of the 2 r flux quanta
and the electron) is manifested by the exchange phase. The exchange phase of
the composite particle will be the sum of the exchange phase of an electron (π)
plus a multiple of π for each flux quantum. Hence, the total exchange phase of
the composite particle wave function will be

π + 2rπ = (2r + 1)π (1728)


183 J. K. Jain, Phys. Rev. Lett. 63, 199 (1989).

521
and, thus, the composite particle is a fermion.

Consider the state with the general filling factor


n
ν = (1729)
2rn ± 1
If the electrons are attached to 2 r flux tubes, one has composite fermions. This
has the effect of reducing the magnetic field from Φ to a value Φ∗ given by

Φ∗ = Φ − 2 r N e Φ0 (1730)

This effective free flux can be either positive or negative. The fractional quantum
Hall effect of electrons with filling factor ν can be related to the quantum Hall
effect of composite fermions with filling factor ν ∗ . The relation between ν and
ν ∗ is found by first inverting the definitions
N e Φ0
ν =
Φ
N e Φ0
ν∗ = (1731)
Φ∗
in which Φ∗ and, therefore, ν ∗ are assumed to be positive. Then, substituting
Φ∗ and Φ into the relation given by eqn(1730) yields
1 1

= − 2r (1732)
ν ν
or
ν∗
ν = (1733)
2ν∗ r + 1
Hence, the fractional quantum Hall effect at general fillings can be related to
the integer quantum Hall effect, with integer fillings ν ∗ , for composite fermions
where 2 r flux tubes attached to each electron. Using r = 1, this process
yields the hierarchy of ν = 31 , 25 , 37 , 49 , . . . , which is the most easily
observable sequence of fractional quantum Hall plateaus. The expression for
the filling fractions with the minus sign in the denominator can be obtained by
considering negative values of Φ∗ , for which a minus sign has to be inserted into
the definition of ν ∗ in order to keep the filling fraction positive.

522
12 Insulators and Semiconductors
The existence of band gaps is a natural consequence of Bloch’s theorem for pe-
riodic crystals. However, the existence of band gaps is a much more universal
phenomenon, for example it also appear in amorphous materials. In these cases,
the existence of band gaps can be traced back to the energy gaps separating the
discrete bound state energies of isolated atoms. When the atoms are brought
together to form a solid, each electron will be shared with all the atoms in a
crystal like in a giant molecule of N atoms. The set of discrete energy lev-
els from each of the N atoms, are nearly degenerate. The binding of the N
atomic degenerate atomic states into a N molecular states will involve bonding
/ anti-bonding splittings that raises the degeneracy. As the energy spread of
the bonding anti-bonding states are fixed, the levels form a dense set of discrete
levels, which can be approximated by a continuous energy band. The separation
between consecutive bands is roughly determined by the energy separation of
the discrete levels the isolated atom. Generally, the low-energy bands have a
small energy spread, and a clear correspondence with the atomic levels can be
established. However, the higher energy bands tend to have larger band widths,
so the bands overlap and the correspondence with the atomic orbitals becomes
more obtuse.

An example is given by the ionic compound LiF , in which the Li ion loses
an electron, and the F ion gains an electron in order that each ion only have
completely filled atomic shells. The Li 1s and F 1s levels are completely filled
and are well separated forming the core levels. The F 2s and 2p levels are also
occupied, but have broader band widths and form the occupied valence bands.
The unoccupied Li 2s and the unoccupied F 3s and 3p levels have large band
widths which strongly overlap yielding a conduction band which has mixed char-
acter. The density of states from the completely filled valence band states are
separated by an energy gap from the completely empty conduction band portion
of the density of states. By definition, the Fermi energy or chemical potential
lies somewhere in the energy gap. The existence of a gap in the density of
states at the Fermi energy is the characteristic feature that defines an insulator
or semiconductor.

Elementary semiconductors or insulators, such as diamond C, Si and Ge,


all belong to group IV of the periodic table. Group IV elements are semicon-
ducting since the outer shells of the individual atoms are half filled. By forming
tetrahedral bonds with neighboring atoms, the electronic states form bonding
and anti-bonding levels. The electrons completely fill the bonding levels which
are separated from the empty anti-bonding levels by a gap. These are examples
of the 8N rule, which states stable compounds are formed by sharing electrons
with atoms so that the outer shells are complete. Compound semiconductors
can be formed by combining group III and group V, elements or group II with
group VI elements. The most commonly used compound semiconductor is the
III-IV compound GaAs, whereas CdSe and CdT e are group II-IV compounds.

523
Group I and Group VII elements are usually ionic insulators, which have large
gaps.

IV Element Lattice Constant Å Energy Gap (eV) Structure

C 3.57 5.48 Cubic


Si 5.43 1.12 Cubic
Ge 5.66 0.664 Cubic

III-V Compound Lattice Constant Energy Gap (eV) Structure

BN 3.62 6.4 Cubic


BP 4.54 2.4 Cubic
AlN 3.11 , 4.98 6.2 Hexagonal
AlP 5.46 2.5 Cubic
AlAs 5.66 2.15 Cubic
GaN 3.18 , 5.17 3.44 Hexagonal
GaP 5.45 2.27 Cubic
GaAs 5.65 1.42 Cubic
InN 3.54 1.89 Cubic
InP 5.87 1.34 Cubic
InAs 6.06 0.354 Cubic

524
II-IV Compound Lattice Constant Energy Gap (eV) Structure

ZnS 5.41 3.68 Cubic


ZnO 3.25 , 5.21 3.44 Hexagonal
ZnSe 5.67 2.7 Cubic
ZnT e 6.10 2.28 Cubic
CdS 5.82 2.55 Cubic
CdSe 6.05 1.9 Cubic
CdT e 6.48 1.475 Cubic
α − HgS 4.15 , 9.5 2.1 Trigonal

The distinction between a semiconductor and insulator is only by the mag-


nitude of the energy gap between the lowest unoccupied state and the highest
occupied state. In insulators this energy gap is large, perhaps larger than 3
eV. In semiconductors this energy gap is small, so the electronic properties are
determined by the electronic states close to the bottom of the conduction band
and the states close to the top of the valence band. The value of the gap de-
creases along the column of the periodic table. This can be considered to be a
consequence of the decrease of the pseudo-potential due to the screening by the
core electrons. Alternatively, from the view-point of the tight-binding model,
the decrease in the band gap can be associated with the increase of the lattice
constants. The value of the gap also increases with the degree of ionicity.

Semiconductors can also be formed from transition metals, such as F eSe, or


lanthanide materials such as SmB6 , Y bB12 or Ce3 Bi4 P t3 , and actinide materi-
als such as U N iSn. In these materials the energy gaps can be as small as a few
meV, and are strongly temperature dependent. It is believed that the smallness
of the energy gaps in these materials is intimately connected to the strength of
electron-electron interactions.

The density of states close to the band gap can usually be parameterized by a
few quantities, such as the value of the band gap, and the effective masses me and
mh for the valence and conduction bands. This is true, since the discontinuities
at the band edges are van Hove singularities. Due to the symmetry one can
represent the single-particle Bloch energies of the valence band and conduction
band states as
d
X
Ec (p) = Eg + ai p2i
i=1

525
d
X
Ev (p) = − bi p2i (1734)
i=1

where the zero of energy was chosen to be at the top of the valence band. Using
the definition of the effective mass as
1 ∂ 2 E(p)
= (1735)
mα,β ∂pα ∂pβ
one finds an effective mass tensor mα,β . The effective mass tensor is symmetric
and can be diagonalized. The diagonal elements are positive for the conduction
band and negative for the valence band. Typical values of the effective mass are
less than unity ∼ 0.04 me , but increase with increasing energy gaps.

Material Eg (eV) m∗e m∗h

Diamond 5.48 0.36 - 1.08 , - 0.36 , - 0.154


Silicon 1.12 0.191 - 0.537 , - 0.153 , - 0.234
Germanium 0.664 0.081 - 0.284 , - 0.044 , - 0.095

There are two types of semiconductors that are frequently encountered. In-
trinsic semiconductors and extrinsic semiconductors.

A Intrinsic Semiconductors.

These are pure semiconductors, where the density of states consists of a


completely filled valence band and a completely empty conduction band at T =
0, which are separated by a band gap Eg . At temperatures comparable to the
band gap
Eg ∼ kB T (1736)
a finite number of electrons can be excited from valence band states to the
conduction band. The thermal excited electrons in the conduction band are
associated with empty states in the valence band. For each conduction electron
there is one empty valence band state. A hole is defined as the absence of an
electron in a valence band state. Thus, at finite temperatures, one has a finite
number of electron - hole pairs. For materials such as Si or Ge the gap in the
density of states is of the order of 1.12 to 0.66 eV. Thus, the number of thermally
activated electron hole pairs is expected to be extremely small under ambient
conditions.

526
B Extrinsic Semiconductors.

Extrinsic semiconductors are a type of semiconductor that contain impuri-


ties. Semiconductors with impurities can have discrete atomic levels that have
energies which are lower than the empty conduction band and higher than the
full valence band. That is the impurity level lie within the gap.

There are two types of extrinsic or impurity semiconductors.

N type semiconductors have impurities with levels which at T = 0


would be filled with electrons. At higher temperatures the electrons can be
excited from the levels into the conduction band. These types of impurities are
called donors, since at high temperatures they donate electrons to the conduc-
tion band. An example of an N type semiconductor is given by semiconducting
Si in which As impurities are substituted for some Si ions or alternatively semi-
conducting Ge substitutionally doped with P impurities. These are examples
of elemental semiconductors from the IV column of the periodic table doped
with impurities taken from the V column. Since the impurity ion has one more
electron than the host material, the host bands are completely filled by taking
four electrons from each impurity, but the impurity ion can still release one
extra electron into the conduction band.

For Si doped with a low concentration of As impurities, each As impurity


can be considered individually. The As atom contains 5 electrons, while the
perfect valence band only contains 4 electrons per site. Thus, the As ion has
one extra electron which, according to the Pauli exclusion principle has to be
placed in states above the valence band. In the absence of the impurity potential
of the ionized As atom, this extra electron would go into the conduction band
and would behave very similarly to a free electron with mass me . However,
one must consider the effect of the potential produced by the As+ ion and the
spatial correlation that this imposes on the extra electron.

The As+ ions has a positive charge, which affects the free conduction elec-
tron much the same way as the positive nuclear charge effects the electron in a
H atom. The extra electron becomes bound to the donor atom. In the semi-
conductor, the binding energy is extremely low and the radius of the orbit is
large. This can be seen by examining the potential produced by the As+ ion

e2
V (r) = − (1737)
εr
which is screened by the dielectric constant ε. The dielectric constant for Si
has a value of about 70. The mass of the electron is the effective mass me . The
Bohr radius of the donor orbital is given by

ε h̄2 2
ad (n) = n (1738)
me e2

527
where n is the principal quantum number. The energy of the donor levels below
the bottom of the conduction band are given by the expression
me e4 1
Ed (n) = − (1739)
2 ε2 h̄2 n2
The lowest state of the donor level is occupied at T = 0, where the electron
is located in an orbit of radius ad (1) ∼ 30 A and the energy of the donor
level Ed (1) ' − 0.02 eV. Thus, there are shallow impurity levels just below
the bottom of the conduction band, one of these set are occupied at T = 0
K. These discrete levels can be represented by a set of delta functions in the
density of states. For sufficiently large concentrations these impurities can be
represented in terms of a finite impurity band which appears inside the gap in
the density of states of the pure host material. Since the value of the gap is
comparable to room temperature, one expects that under ambient conditions
there are a finite number of thermally activated conduction electrons available
for carrying current.

P type semiconductors have impurities with levels that at T = 0 would


be empty of electrons. At higher temperatures, electrons from the filled valence
band will be excited into the empty impurity levels. These thermally excited
levels will be localized on the impurities (for small concentrations of impurities).
However, the holes present in he valence band allows the valence electrons to
conduct electricity and contribute to the properties of the semiconductor. The
impurities in P type semiconductors are called acceptors as they accept electrons
at finite temperatures. Examples of P type semiconductors are Ga impurities
doped substitutionally on the sites of a Si host, or Al impurities substituted
for the atoms in a Ge crystal. These are examples of impurities from the III
column of the periodic table being substituted for the atoms in a semiconductor
composed of an element from the IV column of the periodic table. In this case,
the type III impurity provides only 3 electrons to the host conduction band,
which at finite T contains one hole per Ga impurity.

The Ga impurity atom shares the electrons of the surrounding Si atom and
becomes negatively charged. The extra hole orbits around the negative ion pro-
ducing acceptor levels that lie just above the top of the valence band. A finite
concentration of acceptor levels is expected to produce a smeared out acceptor
band just above the top of the valence band. Due to the smallness of the gap
between the valence band and the acceptor levels, at room temperature an ap-
preciable number of electrons can be excited from the valence band onto the
acceptor levels.

12.1 Thermodynamics
The thermodynamic and transport properties of semiconductors are governed
by the excitations in the filled valence band or empty conduction band, as the

528
fully filled or empty bands are essentially inert. The excitations of electrons
on the localized impurity levels of extrinsic semiconductors, do not directly
contribute to physical properties. However, the electrons in the conduction
band and the valence band are itinerant and do contribute. To develop a theory
of the properties of semiconductors it is convenient to focus attention on the
few unoccupied states of the valence band rather than the many filled states.
This is achieved by reformulating the properties in terms of holes.

12.1.1 Holes
A hole is an unoccupied state in an otherwise completely occupied valence band.
The probability of finding a hole in a state of energy , Ph () is given by the
probability that an electron is not occupying that state

Ph () = 1 − f () (1740)

Thus, as
1
Ph () = 1 − f () = (1741)
1 + exp[ −β ( − µ) ]
one finds that the probability of finding a hole in a state of energy  is also given
by the Fermi function except that  → −  and µ → − µ. That is the energy
of the hole is the energy of a missing electron.

The momentum of a hole can be found as the momentum of a completely


occupied band is zero, since inversion symmetry dictates that for each state with
momentum k e there is a state with momentum −k e with the same energy, and
by assumption both are occupied. Then by definition the momentum of a hole
in k e is that of the full band with one electron missing

kh = 0 − ke (1742)

Thus, the momentum of a hole is opposite to that of the missing electron

kh = − ke (1743)

The charge on the electron is − | e |. The charge on the hole can be obtained
from the quasi-classical form of Newton’s laws applied to an electron in an
electric field E,
dk e
h̄ = −|e|E (1744)
dt
As the momentum of a hole is given by

kh = − ke (1745)

one finds that


dk h
h̄ = +|e|E (1746)
dt

529
Thus, the charge of the hole is just | e |. This is consistent with considerations
of electrical neutrality. In a solid in equilibrium, the charge of the nuclei is equal
in magnitude to the charge of the electrons. Thus
 
| e | Z N n − Ne = 0 (1747)

Thus, in equilibrium the number of electrons is related to the number of nuclei


via Ne = Z Nn . Now if one electron is removed from the valence band and
removed from the solid, there is just one hole. The total charge of the one hole
state is given by
 
Qh = | e | Z Nn − ( Ne − 1 ) (1748)

where the number of electrons is now Ne − 1 = Z Nn − 1. Hence,

Qh = | e | (1749)

Thus, the charge on the hole is minus the charge of the electron.

The velocity of a hole v h can be found by considering the electrical current


carried by the electrons. The electrical current carried by a full band is zero.
The electrical current carried by a hole is the electrical current carried by a full
band minus one electron.

j h = | e | vh = 0 − ( − | e | ve ) (1750)

Thus,
vh = ve (1751)
Thus, the velocity of the hole is the same as the velocity of the missing electron.

The above results indicate that the velocity of the hole is in the same di-
rection as the velocity of the missing electron, but the momenta have opposite
directions. This implies that the sign of the hole mass should be opposite of the
mass of the missing electron. This can also be seen by considering the alternate
form of the quasi-classical equation of motion
dv e
me = −|e|E (1752)
dt
and since
ve = vh (1753)
and the charge of the hole is | e |, one has

dv h
− me = |e|E (1754)
dt

530
Hence,
mh = − me (1755)
the mass of the hole is the negative of the mass of the missing electron.

The most important number for the thermodynamic and transport proper-
ties of a semiconductor is the number of charge carriers. This can be obtained
from analysis of the appropriate semiconductor.

12.1.2 Intrinsic Semiconductors


The number of electrons thermally excited into the conduction band of an intrin-
sic semiconductor can be calculated with knowledge of the chemical potential
µ(T ). Consider an intrinsic semiconductor like pure Si which has an energy
gap Eg of the order of 1 eV. The energy of the top of the valence band shall be
set to zero. The energy of the hole is zero at the top of the valence band and
increases downwards, as the energy is that of the missing electron. This agrees
with the fact that mh < 0, as
∂ 2 Ek 1
2
= < 0 (1756)
∂ k mh
The energy of the conduction electron is given by
h̄2 k 2
Ek = E g + (1757)
2 me
The chemical potential can be obtained from considerations of charge neutrality,
which implies that the number of conduction electrons is equal to the number
of holes in the valence band. The number of electrons can be calculated from
the electron density of states
  32
V 2 me 1
ρ() = 2 (  − Eg ) 2 f or  > Eg (1758)
2 π2 h̄
The number of electrons in the conduction band, Nc is given by
Z ∞
Nc = d ρ() f () (1759)
Eg

The Fermi function can be expressed in terms of (  − Eg ) as


1
f () = (1760)
exp[ β (  − Eg ) ] exp[ − β ( µ − Eg ) ] + 1
E
If it is assumed that µ = 2g then, as the lowest conduction band state has
the energy  = Eg , the Fermi function for the conduction electrons is almost
classical since both exponents are positive. Thus, with the assumption
 
f () ∼ exp − β (  − µ ) (1761)

531
As shall be shown later the above assumption is valid. The number of thermally
activated conduction electrons can now be obtained by evaluating the integral
over the classical Boltzmann distribution by changing variables from  to the
dimensionless variable x = β (  − Eg ). Then
 3   Z ∞  
V 2 me kB T 2 1
Nc ∼ exp − β ( Eg − µ ) dx x 2 exp − x
2 π2 h̄2 0
 3  
2 π me kB T 2
= 2V exp − β ( Eg − µ )
h2
(1762)
which is usually expressed in terms of the thermally De Broglie wavelength of
the conduction electrons λe defined by
h
λe = √ (1763)
2 π me kB T
as  
V
Ne = 2 3 exp − β ( Eg − µ ) (1764)
λe
The number of thermally excited conduction electrons depends exponentially
on the unknown quantity µ.

The number of holes in the valence band can be found from a similar cal-
culation. The occupation number for the holes in the valence band is given
by
Ph () = 1 − f ()
1
=
1 + exp[ −β ( − µ) ]
which is the Fermi function except that  → −  and µ → − µ. The density
of states for the holes in the valence band is given by
  32
V − 2 mh 1
ρ() = ( −  )2 for  < 0 (1765)
2 π2 h̄2
E
Since the value of µ is assumed to be positive and of the order of 2g the Fermi
function can also be treated classically. Thus, the number of holes Nv in the
valence band is given by the integral
Z 0
Nv = d Ph () ρ()
−∞
  32 Z 0 1
V − 2 mh kB T ( −  )2
= d
2 π2 h̄2 −∞ exp[ − β ( − µ)] + 1
  32  
− 2 π mh kB T
∼ 2V exp − β µ
h2
(1766)

532
which also depends on µ.

Since in an intrinsic semiconductor the electrons and holes are created in


pairs, one has
Nc = Nv (1767)
Therefore, the equation for the unknown variable µ is given by
    32  
− mh
exp 2βµ = exp β Eg (1768)
me

or on taking the logarithm, one has



Eg 3 − mh
µ = + kB T ln (1769)
2 4 me

Hence, at T = 0, the chemical potential lies half way in the gap and only
acquires a temperature dependence if there is an asymmetry in the magnitude
of the conduction band density of states and the valence band density in the
vicinity of the band edge. This result justifies the previous assumption that the
distribution functions can be treated classically. The result also shows that the
number of thermally activated electrons or holes crucially depends on the gap
via the exponential factor  
Eg
exp − β (1770)
2
and is extremely small at room temperature for a semiconductor with a gap
that has a magnitude of the order of electron volts.

12.1.3 Extrinsic Semiconductors


In an intrinsic semiconductor the gap in the host material is usually much larger
than the gap between the impurity levels and the band edges. Therefore, under
ambient temperatures one can neglect one of the bands, as the number of ther-
mally excited electrons or holes in that band is small.

For concreteness, consider an N type semiconductor. The bottom of the


conduction band is defined as the reference energy E = 0 and the energy of
the donor levels is defined as − Ed . Let Nd be the number of donor atoms,
and nd be the number of electrons remaining in the donor atoms, and nc be the
number of electrons thermally excited to the conduction band. The important
physical properties are determined by the number of conduction electrons. This
can be calculated from the free energy F .

The free energy F is given in terms of the energy and entropy via the relation

F = E − T S (1771)

533
where the total energy is
E = − nd Ed (1772)
as the energy of the thermally activated conduction electrons E ∼ 0 as they
occupy states at the bottom of the band. The entropy S is given by
S = kB ln Ω (1773)
where Ω is the number of possible states of the system. This is just the number
of ways of distributing nd electrons in the 2 Nd states of the donor atoms. It
is assumed that the donor atoms only have a two-fold spin degeneracy, and
that there is no interaction energy between two electrons occupying the two
spin states on the same donor atom. With these assumptions, the number of
accessible states is given by
( 2 Nd )!
Ω = (1774)
nd ! ( 2 Nd − nd )!
Hence, the Free energy can be calculated in the thermodynamic limit as
F = − Ed nd
 
− kB T 2Nd ln 2Nd − nd ln nd − ( 2Nd − nd ) ln( 2Nd − nd )

(1775)
The chemical potential is found from minimizing the free energy with respect
to nd
∂F
µ =
∂nd
 
= − Ed − kB T − ln nd + ln( 2Nd − nd ) (1776)

Thus, on exponentiating
 
nd
exp β (µ + Ed ) = (1777)
2 N d − nd
which leads to the number of electrons occupying the donor orbitals as
2 Nd
nd = (1778)
exp[ β ( − Ed − µ ) ] + 1
which is just governed by the Fermi function f (−Ed ) and the number of orbitals
2 Nd .

The number of thermally excited conduction electrons is given by


Z
1
nc = 2 V d3 k
exp[ β ( E(k) − µ ) ] + 1
 3  
2 π me kB T 2
= 2V exp β µ (1779)
h2

534
The unknown chemical potential µ can be eliminated from the above two equa-
tions and thereby a relation between nc and nd can be found
    32
nc ( 2 N d − nd ) 2 π me kB T
= exp − β Ed 2V (1780)
nd h2

This is the law of mass action for the dissociation reaction

nd → nc + ( N d − nd ) (1781)

in which filled donors dissociate into conduction electrons and unoccupied donor
atoms. This relation can be written as
 
nc ( 2 N d − nd ) V
= exp − β Ed 2 3 (1782)
nd λ

where λ is the thermal De-Broglie wavelength,


h
λ = √ (1783)
2 π me kB T
On using the condition of electrical neutrality

N d = nd + nc (1784)

one can eliminate nd to find the number of conduction electrons as


 
nc ( N d + nc ) 2V
= exp − β Ed (1785)
( N d − nc ) λ3

This is a quadratic equation for nc which can be solved to yield the positive
root
  
1 2V
nc = − Nd + 3 exp − β Ed
2 λ
s   2  
1 2V 8 Nd V
+ Nd + 3 exp − β Ed + exp − β Ed
2 λ λ3
(1786)

With this expression for the number of conduction electrons one can solve for µ
from
nc λ 3
 
µ = kB T ln (1787)
2V
At sufficiently low temperatures when

λ3 N d
 
 exp − β Ed (1788)
2V

535
one finds that
µ = − Ed (1789)
as at T = 0 the donor level is only partially occupied as there are 2 Nd orbitals
including spin and Nd electrons. At sufficiently high temperatures, the donor
levels are almost completely ionized and

nc ' N d (1790)

and the chemical potential is found as

N d λ3
 
µ = kB T ln (1791)
2V
——————————————————————————————————

12.1.4 Exercise 73
Assume that if the donor levels are localized to such an extent that the Coulomb
repulsion between two opposite spin electrons in the same donor atom is ex-
tremely large. This assumption makes the occurrence of doubly occupied donor
atoms extremely improbable. Show that the number of accessible states is given
by
Nd !
Ω = 2Nd (1792)
nd ! ( Nd − nd )!
Hence, show that
Nd
nd =   (1793)
1
1 + 2 exp − β ( Ed + µ)

Thus, the interaction affects the statistics of the occupation numbers. Also show
that the law of mass action becomes
 
nc ( N d − nd ) V
= 3 exp − β Ed (1794)
nd λ
——————————————————————————————————

12.2 Transport Properties


Transport in doped semiconductors is mainly dominated by scattering from
donor impurities. The potential due to the isolated donor impurities is a
screened Coulomb interaction
Z e2
 
V (r) = − exp − kT F r (1795)
εr

536
The transport scattering rate is given by
2
4 π Z e2
Z Z 
1 c 0 02 0
= dk k δ( E(k) − E(k ) ) d cos θ ( 1 − cos θ )
τ (E(k)) h̄ ε ( (k − k 0 )2 + kT2 F )
(1796)
where c is the impurity concentration and θ is the scattering angle. The inte-
gration over the magnitude of k 0 can be performed over the delta function. The
integration over the angle can be performed as
2
4 π Z e2
Z 
1 2mc 1
= d cos θ ( 1 − cos θ )
τ (E(k)) h̄3 k 3 )2 )
ε ( 1 − cos θ + ( kT F
k
(1797)
Hence, τ (k) ∼ k 3 . The conductivity can be evaluated from the formulae

2 e2 X h̄2 2 ∂f (E(k))
σx,x = − k τ (k)
V m2 x ∂E(k)
k

2 e2 X h̄2 2
 
= k τ (k) β exp − β ( E(k) − µ )
V m2 x
k

β h̄2 2
  Z  
∼ β exp β µ dk k 7 exp − k
2 me
(1798)

On changing the variable of integration from k to the dimensionless variable

h̄2 k 2
x = (1799)
2 me kB T
one finds that the temperature dependence of the conductivity is governed by
 
σx,x ∼ ( kB T )3 exp β µ (1800)

The conductivity is proportional to the electron density and the scattering time.
3
For Coulomb scattering, the average scattering time is proportional to T 2 . On
the other hand, scattering from lattice vibrations gives rise to a conductivity
that is just proportional to exp[ β µ ].

12.3 Optical Properties

537
13 Phonons
14 Harmonic Phonons
The Hamiltonian describing the motion of the ions can be formulated by as-
suming that the ions move slowly compared with the electrons. Thus, at any
instant of time, the electrons have relaxed into equilibrium positions and the
ions are frozen into their instantaneous positions184 .

It is assumed that the ions have definite mean equilibrium positions and
that the displacements of the ions from these equilibrium positions are small.
First, the situation in which the crystal lattice can be described by a Bravais
lattice with a one atom basis is considered. In a later section, the effects of a
multi-atom basis will be described. For the case of a one atom basis, the energy
of the solid can be calculated in the which the ionic positions are displaced by
amounts ui ,
r(Ri ) = Ri + ui (1801)
where ui is the deviation from the equilibrium position Ri . It is assumed that

ui

r(Ri)

Ri

Figure 228: The atoms are moved from their equilibrium positions Ri by the
displacements ui .

the total energy can be formulated as a constant plus pair-wise interactions


which depends on r(Ri ) − r(Rj ). The pair-wise interaction is given in terms of
the pair-potentials Θ(r(Ri ) − r(Rj )) via

1 X
V̂ = Θ(r(Ri ) − r(Rj ))
2 i,j
1 X
= Θ(Ri − Rj + ui − uj ) (1802)
2 i,j

184 M. Born and R. Oppenheimer, Ann. Phys. (Leipzig), 84, 457 (1927).

538
The Hamiltonian governing the motion of the ions is just given by the sum of
the kinetic energies of the ions of mass M and the pair-wise interactions

X P̂ 2
i
Ĥ = + V̂ (1803)
i
2 M

The harmonic approximation assumes that the displacements ui are sufficiently


small so that Ĥ can be expanded in powers of ui . The terms involving ui
describe the change in energies of the ions due to the lattice vibrations. The
expansion is

X P̂ 2 1 X
i
Ĥ = + Θ(Ri − Rj )
i
2M 2 i,j
1 X
+ (ui − uj ) . ∇ Θ(Ri − Rj )
2 i,j
 2
1 1 X
+ (ui − uj ) . ∇ Θ(Ri − Rj ) + ....
2 2! i,j
(1804)

However, since the Ri are equilibrium positions, and the total force on the atom
located at Ri is given by X
∇ Θ(Ri − Rj ) (1805)
j

the total force must vanish in equilibrium185 . Hence, the potential to second
order in u is just given by

V̂ = Veq + V̂harmonic (1806)

where the harmonic potential is given by


1 X X µ
V̂harmonic = (u − uµj ) Θνµ (Ri − Rj ) (uν,i − uν,j ) (1807)
4 i,j µ,ν i

In the above equation, the quantities Θνµ are defined in terms of the second
derivatives of the pair-potential

ν ν

Θµ (Ri − Rj ) = ∇µ ∇ Θ(Ri − Rj + ui − uj ) (1808)
u≡0

185 Jahn and Teller have shown, by only considering symmetry, that a crystal structure is
generally not stable if the electronic ground state is degenerate. The only exceptions occur
when the degeneracy is a two-fold degeneracy associated with an odd-number of electrons, or
when the system is linear. (H. A. Jahn and E. Teller, Proc. Roy. Soc. A 161, 220 (1937).)

539
The harmonic potential V̂harmonic is usually expressed directly in terms of the
displacements, and not their differences. The harmonic potential is written as
1 X X µ ν
V̂harmonic = u D (R − Rj ) uν,j (1809)
2 i,j µ,ν i µ i

where the dynamical matrix is given by


X
Dµν (Ri − Rj ) = δRi ,Rj Θνµ (Ri − R00 ) − Θνµ (Ri − Rj ) (1810)
R00

In general, when the interactions are not just pair-wise, the harmonic potential
can still be defined. The dynamical matrix is defined as a second derivative of
the total energy, with respect to the lattice displacements

∂2E
Dµν (Ri − Rj ) = (1811)
∂uµi ∂uν,j

The dynamical matrix Dµν (R − R0 ) is a symmetric matrix

Dµν (R − R0 ) = Dνµ (R − R0 ) (1812)

due to the analyticity of the pair-potential. Also, for a crystal with a monoatomic
basis, one has
Dµν (R − R0 ) = Dµν (R0 − R) (1813)
which follows since every Bravais lattice has inversion symmetry. Due to trans-
lational invariance, one has the sum rule
X
Dµν (R) = 0 (1814)
R

This follows from consideration of the absence of energy change due to a uniform
displacement of the solid

r(Ri ) = Ri + ∆ ∀i (1815)

Under this displacement, the energy change of the solid is zero, and so
1 X X µ ν
0 = ∆ Dµ (Ri − Rj ) ∆ν (1816)
2 i,j µ,ν

for an arbitrarily chosen ∆µ .

Since ui,µ represents the displacement canonically conjugate to P̂i,µ , the


momentum and displacement operators satisfy the commutation relations

[ ui,µ , P̂j,ν ] = i h̄ δi,j δµ,ν (1817)

540
and
[ ui,µ , uj,ν ] = [ P̂i,µ , P̂j,ν ] = 0 (1818)

To diagonalize the harmonic Hamiltonian, first a canonical transformation


is performed to a representation in which the periodic translational invariance
of the lattice is explicit. The displacement is expressed as
 
1 X
ui = √ uq exp i q . Ri (1819)
N q

and the momentum operator becomes


 
1 X
P̂ i = √ p̂q exp i q . Ri (1820)
N q

where N = N1 N2 N3 is the number of unit cells in the crystal186 On assuming


that the lattice displacements satisfy Born-von Karman boundary conditions187 ,
the wave vectors are quantized as
n1 n2 n3
q = b + b + b (1821)
N1 1 N2 2 N3 3
where ni are integers such that 0 < ni < Ni . In the Fourier transformed
basis, the commutation relations become

[ uq,µ , p̂q0 ,ν ] = i h̄ ∆q+q0 δµ,ν (1822)

and
[ uq,µ , uq0 ,ν ] = [ p̂q,µ , p̂q0 ,ν ] = 0 (1823)
The quantity ∆q is the Kronecker delta function which conserves wave vectors,
modulo reciprocal lattice vectors. The Kronecker delta function is defined via
X  
exp i q . R = N ∆q (1824)
Ri

186 It should be noted that the operators uq and p̂q are no longer Hermitean due to the
appearance of the phase factor. However,
u†q = u−q

and
p̂† = p̂
q −q
.
187 M. Born and Th. von Karman, Zeit. für Physik, 13, 297 (1912), M. Born and Th. von

Karman, Zeit. für Physik, 14, 15 (1913).

541
This is non-zero if q is a reciprocal lattice vector Q, and is zero otherwise. In
terms of the new coordinates, the Hamiltonian becomes
X  p̂†q,µ δµ,ν p̂q,ν 1

Ĥ = + u†q,µ Dνµ (q) uq,ν (1825)
q
2M 2

where D(q) is the Fourier Transform of the dynamical matrix


X  
D(q)
e = D(R)
e exp − i q . R (1826)
R

Thus, the harmonic Hamiltonian is diagonal in the quantum number q.

The symmetries of the dynamical matrix D(R)e can be used to show that
X  
D(q)
e = D(R
e
i − R j ) exp − i q . ( R i − R j )
i
    !
1 X e
= D(R) exp − iq.R + exp + iq.R − 2
2
R
q.R
X  
= −2 D(R)
e sin2 (1827)
2
R

Using the Laue condition, one finds that D(q)


e is invariant under translations
through reciprocal lattice vectors q → q + Q. Also, since D(R)
e is a real and
symmetric, then D(q) is a real and symmetric matrix. Every real 3×3 symmetric
e
matrix has three real eigenvalues, thus, one may find three eigenfunctions
D(q)
e α (q) = M ωα2 (q) α (q) (1828)
where α (q) are the eigenvectors and the eigenvalues have been written as
M ωα2 (q). Since the eigenvalues are real, the frequencies ωα (q) are either real or
purely imaginary. It is necessary that the eigenvalues are positive if the lattice
is to be stable.

The eigenvectors are usually normalized, such that


α (q) . β (q) = δα,β (1829)
Since the eigenvectors form a complete orthonormal set, one may expand the
displacements and the momentum operators in terms of the eigenvectors as
X
uq = Qαq α (q) (1830)
α

and X
p̂q = P̂qα α (q) (1831)
α

542
This transformation diagonalizes the Hamiltonian in terms of the polarization
index α. This can be seen as
X
u†q D(q)
e uq = Q†q β β (q) D(q)
e α (q) Qα
q
α,β
X
= M ωα2 (q) Q†q β
β (q) . α (q) Qα
q
α,β
X
= M ωα2 (q) Q†q α

q (1832)
α

and also X
p̂†q . p̂q = P̂q† α
P̂qα (1833)
α

Hence, the transformed Hamiltonian is diagonal in the polarization indices α

P̂q† α
P̂qα
" #
X M ωα2 (q) † α
Ĥ = + Qq Qα
q (1834)
q,α
2M 2

The Hamiltonian has the form of 3 N independent harmonic oscillators. The


eigenvalues of the Hamiltonian can be found by introducing boson annihilation
and creation operators. The annihilation operator is defined by
s s
M ωα (q) α 1
aq,α = Qq + i Pα (1835)
2 h̄ 2 h̄ M ωα (q) q

and the creation operator is the Hermitean conjugate of the annihilation oper-
ator s s

M ωα (q) † α 1
aq,α = Qq − i P† α (1836)
2 h̄ 2 h̄ M ωα (q) q
The commutation relations for the boson operators can be calculated from the
commutation relations of Pqα and Qα
q , so

[ aq,α , a†q0 ,β ] = ∆q−q0 δα,β (1837)

These are the usual commutation relations for boson creation and annihilation
operators. In second quantized form, the Hamiltonian is expressed as
X h̄ ωα (q)  
Ĥ = a†q,α aq,α + aq,α a†q,α (1838)
q,α
2

Due to the commutation relations, one can show that when the operator a†q,α
acts on an energy eigenstate it produces an energy eigenstate in which the energy
eigenvalue is increased by an amount h̄ ωα (q). Likewise, when the annihilation

543
operator acts on an energy eigenstate it produces an energy eigenstate with an
eigenvalue which is lower by h̄ ωα (q). If one assumes the existence of a ground
state of the oscillator | 0q,α > such that

aq,α | 0q,α > = 0 (1839)

then the energy eigenvalue of the ground state is just 12 h̄ ωα (q). The excited
states can be found by the raising operator to be ( nq,α + 12 ) h̄ ωα (q), where
nq,α is a positive integer. Thus, the Hamiltonian may be expressed as
 
X 1
Ĥ = h̄ ωα (q) nq,α + (1840)
q,α
2

This implies that each normal mode (q, α) has quantized excitations. These
quantized lattice vibrations are known as phonons. Ab-initio pseudo-potential
calculations of the pair-potential for simple metals lead to phonon frequencies
ωα (q) which are in good agreement with the results of inelastic neutron scat-
tering experiments188 . This agreement suggests that the effect of multi-atom
interactions on the phonon dispersion relations are relatively small.

The completeness relation for the polarization vectors is just


X
µα (q) αν (q) = δνµ (1841)
α

therefore, an arbitrary displacement can be expanded in terms of the polar-


ization vectors. The original displacements and momenta can be expressed in
terms of the phonon creation and annihilation operators via
s    
1 X h̄ †
ui = √  (q) aq,α + a−q,α exp i q . Ri
N q,α 2 M ωα (q) α

(1842)

and
s
h̄ M ωα (q)
   
i X
P̂ i = √ α (q) a†−q,α − aq,α exp i q . Ri
N q,α 2

(1843)

In the Heisenberg representation, the displacements and momentum operators


become time-dependent. The time dependence occurs through factors of
 
exp ± i ωα (q) t (1844)

188 L. J. Sham, Proc. Roy. Soc. A, 283, 33 (1965) and E. G. Brovman and Yu. M. Kagan,

Sov. Phys. J.E.T.P. 52, 557 (1967).

544
0.03

0.025

0.02
hωq

0.015

0.01

0.005

0
Γ H P Γ N H

Figure 229: A model calculation of the phonon dispersion relations for a hypo-
thetical metal with an unstable b.c.c. crystal structure. A phonon frequency
become imaginary close to the N point, signalling that the structure is unstable.

appearing along with the phonon creation and annihilation operators.

Excited lattice vibration normal modes that resemble classical waves, in that
they have a definite displacement and definite phase, cannot be energy eigen-
states of the Hamiltonian as they are not eigenstates of the number operator.
The classical lattice vibrations are best described in terms of coherent states
which are a superposition of states each containing large numbers of phonons.
The time dependence contained in the exponential phase factors has the effect
that the displacements associated with each excited normal mode oscillate peri-
odically with time. In general, the time-dependent factors have the effect that,
if the eigenvalues of the dynamical matrix are such that ωα2 (q) > 0, the various
expectation values of the displacements can simply be represented as a sum of
periodic oscillations. On the other hand, if ωα2 (q) < 0, the displacements have
unbounded exponential growth, the harmonic approximation fails, and the lat-
tice becomes unstable.

Sometimes, gradual structural changes are accompanied by softening of the


phonon dispersion relation for some values of q as the temperature is decreased.
For example, in SrT iO3 a phonon mode softens and at the transition tem-
perature, the associated displacement becomes static and is only limited by
anharmonic interactions189 . However, the phonon frequency does not need to
soften all the way down to zero if a structural changes is to occur. The tempera-
ture induced change of stability of the low-temperature close-packed structures
of the light alkaline metals to b.c.c. structures at higher temperatures has been
189 R. A. Cowley, W. J. L. Buyers and G. Dolling, Sol. State. Commun. 7, 181 (1969).

545
peak at low temperature indicates the martensitic transformation. reduced zone scheme for the B2 structure.
T M ↑ and T M ↓ indicate the temperatures at which the martensitic Most modes demonstrate only slight temperature depen-
phase disappears on warming up and first appears on cooling. ~b! dence easily explained by the overall stiffening of the lattice
Temperature dependence of the intensity in the ( 31, 13,0) elastic at low temperatures. The most striking feature, however, is
satellite in cubic Ni 2 MnGa. The peak is characteristic of the pre- the strong temperature dependence of the ( z , z ,0) TA 2
martensitic phase. branch with polarization along (110), discussed in detail in
our previous paper.17 As shown in the blowup in Fig. 3~a!
H4M and H8 triple-axis spectrometers at the HFBR at ~dashed lines!, the dispersion curve has a wiggle at z 0 ' 31 at
Brookhaven National Laboratory. The use of standard and room temperature, which develops into a distinct minimum
high-temperature Displex refrigerators enabled us to perform as the temperature is decreased. In the limit z →0 the
the measurements over a wide temperature range 12–400 K. ( z , z ,0) TA 2 mode corresponds to the elastic constant c 8 5
2 (c 112c 12), which is known to be reduced in bcc
1
Two sample settings were used. The crystal was aligned to
have either the (100) or the (110) planes coincide with the materials,20 and shows an anomalous decrease as T→T M in
scattering plane of the spectrometer. The sample mosaic re- Ni 2 MnGa ~Refs. 14 and 13!. For the (110) direction, the
veals two monocrystalline grains, misaligned by 0.5°, giving slopes in the longitudinal ( v L '5.43105 cm/s! and (001)
a full width at half maximum ~FWHM! of 1°. Neutron scat- transverse ( v T1 '3.53105 cm/s! dispersion curves at q→0
tering experiments were performed with fixed final or initial are in good agreement with experimental ultrasound

FIG. 2. Measured acoustic-phonon dispersion curves for the parent L2 1 phase of Ni 2 MnGa; the dash-dotted lines show the fcc zone
Figure 230: The measured acoustic phonon dispersion relations for N i M nGa
boundaries. Solid lines are guides to the eye and arrows indicate the observed phonon anomalies. 2
which undergoes a transition between an f.c.c. crystal structure at room tem-
perature to a teragonal low-temperature structure with a modulated distortion.
The phonon frequency at the modulation vector softens almost to zero as the
temperature is lowered to T = 240 K. [After Zheludev et al. Phys. Rev. B 54,
15045 (1996).]

[111] cos α = 1/3

α (1-10) α
[11-1]

(112)
[11-1]
√2a

[111] (-1-12)
√2a

√3a/2

Figure 231: A prism of the b.c.c. lattice which becomes the parent of an h.c.p.
unit cell in the b.c.c. → h.c.p. transition. [After W.G. Burgers. Physica 1, 561
(1934).]

546
cos α = 1/3
[111] (1-10) α

(112)
(-1-12)
[11-1]

a'

π/3

c'=√8/3 a'

Figure 232: The b.c.c. → h.c.p. transition involves a combination of a uniform


shear and a shuffle from the quasi-static transverse 12 (1, 1, 0) phonon modes.
[After W.G. Burgers. Physica 1, 561 (1934).]

547
attributed190 to the existence of a low-energy transverse phonon mode in the
open b.c.c. structure. Due to the low-energy of the phonon mode, an increase
in temperature produces a rapid increase in entropy which stabilizes the high-
temperature b.c.c. phase. Henceforth, the discussion will be restricted to the
case of stable lattice structures.

14.1 Lattice with a Basis


When the lattice has a basis composed of p atoms, the analysis of the phonon
excitations is similar to the analysis for a crystal with a mono-atomic basis. The
displacements are labelled by the lattice vector index i and also by an index j
which labels the atoms in the basis. Also, the dynamical matrix becomes a
3 p × 3 p matrix and there are 3 p normal modes labelled by α. Thus, one has
 
1 XX α j
uji = √ Qq α (q) exp i q . Ri (1845)
N q α

The polarization vectors jα (q) generally are complex for a lattice without inver-
sion symmetry. The polarization vectors satisfy the generalized orthonormality
condition
j=p
X
jβ (q)∗ . jα (q) Mj = δα,β (1846)
j=1

where Mj is the mass of the j-th atom in the basis.

The dynamical matrix for an lattice with a basis can be expressed in terms
of the pair-potential. The Fourier transform of the pair potential Θ(k) is intro-
duced via the definition
X  
Θ(R) = Θ(k) exp i k . R (1847)
k

so that Z  
1
Θ(k) = d3 R Θ(R) exp − ik.R (1848)
V
For a simple metal, it might be permissable to approximate the pair potential
by the screened Coulomb interaction

1 4 π Z 2 e2 e
 
Θ(k) = V0 (k)2 (1849)
V k 2 (k)

where (k) is the static dielectric constant. The real space dynamical matrix
can be expressed in terms of the real space pair-potential, which in turn can be
190 C. Zener, Phys. Rev. 71, 846 (1947).

548
expressed in terms of the definition of its Fourier transform. On utilizing the
Laue identity
X   X
exp i ( k − q ) . R = N δk−q,Q (1850)
R Q

one finds that the dynamical matrix can then be written in terms of a sum over
reciprocal lattice vectors as
X  
µ,j µ
Dν,j 0 (q) = N ( Q + q ) exp i ( Q + q ) . ( rj − rj 0 ) ( Q + q )ν Θ(Q + q)
Q
X  
−N δjj0 (Q) µ
exp i Q . ( rj − rj 00 ) ( Q )ν Θ(Q) (1851)
Q,00

which now involves the set of p basis vectors rj . The eigenvalues of the 3p × 3p
dynamical matrix yields the squares of the frequencies, ωα (q) for the 3p phonon
modes.

14.2 A Sum Rule for the Dispersion Relations


The eigenvalue equation for the phonon frequencies

D(q)
e α (q) = M ωα2 (q) α (q) (1852)

can be related to the Fourier transform of the pair-potential. The Fourier Trans-
form of the dynamical matrix is given by
X  
D(q)
e = D(R)
e exp − i q . R (1853)
R

Hence, one has


X  
D(R)
e exp − iq.R α (q) = M ωα2 (q) α (q) (1854)
R

The dynamical operator is given in terms of the pair-potential by


X
Dµν (Ri ) = δRi ,0 ∇µ ∇ν Θ(0 − R00 ) − ∇µ ∇ν Θ(Ri )
R00
X
= δRi ,0 ∇µ ∇ν Θ(R00 ) − ∇µ ∇ν Θ(Ri ) (1855)
R00

and the Fourier Transform of the pair-potential is given by


X  
Θ(R) = exp i k . R Θ(k) (1856)
k

549
On substituting the expression for the Fourier transform of the pair-potential
into the expression for the Fourier transform of the dynamical matrix, one ob-
tains
X X  
ν ν
Dµ (q) = kµ k Θ(k) exp i ( k − q ) . Ri
Ri k
X X  
ν 00
− kµ k Θ(k) exp ik.R (1857)
R00 k

Thus, the phonon frequencies satisfy the eigenvalue equation


X X  
M ωα2 (q) α (q) = k Θ(k) exp i ( k − q ) . Ri k . α (q)
i k
X X  
− k Θ(k) exp i k . Ri k . α (q)
i k

(1858)

On making use of the conservation of momentum, modulo the reciprocal lattice


vectors Q, one has
X
M ωα2 (q) α (q) = N ( q + Q ) Θ(q + Q) ( q + Q ) . α (q)
Q
X
−N Q Θ(Q) ( Q . α (q) ) (1859)
Q

It can be seen that the transverse modes only exist because of the periodicity of
the lattice. That is, if only the Q = 0 reciprocal lattice vector were included
in the sum, the eigenvectors would be longitudinal as (q) would be parallel to
q. In this case,

M ωα2 (q) α (q) = N q Θ(q) q . α (q) (1860)

which are the longitudinal sound modes. Hence, the transverse modes only ex-
ist because of the periodicity of the lattice. Furthermore, by letting q → 0,
one finds that the phonon frequencies M ωα2 (0) must vanish in this limit, if the
interactions are sufficiently short ranged so |Θ(0)| < ∞. The transverse modes
are examples of Goldstone modes. The transverse modes occur because the
continuous translational symmetry of the Hamiltonian is spontaneously broken
at the phase transition when the solid is formed. Goldstone’s theorem191 may
be roughly stated as “When a continuous symmetry of the Hamiltonian is spon-
taneously broken in a phase transition, a continuous branch of normal modes
appears which extend to zero-energy. These normal modes dynamically restore
191 J. Goldstone, Nuovo Cimento 19, 154 (1961), Y. Nambu, Phys. Rev. Lett. 4, 380 (1960).

550
the broken symmetry.” Goldstone’s theorem also depends on the condition that
long-ranged forces are not present. If long-ranged forces are present, the sym-
metry restoring mode may acquire a finite frequency at q = 0 through the
Kibble-Higgs mechanism192 . In this case, the resulting modes are called Higgs
bosons193 .

The Sum Rule.

The Sum Rule is obtained by using the orthogonality relations for the polar-
ization vectors. On taking the scalar product of the eigenvalue equation for the
dispersion relation with the eigenvector eα (q) and summing over the polarization
index α, one obtains
X X X
M ωα2 (q) = N α (q) . ( q + Q ) Θ(q + Q) ( q + Q ) . α (q)
α α Q
X X
−N α (q) . ( Q ) Θ(Q) ( Q ) . α (q)
α Q

(1861)

Utilizing the completeness relation for the components of the polarization vec-
tors X
µα (q) α,ν (q) = δνµ (1862)
α

one finds the sum rule


X X
M ωα2 (q) = N Θ(q + Q) ( q + Q )2
α Q
X
−N Θ(Q) ( Q )2 (1863)
Q

for the phonon frequencies.

——————————————————————————————————

14.2.1 Exercise 74
Show that if the pair-potential is approximated by the non-screened Coulomb
potential between the charged nuclei, one finds
X 4 π N Z 2 e2
ωα2 (q) = (1864)
α
M V
192 T. W. B. Kibble, Proc. Oxford Int. Conf. on Elem. Particles, (1965).
193 P. W. Higgs, Phys. Rev. Lett. 13, 508, (1964).

551
which defines the plasmon frequency for the ions. In the long wavelength limit,
one has the longitudinal plasmon mode and at most, two transverse modes de-
pending on the presence of long-ranged order. As the longitudinal plasmon
mode saturates the sum rule at q = 0, the transverse modes, if they exist,
must be acoustic.

——————————————————————————————————

14.3 The Nature of the Phonon Modes


The long wavelength form of the dynamical matrix can easily be calculated from
q.R
X  
D(q)
e = −2 D(R)
e sin2 (1865)
2
R
as  2
1 X e
D(q) = −
e D(R) q . R (1866)
2
R
This implies that ωα (q)2 ∝ q 2 as q → 0, which gives rise to acoustic modes.
These acoustic modes include the two Goldstone modes as well as the longitu-
dinal sound mode, which can be considered as a density fluctuation similar to
the sound waves found in fluids. If the crystal is anisotropic, the frequency or
sound velocity may depend on the direction of propagation.

In an isotropic solid, there should be one longitudinal and two transverse


polarizations ( k q) and ( ⊥ q). In an anisotropic solid, the relation
between  and q is not so simple, except at high-symmetry points. However,
because the polarization vectors are continuous functions of q, one may still use
the terminology of longitudinal and transverse polarizations in the vicinity of
the high-symmetry points.

A solid with a p atom basis has 3 N p degrees of freedom, 3 N of which


are tied up in the acoustic phonon branches. The other 3 ( p − 1 ) N modes
appear as optic branches.

The phonon density of states is given by an integral over the first Brillouin
zone, and a summation over the polarization index α.
V X Z
ρ(ω) = d3 q δ( ω − ωα (q) ) (1867)
( 2 π )3 α

This can be written as an integral over a surface of constant ωα (q). This surface
is denoted by Sα (ω) which consists of the points ω = ωα (q), where q is in the
first Brillouin zone. This yields
V X Z 1
ρ(ω) = 3
d2 S (1868)
(2π) α Sα (ω) | ∇ωα (q) |

552
(1,0,0) (1/2,1/2,1/2) (1/2,1/2,0)
0.015

L
L
0.01 T
hωq [eV]

0.005 L
T
T
L

T
0
Γ H P Γ N
Figure 233: A model calculation of the phonon dispersion relations for a metal
with a b.c.c. crystal structure. The screening was treated in the Random Phase
Approximation.

(1,0,0) (1,1/2,0) (3/4,3/4,0) (1/2,1/2,1/2)


0.06

0.04
hωq [eV]

0.02

0
Γ X W K Γ L
Figure 234: A model calculation of the phonon dispersion relations for a metal
with an f.c.c. crystal structure. The screening was treated in the Random Phase
Approximation.

553
Figure 235: The phonon dispersion relations for h.c.p. M g measured along the
ΓK direction. [After R. Pynn and G. L. Squires, Proc. Roy. Soc. 326, 347
(1972).]

The van Hove singularities occur when the group velocity vanishes,

∇ωα (q) = 0 (1869)

The occurrence of these singularities in the phonon density of states has been
studied via topological arguments194 . The van Hove singularities are usually
integrable in three dimensions, but still give rise to anomalous slopes or discon-
tinuities in the derivatives of ρ(ω). An example of the van Hove singularities in
the phonon modes is given by Cu which has two transverse and one longitudi-
nal contribution to the density of states. The phonon density of states has been
constructed from the results of inelastic neutron scattering measurements195 .

——————————————————————————————————

14.3.1 Exercise 75
Consider a one-dimensional linear chain, with a unit cell composed of two atoms,
one with mass M1 and the other with mass M2 . The atoms interact with their
nearest neighbors via a harmonic force, with force constant γ. Find the phonon
dispersion relation.

——————————————————————————————————

194 L. van Hove, Phys. Rev. 89, 1189 (1953).


195 E. C. Svensson, B. N. Brockhouse and J. M. Rowe, Phys. Rev. 155, 619 (1967).

554
Figure 236: The phonon density of states ρ(ω) constructed from the data ob-
tained from inelastic neutron scatteringe experiments on Cu. [After Svensson
et al., Phys. Rev. 155, 619 (1967).]

14.3.2 Exercise 76
Consider a one-dimensional line of ions, with equal masses but alternating
charges, such that the charge on the n-th ion is
en = e ( − 1 )n (1870)
Assume that the inter-atomic potential has two contributions:

(A) A short-ranged force between nearest neighbors with a force constant


C1 = γ.

e2
(B) A Coulomb interaction between all the ions Cn = 2 ( − 1 )n n3 a3
where a is the atomic spacing.

(i) Show that



ω(q)2 ( − 1 )n
 
2 q a
X
= sin + σ 1 − cos q n a (1871)
ω02 2 n=1
n3

γ e2
where ω02 = 4 M and σ = γ a3 .

π 4
(ii) Show that ω 2 (q) becomes soft ω 2 (q) = 0 at q = a if σ > 7 ζ(3).

1
(iii) Show that the speed of sound becomes imaginary if σ > 2 ln 2 .

555
Thus, ω 2 goes to zero and the lattice becomes unstable for q in the interval
(0, π) if σ lies in the range 0.475 < σ < 0.721.

——————————————————————————————————

14.3.3 Exercise 77
Consider a two-dimensional square lattice with a mono-atomic basis. The atoms
have mass M and interact with their nearest neighbors and next nearest neigh-
bors through a harmonic force of strength γ1 and γ2 , respectively. Calculate
the frequencies of the longitudinal and transverse phonons at q = ( πa ) (1, 0).

——————————————————————————————————

14.3.4 Exercise 78
(i) Show that the linear chain with nearest neighbor (harmonic) interactions has
a dispersion relation
qa
ω(q) = ω0 | sin | (1872)
2
and that the density of states is given by
2 1
ρ(ω) = p (1873)
πa ω02 − ω 2

which has a van Hove singularity at ω = ω0 .

(ii) Show that in three dimensions, the van Hove singularities near a maximum
of ωα (q) gives rise to a term in the density of states that varies as
q
ρ(ω) ∝ ω02 − ω 2 (1874)

and, thus, has a singularity in the first derivative of the density of states with
respect to ω.

——————————————————————————————————

14.3.5 Exercise 79
(i) Show that if the wave vector q lies along a 3 , 4 or 6 fold axis, then one
normal mode is polarized along q and the other two modes are degenerate and
polarized perpendicular to q.

556
(ii) Show that if q lies in a plane of mirror symmetry, then one mode has a
polarization perpendicular to q and the plane, and the other two modes have
polarizations within the plane.

(iii) Show that if q lies on a Bragg plane that is parallel to a plane of mirror
symmetry, then one mode is polarized perpendicular to the Bragg plane, while
the other two modes have polarizations lying within the plane.

——————————————————————————————————

14.3.6 Exercise 80
Consider an f.c.c. mono-atomic Bravais lattice in which the atoms interact via
a nearest neighbor pair-potential Θ.

(i) Show that the frequencies of the phonon modes are given by the eigen-
values of a 3 × 3 matrix given by
q.R
X   
2
D(q) =
e D(R) sin
e A I + B R̂ R̂
e (1875)
2
R

where the sum over R runs over the 12 lattice sites closest to the site R = 0,
and the constants A and B are given in terms of the pair-potential and its
derivatives at the nearest neighbor separation d = √a2 via

A = 2 Θ0 (d)/d (1876)

and  
B = 2 Θ”(d) − Θ0 (d)/d (1877)

(ii) Show that when q = (q, 0, 0), the longitudinal and transverse acoustic
phonon frequencies are given by
r
8A + 4B qa
ωl (q) = sin (1878)
M 4
and r
8A + 2B qa
ωt (q) = sin (1879)
M 4

(iii) Find the frequencies when q = √q (1, 1, 1).


3

557
(iv) Show that when q = √q2 (1, 1, 0) then the degeneracy between the
transverse modes is lifted and the frequencies are given by
s
8A + 2B qa 2A + 2B qa
ωl (q) = sin2 √ + sin2 √ (1880)
M 4 2 M 2 2

and the two transverse modes are


s
8A + 4B qa 2A qa
1
ωt (q) = sin2 √ + sin2 √ (1881)
M 4 2 M 2 2

and s
8A + 2B qa 2A qa
ωt2 (q) = sin2 √ + sin2 √ (1882)
M 4 2 M 2 2

——————————————————————————————————

14.3.7 Exercise 81
Consider a phonon with a wave vector along the axis of a cubic crystal. Then
consider the sums in
q.R
X  
D(q)
e = D(R)
e sin2 (1883)
2
R

be restricted to the sites in two planes perpendicular to q separated by a distance


q̂ . R. In metals, there exists a long-ranged interaction between the planes

q.R
X  
D(q)
e = D(R)
e sin2
2
R
sin 2kF q̂ . R
= A (1884)
2kF q̂ . R
where A is a constant.
∂ω 2 (q)
(i) Find an expression for ω 2 (q) and ∂q .

∂ω 2 (q)
(ii) Show that ∂q is infinite at q = 2 kF . The kink in the dispersion
relation at the Fermi wave vector is the Kohn anomaly.

——————————————————————————————————

558
14.4 Thermodynamics
A harmonic lattice has an energy given by
 
X 1
Eharmonic = Ecl + h̄ ωα (q) nq,α + (1885)
q,α
2

where Ecl is the ground state energy of the lattice in the classical approximation.
At finite temperatures, the energy is to be replaced by a thermal average. The
thermal average of the number of phonons in mode (q, α) is denoted by nq,α
and is calculated as the Boltzmann weighted average. Since the energies of the
harmonic phonons are additive, then the Boltzmann factor can be expressed as
a product of Boltzamnn factors for each normal mode.
1 Y 1 1
exp[ − β Eharmonic ] = exp[ − β h̄ ωα (q) ( nq,α + ) ] (1886)
Z q,α
Zα (q) 2

This implies that the thermal average can be performed independently for each
phonon mode. Therefore, we can write

1 X 1
nq,α = n exp[ − β ( n + ) h̄ ωα (q) ]
Zα (q) n=0 2
3
1 exp[ − 2 β h̄ ωα (q) ]
= 2
Zα (q)

1 − exp[ − β h̄ ωα (q) ]

(1887)

However, the partition function for a single phonon mode Zα (q) is given by the
normalization condition for the probability

X 1
Zα (q) = exp[ − β (n + ) h̄ ωα (q) ]
n=0
2
exp[ − 12 β h̄ ωα (q) ]
= (1888)
1 − exp[ − β h̄ ωα (q) ]

We note that the contribution from the zero point energy drops out of the ratio.
Thus, the thermal average of the number of phonons is given by
1
nq,α = N (ωα (q)) = (1889)
exp[ β h̄ ωα (q) ] − 1

where N (ω) is the Bose-Einstein distribution function. Therefore, the thermal


averaged energy of the harmonic lattice is given by
 
X 1
Eharmonic = Ecl + h̄ ωα (q) N (ωα (q)) + (1890)
q,α
2

559
The Helmholtz free energy F of the lattice is defined in terms of the partition
function, Z, by
 
Z = exp − β F
  Y nq,α =∞  
X 1
= exp − β Ecl exp − β h̄ ωα (q) ( nq,α + )
q,α nq,α =0
2
!
exp[ − 12 β h̄ ωα (q) ]
  Y 
= exp − β Ecl (1891)
q,α
1 − exp[ − β h̄ ωα (q) ]

Thus, the Helmholtz free energy F is given by


X h̄ ωα (q) X   
F = Ecl + + kB T ln 1 − exp − β h̄ ωα (q) (1892)
q,α
2 q,α

The pressure P is found from the infinitesimal thermodynamic relation


dF = dE − T dS − S dT
dF = − S dT − P dV (1893)
Hence, the pressure is determined by
 
∂F
P = −
∂V T,N
1 X ∂h̄ωα (q)
   
∂Ecl
= − −
∂V T,N 2 q,α ∂V T,N

X ∂h̄ωα (q)
 
1
−   (1894)
∂V T,N exp β h̄ ω (q)
q,α α − 1

The first two terms are temperature independent, and the last term depends on
temperature through the average phonon occupation numbers. The pressure is
only temperature dependent if the phonon frequencies depend on the volume V .

The thermal volume expansion coefficient α is defined by


 
1 ∂V
α = (1895)
V ∂T P
As the equation of state is a relation between pressure, temperature and volume
P = P (T, V ) (1896)
then the infinitesimal derivatives are related by
   
∂P ∂P
dP = dT + dV (1897)
∂T V ∂V T

560
For a process at constant P , dP = 0, so one has the relation
 
∂P
∂T
 
∂V
= −  V (1898)
∂T P ∂P
∂V
T

The bulk modulus, B, defined by


 
∂P
B = −V (1899)
∂V T

is finite as Ecl is expected to be volume-dependent. Hence, the denominator is


finite. The thermal expansion coefficient is non-zero, only if ( ∂P
∂T )V 6= 0. Using
the harmonic approximation, the frequencies must be functions of the volume
V if the solid is to undergo thermal expansion.

The specific heat at constant pressure is different from the specific heat at
constant volume. The difference is found by relating the temperature derivative
of the entropy with respect to temperature, at constant volume, to the tempera-
ture derivative of the entropy with respect to temperature, at constant pressure.
The relation is found by considering the infinitesimal change in entropy, with
either the change in volume or pressure
   
∂S ∂S
dS = dT + dV
∂T V ∂V T
   
∂S ∂S
= dT + dP (1900)
∂T P ∂P T

Using the equation of state relating P and V

V = V (T, P ) (1901)

one finds    
∂V ∂V
dV = dT + dP (1902)
∂T P ∂P T
Thus, on combining the expression for dS in terms of dV and dT with the above
equation for dV , one obtains the expression
          
∂S ∂S ∂V ∂S ∂V
dS = + dT + dP
∂T V ∂V T ∂T P ∂V T ∂P T
(1903)

Thus, one has the relation


       
∂S ∂S ∂S ∂V
= + (1904)
∂T P ∂T V ∂V T ∂T P

561
∂S

A Maxwell relation can be used to eliminate ∂V T
. The Maxwell relation comes
from the analyticity condition on a thermodynamic function with independent
variables (T, V ), which is F (T, V ). Hence, one has
   
∂S ∂P
= (1905)
∂V T ∂T V

and, thus,
        
∂S ∂S ∂P ∂V
T = T +
∂T P ∂T
V ∂T ∂T P
   V
∂P ∂V
CP = CV + T
∂T V ∂T P
 2
∂P
∂T
CP = CV − T  V (1906)
∂P
∂V
T

Therefore, if there is a difference between CP and CV , then the phonon frequen-


cies must be dependent on V .

14.4.1 The Specific Heat


The specific heat at constant volume can be found from the entropy of the
phonon gas
X     
S = kB N (ωα (q)) + 1 ln N (ωα (q)) + 1 − N (ωα (q)) ln N (ωα (q))
q,α
(1907)
In this expression N (ω) is the Bose-Einstein distribution function given by
1
N (ω) = (1908)
exp[ β h̄ ω ] − 1

and β is proportional to the inverse temperature, i.e., β −1 = kB T . The


specific heat is given by
 
∂S
CV = T
∂T V
!
X ∂N (ωα (q)) N (ωα (q)) + 1
= kB T ln
q,α
∂T N (ωα (q))
X ∂N (ωα (q))
= h̄ ωα (q)
q,α
∂T

562
X  2  
= kB β h̄ ωα (q) N (ωα (q)) N (ωα (q)) + 1
q,α

(1909)
This can be expressed as an integral over the phonon density of states ρ(ω) via
Z ∞  2  
CV = kB dω ρ(ω) β h̄ ω N (ω) N (ω) + 1 (1910)
0

The experimentally determined temperature dependence of the specific heat


may be compared with the results of the above formula, where the phonon den-
sity of states is inferred from experiments196 . There is a small discrepancy due
to anharmonic effects which have been neglected in the above analysis. We shall
examine the specific heat using some approximate models of the phonon density
of states.

14.4.2 The Einstein Model of a Solid


The Einstein model of a solid considers the phonons to have a fixed frequency
ω0 for all q vectors, and is an approximate representation of the optic phonons.
In the Einstein model, the phonon density of states is given by
ρ(ω) = 3 N δ(ω − ω0 ) (1911)
where there are 3 modes per atom. The specific heat is given by
 2  
CV = 3 N kB β h̄ ω0 N (ω0 ) N (ω0 ) + 1 (1912)

This vanishes exponentially at low temperatures, kB T  h̄ ω0 , where


 
N (ω0 ) ≈ exp − β h̄ ω0 (1913)

and at high temperatures kB T  h̄ ω0


kB T
N (ω0 ) ≈ (1914)
h̄ ω0
so the specific heat saturates to yield the classical result
lim CV → 3 N kB (1915)
T → ∞

The Einstein model of the specific heat fails to describe the lattice contribution
to low-temperature specific heat of a solid. This is because it fails to describe the
low-energy acoustic phonon excitations which gives rise to a power law tempera-
ture variation. The Debye model of a solid provides an approximate description
of the low-temperature specific heat of a solid.

196 E. C. Svensson, B. N. Brockhouse and J. M. Rowe, Phys. Rev. 155, 619 (1967).

563
But Let’s Take a Closer Look:
High T beha
Reasonable
agreement w
experiment

Low T beha
CV → 0 too
as T → 0 !
Figure 237: The specific heat of diamond compared with the results of the
Einstein Model, with ΘE = 1320. [After A. Einstein, Ann. Physik 22, 180
(1907).]

14.4.3 The Debye Model of a Solid


The Debye model of a solid approximates the phonon density of states for the
two transverse acoustic mode and longitudinal acoustic mode in an isotropic
solid. The dispersion relations of the phonon modes are represented by the
two-fold degenerate transverse mode
ωT (q) = vT q (1916)
and the singly degenerate longitudinal mode
ωL (q) = vL q (1917)
The transverse sound velocity, in general, will be different from the longitudinal
sound velocity, vL 6= vT . There are 3N such phonon modes in the first Brillouin
zone. The phonon density of states is given by the integral over a surface area
in the first Brillouin zone
X Z  −1
V 2 dωα
ρ(ω) = d Sα (ω) (1918)
( 2 π )3 α dq

If the Brillouin zone is approximated as a sphere of radius qD , then the density


of states is given by
 −1
V X
2 dωα
ρ(ω) = 4 π q (1919)
( 2 π )3 α dq

for q < qD . Using the form of the dispersion relation, the density of states can
be re-written as
V X ω2
ρ(ω) = Θ( vα qD − ω ) (1920)
( 2 π 2 ) α vα3

564
This can be further approximated by requiring that the upper limit on the
frequency of all three phonon modes be set to the Debye frequency ωD . In this
case, the density of states is simply given by
V X ω2
ρD (ω) = Θ( ωD − ω ) (1921)
( 2 π 2 ) α vα3

The value of the Debye frequency is determined by the condition


Z ωD
dω ρD (ω) = 3 N
0
V X ω3
D
= (1922)
( 6 π 2 ) α vα3

Using this condition, the Debye model for the phonon density of states is written
as
ω2
ρD (ω) = 9 N 3 Θ( ωD − ω )
ωD
(1923)

Thus, in the Debye model, the phonon density of states varies as ω 2 at low
frequencies and has a cut-off at the maximum frequency ωD .

The temperature dependence of the specific heat of the Debye model is given
by
Z ωD  2  
9 N kB 2
CV = 3 dω ω β h̄ ω N (ω) N (ω) + 1 (1924)
ωD 0

The asymptotic low-temperature variation of the specific heat can be found by


changing variable x = β h̄ω. The specific heat can be written in the form
 3 Z xD
kB T exp[ x ]
CV = 9 N kB dx x4  2 (1925)
h̄ ωD 0
exp[ x ] − 1

where the upper limit of integration is given by xD = β h̄ωD . At sufficiently


low temperatures, kB T  h̄ ωD , the upper limit may be set to infinity yielding
 3 Z ∞
kB T exp[ x ]
CV = 9 N kB dx x4  2
h̄ ωD 0
exp[ x ] − 1
3
12 π 4

kB T
= N kB (1926)
5 h̄ ωD
4
where the integral has been evaluated as 415π . Thus, the low-temperature spe-
cific heat varies as T 3 in agreement with experiment. The asymptotic high

565
Debye Model at
low T: Theory
vs. Expt.
Quite impressive
agreement with
predicted CV ∝ T3
dependence for Ar!
(noble gas solid)

(See SSS program


debye to make a
similar comparison
for Al, Cu and Pb)
Figure 238: The low temperature specific heat of solid argon plotted against
T 3 . [After L. Finegold and N. E. Phillips, Phys. Rev. 177, 1383 (1969).]

temperature specific heat, for kB T  h̄ ωD , is found from


Z ωD  2  
9 N kB 2
CV = 3 dω ω β h̄ ω N (ω) N (ω) + 1 (1927)
ωD 0
kB T
noting that the number of phonons is given by N (ω) = h̄ ω , so
Z ωD
9 N kB
CV = 3 dω ω 2
ωD 0
= 3 N kB (1928)
which is the classical limit. Thus, the Debye approximation provides an interpo-
lation between the low-temperature limit and the high-temperature limit, which
is only governed by one parameter, the Debye temperature kB TD = h̄ ωD .

——————————————————————————————————

14.4.4 Exercise 82
Evaluate the integral
Z ∞
exp[ x ]
dx x4  2 (1929)
0
exp[ x ] − 1

needed in the low-temperature limit of the specific heat for the Debye model.

566
e Model:
y vs. Expt.
greement
stein
t low T
Figure 239: The specific heats of Cu, Ag, P b and C. Diamond has an unusually
high Debye temperature.

al behavior ——————————————————————————————————
olids!
14.4.5 Exercise 83
Generalize the Debye model to a d-dimensional solid. Determine the high-
temperature and leading low-temperature variation of the specific heat due to
emperature lattice vibrations.
d to ——————————————————————————————————
s” of solid,
ted 14.4.6 Exercise 84
Show that the leading high temperature correction to the Dulong and Petit
value of the specific heat due to lattice vibrations is given by
 2
h̄ ω
R
dω kB T ρ(ω)
∆CV 1
= − R (1930)
CV 12 dω ρ(ω)

Also, evaluate the moment of the phonon density of states


Z
dω ω 2 ρ(ω) (1931)

in terms of the pair-potentials between the ions.

——————————————————————————————————

567
14.4.7 Exercise 85
Numerically calculate the phonon density of states for a single phonon mode for
a two-dimensional lattice with a dispersion
 
2 2
ω (q) = ω0 2 − cos qx a − cos qy a (1932)

and hence, obtain the temperature dependence of the specific heat. Compare
this with the numerical evaluation of an appropriate Debye model.

——————————————————————————————————

14.4.8 Lindemann Theory of Melting


Lindemann197 assumed that a lattice melts when the displacements due to lat-
tice vibrations become comparable to the lattice constants. Although this the-
ory does not address the appropriate mechanism, it does give the right order of
magnitude for simple metals and transition metals. It is assumed that melting
occurs at a critical value of the ratio

u2i
γ = (1933)
a2i

The melting occurs when the temperature dependent function γ is equal to a


critical value γc which is expected to be of the order of unity. The function γ is
given by
h̄ X 2 nq,α + 1
γ = (1934)
2 M N a2 q,α ωα (q)

in which nq,α is the thermal average of the number of phonons in the mode
(q, α). The thermally averaged number of phonons of frequency ωα (q) is given
by the Bose-Einstein distribution function
1
N (ωα (q)) = (1935)
exp[ β h̄ ωα (q) ] − 1
so
β h̄ ωα (q)
 
2 N (ωα (q)) + 1 = coth (1936)
2
Hence, the temperature dependent function γ can be expressed as

h̄ X coth βh̄ωα (q)


2
γ = (1937)
2 M N a2 q,α ωα (q)

197 F. A. Lindemann, Z. Phys. 11, 609 (1910).

568
Using the Debye model, the right hand side can easily be evaluated in two
limits: the zero temperature limit and the high temperature limit. In the limit
of zero temperature, the mean squared deviation is only due to the zero point
fluctuations of the harmonic phonons. The low temperature limit of the relative
mean squared deviation is given by
h̄ X 1
γ =
2 M N a2 q,α vα q
Z
3 h̄ V 1
= 2 3
d3 q
2M N va (2π) q
3 h̄ V 2
= 2 π qD (1938)
2 M N v a2 ( 2 π )3

Using the relation


3
V qD 3V
N = 2
= (1939)
6π 4 π a3
one obtains
  13
1 9 h̄ qD
γ ∼
2 2 π2 M v
h̄ qD kB TD
∼ 0.4 = 0.4 (1940)
M v M v2
1
It can be shown that the right hand side is proportional to M − 2 . Therefore, the
above relation implies that if an element is to solidify, it has to have an atomic
mass which is greater than a critical value. At high temperatures, the relative
mean squared displacement is dominated by the thermally activated phonons
and is given by
Z qD
h̄ V 6 kB T
γ = 2 2
dq q 2
2M N a 2π 0 h̄ v 2 q 2
3 V kB T qD
=
M a2 2 π 2 N v2
  13
9 kB T 36 kB T
= 2 =
M a2 qD v2 π2 M v2
kB T h̄2 qD
2
∼ 1.54 = 1.54 2 T 2 kB T (1941)
M v2 M kB D

The Lindemann criterion provides a relation between the Debye temperature


and the melting temperature. The experimental data for alkaline metals and
1
transition metals suggest that γc has a value near 16 , independent of the
198
metal . Of course, one expects that anharmonic effects may become impor-
tant for large displacements of the ions from their equilibrium positions, and
198 G. Grimvall and S. Sjodin, Physica Scripta, 10, 340, (1974).

569
1.5

1
Tc

γ/γc
0.5

0
0 0.5 1 1.5 2 2.5 3
(T/TD)

Figure 240: The schematic temperature variation of the function γ used in the
Lindemann criterion for melting.

Figure 241: The variation of the melting temperature with the square of the
Debye temperature. [After Grimvall and Sjodin (1974).]

570
one expects that the phase transition will involve the collective behavior of the
atoms. In particular, one might expect that the Debye temperature should de-
crease as the phase transition is approached199 .

Mermin-Wagner Theorem.

The Lindemann theory of melting may be extended to provide an example


of the Mermin-Wagner theorem200 . The Mermin-Wagner theorem states that
finite temperature phase transitions, in which a continuous symmetry is sponta-
neously broken, cannot occur in lower than three dimensions. Basically, if such
a transition occurs, then there should be a branch of Goldstone modes that dy-
namically restores the spontaneously broken symmetry201 . These normal modes
produce fluctuations in the order parameter. In a periodic solid where continu-
ous translational invariance is broken, the Goldstone modes are the transverse
sound waves. The transverse sound modes have dispersion relations of the form
ω(q) = v q. The fluctuations in the order parameter are the fluctuations in the
choice of origin of the lattice and, therefore, are just the fluctuations in positions
of any one ion. In d dimensions, at finite temperatures, the fluctuations have
contributions from the region of small q which are proportional to
Z qD
h̄ kB T
2
ui ∼ dd q
0 2 M ω(q) h̄ ω(q)
Z qD
∼ dq q d−3 (1942)
0

where qD is a cut off due to the lattice. The integral diverges logarithmically
for d = 2, indicating that the fluctuations in the equilibrium lattice positions
will be infinitely large, thereby preventing the solid from being formed. Like-
wise, for lower dimensions such as one dimension, the integral will also diverge
at the lower limit. Therefore, no truly one-dimensional solid is stable against
temperature-induced fluctuations. An analysis of the zero point fluctuations
also rules out the possibility of a one-dimensional lattice forming in the limit of
zero temperature.

For a harmonic solid, the phonon frequencies are independent of the volume
V . This can be seen by considering the energy of a solid which has expanded
in the linear dimensions by an amount proportional to .

199 An alternate melting criterion was suggested by Born. Born’s criterion [M. Born, J.

Chem. Phys., 7, 591, (1939).] is based on the observation that solids differ from liquids
by their rigidity or resistance to shear stresses. Born postulated that the velocity of shear
waves would go to zero at the melting temperature. However, measurements of the shear force
constant shows that although it decreases as the temperature is increased towards the melting
temperature, it does remains finite at the melting temperature.
200 N. D. Mermin and H. Wagner, Phys. Rev. Lett. 17, 1133 (1966).
201 J. Goldstone, Nuovo Cimento 19, 154 (1961).

571
The energy of a harmonic solid with static displacements about the original
equilibrium position is given by the harmonic expression
1 X
E = Eeq + u D(R i − R j ) uj (1943)
e
2 i,j i

Now consider the expanded lattice in which the displacements are given by

ui =  Ri + ũi (1944)

Here, ũi are the new displacements from the new lattice sites of the lattice which
has undergone an increase in volume of ( 1 +  )3 , through the application of
external forces. The expanded solid has an energy given by
2 X 1 X
E = Eeq + Ri D(R i − Rj ) Rj + ũ D(R i − Rj ) ũj
e e
2 i,j 2 i,j i
(1945)

The terms linear in  vanish identically, as the total force on an ion must vanish
in equilibrium. The total force is the sum of the internal forces opposing the
expansion and the applied external forces that result in the expansion. Since
the dynamical matrix that governs the lattice displacements ũi is unchanged,
its eigenvalues, which are the phonon frequencies, are unchanged by expansion
of a harmonic solid.

Thermal expansion only occurs for an anharmonic lattice. Thermal expan-


sion provides a measure of the volume dependence of the phonon frequencies
∂ h̄ ω
∂ V or the anharmonicity.

14.4.9 Thermal Expansion


The coefficient of thermal expansion of an insulator can be evaluated from
     
∂V ∂P ∂P
= − (1946)
∂T P ∂T V ∂V T
 
∂F
where the pressure is found from P = − ∂V . Using the expression for
T
the free energy of the lattice, one finds that the coefficient of thermal expansion
can be written as
!
∂ h̄ ωα (q) ∂N (ωα (q))

1 X
α = − (1947)
B q,α ∂V ∂T

where B is the bulk modulus and N (ω) is the Bose-Einstein distribution func-
tion.

572
The specific heat can be written as

∂N (ωα (q))
X  
CV = h̄ ωα (q) (1948)
q,α
∂T

On identifying the contributions from each normal mode, one can define a
Gruneisen parameter for each normal mode

∂ ωα (q)
 
V
γα (q) = − (1949)
ωα (q) ∂V
α B V
which is a dimensionless ratio of CV . Thus,
!
∂ ln ωα (q)
γα (q) = − (1950)
∂ ln V

The Gruneisen parameter for the entire solid can be expressed as a weighted
average of the Gruneisen parameter of each normal mode
P
q,α γα (q) Cq,α
γ = P (1951)
q,α Cq,α

with weights given by Cq,α . This is consistent with the definition of the Gruneisen
parameter in terms of thermodynamic quantities
αBV
γ = (1952)
CV
For most models, γα (q) is roughly independent of T and is a constant.
!
∂ ln ωD
γα (q) ∼ γ = − (1953)
∂ ln V

Hence, as B is roughly T independent, the specific heat CV tracks the coefficient


of thermal expansion α. A typical Gruneisen parameter has a magnitude of ∼
1 or 2, and a slow temperature variation, which changes on the scale of TD .

14.4.10 Thermal Expansion of Metals


For a metal, there is an additional contribution to the pressure from the elec-
trons. The electronic contribution to the pressure is calculated as

2 EVel
Pel = (1954)
3 V

573
and as the electronic energy is temperature dependent, the electronic contribu-
tion to the pressure is also temperature dependent. This gives an additional
contribution to the rate of change of pressure with respect to temperature
 
∂Pel 2 el
= C (1955)
∂T 3 V

Hence, the coefficient of thermal expansion for a metal is determined from


 
∂P
∂T
 
∂V
= −  V (1956)
∂T P ∂P
∂V
T

Hence, !
1 2 el
α = γ CVlatt + C (1957)
B 3 V
2
where 3 is the electronic Gruneisen parameter.

14.5 Anharmonicity
Anharmonic interactions give rise to a coupling of the normal modes. They can
be described a collision process in which the number of phonons may not be
conserved. For example, two phonons may interact and combine to produce a
single phonon. The interaction which describes this process is the cubic term
in the Taylor expansion of the potential in powers of the lattice displacements.
For many properties, it is also necessary to also consider the effect of the quartic
terms in the expansion, as these give rise to perturbative corrections which are
of the same order as the first non-zero corrections produced by the cubic terms.
Furthermore, if the lattice is stable against large amplitude lattice vibrations,
the quartic terms are non-zero. The anharmonic interactions give rise to the
lifetime of phonons, provide temperature dependent corrections to the phonon
dispersion relations, and contribute to the specific heat. The theory of anhar-
monic interactions has been reviewed by Cowley202 and the experimental man-
ifestations have been reviewed by Martin203 . The anharmonic interactions may
usually be thought of as producing small corrections to the harmonic phonons,
except when the systems is on the verge of a structural instability where they
play an important role. The phonon modes are not the only excitations of the
crystalline lattice, there are also large amplitude excitations like dislocations.
Although these excitations may have a large (macroscopic) spatial extent they
do not extend all through the crystal, like the phonon modes, and the deviations
of the atoms from the ideal equilibrium positions can be large, comparable to
202 R. A. Cowley, Adv. in Phys. 12, 421 (1963).
203 D. H. Martin, Adv. in Phys. 14, 39 (1965).

574
the lattice spacing. If the lattice displacements in the dislocations were consid-
ered to be made up of a superposition of coherent states for each phonon mode,
in the absence of the anharmonic interactions, the distortions would disperse
and the dislocations would lose their shape. The anharmonic interactions are
responsible for stabilizing these large amplitude, spatially localized, excitations
by balancing the effects of dispersion of the phonon modes. These excitation
do have macroscopically large excitation energies but they do also have macro-
scopically large effects. In essence, these dislocations are non-linear excitations,
like solitons, and play an extremely important role in determining the actual
mechanical properties of any real solid.

——————————————————————————————————

14.5.1 Exercise 86
The full ionic potential of a mono-atomic Bravais lattice has the form
1 X X
V̂ = Veq + uµ (R1 ) Dµ ν (R1 − R2 ) uν (R2 )
2! µ,ν
R1 ,R2
1 X X
+ uµ (R1 ) uν (R2 ) uλ (R3 ) Dµ ν λ
(R1 , R2 , R3 )
3!
R1 ,R2 ,R3 µ,ν,λ
1 X X
+ uµ (R1 ) uν (R2 ) uλ (R3 ) uρ (R4 ) Dµ ν λ ρ
(R1 , R2 , R3 , R4 )
4!
R1 ,R2 ,R3 ,R4 µ,ν,λ,ρ

(1958)
where u(R) is the displacement from the equilibrium position R.

(i) The sites of the expanded lattice are defined by


R = (1 + )R (1959)
Show that if an expansion is made about the sites of the expanded lattice, then
the dynamical matrix is changed to
Dµ,ν (R − R0 ) = Dµ,ν (R − R0 ) +  δDµ,ν (R − R0 ) (1960)
where the change in the dynamical matrix is given by
X
δDµ,ν (R − R0 ) = Dµ ν λ (R, R0 , R”) Rλ ” (1961)
λ,R”

(ii) Show that the Gruneisen parameter is given by

α (q) δ D(q)
e α (q)
γα (q) = 2
(1962)
6 M ωα (q)

575
——————————————————————————————————

576
15 Phonon Measurements
The spectrum of phonon excitations in a solid can be measured directly, via
inelastic neutron scattering or Raman scattering of light.

15.1 Inelastic Neutron Scattering


The neutrons interact with the atomic nuclei by a very short-ranged contact
interaction
X 2 π h̄2
Ĥint = b δ 3 ( r − r(Ri ) ) (1963)
i
m n

where r is the position of the neutron, and r(Ri ) are the positions of the ions.
The inelastic neutron scattering cross-section contains information about the
ground state and all the excited states of the lattice. The various contributions
to the spectrum are analyzed by use of the conservation laws.

In inelastic neutron scattering experiments, the incident neutron energy is


given by
P2 h̄2 k 2
E = = (1964)
2 mn 2 mn
and the final energy is given by

P 02 h̄2 k 02
E0 = = (1965)
2 mn 2 mn
The energy transfer from the neutron to the sample is given by

h̄ ω = E − E 0 (1966)

The law of conservation of energy demands that this energy is transferred to


the phonon excitations. The conservation law is written as
X
h̄ ω = h̄ ωα (q 0 ) ( n0q0 ,α − nq0 ,α ) (1967)
q 0 ,α

where nq,α is the number of phonons initially in the mode with wave vector and
polarization (q 0 , α) in the initial state and n0q0 ,α is the number in the final state.

Due to the periodic translational invariance of the crystal, the neutron trans-
fers momentum to the phonon modes and the crystal. The principle of conser-
vation of momentum yields
X  
P − P 0 = h̄ k − h̄ k 0 = h̄ q 0 n0q0 ,α − nq0 ,α + h̄ Q (1968)
q 0 ,α

577
where Q is a reciprocal lattice vector. Thus, even if the scattering is elastic, the
neutron may still be diffracted. The use of the two conservation laws allows the
dispersion relation ωα (q 0 ) to be determined.

15.2 The Scattering Cross-Section


2
d σ
The inelastic scattering cross-section dΩdω depends on the scattering geometry
through the scattering angle θ and dΩ. The solid angle dΩ is the angle subtended
by the detector to the target. The inelastic neutron scattering cross-section can
be calculated from the Fermi Golden rule for the neutron scattering rate. The
Fermi Golden rule expression for the rate at which the neutron is scattered from
state k to state k 0 is given by
2  
1 2 π Y
0
Y
0
X
0 0
= < k nq0 ,α | Ĥint | k nq0 ,α > δ h̄ω +
h̄ ωα (q )(nq0 ,α −nq0 ,α )
τk→k0 h̄ 0 0 0
q ,α q ,α q ,α
(1969)
where the phonon occupation numbers in the initial state are given by nq0 ,α and
are n0q0 ,α in the final state. The energy conserving delta function can be written
in terms of the integral
Z ∞
dt0
 X    X 
0 0 0 0 0
δ h̄ω + h̄ ωα (q )(nq0 ,α −nq0 ,α ) = exp − i t ω + ωα (q )(nq0 ,α −nq0 ,α )
0
q ,α −∞ ( 2 π h̄ ) 0 q ,α
(1970)
The exponential factor involving the energy difference between the initial and
final state can be associated with one factor of the matrix elements of the
interaction Hamiltonian. Thus, the transition rate can be interpreted as an
integral over time of a two-time correlation function
 Z ∞
1 1 Y Y
= 2 dt0 < nq0 ,α | < k | Ĥint | k 0 > | n0q0 ,α > ×
τk→k0 h̄ −∞ q 0 ,α q 0 ,α
0 0
    
Y
0 0 it it Y
< nq0 ,α | < k | exp + Ĥ0 Ĥint exp − Ĥ0 | k > | nq0 ,α >
0
h̄ h̄ 0
q ,α q ,α

(1971)
The interaction Hamiltonian Ĥint is expressed in the interaction representation
as a time-dependent operator, via
i Ĥ0 t0 i Ĥ0 t0
   
Ĥint (t0 ) = exp + Ĥint (0) exp − (1972)
h̄ h̄
Hence, the scattering rate is expressed as
 Z ∞
1 1 Y Y
= 2 dt0 < nq0 ,α | < k | Ĥint (0) | k 0 > | n0q0 ,α > ×
τk→k 0 h̄ −∞ 0 0
q ,α q ,α

578
Y Y 
× < n0q0 ,α | < k 0 | Ĥint (t0 ) | k > | nq0 ,α >
q 0 ,α q 0 ,α

(1973)
The matrix elements involve the initial and final states each of which are prod-
ucts of the neutron states k and the states of the lattice. The initial and final
states of the lattice are defined by the
Q number of quanta Q in each normal mode
0
and are written, respectively, as | q 0 ,α nq ,α > and |
0
q 0 ,α nq 0 ,α >. The
above expression involves the matrix elements of the interaction between the
neutron and the nuclei in the solid, but unlike the elastic scattering cross-section
derived previously, the nuclei may be displaced from their equilibrium positions
by ui according to
r(Ri ) = Ri + ui (1974)
Hence, the interaction Hamiltonian is given by
X 2 π h̄2 b
Ĥint (t0 ) = δ 3 ( r − Ri − ui (t0 ) ) (1975)
i
m n

The matrix elements of the interaction between the initial and final states of
the neutron, respectively, labelled by momentum k and k 0 , are given by
X 2 π h̄2 b  
0 0 0 0 0
< k | Ĥint (t ) | k > = exp i (k − k ) . (Ri + ui (t )) − i ω t
i
mn V
(1976)
Hence, the scattering rate is given by
 2 X   Z ∞  
1 2 π h̄ b
= exp i (k − k 0 ) . (Ri − Rj ) dt0 exp − i ω t0
τk→k0 mn V i,j −∞

Y   Y
× < nq0 ,α | exp − i (k − k 0 ) . uj (0) | n0q0 ,α >
q 0 ,α α,q 0
Y   Y 
× < n0q0 ,α | exp i (k − k 0 ) . ui (t0 ) | nq0 ,α >
q 0 ,α q 0 ,α

(1977)
Since the final states of the phonon modes are not measured, the final states
must be summed over. On summing over all the final states of the phonon
modes, one obtains the expression for the transition rate
 2 X   Z ∞  
1 2 π h̄ b 0 0 0
= exp i (k − k ) . (Ri − Rj ) dt exp − i ω t
τk→k0 mn V i,j −∞

Y     Y 
0 0 0
× < nq0 ,α | exp − i (k − k ) . uj (0) exp i (k − k ) . ui (t ) | nq0 ,α >
q 0 ,α q 0 ,α

(1978)

579
The initial occupations of the phonon modes (q 0 , α) are to be thermally averaged
over. Since in the harmonic approximation the exponential Boltzmann factor
factorizes, the thermal average for each phonon mode can be performed inde-
pendently. The thermal average over the mode (q 0 , α) corresponds to the sum
over the occupation number nq0 ,α with the Boltzmann probability pq,α where
 
1 0
pq,α (nq0 ,α ) = exp − β h̄ ω α (q ) n 0
q ,α (1979)
Zα (q 0 )
The normalization of the probability is given by the partition function for the
phonon mode (q 0 , α)

X  
Zα (q 0 ) = exp − β h̄ ωα (q 0 ) nq0 ,α
nq0 ,α =0

1
= (1980)
1 − exp[ − β h̄ ωα (q 0 ) ]

To simplify notation, the thermal average of an arbitrary operator  is denoted


by angular brackets
Y  X ∞  Y Y
< | Â | > = pq ,α1 < nq,α | Â | nq0 ,α0 > (1981)
1
q ,α1
1
nq ,α1 =0 q,α q 0 ,α0
1

The thermal averaged transition rate is given by


 2 X   Z ∞  
1 2 π h̄ b
= exp i (k − k 0 ) . (Ri − Rj ) dt0 exp − i ω t0
τk→k0 mn V i,j −∞
   
× < | exp − i (k − k 0 ) . uj (0) exp i (k − k 0 ) . ui (t0 ) | >

(1982)

This transition rate is to be related to the inelastic scattering cross-section.

The inelastic scattering cross-section is defined in terms of the flux F , the


scattering rate and the density of final states
V
ρdΩ (k 0 ) dk 0 dΩ = k 02 dk 0 dΩ (1983)
( 2 π )3
via
d2 σ 1
F dΩ dω = ρdΩ (k 0 ) dk 0 dΩ (1984)
dΩdω τk→k0
The flux, for a beam with a density of one neutron per volume V , is expressed
by
h̄ k
F = (1985)
V mn

580
The density of the neutron’s final states is given by
mn V
ρdΩ (k 0 ) dk 0 dΩ = k 0 dω dΩ (1986)
( 2 π )3 h̄
Hence, the neutron’s scattering cross-section is given by
2
d2 σ k 0 mn V

1
=
dΩdω k 2 π h̄ 2 π τk→k0
k0 2
= b S(k − k 0 , ω) (1987)
k
where S(q, ω) is the structure factor. The structure factor is given by the Fourier
transform of the time-dependent correlation function
Z ∞  
dt
S(q, ω) = exp − i ω t S(q, t) (1988)
−∞ 2 π

where the correlation function depends exponentially on the lattice displace-


ments
X      
S(q, t) = exp i q . Ri,j < | exp − i q . uj (0) exp + i q . ui (t) | >
i,j
(1989)
This is recognized as the spatial Fourier transform of the nuclear density-density
correlation function. The above expression relating the inelastic scattering cross-
section to the structure factor was first derived by van Hove204 .

It is to be assumed that the displacements ui are sufficiently smaller than


the lattice spacing so that the correlation function can be expanded in powers
of ui .
   
= exp − i q . uj (0) exp + i q . ui (t)

= 1 − i q . ( uj (0) − ui (t) )
 
1 2 2
− ( q . uj (0) ) + ( q . ui (t) ) − 2 ( q . uj (0) ) ( q . ui (t) ) + ...
2!
(1990)
The lattice displacements are then to be expressed in terms of the phonon
creation and annihilation operators, and the thermal average of the terms will
be evaluated. The displacement operators are expressed in terms of the phonon
creation and annihilation operators via
s  
0 1 X h̄ 0 0
ui (t ) = √  (q ) exp − i q . Ri ×
N 0 2 M ωα (q 0 ) α
q ,α

204 L. van Hove, Phys. Rev. 95, 249 (1954).

581
 
× a−q0 ,α exp[ − i ωα (q 0 ) t0 ] + a†q0 ,α exp[ + i ωα (q 0 ) t0 ]

(1991)

where α (q 0 ) are the phonon polarization vectors. The expectation values of


the various powers of u in the expansion can be characterized by the number
of phonons that are excited in the scattering process. Since the displacement
operator either creates or destroys a phonon, the only non-zero terms in the
expansion involve even powers of u.

15.2.1 The Zero-Phonon Scattering Process


The zero-th order term in the expansion of the structure factor corresponds to
the term of unity in eqn(1990). The zero-th order contribution to the structure
factor is given by
  Z ∞  
0
X dt
S (q, ω) = exp i q . Ri,j exp − i ω t (1992)
i,j −∞ 2 π

The integration over t can be performed leading to a zero-frequency delta func-


tion  
X
0
S (q, ω) = exp i q . Ri,j δ( ω ) (1993)
i,j

expressing the elastic scattering condition. This yields the dominant contribu-
tion to the elastic scattering cross-section

d2 σ X  
2
= b exp i q . Ri,j δ( ω ) (1994)
dΩdω i,j

since k = k 0 for elastic scattering. On integrating over an infinitesimal fre-


quency interval ω, around ω = 0, one finds a contribution to the elastic
scattering cross-section given by
Z 
dσ d2 σ
= dω
dΩ − dΩdω
X  
2
= b exp i q . Ri,j (1995)
i,j

If q corresponds to a reciprocal lattice vector Q, then the scattering is coherent


and the intensity of the Bragg peak is proportional to N 2 . As we shall see later,
the intensity of the elastic scattering peak is reduced by quantum and thermal
vibrations of the lattice. The intensity of the elastic peak is reduced by the
Debye-Waller factor.

582
15.3 The Debye-Waller Factor
There are other terms in the expansion in eqn(1990) which are also independent
of time and, hence, lead to contributions to the inelastic scattering spectrum.
For example, the expectation value of the two terms
 
1 2 2
− ( q . uj (0) ) + ( q . ui (t) ) (1996)
2!

in eqn(1990) are time independent. On utilizing the expression for the lattice
displacements in terms of the phonon creation and annihilation operators, one
finds that the thermal average involves expectation values such as

< | ( a†q ,α1 + a−q ,α1 ) ( a†q ,α2 + a−q ,α2 )| > (1997)
1 1 2 2

The terms involving the product of two phonon creation operators or two phonon
destruction operators are identically zero. The remaining terms involve the
product of a phonon creation and annihilation operator. These terms are non-
zero, only if q 1 = − q 2 and α1 = α2 , in which case they can be expressed
directly in terms of the phonon occupation numbers. The non-zero contribution
is given by  
δq +q δα1 ,α2 nq ,α1 + ( 1 + n−q ,α1 ) (1998)
1 2 1 1

Hence, the expectation value of the term proportional to ( q . uj (0) )2 is evalu-


ated as
 2  
1 X h̄
− q . α1 (q 1 ) 1 + 2 N (ωα1 (q 1 )) (1999)
2 N q ,α 2 M ωα1 (q 1 )
1
1

where the thermal average of the phonon occupation numbers have been re-
placed by the Bose-Einstein distribution function N (ω). Due to the time and
spatial homogeneity of the system, the expectation value of ( q . ui (t) )2 is
identical to that of ( q . uj (0) )2 . The two second-order time independent
contributions can be combined with the zero-th order term 1, to give the first
two terms in the expansion of the elastic scattering cross-section. The inelastic
scattering cross-section can be written as

d2 σ X  
= b2 exp i q . Ri,j δ( ω ) W (q) (2000)
dΩdω i,j

where W (q) is of the form


 2  
1 X h̄
W (q) = 1 − q . α1 (q 1 ) 2 N (ωα1 (q 1 )) + 1 + ...
N q ,α 2 M ωα1 (q 1 )
1
1

(2001)

583
The series expansion for W (q) can be exponentiated to yield the expression for
the Debye-Waller factor
2
β h̄ ωα1 (q 1 )
  
1 X h̄
W (q) = exp − q . α1 (q 1 ) coth
N q ,α 2 M ωα1 (q 1 ) 2
1
1
(2002)
The Debye-Waller factor represents the reduction of the intensity of the Bragg
peak due to the loss of coherent scattering caused by the displacement of the nu-
clei from their equilibrium positions205 . The Debye-Waller factor represents the
combined effect of thermal206 and quantum fluctuations207 of the nuclei. The
intensities of the Bragg peaks are usually reduced on increasing temperature.
However, if the solid undergoes a structural instability, the phonon dispersion
relation may soften as the temperature is lowered and the instability is ap-
proached208 . The softening of the phonon modes may lead to a reduction in
the Debye-Waller factor with decreasing temperatures. The form of the expo-
nent is similar to the factor appearing in the Lindemann criterion for melting,
however, instead of depending on the inverse square of the lattice constant, the
Debye-Waller factor is proportional to the square of the momentum transfer
q = k − k 0 . Due to the dependence on the momentum transfer, the Debye-
Waller factor is expected to preferentially reduce the intensity of the Bragg peaks
with large Q, while the intensity of the low Q Bragg peaks are not expected to
change appreciably. The Debye-Waller factor also modifies the intensity of the
Bragg peak in x-ray scattering.

The Debye-Waller factor multiplies the intensities of the multi-phonon pro-


cesses, of all orders. This can be ascertained by examining the correlation
function in the structure factor
   
< | exp − i q . uj (0) exp + i q . ui (t) | > (2003)

For harmonic phonons, one can express the correlation function as


   
< | exp − i q . uj (0) exp + i q . ui (t) | >
    
= exp + < | q . uj (0) q . ui (t) | > ×
  2    2 
1 1
× exp − < | q . uj (0) | > exp − < | q . ui (t) | >
2 2
205 In the case of anomalous scattering in Ge caused by the accumulation of charge associated

with bonding, the intensity of the forbidden Bragg reflections at 2πa


(2, 2, 2) is not governed
by the Debye-Waller factor. The intensity of this peak decreases more rapidly with increasing
temperature, than is predicted by the Debye-Waller factor. (J. B. Roberto, B. W. Batterman
and D. J. Keating, Phys. Rev. B 9, 2599 (1974).)
206 P. Debye, Ann. Phys. (Leipzig) 43, 49 (1914) see also I. Waller, Z. Phys. 17, 398 (1923).
207 H. Ott, Ann. Phys. (Leipzig), 23, 169 (1934).
208 G. Shirane, Rev. Mod. Phys. 46, 437 (1974).

584
(2004)

The proof of this identity proceeds with the aid of the Baker-Hausdorff theorem
and can be found in the review article of A. A. Maradudin, E. W. Montroll,
G. H. Weiss and I. P. Ipatova, Solid State Phys. Suppl. 3, 307 (1971). The
product of the last two factors is identified with the Debye-Waller factor, which
is given by
  2 
W (q) = exp − < | q . ui (0) | >
2
β h̄ ωα (q 0 )
  
1 X h̄ 0
= exp − q . α (q ) coth
N 0 2 M ωα (q 0 ) 2
q ,α

(2005)

The time-dependent term in eqn(2004) can be expanded in powers of the ex-


pectation values of the correlation of the displacements
 
= W (q) exp < | ( q . uj (0) ) ( q . ui (t) ) | >
!
= W (q) 1 + < | ( q . uj (0) ) ( q . ui (t) ) | > + . . .

(2006)

Apart from the first, all the terms in this expansion have a non-trivial expo-
nential time dependence and, thus, do not contribute to the elastic scattering.
Therefore, all the contributions to the scattering cross-section are reduced in
intensity by the Debye-Waller factor. The first term in the expansion is found
to be proportional to δ(ω) and gives rise to the elastic scattering. The second
term is just the one-phonon contribution to the scattering cross-section.

15.3.1 The One-Phonon Scattering Processes.


The one-phonon creation and annihilation contribution to the scattering cross-
section originates from the term

W (q) < | ( q . uj (0) ) ( q . ui (t) ) | > (2007)

in the expansion of eqn(2004). On expressing the displacements uj (0) and


ui (t) in terms of the phonon creation and annihilation operators, eqn(2007)
is evaluated as
 2
X h̄ 0
= W (q) q . α (q ) ×
0
2 N M ωα (q 0 )
q ,α

585
  
× [ nq0 ,α + 1 ] exp − i ( q 0 . Ri,j − ωα (q 0 ) t )
 
+ nq0 ,α exp + i ( q 0 . Ri,j − ωα (q 0 ) t )

(2008)

Hence, the one-phonon contribution to the inelastic scattering cross-section is


found to be
2
d2 σ

k 2 X h̄ 0
= b W (q) q . α (q ) ×
dΩdω k0 2 M N ωα (q 0 )
q 0 ,α
 X  
0 0
× [ nq ,α + 1 ] δ( ω − ωα (q ) )
0 exp i ( q − q ) . Ri,j
i,j
X  
0 0
+ n−q0 ,α δ( ω + ωα (q ) ) exp i ( q − q ) . Ri,j
i,j
(2009)

The displacements have been expressed in terms of the normal modes, and
nq0 ,α is just the number of phonons with wave vector q 0 and polarization α in
the initial state. On performing the sum over lattice vectors Ri , one finds the
condition for conservation of crystal momentum,
X  
exp i ( q − q 0 ) . Ri = N ∆q−q0 (2010)
i

modulo Q. Thus, the summation over q can be trivially performed. Then,


the second and third terms involve the absorption or emission of a phonon of
wave vector q 0 = (q + Q) where Q is a reciprocal lattice vector. These terms
are smaller than the coherent Bragg terms by a factor of N1 . The inelastic
one-phonon contributions are coherent as they involve the conservation of mo-
mentum, but have intensities that are only proportional to N .

The two contributions to the one-phonon scattering cross-section correspond


to different signs of the energy transfer ω. The contribution which is propor-
tional to the Bose-Einstein distribution function corresponds to a negative en-
ergy transfer.
X  
N (ωα (q 0 )) δ( ω + ωα (q 0 ) ) exp i ( q + q 0 ) . Ri,j (2011)
i,j

It represents processes in which a phonon, that is thermally excited in the initial


state, scatters with the neutron and is subsequently destroyed. The factor
N (ωα (q 0 )) represents the probability that a phonon is thermally excited in the

586
initial state. The other contribution corresponds to a positive energy transfer
and has the form
X  
[ N (ωα (q 0 )) + 1 ] δ( ω − ωα (q 0 ) ) exp i ( q − q 0 ) . Ri,j (2012)
i,j

It represents processes whereby the neutron interacts with the solid and emits
a phonon. The term proportional to unity corresponds to the spontaneous
emission process, whereas the term proportional to N (ωα (q 0 )) represents the
stimulated emission process. The presence of a phonon in the mode (q 0 , α) en-
hances the probability that subsequent phonons will be created in that mode.

At low temperatures, the number of thermally-activated phonons is small,


therefore, the inelastic scattering intensity for processes which lead to an increase
in the energy of the neutron due to absorption of phonons is small. On the other
hand, the intensity of processes which involve the energy loss by the neutron
beam due to creation of individual phonons has an intensity governed by the
1 + N (ωα (q 0 )) which is almost unity at low temperatures. The rate for inelastic
transitions of the incident neutrons obeys the principle of detailed balance. That
is, although the neutron beam is not in equilibrium with the solid, the transition
rate is such that it drives the beam towards equilibrium. This can be seen by
inspection of the one-phonon contribution to the spectrum. The one-phonon
absorption and emission spectrum is proportional to

[ 1 + N (ωα (q 0 )) ] δ( E − E 0 − h̄ ωα (q 0 ) ) + N (ωα (q 0 )) δ( E − E 0 + h̄ ωα (q 0 ) )
(2013)
The first term represents processes in which the neutron loses energy due to the
emission of a phonon, whereas the second term represents processes in which
the neutron gains energy due to the absorption of a phonon. The ratio of the
rate at which the neutron beam gains energy to the rate at which the neutron
beam loses energy is given by
 
W (E → E + h̄ω) N (ω)
= = exp − β h̄ ω (2014)
W (E + h̄ω → E) [ N (ω) + 1 ]
If equilibrium with the beam were to be established, the kinetic energy of the
neutron beam would be distributed according to the Boltzmann formula
 
1
P (E) = exp − β E (2015)
Z
such that dynamic equilibrium would be established. In this case, the total
number of transitions from E → E + h̄ ω precisely equals the number of
transitions in the reverse direction E + h̄ ω → E

P (E) W (E → E + h̄ω) = P (E + h̄ω) W (E + h̄ω → E) (2016)

However, the beam produced by the neutron source is not in equilibrium with
the sample, and would only equilibrate if the beam traverses an infinite path

587
Figure 242: The phonon dispersion relation inferred from inelastic neutron scat-
tering experiments on f.c.c. Cu. [After Svensson et al., Phys. Rev. 155, 619
(1967).]

length through the sample.

The phonon dispersion relation can be inferred from a measurement of the


single-phonon scattering peak209 . The scattering cross-section for processes in
which a single phonon is emitted have to satisfy the energy and momentum
conservation laws
h̄2 k 2 h̄2 k 02
= + h̄ ωα (q 0 ) (2017)
2 mn 2 mn
and
k = k0 + q0 + Q (2018)
0
since ω(q ) is periodic with a periodicity of the reciprocal lattice vectors

ωα (q 0 ) = ωα (q 0 + Q) (2019)

One can combine the equations as

h̄2 k 2 h̄2 k 02
= + h̄ ωα (k − k 0 ) (2020)
2 mn 2 mn
In the scattering experiments, the beam of neutrons is generally collimated to
have a definite direction of the k vector, and also to have a definite initial energy.
For a given k, the solution of the above equation for the three components of k 0
form a two-dimensional surface. For a detector placed in a particular scattering
direction, the solution only exists at isolated points. On measuring the scat-
tering cross-section at the various magnitudes of the final momentum, k 0 , one
finds sharp peaks in the spectrum. In general, these peaks occur at different
wave vectors than the Bragg peaks. With knowledge of the magnitude of the
209 R. Weinstock, Phys. Rev. 65, 1 (1944), A. D. B. Woods, B. N. Brockhouse, R. A. Cowley

and W. Cochran, Phys. Rev. 131, 1025 (1963).

588
Figure 243: The single phonon peak in the frequency distribution of the scatter-
ing cross-section of KBr, at constant momentum transfer. The intensity of the
single phonon peak is temperature dependent, as is the intensity of the smooth
multi-phonon background. [After Woods et al., Phys. Rev. 131, 2025 (1963).]

final momentum k 0 , one can construct k 0 − k, and also E 0 − E and, hence, find
h̄ ωα (q 0 ) for the normal mode. By varying the direction of k 0 and the magnitude
of E, one can map out successive surfaces and, therefore, obtain the dispersion
relation.

k=k'+q+Q

h2k'2/2mn

k'=k
h2k2/2mn - hω(k-k')

0 k'

Figure 244: A graphical solution of the conditions of conservation of energy and


momentum for a one-dimensional lattice, in which a neutron interacts with the
solid and emits a phonon.

Information about the polarization of the phonon modes can be obtained


from the dependence of the intensity on the scattering wave-vector k − k 0 as the
scattering cross-section is proportional to
2
( k − k 0 ) . α (q 0 )

(2021)

589
The width of the single-phonon peak obtained in experiments have two ori-
gins, one is the experimental resolution and the other component is not reso-
lution limited. The second component is due to the lifetime of the phonon, τ .
The phonon lifetime, according to the energy-time uncertainty principle, gives
rise to an energy width of h̄τ . The lifetime occurs because the phonons are
scattered either by anharmonic processes or by electrons. The small magnitude
of the width of the phonon peaks attests to the effectiveness of the harmonic
approximation and the Born-Oppenheimer approximation.

15.3.2 Multi-Phonon Scattering


Processes in which two phonons are absorbed or emitted satisfy the two conser-
vation laws
h̄2 k 2 h̄2 k 02
= ± h̄ ωα1 (q 1 ) ± h̄ ωα2 (q 2 ) (2022)
2 mn 2 mn
and
k = k0 ± q1 ± q2 + Q (2023)
Conservation of momentum can be used to express q 2 in terms of q 1 . Conser-
vation of momentum gives rise to the restriction

h̄2 k 2 h̄2 k 02
= ± h̄ ωα1 (q 1 ) ± h̄ ωα2 (k − k 0 ± q 1 ) (2024)
2 mn 2 mn

Since there are six quantities k 0 and q 1 and only one remaining constraint, the
values of k’ and q 1 are still undetermined. Even if the direction of k 0 is fixed in
an experiment, there still remains three unknown quantities q 1 . The variation

Phonon absorption
k+q'=k'+Q
h2k'2/2mn = h2k2/2mn + hω(k'-k) h2k2/2mn + hω(k'-k)

h2k'2/2mn

0 k'=k k'

Figure 245: A graphical solution of the conditions of conservation of energy


and momentum for a one-dimensional lattice, in which a phonon present in the
initial state is absorbed by a neutron.

590
k+q'=k'+Q
Q
Q q'
q'
εT
k-k' εT k-k'

O O

Figure 246: The sensitivity of the inelastic neutron scattering intensity to the
polarization vector T for a transverse phonon of wave vector q 0 . In the scatter-
ing process a transverse phonon present in the initial state is absorbed by the
neutron.

(3/4,3/4,0) (1,1,0) (1,1/2,0) (1,0,0)


0.06

0.05

0.04

0.03

0.02

0.01

0
Γ K X W X Γ

Figure 247: Calculated phonon dispersion relations for an f.c.c. solid, showing
the periodicity of the reciprocal lattice along the XW X high-symmetry direc-
tion.

of q 1 produces a continuously varying final neutron energy. Hence, one obtains


a continuous spectrum. A similar analysis of the higher order multi-phonon
processes also yields a continuous spectrum. Only the one-phonon spectrum
gives rise to a single peak.

Thus, in a general coherent scattering experiment with a specific scattering


direction, the analysis of the scattered neutrons energy provides a spectrum
which contains a continuous portion superimposed with sharp peaks. The spec-
trum may show an elastic Bragg peak depending on the magnitude of k and
θ, or if there is isotopic disorder, one may observe incoherent nuclear scatter-
ing at zero-energy transfer. The peaks of the one-phonon scattering can be
used to map out the dispersion relations. This has been performed for f.c.c.
lead. However, some branches were not observed. The intensity of the one-
phonon absorption peak is proportional to the Bose-Einstein distribution func-
tion N (ωα1 (q 1 )), whereas the one-phonon emission process has intensity pro-

591
portional to [ N (ωα1 (q 1 )) + 1 ]. Thus, it is usual to measure phonon emission
at low temperatures.

The phonon density of states can be obtained directly, if the incoherent


scattering is sufficiently strong. In this case, the isotopic disorder produces
scattering from individual ions. The resulting incoherent scattering is, there-
fore, proportional to the nuclear density-density correlation function evaluated
at the same lattice site. Hence, the incoherent one-phonon scattering process
averages over all the phonon wave vectors. A measurement of the incoherent
inelastic scattering spectrum provides a direct measure of the phonon density
of states210 .

——————————————————————————————————

15.3.3 Exercise 87
(i) Find a graphical description of the conservation laws for the phonon emission
process.

(ii) Show that there is a minimum or threshold energy required for phonon emis-
sion.

——————————————————————————————————

210 A. T. Stewart and B. N. Brockhouse, Rev. Mod. Phys. 30, 250 (1958), B. N. Brockhouse,

Can. J. Phys. 33, 889 (1953).

ez

1/2(1,1,1)
ey
(1,-1,0)
(1,0,0)
ex

Figure 248: The XW X high-symmetry line, in the extended zone scheme. Since
the point X = (1, 1, 0) is equivalent to X = (1, 0, 0), the dispersion relations for
an f.c.c. crystal must be symmetric along the XW X line.

592
15.3.4 Exercise 88
(i) Evaluate the Debye-Waller factor for a one, two or three dimensional system
of acoustic phonons.

(ii) Determine the temperature dependence of the integrated intensity of the


scattering cross-section, defined by
Z +∞
k 0 d2 σ
I(q) = dω (2025)
−∞ k dωdΩ
——————————————————————————————————

15.3.5 Exercise 89
Consider inelastic neutron scattering from a perfect fluid, described by the
Hamiltonian
X P̂ 2
i
Ĥ0 = (2026)
i
2 M
Show that the inelastic scattering cross-section is proportional to
 12 2 
d2 σ h̄2 q 2
  
βM βM
∝ exp − h̄ ω − (2027)
dωdΩ 2 π h̄2 q 2 2 h̄2 q 2 2M

——————————————————————————————————

15.4 Raman and Brillouin Scattering of Light


Since the energy of visible light is of the order of eV and the energy of a typical
phonon is of the order of meV, ( 10−3 eV), it is not possible to observe phonons
by direct absorption or emission of light. However, it is possible to observe the
phonons in a solid via light scattering. Even though the scattering processes
proceed via the same mechanism, the scattering from optical phonons is called
Raman scattering211 and scattering from acoustic phonons is called Brillouin
scattering212 .

As in neutron scattering, the basic process may involve emission of phonons


or absorption of phonons. The conservation laws for the one-phonon absorption
or emission processes are conservation of energy

h̄ ω 0 = h̄ ω ± h̄ ωα (q) (2028)
211 C. V. Raman, Nature, 121, 619 (1928).
212 L. Brillouin, Ann. de Phys., (Paris), 17, 88 (1922).

593
Figure 249: The Stokes and anti-Stokes lines in Si as observed in Raman scat-
tering experiments. The measurements were performed at temperatures of 20,
460 and 770 K. Note the temperature dependence of the ratio of the intensities
of the Stokes and anti-Stokes lines. [After T. R. Hart, R. L. Aggarawal and B.
Lax, Phys. Rev. B 1, 638 (1970).]

and conservation of momentum

h̄ k 0 n = h̄ k n ± h̄ q + h̄ Q (2029)

In these expressions (k, ω) and (k 0 , ω 0 ) are, respectively, the momentum and


energy of the incident beam of photons and the scattered photons, and n is the
refractive index of the media. The refraction index reflects the change in the
wavelength of the light as it enters the solid. The phonon absorption process (+)
gives rise to the anti-Stoke’s shifted line, which has an intensity proportional to
the number of activated phonons

∝ N (ωα (q)) (2030)

The phonon emission process (−) gives rise to the Stoke’s line which has an
intensity proportional to

∝ [ 1 + N (ωα (q)) ] (2031)

as it has contributions from spontaneous and stimulated emission of phonons.

The Raman scattering process allows parts of the phonon dispersion relations
to be mapped out. Since the characteristic phonon frequency is given by the
Debye frequency h̄ ωD ∼ 10−2 eV, which is small compared with a typical
photon energy h̄ c k n ∼ 1 eV, the change in photon wave vector k − k 0 is

594
q
k

θ
k'

Figure 250: The relation between the scattering angle θ, the initial and final
wave vector of the light and the phonon wave vector q. The magnitudes of the
wavevectors k and k 0 of the incident and scattered light are almost equal.

small. Therefore, the triangle formed by the initial and final wave vectors is
almost isosceles. The momentum transfer q is given by

θ
| q | = 2 n k sin
2
ω θ
= 2n sin (2032)
c 2
Since the direction of k and k 0 are known from the experimental geometry, the
direction of q can be inferred. Thus, the phonon momentum momentum and
the phonon energy are known, if the small change in the photon energy, h̄ ∆ω,
is measured.

For Brillouin scattering, the phonon energy is given by

ωα (q) = vα q (2033)

where vα is the velocity of sound. The magnitude of the phonon’s momentum


is given by
ωα (q) ωn θ
q = = 2 sin (2034)
vα (q) c 2
However, the energy of the acoustic phonon is equal to the change in photon
energy, ∆ ω,
ωα (q) = ∆ω (2035)
Thus, the velocity of the acoustic phonon is found as
∆ω c θ
vα (q) = csc (2036)
2ω n 2
The experimentally determined spectra has the form of a strong un-scattered
laser line, surrounded by a small anti-Stoke’s line at higher frequencies, and a
slightly more intense Stoke’s line at lower frequencies. The Stokes and anti-
Stoke’s line are both separated from the main line by the same frequency shift
∆ω. This technique can be used to determine the phonon frequencies. The
widths of the phonon peaks provide a measure of the imaginary part of the

595
dielectric constant.

The quantum mechanical theory of the Raman effect was first formulated
by Loudon213 . Loudon emphasized the role of electrons in mediating the Ra-
man scattering. In particular, Loudon noted that although there are Raman
scattering processes which do not involve electrons, their intensities are negli-
gible unless they are nearly resonant in which case the photon frequency must
be comparable to the phonon frequency. Since typical phonon frequencies are
of the order of meV, this process will be limited to the far infrared region.
For typical ranges of photon frequencies, Raman scattering is dominated by
electron-assisted processes.

The Raman scattering cross-section is calculated with third-order time-


dependent perturbation theory, with two powers of the paramagnetic electron-
photon interaction (i.e., only the A . p̂ terms) and one power of the electron-
phonon interaction. In the Raman process, a photon is absorbed by the solid
creating a virtual electron-hole pair. The pair either emits or absorbs a phonon.
The pair recombines by emitting a photon. The transition amplitude consists of
six terms corresponding to the various possible time-orderings of the three indi-
vidual processes. The transition amplitude is conventionally written in terms of
the Raman tensor. The scattering cross-section is proportional to the squared
modulus of the Raman tensor. The Raman tensor is both frequency and wave
vector-dependent, it is symmetric if the phonon frequency is negligible compared
with the frequency of the incident light. The Raman tensor involves the sums
of products of matrix elements. One factor comes from the matrix elements of
the electron’s momentum along the direction of the initial photon’s polarization
and a second factor involves matrix elements of the electrons momentum along
the polarization of the final photon. The third and last factor represents the
matrix elements of the interaction Hamiltonian between electronic states when
a phonon with a specific polarization is present.

If one neglects the difference between the wavelengths of the incident and
scattered scattered light, then there are selection rules for the phonon modes
that can be Raman scattered. Loudon has examined the form of the Raman
tensor for the different point groups. In crystals which do have a center of in-
version symmetry, only even-parity phonons are Raman active. The odd-parity
phonons are, however, infra-red active and can be observed in optical absorp-
tion measurements. From the symmetry, it is seen that the lowest order Raman
transitions are forbidden if each lattice site is an inversion center. Hence, N aCl
does not exhibit first order Raman scattering214 but diamond does215 .

Loudon’s analysis assumes the existence of a virtual electron-hole pair in the


213 R. Loudon, Proc. Roy. Soc. A 275, 218 (1963), R. Loudon, Adv. in Phys. 13, 423
(1964).
214 M. Born and M. Bradburn, Proc. Roy. Soc. A 188, 161 (1947).
215 H. M. J. Smith, Phil. Trans. Roy. Soc. A 241, 105 (1948).

596
(ω',k') (ω',k')
(ω(q),q) (ω(q),q)
(ω,k)
(ω,k)
(ω(q),q) (ω',k') (ω',k') (ω(q),q)

(ω,k) (ω,k)
space

(ω',k') (ω(q),q) (ω(q),q)


(ω',k')
(ω,k)
(ω,k)

time

Figure 251: A diagramatic depiction of the virtual processes involved in one-


phonon Raman scattering. The incident photon (ω, k) and the scattered photon
(ω 0 , k 0 ) are depicted by red wavy lines. We have depicted a process in which a
phonon (ω(q), q) is emitted by an electron (or a hole). Since the process is a
virtual process, it is only necessary for energy to be conserved in the initial and
final states, and not in the intermediate states.

597
intermediate states, and needs modification when the scattering is in resonance
with the intermediate states216 . In such cases, the resonance can enhance the
Raman scattering by factors which can be as large as 102 , as is found in CdS 217 .

216 J. L. Birman and A. K. Ganguly, Phys. Rev. Lett. 17, 647 (1966), D. L. Mills and E.

Burstein, Phys. Rev. 188, 1465 (1969).


217 R. C. Leite and S. P. S. Porto, Phys. Rev. Lett. 17, 10 (1966).

598
16 Phonons in Metals
An alternate approach to the phonon dispersion in metals is based on a two-
component plasma composed of electrons and ions. The approach starts by
consideration of a plasma composed of the positively charged ions with charge
Z | e | and mass M . The plasma of ions support longitudinal charge density
oscillations which occur in the absence of an external potential. Since the total
potential is related to the external scalar potential via
φext (q, ω)
φ(q, ω) = (2037)
ε(q, ω)

then, if φext (q, ω) = 0, one must have ε(q, ω) = 0 for φ(q, ω) 6= 0. In


this case, one has spontaneous density fluctuations and an induced longitudinal
current. On using Poisson’s equation and the condition of continuity of the
charge density, one finds the induced longitudinal current in the form
" #

jL (q, ω) = φ(q, ω) − φext (q, ω)

 

= 1 − ε(q, ω) φ(q, ω) (2038)

Using the definition of the longitudinal conductivity, one recovers the expression
for the dielectric constant
4 π σ(ω)
ε(q, ω) = 1 − (2039)

The Drude expression for the conductivity of a gas of ions of charge Z | e | and
mass M is given by

Z 2 e2 ρions τ 1
σ(ω) = (2040)
M 1 − iωτ
On substituting the ionic Drude conductivity in the expression for the dielectric
constant of the ions, which on assuming the limit ω  τ1 reduces to

4 π Z e2 ρ
ε(q, ω) = 1 − (2041)
M ω2
where the density of ions is given in terms of the electron density ρ via
ρ
ρions = (2042)
Z

The condition for plasmon oscillations is given by

ε(q, ω) = 0 (2043)

599
the solution for ω is defined as the ionic plasmon frequency Ωp . The ionic
plasmon frequency may be written in terms of the plasmon frequency
Z m 2
Ω2p = ωp (2044)
M
which is independent of q. The ionic plasmon frequency corresponds to an un-
screened phonon frequency. Since the factor ZMm ∼ 4000
1
and h̄ ωp ∼ 10 eV,
1
the unscreened phonon frequency is approximately ∼ 10 eV.

16.1 Screened Ionic Plasmons


The above model is inadequate as it neglects the effects of the conduction elec-
trons. This effect of the electrons can be included by screening the Coulomb
interactions between the charged nuclei

4 π Z 2 e2
(2045)
q2
with the dielectric constant of the electron gas. In the Thomas-Fermi approxi-
mation, the dielectric constant is given by

kT2 F
εeg (q, ω) = 1 + (2046)
q2
Thus, within the Born-Oppenheimer approximation, one obtains the frequency-
dependent dielectric constant as

4 π Z e2 ρ
ε(q, ω) = 1 − 2
kT
(2047)
M (1 + q2
F
) ω2

The screened ionic plasmons have frequencies which are given by

4 π Z e2 ρ
ε(q, ω) = 1 − 2
kT
= 0 (2048)
M (1 + q2
F
) ω2

Thus,
Z m 2 q2
ω2 = ωp 2 (2049)
M q + kT2 F
This is the Bohm-Staver model of the longitudinal phonons in a metal218 . This
model results in a linear dispersion relation ω(q) ≈ v q, where the velocity v
is defined as
Z m ωp2
v2 = (2050)
M kT2 F
218 D. Bohm and T. Staver, Phys. Rev. 84, 836 (1950).

600
As the Thomas-Fermi wave vector is given in terms of the Fermi wave vector by

4 π e2 h̄2 π
2 ≈ (2051)
kT F m kF

and the electron density is expressed as

kF3
ρ = (2052)
3 π2

the velocity of sound is related to the Fermi velocity vF = m kF via

1 Z m 2
v2 = v (2053)
3 M F
m
Thus, the velocity of sound v is reduced below the Fermi velocity vF as M ∼
10−3 − 10−5 .

16.1.1 Kohn Anomalies


A more accurate treatment of the phonon frequency replaces the Thomas-Fermi
dielectric function with the Lindhard expression

4 π e2 d3 k f (Ek+q ) − f (Ek )
Z
εeg (q, ω) = 1 − (2054)
q2 4 π 3 Ek+q − Ek + h̄ ω

where f (x) is the Fermi-Dirac distribution function. For free electrons, the
dielectric function has singularities in the derivative at q = 2 kF . These
singularities correspond to the extremal diameters of the Fermi surface. Walter
Kohn showed that these singularities should appear in the phonon spectrum219
by producing kinks or infinities in the derivative
 
∂ω
(2055)
∂q q=2kF

The Kohn anomalies have been observed in some metals by inelastic neutron
scattering measurements220 . The Fermi surface in Lead has been mapped out
by this indirect method221 . The results are in fair agreement with the Fermi
surface inferred from de Haas - van Alphen measurements222 .

219 W. Kohn, Phys. Rev. Lett. 2, 393 (1959), E. J. Woll Jr. and W. Kohn, Phys. Rev. 126,

1693 (1962).
220 B. N. Brockhouse, K. R. Rao and A. D. B. Woods, Phys. Rev. Lett. 7, 93 (1961).
221 R. Stedman, L. Almquist, G. Nilsson and G. Raunio, Phys. Rev. 163, 567 (1967).
222 J. R. Anderson and A. V. Gold, Phys. Rev. 139, 1459 (1965).

601
16.2 Dielectric Constant of a Metal
The dielectric constant of a metal represents the process in which an external
charge is screened by the combined effects of the electrons and the ions

φext (q, ω) = φ(q, ω) ε(q, ω) (2056)

A dielectric function can be defined for just the electrons in which the total
potential φ(q, ω) is produced as the response to a total external potential which
is external to the electron gas. That is, the total external potential is considered
to be the sum of the applied external potential and the total potential due to
the ion charge density

φext (q, ω) + φions (q, ω) = φ(q, ω) εel (q, ω) (2057)

Analogously, a dielectric function can be defined for the ions as the response
of the ions to an external potential composed of the applied potential and the
electrons
φext (q, ω) + φel (q, ω) = φ(q, ω) εions (q, ω) (2058)
This goes beyond the Born-Oppenheimer approximation. The total potential is
given by the sum of the potentials due to the external, electron and ion charges

φ(q, ω) = φext (q, ω) + φions (q, ω) + φel (q, ω) (2059)

The dielectric constant of the metal is given in terms of the dielectric constant of
the electrons and the dielectric constant of the ions, by adding the two equations
defining the electronic and ionic dielectric constants
 
εions (q, ω) + εel (q, ω) φ(q, ω) = φ(q, ω) + φext (q, ω) (2060)

Then, with the definition of the total dielectric constant, one has the relation
 
ε(q, ω) = εions (q, ω) + εel (q, ω) − 1 (2061)

The dielectric constant of the ions goes beyond the Born-Oppenheimer approxi-
mation. It describes how the ions, alone, screen the potential due to the applied
potential and the potential due to the electrons. The dielectric constant due to
the ions alone is approximated by

Ω2p
εions (q, ω) = 1 − (2062)
ω2
and the electronic dielectric constant (at low frequencies) is given by the Thomas-
Fermi approximation
k2
εel (q, ω) = 1 + T2F (2063)
q

602
Hence, the low-frequency dielectric constant is given by the approximate ex-
pression
k2 Ω2p
ε(q, ω) = 1 + T2F − 2 (2064)
q ω
for ωp  ω.

An alternate definition of the dielectric constant of the ions may be intro-


duced in which one considers the external potential to be first screened by the
electron gas. Secondly, the resulting dressed-external potential is screened by
the ions. That is, instead of the electron gas screening the external potential of
the ions and the applied potential
φext (q, ω) φions (q, ω)
φ(q, ω) = + (2065)
εel (q, ω) εel (q, ω)

one considers only the dressed-external potential


φext (q, ω)
φdressed (q, ω) = (2066)
εel (q, ω)

It is this dressed-external potential that is screened by the ions to produce the


total potential. This relation defines the dressed dielectric constant of the ions
φdressed (q, ω)
φ( q, ω) =
εdressed
ions (q, ω)
φext (q, ω)
=
εel (q, ω) εdressed
ions (q, ω)
(2067)

Hence, the electronic and dressed-ionic dielectric constants are related to the
dielectric constant via

ε(q, ω) = εel (q, ω) εdressed


ions (q, ω) (2068)

Combining this with the relation of the dielectric constant in terms of dielectric
constants of the electrons and ions
 
ε(q, ω) = εions (q, ω) + εel (q, ω) − 1 (2069)

one finds that the dressed-ionic dielectric constant is defined by


 
dressed 1
εions (q, ω) = εions (q, ω) + εel (q, ω) − 1
εel (q, ω)
 
1
= 1 + εions (q, ω) − 1
εel (q, ω)
(2070)

603
The dressed-ionic dielectric constant is calculated as
 
dressed 1
εions (q, ω) = 1 + k2
εions (q, ω) − 1
1 + qT2F
 2 
1 Ωp
= 1 − 2
1 + q2
kT F ω2
(2071)
This can be written in terms of the phonon dispersion relation ω(q)2

ω(q)2
εdressed
ions (q, ω) = 1 − (2072)
ω2
since the phonon oscillations occur when the dielectric constant vanishes
εdressed
ions (q, ω(q)) = 0 (2073)
By inspection of the dressed dielectric constant, the phonon frequency is found
as
q2
ω(q)2 = 2 Ω2 (2074)
q + kT2 F p
The introduction of screening by the electron gas has reduced the frequency of
the ionic density oscillations from the ionic plasmon frequency to a branch of
longitudinal acoustic phonons. The total dielectric constant, which is a product
of the dressed dielectric constant and the Thomas-Fermi dielectric constant of
the electron gas, can now be written in terms of the phonon frequencies as
1 1 1
= 2
kT ω(q)2
ε(q, ω) 1 + F
1 −
q2 ω2
2
1 ω
= 2
1 +
kT F ω 2 − ω(q)2
q2
(2075)
This is in agreement with the expression discussed earlier.

16.3 The Retarded Electron-Electron Interaction


Consider the screening of the Coulomb interaction between a pair of electrons
via the dielectric constant
4π 4π

q2 ε(q, ω) q 2
ω(q)2
 

= 1 +
kT2 F + q 2 ω − ω(q)2
2

(2076)

604
Thus, there is an additional contribution in the effective interaction due to the
screening by the ions. The ω dependence of the interaction represents the fact
that the effective interaction is not instantaneous but instead is a retarded in-
teraction223 . The retarded nature of the attractive interaction between two
electrons is caused by the involvement of the polarization of the lattice. One
electron produces a polarization of the ions in its vicinity, which evolves on
a time scale that is determined by the energy transfer ∼ h̄ ωD . Due to the
large difference between the Fermi velocity an the speed of sound, the original
electron will have moved considerable distances before the polarization is fully
developed. After the dynamical lattice distortion has been created, a second
electron is attracted by the remanent of the lattice deformation left behind by
the original electron. The effective interaction between a pair of electrons in-
volves a momentum transfer q = k − k 0 and energy transfer h̄ ω = Ek − Ek0 .
The effective interaction has the following limits:

(i) This interaction reduces to the Thomas-Fermi screened electron-electron


interaction when the electron energy transfer is greater than the typical phonon
frequency ωD ∼ Ωp . In this case, when ω > ωD , the phonon correction is
unimportant.

(ii) The electron-electron interaction is strongly modified at low frequencies,


where ω < ωD . The contribution from the phonons is large and of opposite
sign to the direct Coulomb repulsion, and exactly cancels at ω = 0. The
important point, however, is that the retarded interaction is attractive at low
frequencies. It exhibits the phenomenon of over-screening and can give rise to
superconductivity.

16.4 Phonon Renormalization of Quasi-Particles


The electron-phonon interaction can give rise to a change in the quasi-particle
dispersion relation. The Hartree-Fock contribution to the quasi-particle energy
from the un-screened electron-electron interaction is
X e2
∆E(k) = f (Ek0 ) < k k 0 | | k k0 >
0
| r − r0 |
k

e2
Z Z   
1 X 3 3 0 0 0
= f (E k 0 ) d r d r 1 − exp i ( k − k ) . ( r − r )
V2 0 | r − r0 |
k

1 X 4 π e2
= ∆EH − f (Ek0 )
V 0
| k − k 0 |2
k

(2077)
223 H. Frohlich, Phys. Rev. 79, 845 (1950), also see J. Bardeen and D. Pines, Phys. Rev.
99, 1140 (1955).

605
The first term is the Hartree term which is k independent and can be absorbed
into a shift of the chemical potential. The second term is the exchange term
which depends on k. The exchange term affects the quasi-particle dispersion re-
lation. If the effect of phonon screening is included, the exchange term becomes
" #
1 X 4 π e2 h̄2 ω(k − k 0 )2
− f (Ek0 ) 1+
V 0
| k − k 0 |2 + kT2 F ( Ek − Ek0 )2 − h̄2 ω(k − k 0 )2
k
(2078)
In this expression the electronic screening of the exchange interaction is treated
in the Thomas-Fermi approximation, and the screening due to the phonons has
also been included.

On utilizing the smallness of the Debye frequency with respect to the Fermi
energy, and integrating over the magnitude of k 0 , one can show that the change
in energy due to the electron-phonon interaction is given by
d2 S 0 4 π e2 h̄ ω(k − k 0 ) µ − Ek − h̄ω(k − k 0 )
Z
1
− ln
8 π 3 h̄ v(k) | k − k 0 |2 + kT2 F 2 µ − E + h̄ω(k − k 0 )
k
(2079)
where k 0 lies on the Fermi surface. Substitution of Ek = µ immediately demon-
strates that the value of the Fermi energy µ and the shape of the Fermi surface
are unaltered by the coupling to the phonons which, in the approximation un-
der consideration, is given by the Thomas-Fermi quasi-particle theory. Secondly,
when the quasi-particle energy is within h̄ ωD of µ, | Eqp (k) | < h̄ ωD , the
logarithmic term can be expanded in inverse powers of h̄ ω. Then, after invoking
self-consistency, it is seen that the phonon contribution to the screening changes
the dispersion relation from that of the Thomas-Fermi screened theory to
EkT F − µ
Eqp (k) = (2080)
1 + λ
where 1 + λ is the wave function renormalization due to the phonons. That
is the wave function contains a coherent quasi-particle component that has a
relative weight of
1
(2081)
1 + λ
the other component is incoherent as the electron is in a linear combination
of other momentum states since it has been scattered by a phonon. The wave
function renormalization is given by the expression
d2 S 0 4 π e2
Z
1
λ = (2082)
8π 3 h̄ v(k 0 ) | k − k 0 |2 + kT2 F
This has the result that the quasi-particle velocity is given by
1
v(k) = ∇ Eqp (k)

1 1
= ∇ EkT F (2083)
1 + λ h̄

606
Thus, the quasi-particle contribution to the density of states is enhanced by a
factor of 1 + λ
ρ(µ) = ( 1 + λ ) ρT F (µ) (2084)
An upper bound to the coupling constant is provided by the inequality
4 π e2 d2 S 0
Z
1
λ < (2085)
2
kT F 8 π 3 h̄ v(k 0 )
however, the Thomas-Fermi screening length is defined by
 −1
4 π e2 ∂ρ 1
= =
kT2 F ∂µ ρ(µ)
−1
d2 S 0
 Z
= (2086)
4 π 3 h̄ v(k 0 )
Hence, the phonon renormalization factor is usually less than unity
λ < 1 (2087)
Typical values of λ are in the range of 0.2 to 0.8. Estimates of λ for various
metals are given below224

Metal λ Metal λ Metal λ

Li 0.41 Be 0.23
Na 0.16 Mg 0.36 Al 0.43-0.38
Zn 0.38 Ga 0.40
In 0.69-0.8 Sn 0.6
Hg 1.0 Tl 0.71 Pb 1.1-1.5
Ti 0.38 V 0.60
Zr 0.41 Nb 0.82 Mo 0.41

Finally, the phonon corrections are negligible for electron energies far from the
Fermi energy. For example, when
| Eqp (k) | > h̄ ωD (2088)
then the dispersion relation suffers only small corrections
 2
h̄ ωD
Eqp (k) = EkT F − µ + O (2089)
EkT F − µ
224 W. L. McMillan, Phys. Rev. 167, 331 (1968).

607
Thus, there has to be a kink in the quasi-particle dispersion relation at energies
close to the Fermi energy.

16.5 Electron-Phonon Interactions


The effect of coupling with the phonons on the quasi-particle spectrum can be
used to deduce the form of the electron-phonon interaction. The change in the
ground state energy of a metal due to the electron phonon interaction, Ĥint ,
can be estimated from second order perturbation theory as
X | < Ψ0 | Ĥint | Ψm > |2
∆2 E = (2090)
i
E0 − Em
It is assumed that the form of the electron - phonon interaction is dominated by
the first non-trivial term in the expansion of potential acting on the electrons
in powers of the ionic displacements
X
Ĥint = ûi . ∇Ri Vions (r) (2091)
i

Thus, the most important excitation process comes from excited states | Ψm >
in which an electron has been scattered from state k to k − q. Also a phonon of
wave vector q has been excited, hence,
Em − E0 = Ek−q + h̄ ω(q) − Ek (2092)
Thus, one can express the second order correction to the ground state energy
in a phenomenological manner as
X | λq |2
∆2 E = − f (Ek ) ( 1 − f (Ek−q ) ) (2093)
Ek−q + h̄ ω(q) − Ek
k,q

where f (x) is the Fermi function. One can identify an effective electron-electron
interaction, due to the phonons, from the functional derivative of the energy with
respect to the Fermi functions
δ 2 ∆2 E
Vef f (q) = (2094)
δf (Ek ) δf (Ek−q )
Hence,
| λq |2
Vef f (q) = −
Ek − Ek−q − h̄ ω(q)
| λq |2

Ek−q − Ek − h̄ ω(q)
" #
2 h̄ ω(q)
= | λq |2
h̄2 ω(q)2 − ( Ek − Ek−q )2
(2095)

608
On identifying the above effective potential with the phonon contribution to
the screened interaction between the electrons, one obtains an expression for
the effective coupling constant | λq |2 as

1 4 π e2 h̄ ω(q)
| λq |2 = 2 2 (2096)
V q + kT F 2

For small q, the coupling constant vanishes linearly with q, since

4 π e2 2 µ
= (2097)
kT2 F 3 ρ

for q < kT F , the coupling constant varies as

µ h̄ ω(q)
| λq |2 =
ρV 3
h̄ ω(q) µ
=
3N Z
(2098)

16.6 Electrical Resistivity due to Phonon Scattering


The electron-phonon scattering contributes to the electrical resistivity. The
phonon gas acts as a source or sink for the electron momentum, thus, the interac-
tions with the electron gas reduces the current flow. Hence, the electron-phonon
interaction increases the resistivity. The electron-ion interaction is given
by X
Ĥions = V (r − R) (2099)
R

and as the position of the i-th ion can be written in terms of the equilibrium
position and a displacement

R = R i + ui (2100)

The potential of the ions is expanded up to linear order in the lattice displace-
ments ui
" #
X
Ĥions = V (r − Ri ) − ui . ∇R V (r − Ri ) + . . . (2101)
i

The first term represents the static lattice and the second term is the electron
phonon interaction. The electron phonon interaction is given by
X
Ĥint = − ui . ∇R V (r − Ri ) (2102)
i

609
Thus, the interaction produces scattering of the electrons between Bloch states
and, through ui involves the absorption or emission of phonons. The condition
of conservation of energy yields the selection rule
E(k) = E(k 0 ) ± h̄ ω(k − k 0 ) (2103)
This single restriction leads to a two-dimensional surface of Bloch state wave
vectors k 0 that are allowed final states for the electron initially in Bloch state
k. The momentum transfer for these processes is given by q = k − k 0 . The
surface of allowed final states must be close to the surface of initial energy as
h̄ ω  µ, hence, E(q) ∼ E(k − q). The scattering rate out of the state with
momentum k is given by
1
=
τ (k → k 0 )
 
2π X
| λα
q |2
f (E(k)) 1 − f (E(k + q))
h̄ α
"  
× N (ωα (q)) δ E(k) − E(k + q) + h̄ ωα (q)
   #
+ 1 + N (ωα (q)) δ E(k) − E(k + q) − h̄ ωα (q)

(2104)
The rate for scattering into the momentum state k is given by
1
0 =
τ (k → k)
 
2π X α 2
| λq | f (E(k + q)) 1 − f (E(k))
h̄ α
"  
× N (ωα (q)) δ E(k + q) − E(k) + h̄ ωα (q)
   #
+ 1 + N (ωα (q)) δ E(k + q) − E(k) − h̄ ωα (q)

(2105)
The transport scattering rate is the rate for momentum change of an electron
at the Fermi surface is defined by
  X  
1 1 0 1
( k . E ) f (E(k)) 1 − f (E(k)) = (k.E) − (k .E)
τ 0
τ (k → k 0 ) τ (k 0 → k)
k

(2106)

610
The rate for scattering out of state k will be transformed into a form comparable
to the rate for scattering in. The rate for scattering out of momentum state k
is re-written as
   
2π X α 2
= | λq | f (E(k)) 1 − f (E(k + q)) exp β ( E(k) − E(k + q) )
h̄ α
"   
× 1 + N (ωα (q)) δ E(k) − E(k + q) + h̄ ωα (q)
 #
+ N (ωα (q)) δ E(k) − E(k + q) − h̄ ωα (q)

 
2π X α 2
= | λq | f (E(k + q)) 1 − f (E(k))
h̄ α
"   
× 1 + N (ωα (q)) δ E(k) − E(k + q) + h̄ ωα (q)
 #
+ N (ωα (q)) δ E(k) − E(k + q) − h̄ ωα (q)

(2107)
Thus, the transport scattering rate can be expressed as
 
1
( k . E ) f (E(k)) 1 − f (E(k)) =
τ
(2108)
 
2π X
= ( q . E ) | λα
q |2
f (E(k + q)) 1 − f (E(k))
h̄ α, q
"   
× 1 + N (ωα (q)) δ E(k) − E(k + q) + h̄ ωα (q)
 #
+ N (ωα (q)) δ E(k) − E(k + q) − h̄ ωα (q)

(2109)
Furthermore, as
    
f (E(k + q)) 1 − f (E(k)) 1 + N (ωα (q)) δ E(k) − E(k + q) + h̄ ωα (q)
   
= f (E(k)) 1 − f (E(k + q)) N (ωα (q)) δ E(k) − E(k + q) + h̄ ωα (q)

(2110)

611
the scattering rate can be expressed as
 
1 2π X
( k . E ) f (E(k)) 1 − f (E(k)) = ( q . E ) | λα 2
q | N (ωα (q))
τ h̄ α, q
"    
× f (E(k)) 1 − f (E(k + q)) δ E(k) − E(k + q) + h̄ ωα (q)
   #
+ f (E(k + q)) 1 − f (E(k)) δ E(k) − E(k + q) − h̄ ωα (q)

2π X
= ( q . E ) | λα 2
q | N (ωα (q)) N ( − ωα (q))
h̄ α, q
"   
× f (E(k + q)) − f (E(k)) δ E(k) − E(k + q) + h̄ ωα (q)
   #
+ f (E(k)) − f (E(k + q)) δ E(k) − E(k + q) − h̄ ωα (q)

2π X
= ( q . E ) | λα 2
q | N (ωα (q)) N ( − ωα (q))
h̄ α, q
"   
× f (E(k) + h̄ωα (q)) − f (E(k)) δ E(k) − E(k + q) + h̄ ωα (q)
   #
+ f (E(k)) − f (E(k) − h̄ωα (q)) δ E(k) − E(k + q) − h̄ ωα (q)

(2111)

The summation over q is evaluated by transforming it into an integral

2π X
= ( q . E ) | λα 2
q | N (ωα (q)) N ( − ωα (q))
h̄ α, q
"   2
h̄2 2


× f (E(k) + h̄ωα (q)) − f (E(k)) δ (k.q) + q − h̄ ωα (q)
m 2m
  2 #
h̄2 2


+ f (E(k)) − f (E(k) − h̄ωα (q)) δ (k.q) + q + h̄ ωα (q)
m 2m
(2112)

The integration over the direction of q is performed in spherical polar coordi-


nates, in which the direction of k is fixed as the polar axis. The integral over
the azimuthal angles result in the factors of sin φ and cos φ in

( q . E ) = q cos θ Ez + q sin θ ( sin φ Ey + cos φ Ex ) (2113)

612
vanishing. The sole surviving term, proportional to Ez , can then be written in
a manner independent of the choice of axis as k . E which can be factored out
of the integral
2π 2mπ X Z
= ( 2 2 )(k.E) dq q 2 | λα 2
q | N (ωα (q)) N ( − ωα (q))
h̄ h̄ k α
"  Z 1
m ωα (q)
 
q
× f (E(k) + h̄ωα (q)) − f (E(k)) d cos θ cos θ δ cos θ + −
−1 2k h̄ k q
 Z 1 #
m ωα (q)
 
q
+ f (E(k)) − f (E(k) − h̄ωα (q)) d cos θ cos θ δ cos θ + +
−1 2k h̄ k q
(2114)
On neglecting the term of order vvFα , cancelling the factors of ( k . E ), and Taylor
expanding the Fermi function factors in powers of the phonon frequencies, one
finds the transport scattering rate for electrons on the Fermi surface is given by
 
1
f (E(k)) 1 − f (E(k)) =
τ
Z  
2π 2mπ X ∂f (E(k))
= − ( 2 3 ) dq q 3 | λα q |2
N (ω α (q)) N ( − ω α (q)) h̄ ω α (q)
h̄ h̄ k α
∂E(k)
(2115)
On using    
∂f (E(k))
= − β f (E(k)) 1 − f (E(k)) (2116)
∂E(k)
one finds
4 m π2 X
Z
1
= β( 3 3 ) dq q 3 | λα 2
q | N (ωα (q)) N ( − ωα (q)) h̄ ωα (q)
τ h̄ k α
(2117)
The temperature dependence of the transport scattering rate can be evaluated
using the Debye model for the phonons, and using a linear q dependence of
|λα 2
q | . The integral over q is evaluated through the substitution z = β h̄ ωα (q)
and ωα (q) = vα q to yield
TD
z5
Z
1 T
∝ T5 dz (2118)
τ 0 ( exp[z] − 1 ) ( 1 − exp[−z] )
For this temperature range, the number of thermally exited phonons is propor-
tional to T 3 . One would expect that the scattering rate would be proportional
Z
1
∝ dq q 2 N (ω(q))
τ
∝ T3 (2119)

613
However, as forward scattering is ineffective in transport properties, the trans-
port scattering rate is proportional to the change in momentum along the di-
rection of the electric field and therefore, is proportional to
θ
( 1 − cos θ ) = 2 sin2
2
1 q2
≈ (2120)
2 kF2

which produces an additional T 2 dependence. For low temperatures ( T < TD


) , the upper limit on the integration may be set to infinity yielding
 5
−1 T
σ(T ) ∝ (2121)
TD

Thus, the combined effect of the factor ( 1 − cos θ ) and τ1 ∝ T 2 produces a


T 5 temperature dependence in the low-temperature resistivity.

At high temperatures ( T > TD ), the range of integration is less than


unity so the integrand may be expanded in powers of z yielding
Z TD
T
−1 5
σ(T ) ∝ T dz z 3 = T TD
4
0
 
T
∝ (2122)
TD

which is the result for the classical limit of the scattering. This can be considered
to arise merely as the number of thermally activated phonons is given by the
classical expression
kB T
N (ωα (q)) = (2123)
h̄ ωα (q)
The above results were first derived independently by Bloch and Gruneisen and
the resulting formula is known as the Bloch - Gruneisen resistivity due to phonon
scattering.

16.6.1 Umklapp Scattering


Umklapp processes may change the leading low-temperature variation of the
resistivity. Umklapp scattering circumvents the factor of ( 1 − cos θ ) which
produces the extra T 2 factor. When kF is close to the zone boundary, a small q
value may couple the sheets of the Fermi surface in neighboring Brillouin zones.
These are the umklapp processes. They produce a large change in the electron
velocity ∆v, by a phonon induced Bragg reflection.

614
16.6.2 Phonon Drag
The resistivity could decrease faster than T 5 if the system was relatively free
of defects and umklapp scattering could be neglected. This would occur if the
phonons were allowed to equilibrate with the electronic system in its steady
state. The combined system of electrons and phonons should have a total mo-
mentum, which is conserved in collisions. As a result, the phonon system would
not be able to momentum (or current) from the electron system as they drift
together.

615
17 Phonons in Semiconductors
17.1 Resistivity due to Phonon Scattering
The transport scattering rate in a semiconductor can be obtained from the
collision integral of the Boltzmann equation
  "
X
2
I f (k) = λq f (k) ( 1 − f (k + q) ) N (ω(q)) δ( E(k) − E(k + q) + h̄ ω(q) )
q
#
− ( 1 − f (k) ) f (k + q) ( 1 + N (ω(q)) ) δ( E(k + q) − E(k) − h̄ ω(q) )

(2124)
in which f () is the non-equilibrium distribution function. On linearizing about
the equilibrium Fermi distribution function
∂f0 (k)
f (k) = f0 (k) + A ( k . E ) (2125)
∂E
yields the linearized collision integral.

Using the identity


   
( 1 − f (E(k)) ) f ((E(k)+h̄ω(q)) = 1 − exp β h̄ ω(q) f (E(k)) − f (E(k)+h̄ω(q))
(2126)
one finds the result
    Z 2k
2mAV dq 2
I f (k) = 2 exp − β ( E(k) − µ ) q Ez λ2q N (ω(q)) N (−ω(q))
h̄ kB T k 0 2π
"
m ω(q)
  
q
× + 1 − exp − β h̄ω(q)
2k h̄ k q
#
m ω(q)
  
q
− − 1 − exp + β h̄ω(q)
2k h̄ k q
(2127)
For low frequency acoustic phonons, the Bose-Einstein distribution can be ap-
proximated by its high temperature form leading to the collision integral
    Z 2k  
∂f (k) m V dq 2 2 q
I f (k) = A ( k . E ) − q λq −
∂E k3 0 2π 2 β h̄ ω(q)
(2128)
The transport scattering rate is found by factoring out the non-equilibrium part
of the distribution function
Z 2k
1 mV dq 3 N
= kB T q q 2 | V (q) |2
τ (E) k3 0 4 π 2 M ω(q)2

616
N V m
= k | V (0) |2 kB T (2129)
4 π c2 M
Hence, the conductivity in a semiconductor, in which the scattering rate is
dominated by phonon scattering, is given by
 Z  
2 3
σx,x ∼ β exp β µ dk k exp − β E(k) (2130)

Thus, the conductivity has a temperature dependence given by


 
σx,x ∼ exp β µ (2131)

Thus, as expected, the conductivity is still dominated by the number of carriers,


3
but the conductivity has an additional T dependence of T − 2 above and beyond
the prefactor in the number of carriers.

17.2 Polarons
Electron-phonon coupling in semiconductors can be large. For low density of
carriers, each carrier can cause a distortion of the lattice. The carrier and the
surrounding distortion forms an excitation which is known as a polaron. At low
temperatures the polaron appears to have a large effective mass, as the motion
of the carrier is hindered by the need to drag the surrounding lattice distortion.
Thus, there is a low-temperature regime in which the conductivity is governed
by the motion of the heavy quasi-particles with an extremely large and tem-
perature dependent effective mass. At high temperatures, the conductivity is
dominated by incoherent hopping processes, which are thermally assisted by the
presence of a thermal population of phonons.

17.3 Indirect Transitions


In a semiconductor, light can be absorbed in processes where by an electron
is excited from the filled valence band into the empty conduction band. The
minimum energy of the photon must be greater than the band gap between the
conduction and valence band density of states. Since the speed of light c is
so large, the wave vector of the photon absorbed in a transition between two
states with energy difference of the scale of eV is extremely long. Thus, the
momentum of the photon is negligible on the scale of the size of the Brillouin
zone. This means that in a semiconductor, if only a photon and an electron are
involved, momentum conservation only allows transitions in which the initial
and final state of the electron have the same k value. This type of transition is
called a direct transition. In some semiconductors the minimum of the conduc-
tion band dispersion relation lies vertically above the maximum of the valence
band, and the band gap is called the direct gap. In this case, the threshold for

617
direct absorption should coincide with the gap observed in the density of states.
On the other hand, if the energetic separation between the maximum of the va-
lence band and the minimum of the conduction band dispersion relations occur
at different k values, then the threshold energy for the absorption of a photon
in a direct transition should be greater than the separation inferred from just
considering the density of states alone. This second type of semiconductor has
two gaps, the indirect band gap inferred from the density of states and a direct
gap inferred for q = 0 transitions by consideration of the dispersion relations.

If the ions of the lattice are displaced from their equilibrium positions, simple
conservation of momentum arguments do not apply. In this case, it is possible
to have absorption at the indirect gap. At the threshold for indirect transitions,
the absorption process involves the absorption or emission of a phonon with
wave vector equal to the wave vector Q separating the valence band maximum
to the conduction band minimum. The transition rate has to be calculated via
second order perturbation theory, one power of the interaction involves the ab-
sorption of one photon and the other power of the interaction involves either
the absorption or emission of one phonon.

The state of the joint system composed of an electron, phonons of wave


vector Q and photons of frequency ω is denoted by | Ψ >. This state satisfies
the equation of motion
 

i h̄ |Ψ > = Ĥ0 + Ĥint | Ψ > (2132)
∂t
The state is decomposed in terms of eigenstates of Ĥ0 , | φn > with energy En
 
X En
|Ψ > = Cn (t) exp − i t | φn > (2133)
n

Then, one finds that the expansion coefficients Cn (t) satisfy the equation
 
∂ X Em − E n
i h̄ Cn (t) = < φn | Ĥint | φm > exp i t Cm (t) (2134)
∂t m

Since the system is initially in the ground state, then the state is subject to the
initial condition given by
Cn (0) = δn,0 (2135)
To first order, one has
Z t  
i E0 − En
Cn1 (t) = − dt0 < φn | Ĥint | φm > exp i t0 (2136)
h̄ 0 h̄
We assume the perturbation has no diagonal elements, therefore, C01 (t) = 0.
To second order, one has
 
∂ 2 X Em − E n 1
i h̄ Cn (t) = < φn | Ĥint | φm > exp i t Cm (t) (2137)
∂t h̄
m6=0

618
18 Impurities and Disorder
If an isolated impurity is introduced into a solid, and the impurity has no low-
energy degrees of freedom which can be excited, then it can be treated as an
impurity potential Vimp (r). The total Hamiltonian Ĥ is written as
Ĥ = Ĥ0 + V̂imp (2138)
where Ĥ0 describes the conduction band states of the pure metal. The eigen-
states of Ĥ0 are the Bloch states φk with energy eigenvalues Ek .
Ĥ0 φk = Ek φk (2139)
Since the impurity breaks the periodic translational invariance of the solid, the
impurity potential will scatter an electron between Bloch states with different
Bloch wave vectors. The non-zero matrix elements of the potential can be
written as
Z
d3 r φ∗k0 (r) Vimp (r) φk (r) = < k 0 | Vimp | k > (2140)

If the wave function, in the presence of an impurity, is written as a superposition


of Bloch states X
ψα (r) = Cα (k) φk (r) (2141)
k

then the energy eigenvalue Eα can be expressed as


X
( Eα − Ek ) Cα (k) = Cα (k 0 ) < k | Vimp | k 0 > (2142)
k0

If the quantity k0 Cα (k 0 ) < k | Vimp | k 0 > is well defined and a non-


P
zero function of k, then there exist eigenvalues Eα between every consecutive
pair of values of the Bloch energies Ek . For a large system, where Ek are very
closely spaced, the eigenvalues form a continuum. These eigenstates correspond
to weakly perturbed Bloch states. On the other hand, if the potential is attrac-
tive, and there is a minimum value of Ek (say Ek = 0), below which there can
be bound states with energies Eα .

The dependence of the bound state energy on the density of states of the
ordered material can be easily found, for the case where the potential has the
property that the matrix elements are independent of k and k 0 . This corresponds
to a short-ranged potential. To simplify notation, we shall introduce the matrix
elements
< k | V̂imp | k 0 > = Vimp (2143)
which are of the order of the inverse of the sample’s volume, V −1 . In this case,
one can easily solve for the bound states. The above equations can be solved
by introducing the quantity γ, defined by
X
γ = Cα (k 0 ) (2144)
k0

619
4

Im z
3

0
-4 -3 -2 -1 0 1 2 3 4 5 6

-1 Re z

-2

-3

-4

Figure 252: The analytic structure of the Green’s function. The Green’s Func-
tion has isolated poles on the negative real axis, corresponding to the bound
state energies Eα and a branch cut across the positive real axis..

The expansion coefficients in the eigenvalue equation can be expressed in terms


of γ through
Vimp
Cα (k) = γ (2145)
Eα − Ek
The above two equations leads to a self-consistency condition for the bound
state energy Eα
X Vimp
1 = (2146)
Eα − Ek
k

which shows that, for an attractive potential, there may be a critical value of
Vimp needed for a bound state to form.

A more powerful way of solving the same problem involves use of the one-
particle resolvent Green’s function. The resolvent Green’s function is defined
by the operator
Ĝ(z) = ( z − Ĥ )−1 (2147)
where z is a complex number. Since Ĥ is a Hermitean operator, the matrix
elements of the Green’s function can be expressed in terms of a sum of simple
poles at the energy eigenvalues. Since the eigenvalues of the Hamiltonian are
composed of discrete bound states at negative energies and a semi-continuous
spectrum at positive energies, the Green’s function has discrete poles along the
negative z axis and a branch cut across the positive real axis, x = <e z > 0.
The imaginary part of the Green’s function is discontinuous along the real axis.
The discontinuity is given by
  X
< Ψ | Ĝ(x−iη) − Ĝ(x+iη) | Φ > = 2 π i < Ψ | En > < En | Φ > δ( x − En )
n
(2148)
where the | En > are the energy eigenstates of Ĥ corresponding to the energy
eigenvalues En .

620
The density of states ρ() is given by the trace of the Green’s function.
 X 
1
ρ() = − =m < Ψn | Ĝ( + iη) | Ψn > (2149)
π n

where | Ψn > form a complete orthonormal basis. The identity can be proved
by using a specific basis for evaluating the trace. The trace can be evaluated
in the basis of energy eigenstates, since they form a complete orthonormal set.
The right hand side of the identity can be evaluated as
 X 
1
ρ() = − =m < En | Ĝ( + iη) | En > (2150)
π n

In the limit η → 0, the matrix elements of the Green’s function are given by
 X 
1 1
ρ() = − =m < En | | En >
π n  + i η − Ĥ
 X 
1 1
= − =m < En | | En >
π n
 + i η − En
1 X η
= < En | | En >
π n (  − En )2 + η 2
X
= δ(  − En ) (2151)
n

which is the definition of the density of states. Hence, the density of states is
given by the trace of the imaginary part of the Green’s function.

The resolvent Green’s function can be obtained by expressing the Hamilto-


nian in terms of the unperturbed Hamiltonian Ĥ0 and the interaction due to
the impurity potential, Ĥint ,

Ĥ = Ĥ0 + Ĥint (2152)

Then, the Green’s function


 −1  −1
Ĝ(z) = z − Ĥ = z − Ĥ0 − Ĥint (2153)

is identically equal to
 −1  −1
Ĝ(z) = z − Ĥ0 + z − Ĥ0 Ĥint Ĝ(z) (2154)

This identity can be expressed in terms of the non-interacting resolvent Green’s


function, Ĝ0 (z), as

Ĝ(z) = Ĝ0 (z) + Ĝ0 (z) Ĥint Ĝ(z) (2155)

621
Similarly, one can also prove the analogous identity

Ĝ(z) = Ĝ0 (z) + Ĝ(z) Ĥint Ĝ0 (z) (2156)

The matrix elements of the non-interacting resolvent Green’s function are easily
evaluated in terms of the matrix elements between the eigenstates of Ĥ0 , | En0 >
.
1
< En0 | Ĝ0 (z) | Em
0
> = < En0 | | Em0
>
z − Ĥ0
1
= < En0 | 0
| Em0
>
z − Em
< En0 | Em 0
>
= 0
z − Em
δn,m
= 0
(2157)
z − Em
which is diagonal. Since Bloch states of the host material are eigenstates of the
unperturbed Hamiltonian Ĥ0 , the non-interacting resolvent Green’s function
only has diagonal matrix elements between Bloch states. The matrix elements
are evaluated as
1 < k0 | k >
< k0 | |k > =
z − Ĥ0 z − Ek
δk,k0
= (2158)
z − Ek

The interacting Green’s function can be expressed in terms of the T̂ (z) matrix
as
Ĝ(z) = Ĝ0 (z) + Ĝ0 (z) T̂ (z) Ĝ0 (z) (2159)
where the T-matrix is defined as
 −1
T̂ (z) = Ĥint 1 − Ĝ0 (z) Ĥint (2160)

Thus, the poles of the T-matrix are related to the poles of the Green’s function.
For a sufficiently short-ranged potential, the matrix elements of Ĥint are inde-
pendent of k. In this case, the matrix elements of the T-matrix between any
pair of Bloch states can be evaluated as
 −1
0
X Vimp
< k | T̂ (z) | k > = Vimp 1 − (2161)
z − Ek”
k”

Since the density of Bloch states ρ0 () is defined as


 
1 X
ρ0 () = − δ  − Ek (2162)
π
k

622
0.50

πρ0(ε)
0.25

0.00
-12 -9 -6 -3 0 3 6 ε/t 9 12

En -0.25
Vimp-1

-0.50

Figure 253: A graphical solution of the equation for the bound state energy En .
The Hilbert transform of the density of states is shown in blue. The imaginary
part of the Hilbert transform is shown in red.

one can express the function in the denominator of the T-matrix as an integral
X Vimp Z ∞
Vimp ρ0 ()
= d (2163)
z − Ek 0 z − 
k

This has a discontinuous imaginary part on the positive real axis, therefore, the
T-matrix is non-analytic for z on the positive real axis corresponding to the
continuous spectra of energy eigenvalues. The T-matrix also has isolated poles
at the negative energies z = En which are given by the solutions of
Z ∞
1 ρ0 ()
= d (2164)
Vimp 0 En − 

These energies are the energies of bound states225 . The bound states are expo-
nentially localized around the impurity site. The minimum value of the attrac-
tive potential Vimp that produces a bound state strongly depends on the form
of the density of states at the edge of the continuum. The critical value of Vimp
denoted as Vc is given by the condition that the bound state energy is zero, i.e.
E0 = 0 Z ∞
1 ρ0 ()
= − d (2165)
Vc 0 
d−2
Since ρ0 () ∝ ( 2 ) near the band edges, the integral converges for three
dimensions and higher, but diverges for one and two dimensions. Hence, an
attractive short-ranged interaction will produce a bound state in one and two-
dimensions, irrespective of the strength of Vimp . However, in three dimensions,
the attractive potential has to exceed a critical value before a bound state can
be formed.

225 G. F. Koster and J. C. Slater, Phys. Rev. 96, 1208 (1954), P. A. Wolff, Phys. Rev. 124,

1030, (1961).

623
0.50

πρ0(ε)
0.25

0.00
-12 -9 -6 -3 0 3 6 ε/t 9 12

-0.25 Vimp-1

E0

-0.50

Figure 254: A graphical solution of the equation for the energy of a resonance
E0 . The Hilbert transform of the density of states is shown in blue. The
imaginary part of the Hilbert transform is shown in red.

If the potential is not strong enough to produce a bound state, then the real
part of the denominator of the T-matrix, i.e.,
Z ∞
ρ0 ()
1 − Vimp <e d (2166)
0 z − 
may vanish for some values of z on the positive real axis at which the unper-
turbed density of states is finite. The solutions usually occur in pairs. If the
denominator vanishes at z = E0 > 0, then the T-matrix has a resonance
at this energy where it varies rapidly with z. The T-matrix does not diverge
since the denominator has an imaginary part proportional to the unperturbed
density of states
Vimp π ρ0 (E0 ) (2167)
which is finite if E0 > 0. The change in the density of states due to the
impurity potential can be found by taking the trace of the equation
Ĝ( + iη) = Ĝ0 ( + iη) + Ĝ0 ( + iη) T̂ ( + iη) Ĝ0 ( + iη) (2168)
with the Bloch states of the host material. The first term gives rise to the
density of states of the host material. Hence, the impurity density of states is
given by the imaginary part of the trace of the remaining term which involves
the T-matrix. The impurity density of states can be written in terms of an
energy derivative
1 X ∂ 1
ρimp () = =m < k | T̂ ( + iη) | k > (2169)
π ∂  − Ek + i η
k

On using the explicit form of the T-matrix, the impurity density of states is
found to have the form of a logarithmic derivative
" Z ∞ #
ρ0 (0 )

1 ∂ 0
ρimp () = − =m ln 1 − Vimp d (2170)
π ∂ −∞  + i η − 0

624
The resonances in the T-matrix may cause the density of states to build up
close to the resonance near the band edge and be depleted at the higher energy
(anti-)resonance.

18.1 Scattering by Impurities


The exact eigenstates of a Hamiltonian containing a scattering potential V̂imp
satisfies the equation

Ĥ | Ψ+ > = ( Ĥ0 + V̂imp ) | Ψ+ > = E | Ψ+ > (2171)

This can be re-expressed as an integral equation with an initial state given by


the incident state is an eigenstate of Ĥ0 corresponding to the plane wave | k >
as
( E − Ĥ0 + i η ) | Ψ+ > = V̂imp | Ψ+ > (2172)
This equation has general solutions which are the superposition of the solutions
of the homogeneous equation and a particular solution of the inhomogeneous
equation
1
| Ψ+ > = | k > + V̂imp | Ψ+ > (2173)
E − Ĥ0 + i η
where E = Ek . Or, more formally,

| Ψ+ > = | k > + Ĝ0 (E + iη) V̂imp | Ψ+ > (2174)

where the unperturbed Green’s function Ĝ0 (z) is defined as

Ĝ0 (z) = ( z − Ĥ0 )−1 (2175)

For future reference, we shall also give the analogous expression

| Ψ+ > = | k > + Ĝ(E + iη) V̂imp | k > (2176)

involving the exact Green’s function, Ĝ(z) defined by

Ĝ(z) = ( z − Ĥ )−1 (2177)

and the asymptotic incident state | k > . On inserting a complete set of


eigenstates of Ĥ0 in eqn(2174), one finds
X
| Ψ+ > = | k > + | k 0 > < k 0 | Ĝ0 (E + iη) V̂imp | Ψ+ >
k0
X 1
= |k > + | k0 > < k 0 | V̂imp | Ψ+ >
E − Ek 0 + i η
k0

(2178)

625
k'

k+iδ
-∞ -k-iδ +∞

Figure 255: The contour used in the evaluation of the asymptotic behavior of
the scattered wave.

To ensure that | Ψ+ > − | k > is an outgoing wave, η must be chosen as a


positive infinitesimal constant. The asymptotic behavior of the scattered wave
function can be expressed as
 
1
Ψ+ (r) = √ exp i k . r
V
 
0 0
Z exp i k . ( r − r )
1 X
+ d3 r 0 Vimp (r0 ) Ψ+ (r0 )
V 0
E − Ek 0 + i η
k

(2179)

The sum over k 0 in the second term is evaluated as


 
0
exp i k . R
exp[ i k 0 . R ]
Z
1 X 1 3 0
= d k
V E − Ek 0 + i η ( 2 π )3 E − Ek 0 + i η
k0
Z ∞ 0 0
2π 2m 0 0 exp[ i k R ] − exp[ − i k R ]
= 2 dk k 2 m
( 2 π )3 i R h̄ 0 k 2 − k 02 + i h̄2 η
Z ∞
m 0 0 exp[ i k 0 R ]
= dk k
i 2 π 2 h̄2 R −∞ k 2 − k 02 + i 2h̄m 2 η
m
= − exp[ i k R ] (2180)
2 π h̄2 R
The integral in the third line has been evaluated by Cauchy’s theorem, in which
the contour along the real k axis is closed by a semicircle at infinity in the
upper-half complex plane. This contour encloses the pole at k 0 = k + i δ,
but excludes the pole at k 0 = − k − i δ. Thus, the wave function for the

626
k'
k
θ

Figure 256: The asymptotic behavior of the stationary state corresponding to


an incoming plane wave and an outgoing spherical wave.

stationary scattering state has the solution


 
1
Ψ+ (r) = √ exp i k . r
V
 
Z exp i k | r − r0 |
m
− d3 r 0 Vimp (r0 ) Ψ+ (r0 )
2 π h̄2 | r − r0 |
(2181)

This corresponds to a linear superposition of the unscattered wave and a spher-


ical outgoing wave emanating from the impurity. Far from the impurity, the
wave may be expressed in terms of the scattering amplitude f (k, θ) via
 
  exp i k r !
1
lim Ψ+ (r) → √ exp i k . r + f (k, θ)
r → ∞ V r
(2182)

where the direction of the incident beam defines the z axis, and θ is the angle
between the z axis and r. The asymptotic form can also be expressed in terms
of the phase-shifts δl (k) via a partial-wave analysis. Far from the impurity
potential, r  r0 , the particles undergoing the scattering are asymptotically
free. In this region of space, the wave function can be expressed as a linear

627
superposition of energy eigenstates of the free particle Hamiltonian, with energy
2
k2
E = h̄2 m . Thus,
∞  
1 X
lim Ψ+ (r) ∼ √ ( 2 l + 1 ) al jl ( k r) + bl ηl ( k r ) Pl (cos θ)
r → ∞ V l=0
(2183)
where al and bl are coefficients that are to be determined, and jl (x) and ηl (x) are
the Riccati Bessel functions. The Riccati Bessel functions have the asymptotic
forms
sin( k r − l π2 )
jl (kr) ∼
kr
cos( k r − l π2 )
ηl (kr) ∼ − (2184)
kr
Hence, if one defines the phase shift, δl (k) via
bl
tan δl (k) = − (2185)
al
then the asymptotic form of the solution can be simply written as

1 X (2l + 1) π
lim Ψ+ (r) ∼ √ Al sin( k r − l + δl (k) ) Pl (cos θ)
r → ∞ V l=0 kr 2
(2186)
which is similar to the general solution for the free particle, except that the short-
ranged potential has introduced a phase shift in the argument of the trigono-
metric function.

The incident particle’s state can be expanded in terms of the free particle
states with energy E and angular momentum l, as
  ∞
1 1 X
√ exp i k . r = √ ( 2 l + 1 ) il jl (kr) Pl (cos θ) (2187)
V V l=0

Also, the scattering amplitude f (k, θ) can be expanded in terms of the Legendre
polynomials
X∞
f (k, θ) = ( 2 l + 1 ) fl (k) Pl (cos θ) (2188)
l=0
On combining the above expressions with the two asymptotic expression for
Ψ+ (r) and equating the coefficients of exp[ikr] and exp[−ikr], one finds the two
conditions
 
−l
i Al exp − i δl (k) = 1
 
i−l Al exp i δl (k) = 1 + 2 i k fl (k) (2189)

628
Hence, the partial wave scattering amplitudes fl (k) are given by
 
( 1 + 2 i k fl (k) ) = exp 2 i δl (k) (2190)

Therefore, the partial-wave scattering amplitudes fl (k) are given by


exp[ 2 i δl (k) ] − 1
fl (k) =
 2 i k
sin δl (k)
= exp i δl (k) (2191)
k
Thus, the amplitude of the asymptotic scattered wave is also determined by the
phase shift. The asymptotic form is given by
∞  
1 X (2l + 1) l lπ
lim Ψ+ (r) ∼ √ i exp i δl Pl (cos θ) sin( k r − + δl )
r → ∞ V kr 2
l=0
(2192)

Since the on energy-shell T-matrix has matrix elements which satisfy

< k 0 | T̂ | k > = < k 0 | V̂imp | Ψ > (2193)

one finds that


 
∞ exp i 2 δl − 1
mV 0
X
< k | T̂ | k > = Pl (cos θ) (2194)
2 π h̄2 l=0
2ik

The differential scattering cross-section is given in terms of the on shell T-matrix


by
 2 2
dσ mV 0

= 2
< k | T̂ | k > (2195)
dΩ 2 π h̄
Therefore, in general, the angular dependence is expressible in terms of Legendre
polynomials and the phase shifts. In the limit k → 0, only the s-wave phase
shift δ0 is significant. Therefore, in the limit k → 0, one finds
 
exp i 2 δ0 − 1
mV 0
< k | T | k > = (2196)
2 π h̄2 2ik
Thus, the scattering cross-section is given by
dσ sin2 δ0 (k)
= (2197)
dΩ k2
and the total cross-section σ is given by
4 π sin2 δ0 (k)
σ = (2198)
k2

629
The impurity scattering cross-sections are independent of the volume of the
sample, and give the largest contribution to the scattering at energies where
δ0 (k) = π2 .

The density of states due to the impurity can be expressed in terms of the
phase shift δ0 (k). The impurity is assumed to be contained in the host sample
which has the form of a sphere of radius R. The wave functions are required to
vanish at the surface of the sample, r = R. Hence, the phase shift must satisfy
the condition
k R + δ0 (k) = n π (2199)
This condition quantizes the allowed values of k. Since successive states satisfy
this condition with consecutive integers n and n + 1, then the difference in the
k values of any two consecutive states is given by
∂δ0
∆k ( R + ) = π (2200)
∂k
Thus, the number of states per k interval is determined by
 
1 1 dδ0 (k)
= R + (2201)
∆k π dk
dk
On multiplying this by dEk , one obtains the impurity density of states, per spin,
226
as  
1 ∂δ0
ρimp () = (2202)
π ∂
Systems which have a rapid variation of the phase shift at the Fermi energy,
have a large impurity density of states. On integrating the impurity density of
states with respect to , one finds that the number of states due to the impurity
with energy less than , N (), is given by
1
N () = δ0 () (2203)
π
The condition for electrical neutrality for a charge Z | e | determines the phase
shift at the Fermi energy, through Friedel’s sum rule227
2
Z = δ0 (µ) (2204)
π
where the factor of two occurs due to the spin degeneracy.

The T-matrix for the short-ranged potential is given by


 −1
X Vimp
< k 0 | T̂ (z) | k > = Vimp 1 − (2205)
z − Ek”
k”

226 In general, the impurity density of states has contributions from all the partial wave
∂δ
channels ρimp () = π1
P
l
( 2 l + 1 ) ( ∂l ).
227 J. Friedel, Phil. Mag. 43, 153 (1952).

630
δ0(ε)/π
0.5 πρ0(ε) t
ρimp(ε) t
F(ε) t
0.3

0.1

-6 -4 -2 -0.1 0 2 4 6

-0.3
t/V

-0.5
ε/t

Figure 257: The energy variation of the s-wave phase shift δ0 () and impurity
density of states ρimp () [in units of t ], for an s-wave resonance.

Hence, on using the partial-wave expansion for the T-matrix


 
∞ exp i 2 δl − 1
mV 0
X
< k | T̂ | k > = P l (cos θ) (2206)
2 π h̄2 l=0
2ik

one finds that the short-ranged potential only produces an l = 0 phase shift.
The phase shift is given by

π Vimp ρ0 ()
tan δ0 (k) = − P 1 (2207)
1 − Vimp k0  − E k0

It can be seen that at a resonance, the denominator vanishes and the phase shift
is equal to π2 , modulo π. Hence, the scattering cross-section is maximized at a
resonance. Furthermore, since the change in the density of states is given by

1 ∂δ0 ()
ρimp () = (2208)
π ∂
then one has
 
1 ∂ π Vimp ρ0 ()
ρimp () = − tan−1 P 1 (2209)
π ∂ 1 − Vimp k0  − E k0

which is consistent with eqn(2170). The impurity density of states can be ex-
pressed in terms of the phase shift as
   Z ∞ 0

1 ∂ρ0 1 ∂ 0 ρ0 ( )
ρimp () = ρ−1
0 () sin 2 δ 0 () − sin2
δ 0 () d
2π ∂ π2 ∂ −∞  − 0
(2210)
On applying Friedel’s sum rule

631
πZ
δ0 (µ) = (2211)
2
one finds that the density of states at the Fermi level is given by
     Z ∞ 0

−1 1 ∂ρ0 1 2 πZ ∂ 0 ρ0 ( )

ρimp (µ) = ρ0 (µ) sin( π Z ) − 2 sin d
2π ∂ π 2 ∂ −∞  − 0 µ
(2212)
The first term corresponds to the change in the density of states found in the
rigid band approximation. As the impurity adds Z electrons to the alloy, the
Fermi energy changes by an amount
Z −1
∆µ = ρ (µ) (2213)
2 0
In the rigid-band approximation, the change in the density of states at the Fermi
energy is given by
ρimp (µ) = ρ0 (µ + ∆µ) − ρ0 (µ)
 
∂ρ0
≈ ∆µ + ...
∂
 
−1 Z ∂ρ0
≈ ρ0 (µ) (2214)
2 ∂
Therefore, the change in the density of states at the Fermi energy, when calcu-
lated in the rigid-band approximation, is finite for all values of Z. On the other
hand, the exact result vanishes for even integer values of Z. In the cases of
odd integer values of Z, the leading term in the exact impurity density of states
comes from the second term of eqn(2212). The second term is independent of
the sign of Z, and decreases as the average value of Z is decreased. The finite
impurity density of states at the Fermi energy gives rise to an impurity contri-
bution to the specific heat and the susceptibility. These impurity contributions
have been observed in alloys where the charge difference between the host and
impurity ions have different signs and the results agree with the theoretical pre-
dictions228 .

——————————————————————————————————

18.1.1 Exercise
Consider two impurities which act as s-wave scattering centers that are sepa-
rated by a distance R. Determine a formal expression for the exact Green’s
function for the conduction electrons.

——————————————————————————————————

228 A. M. Clogston, Phys. Rev. 125, 439 (1962), A. M. Clogston and V. Jaccarino, Phys.

Rev. 121, 1357 (1961).

632
Figure 258: The susceptibility due to non-magnetic impurities in V3 Ga versus
impurity concentration x. The susceptibility decreases for both types of impu-
rities, even though Z has opposite signs for Cr and T i. [After A.M. Clogston
(1962).].

18.2 Virtual Bound States


The virtual bound state can be envisaged as an (almost) localized level that has
a finite probability amplitude for transitions into the conduction band states.
These virtual bound states are most frequently found for 3d transition metal
impurities in metals or in mixed valent lanthanide element impurities in metals.
In both these cases, the potential well has a large centrifugal barrier

h̄2 l ( l + 1 )
Vl (r) = (2215)
2 m r2
The centrifugal potential prevents the 3d states from being filled until after the
4s states are filled or, in the case of the lanthanide elements, the 4f states re-
main unfilled until after the 6s, 5p and 5d states are all occupied. When the
nuclear potential is strong enough, such that the 3d or 4f states can be occupied
in the ground state, the ion localizes an electron within the centrifugal barrier
in an inner ionic shell. For example, in the Ce atom the 4f wave function is
localized, in that it has a spatial extent of 0.7 a.u. which lies inside the core-like
5s and 5p orbitals. However, its’ ionization energy is small and comparable to
the ionization energy of the band-like 6s and 6p orbitals. As the localized state
is degenerate with the conduction band states, there is a finite probability am-
plitude for an electron in the 4f level to tunnel through the barrier. The virtual
bound state describes an extended state which, through resonant scattering,
builds up a significant local character. The virtual bound state in a metal may
be modelled by a Hamiltonian which is the sum of three terms

Ĥ = Ĥ0 + ĤV = Ĥc + Ĥd + ĤV (2216)

where Ĥc describes the electrons in the conduction band, the Hamiltonian Ĥd
represents the (isolated) localized d level on the impurity and the term ĤV

633
0.08

4f
5d
0.06
6s

P(r) r2
0.04

0.02

0
0 20 40 60 80

Z r / a0

Figure 259: The schematic spatial dependence of 4f electron densities relative


to the 5d and 6s electron densities.

describes the coupling. The conduction band Hamiltonian is expressed in terms


of the number of conduction electrons in the Bloch states (k, σ) with dispersion
relation Ek through
X
Ĥc = Ek n̂k,σ
k,σ

Ek c†k,σ ck,σ
X
= (2217)
k,σ

where c†k,σ and ck,σ , respectively, create and annihilate an electron in the con-
duction band state with Bloch wave vector k and spin σ. Likewise, the energy
for an electron in the localized d state is given by the binding energy Ed times
the number of d electrons of spin σ,
X
Ĥc = Ed n̂d,σ
σ
X
= Ed d†σ dσ (2218)
σ

where d†σ and dσ respectively create and annihilate an electron of spin σ in


the localized d state. The hybridization or coupling term is given by the spin
conserving Hamiltonian
 
1 X
ĤV = √ V (k) c†k,σ dσ + V ∗ (k) d†σ ck,σ (2219)
N k,σ

The first term represents a process whereby an electron in the d orbital tun-
nels into the conduction band, and the Hermitean conjugate term represents
the reverse process. It is assumed that the conduction band states have been
orthogonalized to the localized states, so that the conduction band fermion op-
erators anti-commute with all the local fermion operators.

634
The Resolvent Green’s function can be calculated from the expression
1 1
( z − Ĥ0 ) = 1 + ĤV (2220)
z − Ĥ z − Ĥ
Evaluating the matrix elements of this equation between the one-electron eigen-
states of Ĥ0 yields the coupled equations
1 1 X 1
( z − Ed ) < d | |d > = 1 + √ V (k) < d | |k >
z − Ĥ N k
z − Ĥ
(2221)
and
1 1 1
( z − Ek ) < d | |k > = √ V ∗ (k) < d | |d >
z − Ĥ N z − Ĥ
(2222)
These equations can be combined to yield the matrix elements of the resolvent
Green’s functions as
1
Gd,d (z) = < d| |d >
z − Ĥ
1
= (2223)
z − Ed − Σ(z)
where the d-electron self-energy Σ(z) is defined by
1 X | V (k) |2
Σ(z) = (2224)
N z − Ek
k

The real part of the self-energy can be interpreted as producing a renormaliza-


tion of the energy of the localized level Ed . The imaginary part of the self-energy
can be interpreted as giving rise to an width or lifetime τ such that

= − =m Σ(Ed + iη) (2225)

The conduction band Resolvent Green’s function is evaluated, from a similar
set of coupled equations as
1
Gk,k0 (z) = < k| | k0 >
z − Ĥ
δk,k0 1 V (k) V ∗ (k 0 )
= + Gdd (z)
z − Ek N z − Ek z − Ek 0
(2226)
The matrix elements of the T-matrix between different Bloch states is identified
as
< k | T (z) | k 0 > = V (k) Gd,d (z) V ∗ (k 0 ) (2227)

635
3.0

2.0

ρd(ε)
1.0

2∆

0.0
-6 -5
Ed
-4 -3 ε -2

Figure 260: The impurity d-density of states for a virtual bound state.

From these equations, it can be seen that the density of states of the impurity
d level is given in terms of the imaginary part of Σ( + iη) via
 
1
ρd () = − =m Gd,d ( + iη)
π
1 =m Σ( + iη)
= − 2 2
π
 
 − Ed − <e Σ( + iη) + =m Σ( + iη)

(2228)

The impurity density of states is approximately in the form of a Lorentzian


centered on Ed , and has a width given by =m Σ( − iη). The width is given by
π X
=m Σ(Ed − iη) = | V (k) |2 δ( Ed − Ek )
N
k
1
≈ π | V |2 ρ0 (Ed ) (2229)
N
which is related to the Fermi Golden rule expression for the rate for the local-
ized electron to tunnel into the conduction band density of states ρ0 (). Thus,
the virtual bound state can be interpreted in terms of a narrow band density of
states which is weakly coupled to the extended conduction band states.

——————————————————————————————————

18.2.1 Exercise
Determine a formal expression for the change in the conduction band density
of states due to the presence of a single impurity with a virtual bound state.

636
——————————————————————————————————

18.3 Disorder
Give a distribution of impurities in a solid, the potential in the solid will be
non-uniform. The thermodynamic properties of the solid can be expressed in
terms of the energy eigenvalues, or alternatively the poles of the Green’s func-
tion. For a macroscopic sample, the exact distribution of impurities will not be
measurable and the thermodynamic properties are expected to be representa-
tive of all distributions of impurities. Therefore, the average value of a quantity
can be represented by averaging over all configurations of the impurities. It can
easily be shown that the configurational averaged density of states is given by
the discontinuity across the real axis of the configurational averaged resolvent
Green’s function.

The Hamiltonian of a binary (A-B) alloy, with site disorder, may be repre-
sented by
Ĥ = Ĥ0 + V̂ (2230)
The Hamiltonian H0 describes the tight-binding bands of a pure metal with a
dispersion relation
Xd
Ek = − t cos ki ai (2231)
i=1
The randomness appears as a shift of the binding energies of the atomic orbitals,
due to the potential operator V̂ . The potential V̂ is written as the sum of local
or on-site potentials
X
V̂ = ER | φ̃R > < φ̃R | (2232)
R

where the state | φ̃R > represents the single electron Wannier state at site R.
The single site energies ER can take on the values EA or EB depending on the
type of atom present at site R.

The average Green’s function is given by


 −1
G(z) = z − Ĥ0 − V̂ (2233)

which can be expressed as


 −1
G(z) = z − Ĥ0 − Σ(z) (2234)

where the operator Σ(z) is complex and is known as the self-energy due to
disorder. Since the configurational averaged Green’s function has translational

637
invariance, then so does the self-energy. The disorder self-energy represents the
effect that the randomly distributed impurities have on the eigenvalue spectrum.
The spectrum may be composed of extended states and localized states229 . Due
to the fluctuations in the random potential, the energy eigenvalues correspond-
ing to localized eigenstates may be broadened to form continua.

The averaged Green’s function can be calculated by expanding the Green’s


function in powers of the potential and then performing the configurational aver-
age. For strongly fluctuating potentials, the resulting power series may be slowly
convergent, or it may not even be convergent at all. Therefore, to increase the
rate of convergence, it may be preferable to expand the Green’s function about
the self-energy. This procedure leads to the coherent potential approximation.

18.4 Coherent Potential Approximation


The potential difference between a specific realization of the potential V̂ due to
the impurities and the self-energy can be expressed as
d (z) = V̂ − Σ(z)
∆V (2235)
The resolvent Green’s function for this type of disordered impurity problem can
be expressed as
 −1
Ĝ(z) = z − Ĥ0 − Σ(z) − ∆V d (z) (2236)

which can be expressed in terms of the T-matrix via


Ĝ(z) = G(z) + G(z) T̂ (z) G(z) (2237)
where the T-matrix is given by
 −1
T̂ (z) = ∆V
d (z) 1 − G(z) ∆V
d (z) (2238)

On taking the configurational average, one finds that the averaged T-matrix
must be zero
 −1
T (z) = ∆V (z) 1 − G(z) ∆V (z)
d d = 0 (2239)

This equation can be used to obtain the self-energy.

For the A − B alloy the effective potential is


X
∆V
d (z) = ( ER − Σ(z) ) | φ̃R > < φ̃R | (2240)
R

229 P. W. Anderson, Phys. Rev. 109, 1492 (1958).

638
The concentration of A atoms is denoted by c, so the concentration of B atoms
is ( 1 − c ). It is assumed that the two types of atoms are randomly dis-
tributed on the lattice sites, such that there is one atom at each lattice site.
It is also assumed that the T-matrix can be represented as a sum of single-site
T-matrices, in which the scattering is referenced to an appropriately chosen av-
eraged medium. This is the single-site approximation. The averaged T-matrix
can be written as
EA − Σ(z)
T (z) = c  
1 − EA − Σ(z) < R0 | G(z) | R0 >

EB − Σ(z)
+(1 − c)  
1 − EB − Σ(z) < R0 | G(z) | R0 >

(2241)

The Coherent Potential Approximation230 (C.P.A.) sets

T (z) = 0 (2242)

The resulting equations are non-trivial to solve since the Green’s function in the
denominator is formed from a sum over the Bloch states and also involves the
self-energy.
1 X 1
< R0 | G(z) | R0 > =
N z − Σ(z) − Ek
k
= < R0 | G0 ( z − Σ(z) ) | R0 > (2243)

where Ĝ0 (z) is the Green’s function for the tight-binding Hamiltonian. Never-
theless, this can be solved numerically or alternatively, if the sum over Bloch
energies can be evaluated analytically, an analytic solution may be found.

The C.P.A. is expected to be valid in various limits. These include the limit
of a dilute concentration of impurities 1  c, weak scattering t  | EA − EB |
and trivially in the atomic limit, where the single-site approximation is exact.
In general, the C.P.A. may be only trusted to yield the density of states and
thermodynamic properties 231 , and not transport properties. The density of
states obtained from this method resembles a smeared version of the weighted
sums of the density of states of a solid composed of A atoms and the density of
states composed of B atoms. For small magnitudes in the differences of the site
energies, the two components overlap, but they separate for large differences
in the site energies. When the bands are split, the widths of the component
bands are drastically modified from the ideal superposition. The change in the
230 P. Soven, Phys. Rev. 156, 809 (1967), B. Velický, S. Kirkpatrick and H. Ehrenreich, Phys.

Rev. 175, 747 (1968).


231 R. J. Elliott, J. A. Krumhansl and P. L. Leath, Rev. Mod. Phys. 46, 465 (1975).

639
0.8

c = 0.1 (Ea-Eb)/W
0.6 0.0
0.5
1.0

ρ(ω)
1.5
0.4

0.2

0
-2 -1.5 -1 -0.5 0 0.5 1 1.5

ω/W

Figure 261: The dependence of the CPA density of states ρ() on the strength
−Eb
of the disorder ( EaW ), where 2W is the band width of the pure material.

widths of the split bands reflects the increasing separation between sites of the
same type which decreases the tendency to form bands. The effect of the impu-
rity scattering is to produce a smearing, which washes out any structure such
as van Hove singularities. Transport properties crucially depend on the spa-
tial extended nature of the energy eigenstates, which may be destroyed by the
fluctuations in the random potentials. This type of phenomenon is completely
absent in C.P.A., and can lead to the energy eigenstates becoming localized 232 .

——————————————————————————————————

18.4.1 Exercise
Assume that after diagonalizing the tight-binding Hamiltonian Ĥ0 , one obtains
a density of states ρ0 (z) of the form
2 p 2
ρ0 (z) = W − z2 (2244)
π W2
Determine the form of associated Green’s function G0 (z). Show that the CPA
equations can be reduced to a cubic equation for x() =  − Σ(). (Be sure to
check whether the cubic equation yields spurious solutions not present in Soven’s
initial equation.) Determine the bounds of the CPA spectra and compare them
with the Saxon-Huttner bounds of

| EX −  | < W (2245)

where X is either A or B. Determine the condition for the solution of the CPA
equations to exhibit a pole in z().

232 P. W. Anderson, Phys. Rev. 109, 1492 (1958).

640
——————————————————————————————————

18.5 Localization
The phenomenon of disorder induced localization is easiest to understand in
terms of states at the tail edge of a band. Just as one impurity with a suffi-
ciently strong attractive potential may cause a bound state to form around it,
a bound state may also be formed for a number of nearby atoms with weaker
attractive interactions, in which case the bound state may be of larger spatial
extent. In both cases, they will produce localized states with energies below
the continuum of the density of states. A distribution in the spatial separation
of the impurity atoms will smear the spectrum of these discrete bound states.
The localized states manifest themselves as low-energy tails to the density of
states233 . As the strength of the disorder is increased, the number of localized
states in the tails of the density of states will increase. One surprising feature
is that a sharp energy, the mobility edge234 , separates the states that extend
throughout the crystal from the localized states. The length scale over which
the states on the localized side of the mobility edge decay are extra-ordinarily
long235 . These states and cannot be treated by perturbation methods but re-
quire renormalization group types of approach.

On using the electron-hole symmetry for states at the top of the band, one
discovers that the states at the top edge of the band will also become localized
due to disorder, and also have a mobility edge. On increasing the strength of
the disorder, the mobility edges will move towards the middle of the bands.
A disorder driven metal insulator transition will occur when the mobility edge
crosses the Fermi energy. This type of transition is known as the Anderson
transition236 . The effect of many-body interactions complicate the physics on
the metallic side of the Anderson transition, where weak localization occurs237 .
On the insulating side of the transition, conduction will still be possible but
only due to the thermal excitation of electrons to the itinerant states above the
mobility edge, or by thermally assisted tunnelling processes. For sufficiently
strong disorder all the states in the band will become localized. All states in
one-dimensional and two-dimensional systems must become localized, for arbi-
trarily small strengths of disorder. However, this localization will only show up
in experiments if the length scale over which the states are localized is smaller
than the sample size.

233 P. W. Anderson, Phys. Rev. 109, 1492 (1958).


234 N. F. Mott, Adv. in Phys. 16, 49 (1967).
235 D. J. Thouless, Phys. Rev. Lett. 39, 1167 (1977).
236 N. F. Mott, Metal-Insulator Transitions, Taylor and Francis, London (1974).
237 P. A. Lee and T. V. Ramakrishnan, Rev. Mod. Phys. 57, 287 (1985).

641
0.5

0.4

0.3

ρ(ε) 0.2 εF

0.1

ε− ε+
0
-4 -2 0 2 4

ε/W

Figure 262: The density of states of a disordered band ρ(), showing the upper
and lower mobility edges ± and the Fermi-energy F .

18.5.1 Anderson Model of Localization


In a doped semiconductor such as P doped Si, as the impurity concentration
is increased, it is expected that the energy levels of the isolated impurities will
broaden and form bands. For large concentrations, the impurity level wave func-
tions are expected to overlap and become extended. Thus, it is expected that
a metal-insulator transition will occur as a function of impurity concentration.
The metal insulator transition can be described by a tight-binding model of a
disordered system

εi c†i,σ ci,σ − t c†i,σ cj,σ


X X
Ĥ = (2246)
i,σ i,j,σ

where t are the nearest neighbor tight-binding hopping matrix elements and the
sum over (i, j) are assumed to run over pairs of nearest neighbors lattice sites.
The site energies εi are assumed to be random variables uniformly distributed
over an energy width ∆V
1 ∆V ∆V
P (ε) = for − < ε <
∆V 2 2
= 0 otherwise (2247)
The degree of disorder is measured by the dimensionless parameter ∆V /t.

For sufficiently large ∆V /t the states are expected to all be localized. The
critical value of (∆V /t)c is expected to be dependent on the dimensionality of
the lattice. In three dimensions, the critical value is estimated as (∆V /t)c ∼ 15.
For ∆V /t less than the critical value, the states around the center of the tight-
binding bands are extended while states near the band edges are localized. There

642
+ exp[-|r|/ξ]

Re [φk(r) ]
- exp[-|r|/ξ]

r/a

Figure 263: A schematic depiction of a the Real part of a wave function φk (r)
localized due to randomness. The exponential envelope of the wave function is
depicted by the dashed red lines.

are energies Ec , called mobility edges that separate the localized and extended
states. When the chemical potential µ crosses the mobility edge, the states
at the Fermi energy change their characters and a metal non-metal transition
occurs. This is known as the Anderson transition.

The wave functions corresponding to extended and localized states have


different characters. A wave function for the disordered solid can be expressed
as a linear combination of atomic wave functions
X
ψ(r) = C(R) φ(r − R) (2248)
R

A delocalized wave function has an amplitude C(R) which does not decay to
zero at large distances. A localized wave function is expected to decay to zero
with an exponential envelope
 
| C(R) | ∼ exp − | R | /ξ (2249)

The spatial extent of the envelope is given by the correlation length ξ. The cor-
relation length is expected to depend on the energy E of the energy eigenstate.
The correlation length is expected to diverge as E approaches the mobility edge
Ec . In the Anderson transition, the spectral density of localized states is ex-
pected to be continuous. Numerical studies 238 show that the wave function
exhibits long-ranged fluctuations close to the critical value of ∆V /t, and ap-
pears to be self-similar when viewed at all (sufficiently long) length scales.

238 J. T. Edwards and D. J. Thouless, 5, 807 (1972).

643
18.5.2 Scaling Theories of Localization
Since numerical studies of Anderson localization are hampered by finite-size
effects which tend to obscure the effect of localization, Licciardello and Thou-
less239 introduced a number g(L) which describes the sensitivity of energy eigen-
values on the boundary conditions, for a system with linear dimension L. The
Thouless number is defined as the ratio
∆E
g(L) = (2250)
δE
where ∆E is the shift in energy levels that occurs when the boundary conditions
on the wave function are changed from periodic to anti-periodic. The quantity
δE is the mean spacing of the energy levels of the finite size sample. If the wave
functions are exponentially localized, it is expected that
 
2L
g(L) ∝ exp − (2251)
ξ(E)

Hence in localized states, g(L) tends to zero for ξ(E) < L since the wave
functions of localized state are insensitive to the choice of boundary condition.
On the other hand, if the wave functions are extended, the energy shift due to
the different boundary conditions should be proportional to

(2252)
τ
where τ is the time required for the electron to diffuse to the boundary of
the sample. This diffusion time is at most algebraically dependent on L. The
different dependencies of g(L) on L provide a simple criterion in numerical
studies as to whether the states are extended or localized. An elegant argument
has been put forth240 which indicates that the quantity g(L), the Thouless
number, is proportional to the conductance, G(L)

G(L) h
g(L) = (2253)
2 e2
The conductance G(L) is related to the conductivity σ via a factor of the area
divided by the length
σ
G(L) = Ld−1 (2254)
L
Hence, the Thouless number g(L) is related to the conductivity by

g(L) ∝ Ld−2 σ (2255)

239 D. C. Licciardello and D. J. Thouless, Phys. Rev. Lett. 35, 1475 (1975), D. C. Licciardello

and D. J. Thouless, J. Phys. C, 8, 4157 (1975).


240 J. T. Edwards and D. J. Thouless, J. Phys. C 5, 807 (1972).

644
The scaling theory of localization241 is based upon the length dependence
of g(L). A scale change, of a d-dimensional system with linear dimension L, is
produced when the length scale L is changed to b L. It is expected that g(bL)
is related to g(L) and the factor b, and nothing else. This is summarized in the
formula
g(bL) = f [b, g(L)] (2256)
where f (x) is a universal scaling function, which only depends on the dimen-
sionality d of the lattice. An infinitesimal scale change is defined by
dL
b = 1 + (2257)
L
so that the scale can be changed continuously. A scaling function β[g(L)] is
introduced via the definition
d ln g(L) ∂f (b, g)/∂b
β[g(L)] = = (2258)
d ln L g(L)

The functional β[g] completely specifies the scaling property of the conductivity
in disordered systems. It is assumed that β(g) is a smooth continuous function
of g which is independent of L. This implies that the change in the effective
disorder of the system can be uniquely determined at one length scale from
knowledge of its value at a smaller length scale.

The asymptotic forms of β can be found in the asymptotic limits g → 0


and g → ∞. In the strongly localized regime g → 0 where the wave function
is exponentially localized, one finds that since
 
L
g(L) ∝ exp − 2 (2259)
ξ

for L  ξ, then
L
β(g) = − 2 ∝ ln g + Const. (2260)
ξ
Thus, β(g) tends to − ∞ as g → 0. In the metallic limit g → ∞ and Ohm’s
law applies so σ is finite and independent of L, if the length scale is greater than
the mean free path λ. Therefore, in this region one has

β(g) → ( d − 2 ) (2261)

The qualitative dependence of β(g) on g can be determined from continuity and


the use of perturbation expansions in g and g −1 . The variation of β(g) with
ln g is shown in Fig(264). From this, one finds that the system is localized for
all spatial dimensionalities less than or equal to two, d < 2. For d < 2,
241 F. J. Wegner, Zeit. für Physik 25, 327 (1976).

E. Abrahams, P. W. Anderson, D. C. Licciardello and T. V. Ramakrishnan, Phys. Rev. Lett.


42, 673 (1979).

645
Figure 264: A sketch of the universal curve showing the dependence of β(g) on
ln g. The dashed line for d = 2 represents the un-physical variation required for
there to be a jump in the conductivity. [After Abrahams et al. (1979).]

as β(g) is always negative, then g(L) scales to zero on increasing L. In two-


dimensions, the conductivity decreases with increasing L. The decrease of σ
is logarithmic at large values of g and exponential at small values of g. Thus,
in two-dimensions and in the limit of large systems, no matter how weak the
randomness is, the states are always localized. By contrast, for d > 2, there
is a critical value of gc such that for g > gc the system scales to the metallic
limit where β(g) = ( d − 2 ). As β(g) is positive for g > gc , then when L is
increased above λ, limL → ∞ g(L) → ∞. Since β(g) is negative for g < gc ,
then g(L) scales to zero on increasing L. At the critical value of gc , the scaling
function is zero, β(gc ) = 0, and varies approximately linearly with ln g/gc so it
can be shown that the slope s determines the exponent of the correlation length
as s−1 . From the scaling theory, one can infer the dependence of conductivity
on the concentration of impurities, c. For c > c0 , close to the metal insulator
transition, the conductivity scales as

σ = σ0 ( c − c0 )1 (2262)

where the exponent of unity can be obtained exactly via perturbation theory242 .

242 L. P. Gor’kov, A. I. Larkin and D. E. Khmel’nitskii, J.E.T.P. Lett. 30, 228 (1979).

646
19 Magnetic Impurities
19.1 Localized Magnetic Impurities in Metals
When transition metal or rare earth impurities are dissolved in simple metals,
the electronic states on the impurities hybridize with the conduction band states
and form a Friedel virtual bound state. Since the impurity states are localized
the Coulomb interaction U between two electrons occupying these states is large
and has to be taken into consideration. The Hamiltonian can be expressed as

Ĥ = Ĥ0 + Ĥint (2263)

where the Hamiltonian Ĥ0 represents the non-interacting conduction band and
the virtual bound state

Ek c†k,σ ck,σ +
X X
Ĥ0 = Ed d†σ dσ
k,σ σ
" 
c†k,σ
X

+ V (k) dσ + V (k) d†σ ck,σ
k,σ

(2264)

and the Coulomb interaction U between a pair of electrons in the (spin only
degenerate) impurity state is given by

Ĥint = U d†↑ d†↓ d↓ d↑ (2265)

This is the Anderson impurity Hamiltonian. The Anderson impurity model


is exactly soluble using numerical renormalization group243 , or Bethe-Ansatz
techniques244 . The mean-field solution will be outlined below.

19.2 Mean-Field Approximation


The interaction term Ĥint can be expressed in terms of fluctuations of the
number of d-electrons with spin σ

∆n̂σ = d†σ dσ − < | d†σ dσ | > (2266)

and the average value


nσ = < | d†σ dσ | > (2267)
When written in terms of the fluctuations of the occupancy of the d-orbitals,
the Hamiltonian takes the form
X
Ĥint = U ∆n̂↑ ∆n̂↓ + U ∆n̂σ n−σ + U n↑ n↓ (2268)
σ
243 K. G. Wilson, Rev. Mod. Phys. 47, 773 (1975).
244 N. Andrei, K. Furuya and J. H. Lowenstein, Rev. Mod. Phys. 5, 331 (1983), A. M.
Tsvelik and P. B. Weigmann, Adv. in Phys. 32, 453 (1983).

647
In the mean-field approximation, the term quadratic in the occupation number
fluctuations is neglected, yielding
X
ĤM F = U n̂σ n−σ − U n↑ n↓ (2269)
σ

The localized electrons experience an effective spin-dependent binding energy


given by X
Ĥd = ( Ed + U n−σ ) d†σ dσ (2270)
σ
where n−σ is the average number of electrons in the localized level of spin −σ.
The spin-dependent occupation number is found as an integral over the density
of states of the virtual bound state, and is given by
Z ∞
nσ = d f () ρσd () (2271)
−∞

where the spin-dependent impurity density of states is given by


 
σ 1 1
ρd () = − =m (2272)
π  + i η − Ed − U n−σ − Σ( + iη)
The self-energy can be approximated by a constant imaginary part with value ∆
and a small energy shift that can be absorbed into the definition of Ed . Hence,
the spin-dependent density of states can be approximated by a Lorentzian
1 ∆
ρσd () ≈ 2 (2273)
π ε − Ed − U n−σ + ∆2
where the width of the Lorentzian is given by
X
∆ = π | V (k) |2 δ(Ed − Ek )
k

∼ π | V |2 ρ0 (Ed ) (2274)

Thus, at T = 0, one finds that the average occupation of the localized state
with spin σ is given by the expression
 
1 −1 Ed − µ + U n−σ
nσ = cot (2275)
π ∆
Since cot θ is defined for θ on the interval 0 to π and runs between ∞ and − ∞,
then cot−1 x has values that run from π to 0. The two coupled equations for nσ
and n−σ have to be solved self-consistently. This can be done by introducing
the parameters
µ − Ed
x =

U
y = (2276)

648
1 1

m=0
− ρ↓(E) ρ↑(E)
− ρ↓(E) ρ↑(E) m>0
0.5 0.5

Ed+Un-µ

E-µ

E-µ
U(n↑-n↓)
0
0

-0.5
-0.5

-1
-1

Figure 265: The spin-dependent Anderson-impurity d-density of states. The


local density of states of the up-spin sub-band is degenerate with the down-spin
sub-band, when there is no local-moment is present. In the case where a local-
moment has been formed, the up-spin sub-band (blue) is shifted by an energy
U m relative to the down-spin sub-band (red). [After P.W. Anderson (1961).].

which are dimensionless measures of the position of the Fermi energy relative to
the d level and the Coulomb interaction. The pair of self-consistency equations
become

cot π n↑ = ( y n↓ − x )
cot π n↓ = ( y n↑ − x )
(2277)

The non-magnetic phase is described by

n↑ = n↓ = n (2278)

This has a unique solution for n, in the range 0 < n < 1 which is given by
the solution of
cot π n = ( y n − x ) (2279)
The non-magnetic solution corresponds to a partial occupation of the localized
levels. In this case, the virtual bound state does not posses a magnetic moment.
However, if y is large the equations have two degenerate magnetic solutions.
The magnetic solutions only occur for sufficiently large values of y, and when
they occur they are stable because they minimize the energy. The boundary
which separates the regions were magnetic moments can occur from regions were
the magnetic moment is zero, can be found by expressing the self-consistency
equations in terms of the variable m defined by
1
nσ = n + σm (2280)
2
(σ = ±1) so that the magnetization is simply given by

m = n↑ − n↓ (2281)

649
1

m=0
0.8

0.6

(µ−Ed)/U m>0
0.4

0.2

0
0 0.1 0.2 0.3 0.4

∆/U
Figure 266: The mean-field phase diagram of the single-impurity Anderson
Model. The phase space is separated into two regions, one where there is a
finite local-moment and the other where there is no local-moment. [After P.W.
Anderson (1961).].

The condition that a solution with an infinitesimal value of m first occurs can be
found by linearizing the self-consistency equations in powers of m. On equating
the coefficients of the first two terms in the expansion in m to zero, one finds

cot π n = (yn − x)
π
2 = y (2282)
sin π n
The first equation defines the non-magnetic solution, and the vanishing of the
coefficient of the term linear in m allows a solution with an infinitesimal non-zero
value of m to occur. The above equations can be re-written as
x 1
= ( θ − sin θ )
y 2π
1 1
= ( 1 − cos θ ) (2283)
y 2π
where θ = 2 π n. This pair of equations define the phase boundary line sepa-
rating the areas of phase space in which the impurity is magnetic from the area
in which the impurity is non-magnetic. The tendency for magnetism is strongest
when the d-d interaction, U , is large and when n is close to 12 , i.e., when Ed
and Ed + U are positioned symmetrically about the Fermi level. In this case,
the total number of d-electrons of both spins is almost unity. The non-magnetic

650
Figure 267: The magnitudes of the local magnetic moments of F e (as determined
from the high-temperature Curie-Weiss susceptibility) dissolved in various tran-
sition metal alloys. [After A.M. Clogston et al. (1962).].

solution occurs when U is small or when the d-level is either almost completely
filled or almost completely empty.

For large y, the magnetic solutions are described by


1
nσ ≈ 1 −
πy
1
n−σ ≈ (2284)
π(y − x)

These local moment solutions are doubly-degenerate and correspond to the spin-
up and spin-down states of the impurity. It is to be expected that the solution
should have a continuous symmetry with respect to the orientation of the impu-
rity spin. However, the spin-rotational invariance has been specifically broken
by the mean-field approximation through the choice of a specific quantization
axis. The spin-rotational invariance is expected to be restored by interactions
with conduction electrons which results in resonant scattering processes that
successively flip the impurity spin.

Thus, the mean-field solution of the Anderson model contains solutions with
and without local magnetic moments245 . The appearance of magnetic moments
of transition metal impurities in metals can be interpreted in terms of the change
of position and width of the virtual bound state246 .

245 P.W. Anderson, Phys. Rev. 124, 41 (1961).


246 A.M. Clogston, B. T. Matthias, M. Peter, H. J. Williams, E. Corenzwit and R. C.
Sherwood, Phys. Rev. 125, 541 (1962).

651
19.3 The Atomic Limit
In the case when the hybridization is set to zero, the d-orbital is entirely local-
ized. The local level is entirely decoupled from the conduction band and the
model is exactly soluble. The local d level can be described in terms of the
eigenstates of the d number operator. The four basis states that correspond to
the d level being unoccupied, with energy 0, two states which correspond to the
d-state being occupied by one electron, with energy Ed and one state in which
the d-level is occupied by two electrons. The doubly-occupied state has energy
2 Ed + U .

The projection operators for the atomic energy states can be expressed in
terms of the electron number operators. The projection operator for the unoc-
cupied state P̂0 is given by

P̂0 = ( 1 − n̂σ ) ( 1 − n̂−σ ) (2285)

The projection operator for the singly occupied state with spin σ, P̂σ is given
by
P̂σ = n̂σ ( 1 − n̂−σ ) (2286)
while the projection operator for the doubly-occupied state P̂σ,−σ is simply given
by
P̂σ,−σ = n̂σ n̂−σ (2287)
The projection operators P̂α satisfy the definition of a projection operator

P̂α P̂α = P̂α (2288)

as can be explicitly seen by noting that the number operators satisfy n̂σ n̂σ =
n̂σ . The projection operators have eigenvalues of either one or zero. Further-
more, since the four states are all the energy eigenstates of Ĥ, and hence form
an orthogonal and complete set, one has

P̂α P̂β = 0 (2289)

if α 6= β and X
P̂α = Iˆ (2290)
α

The time-ordered single-electron Green’s function Gdd;σ (t) is defined as


i
Gdd;σ (t, t0 ) = − < | T̂ dσ (t) d†σ (t0 ) | > (2291)

where the d-electron creation and annihilation operators are given in the Heisen-
berg representation so
   
i i
Â(t) = exp + Ĥ t  exp − Ĥ t (2292)
h̄ h̄

652
and T̂ is Wick’s time-ordering operator which is defined as

T̂ Â(t) B̂(t0 ) = Θ(t − t0 ) Â(t) B̂(t0 ) − ( 1 − Θ(t − t0 ) ) B̂(t0 ) Â(t) (2293)

where  and B̂ are both fermion creation and/or annihilation operators. There-
fore, the Green’s function can be expressed as
i
Gdd;σ (t, t0 ) = − Θ(t − t0 ) < | dσ (t) d†σ (t0 ) | >

i
+ ( 1 − Θ(t − t0 ) ) < | d†σ (t0 ) dσ (t) | > (2294)

The first contribution to the time-ordered Green’s function, describes the prob-
ability amplitude that an electron of spin σ added to the d state at time t0 is
still found in the state at a later time t. The second contribution describes
an analogous probability amplitude that the whether the state from which an
electron of spin σ is removed at time t is still empty at a later time t0 .

In the atomic limit, the Green’s function can be explicitly evaluated from
the definition by projecting the state | > onto each of the four atomic energy
eigenstates | 0 >, | dσ >, | d−σ > and | dσ d−σ >. If the state | > is projected
onto the state | 0 > where the d-level unoccupied, the Green’s function can be
evaluated as
 
i i
Gdd;σ (t, t0 ) = − Θ(t − t0 ) < 0 | exp − E d ( t − t0 ) | 0 >
h̄ h̄
 
i i
= − Θ(t − t0 ) exp − Ed ( t − t0 ) (2295)
h̄ h̄

whereas if it is projected onto the state which is singly-occupied by an electron


of spin σ, | dσ > , one finds
 
0 i 0 i 0
Gdd;σ (t, t ) = ( 1 − Θ(t − t ) ) < dσ | exp − Ed ( t − t ) | d σ >
h̄ h̄
 
i 0 i 0
= ( 1 − Θ(t − t ) ) exp − Ed ( t − t ) (2296)
h̄ h̄

These results are not dependent on the interaction U since the excited states,
respectively, consist of a single added electron or the unoccupied state. On the
other hand, if one projects onto a state which is initially singly occupied by an
electron of spin −σ, |d−σ > , the Green’s function is evaluated as
 
0 i 0 i 0
Gdd;σ (t, t ) = − Θ(t − t ) < d−σ | exp − ( Ed + U ) ( t − t ) | d−σ >
h̄ h̄
 
i 0 i 0
= − Θ(t − t ) exp − ( Ed + U ) ( t − t ) (2297)
h̄ h̄

653
whereas if the state is projected onto the doubly occupied state | dσ d−σ >,
the Green’s function is expressed as
 
0 i 0 i 0
Gdd;σ (t, t ) = ( 1 − Θ(t − t ) ) < dσ d−σ | exp − ( Ed + U ) ( t − t ) | dσ d−σ >
h̄ h̄
 
i 0 i 0
= ( 1 − Θ(t − t ) ) exp − ( Ed + U ) ( t − t ) (2298)
h̄ h̄
Hence, on inserting the completeness relation for the projection operators in the
definition of the Green’s function and then evaluating the expectation value for
each of the terms, one finds that the time-dependent Green’s function can be
expressed as
  
0 i 0 i 0
Gdd;σ (t, t ) = − Θ(t − t ) (1 − nd,σ ) (1 − nd,−σ ) exp − Ed (t − t )
h̄ h̄
 
i 0
+ nd,−σ exp − (Ed + U ) (t − t )

  
i 0 i 0
+ ( 1 − Θ(t − t ) ) nd,σ (1 − nd,−σ ) exp − Ed (t − t )
h̄ h̄
 
i
+ nd,−σ exp − (Ed + U ) (t − t0 ) (2299)

Due to the time homogeneity of the ground state, the Green’s function only
depends on the time difference (t − t0 ). The Fourier transform of the Green’s
function is calculated from
Z ∞  
η
Gdd;σ (ω) = dt exp i ω ( t − t ) Gdd;σ (t − t0 ) exp[ − i | ( t − t0 ) | ]
0
−∞ h̄
(2300)
where a factor of exp[ − i h̄η | ( t − t0 ) | ] has been inserted to make the integral
convergent. Hence, one obtains the frequency-dependent Green’s function as
 
1 − nd,−σ nd,−σ
Gdd;σ (ω) = (1 − nd,σ ) +
h̄ ω − Ed − i η h̄ ω − Ed − U − i η
 
1 − nd,−σ nd,−σ
+ nd,σ +
h̄ ω − Ed + i η h̄ ω − Ed − U + i η
(2301)

The Green’s function can also be obtained by using the equations of motion.
It can be shown that

i h̄ Gdd:σ (t) = δ(t) < | [ dσ (t) d†σ (0) + d†σ (0) dσ (t) ] | >
∂t
i
− < | T̂ [ dσ (t) , Ĥ(t) ]− d†σ (0) | > (2302)

654
where the first term on the right-hand side originates from the explicit time-
dependence in the definition of Wick’s time-ordering operator. Since the expec-
tation value is to be evaluated at equal times, and the equal-time creation and
annihilation operators satisfy the anti-commutation relation

[ dσ , d†σ ]+ = 1 (2303)

one finds
∂ i
i h̄ Gdd:σ (t) = δ(t) − < | T̂ [ dσ (t) , Ĥ(t) ]− d†σ (0) | > (2304)
∂t h̄
The second term comes from the time-dependence of the annihilation operator
in the Heisenberg representation. If the atomic Hamiltonian is decomposed as

Ĥ = Ĥ0 + Ĥint (2305)

where Ĥ0 described the atomic binding energy


X
Ĥ0 = Ed d†σ dσ (2306)
σ

and Ĥint describes the Coulomb interaction between a pair of opposite spin
electrons
U X † †
Ĥint = dσ d−σ d−σ dσ (2307)
2 σ
Due to the commutation relation

[ dσ (t) , Ĥ0 (t) ]− = Ed dσ (t) (2308)

and
[ dσ (t) , Ĥint (t) ]− = U dσ (t) d†−σ (t) d−σ (t) (2309)
The equation of motion becomes
 

i h̄ − Ed Gdd;σ (t) = δ(t) + U Fdd;σ (t) (2310)
∂t

where Fdd;σ (t) is defined as


i
Fdd;σ (t) = − < | T̂ dσ (t) n̂d,−σ (t) d†σ (0) | > (2311)

The equation motion for the two-particle Green’s function Fdd;σ (t) is evaluated
as
 
∂ i
i h̄ − Ed Fdd;σ (t) = δ(t) nd,−σ (t) − U < | T̂ dσ (t) n̂d,−σ (t) n̂d,−σ (t) d†σ (0) | >
∂t h̄
(2312)
On using the identity
n̂d,−σ n̂d,−σ = n̂d,−σ (2313)

655
the equation reduces to
 

i h̄ − Ed Fdd;σ (t) = δ(t) nd,−σ (t) + U Fdd;σ (t) (2314)
∂t
On Fourier transforming the pair of equations of motion for Fdd;σ and Gdd;σ ,
one finds the pair of coupled algebraic equations

( h̄ ω − Ed ∓ i η ) Gdd;σ (ω) = 1 + U Fdd;σ (ω)


( h̄ ω − Ed − U ∓ i η ) Fdd;σ (ω) = nd,−σ (2315)

where the plus or minus signs, respectively, pertain to the cases where nd,σ is
either unity or zero. This set of algebraic equations is closed and can be solved
to yield
nd,−σ
Fdd;σ (ω) = (2316)
h̄ ω − Ed − U ∓ i η
and
1 U nd,−σ
Gdd;σ (ω) = +
h̄ ω − Ed ∓ i η (h̄ ω − Ed ∓ i η ) (h̄ ω − Ed − U ∓ i η )
1 − nd,−σ nd,−σ
= + (2317)
h̄ ω − Ed ∓ i η h̄ ω − Ed − U ∓ i η
Hence, the equations of motion yields recovers the explicit form of the solution
as
 
1 − nd,σ nd,−σ
Fdd;σ (ω) = nd,−σ +
h̄ ω − Ed − U − i η h̄ ω − Ed − U + i η
(2318)
and
 
1 − nd,−σ nd,−σ
Gdd;σ (ω) = (1 − nd,σ ) +
h̄ ω − Ed − i η h̄ ω − Ed − U − i η
 
1 − nd,−σ nd,−σ
+ nd,σ +
h̄ ω − Ed + i η h̄ ω − Ed − U + i η
(2319)

The poles of the Green’s function represent the energies required to either add
or remove an electron of spin σ from the d-state the system. The excitation
energy required to put an additional particle in the d shell is, therefore, either
Ed or Ed + U depending on whether the d-state of the impurity is initially
unoccupied or singly occupied.

19.4 The Schrieffer-Wolf Transformation


If the local magnetic impurity has a narrow width and is almost completely
occupied by one electron, then the Anderson Model can be mapped onto a

656
model of a localized magnetic moment by the Schrieffer-Wolf transformation247 .
The zero-th order Hamiltonian can be considered to be the terms in which the
hybridization is set to zero. Thus, for the present purposes one may write

Ĥ = Ĥ0 + ĤV (2320)

where Ĥ0 describes the ionic d states and the conduction band states.

Ek c†k,σ ck,σ +
X X
Ĥ0 = Ed d†σ dσ
k,σ σ

+ U d†↑ d†↓ d↓ d↑ (2321)

The Hamiltonian ĤV is the hybridization which couples the local and conduction
band states.
" 
V (k) c†k,σ dσ + V (k)∗ d†σ ck,σ
X
ĤV = (2322)
k,σ

The Schrieffer-Wolf transformation is based on a canonical transformation which


acts on the operators  and is of the form
   
Â0 = exp + Ŝ Â exp − Ŝ (2323)

where Ŝ is an anti-Hermitean operator. That is, the operator Ŝ satisfies

Ŝ † = − Ŝ (2324)

Thus, if the operator  is Hermitean then Â0 is also Hermitean. The canonical
transformation leads to the same expectation values if the states | Ψ > are
also transformed as
 
| Ψ0 > = exp + Ŝ | Ψ > (2325)

In particular the eigenvalues of Ĥ 0 and Ĥ are identical. The Schrieffer-Wolf


transformation S is chosen such that terms linear in the hybridization ĤV vanish
in the transformed Hamiltonian Ĥ 0 . This can only be achieved if Ŝ is assumed
to be of the same order as ĤV . In this case, the transformed Hamiltonian can
be expanded in powers of ĤV and Ŝ. On retaining the terms up to second-order,
one finds

Ĥ 0 = Ĥ0 + ĤV + [ Ŝ , Ĥ0 ]


1
+ [ Ŝ , ĤV ] + [ Ŝ , [ Ŝ , Ĥ0 ] ] + . . . (2326)
2!
247 J. R. Schrieffer and P. A. Wolff, Phys. Rev. 149, 491 (1966).

657
The operator Ŝ is chosen such that the terms linear in V (k) vanish. Hence, it
is required that Ŝ satisfies the linear equation

[ Ŝ , Ĥ0 ] = − ĤV (2327)

This is an operator equation, and Ŝ is determined if all its matrix elements are
known. This requires that a complete set of states be used. The simplest set of
complete sets correspond to the eigenstates of Ĥ0 , | φn > with eigenvalues En .
In this case, matrix elements are found as

< φm | ĤV | φn >


< φm | Ŝ | φn > = (2328)
Em − En

Thus, the operator Ŝ connects states which differ through the presence of an
additional conduction electron and a deficiency of an electron in the local orbital,
and vice versa. The energy denominators are of the form Ek − Ed or Ek −
Ed − U depending on the state of occupation of the local level. Thus, the anti-
Hermitean operator Ŝ can be expressed in terms of the four creation operators
for the localized level. The operator Ŝ is found as

c†k,σ dσ
"  
1 X
Ŝ = √ V (k) 1 − d†−σ d−σ
N Ek − Ed
k,σ

c†k,σ dσ
 
+ V (k) d−σ d†−σ
Ek − Ed − U
d†σ ck,σ
 
− V (k)∗ 1 − d†−σ d−σ
Ek − Ed
#
d†σ ck,σ

∗ †
− V (k) d−σ d−σ (2329)
Ek − E d − U

Having determined the operator Ŝ, the Hamiltonian to second-order in V is


given by
1
Ĥ 0 = Ĥ0 + [ Ŝ , ĤV ] + [ Ŝ , [ Ŝ , Ĥ0 ] ] + . . .
2!
1
= Ĥ0 + [ Ŝ , ĤV ] + . . . (2330)
2!

The transformed Hamiltonian Ĥ 0 contains an interaction term whereby the


conduction electrons are scattered from the different singly occupied states of
the d impurity. On expressing the conduction band factors in terms of the
matrix elements of the Pauli-spin matrices
1 X †
ŝα
k,k0 = ck,δ < δ | σ α | γ > ck0 ,γ (2331)
2
δ,γ

658
and likewise for the local operators
X †
Ŝ α = dδ < δ | S α | γ > dγ (2332)
δ,γ

one finds that in addition to a potential scattering term there is also an interac-
tion between the components of the spin density operators. Since the direction
of the axis of spin quantization can be chosen arbitrarily, the interaction must
be invariant under global rotations of the spin directions. The spin-flip contri-
bution of the interaction is of the form
"   
0 1 X 1 1
Ĥspin−f lip = −
2 0 Ek 0 − E d Ek 0 − Ed − U
k,k ,σ
 
0 ∗ † † ∗ 0 † †
× V (k) V (k ) ck0 ,−σ ck,σ dσ d−σ + V (k) V (k ) ck,σ ck ,−σ d−σ dσ
0

(2333)

To describe scattering of electrons close to the Fermi energy one may set Ek =
Ek0 , then the effective exchange interaction has the strength
 " #
0 ∗ 1 1
Jk,k0 = 2 <e V (k) V (k ) −
Ed − Ek U + Ed − Ek
(2334)

The total spin-dependent part of the interaction is recognized as just involving


the scalar product of the Fourier components of the two spin densities.
X
Ĥint = − Jk,k0 Ŝ . σ̂ k,k0 (0) (2335)
k,k0

For a singly occupied level, where Ed − µ is negative, the coefficient Jk,k0 also
has a negative sign if U is sufficiently large, so that the energy is lowered when-
ever the expectation values of both the spin density operators are anti-parallel.
Thus, classically, the energy is lowered whenever the polarization produced by
the conduction electron gas is anti-parallel to the spin of the local moment. This
type of coupling is known as an anti-ferromagnetic interaction. The alternative
type of coupling occurs when the sign of Jk,k0 is positive, and the ferromagnetic
interaction attempts to polarize the conduction electron spin density to be par-
allel to the local spin density.

19.4.1 The Kondo Hamiltonian


The resulting Hamiltonian is the Kondo Hamiltonian248 , it contains an inter-
action between the localized magnetic moment and the spins of the conduction
248 J. Kondo, Prog. Theor. Phys. 32, 37 (1964).

659
electrons. The Hamiltonian can be expressed as

Ĥ = Ĥ0 + Ĥint (2336)

where Ĥ0 represents the Hamiltonian for the conduction electrons

Ek c†k,σ ck,σ
X
Ĥ0 = (2337)
k,σ

and the interaction is given by

Ĥint = − J S . ŝ(0) (2338)

where S is a local moment and ŝ(0) is the spin of the conduction electrons at
the position of the impurity spin. The components of the conduction electron
spin is given in terms of matrix elements of the Pauli-spin matrices
1
c†k,δ < δ | σ α | γ > ck0 ,γ
X
ŝα (0) = (2339)
2N 0
k,k ;γ,δ

It is convenient to write the spin-dependent interaction in terms of the spin


raising and lowering operators for the local spin and the conduction electron
spin density

Ŝ ± = Ŝ x ± i Ŝ y
ŝ± = ŝx ± i ŝy (2340)

with the aid of the identity


1
Ŝ . ŝ = Ŝ z ŝz + ( Ŝ + ŝ− + Ŝ − ŝ+ ) (2341)
2
Hence, the interaction is written as

J 1 X
Ĥint = − S z ( c†k,↑ ck0 ,↑ − c†k,↓ ck0 ,↓ )
N 2
k,k0

+ † − †
+ S ck,↓ ck0 ,↑ + S ck,↑ ck0 ,↓ (2342)

where J is expected to be negative.

19.5 The Resistance Minimum


The Kondo effect249 results in a minimum in the resistivity of metals. The mini-
mum in the resistivity is due to the increasing T 5 resistivity caused by electron-
phonon scattering and a decreasing contribution from the impurity spin-flip
scattering, which in an intermediate temperature regime follows a ln T varia-
tion

ρ(T ) = ρ(0) + b T 5 + c ρ1 Jρ(µ) S ( S + 1 ) ln kB T ρ(µ) (2343)

660
Figure 268: The minimum in the resistivity of alloys containing F e impuri-
ties. The resistivities are normalized at T = 4.2 K. [After M.P. Sarachik et al.
(1964).].

k' k k'
k
k
k

S+ S- S- S+

Figure 269: The second-order contributions to the T-matrix for spin-flip scat-
tering of an up-spin electron (solid blue) by a localized spin (red). The dotted
blue line indicates an intermediate state with a hole in state k 0 .

where c is the concentration of impurities. Then, for negative J, the resistivity


shows a minimum with depth proportional to c which occurs at a concentration-
dependent temperature
  15
ρ1 |J| ρ(µ) S(S + 1) 1
Tmin = c5 (2344)
5b

in agreement with experimental findings250 .

The ln T term in the resistivity comes from scattering process to third-order


in J. This can be seen by considering the T-matrix for non spin-flip scattering
of an up-spin electron in second-order. The T-matrix will be evaluated on the
energy shell Ek = Ek0 , and E will be set to the ground state energy. To lowest
order, the non spin-flip scattering matrix elements are given by
J
< k 0 ↑ | T (1) (E + iη) | k ↑ > = − Sz (2345)
2N
whereas to second-order, one finds four non-zero contributions, two contribu-
tions from the spin-flip part ( Ŝ ± ) of the interactions and two contributions
249 J. Kondo, Prog. Theor. Phys. 32, 37 (1964).
250 A. D. Caplin and C. Rizzuto, Phys. Rev. Lett. 21, 746 (1968).

661
from the non spin-flip part ( Ŝ z ). The non spin-flip part gives rise to a term in
T (2) (E + iη) of
 2  z 2 X
J S
< k 0 ↑ | Tzz(2)
(E + iη) |k ↑ > = < k 0 ↑ | ( c†k ,↑ ck01 ,↑ − c†k ,↓ ck01 ,↓ )
N 2 1 1
k1 ,k2
1
× ( c†k ,↑ ck02 ,↑ − c†k ,↓ ck02 ,↓ ) | k ↑ >
E − Ĥ0 + iη 2 2

(2346)
As only the spin-up terms contribute to the scattering of the spin-up electron
the term simplifies to yield
 2  z 2 X
J S 1
= < k 0 ↑ | c†k ,↑ ck01 ,↑ c†k ,↑ ck02 ,↑ | k ↑ >
N 2 1
E − Ĥ0 + iη 2
k ,k1 2

(2347)
This has two contributions, one which corresponds to k 0 = k 1 and k = k 02 and
the other with k 0 = k 2 and k = k 01 . The sum of these terms are evaluated as
 2  z  2 X
J S 1 − f (Ek2 )
=
N 2 Ek − Ek2 + iη
k2
2  z
2 X
f (Ek1 )

J S

N 2 Ek1 − Ek + iη
k1
 2  z
2 X
J S 1
= (2348)
N 2 Ek − Ek1 + iη
k1

The singularity at Ek1 = Ek yields a finite result when integrated over k 1 .


Thus, there is no non-analytic behavior originating from the Ŝ z terms in the
interaction, which is just of the order J 2 ρ(µ) which is just a factor of J ρ(µ)
smaller than the leading contribution to the T-matrix.

The two spin-flip contributions to the T-matrix are given by


 2 X
(2) J 1
0
< k ↑ | T+− (E + iη) | k ↑ > = < k 0 ↑ | Ŝ + c†k ,↓ ck01 ,↑ Ŝ − c†k ,↑ ck02 ,↓ | k ↑ >
2N 1
E − Ĥ 0 + iη 2
k ,k 1 2

(2349)
and
 2 X
(2) J 1
0
< k ↑ | T−+ (E + iη) | k ↑ > = < k 0 ↑ | Ŝ − c†k ck01 ,↓ Ŝ + c†k ck02 ,↑ | k ↑ >
2N 1 ,↑ 2 ,↓
k1 ,k2
E − Ĥ0 + iη
(2350)

662
respectively. These terms are calculated to be
2 X
Ŝ + Ŝ −

(2) J
0
< k ↑ | T+− (E + iη) | k ↑ > = < k 0 ↑ | c†k ,↓ ck01 ,↑ c†k ,↑ ck02 ,↓ | k ↑ >
2N 1
E − Ĥ 0 + iη 2
k1 ,k2
2
f (Ek2 )

J X
= Ŝ + Ŝ − (2351)
2N Ek − Ek2 + iη
k2

and
2 X
Ŝ − Ŝ +

(2) J
0
< k ↑ | T−+ (E + iη) | k ↑ > = < k 0 ↑ | c†k ,↑ ck1 ,↓
0 c†k ,↓ ck2 ,↑ |
0 k↑>
2N 1
E − Ĥ0 + iη 2
k1 ,k2
2
1 − f (Ek1 )

J X
= Ŝ − Ŝ + (2352)
2N Ek − Ek1 + iη
k1

In this case, the two terms cannot be combined to give a result independent of
the Fermi function, as Ŝ + and Ŝ − do not commute. In this case, one can use
the identities
Ŝ + Ŝ − = S ( S + 1 ) − ( Ŝ z )2 + Ŝ z

Ŝ − Ŝ + = S ( S + 1 ) − ( Ŝ z )2 − Ŝ z (2353)
The terms proportional to S ( S + 1 ) − ( Ŝ z )2 combine to yield an analytic
contribution to the T-matrix of
 2
(2) J X 1
< k 0 ↑ | Tsf (E + iη) | k ↑ > = ( S ( S + 1 ) − ( Ŝ z )2 )
2N Ek − Ek1 + iη
k1

(2354)
whereas the remaining contribution is proportional to S z and the integration is
divergent at Ek = Ek1 but the integration is cut off by the Fermi function.
 2 X 2 f (Ek ) − 1
0 (2) J
< k ↑ | Tsf (E + iη) | k ↑ > = Ŝ z 1

2N Ek − Ek1 + iη
k1

(2355)
At finite temperatures, either the Fermi function provides a cut-off for the sin-
gularity when the scattered particle is on the Fermi surface Ek = µ, or if the
scattered particle is off the Fermi surface, the excitation energy acts as a cut off.
In the latter case, the second-order contribution to the real part of the T-matrix
can be evaluated as
 2
(2) 2 J
< k 0 ↑ | Tsf (E + iη) | k ↑ > ∼ − Ŝ z ρ(µ) ln ( Ek − µ ) ρ(µ)

N 2
(2356)

663
Figure 270: The impurity (F e) contribution to the electrical resistivity of CuF e
alloys as a function of ln T . [After M.D. Daybell and W.A. Steyert (1967).].

which is divergent when Ek approaches µ. Thus, this second-order term can be


as large as the first-order term which is also proportional to Sˆz . The scattering
rate which enters into the resistivity is proportional to the thermal average of
the square of the T-matrix, and involves all possible final spin states of the
scattered electron (i.e. k 0 , ↑ and k 0 , ↓). On noting that for non-polarized spins
1
< ( Ŝ z )2 > = S(S + 1) (2357)
3
one finds that the rate of scattering from magnetic impurities with concentration
c is given by
 2  
1 2π J S(S + 1)
= 3 c ρ(µ) 1 + 2 J ρ(µ) ln kB T ρ(µ) + . . .

τ h̄ 2 3
(2358)

The factor of three in front occurs since the spin-flip scattering rate (k, ↑) →
(k 0 , ↓) is twice as large as the non-spin-flip scattering rate (k, ↑) → (k 0 , ↑). The
above scattering rate gives rise to a logarithmically increasing resistivity for
magnetic impurities in simple metals. Since the logarithmic divergence is caused
by spin-flip scattering in the intermediate states, the application of a field should
suppress the Kondo effect. The resistivity does not diverge at T = 0 and, hence,
the exact T-matrix also does not diverge. The logarithmic dependence found in
perturbation theory saturates when all the logarithmically divergent scattering
processes are taken into account. The leading-order logarithmic coefficient of
each term in the perturbation expansion series (in powers of J ρ(µ)) can be
calculated by various means251 . In the ferromagnetic case, where J > 0, the
saturation occurs at a characteristic Kondo energy or Kondo temperature TK
given by


J ρ(µ) ln kB TK ρ(µ) = − 1
(2359)

251 A. A. Abrikosov, Physics 2, 5 (1965).

664
Fixed Point J+2- Jz2 = Const
1

0.75

J+ ρ(µ)
0.5

0.25

0
-1 -0.5 0 0.5 1
AF Jz ρ(µ) F

Figure 271: The renormalization flow of the anisotropic Kondo Hamiltonian


as calculated with a cut-off renormalization scheme. The system is shown to
either scale to ferromagnetic Ising fixed points (F) or to the strong-coupling
antiferromagnetic fixed point. [After Anderson (1970).]

or  
1
kB TK = ρ(µ)−1 exp − (2360)
|J| ρ(µ)
and all the results are finite. The above formula also defines the Kondo tem-
perature when the local exchange interaction J is antiferromagnetic.

For the case of anti-ferromagnetic coupling, the physics scales to a strong


coupling fixed point252 so the solution must be obtained by other means such
as Bethe-Ansatz253 . The properties of the anti-ferromagnetic solution include
the cross-over from a high temperature (T > TK ) Curie susceptibility for the
free impurity moments to a Pauli paramagnetic susceptibility for T < TK .
Also, the specific heat originating from the impurity changes from a constant
value at high temperatures to a low-temperature form having a linear T de-
pendence. This indicates that the magnetic moments of the impurity are being
removed and that at low temperatures, the properties are those of a narrow
virtual bound state of width kB TK located near the Fermi energy254 . In fact
a variational analysis shows255 that the magnetic moments are being screened
by a compensating polarization of conduction electrons, and that the cloud and
moment form a singlet bound state of binding energy kB TK . For T < TK
the conduction electrons occupy the bound state and the moment is screened,
for T > TK the bound state is thermally depopulated and the system ex-
hibits properties of the free moments. From the perspective of the Anderson
252 P. W. Anderson, J. Phys. C, 3, 2436 (1970).
253 N. Andrei, K. Furuya and J. H. Lowenstein, Rev. Mod. Phys. 5, 331 (1983), A. M.
Tsvelik and P. B. Weigmann, Adv. in Phys. 32, 453 (1983), P. Schlottmann, Phys. Reps.
181, 1 (1989).
254 M. A. Daybell and W. A. Steyert, Rev. Mod. Phys. 40, 380 (1968), D. K. Wohlleben

and B. R. Coles, in Magnetism, eds. G. T. Rado and H. Suhl, Academic Press, N.Y. 1973.
255 K. Yosida, Phys. Rev. 147, 223 (1966).

665
0.2

0.15

kBT χ/(gμB)2
0.1

0.05

0
0 5 10 15 20 25
T/TK

Figure 272: The universal temperature dependence of the effective local mag-
netic moment squared (blue) and kB TK times the reduced impurity suscepti-
bility (red) for the S = 12 Kondo Model. The Kondo effect quenches the local
moments at low temperatures.

T = 0 Density of States
20
ρ(E) [ States / Energy ]

15

kBTK
10

5

Ν∆
0
0 0.5 E1f µ
1.5 2 Ef +
2.5 U 3
E [ Energy ]

Figure 273: The schematic (T = 0) electronic density of states for an N -fold


spin and orbitally degenerate Anderson impurity model, as appropriate for Ce
impurities. The Coulomb interaction splits the localized density of states ρ(E)
into a singly-occupied level at Ef and an unoccupied level at Ef +U . In addition,
there is a narrow Abrikosov-Suhl resonance just above the Fermi-energy µ.

666
impurity model, the density of states that is found at high temperatures follows
directly from Anderson’s picture of a spin-split virtual bound state. However,
as T decreases below TK , the density of states shows a sharp peak of width
kB TK growing in the vicinity of µ. In the low-temperature limit, the height of
the Abrikosov-Suhl256 peak saturates on the order of ( kB TK )−1 . Thus, the
low-temperature properties can be directly understood in terms of the virtual
−1
bound state with a density of states which is very large ∝ TK . The properties
of this low-temperature Fermi liquid were established by Nozières257 .

256 A. A. Abrikosov, Physics 2, 5 (1965), H. Suhl, Physics 2, 39 (1965).


257 P. Nozières, Ann. Phys., 10, 19 (1985).

667
20 Collective Phenomenon
21 Itinerant Magnetism
21.1 Stoner Theory
The Stoner theory of itinerant magnetism258 examines the stability of a band
of electrons to Coulomb interactions. The Hamiltonian is expressed as the sum
of two terms, Ĥ0 the non-interacting electrons in the Bloch states and Ĥint
describing the Coulomb repulsion between the electrons
Ĥ = Ĥ0 + Ĥint (2361)
The Hamiltonian for the non-interacting electrons in the Bloch states is written
as X
Ĥ0 = Ek nk,σ (2362)
k,σ

The interaction Hamiltonian is given by


U X
Ĥint = n̂i,σ n̂i,−σ (2363)
2 i,σ

where U represents the short-ranged Coulomb interaction between a pair of


electrons occupying the orbitals on the i-th lattice site. The operator n̂i,σ cor-
responds to the number of electrons of spin σ which occupy the i-th lattice site.
It is assumed that the band is non-degenerate, therefore, there is only one or-
bital per lattice site which due to the limitations imposed by the Pauli exclusion
principle can only hold a maximum of two electrons.

The interaction is treated in the mean-field approximation. First it shall be


assumed that translational invariance holds, so that the orbitals in each unit
cell have the same occupation numbers. Also the Hamiltonian is expanded in
powers of the fluctuation operator ∆n̂i,σ = n̂i,σ − nσ so that
 
U X
Ĥint = ∆n̂i,σ ∆n̂i,−σ + nσ ∆n̂i,−σ + n−σ ∆n̂i,σ + n−σ nσ (2364)
2 i,σ

and then the second-order fluctuations are ignored. This leads to the interaction
energy being approximated in terms of single-particle operators
 
U X
Ĥint ≈ n̂i,σ n−σ + n̂i,−σ nσ − n−σ nσ
2 i,σ
 
U X
= n̂k,σ n−σ + n̂k,−σ nσ − n−σ nσ
2
k,σ

(2365)
258 E. C. Stoner, Rep. Prog. in Phys. 11, 43 (1948).

668
Thus, in the mean-field approximation, the Hamiltonian is given by
X U X
ĤM F = ( Ek + U n−σ ) nk,σ − N n−σ nσ (2366)
2 σ
k,σ

The single particles have the spin-dependent energy eigenvalues

Eσ (k) = Ek + U n−σ (2367)

The magnetization is given by


 
1
Mz = g µB n↑ − n↓
2
Z ∞  
= µB d f () ρ( − U n↓ ) − ρ( − U n↑ )
−∞
(2368)

This equation has non-magnetic solutions with n↑ = n↓ and may have fer-
romagnetic solutions in which the number of up-spin electrons is greater than
the number of down spin electrons n↑ 6= n↓ . In the ferromagnetic state, the
Stoner model predicts that the up-spin sub-bands are rigidly shifted relatively
to the down-spin bands by the exchange splitting ∆ which has a magnitude of
U ( n↑ − n↓ ). On increasing U from zero, the ferromagnetic solutions first

1.5

1
ρ↑(ε)
0.5

εF
0
-4 -2 0 2 4 6
ε
-0.5

-1
∆ −ρ↓(ε)
-1.5

Figure 274: A schematic depiction of the spin sub-band density of states, ex-
pected for a ferromagnetic solution of the Stoner model.

become stable at T = 0 when Mz ∼ 0. For an infinitesimal magnetization, the


equation can be linearized to yield
  Z ∞

( n↑ − n↓ ) = U n↑ − n↓ d f () ρ() (2369)
−∞ ∂

669
since the higher-order terms are negligible. On canceling a factor of M z from
both sides of the equation, one finds a self-consistency equation which is satisfied
for infinitesimal magnetizations. The ferromagnetic state has the lowest energy
when the self-consistency equation is satisfied. The integral in the resulting
self-consistency equation can be performed via integration by parts yielding
Z ∞

1 = U d f () ρ()
−∞ ∂
Z ∞

= −U d ρ() f () (2370)
−∞ ∂

At low temperatures, the derivative of the Fermi function can be replaced by a


delta function at the Fermi energy.

− f () = δ( − µ) (2371)
∂
This yields the Stoner criterion for ferromagnetism as

1 < U ρ(µ) (2372)

where ρ(µ) is the density of states per spin at the Fermi energy.

If the Stoner criterion is satisfied the paramagnetic state is unstable to the


ferromagnetic state, and a spontaneous magnetic moment Mz occurs in the tem-
perature range 0 ≤ T ≤ Tc . The magnetization is given by the solution of the
non-linear equation, eqn(2368). The non-linear equation shows that the magne-
tization increases with increasing U , and saturates to a value which is one Bohr
magneton per electron, for low density materials which the bands have a filling
of less than one electron per atom. In systems which have bands that are more

0.8

0.6 Uρ(μ) > 1

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
M/μB

Figure 275: A graphical solution of the non-linear equation for the magnetiza-
tion M , when the Stoner criterion U ρ(µ) > 1 is satisfied.

than half-filled, the saturation magnetic moment is equal to a Bohr magneton

670
per unoccupied state. At finite temperatures, the value of the magnetization
is reduced and disappears at a critical temperature Tc . Unfortunately, Stoner
theory does not predict reasonable values for the critical temperatures.

In the paramagnetic state, Stoner theory predicts that the (T = 0) suscep-


tibility should be exchange enhanced over the non-interacting susceptibility χ0p
via
χ0p
χp = (2373)
1 − U ρ(µ)
For systems which are close to the ferromagnetic instability, the susceptibility
should take on large values. This is the case for P d in which the d band is
almost completely occupied.

——————————————————————————————————

21.1.1 Exercise 90
Determine the critical temperature Tc predicted by Stoner theory.

——————————————————————————————————

21.1.2 Exercise 91
Determine the paramagnetic susceptibility by using Stoner theory.

——————————————————————————————————

21.2 Linear Response Theory


The spatially varying magnetization Mz (r) of a paramagnetic system produced
by a spatially varying applied magnetic field Hz (r) can be expressed in terms
of the z − z component of the magnetic susceptibility tensor, through the linear
relationship Z
Mz (r) = d3 r0 χz,z (r, r0 ) Hz (r0 ) (2374)

This is a special case of the more general relation


Z X
Mα (r) = d3 r 0 χα,β (r, r0 ) Hβ (r0 ) (2375)
β

For translational invariant systems, the expression for the response function is
only a function of the difference r − r0 . Also for non-magnetic systems, that

671
possess spin rotational invariance, the susceptibility tensor is diagonal and the
diagonal components are related via
χx,x (r − r0 ) = χy,y (r − r0 ) = χz,z (r − r0 ) (2376)
The relation between the magnetic response and the applied field becomes sim-
pler, after Fourier transforming. The Fourier transform of the magnetization is
defined as Z  
M (q) = d3 r exp − iq.r M (r) (2377)

The Fourier transform of the magnetization is related to the Fourier transform


of the applied field via
X
Mα (q) = χα,β (q) Hβ (q) (2378)
β

The response function can be evaluated from perturbation theory, in which the
Zeeman interaction
Z
ĤZeeman = − d3 r0 M (r0 ) . H(r0 )
Z
= − d3 q M (q) . H(q) (2379)

is treated as a small perturbation.

For convenience, χz,z (q) shall be calculated by reducing it to a previously


known case. The change in density of electrons of spin σ, with Fourier com-
ponent q, produced in response to an applied spin-dependent potential. The
Fourier component of the potential is given by
gµB z
Vσ (q) = − H (q) σ + U ρ−σ (q) (2380)
2
Thus, the charge density is given by the two coupled equations
 
0 g µB
ρσ (q) = χ (q) − Hz σ + U ρ−σ (q) (2381)
2
one for each spin polarization. In the above expression, χ0 (q) is the Lind-
hard density-density response function, per spin. The z-component of the q-
dependent magnetization is defimed as
  
g µB
Mz (q) = ρ↑ (q) − ρ↓ (q) (2382)
2
On combining these equations, one finds that the z-component of the q-dependent
magnetization produced by a magnetic field applied along the z-direction is given
by
2 Mz (q)
 
g µB 0
Mz (q) = − χ (q) g µB Hz (q) + U (2383)
2 g muB

672
Thus, it is found that the Pauli paramagnetic susceptibility is given by
Mz (q)
χz,z
p (q) =
Hz (q)
g 2 µ2B 2 χ0 (q)
= − (2384)
4 1 + U χ0 (q)

It is usual to use re-write this expression in terms of the reduced non-interacting


magnetic susceptibility defined by
χz,z
0 (q)
ez,z
χ0 (q) = (2385)
g 2 µ2B
which results in
1 0
ez,z
χ0 (q) = − χ (q) (2386)
2
instead of the density-density response function χ0 (q). This yields the result

ez,z
2
4χ0 (q)

g µB
χz,z
p (q) = (2387)
2 1 − 2U χ ez,z
0 (q)

Since the reduced non-interacting magnetic susceptibility is positive, and U is


positive, the paramagnetic susceptibility is enhanced for sufficiently small values
of U .

21.3 Magnetic Instabilities


The reduced non-interacting susceptibility, χ ez,z
0 (q), may have maxima at cer-
tain values of q, say Q, which are determined by the band structure and the
occupancy of the non-interacting bands. If the values of the non-interacting
susceptibility at these maxima are finite, then the denominator of the Pauli-
paramagnetic susceptibility may become small at these q values, for sufficiently
small values of U . This has the effect that, for small U , the Pauli-paramagnetic
susceptibility is enhanced at these Q values. If U is increased further, there
will be a critical value of U , Uc at which point the denominator will fall to
zero and the susceptibility at Q will become infinite. The divergence of the
susceptibility at Q indicates that an infinitesimal applied field can produce a
finite staggered magnetization M z (Q). Although this argument assumed that
the z-axis was the axis of quantization, the argument could have been made if
the quantization axis was assumed to be along any arbitrarily chosen direction.
The infinitesimal field may be produced by a spontaneously statistical fluctua-
tion, and have an arbitrary direction. This field will force the system to order
magnetically by having a finite M (Q) in the spontaneously chosen direction.
The system, by spontaneously choosing a direction for the magnetization, has
spontaneously broken the symmetry of the Hamiltonian.

673
The critical value of U , above which the paramagnetic state becomes unsta-
ble to a state with a modulated spin density M (Q), is given by

ez,z
1 = 2 Uc χ0 (Q) (2388)

If a non-interacting system is considered which has a maximum in χ ez,z


0 (Q) at
Q = 0 the above expression reduces to the Stoner criterion for ferromagnetism
as
1
lim χ ez,z
0 (Q) → ρ(µ) (2389)
Q → 0 2
where ρ(µ) is the density of states, per spin, at the Fermi energy. Thus, when
Q = 0, it is found the critical value of U is given by the criterion

1 = Uc ρ(µ) (2390)

and for values of U larger than the critical value the paramagnetic state is un-
stable to the formation of a ferromagnetic state.

For values of U greater than Uc , the mean-field analysis has to be modified


to include the effect of the spontaneous magnetization. For a ferromagnet, the
interaction produces a rigid splitting between the up-spin bands and down-spin
bands by an amount ∆ = U ( n↑ − n↓ ) called the exchange splitting. For
isotropic systems, the magnetic response will crucially depend on the direction
of the applied field compared to that of the spontaneous magnetization. For
a ferromagnet, the longitudinal response (produced by a field which is parallel
to the spontaneous magnetization M ) will be finite, as this corresponds to pro-
cesses which excites the system as it stretches the magnitude of M . However,
the transverse response will be infinite as this corresponds to applying a field
that will rotate the direction of the spontaneous magnetization until it aligns
with the applied field. As the system is isotropic, this global spin rotation can
be achieved without requiring any finite energy excitations. The zero-energy
excitations that uniformly rotate the magnetization in a ferromagnet are the
q = 0 Goldstone modes associated with the spontaneously broken continuous
spin rotational invariance of the Hamiltonian.

In three-dimensional systems with almost spherical Fermi surfaces, the in-


stability can only occur at 2 kF . This can lead to a spin density wave which
has a periodicity which is incommensurate with the underlying lattice. In low-
dimensional systems, such as two-dimensional and one-dimensional organic ma-
terials, there can be large sheets of the Fermi surface which can produce a large
non-interacting susceptibility at the Q value connecting these sheets. The large
response of the system to the interaction can produce a spin density wave in
which the magnetization is modulated with this wave vector.

For tight-binding bands which satisfy the perfect nesting condition

Ek+Q = − Ek (2391)

674
1

0.5
Q

kya/π -1 -2 -3
0

-0.5

3 2 1 0
-1
-1 -0.5 0 0.5 1
kxa/π

Figure 276: The nesting vector Q = 2π 1 1


a ( 2 , 2 .0) for a perfectly nested two-
dimensional tight-binding model. The constant energy surfaces are depicted as
contours.

for some Q, the non-interacting susceptibility can be evaluated as an integral


over the density of states
1 X f (−Ek ) − f (Ek )
ez,z
χ0 (Q) =
2 2 Ek
k
Z ∞
1 f (−) − f ()
= d ρ()
2 −∞ 2
Z ∞
1 f ()
= − d ρ() (2392)
2 −∞ 
sine ρ(−) = ρ(). From this one can see that if the density of states ρ(0) is
ez,z
non-zero, the susceptibility χ 0 (Q) will diverge logarithmically or faster than
logarithmically when µ → 0. The divergence occurs when two large portions of
the Fermi surface are connected by the wave vector Q, which allows the system to
rearrange the electrons at the Fermi surface by zero-energy excitations involving
a momentum change Q. Thus, in this case, there is no energy penalty to be
incurred in producing a spin density wave M (Q). The perfect nesting condition
occurs at Q = πa (1, 1, 1) for non-degenerate tight-binding bands with a simple
cubic lattice, where
i=3
X
Ek = − 2 t cos ki a (2393)
i=1
Since the bands are symmetric around  = 0, the non-interacting susceptibil-

675
0.50

0.40

0.30

0.20 U-1

0.10

0.00
-6 -3 0 3 6

µ/t

Figure 277: The condition that a paramagnet becomes unstable relative to a


Neel state with ordering vector 2π 1 1 1
a ( 2 , 2 , 2 ), for three-dimensional tight-binding
band, is satisfied for µ values close to half-filling.

ity diverges for half-filled bands. In this case, the critical value of U is zero.
Hence, the paramagnetic state will become unstable to a state in which the
magnetization exhibits spatial oscillations with wave vector Q, even with an
infinitesimally small value of U . In real space, the staggered magnetization of
this ordered state is given by
X ri
M (r) = M Q cos π (2394)
i
a

The magnetic moments on the nearest-neighboring lattice sites are antiparal-


lel, which constitutes an anti-ferromagnetic ordering. Since anti-ferromagnetic
ordering was first proposed by Louis Neel to describe classical magnets259 , this
type of ordering is known as Neel ordering. Unfortunately, the Neel state is not
an exact ground state for a quantum system.

The occurrence of an anti-ferromagnetically ordered state may be accompa-


nied by a metal-insulator transition. This process was first discussed by J. C.
Slater260 . Physically, the appearance of anti-ferromagnetic order could result in
a doubling of the size of the real space unit cell. The electrons of spin σ traveling
in the solid experience a periodic potential which contains a contribution due to
the interaction with electrons of opposite spin. The doubling of the size of the
real space unit cell, produced by the magnetic order, results in the volume of the
Brillouin zone being halved. The new periodicity caused by the sub-lattice mag-
netization shows up as a spin-dependent contribution to the potential, and may
produce gaps (∆ ∝ U MQ ) in the electronic dispersion relations at the surface
259 L. Neel, Ann. de Physique, 17, 64 (1932), Ann. de Physique 5, 256 (1936).
260 J. C. Slater, Physical Review, 82, 538, (1951).

676
2.5

1.5
Ek+

0.5

Ekα/t
2∆
-0.5

Ek-
-1.5

-2.5
-1 -0.5 0 0.5 1

ka/π

Figure 278: A schematic description of the mean-field electronic bands for a Neel
antiferromagnet. The dashed lines indicates the electronic bands of the high-
temperature paramagnetic state. In the Neel state, the size of the Brillouin zone
is halved, and an energy gap of 2 ∆ appears at the boundary of the magnetic
Brillouin zone.

of the new Brillouin zone. If the magnitude of the spin-dependent potential is


large enough, a gap may occur all around the Brillouin zone resulting in a gap
in the density of states. If the Fermi energy lies, in the gap, as is expected for a
half-filled band, the state will be insulating. Such insulating anti-ferromagnetic
states occur in undoped La2 CuO4 , which is the parent material of some high-
temperature superconductors. Although the insulating anti-ferromagnetic state
does not have low-energy electronic excitations, it does have low-energy spin
excitations in the form of Goldstone modes. These are spin waves, which have
the dispersion relation ω = c q.

21.4 Spin Fluctuations near Ferromagnetic Instabilities


The dynamical magnetic response of a paramagnetic system to a time and spa-
tially varying applied magnetic field of wave-vector q and frequency ω is given
by the dynamical response χz,zp (q; ω). The imaginary part of this response func-
tion yields the spectrum of magnetic excitations. The imaginary part of the
reduced susceptibility can be measured directly by inelastic neutron scattering
experiments, in which the neutron’s spin interacts with the electronic spin den-
sity via a dipole-dipole interaction. A simple extension of our previous analysis
shows that, in the mean-field approximation, the response for a paramagnetic
material is given by

ez,z
2
4χ0 (q, ω)

g µB
χz,z
p (q; ω) = (2395)
2 1 − 2U χ ez,z
0 (q; ω)

677
Let us examine the imaginary part of the response function for a paramagnetic
metal, such as P d, which is on the verge of an instability to a ferromagnetic
state. Then,
ez,z
2
4 =m χ 0 (q, ω)
  
z,z g µB
=m χp (q; ω) = 2 2
2
 
z,z z,z
1 − 2 U <e χ e0 (q; ω) + 2 U =m χ e0 (q; ω)
(2396)
which on using the approximation to the Lindhard susceptibility,
1
ez,z
<e χ0 (q; ω) ≈ ρ(µ)
2
π ω
ez,z
=m χ0 (q; ω) ≈ ρ(µ) (2397)
8 q vF
shows that the system exhibits a continuum of quasi-elastic magnetic excita-
tions. As the value of U is increased so that a ferromagnetic instability is
approached, the spectrum is enhanced at low frequencies. These magnetic exci-

5
Im χ(q;ω) [units of εF]

4
q/kF=0.7
3
U/Uc
0.95
2
0.9
0.85
0.7
1 0.0

0
0 0.5 1 1.5 2
ω/εF

Figure 279: The spectrum of magnetic excitations with momentum transfer


q/kF = 0.7 for a paramagnet for various values of the Coulomb interaction U .

tations are known as paramagnons. The lifetime of the paramagnon excitations


and the frequency of the excitations soften as the value of U is increased towards
the critical value Uc . Basically, this represents a slowing down of the rate at
which a small region of ferromagnetically aligned spins relax back to the equi-
librium (paramagnetic) state. The existence of large amplitude paramagnon
fluctuations not only manifest themselves in the inelastic neutron scattering
cross-section (which is directly proportional to =m [ χ eα,β (q; ω) ]), and an en-
hanced susceptibility but also leads to a logarithmic enhancement of the linear

678
2
0.5 0.25

1.5

1 1.0
ω/εF

0.5 1.5
U=0.8Uc
2.0
0
0 0.5 1 1.5 2 2.5 3
q/kF
-0.5

-1

Figure 280: The phase space of magnetic excitations for a paramagnet with
the Coulomb interaction U = 0.8 U c. The values of Ime χ(q; ω), in units of
F , are depicted as contours. The paramagnon-excitations are revealed as a
broadened low-energy resonance with small momentum transfers. Due to the
Pauli exclusion principle, only excitations with ω > 0 are allowed.

T term in the electronic specific heat261 , and an enhancement in the T 2 term in


the electrical resistivity262 . These characteristics have been observed in metallic
P d 263 .

The above mean-field type of analysis has shown that close to a magnetic
instability, there will be large-amplitude Gaussian fluctuations. This continuous
spectrum of excitations is expected to soften as the instability is approached,
261 N. F. Berk and J. R. Schrieffer, Phys. Rev. Lett. 17, 433 (1966), S. Doniach and S.

Engelsberg, Phys. Rev. Lett. 17, 750 (1966).


262 P. Lederer and D. L. Mills, Phys. Rev. Lett. 20, 1036 (1968).
263 A. I. Schindler and B. R. Coles, J. Appl. Phys. 39, 956 (1968).

Figure 281: The magnetic inelastic neutron scattering cross-section for the en-
hanced paramagnet N i3 Ga. [After B. R. Bernhoeft, S. M. Hayden, G. G. Lon-
zarich, D. McK. Paul and E. J. Lindley, Phys. Rev. Lett. 62, 657 (1989).]

679
say by increasing U . The fluctuations are expected to be long-ranged and long-
lived. However, the above mean-field analysis is expected to fail close to the
transition, where critical fluctuations should be taken into account. Unlike most
other phase transitions, the phase transition that has just been described first
occurs at T = 0. The critical fluctuations are not thermally excited and can-
not be treated classically but are zero point fluctuations associated with the
existence of a quantum critical point264 .

21.4.1 Ferromagnetic Spin Waves


For values of U greater than Uc where the system is ferromagnetic, the q = 0
transverse response shows a sharp zero-energy mode that represents the Gold-
stone mode of the system. In the ferromagnetically ordered state, these excita-
tions form a sharp (delta function-like) branch of spin waves which stretch up
from ω = 0 at q = 0. The transverse response functions are equal

χx,x (q; ω) = χy,y (q; ω) (2398)

but, in the ferromagnetic state, differ from the longitudinal response

χx,x (q; ω) 6= χz,z (q; ω) (2399)

The transverse response can be expressed in terms of the spin-flip response


function involving the spin raising and lowering operators

M̂ ± (q) = M̂ x (q) ± i M̂ y (q) (2400)

The spin-flip response functions are

χ+,− (q; ω) = χ−,+ (q; ω) = 2 χx,x (q; ω) (2401)

In the random phase approximation265 , the reduced transverse dynamic suscep-


tibility has the form

e+−
χ0 (q; ω)
e+− (q; ω) =
χ (2402)
e+−
1 − U χ0 (q; ω)

where χe+−
0 (q; ω) is the mean-field transverse susceptibility. In the ferromagnetic
state, the mean-field susceptibility is given by
fk↑ − fk+q↓
 
1 X
e+−
χ0 (q; ω) = (2403)
N ∆ + εk+q − εk − h̄ ω
k

which involves the spin-split sub-bands. The spectra of spin-flip excitations is


proportional to the imaginary part of the transverse susceptibility. The spectra
264 J. A. Hertz, Phys. Rev. B, 14, 1165 (1976).
265 T. Izuyama, D.J. Kim and R. Kubo, J. Phys. Soc. Jpn. 18, 1025 (1963).

680
10
M = 1/4

2
2m a Eσ(k) / ћ
kF↓
0
-4 -3 -2 -1 0 1 2 3 kF↑ 4

2
-5
E↓(k)

-10

E↑(k)

-15

ka

Figure 282: A schematic description of the mean-field electronic bands for


a ferromagnet. The dashed lines indicates the electronic bands of the high-
temperature paramagnetic state. In the ferromagnetic state, the up-spin sub-
band is lower than the down-spin sub-band by an energy ∆.

for spin-flip processes, which decreases the z-component of the magnetization


by h̄/N , is given by

e+−
Im χ0 (q; ω)
 
Im e+− (q; ω)
χ =
[ 1 − U Re χ e+− 2
0 (q; ω) ] + [ U Im χ e+−
0 (q; ω) ]
2

(2404)
The spectra is composed of a continuum of slightly modified Stoner excitations,
which exist throughout the (q, ω) region where

e+−
Im χ0 (q; ω) 6= 0 (2405)

For small q, this region has the form of a distorted triangle with an apex at
ν = ∆ at q = 0, as seen in Fig(283). For free electrons, the energy of the upper
boundary of the Stoner continuum is given by

h̄2
h̄ω = (k 2 − kF2 ↓ + 2kF ↑ q + q 2 ) (2406)
2m F ↑
The energy of the lower boundary of the Stoner continuum is given by

h̄2
h̄ω = (k 2 − kF2 ↓ − 2kF ↑ q + q 2 ) (2407)
2m F ↑
For weak ferromagnets (for which M/µB is small), the Stoner continuum first
reaches ω = 0 at the wave vector qc = kF ↑ −kF ↓ which connects the up-spin and
down-spin Fermi surfaces. Nevertheless, the imaginary part of the susceptibility
is non-zero for smaller q values at energies which are below the energy of the
Stoner continuum even though Ime χ+−
0 (q : ω) = 0. The ω values of these small
q excitations are given by the solutions of the equation

e+−
1 − U Re χ0 (q; ω) = 0 (2408)

681
2
0.5
0.25

1
1
∆/εF 2
M=0.5
4
ω/εF

Spin Waves 8
0
0 0.5 1 1.5 2 2.5 3

q/kF
-1

Figure 283: The phase-space for spin-flip excitations. Since the mean-field
state is stable against single-particle excitations, the allowed spin-flip (Stoner)
excitations must have ω > 0. The upper and lower boundaries of the Stoner
continuum are depicted by the blue lines. The contours of the imaginary part
of the transverse susceptibility are depicted as black lines. The branch of spin
wave excitations is marked by a red line.

The spectrum of magnetic excitations contains a sharp delta function located at


these points in (q, ω) space. The solutions of eqn.(2408) describes the dispersion
relation of a branch of collective modes. The above equation is guaranteed to
have a solution at ω = 0 and q = 0, since for q = 0 the above equation reduces
to

1 e+−
= U Re χ0 (0, ω)
 
U X fk↑ − fk↓
=
N ∆ − h̄ω
k
( n↑ − n↓ )
= U (2409)
∆ − h̄ ω
which at ω = 0 reduces to
U ( n↑ − n↓ )
1 = (2410)

The above condition is identified with the self-consistency condition for the ex-
change splitting in Stoner-Wohlfarth theory. The zero-energy collective magnetization-
lowering excitations are Goldstone modes266 since, when they are combined co-
herently, they produce a uniform rotation of the magnetization, and, thereby,
restore the spontaneously broken symmetry. By expanding eqn.(2408) for small
266 J. Goldstone, Nuovo Cimento, 19, 154 (1961).

682
Figure 284: The spin wave dispersion relation inferred from inelastic neutron
scattering experiments on an f.c.c. Cobolt alloy. [After R.N. Sinclair and B.N.
Brockhouse, Phys. Rev. 120, 1638 (1960).]

(q, ω), the collective spin-wave excitations are found to obey the dispersion re-
lation
h̄ ω = D q 2 (2411)
where the spin-wave stiffness constant D is given by the expression,
    
U X fk↑ + fk↓ fk↑ − fk↓
D = ∇2 εk − |∇εk |2 (2412)
3N ∆ 2 ∆
k

This dispersion relation differs from the dispersion relation of most other Gold-
stone modes, which are usually linear in q, since for a ferromagnet the order
parameter is a conserved quantity. At q = 0, it is easy to see that the spectra
from the Goldstone modes consists of a delta function located at ω = 0 with a
strength governed by the value of the magnetization. From the Kramers-Kronig
relations, one finds that the real part of the q = 0 response has a simple pole at
ω = 0, therefore the static transverse response is divergent as is expected for an
isotropic ferromagnet. For qc > q, the Goldstone modes have lower energies
than the threshold for the Stoner excitations. For large q values, the branch of
spin-wave modes merges with the continua of Stoner excitations as a resonance
which is then rapidly damped out.

By contrast, the uniform static longitudinal response is not divergent in the


ferromagnetic state. The reduced longitudinal susceptibility, evaluated within

683
2 2 2
2qkF↑/kF +(q/kF) 2 2
2qkF↓/kF +(q/kF) 0.25 0.125

0.125
1
0.5
ω/εF

1 M=0.5
2
0
0 0.5 1 1.5 2 2.5 3
-2qkF↑/kF +(q/kF)2
2

-2qkF↓/kF +(q/kF)2
2

-1

q/kF

Figure 285: The phase-space for longitudinal magnetic excitations. Since the
mean-field state is stable against single-particle excitations, the allowed non-
spin-flip excitations must have ω > 0. The upper and lower boundaries of
the continuum excitations are depicted by the blue lines. The contours of the
imaginary part of the longitudinal susceptibility are depicted as black lines.

the Random Phase Approximation, is given by the expression


χ
eσ (q; ω) ( 1 + U χ e−σ (q; ω) )
 
1 X
ezz (q; ω) =
χ (2413)
4 σ 1 − U2 χ eσ (q; ω) χ
e−σ (q; ω)

where χ
eσ (q; ω) represents the mean-field response of the electrons with spin σ

fk,σ − fk+q;σ
 
1 X
χ
eσ (q; ω) = (2414)
N k+q − k − h̄ ω
k

The phase space for longitudinal magnetic excitations resembles the phase space
for paramagnetic fluctuations in that the continuum extends to ω = 0 at q = 0.
Although there does exist a low-energy resonance in the continuum, for appre-
ciable values of Mz /µB , its intensity is greatly reduced below the intensity of
the resonance in the transverse fluctuations.

The Hartree-Fock approximation to the strong ferromagnetic state Mz /µB =


1 is an exact eigenstate of the Hamiltonian, since when there are no down-spin
electrons the effect of Coulomb interaction is zero. Because kF ↓ = 0, one has

h̄2 kF2 ↑
∆ = U n↑ ≥ (2415)
2m
and, therefore, the spectrum of Stoner excitations is either gapped for all mo-
mentum transfers or at most reaches ω = 0 at the isolated point q = kF ↑ .

684
2.5

2
Stoner
1.5 ∆/εF↑ Continuum

hω/εF↑
1

0.5
Spin Waves?
0
0 0.5 1 1.5 2

q/kF↑

Figure 286: The possible phase-space for the spin-flip excitations of a Strong Fer-
romagnet. A Strong Ferromagnet is defined as being fully polarized, Mz /µB =
1.

However, even though the Hartree-Fock state is an exact energy eigenstate for
Mz /µB = 1, it may not be the lowest-energy state or ground state. In par-
ticular, the strong ferromagnetic state is unstable for a half-filled tight-binding
band, as can be seen by examining the above expression for the spin-wave stiff-
ness constant D. Since fk↑ = 1 and fk↓ = 0, the spin-wave stiffness constant
reduces to    
U X 1 2 1 2
D = ∇ εk − |∇εk | (2416)
3N ∆ 2 ∆
k

The first-term vanishes on summing over the entire Brillouin zone and the
second-term is manifestly negative. Hence, the spin wave stiffness constant
is negative. Therefore, the strong ferromagnetic state is unstable, since by
emitting spin waves with finite q the system can lower its energy. This result
is consistent with the prediction that the ground state of the half-filled tight-
binding band is the Neel insulating state.

An anti-ferromagnet also has Goldstone modes, but unlike the ferromagnet


the order parameter (the sub-lattice magnetization) for the anti-ferromagnet is
not a constant of motion. This results in the dispersion of the Goldstone modes
being linear in q, ω = c q, similar to the transverse sound waves in a crystalline
solid.

21.5 The Slater-Pauling Curves


The reduced magnetization can be obtained normalized saturated magnetization
(T → 0, H → ∞) for AB alloys containing 3d atoms. When the reduced
magnetization µ/µB is plotted against the total number of 3d and 4s electrons

685
n, the curves fall on-top of straight-lines with slopes of unit magnitude267 known
as the Slater-Pauling curves268 .
2.5
FeNi

FeCo
2 FeCr

FeV

FeMn
1.5

μ/μB
CoNi

NiCu
1
NiCr

NiV

0.5 NiMn

0
7 8 9 10 11
Mn Fe Co Ni Cu

Figure 287: The reduced magnetization of 3d AB alloys versus the total number
n of 3d and 4s electrons. [After J. Crangle and G. C. Hallam (1963).]

The Slater-Pauling curves indicate whether the magnetic properties of these


alloys can be described by an itinerant electron model in which there is one
common rigid band and a common exchange splitting269 . The line with positive
slope can be considered as the result of the chemical potential being located in
the minimum in the down-spin density of states. A typical example of a density
of states with this minimum270 is given by spin-polarized b.c.c. Fe. The pinning
J B Staunton et al

The theory proves that the relevant one-body effective 30


Density of States (states per Ry per spin)

potential can, in principle, contain all the effects of electron


correlations although, in practice, approximations must
be made. All the theorems and methods of the density 20
functional (DF) formalism were soon generalized [21, 22] Majority Spin
to deal with cases in which spin-dependent properties play 10
an important role, such as in magnetic systems (SDF). The
energy thus becomes a functional both of the density and
0
of the local magnetic density m(r). The proofs of these
theorems are provided in the original details and there have
been many developments of a formal nature [12, 23]. 10
The many-body effects of the complicated quantum- Minority Spin
mechanical problem are buried in the so-called exchange-
20
correlation part of the energy functional Exc [n(r), m(r)].
The exact solution is intractable for macroscopic systems
and some approximation must be made. The so-called 30
-0.6 -0.4 -0.2 0 0.2 0.4
‘local approximation’ is the most widely used owing to
its simplicity and its success in describing the ground state Energy (Ry)
and equilibrium properties of a great many different types Figure 1. The spin-polarized electronic density of states of
of materials. The local approximation (LSDA) takes as BCC Fe in units of states Ry−1 using the one-electron
potentials from [15].
Figure 288: The spin-split density of states of b.c.c. Fe. [After J. B. Staunton
its starting point the energy of a uniformly spin-polarized
homogeneous electron gas εxc (n(r), |m(r)|) [21, 24] so
et sl., (1998).]
that Exc can be written in the form
Z obtained which are expected from the simpler Hubbard-
Exc [n, m] ≈ dr n(r)εxc (n(r), |m(r)|). (7) model treatment described in the last section. Examples
of the chemical potential to the minimum constrains the number of down-spin
of band theory calculations for the magnetic 3d transition
metals BCC iron and FCC nickel can be found in the book
3d electrons to be three. Hence, the reduced magnetization is given by
The functional derivative of this quantity with respect to
m(r) provides the effective magnetic fields for the single- by Moruzzi et al [15]. The figure on p 170 of their book
electron equations, namely the spin-polarized band structure
  (we show a similar figure in figure 1) shows the density of
discussed in the introduction: µ states of BCC iron as a function of energy. The densities
Z
n(r)
= n − 6 − 2n (2417)
of states for the two spins are almost (but not quite)4s↓
rigidly
v eff [n, m; r] = V ext (r) + e2 dr 0
|r − r 0 |
µ shifted. AsBis typical for BCC structures, the d band has
three major peaks. The Fermi energy resides in the top
δE xc 267 J. Crangle and G.
+ [n, m] (8) C. Hallam,
of the d bandsProc. Roy.
for the Soc.majority
so-called 272,spins119between
(1963).
δn(r) 268 J. C. Slater, J. Appl. Phys. the upper8, two385
peaks. The saturation magnetization, Ms ,
δE xc (1937).
B eff [n, m; r] = B ext (r) + [n, m] that results from this ‘self-consistent’ calculation is 2.2µB ,
269 J. Friedel,
δm(r) Nuovo(9)Cimento, which is2, 287 agreement
in good (1958).with experiment.
ext
270 J. B. Staunton, S. S. A, Razee, M. F. Ling, D. D. Johnson and F. J. Pinski, J. Phys. D,
where V describes an external potential such as a lattice In nickel and cobalt the majority spin d bands are
array of nuclei and B extAppl.
an external Phys.magnetic 31,field.
2355The(1998).
completely occupied and the Fermi energy lies in the
electron density and magnetization are given by prominent peak in the minority spin d density of states.
Z εF X The d band width has been a topic for close scrutiny over
n(r) = dε tr(φi∗ (r, ε)φi (r, ε)) (10) the years owing to the width extracted from photoemission
i measurements being much smaller than that of band 686
Z εF X structure calculations. This serves to emphasize the fact
m(r) = dε tr(φi∗ (r, ε)σ̃φi (r, ε)) (11) that the SDF theory is a theory for the ground state whereas
i excited states are probed by spectroscopic measurements.
In particular the theory does not correctly describe the
where the φi (r, ε) obey the Schrödinger–Pauli (Kohn–
correlated motion of the electrons as they are excited into
Sham) equation
states necessary for them to leave the metal. On the
(−∇ 2 + v eff (r)1̃ − σ̃ · B eff (r))φi (r, ε) = εφi (r, ε) (12) other hand all the ground state properties such as Ms
and the lattice spacing are in very good agreement with
where εF is the system’s Fermi energy. experimental values.
The nearly rigidly split spin-polarized bands of these
3d elemental transition metal magnets are actually special
3.1. Elemental metals
cases. We now show how this simple picture is lost as
The line with negative slope can be described as having a completely occupied
up-spin 3d band, containing five electrons, and the magnetization decreases since
the number of down-spin holes increases as n increases. Therefore, one expects
that  
µ
= 10 − n + 2 n4s↑ (2418)
µB
This describes the dominant linear variations. The assumption of one common
rigid band for the AB alloy can be quantified by using the C.P.A.271 . The den-
sity of states of the two elements are centered on A and B and, if |A − B |
is smaller than the band width W , one can expect the density of states will be
unsplit. On the other hand, if |A − B | is greater than the band width W , one
can expect the density of states will be split and one expects deviations from
the dominant linear variation. If the exchange splittings are markedly different
between the A and B atoms, then one expects that the subbands of a specific
spin may resemble unsplit bands while the subbands of the other spin may be
considered as split. Therefore, in this case, the polarized bands are not sub-
jected to a rigid splitting.

21.6 The Heisenberg Model


The above model of itinerant magnetism is believed to be appropriate for tran-
sition metals only involves one type of electrons. Another model is appropriate
for materials which contain two types of electrons, such as rare earth materials,
in which the magnetic moments occur in the f states which are inner orbitals
buried deep inside the f ion and the interaction is mediated by the itinerant
conduction electrons.

The spin localized at site Ri is denoted by S i . The spin at site i interacts


with the conduction electrons spin σ near site i via a local exchange interaction
X
Ĥint = − J S i . σi (2419)
i

which acts like a localized magnetic field of g 2µB J S i . This localized magnetic
field polarizes the conduction electrons, producing a polarization at site j of
 
2
J S i χz,z (Ri − Rj ) (2420)
g µB
This polarization then interacts with the spin at site j via the local exchange
interaction leading to an oscillatory interaction between pairs of localized spins
of the form
X
Ĥ = − J(Ri − Rj ) S i . S j
i,j

271 H. Hasegawa and J. Kanamori, J. Phys. Soc. Japan, 31, 383 (1971).

687
0.01

0.005

J(r)/J0
0
0 2 4 6 8 10 12

kFr
-0.005

-0.01

Figure 289: The form of the oscillatory R.K.K.Y. interaction J(r) between two
local moments mediated via the conduction band.

X  2 2
= − J 2 χz,z (Ri − Rj ) S i . S i (2421)
i,j
g µB

The oscillations in the interaction are produced by the oscillations of the re-
sponse function of the conduction electrons. The Fourier transform of J(R)
shows the oscillation frequency is 2 kF . This interaction was discovered inde-
pendently, by Ruderman and Kittel, Kasuya and Yosida272 .

22 Localized Magnetism
The nearest neighbor Heisenberg exchange interaction couples spins localized
on adjacent lattice sites
J X
Ĥ = − S(R + δ) . S(R) (2422)
2
R,δ

where δ are the vectors connecting pairs of adjacent sites. This interaction
Hamiltonian can be derived from the model of itinerant magnetism, for large U
in the case when the bands are half filled. In this case, there is a spin at each
lattice site and the exchange between the spins is the anti-ferromagnetic super
exchange interaction found by Anderson273 . The exchange constant is given in
terms of the tight-binding matrix element t and the Coulomb repulsion via
t2
J = −4 (2423)
U
272 M. Ruderman and C. Kittel, Phys. Rev. 96, 99 (1954), T. Kasuya, Progr. Theor. Phys.

16, 45 (1956), K. Yosida, Phys. Rev. 106, 803 (1957).


273 P.W. Anderson, Phys. Rev. 115, 2 (1959).

688
The Heisenberg Hamiltonian can be expressed as
"  #
J X 1
Ĥ = − S z (R) S z (R+δ) + S + (R) S − (R+δ) + S − (R) S + (R+δ)
2 2
R, δ
(2424)
For a ferromagnetic exchange, J > 0, the ground state of a three-dimensional
lattice of spins of magnitude S consists of parallel aligned spins. The direction
of quantization is chosen as the direction of the magnetization. All the spins
have their z-components of the spin maximized
Y
| Ψg > = | ( mR = S ) > (2425)
R

This state has a total magnetization proportional to N S, and is an eigenstate


of the Hamiltonian, with energy Eg = − 3 N J S 2 since the z component of
the Hamiltonian is diagonal and the spin flip terms vanish as the effect of the
raising operators acting on the fully-polarized state

S + (R) | mR = S > = 0
S + (R + δ) | mR+δ = S > = 0 (2426)

are both zero. The excitations of this system are the spin waves. The excited
state wave function corresponding to a single spin wave is given by
 
1 X
|q > = √ exp − i q . R S − (R) | Ψg > (2427)
N R

It corresponds to a state with total spin N S − 1 since one spin is flipped


over. This state is a coherent superposition of all states in which only one spin
is flipped and has total momentum of q. The excitation energy is found from
 
( Ĥ − Eg ) | q > = 2 J S 3 − cos qx a − cos qy a − cos qz a | q > (2428)

At long wave-lengths, the spin wave excitation energy h̄ω(q) is found as


 
h̄ ω(q) = 2J S 3 − cos qx a − cos qy a − cos qz a

∼ J S q 2 a2 (2429)

which is the branch of collective Goldstone modes that dynamically restore


the spontaneously broken spin rotational invariance of the ferromagnet. The
quadratic variation of the frequency near q = 0 is a general consequence of the

689
total spin being a constant of motion. In the limit q → 0, the spin wave state
just corresponds to a uniform reduction of the total magnetization since
X  
exp − i q . R S − (R) → ST−ot (2430)
R

The above excitations of the spin system are small amplitude excitations
that have a close resemblance to the harmonic phonons of a crystalline lattice.
The effects of the interactions could be expected to produce small anharmonic
corrections to these excitations, providing them with a lifetime and a renor-
malized dispersion relation. Not all the excitations can be expressed as small
amplitude excitations, some systems have large amplitude soliton excitations
that cannot be treated by perturbation theory. However, the small amplitude
excitations can be adequately treated as harmonic modes, as can be seen from
an analysis based on the Holstein-Primakoff transformation.

22.1 Holstein-Primakoff Transformation


The Holstein-Primakoff transformation provides a representation of localized
spins, which enables the low-temperature properties of an ordered spin system
to be analyzed in terms of boson operators. The technique is particularly useful
for systems where the magnitude of the spin S is large S > 1. The Hamilto-
nian can be expanded in terms of boson operators, providing a description of the
ground state, small amplitude spin fluctuations and the anharmonic interaction
between them.

The Holstein-Primakoff transformation274 of the spins represents the effect


of the spin operators by a function of bosons operators. The components of
the spin operators at the i-th site Ŝiα can be defined by their action on the
eigenstates of Ŝiz
Ŝiz | mi > = mi h̄ | mi > (2431)
In particular the spin raising and lowering operators
Ŝi± = Six ± i Siy (2432)

have the commutation relations with Ŝ x


 
±
z
Ŝi , Ŝi = ± h̄ Ŝi± (2433)

This can be used to show that the operators Ŝi± have the effect of raising and
lowering the magnitude of the eigenvalue S z by one unit of h̄
Ŝi± | mi > =
p
S ( S + 1 ) − mi ( mi ± 1 ) h̄ | mi ± 1 > (2434)
274 T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 (1940).

690
The Holstein-Primakoff transformation represents the local ( 2 S + 1 ) basis
states | mi > by the infinite number of basis states of a local boson number
operator a†i ai . The local boson basis states, | ni >, are defined through

a†i ai | ni > = ni | ni > (2435)

The relation between the local spin basis states and boson states is provided by
the boson representation of the z-component of the spin operator

Ŝiz = S − a†i ai (2436)

Thus, the state where the spin is aligned completely along the z-axis (mi = S)
is the state with boson occupation number of zero, and the state with the
lowest eigenvalue of the spins z-component (mi = − S) corresponds to the
state where 2 S bosons are present. The states with higher number of bosons
are un-physical and must be projected out of the Hilbert space. The effects
of the spin raising and lowering operators are similar to the boson annihilation
and creation operators, within the space of physical states. The correspondence
between the spin and raising and lowering operators and the boson operators
can be made exact, by multiplying the boson creation and annihilation operators
with a function that ensures that only the physical acceptable boson states form
the Hilbert space of the spin system. This is achieved by representing the spin
lowering operator as
q 
Ŝi− = a†i 2 S − a†i ai (2437)

which decouples states with more than 2 S bosons present. The raising operator
is given by the Hermitean conjugate
q 
Ŝi+ = 2 S − a†i ai ai (2438)

This transformation respects the spin commutation relations.

The nearest neighbor ferromagnetic (J > 0) Heisenberg Hamiltonian

J X
Ĥ = − Ŝ(R) . Ŝ(R + δ) (2439)
2
R,δ

can be expressed as
"  #
J X z z 1 + − − +
Ĥ = − Ŝ (R) Ŝ (R + δ) + Ŝ (R) Ŝ (R + δ) + Ŝ (R) Ŝ (R + δ)
2 2
R,δ

(2440)

691
S ω

Figure 290: The coherent spin precession of a classical spin wave, as seen along
its direction of propagation q.

On representing this Hamiltonian in terms of the boson operators, and expand-


ing in powers of S1 , one finds
"  #
J X † † † †
Ĥ ≈ − ( S − aR aR ) ( S − aR+δ aR+δ ) + S aR aR+δ + aR+δ aR
2
R,δ

(2441)

The terms of order S 2 just represents the classical ferromagnetic ground state, in
which all the spins are aligned. The terms of order S represent excitations from
the ground state and can be put in diagonal form by expressing them in terms
of the Fourier transformed boson operators. The spatial Fourier transform of
the boson operators are defined as
 
1 X
aR = √ exp i q . R aq (2442)
N q

and the creation operator is given the Hermitean conjugate expression


 
1 X
a†R = √ exp − i q . R a†q (2443)
N q

Substitution of the creation and annihilation operators in the Hamiltonian, and


performing the sums over the spatial index, yields the following approximate
expression for the Hamiltonian
J X X  
2 †
Ĥ ≈ − N Z S + J S aq aq 1 − cos q . δ (2444)
2 q δ

The above expression is the sum of the ground state energy and the energies of
harmonic normal modes that represent the excitations of the spin waves from
the ferromagnetic ground state. The terms of higher-order in S1 yield quantum
corrections to the ground state energy, the spin wave energies and also produce
anharmonic interactions between the spin waves.

692
3
Figure 291: The approximate T 2 decrease of the magnetization Mz of a single
crystal of N i at low temperatures. [After B. E. Argyle, S. H. Charap and E. W.
Pugh, Phys. Rev. 132, 2051 (1963).].

The thermally excited spin waves have the effect of reducing the magnetiza-
tion from the fully saturated T = 0 value
X
M z (T ) − N S = N (ω(q)) (2445)
q

where N (ω) is the Bose-Einstein distribution function. Since the ferromagnetic


spin waves have a dispersion relation which is ω(q) ∼ q 2 at small q, one finds
that the temperature induced change in the magnetization is given by
  32
kB T
M (T ) − M (0) ∼ (2446)
J S
for a three-dimensional lattice at low temperatures. Likewise, the thermal av-
erage value of the energy can be calculated as
J 2 X
E(T ) + N Z S = h̄ ω(q) N (ω(q))
2 q
  52
kB T
E(T ) − E(0) ∼ N J S (2447)
J S
This should result in the low-temperature specific heat being proportional to

693
3
T 2 for a long-range ordered insulating ferromagnet.

22.2 Spin Rotational Invariance


In the limit of zero applied magnetic field, the Heisenberg exchange Hamiltonian
is invariant under the simultaneous continuous rotation of all the spins. As there
is no preferred choice of z-axis, the energy of the ferromagnetic state of a fully
polarized spin system with a total spin S T = N2 has the same energy as the state
where all the spins are oriented along the direction (θ, ϕ). The ferromagnetic
state where the spins are fully-polarized along the z-axis is given by
Y 1
|0 >= | mj = > (2448)
j
2

The state where the polarization is rotated through (θ, ϕ) is given by


Y  θ
 
θ

| θ, ϕ > = cos îj + exp i ϕ sin Ŝj− | 0 > (2449)
j
2 2

This can be proved by representing the spin vector operator for one site σ j in
terms of its component along the unit vector η̂ in the direction (θ, ϕ)

η̂ . σ j = sin θ cos ϕ σx j + sin θ sin ϕ σy j + cos θ σz j (2450)

where σx j , σy j and σz j are the three Pauli spin matrices for the spin at site
j. Thus, the representation of the operator for the η̂ component of the spin at
site j is found as
   
 cos θ sin θ exp − i ϕ 
 
 
η̂ . σ j = 
   

 sin θ exp + i ϕ
 − cos θ 

This above component of the spin operator has two eigenstates. The eiegen-
state
 
θ θ
| θ, ϕ + > = cos | + > + exp + i ϕ sin | − > (2451)
2 2

corresponds to the eigenvalue of +1 and


 
θ θ
| θ, ϕ − > = cos | − > − exp − iϕ sin |+ > (2452)
2 2

694
corresponds to the eigenvalue of −1. The un-rotated ferromagnetic state has all
the spins in the | + > spinor state and after rotation all the spins have maximal
eigenvalue along the direction (θ, ϕ). In the rotated state each spin is described
by the | θ, ϕ + > spinor state. Therefore, the rotated ferromagnetic state has
all the spins aligned along the same direction. This classical ferromagnetic state
has an infinitesimal overlap with the un-rotated ferromagnetic state, since
 N  
θ θ
< 0 | θ, ϕ > = cos = exp N ln cos (2453)
2 2
which vanishes in the limit N → ∞. The rotated state can be considered
to be a Bose-Einstein condensate of the q = 0 spin waves. For example on
expanding the rotated state in powers of exp[ i ϕ ], one finds

( ST−ot )n
X  
| θ, ϕ > = A(n) exp i n ϕ |0 >
n=0
n!

( ST−ot )n
 
(N −n) θ n θ
X
= cos sin exp i n ϕ |0 >
n=0
2 2 n!
(2454)

where the total spin operator is defined by


X
ST−ot = Ŝj− (2455)
j

since ( Sj− )2 ≡ 0. The un-normalized states with n spin flips present are
defined as
( ST−ot )n
|n > = |0 > (2456)
n!
These states have the normalization
 
N
< n|n >= C (2457)
n
Hence, the rotated spin state can be expressed as
∞  
X
(N −n) θ θ
| θ, ϕ > = cos sinn exp inϕ |n > (2458)
n=0
2 2

The number of q = 0 spin flips, n, are distributed with the binomial


probability
| < n | θ, ϕ > |2
P (n) = (2459)
< n|n >
which yields  
N θ θ
P (n) = C cos2(N −n) sin2n (2460)
n 2 2

695
Since N is a macroscopic number, the distribution of spin flips is a sharp Gaus-
sian distribution with a peak at nmax = N sin2 θ2 , and a width given by
1
∆n = N 2 sin θ2 cos θ2 . The number of bosons in the q = 0 spin wave mode
is macroscopic and of the order N . This is a coherent representation, and it is
the quantum state that is closest to a classical state as possible.

The coherent state corresponds to the classical state in which the total spin
is oriented along the direction (θ, ϕ). This can be established by examining the
matrix elements of the spin operators. The matrix elements of the total spin
lowering operator between states with total numbers of q = 0 spin flips close
to the maximum of the wave packet are found first by noting that
   
− N
< n + 1 | ŜT ot | n > = N − n C (2461)
n

since Ŝ − can only have a non-zero effect on the N − n up-spins and creates
an extra down-spin. The expectation value of the spin lowering operator in the
rotated state is found to be
X  

< θ, ϕ | ST ot | θ, ϕ > = exp − i ϕ A(n + 1) A(n) < n + 1 | ŜT−ot | n >
n
  X  
θ N
= tan exp − iϕ A(n)2 ( N − n ) C
2 n
n
 
N
= sin θ exp − iϕ (2462)
2
and likewise, the matrix elements of the spin raising operator are given by the
complex conjugate. Finally, the matrix elements of the ŜTz ot is given by
 
N N
< n | ŜTz ot | n > = ( − n) C (2463)
2 n
Thus, one has
X
< θ, ϕ | STz ot | θ, ϕ > = A(n)2 < n | ŜTz ot | n >
n
 
X (N − 2n)
2 N
= A(n) C
2 n
n
N
= cos θ (2464)
2
Therefore, the components of the total spin operator have matrix elements be-
tween the coherent state that exactly corresponds to the components of the
classical vector.

Thus, the different classical ferromagnetic states are represented by coher-


ent states which are superpositions of states with arbitrary numbers of excited

696
Goldstone modes. The thermal average in a ferromagnetic state should yield
a zero magnetization for a system in the thermodynamic limit. The thermal
average has to be taken, in the thermodynamic limit, in the presence of an ar-
bitrary small magnetic field. In this case, the different classical states have zero
overlap and can be considered as being in disjoint portions of Hilbert space. In
this quasi-static state the field may then be driven to zero leading to a non-
vanishing vector order parameter.

——————————————————————————————————

22.2.1 Exercise 92
Determine the spin wave spectrum for an isotropic Heisenberg ferromagnet in
the presence of an applied magnetic field. Do the conditions of Goldstone’s the-
orem apply, and what happens to the excitation energy of the q = 0 spin wave?

——————————————————————————————————

22.3 Anti-ferromagnetic Spinwaves


One can obtain an approximate spin wave spectrum for an anti-ferromagnet (
J < 0 ) in a Neel state. Neel ordering275 shall be considered on a crystal
structure that can be decomposed into two interpenetrating sub-lattices. The
spins on one sub-lattice (the A sub-lattice sites) shall be oriented parallel to
the z-axis, and the spins on the second sub-lattice (the B sub-lattice) are anti-
parallel to the z-axis. In order for the bosons to represent excitations, it is
necessary to switch the directions of the spins on the B sub-lattice Siz → − Siz ,
Six → Six and Siy → − Siy . This is a proper rotation of π about the x-axis
so that the commutation relations remain the same. The Holstein-Primakoff
transformation for the operator representing the z-component of the B spins is
of the form
Ŝiz = b†i bi − S (2465)
and the spin raising operators for the B spins are
q 
+ † †
Ŝi = bi 2 S − bi bi (2466)

which decouples states with more than 2 S bosons present. The lowering oper-
ator is given by the Hermitean conjugate expression
q 
− †
Ŝi = 2 S − bi bi bi (2467)

275 L. Neel, Ann. de Physique, 17, 64 (1932), Ann. de Physique 5, 256 (1936).

697
2. Crystal and magnetic structures of MnF2

The crystal structure of MnF2 belongs to the tetragonal space group D14 4h with two molecules
per unit cell. The lattice constants at room temperature are a = 4.8734 Å and c = 3.3103 Å
(Griffel and Stout 1950).
From the neutron scattering study of Erickson (1953) below the Néel temperature
(TN = 67.34 K) (Heller and Benedek 1962), the magnetic structure of MnF2 was determined.
In the ordered phase, the spins at body centre sites point antiparallel to those at the corner
sites with the spin easy axis parallel to the c axis. The main origin of the magnetic anisotropy
is the dipole–dipole interaction.

Figure 1. The temperature dependence of the molar magnetic susceptibilities parallel and
Figure 292: The temperature-dependence
perpendicular to the c axis of MnFof the anisotropic magnetic suscepti-
2 measured at 100 Oe.
bility of M nF2 parallel and perpendicular to the c-axis. [After M. Hagiwara,
K. Kamatsumata, I. Yamada and H. Suzuki, J. Phys. C.M. 8, 7349 (1996).].
3. Experimental details
for the3.1.
B sub-lattice.
Sample preparation and characterization

AnSingle crystals ofspin


semi-classical 2 wereHamiltonian
MnFwave grown by the can
Bridgman method.
be found 276 A commercially available
by expanding the
powder of MnF2 with 4 −1 N purity was placed in a Pt crucible, melted at about 900 ◦ C and
Hamiltonian in powers of◦S −1. The expansion yields the expression
cooled at the rate of 1 C h under an atmosphere of mixed N2 and HF gas. A transparent
pink X  single crystal was obtained. A thin disc of MnF2 with the c axis
coloured  perpendicular
Ĥ ≈toJthe plane (was a†i aia)larger
− from
S cut (S − b†j bj ) − S ( a†i b†j + ai bj ) (2468)
crystal.
Ini,jorder to characterize the crystal, we have measured the temperature dependence of
the magnetization (M) under an external magnetic field (H ) using a SQUID magnetometer
On defining
(Quantum the Design’s
Fourier MPMS2).
transformedThe operators
results are shown in figure 1. There is no anisotropy
in the susceptibility (M/H
r ) along the  c plane (χ⊥ ) above about 67 K.
 c axis (χk ) and
On the other hand, χk and 2 Xχ⊥ behave quite differently below ∼67 K, which is a typical
aq = exp − i q . Ri ai
N i
r  
2 X
bq = exp − i q . Rj bj (2469)
N j

then the Hamiltonian can be reduced to


X 
Ĥ = N z J S2 − J S z ( a†q aq + b†−q b−q )
q
  
b†−q
X
+ exp − iq.δ ( a†q + aq b−q ) (2470)
δ

Since this form is not diagonal in the a and b operators, it is necessary to use
the Bogoliubov canonical transformation to diagonalize the Hamiltonian. The
276 P.W. Anderson, Phys. Rev. 86, 694 (1952).

698
form of the Bogoliubov transformation is given by
   
αq = exp + Ŝ aq exp − Ŝ
   
β−q = exp + Ŝ b−q exp − Ŝ (2471)

where the anti-Hermitean operator Ŝ is given by


X θq  † 

Ŝ = b−q aq − aq b−q (2472)
q
2

and where θq still has to be determined. The transformation is evaluated as


θ θ †
αq = cosh aq − sinh b
2 2 −q
θ θ †
β−q = cosh b−q − sinh a (2473)
2 2 q
in which θq is to be chosen so that the Hamiltonian is diagonal. The inverse
transformation is given by
θ θ †
aq = cosh αq + sinh β
2 2 −q
θ θ †
b−q = cosh β−q + sinh α (2474)
2 2 q
After substitution of the inverse Bogoliubov transformation into the Hamilto-
nian, the Hamiltonian takes the form
Ĥ = N z J S2   
X X
† †
−J S z cosh θq − exp[−iq . δ] sinh θq αq αq + βq βq
q δ
X  X 


−J S − z sinh θq + exp[−iq . δ] cosh θq αq† β−q + βq α−q
q δ

(2475)
The value of θq which eliminates the off-diagonal terms in the Hamiltonian is
found as  
1 X
tanh θq = exp − i q . δ (2476)
z
δ
The resulting approximate Hamiltonian can be interpreted in terms of a zero-
point energy and a sum of harmonic normal modes, as was done for the ferro-
magnet. However, unlike the case of ferromagnetic spin waves, the amplitude
of anti-ferromagnetic spin waves can be anomalously large.

——————————————————————————————————

699
i
2 l
2 l
2

000 001

Figure 1. The reciprocal lattice diagram of a simple cubic antlferromagnet m a


110 plane. N and M mark the elastic nuclear and magnetic reflectlon pomts.
T h e thm lines show the corresponding zone boundaries. Heavy lines show a
possible neutron-scattering hagram for the experimental settings used. k and k
represent the mcident and scattered neutron wave vectors. K represents the wave
vector transfer k -k, while q represents the transfer relative to the centre of the
zone.

0 0.2 0.4 0.6


Wave vector Iq I (1-1)
Figure 2. Spin waves m RbMnFs at 4 2 OK wlth q vectors dstnbuted over a 1l0
Plane. The smooth curves show the calculated dispersion along Pamcdar
directions w t h JI = 3 4 OK, J2 = J 3 = 0.0 OK. These values were found from a lest-
Figure 293: The spin wave dispersion relation of the antiferromagnet RbM nF3
Squares analysis, the exact direction of all the q vectors bemg taken into accounl*
The fact that the linear part of the curve extends so close to the o r i p reflectsthe
in the 110 plane. [After C. G. Windsor, and R. W. H. Stevenson, Proc. Phys.
very small anisotropy field.
Soc. 87, 501 (1966).].

22.3.1 Exercise 93
Find the approximate dispersion relation for spin waves of a Heisenberg anti-
ferromagnet, J < 0, for spins of magnitude S. The Hamiltonian describes
interactions between nearest neighbor spins arranged on a simple cubic lattice,
X
Ĥ = − J S(R) . S(R + δ) (2477)
R, δ

Assume that S is large so that the classical Neel state can be considered as
being stable. Also calculate the zero point energy.

——————————————————————————————————

Since tanh θ → 1 in the limit q → 0 then θ → ∞ so both sinh θ2 and


cosh θ2 diverge. In one dimension, the change in the sub-lattice magnetization
< ψ | a†i ai | ψ
P > diverges logarithmically. In two and three dimensions, the

divergence in q < ψ | aq q | ψ > is integrable and converges. There is an
a
energy change relative to the nominal classical energy of the Neel state, given
by the sum of the zero point energies,
 
1
Eg = J z N S 2 1 + ( 1 − Id ) (2478)
S

where v
u  Xd 2
2 X ut d2 −
Id = cos qi a (2479)
N d q i=1

700
Spin waves in antiferromagnetic FeFz 315
80r I

70- -7
v

h
P
c
\
W

<I 0o> <OOl>


501--1--- I 1- ~~~~~ 1--1~L--- _L.--
A d ~

0.2 0.4 0.6 0.8 I .o


9, (i-')
Figure 5. The points are the energies of those spin waves observed which lie close to two
symmetry directions. The error bars represent estimates of the probable uncertainty in
Figure 294: The
the spin wave
measurements. dispersion
The full relation
curves are calculated fromofequations
the antiferromagnet F eF2 for
(6) and (7) using
parameters giving the best fit to the data throughout the zone.
two high-symmetry directions. The system has a large gap due to single site
Table 2. Values obtained for parameters (in cm-1 )in the spin Hamiltonian for FeFz by fitting to the
anisotropy. [After M. T. Hutchings, B. D.approximations
neutron data with different Rainford and H. J. Guggenheim, J.
Phys. C. 3, 307 (1970).]. D Jl J2 J3

A 6.46 +
0.29 -0.048 k 0.060 3.64 k 0.10 0.194 i 0.060
- 0.10
B(i) 6.69 - 0.074 3.57 0.159
The amplitude
B(ii) of the
6.34q = 0 spin wave is divergent,
0.0 12 3.73 but be neglected in 0.209
can
B(iii) 6.8 1 - 3.6 1
the limit NC → ∞.6.94As the q -= 0 spin wave 3.5 1 has zero frequency, one can
compose aA, state which
best fit is a superposition
to full Hamiltonian of the
equation (5) including q interaction
dipolar = 0 spin wave
dispersion excitations.
and the
full Oguchi factor. Final result.
The dynamics of the finite frequency spin wave excitations can be examined
B(i), best fit including dipolar interaction and Oguchi factor in full but not fixing E , at 52.7 cm- l,
for
finite timeB(ii),
scales before
fit obtained theOguchi
omitting zero-energy
factor (i.e. a4 =amplitude
0), and omitting wave packet
the dispersion of theof q = 0 spin
dipolar
interaction. E , fixed at 52.7 cm-'.
waves diverges re-orienting
B(iii), fit obtained with clq = the sub-lattice
$: = Jl = J3 = 0. E , ismagnetization.
fixed at 52.7 cm-' and D will include the
dipolar interaction.
C, values obtained with clq = $: = J, = J, = 0. J, and D are calculated from E , = 52.7 cm-',
Just like in
andthe 77.1 cm- '.D will include
E,, =ferromagnetic state, theinteraction.
the dipolar magnetic response of an antiferro-
magnet to an applied magnetic field depends on the direction of the applied
zero or antiferromagnetic value, while J, is larger and antiferromagnetic. These parameters
field relative
will be to the direction
discussed in 4 6. Using ofthemthe
we order
calculateparameter.
the zone boundary In energies
the absence
to be of any
anisotropy, it =is79.2expected
Eool(Z) that
cm-', EIoo(X) the
= 77.7 application
cm-', E,,,(R) = 77.7ofcm-',a magnetic
E,,,(M) = 76.1field
cm-' will cause
and E,,,(A) = 76.1 cm-', with errors of + 2 cm-'. The labels are the symmetry points
the Neeldesignated
state to become
in figure unstable
6. It is sometimes to a to
useful spin-flop state.
represent the In the
dispersion spin-flop state,
by a formula
the sublattice magnetizations
which reproduces the approximatewill rotate
spin-wave to be
energies in aperpendicular
simple manner. The to theof direction
values
J, and D in row C of table 2 are calculated from E , and a weighted average of the above zone
of the applied field, and will also develop
boundary energies at the five special points. small components parallel to the field
direction that are proportional to the field. The presence of uniaxial anisotropy
z 2 1
P
term in the Hamiltonian (such as − i D ( S i ) for S > 2 ) will stabilize
the ground state against the spin-flop transition, for sufficiently small magnetic
fields277 . However, in this case the spin-wave spectrum will develop a gap as
Goldstone’s theorem does not apply278 . For fields applied that are perpendic-
ular to the sublattice magnetization, the transverse susceptibility is constant
below the Neel temperature. On the other hand, for the field direction parallel
to the sublattice magnetization, the longitudinal susceptibility falls to zero as
T 2 below TN .
277 F. Burr Anderson and H. B. Callen, Phys. Rev. 136, A 1068 (1964).
278 R. Kubo, Phys. Rev. 87, 568 (1952).

701
Figure 295: The concentration-temperature phase diagram AuF e alloys, show-
ing the paramagnetic, super-paramagnetic, ferromagnetic, cluster glass and spin
glass phases. [After B. R. Coles, B. V. B. Sarkissian and R. H. Taylor, Phil.
Mag. B 37, 498 (1978).]

23 Spin Glasses
Spin glasses are found when magnetic impurities are randomly distributed in
a metal, such as F e impurities in gold Au or M n in Cu. Due to the random
separations between the moment carrying impurities, the R.K.K.Y. interaction
between the magnetic moments are also randomly distributed and can take on
both ferromagnetic and anti-ferromagnetic signs. The distribution of interac-
tions prevents the local magnetic moments from forming a long-range ordered
phase at low temperatures. Nevertheless, the random spin system may freeze
into a spin glass state below a critical temperature. At high temperatures the
spins are disordered, and as the temperature is reduced, the spins which are
most strongly interacting progressively build up their correlations and freeze
into clusters. The dynamics of the spin clusters slow down as they grow, and at
a critical temperature Tf they lock into the spin glass phase. However, this may
not be the lowest energy state, as in order to reach the ground state there may

702
THE EFFECT OF SPIN-GLASS ORDERING ON THE SPECIFIC HEAT OF CuMn

G.E. BRODALE, R.A. FISHER, W.E. FOGLE, N.E. PHILLIPS a n d J. V A N C U R E N


Department of Chemistry and Matertals and Moleeular Re~eareh Dtwston, Lawren(e Berkeley Laboratory. Unmerstt} of
Cahforma, Berkele), CA 94720. USA

Hlgh-premsmn measurements have shown the nature of the specific heat anomaly associated w, th the transmon to the
spin-glass phase m CuMn, and have been used to determine the entropy as a function of the field and temperature

theoretical d e m o n s t r a t i o n [1] that order in spin 200 I I I


could a p p e a r at a well defined crmcal tempera- 2790 ppm C.~uuMn
~g, explained the observed d i s c o n t l n u m e s m mag-
roperties [2] at least qualitatively, but left the
[50
to observe comparable features in the specific ~,
ore puzzling than ever More recently, the prop-
of spin glasses have usually been interpreted m _~
of the S h e r r m g t o n - K l r k p a m c k mfmlte-range
, 100
treated in various extensions and approximations x:
H (kOe)
particular, the P a r l s l - T o l o u s e hypothesis [4] has -~
idely mvoked According to this hypothesis, the
y, S, is i n d e p e n d e n t of applied field, H, below the
lity line or phase boundary, T~g(H), recognized
AlmeIda and Thouless [5]. This paper is a review
results of a continuing series of calorimetric
vco
50
io
2o
5o
o
45
6o
75
1
t t i
e m e n t s designed to determane the thermody- 0 50 Ioo 15 0 200
properties just above and below T,g(H), and to T (K)
for effects in the specific heat associated with T,g Fig 1 Typical specific heat data for CuMn (T,g = 3 89 K)
Figure 296: The temperature dependence of the magnetic heat capacity of a
I I E I I I i '
CuM n alloy. [After G. E. Brodale, R. A. Fisher, W. E. Fogle, N.E. Phillips and
J. Van Curen, J. Mag. Mag. Mat. 31-34, 1331 (1983).]
0 • •

have to be large scale reorientations of the spin clusters. Thus, the spin glass
state is not unique but instead is highly degenerate. This occurs as a result
of frustration. The concept of frustration is illuminated by imagining that all
the spins on the magnetic sites are frozen in fixed directions, except one, then
there is a high probability that the long-ranged interactions between the spin
5
under consideration and the fixed spins almost average out to zero. At finite
temperatures, the spin under consideration is almost degenerate with respect to
the orientation of the spin as it leads to an insignificant lowering of the energy
of the spin glass state.

The experimental signatures of spin glass freezing are a plateau in the static
0
susceptibility and a rounded peak in the specific heat. The susceptibility follows
a Curie-Weiss law at high temperatures
800
I i I I k I H(Oe) I
2
25 50 55 40 45 50 55 60 6 5µB S(S + 1)
χ(T ) = c (2480)
T (K) 3 kB ( T − Θ )
The field dependence of C / T =- A + BH 2 The error bars in the reset correspond to + 0 01% of the total measured specffm
where Θ is the strength of the resultant interaction on an individual spin. For
an R.K.K.Y. interaction, the Curie-Weiss temperature Θ should be proportional
8853/83/0000 0 0 0 0 / $ 0 3 0 0 © 1983
to c. N At
o r t hlower
- H o l l a temperatures
nd where the spin freeze into clusters, the effective
moment increases, reflecting the growth of the clusters. The peak in the specific
heat, encloses an entropy which is a considerable fraction of

∆S = c kB ln (2S + 1) (2481)

703
Figure 297: The temperature dependence of the a.c. susceptibility of several
AuF e alloys. [After V. Canella and J. A. Mydosh (1972).]

where c is the concentration of magnetic impurities of spin S. This entropy


represents the entropy of the spins gradually freezing into clusters, but does not
contain the entropy of the frustrated spins. This maximum disappears and is
broadened and shifted to higher temperatures as a magnetic field is applied. The
effect of the applied field is to order the spins at higher temperatures. Crude
estimates indicate that about 70 % of the spins are already ordered above Tf .
The temperature dependence of the resistivity shows a sharp drop or knee at the
spin glass freezing temperature. At this temperature, the majority of spin are
frozen in specific directions preventing the logarithmic increase with decreasing
temperature associated with spin flip scattering.

As the spin glass phase is not a ground state but is instead a highly de-
generate meta-stable state, the most unusual properties occur in the dynamical
properties. The low-field a.c. susceptibility shows a very sharp cusp at the spin
glass freezing temperature279 . The cusp becomes rounded and the temperature
of the peak diminishes as the a.c. frequency is lowered. The susceptibility satu-
279 V. Canella and J. A. Mydosh, Phys. Rev. B, 6, 420 (1972).

704
Figure 298: The temperature dependence of the a.c. susceptibility of two AuF e
alloys for zero and various applied fields. [After V. Canella and J. A. Mydosh
(1972).]

rates to a finite value at T = 0 which is roughly half the value of the cusp and
has a T 2 variation on the low-temperature side. The d.c. susceptibility shows
a memory effect, in that, in field cooled samples (H 6= 0) the susceptibility
saturates at Tf , and the curve χ(T ) is reversible as it is also followed for in-
creasing temperature at fixed field. By contrast, the zero field cooled samples
(H = 0) shows a cusp280 at Tf . The susceptibility is zero, by definition, until
the field is applied. When the field is applied, the susceptibility jumps to a value
similar to that found via the a.c. susceptibility measurements and shows the
cusp. However, below Tf , the value of the susceptibility also increases with in-
creasing measurement time. The magnetization is slowly increasing as the spins
slowly adjust to lower energy state in the presence of the applied field. This is
contrasted to the field cooled state in which the spins have already minimized
the field energy before the temperature is lowered and they are frozen into the
spin glass state.

The spin glass freezing resembles a phase transition, but the nature of the
order is unclear as the spin glass state involves disorder and is a highly degener-
ate meta-stable state. Likewise, the description of the low frequency dynamics
of the magnetization is complicated by the existence of long-ranged correlations
between large groups of spins. Since there is no well defined order parame-
ter, there is no well defined low frequency Goldstone mode. Several important
steps in the solution of the thermodynamics of the spin glass problem have been
undertaken; this includes the discovery of the nature of the order parameter,
280 S. Nagata, P. H. Keesom and H. R. Harrison, Phys. Rev. B 19, 1633 (1979).

705
Figure 299: The temperature dependence of the susceptibility of two samples
of CuM n. Curves (b) and (d) were obtained when the samples were cooled in
zero field, before the field was applied. Curves (a) and (c) were obtained when
the samples were cooled in a finite field. [After S. Nagata, P.H. Keesom and
H.R. Harrison (1979).]

by Edwards and Anderson281 , the formulation of a model which is exactly


soluble mean-field theory by Sherrington and Kirkpatrick 282 . The Sherrington-
Kirkpatrick model consists of an Ising interaction
X
Ĥ = − Ji−j Siz . Sjz (2482)
i,j

where Ji−j is a randomly distributed long-ranged interaction between the spins.


The average value of Ji−j is zero

< Ji−j > = 0 (2483)

and the average value of the square is given by

2 J2
< Ji−j >= (2484)
N
The averaging over the randomly distributed interactions is not commutative.

23.1 Mean-Field Theory


The simplest mean-field approximation is based on a representation of the free
energy, for a spin glass with long-ranged interactions between the Ising spin
S = 21 , in which the exact value of the spin on a site i is replaced by the
281 S. F. Edwards and P. W. Anderson, J. Phys. F, 5, 965 (1975).
282 D. Sherrington and S. Kirkpatrick, Phys. Rev. Lett. 35, 1792 (1975).

706
thermal averaged value mi . The mean-field free energy F [mi ] is given by
X X  1 + mi 1 + mi 1 − mi 1 − mi

F [mi ] = − Ji,j mi mj + kB T ln + ln
i,j i
2 2 2 2
(2485)
where the exchange interactions Ji,j are randomly distributed. On minimizing
the Free energy, one finds, the mean-field magnetization at every site. Above the
spin glass freezing temperature, the average magnetization at each site is zero.
Below the freezing temperature the spin on each site has a non-zero average
value, the direction and magnitude varies from site to site and is determined by
the non-trivial solution of

X kB T 1 + mi
0 = −2 Ji,j mj + ln (2486)
j
2 1 − mi

On linearizing in mi (only valid for T ≥ Tf ), one obtains the eigenvalue


equation which determines the spin glass freezing temperature
X
kB Tf mi = 2 Ji,j mj (2487)
j

in terms of the largest eigenvalue of the random matrix Ji,j . This is solved by
finding a basis λ that diagonalizes the matrix
X
Ji,j = Jλ < i | λ > < λ | j > (2488)
λ

The basis gives the set of the spin configurations that the spins will be frozen
into below the spin glass freezing temperature. In the limit N → ∞, the
eigenvalues of the random exchange matrix are distributed according to a semi-
circular law283
1
q
ρ(Jλ ) = 4 J 2 − Jλ2 (2489)
2 π J2
where, obviously, 2 J is the largest eigenvalue. The spin glass freezing temper-
ature is determined as
kB Tf = 4 J (2490)
This mean-field theory predicts a transition temperature which is a factor of
2 too large. This is because the mean-field theory needs to incorporate a self
reaction term. Namely, the reaction term includes the effect of the central spin
on the neighbors back on itself, before the thermal averaging is performed284 .

283 S. F. Edwards and R. C. Jones, J. Phys. A 9, 1591 (1978).


284 D. J. Thouless, P. W. Anderson and R. G. Palmer, Phil. Mag. 35, 593 (1977).

707
23.2 The Sherrington-Kirkpatrick Solution.
The correct mean-field solution for the Sherrington-Kirkpatrick model can be
obtained in a systematic manner, starting from the partition function. Although
the average value of the partition function Z is easily evaluated, the average
value of the Free energy is difficult to evaluate. However, the logarithm of the
free energy can be evaluated with the aid of the mathematical identity
 n 
Z − 1
− β F = lim (2491)
n→0 n
For finite integer n the configurational average over Ji−j can be evaluated lead-
ing to an expression for the partition function for n replicas of the spin system
in which the replicas are interacting. The Gaussian averaged value of Z n is
given by
2
r
Y  N Ji−j
Z  
n N
Z = dJi−j exp − ×
i−j
2 π J2 2 J2
  X  n
× T race exp − β Ji−j Si Sj
i−j
2
r
Y  N Ji−j
Z  
N X
α α
= T race dJi−j exp − − β Ji−j Si Sj
i−j
2 π J2 2 J2 α

( β J )2
 X X 
Zn = T race exp Siα Sjα Siβ Sjβ (2492)
2N i,j α β

where α and β are the indices labelling members of the n different replicas. The
trace can be evaluated for integer n and then the result can be extrapolated to
n → 0. The spin glass order parameter is given by the correlation between the
spins of different replicas
q α,β = < | Siα Siβ | > (2493)
which becomes non-zero below the freezing temperature. The Free energy is
evaluated by re-writing the trace in terms of a Gaussian integral
( β J )2 X X α α β β
 
T race exp Si Sj S i Sj
2N i,j α β
√ !
( β J )2
Y Z dyα,β β J N   X
2 α β
= T race √ exp − N yα,β − 2 yα,β Si Si
2π 2 i
α,β

( β J )2 2
Z  
Y dyα,β β J N
= √ exp − N yα,β ×
2π 2
α,β
"  X #
2 α β
× exp N ln T race exp ( β J ) yα,β S S (2494)
α,β

708
In thie above expression, the thermodynamic limit N → ∞ and the limit n →
0 have been interchanged. Due to the long-ranged nature of the interaction,
the trace is over a single spin replicated n times. In the thermodynamic limit
N → ∞, this integral can be evaluated by steepest descents. The saddle
point value of y α,β is denoted by q α,β . For temperatures above Tf , it is easy
to show that the interaction part of the Free energy originates from the terms
with α = β as the off-diagonal terms of q α,β are all equal and zero. Therefore,
above the freezing temperature, the Free energy is found as
N
− β F = N ln 2 + ( β J )2 (2495)
2
Just below the spin glass freezing transition, the off-diagonal terms q α,β are all
equal and finite. Separating out the terms where α = β and replacing the
integrals by their saddle point values, the n-th power of the partition function
becomes
( β J )2
  
= 2 exp n N 1 − q 2 (n − 1) ×
2
"  X #
2 α β
× exp N ln T race exp (βJ ) qS S
α6=β
2
  
(βJ )
= exp nN 1 − 2 q − q 2 (n − 1) ×
2
"  X #
2 α β
× exp N ln T race exp (βJ ) qS S (2496)
α,β

where in the last line, the sum over pairs of replicas has been extended to include
the term α = β. By introducing a Gaussian integration, the trace over spins in
different replicas can be evaluated in terms of a trace of the spin in one replica
( β J )2
  
= exp n N 1 − 2 q − q 2 (n − 1) ×
2
" #
z2
Z    X
dz p α
× exp N ln T race √ exp − exp (βJ )z 2qS
2π 2 α
( β J )2
  
= exp n N 1 − 2 q − q 2 (n − 1) ×
2
" #
z2
Z   
dz
2n coshn ( β J ) z 2 q
p
× exp N ln √ exp −
2π 2
(2497)
The saddle point value of q is found by differentiating Z n with respect to q.
After an integration by parts and then clearing away fractions, one obtains
z2
  Z    
dz n
p
1 + q (n − 1) √ exp − cosh (βJ )z 2q
2π 2

709
z2
Z      
dz
coshn 1 + (n − 1) tanh2
p p
= √ exp − (βJ )z 2q (βJ )z 2q
2π 2
(2498)

In the limit n → 0, the order parameter is given by the solution of the equation
Z ∞
z2
 
dz p
q(T ) = √ exp − tanh2 β J 2 q(T ) z (2499)
−∞ 2π 2

The temperature variation of the order parameter is given by


  2 
1 T
q(T ) = 1 − for T < Tf
2 Tf
  12
2 T
lim q(T ) = 1 − (2500)
T → 0 3π TF

The finite value of the order parameter produces the cusp in the susceptibility
and the low-temperature saturation, since one can show that

g 2 µ2B
 
χ(T ) = 1 − q(T ) (2501)
3 kB T

Although the long-ranged model is exactly soluble in the mean-field approxi-


mation, the replica symmetric solution does not have the minimum value of the
Free energy. The model is only soluble for all temperatures below the freezing
temperature if the symmetry between the different replicas is broken. Replica
symmetry breaking is specific to interacting random systems285 , and the exact
solution of the mean-field model involves repeated replica symmetry breaking286 .
This repeated replica symmetry breaking has the consequence that the dynam-
ics of the low-temperature system are frozen and no longer consistent with the
ergodic hypothesis.

285 J.
R. L. de Almeida and D. J. Thouless, J. Phys. A, 11, 983, (1978).
286 G.Parisi, Phys. Rev. Lett. 43, 1754 (1979), G. Parisi, J. Phys. A, 13, L-115, 1101 and
1887 (1980).

710
24 Magnetic Neutron Scattering
The excitations of the electronic system can be probed by inelastic neutron scat-
tering experiments. These experiments provide information about the magnetic
character of the excitations, due to the nature of the interaction.

24.1 The Inelastic Scattering Cross-Section


The neutron scattering occurs through the interaction with the magnetic mo-
ments of the electronic system.

24.1.1 The Dipole-Dipole Interaction


A neutron has a magnetic moment given by

µn = gn µn σ n (2502)

where the neutrons gyromagnetic ratio is given by gn = 1.91 and interacts with
the magnetic moments of electrons via dipole-dipole interactions. The magnetic
field produced by a single electron moving with velocity v is a dipole field given
by  
ge µB σ e ∧ r |e| v ∧ r
H = ∇ ∧ − (2503)
| r |3 c | r |3
where r is the position of the field relative to the electron. The interaction
between the neutron and the magnetic field is given by the Zeeman interaction
"   #
σe ∧ r |e| v ∧ r
Ĥint = − gn µn σ n . ∇ ∧ ge µB −
| r |3 c | r |3
"  
σ ∧ r
= gn µn σ n . ∇ ∧ ge µB e 3
|r|
 #
|e| σn ∧ r σn ∧ r
− p. + .p (2504)
2 me c | r |3 | r |3

The first term is a classical dipole - dipole interaction and the second term is a
spin - orbit interaction.

24.1.2 The Inelastic Scattering Cross-Section


The scattering cross-section of a neutron, from an initial state (k, σn ) to a final
state (k 0 , σ 0n ), in which the electron makes a transition from the initial state

711
| φn > to the final state | φn0 > is given by
2 0  2 X
d2 σ
  
k V mn 0 0 σe ∧ r
= ge gn µn µB P (n) < φn0 ; k , σN | σ N . ∇ ∧

dω dΩ k 2 π h̄2 | r |3
n,n0
  2
1 σ ∧ r σn ∧ r
p. n 3

− + 3
. p | φn ; k, σ n > δ( h̄ ω + En − En0 )
2 h̄ |r| |r|
(2505)
Here, the probability that the electronic system is in the initial state is rep-
resented by P (n). The neutron’s energy loss h̄ ω and the momentum loss or
scattering vector h̄q are defined via
h̄ ω = E(k) − E(k 0 )

h̄ q = h̄ k − h̄ k 0 (2506)
As the neutron states are momentum eigenstates, the matrix elements of the
interaction can be easily evaluated. The spin component of the magnetic inter-
action is evaluated by considering the neutron component of the matrix elements
 
0 σe ∧ r
<k |∇ ∧ |k >
| r |3
 
0 1
= − < k |∇ ∧ σe ∧ ∇ |k >
|r|
Z    
1 3 1
= d rn exp + i q . rn ∇ ∧ σe ∧ ∇
V |r|
   

= q ∧ ( σe ∧ q ) exp + i q . re (2507)
V q2
This shows that the neutron only interacts with the component of the electron’s
spin σ perpendicular to the scattering vector. Likewise, the orbital component
can be evaluated as
     
0 σe ∧ r 4πi
< k | pe ∧ |k >= − σ e ∧ ( q ∧ pe ) exp + i q . re
| r |3 V q2
(2508)
 
Furthermore, the operator ( q ∧ pe ) commutes with exp + i q . re as
q ∧ q ≡ 0. Hence, the neutron scattering cross-section from a multi-electron
system can be written as
2 0  2 X
d2 σ

k 2 mn
= ge gn µn µB P (n) δ( h̄ ω + En − En0 )
dω dΩ k h̄2 n,n0
2
σe ∧ q i q ∧ pe
X      
0

× < φn0 ; σn | σ n .
q ∧ 2
− 2
exp + i q . re | φn ; σn >
e
| q | h̄ | q |
(2509)

712
Since the nuclear Bohr magneton has the value

| e | h̄
µn = (2510)
2 mp c

the coupling constant can be simplified

gn e2
 
2 mn
2 g g µ µ
n e n B = = re (2511)
h̄ me c2

to yield re , the classical radius of the electron. Thus, the scattering cross-section
can be written as
d2 σ k0
= re2 S(q; ω) (2512)
dω dΩ k
where the response function is given by
X
S(q; ω) = P (n) δ( h̄ ω + En − En0 )
n,n0
2
σe ∧ q i q ∧ pe
X      
0

× < φn0 ; σn | σ n .
q ∧ − exp + i q . re | φn ; σn >
e
| q |2 h̄ | q |2
(2513)

This expression still depends on the polarization of the neutrons in the incident
beam, and also on the polarization of the detector. Polarized neutron scattering
measurements reveal more information about the nature of the excitations of
a system. However, due to the reduction of the intensity of the incident beam
caused by the polarization process, and the concomitant need to compensate
the loss of intensity by increase the measurements time, it is more convenient
to perform measurements with unpolarized beams. For an un-polarized beam
of neutrons, the initial polarization must be averaged over. The averaging can
be performed with the aid of the identity
X 1
< σn | σnα σnβ | σn > = δα,β (2514)
σ
2
n

which follows from the anti-symmetric nature of the Pauli spin matrices. For
an un-polarized beam of neutrons the response function reduces to
X X  
S(q; ω) = P (n) δ( h̄ ω + En − En0 ) δα,β − q̂α q̂β
n,n0 α,β
X  i q ∧ pe
  
× < φn | σe + exp − i q . re | φn0 >
e
h̄ | q |2 α
X  i q ∧ pe
  
× < φn0 | σe − exp + i q . re | φn >
e
h̄ | q |2 β
(2515)

713
where q̂ is the unit vector in the direction of q. On defining the spin density
operator Ŝα (q) via
X  i q ∧ pe
  
Ŝα (q) = σe + exp − i q . r e (2516)
e
h̄ | q |2 α

then the response function can be expressed as a spin - spin correlation function
X
S(q; ω) = P (n) δ( h̄ ω + En − En0 ) ×
n,n0
X  
× δα,β − q̂β q̂β < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >
α,β
(2517)
Thus, the inelastic neutron scattering measures the excitation energies of the
system, with intensity governed by the matrix elements < φn | Ŝα (q) | φn0 >
which filters out the excitations of a non-magnetic nature. Furthermore, the
scattering only provides information about the magnetic excitations which have
a component of the fluctuation perpendicular to the momentum transfer.

In the case where the spin density can be expressed in terms of the atomic
spin density due to the unpaired spins in the partially filled shells, such as
in transition metals or rare earths, it is convenient to introduce the magnetic
atomic (ionic) form factor F (q). For a mono-atomic Bravais lattice, this is
achieved by decomposing the spin density in terms of the spin density from
each unit cell
X  i q ∧ pe
  
Ŝα (q) = σe + exp − i q . re
e
h̄ | q |2 α

i q ∧ pj
X   X    
= exp − i q . R σj + exp − i q . rj
j
h̄ | q |2 α
R

(2518)
Since the unpaired electrons couple together to give the ionic spin ŜR , the
Wigner - Eckert theorem can be used to express the spin density operator as
X  
Ŝα (q) = exp − i q . R F (q) ŜR (2519)
R

The form factor F (q) is defined as the Fourier transform of the normalized spin
density for the ion. By definition, the form factor is normalized by
F (0) = 1 (2520)
In this case, the inelastic neutron scattering spectrum can be expressed as
d2 σ k0
= re2 | F (q) |2 S(q; ω) (2521)
dω dΩ k

714
where the spin - spin correlation function is expressed in terms of the local ionic
spins ŜR . Of course, it is being implicity assumed that the magnetic scattering
can be completely separated from the phonon scattering. Thus, the analysis has
ignored the existence of phonon excitations, in the case of zero phonon excita-
tions, the intensity of the magnetic scattering is expected to be reduced by the
Debye-Waller factor of the phonons.

24.2 Time-Dependent Spin Correlation Functions


The spin-dependent correlation function measured in scattering experiments is
denoted by S α,β (q; ω) and is defined as
X
S α,β (q; ω) = P (n) δ( h̄ ω + En − En0 ) ×
n,n0
" #
× < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >

(2522)

and where P (n) is the probability that the system is found in the initial state
| φn >. Using the expression for the energy conserving delta function as an
integral over a time variable
Z ∞  
dt i
δ( h̄ ω + En − En0 ) = exp ( h̄ ω + En − En0 ) t
−∞ 2 π h̄ h̄
(2523)

the spin - spin correlation function can be written as a Fourier transform of a


time-dependent correlation function.
Z ∞    
α,β 1 X dt i
S (q; ω) = P (n) exp i ω t exp ( En − E n0 ) t
h̄ −∞ 2 π h̄
n,n0
" #
× < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >
Z ∞    
dt 1 X i
= exp iωt P (n) exp ( E n − En0 ) t
−∞ 2π h̄ h̄
n,n0
" #
× < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn > (2524)

The product of the phase factor and the matrix elements of the Ŝ(q) can be
expressed in terms of an operator in the interaction representation
 
i
exp ( E n − En ) t
0 < φn | Ŝα (q) | φn0 >

715
J. Appl. Phys., Vol. 85, No. 8, 15 April 1999 Goremychkin

FIG. 2. ~a! Energies of crystal field leve


W fixed to 1. ~b! Anisotropy of the mag
v x at T541 K.

FIG. 1. Inelastic neutron scattering data from CeAl3 measured on the HET e
S ~ Q, e ! 5 f
Figurespectrometer
300: The spinwith correlation
an incident energy of 35S(Q,
function meV.)The lines are
inferred the inelastic
from results of neutron 12exp~ 2 e /k B T !
a fit toexperiments
scattering a three-component
on the model as described compound
paramagnetic in Ref. 7. CeAl3 . The spectrum
2 13
where f (Q) is the Ce form f
shows a quasi-elastic peak and an inelastic peak (near 7 meV) due to crystal
Lorentzian
field excitations. [After E. A. Goremychkin, R. Osborn and I. L. Sashin, J. function characterizin
Appl. Phys. 85, 6046 (1999).].
In the experiments on KDSOG-M, approximately 200 g, sition, and x 0 is the static bulk su
of CeAl3 and LaAl3 were sealed in aluminum containers and in Fig. 1 are the calculated S(Q,
placed in a helium cryostat for measurements at 8 K. A py- components of the magnetic resp
rolytic graphite monochromator and beryllium filter in front In the case of interaction of t
of the detector fixes the final energy at 4.8 meV. The INS the static susceptibility x 0 is giv
was measured in neutron energy loss up to 80 meV energy and Van Vleck x mn VV contribution
transfer. The resolution for the
716 elastic scattering was 0.6
meV and the spectra were summed over three scattering x 05 (n x nC 1 mÞn
( x mn
VV ,

angles: 30°, 50°, and 70°.


g 2J m 2B
Comparison of the data for CeAl3 and LaAl3 measured
on KDSOG-M and HET, as well as inspection of the HET
x C5
n
k BT
rn (
a 5x,y,z
u ^ n u J au n
data taken at low ~^f&519°! and high ~^f&5136°! scattering
( a 5x
VV 52g J m B ~ r n 2 r m !
angles, shows that there is one well-defined inelastic mag- x mn 2 2
   
i i
= < φn | exp + Ĥ0 t Ŝα (q) exp − Ĥ0 t | φn0 >
h̄ h̄
= < φn | Ŝα (q; t) | φn0 > (2525)

Hence, the spin - spin correlation function is given by


Z ∞  
dt
S α,β (q; ω) = exp i ω t ×
−∞ 2 π
" #
1 X †
× P (n) < φn | Ŝα (q; t) | φn0 > < φn0 | Ŝβ (q; 0) | φn >
h̄ 0
n,n
(2526)

The final states are a complete set of states, therefore, on using the completeness
relation, one finds
Z ∞  " #
dt 1 X †
S α,β (q; ω) = exp i ω t P (n) < φn | Ŝα (q; t) Ŝβ (q; 0) | φn >
−∞ 2 π h̄ n
Z ∞  
dt 1
= exp i ω t < | Ŝα (q; t) Ŝβ† (q; 0) | >
−∞ 2 π h̄
(2527)

The correlation function S α,β (q; ω) is the Fourier Transform with respect to
time of the thermal averaged spin - spin correlation function. The inverse spatial
Fourier transform of the spin density operator and its Hermitean conjugate are
given by
Z  
1 3
Ŝα (q) = d r exp − i q . r Ŝα (r)
V
Z  
† 1 3 0 0
Ŝα (q) = d r exp + i q . r Ŝα (r0 ) (2528)
V

On inserting the above expressions into S α,β (q; ω) and using the spatial homo-
geneity of the system, one finds that the inelastic neutron scattering spectrum
is related to the spatial and temporal Fourier transform of the spin - spin cor-
relation function
Z ∞ Z  
α,β 1 dt 3 1
S (q; ω) = d r exp i ( ω t − q . r ) < | Ŝα (r; t) Ŝβ (0; 0) | >
V −∞ 2 π h̄
(2529)

where the brackets < | . . . | > represents the quantum mechanical thermal
average. Thus, the inelastic neutron scattering probes the Fourier transform of
the equilibrium spin correlation functions.

717
24.3 The Fluctuation - Dissipation Theorem
The spin - spin correlation function
P (n) δ( h̄ ω + En − En0 ) < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >
X
S α,β (q; ω) =
n,n0
(2530)
is consistent with the principle of detailed balance. If the equilibrium probability
P (n) is given by the Boltzmann expression
 
1
P (n) = exp − β En (2531)
Z
then spin-spin correlation function can be re-written as
P (n) δ( h̄ ω + En − En0 ) < φn0 | Ŝβ† (q) | φn > < φn | Ŝα (q) | φn0 >
X
S α,β (q; ω) =
n,n0
  X
= exp β h̄ ω P (n0 ) δ( En0 − En − h̄ ω ) < φn0 | Ŝβ† (q) | φn > < φn | Ŝα (q) | φn0
n,n0
(25
where we have used the identity exp[−βEn ] δ(h̄ω+En −En0 ) = exp[βh̄ω] exp[−βEn0 ] δ(h̄ω+
En − En0 ). On interchanging the summation indices n and n0 , one finds
  X
= exp β h̄ ω P (n) δ( En − En0 − h̄ ω ) < φn | Ŝβ† (q) | φn0 > < φn0 | Ŝα (q) | φn >
n,n0
 
= exp β h̄ ω S β,α (−q; −ω)

(2533)
which is consistent with the principle of detailed balance for equilibrium pro-
cesses.

The correlation function S α,β (q; ω) is also related to the imaginary part of
the magnetic susceptibility χα,β (q; ω) via the fluctuation dissipation theorem.

The reduced dynamical magnetic susceptibility is given by the expression


 
0 0 i
α,β
χ (r, r ; t − t ) = − < | Ŝα (r, t) , Ŝβ (r , t ) | > Θ( t − t0 ) (2534)
0 0

which can be expressed as
 
i X i
χα,β (r; t) = − P (n) exp ( En − En0 ) t < φn | Ŝα (r) | φn0 > < φn0 | Ŝβ (0) | φn > Θ
h̄ h̄
n,n0
 
i X i
+ P (n) exp ( En0 − E n ) t < φn | Ŝβ (0) | φn0 > < φn0 | Ŝα (r) | φn > Θ
h̄ 0

n,n
(2

718
The Fourier transform is defined as
Z ∞ Z  
α,β 1 dt
χ (q; ω) = d r exp i ( ω t − q . r ) χα,β (r; t)
3
V −∞ 2 π
(2536)

and is evaluated as

< φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >
"
α,β 1 X
χ (q; ω) = P (n)
2π 0
h̄ ω + En − En0 + i δ
n,n

< φn | Ŝβ† (q) | φn0


#
> < φn0 | Ŝα (q) | φn >
− (2537)
h̄ ω + En0 − En + i δ

The imaginary part of the dynamic susceptibility is given by


 
1 X
=m χα,β (q; ω) = − P (n) ×
2
n,n0
"
× δ( h̄ ω + En − En0 ) < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >
#
− δ( h̄ ω + En0 − En ) < φn | Ŝβ† (q) | φn0 > < φn0 | Ŝα (q) | φn >

(2538)

The first term is recognized as being S α,β (q; ω). On replacing P (n) by P (n0 ) exp[−βh̄ω]
in the second term and interchanging the summation indices n and n0 , one recog-
nizes that the second term is just proportional to exp[−βh̄ω] S α,β (q; ω). Hence,
we have found that
  "  #
α,β 1 α,β
=m χ (q; ω + iδ) = − S (q; ω) 1 − exp − β h̄ ω
2
(2539)

or    
S α,β (q; ω) = 2 1 + N (ω) =m χα,β (q; ω − iδ) (2540)

This is the fluctuation - dissipation theorem287 . The fluctuation - dissipation


theorem relates the dynamical response of the system to an external perturba-
tion to the naturally occurring excitations in the system, such as those measured
in neutron scattering experiments.

287 H. B. Callen and T. A. Welton, Phys. Rev. 83, 34 (1951) and R. Kubo, J. Phys. Soc.

Japan 12, 570 (1957).

719
24.4 Magnetic Scattering
The neutron scattering cross-section is given in terms of the components of the
spin spin correlation function.

As can be seen by inspection from the Holstein-Primakoff representation of


the spins and the spin waves, the spin correlation function is a non-linear func-
tion of the spin wave creation operators. The inelastic scattering cross-section
can be expanded in powers of the number of spin waves. The lowest-order term
is time-independent and corresponds to Bragg scattering.

24.4.1 Neutron Diffraction


The ω = 0 component of the inelastic scattering cross-section given by the
ω → 0 limit of
Z ∞  
dt 1
S α,β (q; ω) = exp i ω t < | Ŝα (q; t) Ŝβ† (q; 0) | > (2541)
−∞ 2 π h̄

diverges if the integrand does not decay rapidly as t → ∞. In this case, the
time-independent component of the spin - spin correlation function given by
1
lim < | Ŝα (q; t) Ŝβ† (q; 0) | > (2542)
t → ∞ h̄
produces a Bragg peak with finite intensity since
Z ∞  
dt
δ(ω) = exp i ω t (2543)
−∞ 2 π

Thus, the intensity of the Bragg peak represents the static correlations. If the
ergodic hypothesis holds, then in the long time limit the correlation function
decouples into the product of two expectation values
1
lim < | Ŝα (q; t) Ŝβ† (q; 0) | >
t → ∞ h̄
1
= lim < | Ŝα (q; t) | > < | Ŝβ† (q; 0) | > (2544)
t → ∞ h̄

and for a static system for which

< | Ŝα (q; t) | > = < | Ŝα (q; 0) | > (2545)

then the Bragg peak has an intensity given by


1
= lim < | Ŝα (q; 0) | > < | Ŝβ† (q; 0) | > (2546)
t → ∞ h̄
For a paramagnetic system the (quasi-stationary) average value of the spin

720
Figure 301: Left: The neutron diffraction pattern of M nO at temperatures
below and above the magnetic ordering temperature. Right: The antiferromag-
netic structure of M nO below the Neel temperature. [After C. G. Shull, W. A.
Strauser and E. O. Wollan, Phys. Rev. 83, 333 (1951).].

vector is zero
< | Ŝα (q; 0) | > = 0 (2547)
and, thus, there is no magnetic Bragg peak for a paramagnetic system. On the
other hand, if there is long-ranged magnetic order with wave vectors Q and with
spin oriented along certain directions (say α), then

< | Ŝα (Q; 0) | > 6= 0 (2548)

thus, the magnetic Bragg peaks are non-zero. The temperature dependence of
the intensity of the Bragg peaks provides a direct measure of the temperature
dependence of the magnetic order parameter. For a ferromagnet, the magnetic
Bragg peaks coincide with the Bragg peaks due to the crystalline order, so no
new peaks emerge. The Bragg scattering cross-section is given by
  2 X
  2
dσ 2 ( 2 π N ) 2

= re 1 − q̂z δ( q − Q ) < | Sz | >

dΩ Bragg V
Q
(2549)
For a small single domain single crystal, the magnetic elastic scattering is ex-
tremely anisotropic as the scattering should be zero for momentum transfers
along the direction of the magnetization.

For anti-ferromagnetic or spin density wave order new Bragg peaks may
emerge at the vectors of the anti-ferromagnetic reciprocal lattice. Analysis of
the anisotropy of the neutron scattering intensity for anisotropic single crystals

721
Figure 302: The temperature-dependence of the antiferromagnetic order param-
eter for M nO. [After C. G. Shull, W. A. Strauser and E. O. Wollan, Phys. Rev.
83, 333 (1951).].

leads to the determination of the preferred directions of the magnetic moments.

——————————————————————————————————

24.4.2 Exercise 94
Evaluate the elastic scattering cross-section for a anti-ferromagnetic insulator,
using the Holstein-Primakoff representation of the low-energy spin wave exci-
tations. Discuss the anisotropy and also the temperature dependence of the
intensity of the Bragg peaks.

——————————————————————————————————

24.4.3 Exercise 95
Design a neutron diffraction experiment that will determine if a system has a
spiral spin density wave order, as opposed to a magnetic moment that is mod-
ulated in intensity. How can the direction of spiral be determined?

722
——————————————————————————————————

24.4.4 Spin Wave Scattering


The spin wave excitations of an ordered magnet show up in the inelastic neutron
scattering spectra. In a process whereby a single spin wave is emitted in the
scattering process, conservation of energy leads to the energy difference between
the initial state En and the final state En0 being given by
En0 − En = h̄ ωq (2550)
The matrix elements in the spin - spin correlation function can be evaluated as
Y Y
< φn | Sα (q) | φn0 > = < nq0 | Sα (q) | n0q” > (2551)
q0 q”

where the number of spin waves in the initial state are related to the number
in the final state via
n0q0 = nq0 for q 0 6= q (2552)

and
n0q = nq + 1 (2553)

For a ferromagnet, the matrix elements are evaluated as


< nq | Sz (q) | nq + 1 > = 0 (2554)
while the transverse matrix elements are
1
< nq | Sx (q) | nq + 1 > = < nq | ( S + (q) + S − (q) ) | nq + 1 >
2
1 q √
= nq + 1 2S (2555)
2
and
1
< nq | Sy (q) | nq + 1 > = < nq | ( S + (q) − S − (q) ) | nq + 1 >
2i
1 q √
= + nq + 1 2S (2556)
2i
Thus, the inelastic neutron scattering from the single spin wave excitations of
a ferromagnet is given purely by the diagonal components of the transverse
spin-spin correlation function, as the longitudinal components are zero. The
off-diagonal terms cancel. The cross-section for the spin wave emission process
is given by
 2  2
  X  
d σ 2 ( 2 π) N S 2 0 0 0
= re 1 + q̂z δ( h̄ ω − h̄ ω(q ) ) δ( q − q −Q ) 1 + N (ω(q ))
dω dΩ emit V 2 0
q ,Q
(2557)

723
Likewise, the absorption process has a scattering cross-section given by
 2
( 2 π)2 N S
   X
d σ
= re2 1 + q̂z2 δ( h̄ ω + h̄ ω(q 0 ) ) δ( q − q 0 −Q ) N (ω(q 0 ))
dω dΩ abs V 2 0 q ,Q
(2558)
The intensities of the emission and absorption processes are consistent with the
principle of detailed balance (only valid for total equilibrium), and give rise to
a Stokes and anti-Stokes line in the spectrum of the scattered neutrons.

——————————————————————————————————

24.4.5 Exercise 96
Evaluate the two lowest-order terms in the inelastic scattering cross-section for
an anti-ferromagnetic insulator, using the Holstein-Primakoff representation of
the low-energy spin wave excitations. Discuss the differences between the spec-
trum obtained from magnetic scattering and that found in measurements of the
phonon excitations.

——————————————————————————————————

24.4.6 Critical Scattering


Just above the temperature where magnetic ordering occurs, the inelastic neu-
tron scattering cross-section in the paramagnetic phase shows a softening or
build up at low frequencies and becomes sharply peaked at q values close to
the magnetic Bragg vectors Q. Below the ordering temperature, the intensity
transforms into the Bragg peak. This phenomenon of the build up of inten-
sity close to the Bragg peak is known as critical scattering. The Bragg peak
is extracted from the inelastic scattering spectrum by extracting a delta func-
tion δ(ω), i.e., the inelastic scattering cross-section is integrated over a small
window dω. On invoking the fluctuation dissipation theorem and then noting
that if, in the paramagnetic phase, the main portion of the scattering occurs
with frequencies such that β ω  1, then one finds that by using the Kramers
- Kronig relation, the intensity of the critical scattering is given by the static
susceptibility. For q values close to the Bragg peak, the susceptibility varies as

1
∝ (2559)
( q − Q )2 + ξ −2

where ξ the correlation length, in the mean-field approximation, is given by


s
−1 Tc
ξ = a (2560)
6 ( T − Tc )

724
Thus, the critical scattering diverges as ( T − Tc )−1 as the transition temper-
ature is approached.

725
Figure 303: The electrical resistance of Hg as a function of temperature. [After
H. Kammerlingh Onnes (1911).].

25 Superconductivity
The electrical resistivity ρ(T ) of metals at low temperatures is expected to be
described by the Drude model
m
ρ(T ) ∝ (2561)
e2 τ
The resistivity should vary with temperature according to

ρ(T ) ∼ ρ(0) + A T 2 + B T 5 (2562)

since the scattering rates for scattering from static impurities, electron-electron
scattering and phonon scattering are expected to be additive. The resistivity
of a perfect metal, without impurities, may be expected to vanish at T = 0.
However, it was discovered by Kammerlingh Onnes288 that the resistivity of
a metal may become so small as to effectively vanish for all temperatures be-
low a critical temperature Tc . This indicates that the scattering mechanisms
suddenly becomes ineffective for temperatures slightly below the critical tem-
perature, where the metal seems to act like a perfect conductor. The resistivity
is so small that persistent electrical currents have been observed to flow without
attenuation for very long time periods. The decay time of a super-current in
favorable materials is apparently not less than 10,000 years.

288 H. Kammerlingh Onnes, Comm. Phys. Lab. Univ. Leiden, Nos. 119, 120, 122 (1911).

726
B B

N S

T > Tc T < Tc

Figure 304: A schematic depiction of the Meissner effect. The magnetic induc-
tion field B is excluded from the bulk of a superconducting sample, in contrast
with the normal state.

25.1 Experimental Manifestation


The first manifestation of superconductivity is zero resistance, below Tc . An-
other manifestation of superconductivity was found by Meissner and Ochsen-
feld289 , which is flux exclusion. A superconductor excludes the magnetic induc-
tion field B from its interior, irrespective of whether it was cooled from above
Tc to below Tc in the presence of an applied field, or whether the field is only
applied when the temperature is smaller than Tc . In other words, the Meissner
effect excludes time independent magnetic field solutions from inside the su-
perconductor. The Meissner effect distinguishes superconductivity from perfect
conductivity, as a static magnetic field can exist in perfect conductor.

The perfect conductor has the property that the current produced by an
applied electric field increases linearly with time. Therefore, a perfect conductor
excludes electric fields from within its bulk. Maxwell’s equations reduce to
1 ∂B
− = 0
c ∂t

∇ ∧ B = j
c
∇.B = 0 (2563)

Thus, a perfect conductor only excludes a time varying magnetic field, but not
a static magnetic field.

The Meissner effect shows hat the magnetic induction inside a superconduc-
tor is zero. However, the magnetic induction B can be expressed in terms of
the applied field H and the magnetization M via

B = H + 4πM (2564)
289 W. Meissner and R. Ochsenfeld, Naturwiss. 21, 787 (1933).

727
Type I

Type II

-4πM Hc1 Hc Hc2 H

Figure 305: A schematic depiction of the dependence of the (diamagnetic) mag-


netization on applied magnetic field for a type I and a type II superconductor.

so B = 0 implies that
1
M = − H (2565)

so that perfect diamagnetism implies that the susceptibility is given by
1
χ = − (2566)

The perfect diamagnetism does not hold for arbitrarily large applied magnetic
fields. For fields larger than a critical magnetic field, the induction inside the
superconductor becomes non-zero. For a type I superconductor, the applied
field fully penetrates into the bulk of the superconductor above the critical field
Hc . The magnetization drops discontinuously to zero at Hc . The value of Hc
depends on temperature according to

T2
 
Hc (T ) = Hc (0) 1 − 2 (2567)
Tc

For a type II superconductor, the induction first starts penetrating into the
bulk at the lower critical field Hc1 , For fields larger than the lower critical field,
the magnetization deviates from linear relation associated with perfect diamag-
netism. The magnitude of the magnetization is reduced as the applied field is
increased above Hc1 . The magnetization falls to zero at the upper critical field
Hc2 , at which point the applied field fully penetrates into the bulk.

The experimental observations of a drop in the resistivity and the Meissner


effect demonstrate that the transition to the superconducting state is a phase
transition as the properties are independent of the history of the sample. For
a type I superconductor, the bulk superconductivity is completely destroyed at
Hc (T ).

728
25.1.1 The London Equations
A phenomenological description of superconductivity was developed by the Lon-
don brothers290 . Basically, this description is based on two phenomenological
constitutive equations for the electromagnetic field and its relation to current
and density. The first London equation is of the form

ns e2
j(r, t) = − A(r, t) (2568)
mc
which expresses the extreme diamagnetic response of a superconductor. Here,
ns is density of superfluid electrons. It should be noted that London’s first
equation is not gauge invariant. The vector potential in London’s first equation
has to be calculated with a specific choice of gauge condition. If the supercon-
ducting current is to be conserved, one must require that ∇ . A = 0. The
London equation describes the microscopic current in the superconductor that
screens the applied magnetic field. This is slightly different from the condition
of perfect conductivity in a metal. In a perfect conductor, the time derivative
of the current is related to the electric field. In deriving the London equation
from the condition of perfect conductivity, it has been assumed that the electric
field is transverse. The condition for perfect conductivity has been integrated
with respect to time, and the constant of integration has been chosen to be zero.
The choice of the constant of integration allows a constant current to screen the
static applied magnetic field. In order that the continuity equation be satisfied
in a steady state, a gauge condition must be imposed such that ∇ . A(r, t) = 0
and one also requires that the perpendicular component of A vanish at the sur-
face. This gauge condition defines the London gauge.

The second London equation comes from Maxwell’s equations

4 π j(r, t) 1 ∂
∇ ∧ B(r, t) = + E(r, t) (2569)
c c ∂t
and with the definitions

B(r, t) = ∇ ∧ A(r, t)
1 ∂
E(r, t) = − A(r, t) (2570)
c ∂t
one finds
1 ∂2 4 π ns e2
 
∇ ∧ ∇ ∧ + A(r, t) = − A(r, t) (2571)
c2 ∂t2 m c2
2
This is referred to as the second London equation. The quantity nms ce2 has units
of inverse length squared and is used to define the London penetration depth
290 F. London and H. London, Proc. Roy. Soc. (London), A 149, 71 (1935), F. London,

Phys. Rev. 74, 562 (1948).

729
λL , via
4 π ns e2 1
= (2572)
m c2 λ2L
The second London equation expresses the Meissner effect. Namely, that
a superconductor excludes the magnetic induction field B from the bulk of its
volume. However, the field does penetrate the region at the surface and extends
over a distance λL into the superconductor. This can be seen by examining
various cases in which a static applied magnetic field is produced near a super-
conductor. The geometry is considered in which the applied field is parallel to
the surface.

Let the surface be the plane z = 0, which separates the superconductor


z > 0 from the vacuum z < 0. The external field is applied in the x direction,
so B = B0 x̂ for z < 0. The vector potential inside the superconductor must
satisfy the boundary condition Az (z = 0) = 0 as any current should be per-
pendicular to the surface. The London gauge requires the non-zero components
of the vector potential to be Ax and Ay . Thus, the vector potential must be
parallel to the surface. The static solution for the vector potential that satisfies
the boundary conditions on the current for the semi-infinite solid is
 
z
A(z) = A0 exp − (2573)
λL
An additional boundary condition at z = 0 is that Bx should be continuous.
Hence, as the equation for the magnetic induction simplifies to
∂Ay (z)
Bx (z) = − (2574)
∂z
one finds that the vector potential is directed parallel to the surface, but is also
perpendicular to the applied field. The only non-zero component of the vector
potential in the superconductor is found as
 
z
Ay (z) = + λL Bx (0) exp − for z > 0 (2575)
λL
London’s first equation then implies that a supercurrent, jy (z), flows in a region
near the surface of the superconductor which, through Ampere’s law, produces
magnetic field that screens or cancels the applied field. The magnetic induction
and the supercurrent are non-zero in the superconductor only within a distance
of λL from the surface. Hence, λL is called the penetration depth.

25.1.2 Thermodynamics of the Superconducting State


The phase transition to a superconducting state, in zero field, is a second or-
der phase transition. This can be seen by examining the specific heat which

730
1.2
Vacuum Superconducto
1 r
Ay(z) = A0
0.8
Ay(z) = A0 exp [ -z / λ ]

Ay(z)
0.6

0.4

0.2

0
-1 -0.75 -0.5 -0.25 0 0.25 0.5 0.75 1

z/λ

Figure 306: The spatial dependence of the vector potential A showing that the
magnetic field only penetrates the superconductor over a distance of the order
of λ.

exhibits a discontinuous jump at Tc . The absence of any latent heat implies


that the entropy is continuous, and since the entropy is obtained as a first order
derivative of the Free energy the transition is not first order. The non-analytic
behavior of the specific heat, which is obtained from a second derivative of the
free energy defines the transition to be second order.

In the presence of a field the transition is first order. The thermodynamic


relations are derived from the Gibbs free energy G in which M plays the role of
the volume V and the externally applied field H plays the role of the applied
pressure P . Then, G(T,H) has the infinitesimal change
dG = − S dT − M . dH (2576)
where S is the entropy and T is the temperature. Since G is continuous across
the phase boundary at (Hc (T ), T )
Gn (T, Hc (T )) = Gs (T, Hc (T )) (2577)
which on taking an infinitesimal change in both T and H = Hc (T ) so as to stay
on the phase boundary, one finds
( Ss − Sn ) dT = ( Mn − Ms ) dHc (T ) (2578)
The magnetization in the normal state is negligibly small, but the supercon-
ducting state is perfectly diamagnetic, so
1
Mn − Ms = +Hc (T ) (2579)

This shows that in the presence of an applied field the superconducting transi-
tion involves a latent heat, L, given by
L = T ( Sn − Ss )
T ∂Hc
= − Hc (2580)
4π ∂T

731
Thus, the transition is first order in the presence of an applied field.

In the limit that the critical field is reduced to zero, the transition tempera-
ture T is reduced to the zero field value Tc , and the entropy becomes continuous
at the transition. However, there is a change in slope at Tc , which can be found
by taking the derivative of Sn − Ss with respect to T , and then letting Hc → 0
"  2   2 #
∂ 1 ∂ Hc ∂Hc
( Sn − Ss ) = − Hc +
∂T 4π ∂T 2 ∂T
 2
1 ∂Hc
= − (2581)
4π ∂T
The specific heat may show a discontinuity or jump at Tc that is a measure of
the initial slope of the critical field
 2
T ∂Hc
Cs − Cn = (2582)
4π ∂T
The discontinuous jump in the (zero field) specific heat is a characteristic of a
mean-field transition. For temperatures below Tc the specific heat is exponen-
tially activated  

Cv ∼ γ Tc exp − (2583)
kB T
The activated exponential behavior of the specific heat suggests that there is an
energy gap in the excitation spectrum. The existence of a gap is confirmed by
a threshold frequency for photon absorption by a superconductor. Above Tc ,
the absorption spectrum is continuous and photons of arbitrarily low frequency
can be absorbed by the metal. However, for temperatures below Tc , there is
a minimum frequency above which photons can be absorbed. The threshold
frequency is related to ∆.

In most superconductors, the interaction mechanism that is responsible for


pairing is mediated by the electron-phonon coupling. This was first identi-
fied through the insight of Fröhlich291 , who predicted that the superconducting
transition temperature Tc should be proportional to the phonon frequency. Fur-
thermore, as the square of the phonon frequency is inversely proportional to the
mass of the ions M , the superconducting transition temperature should depend
upon the isotopic mass through
1
Tc ∝ M − 2 (2584)
This isotope effect was confirmed in later experiments by Maxwell 292 and
Reynolds et al.293 on simple metals. However, in transition metals the ex-
ponent of the isotope effect is reduced and may become zero, and in α − U the
291 H. Fröhlich, Phys. Rev. 79, 845 (1950).
292 E. Maxwell, Phys. Rev. 78, 447 (1950), Phys. Rev. 79, 173 (1950).
293 C. A. Reynolds, B. Serin, W. H. Wright and L. B. Nesbitt, Phys. Rev. 78, 487 (1950).

732
Figure 307: The dependence of the superconducting transition temperature Tc
of Sn on the isotopic mass M on a log-log plot. The critical temperature is
assumed to vary as Tc ∝ M −α . The two straight lines yield α ≈ 0.487 and
α ≈ 0.505. [After E. Maxwell, Phys. Rev. 86, 25 (1952).].

exponent is positive. The occurrence of a positive isotope effect does not nec-
essarily signify the existence of alternate pairing mechanisms, but can indicate
the effect of strong electron-electron interactions.

25.2 The Cooper Problem


The electron-electron interaction in a metal, is attractive at low frequencies.
The attractive interaction originates from the screening of the electrons by the
ions, but only occurs for energy transfers less than h̄ ωD . The effective attrac-
tion is retarded, and occurs due to the attraction of a second electron with the
slowly evolving polarization of the lattice produced by the first electron. Cooper
showed that two electrons, which are close to the Fermi energy, will bind into
pairs whenever they experience an attractive interaction, no matter how weak
the interactions is294 .

Consider a pair of electrons of spin σ and σ 0 , excited above the Fermi en-
ergy. Due to the interaction between the pair of particles, the center of mass
momentum q will be a constant of motion, but not the relative motion. Thus,
the wave function of the Cooper pair with total momentum q can be written as
X
| Ψq > = C(k) | σ, k + q σ 0 , −k > (2585)
k

Due to the Pauli exclusion principle the single-particle energies E(−k) and E(k+
q) must both be above the Fermi energy µ. The wave function is normalized
294 L. Cooper, Phys. Rev. 104, 1189 (1956).

733
k' k

µ µ+hωD
kF

-k -k'

Figure 308: The scattering process involved in the pairing problem considered
by Cooper. Two electrons in states (σ, k) and (−σ, −k) above the Fermi-surface
are scattered into states (σ, k 0 ) and (−σ, −k 0 ), if all the states are within an
energy of h̄ωD of the Fermi energy.

such that X
| C(k) |2 = 1 (2586)
k

The wave function must be an energy eigenstate of the Hamiltonian Ĥ and,


thus, satisfies
Ĥ | Ψq > = E(q) | Ψq > (2587)
where E(q) is the total energy of the pair of electrons. On projecting out C(k)
using the orthogonality of the different momentum states | σ, k + q σ 0 , −k > ,
one finds the secular equation
 
1 X
E(q) − Ek − Ek+q C(k) = − V (k, k 0 ) C(k 0 ) (2588)
N 0 k

The attractive pairing potential V , ( − V < 0 ), scatters the pairs of electrons


between states of different relative momentum. The summation over k 0 is re-
stricted to unoccupied Bloch states within h̄ ωD the Fermi surface, where the
interaction is attractive. The above equation has a solution for the amplitude
C(k) which is given by

α(k)
C(k) = (2589)
E(q) − Ek − Ek+q

where α is given by
1 X
α(k) = − V (k, k 0 ) C(k 0 ) (2590)
N 0 k

This equation can be solved analytically in the case where the potential is sepa-
rable, such as the case where V is just a constant. In such cases, the summation

734
over k 0 can be performed to yield a result which is independent of k. For sim-
plicity, the separable potential shall be assumed to be attractive and have a
magnitude of V when both k and k 0 are within h̄ ωD of the Fermi surface.
Then, α is independent of k and
α X V (k, k 0 )
α = − (2591)
N 0
Ek0 + Ek0 +q − E(q)
k

Thus, the energy eigenvalue is determined from the equation


1 X V (k, k 0 )
1 = − (2592)
N 0
Ek0 + Ek0 +q − E(q)
k

For Cooper pairs with zero total momentum q = 0, this equation reduces to
Z µ+h̄ωD
ρ()
1 = V d (2593)
µ 2  − E(0)
The density of states ρ() can be approximated by a constant ρ(µ), and the
integral can be performed, yielding
 
V ρ(µ) 2 h̄ ωD + 2 µ − E(0)
1 = ln (2594)
2 2 µ − E(0)
This can be inverted to give the energy eigenvalue as
2 h̄ ωD
E(0) = 2 µ −  2
 (2595)
exp V ρ(µ) − 1
This eigenvalue is less than the minimum energy of the two independent elec-
trons, thus, the electrons are bound together. It is concluded that, due to the
sharp cut off of the integral at the Fermi energy, the electrons bind to form
Cooper pairs no matter how small the attractive interaction is. The binding
energy is small and is a non-analytic function of the pairing potential V , that
is, the binding energy cannot be expanded as a power series in V .

In the case that the pairing potential is spin rotationally invariant, the total
spin of the pair S is a good quantum number. The pairing states can be catego-
rized by the value of their spin quantum number and the projection of the total
spin along the z-axis. On pairing two spin one-half electrons, there are four
possible state, a spin singlet state S = 0 and a spin triplet state S = 1 which
is three-fold degenerate. The four Cooper pair wave functions corresponding to
these states have to obey the Paul-exclusion principle and are written as
 
X 1
ψS=0 (r1 , r2 ) = CS=0 (k) φk (r1 ) φ−k (r2 ) − φ−k (r1 ) φk (r2 ) ×
2
k
 
× χ+ 1 χ− 2 + χ− 1 χ+ 2

(2596)

735
for the spin singlet pairing. The three spin triplet pair wave functions are
X
ψS=1,m=1 (r1 , r2 ) = CS=1 (k) φk (r1 ) φ−k (r2 ) χ+ 1 χ+ 2
k
 
X 1
ψS=1,m=0 (r1 , r2 ) = CS=1 (k) φk (r1 ) φ−k (r2 ) + φ−k (r1 ) φk (r2 ) ×
2
k
 
× χ+ 1 χ− 2 − χ− 1 χ+ 2
X
ψS=1,m=−1 (r1 , r2 ) = CS=1 (k) φk (r1 ) φ−k (r2 ) χ− 1 χ− 2
k

(2597)

Thus, for singlet pairing one must have

CS=0 (k) = CS=0 (−k) (2598)

which requires that, when expanded in spherical harmonics, the expansion only
contains even components of orbital angular momentum. For triplet pairing,
one has
CS=1 (k) = − CS=1 (−k) (2599)
thus, the triplet pair can only be composed of odd values of orbital angular
momentum.

Most superconductors that have been found have singlet spin pairing and are
in a state which is predominantly in a state of orbital angular momentum l = 0.
The high Tc superconductor such as Sr doped La2 CuO4 found by Bednorz and
Muller295 in 1986 (Tc = 35 K) or Y Ba2 Cu3 O7 (Tc = 90 K) form exceptions
to this rule. These materials evolve from an anti-ferromagnetic insulator phase
at zero doping, but as the doping increases they lose the antiferromagnetism
and become metallic paramagnets. A superconducting phase appears for dop-
ing concentrations above a small critical concentration. The superconductivity
is exceptional, not just in the magnitude of the transition temperature Tc but
also in that the pairing is singlet, but with an appreciable admixture of a com-
ponent with l = 2 in the pair. Due to this admixture, the pairing in high Tc
superconductors is sometimes referred to as d wave pairing. In heavy fermion
superconductors, such as CeCu2 Si2 , U Be13 , U P t3 and U Ru2 Si2 , experimental
evidence exists that these materials do not show exponentially activated behav-
ior characteristic of a gap. Instead, the specific heat and susceptibility show
power law variations296 . This, and the multiple superconducting transitions
found in U P t3 and T h doped U Be13 , indicate that the order parameter is dom-
inated by components with non-zero angular momentum297 . If the symmetry
295 J.G. Bednorz and K. A. Müller, Zeit. für Physik, B, 64, 189 (1986).
296 H. R. Ott, H. Rudigier, T. M. Rice, K. Ueda, Z. Fisk and J. L. Smith, Phys. Rev. Lett.
52, 1915 (1984).
297 H. R. Ott, H. Rudigier, Z. Fisk and J. L. Smith, Phys. Rev. B, 31, 1615 (1985).

736
-r/2

r/2

Figure 309: A schematic depiction of the overlap of Cooper pairs. The coherence
length is denoted by ξ.

of the order parameter is that of a state with definite angular momentum, the
superconducting gap may vanishes at points or lines on the normal state Fermi
surface.

It is customary to represent the wave function of the Cooper pair in terms of


r + r
relative coordinates r = r1 − r2 and center of mass coordinates R = 1 2 2 .
Thus, the Cooper pair wave function is written as

ψ(r1 , r2 ) → ψ(r, R) (2600)

and as the pair usually is in a state with zero total momentum, q = 0, the
center of mass dependence can be ignored.

The mean square radius of the Cooper pair wave function is given by
Z
ξ2 = d3 r r2 | ψ(r) |2 (2601)

but  
X
ψ(r) = C(k) exp ik.r (2602)
k

Thus,
Z X  
2 3 2 ∗ 0 0
ξ = d rr C(k) C (k ) exp i(k − k ).r
k,k0
X
= | ∇k C(k) |2
k
 2
4 h̄ vF
=
3 2µ − E
 2  
4 vF 4
= exp (2603)
3 2 ωD V ρ(µ)

737
For a binding energy of order 10 K and a Fermi velocity vF of the order of 106
m/sec, one obtains a pair size ξ of order 104 Angstroms. The coherence length
ξ is much greater than the average spacing between the electrons.

The standard weak-coupling theory of superconductivity due to Bardeen,


Cooper and Schrieffer, (B.C.S.)298 , treats the Cooper pairing of all electrons
close to the Fermi surface in a self-consistent manner.

25.3 Pairing Theory


25.3.1 The Pairing Interaction
The attractive pairing interaction can be obtained from the electron-phonon
interaction, via an appropriately chosen canonical transform. The energy of the
combined electron phonon system can be expressed as the sum

Ĥ = Ĥ0 + Ĥint (2604)

where the non-interacting Hamiltonian is given by

Ek c†k,σ ck,σ +
X X
Ĥ0 = h̄ ωα (q) a†q,α aq,α (2605)
k,σ q,α

and the interaction term is given by


 
λq c†k+q,σ ck,σ aq,α + a†−q,α
XX
Ĥint = (2606)
k,σ q,α

The Hamiltonian will be transformed via


   
Ĥ 0 = exp + Ŝ Ĥ exp − Ŝ (2607)

where Ŝ is chosen in a way that will eliminate the interaction term (at least to
in first order). The operator Ŝ can be thought of as being of the same order as
Ĥint . The transformation proceeds by expanding the transformed Hamiltonian
in powers of Ŝ
 
Ĥ 0 = Ĥ0 + Ĥint + Ŝ , Ĥ0
      
1 3
+ Ŝ , Ŝ , Ĥ0 + Ŝ , Ĥint + O Ĥint
2
(2608)
298 J. Bardeen, L. N. Cooper and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).

738
and then choosing Ŝ such that
 
Ĥint = Ĥ0 , Ŝ (2609)

Since this is an operator equation, this is solved for Ŝ by taking matrix elements
between a complete set. A convenient complete set is provided by the eigenstates
| φm > of Ĥ0 , which have energy eigenvalues Em . Thus, the matrix elements
of Ŝ are found from the algebraic equation
< φm | Ĥint | φn > = ( Em − En ) < φm | Ŝ | φn > (2610)
The complete set of energy eigenstates are energy eigenstates of the combined
non-interacting electron and phonon Hamiltonians. The non-zero matrix el-
ements only occur between states which involve a difference of unity in the
occupation number of one phonon mode ( either q or − q ) , and also a change
of state of one electron ( k to k + q ). The anti-Hermitean operator Ŝ can be
represented in second quantized form as

λq aq,α λq a†−q,α 
c†k+q,σ
XX
Ŝ = − ck,σ +
Ek+q − Ek − h̄ ω(q) Ek+q − Ek + h̄ ω(q)
k,σ q,α
(2611)
The transformed Hamiltonian contains the effects of the interaction only through
the higher order terms
   
1
Ĥ 0 = Ĥ0 + Ŝ , Ĥint + O Ĥint 3
2
(2612)
On evaluating the commutation relation, one finds a renormalization of the
electron dispersion relation of order | λq |2 , and electron-electron interaction
terms. The electron-electron interaction terms combine and can be written as
| λq |2 h̄ ω(q)
c†k+q,σ ck,σ c†k0 −q,σ0 ck0 ,σ0
X X
0
Ĥint =
k,σ;k0 ,σ 0 q,α
( Ek+q − Ek )2 − h̄2 ω(q)2
(2613)
Thus, for electrons within h̄ ω(q) of the Fermi energy there is an attractive inter-
action between the electrons. This interaction depends on the energy transfer
between the electrons, and the energy transfer corresponds to a frequency. As
the interaction is frequency-dependent, it corresponds to a retarded interaction.
Since the interaction is only attractive at sufficiently low frequencies, the in-
teraction is only attractive after long time delays. In the B.C.S. theory this
interaction is simplified. The simplification consists of only retaining scattering
between electrons of opposite spin polarization and momentum, as this maxi-
mizes the phase space of the allowed final states. That is, the momenta and
spin are restricted such that
k = − k0 (2614)

739
and also
σ = − σ0 (2615)
This procedure produces a pairing between electrons of opposite spins.

The B.C.S. Hamiltonian is composed of the energy of electrons in Bloch


states and an attractive interaction between the electrons mediated by the
phonons. The B.C.S. Hamiltonian is written as

Ek c†k,σ ck,σ − V (k, k 0 ) c†−k0 ,↓ c†k0 ,↑ ck,↑ c−k,↓


X X
Ĥ = (2616)
k,σ k,k0

25.3.2 The B.C.S. Variational State


The pairing theory of superconductivity considers the ground state to be a state
within the grand canonical ensemble. That is, the ground state is composed of
a linear superposition of states with different numbers of particles. If required,
a ground state in the canonical ensemble can be found by projecting the B.C.S.
ground state onto one with a fixed number of particles. The B.C.S. state is
chosen variationally, by minimizing the energy.

The B.C.S. ground state is found from anti-symmetrizing the many-particle


state which is composed of products of pairs of wave functions. Each pair
corresponds to the two Bloch states (k, ↑) and (−k, ↓). For each wave vector k,
the pair state ((k, ↑), (−k, ↓)) is occupied with probability amplitude u(k) and
unoccupied with probability amplitude v(k). The probability amplitudes are
often referred to as coherence factors.
Y  
| ΨBCS > = v(k) + u(k) c†k,↑ c†−k,↓ | 0 > (2617)
k

The amplitudes satisfy the constraint

| u(k) |2 + | v(k) |2 = 1 (2618)

The normal state for non-interacting electrons just corresponds to the special
case,
| u(k) |2 = Θ( µ − Ek ) (2619)
and the phase of the u(k) for each occupied wave function are completely unde-
termined. The functions u(k) and v(k) in the variational ansatz are variational
parameters that are to be found by minimizing the expectation value of the
Hamiltonian, which includes the pairing interaction.

740
The expectation value for the appropriate energy, in the B.C.S. state, is
given by

E − µN = < ΨBCS | ( Ĥ − µ N ) | ΨBCS >


X X
= 2 ( Ek − µ ) | u(k) |2 − V (k, k 0 ) v ∗ (k) u(k) u∗ (k 0 ) v(k 0 )
k k,k0

(2620)

The term involving the double sum is eliminated by introducing a quantity


X
∆(k) = V (k, k 0 ) u∗ (k 0 ) v(k 0 ) (2621)
k0

On minimizing the energy with respect to u(k) and v ∗ (k), subject to the con-
straint of conservation of probability, one finds
 
0 = 2 ( Ek − µ ) + λ u∗ (k) − ∆(k) v ∗ (k)

0 = λ v(k) − ∆(k) u(k) (2622)

where λ is the Lagrange undetermined parameter. These equations can be


solved to yield
Ek − µ
 
2 1
| u(k) | = 1 − (2623)
2 Eqp (k)
and
Ek − µ
 
1
| v(k) |2 = 1 + (2624)
2 Eqp (k)
where we have defined
∆(k)
u(k) v ∗ (k) = (2625)
2 Eqp (k)
The first two equations can be multiplied and equated to the modulus squared
of the third equation according to the identity
 ∗  
2 2 ∗ ∗
| u(k) | | v(k) | = u(k) v (k) u(k) v (k) (2626)

The resulting expression can be solved for the quasi-particle energy, Eqp (k),
resulting in q
Eqp (k) = + ( Ek − µ )2 + | ∆(k) |2 (2627)

The factor | u(k) |2 , is the probability of finding an electron of momentum k


and spin σ in the B.C.S. ground state and, therefore, is just n(k). Unlike a
Fermi liquid, where n(k) is discontinuous at the Fermi surface with magnitude
Z(k)−1 , in the superconductor the distribution drops smoothly to zero as k in-
creases above kF . Thus, the concept of Fermi surface is not well defined in the

741
1.2

|u(k)|2 |v(k)|2
1

0.8

0.6

0.4

0.2

0
-10 -8 -6 -4 -2 0 2 4 6 8 10
(Ek−µ)/∆

Figure 310: The energy dependence of the probability for finding a pair |u(k)|2 ,
and the probability |v(k)|2 that the pair is absent.

superconducting state. The energy Eqp (k), relative to µ, turns out to be the
energy required to create a quasi-particle of momentum k from the ground state.
The quasi-particle is either of the form of an added electron or a hole. With
the B.C.S. ground state both of these quasi-particle excitations leave a single
unpaired electron in an otherwise perfectly paired B.C.S. state. The minimum
energy required to create two quasi-particles, that is two individual electrons,
is just 2 ∆(k F ) . That is, there is a gap of 2 ∆(k F ) in the excitation spectra of
the superconductor.

25.3.3 The Gap Equation


The “energy gap” parameter satisfies the non-linear integral equation
X ∆∗ (k 0 )
∆(k) = V (k, k 0 ) (2628)
2 Eqp (k 0 )
k0

where V (k, k 0 ) is the attractive pairing interaction mediated by the phonons. It


should be noted that the phase of u(k 0 ) v ∗ (k 0 ) is the phase of a pair k 0 relative
to the rest of the ground state. Furthermore, if the phase of the pairs for each
k 0 value were randomly distributed, then ∆(k) would most probably be zero.
Hence, the gap equation implies that there is a relation between the phases of
all the wave functions of the pairs of electrons299 . The pairing interaction can
be approximated by the attractive s-wave potential

V (k, k 0 ) = V for | Ek − µ | < h̄ ωD


299 To be more precise, the spatial extent of a Cooper pair is given by the coherence length

ξ which is extremely long compared to the average separations between adjacent electrons.
Thus, as many as 106 Cooper pairs have their centers of mass located in a volume ξ 3 contained
between the two electrons which form a Cooper pair. However, all the Cooper pairs with their
centers of mass in this spatial region are phase coherent.

742
V (k, k 0 ) = 0 for | Ek − µ | > h̄ ωD (2629)

In this case, one finds that the gap has s-wave symmetry, and is give by

∆(k) = ∆(0) for | Ek − µ | < h̄ ωD

∆(k) = 0 for | Ek − µ | > h̄ ωD (2630)

For other types of potentials, the gap may have p or d wave symmetry and,
therefore, vanish at lines or points on the Fermi surface. For the s-wave case,
the gap in the quasi-particle dispersion relation at the Fermi energy is given by
the solution of
Z h̄ωD
1
1 = V ρ(µ) d p
−h̄ωD 2  + | ∆(0) |2
2
Z h̄ωD
1
= V ρ(µ) d p
2 + | ∆(0) |2
0
h̄ ωD
= V ρ(µ) sinh−1 (2631)
| ∆(0) |

which is solved as
h̄ ωD
| ∆(0) | = 1 (2632)
sinh V ρ(µ)
This gap 2 ∆(0) just corresponds to the minimum energy required to break the
s-wave Cooper pair. At finite temperatures, the superconducting gap satisfies
the equation
X ∆(k 0 )
∆(k) = V (k, k 0 ) ( 1 − 2 f (Eqp (k 0 )) )
2 Eqp (k 0 )
k0
X ∆(k 0 ) βEqp (k 0 )
= V (k, k 0 ) 0 tanh (2633)
2 Eqp (k ) 2
k0

The tanh factor is a decreasing function for increasing temperature, therefore,


for the equation to have a non-trivial solution the denominator has to decrease
with increasing temperature. This can only happen if | ∆(T ) | decreases with
increasing temperature. For sufficiently high temperatures, the equation can be
reduced to
β h̄ωD
∆(T ) = ∆(T ) V ρ(µ) (2634)
2
which only has the trivial solution ∆(T ) = 0, when 2 kB T  h̄ ωD V ρ(µ).
The critical temperature below which the gap is non-zero, ∆(Tc ) = 0, is given

743
by the linearized equation300
h̄ωD
tanh βc2 
Z
1 = 2 ρ(µ) V d
0 2
h̄ωD
tanh βc2 
Z
= ρ(µ) V d
0 
Z βh̄ωD
2 tanh z
= ρ(µ) V dz
z
0 Z ∞ 
βh̄ωD 2
= ρ(µ) V ln − dz ln z sech z
2 0
 
βh̄ωD π
= ρ(µ) V ln − ln (2635)
2 4 exp γ

The critical temperature Tc is given by


 
1
kB Tc = 1.14 h̄ωD exp − (2636)
V ρ(µ)

The critical temperature is a non analytic function of the coupling constant.


The critical temperature is proportional to h̄ωD and, therefore, proportional
1
to M − 2 , as is expected from the isotope effect. Deviations from the classical
isotope effect are found in transition metals such as Ru and Os, and in the
actinides where the Coulomb interaction strength is large. The strength of the
coupling constant can be estimated from knowledge of the Debye frequency and
the critical temperature. Typical values of ρ(µ) V are quite small, of the order
of 0.2.

300 The analysis described here assumes that the density of states at the Fermi energy is

roughly constant. This is not the case for M gB2 [Nagamatsu et al., Nature, 410, 63, (2001)].
Since in M gB2 Boron is isoelectronic with graphite and forms layers which have the same
structure as the layers of Graphite, the electronic density of states is expected to show an
energy dependence similar to that of graphite.

744
Metal ΘD (K) Tc (K) ρ(µ) V

Zn 235 0.9 0.18


Cd 164 0.56 0.18
Hg 70 4.16 0.35
Al 375 1.2 0.18
Ga 320 1.1 0.18
In 109 3.4 0.29
Tl 100 2.4 0.27
Sn 195 3.75 0.25
Pb 96 7.22 0.39

The ratio of the T = 0 energy gap 2 ∆(0) to the superconducting transition


temperature has the theoretical value301
2 ∆(0)
= 3.53 (2637)
kB Tc
whereas the experimentally determined values in Al, Ga, In, T l, Sn and P b are
3.4, 3.5, 3.6, 3.6, 3.5 and 4.4.

Below the critical temperature, the superconducting order parameter ∆(T )


is finite. Above the critical temperature ∆(T ) = 0 and the B.C.S. state reduces
to the normal state. Just below the critical temperature, one has
8 π2 2
∆(T )2 = k Tc ( T c − T ) T → Tc (2638)
7 ζ(3) B
Thus, the order parameter has a typical mean-field variation with an exponent
of β = 21 close to Tc .

25.3.4 The Ground State Energy


The normal state is unstable to the B.C.S. state only if the B.C.S. state has a
lower energy. At T = 0 the stability can be found by examining the energy

E − µN = < ΨBCS | ( Ĥ − µ N ) | ΨBCS >


X X | ∆(k) |2
= 2 | u(k) |2 ( Ek − µ ) −
2 Eqp (k)
k k

301 J. Bardeen and J. R. Schrieffer, Prog. in Low Temp. Phys. 3, 170, (1961).

745
Figure 311: The reduced superconducting gap ∆(T )
∆(0) for In, Sn and P b as a
function of the reduced temperature TTc . The experimental data are compared
g. 25: Temperature dependence of the energy gap according to the BCS theory and comparison
with the prediction of the B.C.S. theory. [After Ivar Giaever and Karl Megerle,
with experimental data
Phys. Rev. 122, 101 (1961).].

h̄ωD
| ∆(0) |2
X Z
2
= 2 | u(k) | ( Ek − µ ) − ρ(µ) d p
One of the fundamental formulaek of the BCS theory is the 0 relation | ∆(0) |2 the energy gap ∆(
between
2 +
= 0, the Debye frequency ωD and Xthe electron-lattice (interaction V 0 :|2
potential| ∆(0)
" #
Ek − µ )2
= ( Ek − µ ) − p −
2 ∆(k) |2 V
k  ( Ek − µ ) + | 
1 (2639)
∆(0) = 2~ωD exp − . (2
V N (E )
F represents the at-
0 last term
The first term represents the kinetic energy. The
tractive interaction. The integral in the last term is evaluated with the aid of
the gap equation. The condensation energy, ∆E, is defined as the difference
e N (EF ) is the densitybetween
of single-electron states of a state
the energy of the superconducting given andspin orientation
the normal state at E = E F (the other sp
ntation is not counted because a Cooper pair consists of twoZ 0electrons with opposite spin). Althou
∆E = < ΨBCS | ( Ĥ − µ N ) | ΨBCS > − 2 d  ρ(µ + )
interaction potential V0 is assumed to be weak, one of the−∞most striking observations is that t
onential function cannot be expanded in a Taylor series around V 0 = 0 because (2640) all coefficients van
tically. This implies that Eq. (27) is
The condensation a truely
energy non-perturbative
is evaluated result.
by writing the sum over The
k as an fact that superconductiv
integral
over the density of states.
not be derived from normal conductivity Z h̄ω
by introducing
" a ‘small’ # interaction potential and applyi
2
D

urbation theory (which is∆E the =usual method
d ρ(µ + )for treating
− p problems of atomic, nuclear and solid sta
0 2 + | ∆(0) |2
sics that have no analytical solution) explains why it took so many decades to find the correct theo
critical temperature is given by a similar expression
746
 
1
kB Tc = 1.14 ~ωD exp − . (2
V0 N (EF )

mbining the two equations we arrive at a relation between the energy gap and the critical temperatu
" #
0
2
Z
+ d ρ(µ + )  − 2 − p
−h̄ωD 2 + | ∆(0) |2
| ∆(0) |2

V " #
Z h̄ωD
2
= d ρ(µ + )  − p
0 2 + | ∆(0) |2
" #
0
2
Z
+ d ρ(µ + ) −  − p
−h̄ωD 2 + | ∆(0) |2
| ∆(0) |2
− (2641)
V
The integral over states below the Fermi energy,  < 0, can be transformed to
an integral over positive . This leads to the condensation energy being given
by
Z h̄ωD " #
2 | ∆(0) |2
∆E = 2 ρ(µ) d  − p −
0 2 + | ∆(0) |2 V

" #
Z h̄ωD p
= 2 ρ(µ) d  − 2 + | ∆(0) |2
0
h̄ωD
| ∆(0) |2 | ∆(0) |2
Z
+ 2 ρ(µ) d p −
0 2 + | ∆(0) |2 V
(2642)
The integrals are evaluated with the aid of the substitution
 = | ∆(0) | sinh θ (2643)
which gives the result
s 2 
| ∆(0) |2 | ∆(0) |2
 
2 2 | ∆(0) |
∆E = h̄ ωD ρ(µ) 1 − 1 + + −
h̄ ωD V V
1
≈ − ρ(µ) | ∆(0) |2 + . . . (2644)
2
The condensation energy comes from the attractive potential which is larger
than the increase in the kinetic energy caused by the confinement of the pair
within the coherence length ξ. The net lowering can be understood in terms of
the quasi-particle dispersion relation. The electrons with energy within | ∆(0) |
of µ have their energy lowered by an amount | ∆(0) |. The net lowering of
energy is just the number of electrons, ρ(µ) | ∆(0) |, times the energy lowering
| ∆(0) |. Therefore, the B.C.S. state has lower energy than the normal state
whenever the gap is non-zero.

747
25.4 Quasi-Particles
The B.C.S. Hamiltonian can be solved for the quasi-particle excitations, in the
mean-field approximation, by linearizing the pairing interaction terms. In a
normal metal, the only allowed matrix elements are between initial and final
states which have the same number of electrons. However, since for a super-
conductor the average is to be evaluated in the B.C.S. ground state, matrix
elements between operators with different numbers of pairs are non-zero. These
give rise to the anomalous expectation values. For example, the anomalous ex-
pectation value associated with adding a pair of electrons ((k 0 , ↑), (−k 0 , ↓)) to
the superconducting condensate is given by the probability amplitude

< ΨBCS | c†k0 ,↑ c†−k0 ,↓ | ΨBCS > = u∗ (k 0 ) v(k 0 ) (2645)

The linearized mean-field Hamiltonian is given by


X  † †

ĤM F − µ N = ( Ek − µ ) ck,↑ ck,↑ + ( E−k − µ ) c−k,↓ c−k,↓
k

V (k, k 0 ) < ΨBCS | c†−k0 ,↓ c†k0 ,↑ | ΨBCS > ck,↑ c−k,↓


X

k,k0

V (k, k 0 ) c†−k0 ,↓ c†k0 ,↑ < ΨBCS | ck,↑ c−k,↓ | ΨBCS >


X

k,k0

V (k, k 0 ) < ΨBCS | c†−k0 ,↓ c†k0 ,↑ | ΨBCS > < ΨBCS | ck,↑ c−k,↓ | ΨBCS >
X
+
k,k0

(2646)

The anomalous expectation value leads to a term in the Hamiltonian with


strength X
∆(k) = V (k, k 0 ) u∗ (k 0 ) v(k 0 ) (2647)
k0

which corresponds to a process in which two electrons ((k, ↑), (−k, ↓)) are ab-
sorbed into the condensate. The mean-field Hamiltonian also contains the Her-
mitean conjugate which represents the reverse process in which two electrons
are emitted from the condensate.
X  
ĤM F − µ N = ( Ek − µ ) c†k,↑ ck,↑ + ( E−k − µ ) c†−k,↓ c−k,↓
k
X  | ∆(0) |2

− ∆(k) ck,↑ c−k,↓ + c†−k,↓ c†k,↑ ∗
∆ (k) +
V
k

(2648)

In the absence of an electromagnetic field, the order parameter ∆(k) can be


chosen to be real. The mean-field Hamiltonian involves terms in which the con-
densate emits or absorbs two electrons. This is reminiscent of the treatment of

748
anti-ferromagnetic spin waves, using the method of Holstein and Primakoff, ex-
cept here the Hamiltonian involves fermions rather than bosons. The quadratic
Hamiltonian can be diagonalized by means of a canonical transformation.

We shall define two new fermion operators via the transformation


   
αk = exp + Ŝ ck,↑ exp − Ŝ (2649)

and    
βk† = exp + Ŝ c†−k,↓ exp − Ŝ (2650)

where Ŝ is an anti-Hermitean operator, Ŝ † = − Ŝ. The energy eigenvalues of


the Hamiltonian can be found directly from the transformed Hamiltonian
   
0
ĤM F = exp + Ŝ Ĥ MF exp − Ŝ (2651)

as they have the same eigenvalues and the eigenstates are related via
 
0
| φn > = exp + Ŝ | φn > (2652)

The operator Ŝ is chosen to be of the form


 
† †
X
Ŝ = θk ck,↑ c−k,↓ − c−k,↓ ck,↑ (2653)
k

Explicitly, the transformation yields

αk = ck,↑ cos θk − c†−k,↓ sin θk

βk† = c†−k,↓ cos θk + ck,↑ sin θk (2654)

Rather than working with the transformed Hamiltonian, we shall express the
original Hamiltonian in terms of the transformed operators. Hence, we shall
require the inverse transformation which expresses the original electron and
holes operators in terms of the new quasi-particles. The inverse transformation
is expressed in terms of the transformation matrix but with θk → − θk so one
has

ck,↑ = αk cos θk + βk† sin θk

c†−k,↓ = βk† cos θk − αk sin θk (2655)

The mean-field Hamiltonian is expressed in terms of the new operators and θk is


chosen so that the terms that are not represented in terms of the quasi-particle

749
number operators vanish. The normal terms in the Hamiltonian are found as
X  † †

( Ek − µ ) ck,↑ ck,↑ + ( E−k − µ ) c−k,↓ c−k,↓
k
"    #
αk αk† βk βk† αk† βk†
X
2 2
= ( (k) − µ ) sin θk + + cos θk αk + βk
k
 
αk† βk† + βk αk
X
+ ( (k) − µ ) sin 2θk (2656)
k

The anomalous terms are evaluated as


X  
− ∆(k) ck,↑ c−k,↓ + c†−k,↓ c†k,↑ ∆(k)
k
   
βk† αk αk†
X
= − <e ∆(k) sin 2θk βk −
k
   
αk† βk† + βk αk
X
+ <e ∆(k) cos 2θk
k

(2657)

The off-diagonal terms can be made to vanish by choosing


 
<e ∆(k)
tan 2θk = −  (2658)
Ek − µ

Thus, θk decreases from a value less than π4 to less than − π4 as (k) varies
from h̄ωD below µ to h̄ωD above µ. The factors cos θk and sin θk are found
to be related to the factors u(k) and v(k) in the B.C.S. ground state wave
function. After this value has been chosen, the Hamiltonian is expressed as the
sum of a constant and terms involving the number operators of the α and β
quasi-particles
 
† †
X
ĤM F = E0 + Eqp (k) αk αk + βk βk (2659)
k

This procedure shows that the excitations are quasi-particles as they are still
fermions. Furthermore, these quasi-particles have excitation energies which have
the dispersion relation
q
Eqp (k) = + ( Ek − µ )2 + | ∆(k) |2 (2660)

The canonical transformation shows that the quasi-particles are part electron
and part hole like. Basically, this is a consequence that the quasi-particle
excitation consists of a single unpaired electron (k, σ), in the presence of the

750
Excitation Energies
2
Allowed
excitations E>0
1
electron addition

E(k)/µ 0
-kF electron removal kF

-1
Forbidden
excitations E<0
-2
-2 -1 0 1 2

k/kF

Figure 312: The excitation energies for adding an electron or removing an elec-
tron (creating a hole) from the normal state of a metal. Electrons can only be
added to states with k > kF , and holes can only be created for k < kF . The
allowed excitation energies are positive.

Quasiparticle Excitation Energies


2

µ
1
E(k)/µ

∆ ∆
0
-kF kF

-1

-2
-2 -1 0 1 2
k/kF

Figure 313: The excitation energies for creating a quasi-particle in the super-
conucting state of a metal. The minimum energy required to create a quasi-
particle is ∆.

751
condensate. This specific state can be produced from the ground state, either
by adding the electron (k, σ) to the system or by breaking a Cooper pair by
removing the partner electron (−k, −σ). We note that the quasi-particles are
eigenstates of the spin operator. The α quasi-particle is a spin-up excitation
as it is composed of an up-spin electron and down-spin hole, whereas the β
quasi-particle is a spin-down excitation. From the dispersion relation, one finds
that the B.C.S. superconductor is actually characterized by the presence of a
gap in the excitation spectrum. That is, there is a minimum excitation energy
2 | ∆(kF ) | corresponding to breaking a Cooper pair and producing two inde-
pendent quasi-particles.

——————————————————————————————————

25.4.1 Exercise 97
Evaluate the constant term in the mean-field B.C.S. Hamiltonian. Show that
the variational B.C.S. ground state is the lowest energy state of the mean-field
Hamiltonian by showing that the quasi-particle destruction operators annihilate
the B.C.S. state

αk | ΨBCS > = 0

βk | ΨBCS > = 0 (2661)

——————————————————————————————————

25.5 Thermodynamics
Since the quasi-particles are fermions, the entropy S due to the gas of quasi-
particles is given by the formulae
X 
S = − 2 kB ( 1 − f (Eqp (k)) ) ln[ 1 − f (Eqp (k)) ] + f (Eqp (k)) ln[ f (Eqp (k)) ]
k
(2662)
By the usual procedure of minimizing the grand canonical potential Ω with
respect to the distribution f (Eqp (k)) , one can show that the non-interacting
quasi-particles are distributed according to the Fermi-Dirac distribution func-
tion. Therefore, the quasi-particle contribution to the specific heat is just given
by
 
X ∂f (Eqp (k))
Cqp (T ) = 2 Eqp (k)
∂T
k

752
6

ρqp(E)/ρ(µ)
3

0
0 1 2 3 4
E/∆

Figure 314: The normalized quasi-particle density of states for an s-wave super-
conductor. An s-wave superconductor has a gap all around the Fermi-surface,
and hence the quasi-particle density of states exhibits a gap.

  
X Eqp (k) ∂Eqp (k) ∂f (Eqp (k))
= 2 Eqp (k)+ −
T ∂T ∂E
k
Z +∞
T ∂∆(T )2
   
2 2 ∂f
= − dE ρqp (E) E −
T −∞ 2 ∂T ∂E
(2663)

The average of the temperature derivative of the square of the quasi-particle en-
ergy is given by the temperature derivative of the gap. In the above expression,
we have introduced the quasi-particle density of states
X  
ρqp (E) = δ E − Eqp (k) (2664)
k

Since, in the mean-field approximation, the square of the gap has a finite slope
for T just below Tc and is zero above,
 
T
∆(T )2 ∼ ∆(0)2 1 − Θ( Tc − T ) (2665)
Tc
the specific heat has a discontinuity at Tc . In B.C.S. theory, the magnitude
of the specific heat jump has the value given by, 3.03 ∆2 (0) ρ(µ) / Tc . Thus,
the value of the specific heat jump found in weak coupling B.C.S. theory, when
normalized to the normal state specific heat, is given by
∆C(Tc ) Cs − Cn
=
C(Tc ) Cn
12
= = 1.43 (2666)
7 ζ(3)

753

∆C

Material C
Tc

Zn 1.3
Cd 1.4
Hg 1.4
Al 1.4
Ga 1.4
In 1.7
Tl 1.5
Sn 1.6
Pb 2.7

This ratio is a measure of the quantity


2
∂∆2 (T )
  
1 ∆(0)
2 T
∼ (2667)
2 kB c ∂T
Tc kB Tc

The values of the specific heat jumps for strong coupling materials tend to be
higher than the B.C.S. value, for example the normalized jump for P b is as large
as 2.71. This trend is understood as being due to inelastic scattering processes
which tend to suppress Tc more than ∆(0). The heavy fermion superconductors
show that the normalized specific heat discontinuities are significantly smaller
than the B.C.S. ratio.

Low Temperatures.

The gap in the quasi-particle density of states could be expected to show up


in an activated exponential dependence of the low-temperature electronic spe-
cific heat, for T  Tc . For these temperatures the order parameter is expected
to have saturated, and so if one considers the Fermi liquid as being well formed
then the quasi-particle contribution is given by
Z +∞  
2 ∂f
Cqp (T ) = − dE ρqp (E) E 2 (2668)
T −∞ ∂E

The B.C.S. quasi-particle density of states is evaluated as


X  
ρqp (E) = δ E − Eqp (k)
k
Z ∞ 
p

= d ρ() δ E − (  − µ )2 + ∆(T )2
−∞

754
Figure 315: The specific heat for normal and superconducting states of Al. The
specific heat shows a jump at the transition temperature Tc = 1.163 K. [After
N. E. Phillips, Phys. Rev. 114, 676 (1959).]

|E|
∼ ρ(µ)
| − µ|
|E|
= ρ(µ) p for | E | > ∆(T )
E − ∆(T )2
2

(2669)

In evaluating the B.C.S. density of states, the conduction band electron density
of states has been approximated by a constant value. The resulting B.C.S.
quasi-particle density of states has a gap of magnitude 2 ∆(T ) around the Fermi
energy. This yield an exponentially activated behavior of the specific heat,
 
∆(0)
Cqp (T ) ∼ 9.17 γ Tc exp − (2670)
kB T

found in B.C.S. theory. For superconductors where order parameter has p or


d wave symmetry, the gap may vanish either on lines or points of the Fermi
surface. The vanishing of the gap can give rise to a power law behavior of the
low temperature specific heat of the superconductor302 .

302 H. R. Ott, H. Rudigier, Z. Fisk and J. L. Smith, Phys. Rev. B, 31, 1615 (1985).

755
25.6 Perfect Conductivity
The current is composed of the sum of a paramagnetic current and a diamagnetic
current. The paramagnetic current can be evaluated from the Kubo formula.
There are two contributions to the paramagnetic current: one contribution is
from the particles in the condensate, the other contribution is from the excited
quasi-particles. The condensate contribution is proportional to
2
e2 h̄2 1 X
 

j p (q; ω) = 2
( 2 k − q ) ( 2 k − q ) . A u(k) v(k + q) − u(k + q) v(k) ×

8m c V
k

1 − f (E(k)) − f (E(k − q)) 1 − f (E(k − q)) − f (E(k))


 
× +
E(k − q) + E(k) + h̄ ω E(k) + E(k − q) − h̄ ω
(2671)

The coherence factor,


 
u(k) v(k + q) − u(k + q) v(k) (2672)

occurs, since, in the B.C.S. ground state, adding an electron with specific spin
and momentum produces the same final state as adding a hole of opposite spin
and momentum. The process of exciting an electron from the Bloch state (k, σ)
to the Bloch state (k + q, σ) occurs with the probability amplitude v(k + q) u(k),
whereas the process of exciting an electron from the Bloch state (−k − q, −σ)
to the Bloch state (−k, −σ) occurs with probability amplitude − u(k + q) v(k).
Since these two processes start from the unique B.C.S. ground state and lead
to exactly the same final state, their probability amplitudes should be added.
The need to include the coherence factors becomes immediately transparent, if
one expresses the electron creation and annihilation operators in terms of the
quasi-particle operators

ck,↑ = αk cos θk + βk† sin θk (2673)

and
c−k,↓ = βk cos θk − αk† sin θk (2674)
From this, one sees that the up-spin electron and down-spin hole components
of the current operator
 
1
(k + q ) a†k+q,↑ ak,↑ − a†−k,↓ a−k−q,↓ (2675)
2

combine to produce the usual quasi-particle contributions to the current, like


 
1 †
(k + q ) ( cos θk+q cos θk + sin θk+q sin θk ) αk+q αk − βk† βk+q
2
(2676)

756
and the anomalous contributions, such as
 
1 †
(k + q ) ( cos θk+q sin θk − sin θk+q cos θk ) αk+q βk† − βk+q αk
2
(2677)
It is the anomalous contributions which give rise to the condensate component
of the paramagnetic current. Due to the appearance of the coherence factor
in the expression for the condensate component of the paramagnetic current,
and the finite value of the denominator when ω < 2 | ∆(0) |, the condensate
contribution vanishes in the limit q → 0. The condensate contribution also
vanishes identically in the normal state. The quasi-particle contribution to the
paramagnetic current has a coherence factor given by
 
u(k) u(k + q) + v(k) v(k + q) (2678)

This coherence factor represents the contribution from thermally activated quasi-
particles in the initial state. One contribution represents a process, in which the
thermally excited quasi-particle is viewed as an unpaired electron in the Bloch
state (k, σ) which is subsequently excited to the Bloch state (k + q, σ). This
process occurs with probability amplitude u(k) u(k +q). Alternatively, the same
quasi-particle may be viewed as a thermally exited single hole in the Bloch state
(−k, −σ) which is subsequently excited to the final Bloch state (−k − q, −σ).
This process is associated with the probability amplitude v(k) v(k + q). The
coherence factor is given by the sum of the two probability amplitudes. This
coherence factor tends to unity as q → 0, since it reduces to the normalization
condition. Hence, the quasi-particle contribution to the paramagnetic current
is given by
2
e2 h̄2 1 X
 

j p (q; ω) = 2
( 2 k − q ) ( 2 k − q ) . A u(k) u(k + q) + v(k) v(k + q) ×
4m c V
k

f (E(k)) − f (E(k − q)) f (E(k − q)) − f (E(k))


 
× +
E(k − q) − E(k) + h̄ ω E(k) − E(k − q) + h̄ ω
(2679)
In this expression E(k) is the quasi-particle energy in the superconductor. In
the static limit with uniform fields, ( ω → 0 , q → 0 ), the paramagnetic
current reduces to
e2 h̄2 1 X
 
∂f (E(k))
j p (0; 0) = 2 2 k(k.A) − (2680)
m c V ∂E
k

The total current is found by combining the paramagnetic current with the
diamagnetic current
e2 h̄2 1 X ρ e2
 
∂f (E(k))
j(0; 0) = − 2 2 k(k.A) − A
m c V ∂E mc
k

757
e2 h̄2 1 ρ e2
Z  
4 ∂f (E(k))
= −2 2 dk k A − A
m c 6 π2 ∂E mc
 2 2
ρ e2
Z   
e h̄ 1 4 ∂f (E(k))
= − dk k + A
m2 c 3 π 2 ∂E mc
(2681)
In the normal state, where the gap in E(k) vanishes, the derivative of the Fermi
function can be approximated as
 
∂f (E(k))
= − δ( E(k) − µ )
∂E
2m
= − 2 δ( k 2 − kF2 ) (2682)

which leads to the vanishing response as
kF3
ρ = (2683)
3 π2
Thus, in the normal state current does not flow in response to a static vector
potential. However, in the superconducting state the total current is given by
ρ e2
 Z  
2µ 4 ∂f (E(k))
j = − A 1 + dk k (2684)
mc kF5 ∂E
and as there is a gap on the Fermi surface, the derivative of the Fermi function
is always exponentially small. Because of the finite superconducting gap, the
second term is small and the cancellation does not occur. In the superconducting
state, this reduces to the London equation
ρ e2
j = − A (2685)
mc
This shows that a current will flow in a superconductor in response to a uniform
static vector potential, that is the current will screen an applied magnetic field.
This leads to the Meissner effect.

25.7 The Meissner Effect


In the superconducting state, the susceptibility is expected to be dominated
by the diamagnetic susceptibility produced by the supercurrent shielding the
external field. The Pauli spin susceptibility will also be modified by the su-
perconductivity, and provide information about the pairing. The zero field
susceptibility is defined as a derivative of the magnetization, χs (T ) = ( ∂M
∂H ).
The magnetization, produced by the electronic spins aligning with a magnetic
field applied along the z-axis, is given by
  " #
g µB X
Mz = f (E↑ (k)) − f (E↓ (k)) (2686)
2
k

758
1.2

0.8

χ(T) / χn
0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2
T / Tc

T
Figure 316: The spin susceptibility as a function of the reduced temperature Tc
for a BCS s-wave superconductor.

which is given in terms of the Fermi distribution for quasi-particles with spin σ
and quasi-particle energy Eσ (k).

For singlet pairing, the magnetic field couples to the spins of the quasi-
particles via the Zeeman energies and, as can be seen from inspection of the
matrix, only the time reversal partners pair. The quasi-particles consist of
broken pairs, i.e. electrons of spin σ and holes of spin − σ. Since a down-
spin hole has the same Zeeman energy as an up-spin electron, the quasi-particle
energies depend on field through
 
g µB σ H
Eσ (k) = EH=0 (k) − (2687)
2
and so the spin susceptibility takes the usual form
 2 Z +∞  
g µB ∂f
χs (T ) = − 2 dE ρqp (E) (2688)
2 −∞ ∂E
which involves the B.C.S. quasi-particle density of states. The Pauli suscepti-
bility tends to zero as T → 0 in an exponentially activated way
 
∆(0)
χp (T ) ∼ exp − (2689)
kB T
The exponential vanishing of the spin susceptibility occurs as the electrons form
singlet pairs in the ground state, and the finite spin moment is caused by ther-
mal population of quasi-particles.

Thus, in the spin-singlet phases, the spin susceptibility could be expected


to vanish as T → 0. However, spin-orbit coupling will produce a residual
susceptibility that depends on the ratio of the superconducting coherence length,
ξ0 to the mean free path due to spin-orbit scattering, lso . In the presence
of spin-orbit coupling, the spin is no longer a good quantum number for the

759
single-particle eigenstates, and the spin-up and spin-down states are mixed. In
→−
− →
the limit that the strength of the spin-orbit coupling λ L . S is so large that
λ  ∆0 , the average value of σz for a single-particle state tends to zero.
The spin susceptibility is, therefore, reduced. The scattering has the effect that
a significant contribution to the normal state χ(T ) comes from single-particle
states separated by an energy of the order of the spin-orbit scattering rate,
which is by our assumption greater than ∆. As an opening up of a gap at the
Fermi energy is not expected to change the contribution of these higher energy
states, one finds that the susceptibility in the superconducting state can remain
comparable in magnitude to the normal state value. According to Anderson,
the normalized susceptibility should have the two limits,

χs (0) 2 lso
= 1 − (2690)
χn π ξ0
for strong spin-orbit scattering and for weak spin-orbit scattering, one has

χs (0) π ξ0
= (2691)
χn 6 lso
Hence, a partial Meissner effect at T = 0 can be found in a conventional
superconductor.

25.8 Landau-Ginzburg Theory


Superconductors can be divided into two categories, which depend on their
macroscopic characteristics when an applied magnetic field is present. The clas-
sification is based on the length scale over which the magnetic field is screened,
λL , relative to the length scale over which the superconducting order parameter
changes, ξ. The latter length is given by the spatial extent of the Cooper pair
wave function or coherence length ξ0
h̄ vF
ξ0 = (2692)
π ∆(0)

Type I Superconductors.

Type I superconductors are materials with short penetration depths λ and


long coherence lengths ξ. In particular, in type I superconductors the ratio must
satisfy the inequality
λ 1
κ = < √ (2693)
ξ 2
Type I superconductors include simple (non-transition) metals, such as Al,
where the penetration depths are of the order of λ ∼ 300 A and the co-
herence lengths are ξ ∼ 1 × 104 A as vF is large.

760
Figure 317: The magnetization curves of annealed polycrystaline P b (A) and
P bIn alloys (B-D). [After J. D. Livingston, Phys. Rev. 129, 1943 (1963).]

Type II Superconductors.

Type II superconductors include materials with long penetration depths and


short coherence lengths. Type II superconductors can be satisfy the inequality
λ 1
κ = > √ (2694)
ξ 2
Type II materials include the transition metals, rare earths and intermetallic
compounds. For example, in V3 Ga, the band mass is very large, so λL is very
large of the order of 2000 A. Since these have high effective masses, the Fermi
velocities are small ( vF ∼ 104 m /s ) and as Tc and ∆(0) are high, then ξ is
small ( ∼ 50 Å).

Since λL and κ diverge the same way at Tc the dimensionless ratio κ is ap-
proximately temperature independent.

Metal ξ (A) λ (A)

Al 16,000 490-515
Sn 2,300 510
Pb 830 390
Cd 7,600 1,100
Nb 380 390

If a magnetic field H < Hc is applied to a small superconductor, the field


is excluded from the superconductor, but if H > Hc the field will penetrate

761
and the superconductor will undergo a transition to the normal state. If a field
is applied normal to the surface of a large superconducting slab then, because
∇ . B = 0, the field has to penetrate the slab.

In a type I superconductor the magnetic field will concentrate into regions


where
| B | = Hc (2695)
which are normal, and regions where

B = 0 (2696)

which are superconducting. The condensation energy density in the supercon-


H2 H2
ducting state is 8 cπ , and the diamagnetic energy of the normal state is also 8 cπ .
These regions are separated by a domain wall which has positive energy. The
energetic cost of forming a domain wall of area A can be estimated as
E H2 Hc2
∼ ξ c − λ (2697)
A 8π 8π
Hc2
The term ξ 8 π is the energetic cost of setting the order parameter to zero. The
H2
diamagnetic energy is reduced by λ 8 πc . Because of the positive domain wall
energy in a type I superconductor, the number of domains and domain walls
will be minimized. The domain pattern will have a scale of subdivision which
is intermediate between ξ and the sample size.

In type II superconductors, a similar separation occurs, but as the domain


wall energy is negative, the superconductor will break up into as many normal
regions as possible. These normal regions have the form of magnetic flux car-
rying tubes that thread the sample, which are known as vortices. Each vortex
carries a minimum amount of flux Φs , the superconducting flux quantum. The
flux quanta in a superconductor differs from the flux quanta in a normal metal,
as the current in a superconductor is carried by Cooper pairs with charge 2 e,
whereas in a normal metal current is carried by electrons with charge e. The
flux in a superconductor is quantized in units of Φs .
hc
Φs = = 2.07 × 10−7 Gauss cm2 (2698)
2e
The vortices first enter the superconductor at a critical field Hc1 . The vortices
are arranged on the sites of a triangular lattice. The superconductor becomes
saturated with vortices when it becomes completely normal at an upper critical
field Hc2 .

The magnetization M is linear in field up to Hc1 with susceptibility − 41π .


At Hc1 the magnitude of the magnetization has a cusp and the magnitude falls
to zero at Hc2 .

762
In order to discuss the spatial variation of superconductivity due to the mag-
netic field, it is necessary to extend the microscopic B.C.S. theory to inhomo-
geneous systems. The microscopic generalization was performed by Gor’kov303 .
This resulted in a set of equations which had previously been proposed by Lan-
dau and Ginzburg304 as a phenomenological description of superconductivity.

The Landau-Ginzburg equations are based on the phenomenological form of


the Helmholtz free-energy functional expressed in terms of a superconducting
order parameter
Z  
β
F [∆] = F0 + d3 r α(T ) | ∆(r) |2 + | ∆(r) |4 + γ | ∇ ∆(r) |2 + . . .
2
(2699)
The gradient term allows a slow variation of the order parameter, and is writ-
ten in a form appropriate for cubic crystals. For a homogeneous system, for
temperatures close to Tc , the order parameter is homogeneous and small. Only
the leading terms in the expansion in powers of ∆(T ) need be retained. On
minimizing the Helmholtz free-energy with respect to ∆∗ , one finds
α(T )
| ∆(T ) |2 = − (2700)
β
and the free-energy simply becomes
α(T )2
F [∆] = F0 − V (2701)

One can identify the bulk superconducting order parameter with the B.C.S.
result
| ∆(T ) |2 = 10.2 kB
2
Tc ( Tc − T ) (2702)
close to the transition temperature. Furthermore, since the difference in the
Free-energy between the normal and superconducting state is given by the crit-
ical field energy density, one has
α(T )2 Hc (T )2
= (2703)
2β 8π
Hence, one can identify the coefficients in the Landau free-energy functional as
 
T − Tc
V α(T ) = ρ(µ) (2704)
Tc
and
V β = 0.098 ρ(µ) ( kB Tc )−2 (2705)
For a pure metal, the coefficient of the gradient term has a magnitude given by
V γ ∼ ρ(µ) ξ02 (2706)
303 L. P. Gor’kov, J.E.T.P. 9, 1364 (1960), L. P. Gor’kov, J.E.T.P. 10, 593 (1960).
304 V. I. Ginzburg and L. D. Landau, Zh. Eksp. i. Teor. Fiz. 20, 1064 (1950).

763
In the presence of a static magnetic field, the vector potential is given by the
solution of
B = ∇ ∧ A (2707)
However, the solution is invariant under the gauge transformation

A → A0 = A + ∇Λ (2708)

Under this transformation, the one-electron wave functions are also transformed
according to
 
|e|
φ(r) → φ0 (r) = exp − i Λ φ(r) (2709)
2 π h̄ c

Hence, the gap also acquires a phase through the gauge transformation
 
|e|
∆(r) → ∆0 (r) = exp − 2 i Λ ∆(r) (2710)
2 π h̄ c

Since the free energy must be gauge invariant, in the presence of fields, the
gradient terms must be replaced by the gauge invariant terms

2|e|
− i h̄ ∇ → − i h̄ ∇ + A (2711)
c
Thus, on adding the field energy, one has the Ginzburg-Landau Free energy
functional
Z 
β
F [∆] = F0 + d3 r α(T ) | ∆(r) |2 + | ∆(r) |4
2
2
B(r)2
  
γ 2|e|
+ − i h̄ ∇ + A ∆(r) +
h̄2 c 8π
(2712)

In this expression B is the internal field. The original formulation of Ginzburg


and Landau only contained a factor of e rather than 2 e in the gradient term.
The factor of two occurs since the charge on a Cooper pair is 2 e.

25.8.1 Extremal Configurations


In equilibrium, the free energy functional is to be minimized. The minimization
of the free energy requires that the variation of both the order parameter and
the field distributions are extrema of the functional. On making the variations,
of the fields from their extremal values

∆(r) = ∆ext (r) + δ∆(r)


A(r) = Aext (r) + δA(r) (2713)

764
one finds the first order functional derivatives are given by
δF
= α ∆ext (r) + β | ∆ext (r) |2 ∆ext (r)
δ∆∗ (r)
 2
γ 2|e|
+ − i h̄ ∇ + Aext ∆ext (r)
h̄2 c
 
δF 1 2|e|γ ∗ 2|e|
= ∇ ∧ B ext (r) + ∆ ext (r) − i h̄ ∇ + Aext ∆ext (r)
δA(r) 4π h̄2 c c
 
2|e|γ 2|e|
+ ∆ ext (r) + i h̄ ∇ + A ext ∆∗ext (r) (2714)
h̄2 c c

if the fields satisfy appropriate boundary conditions at infinity. On utilizing the


static form of Maxwell’s equation

∇ ∧ B = j (2715)
c
the Euler-Lagrange equations become
 2
γ 2|e|
0 = − i h̄ ∇ + A ext ∆ext (r) + α ∆ext (r) + β | ∆ext (r) |2 ∆ext (r)
h̄2 c
8 e2 γ
 
2|e|γ
j = i ∆∗ext (r) ∇ ∆ext (r) − ∆ext (r) ∇ ∆∗ext (r) − | ∆ext (r) |2 Aext
h̄ h̄2 c
(2716)

The extremal fields are the measurable physical fields. Hence, the subscript can
now be dropped without any ensuing ambiguity.

The above equations show that, close so the critical temperature, the gap
parameter satisfies a non-linear Schrodinger equation. That is, the gap satis-
fies a one-particle Schrodinger equation which includes a Hartree-like term due
to point contact interactions. The relation between the gap and the electrical
current is also very similar to the relation between a one-particle wave function
and the current. It is customary to write the gap, ∆, as being proportional to
the Cooper pair wave function Ψ.

25.8.2 Characteristic Length Scales


The Landau-Ginzburg equations contain two characteristic length scales. These
are the coherence length ξ which governs the relaxation of ∆ around an inhomo-
geneity. The coherence length can be found by examining the one-dimensional
Landau-Ginzburg theory. For a homogeneous system in equilibrium, the order
parameter has the value
α
| ∆0 |2 = − > 0 (2717)
β

765
for T < Tc . On scaling the order parameter to the bulk value

∆(z) = Ψ(z) ∆0 (2718)

the Landau-Ginzburg equation


∂2 ∆
−γ + α ∆(z) + β ∆(z) | ∆(z) |2 = 0 (2719)
∂z 2
reduces to
γ ∂2 Ψ
− Ψ(z) + Ψ(z) | Ψ(z) |2 = 0 (2720)
α ∂z 2
The coherence length ξ(T ) is defined as
γ
ξ 2 (T ) = − > 0 (2721)
α
The coherence length is given in terms of the B.C.S. coherence length ξ0 via
  12
Tc
ξ(T ) = 0.74 ξ0 (2722)
Tc − T
Thus, the coherence length diverges at the superconducting transition temper-
ature.

The spatial variation of Ψ can be found by imposing a boundary condition


at an inhomogeneity, say Ψ(0) = 0. The equation can be integrated using the
integrating factor  ∗
∂Ψ
(2723)
∂z
leading to
∂Ψ 2

1 2 1 2 1 4
− ξ (T ) ∂z − 2 | Ψ(z) | + 4 | Ψ(z) | = C
(2724)
2
where C is a constant of integration. It is assumed that as z → ∞, one recovers
the bulk superconductivity. Hence, one has limz → ∞ Ψ(z) → 1. This fixes
the constant of integration C = − 14 . Hence, the equation can be written as a
complete square
∂Ψ 2
 2
2 1 2
ξ (T )
= 1 − |Ψ(z) | (2725)
∂z 2
The pair of first order differential equation have the solutions
 
z
Ψ(z) = ± tanh √ (2726)
2 ξ(T )
Thus, the near a superconducting - normal metal interface, the order parameter
varies over the length scale provided by the temperature dependent coherence

766
length, ξ(T ).

The penetration depth can be obtained by assuming a constant value for the
order parameter. In these circumstances, the Landau-Ginzburg equations yield
a diamagnetic current given by
8 e2 γ 2
j(r, t) = − ∆0 A(r, t) (2727)
h̄ c
This is the London equation, if one identifies the phenomenological constants
ns e2 8 e2 ∆20 γ
= (2728)
mc h̄ c
This provides a local relation between the current density and the electromag-
netic field. On taking the curl of the equation, one obtains
8 e2 γ 2
∇ ∧ j = − ∆0 B(r, t) (2729)
h̄ c
Hence, on using Maxwell’s equations, one finds that the magnetic field will only
penetrate a distance λL inside the superconductor
 
z
Bx (z) = Bx (0) exp − (2730)
λL
where λl (T ) is given by

32 π e2 ∆20 γ
λL (T )−2 = (2731)
h̄ c2
The B.C.S. theory yields, the temperature dependence of the penetration length
as  
1 Tc
λL (T ) = √ λL (0) (2732)
2 Tc − T
which also diverges at Tc .

From the above analysis, once concludes that the ratio of λL (T ) to ξ(T ) is
almost temperature independent close to Tc . The ratio
λL (T )
κ = (2733)
ξ(T )
is the Landau - Ginzburg parameter. For a pure material, the value of κ is
evaluated in terms of the zero temperature values
λL (0)
κ = 0.96 (2734)
ξ0
Hence, there is at most only a slight temperature variation.

767
Type I superconductors are classified as those where κ > √12 , and type II
superconductors are those for which κ < √12 . The value of κ can be obtained
directly from the experimentally measured values of Hc (T ) and λL (T ) from
√ |e|
κ = 2 2 Hc (T ) λ2L (T ) (2735)
h̄ c

25.8.3 The Surface Energy


The Landau-Ginzburg equation can be used to calculate the energy of a sur-
face separating a normal and superconducting region. The normal material is
assumed to be located at z < 0 and the superconducting region is located at
z > 0. The boundary conditions are

for z → − ∞
∆ = 0
B = Hc (2736)

for the normal metal and

for z → + ∞
∆ = ∆0
B = 0 (2737)

in the superconductor.

Inside the bulk superconductor, the free energy is given by


Z 
β
F [∆] = F0 + d r α(T ) | ∆(r) |2 +
3
| ∆(r) |4
2
  2 2

γ 2|e| + B(r)

+ − i h̄ ∇ + A ∆(r)
h̄2 c 8π
(2738)

However, since the field penetrates into the superconductor, there should also be
a magnetic contribution due to the magnetization produced by Hc , This energy
is
B − Hc
− M Hc = − Hc (2739)

The bulk normal and superconducting states have the same Gibbs free-energy,
since B = Hc , therefore they can coexist. At the boundary, the energy will
vary as both ∆(z) and B(z) vary. Hence, the total energy is just
Z 
β
G[∆] = F0 + 3
d r α(T ) | ∆(r) |2 + | ∆(r) |4
2

768
2
B(r)2 Hc2
  
γ 2|e| B Hc
+ 2 − i h̄ ∇ + A ∆(r) +
− +
h̄ c 8π 4π 4π
(2740)

The surface energy, per unit area, is found by subtracting the bulk energy
density which is
Hc2
(2741)

for both phases. After dividing the free energy by the area, the surface energy
σ is found as
Z ∞ 
β
σ = dz α(T ) | ∆(z) |2 + | ∆(z) |4
−∞ 2
∂∆ 2 2 2

+ γ + 4 e γ | ∆(z) |2 A2 + ( B(z) − Hc )
∂z h̄2 c2 8π
(2742)

This energy has different signs depending on whether κ is greater or smaller


than √12 . For a type I superconductor, the surface energy is positive. For a type
II superconductor the surface energy is negative.

The surface energy in the extreme type I limit can be easily calculated. Since,
the penetration depth is vanishingly small, it provides a negligible contribution
to the surface energy. Hence, for κ  1, the surface energy is simply given by
∂∆ 2
Z ∞  2

β + Hc
σ ∼ dz α(T ) | ∆(z) |2 + | ∆(z) |4 + γ
0 2 ∂z 8π
(2743)

Since the condensation energy density is such that

α(T ) 2 Hc (T )2
∆0 (T ) = − (2744)
2 8π
the surface energy can be scaled to
Z ∞   2  2 
∂Ψ 1
σ = − α(T ) ∆20 (T ) dz ξ 2 (T ) + 1 − Ψ2 (z) (2745)
0 ∂z 2
The spatial variation of the order parameter is given by
 
z
Ψ(z) = tanh √ (2746)
2 ξ(T )
After evaluating the integrals, one finds the surface energy is positive

4 2 Hc (T )2
σ = ξ(T ) (2747)
3 8π

769
Hence, a type I superconductor minimizes the number of domain walls.

On the other hand, in the extreme type II limit, the surface energy can be
estimated by using the London equation. In this limit, the superconducting
order parameter can be set to the bulk value. The surface energy reduces to
Z ∞
Hc2 (T ) 4 e2 γ ( B(z) − Hc )2
 
σ = dz − + 2 2 | ∆0 |2 A2 +
−∞ 8π h̄ c 8π
(2748)

Using the definition of the London penetration depth, the surface free energy
can be re-written as
Z ∞
Hc2 (T ) Hc2 (T ) 2 ( B(z) − Hc )2
 
σ = dz − + A (z) +
−∞ 8π 8 π λ2L 8π
(2749)

From the London equations, one finds that the spatial variation of the magnetic
field and vector potential is given by
 
z
Ay (z) = λL Hc (T ) exp −
λL
 
z
Bx (z) = Hc (T ) exp − (2750)
λL

The surface energy is found, from an elementary integration, to be given by

Hc2 (T )
σ = − λL (T ) (2751)

Thus, for κ(T )  1 the surface energy is negative. Hence, in the presence of a
field, a type II superconductor will allow the creation of vortices or form many
microscopic domains.

The macroscopic properties of the two classes of superconductors are dra-


matically different.

25.8.4 The Little-Parks Experiment


The Little-Parks experiment demonstrates flux quantization in a superconduc-
tor305 , and thereby provides a direct experimental confirmation of the charge
on a Cooper pair. Consider a superconductor in the shape of a thin cylindrical
shell of thickness d and radius R. A magnetic field, parallel to the cylindrical
axis, threads through the hollow center of the cylinder. The superconducting
305 W. Little and R. D. Parks, Phys. Rev. 133 A, 97 (1964).

770
gap is assumed to have a constant magnitude, ∆, but has a spatially varying
phase  
∆(r) = | ∆ | exp i θ(r) (2752)

Inside the superconductor, the free energy is given by


Z 
β
F [∆] = F0 + d3 r α(T ) | ∆ |2 + | ∆ |4
2
2
B(r)2
 
γ 2|e| 2
+ h̄ ∇ θ + A |∆| +
h̄2 c 8π
(2753)
The magnetic field induces a current jϕ which wraps around the cylinder. The
current is given by
8 e2 γ
 
2|e|γ ∗ ∗
j = i ∆ (r) ∇ ∆(r) − ∆(r) ∇ ∆ (r) − | ∆(r) |2 A
h̄ h̄2 c
4|e|γ 8 e2 γ
= − | ∆ |2 ∇ θ(r) − | ∆ |2 A
h̄ h̄2 c
(2754)
On integrating the current around a loop encircling the cylinder, one finds
4|e|γ 8 e2 γ
jϕ 2 π R = − | ∆ |2 δθ − | ∆ |2 Φ (2755)
h̄ h̄2 c
where δθ is the change of the phase on going around the loop. In deriving this
expression, we have used Stokes’s theorem
I Z
A . dr = d2 S . B

= π R2 B
= Φ (2756)
where Φ is the total flux enclosed by the cylinder. Since the order parameter
must be single valued, then on performing one loop around the cylinder’s axis
the phase must change by an integer multiple of 2 π,
δθ = 2 π n (2757)
for some integer value of n. Hence, one has
 
8π|e|γ 2|e|
jϕ 2 π R = − | ∆ |2 n + Φ (2758)
h̄ 2 π h̄ c
Hence, on defining the superconducting flux quantum as
2 π h̄ c
Φs = (2759)
2|e|

771
one has  
8π|e|γ Φ
jϕ 2 π R = − | ∆ |2 n + (2760)
h̄ Φs
The expression for the current is equivalent to the gradient terms in the free
energy, since

∇θ = n êϕ
R
RB
A = êϕ (2761)
2
so
Z 
β
F [∆] = F0 + d3 r α(T ) | ∆ |2 +
| ∆ |4
2
2
4 π2 γ B(r)2
 
2|e| 2 2
+ n + πR B |∆| +
R2 2 π h̄ c 8π
(2762)

Thus, one finds that the phase variation of the order parameter is such as to
minimize | n + ΦΦs |. That is, n will jump discontinuously once, every time Φ is
increased by Φs . Hence, the current is a periodic function of the flux Φ. Since n
is now fixed, the free energy can be minimized with respect to the magnitude ∆.
The non-zero kinetic energy term increases the coefficient of the term quadratic
in the order parameter, whereas the quartic term is unaffected.

4 π2 γ
 
Φ 2 β 4
α(T ) + 2
( n + ) ∆2 + ∆ (2763)
R Φs 2
Hence, the gap parameter is given by

4 π2 γ
 
Φ 2
∆2 = − α(T ) + ( n + ) β −1 (2764)
R2 Φs
and is a periodic function of the flux. The variation of the gap is measurable. In
fact, the critical temperature Tc (H) at which the gap vanishes is also a periodic
function of Φ. The periodic variation of resistance as a function of the field was
observed by Little and Parks306 .

25.8.5 The Critical Current


If a current is run through a superconductor, there is a maximum value of the
current that can be drawn while the material remains superconducting. For
currents larger than the critical current, the material transforms to the normal
state.

306 W. Little and R. D. Parks, Phys. Rev. 133 A, 97 (1964).

772
We shall assume the sample has the geometry of a thin film. The supercon-
ducting gap is assumed to have a constant magnitude, ∆, but has a spatially
varying phase  
∆(r) = | ∆ | exp i θ(r) (2765)

Inside the superconductor, the free energy is given by


Z  
3 2 β 4 2 2
F [∆] = F0 + d r α(T ) | ∆ | + |∆| + γ(∇θ) |∆|
2
(2766)
The current is related to the spatially varying phase
4|e|γ
j = − | ∆ |2 ∇ θ(r) (2767)

The relation between the magnitude of the order parameter and the gradient
of the phase can be found by minimizing the free-energy, with respect to the
amplitude at constant pase gradient. This leads to the equation
γ ( ∇ θ )2 + α + β | ∆ |2 = 0 (2768)
which can be solved for ∇ θ as
  12
1 2
∇θ = √ − α − β|∆| (2769)
γ
Hence, the current is given in terms of the amplitude by
√  12
4|e| γ

j = − | ∆ |2 − α − β | ∆ |2 (2770)

This expression shows that there is a maximum value of the current, as a func-
tion of | ∆ |. The maximum value of the current is found for

| ∆ |2 = − (2771)

and is given by
√ 1
8|e| γα

α 2
jc = − (2772)
3 h̄ β 3
Thus, if | ∆ | is finite, there is a maximum value of the current. For zero current,
the magnitude of the order parameter is given by the value
α
| ∆ |2 = − (2773)
β
and decreases with increasing j, until j reaches jc , at which point

| ∆ |2 = − (2774)

For larger values of j, the system becomes unstable to the normal state.

773

Вам также может понравиться