Sei sulla pagina 1di 506

Condensed Matter Physics I

Peter S. Riseborough
November 21, 2002

Contents
1 Introduction 9
1.1 The Born-Oppenheimer Approximation . . . . . . . . . . . . . . 9

2 Crystallography 13

3 Structures 13
3.1 Fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2 Crystalline Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3 The Direct Lattice . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3.1 Primitive Unit Cells . . . . . . . . . . . . . . . . . . . . . 19
3.3.2 The Wigner-Seitz Unit Cell . . . . . . . . . . . . . . . . . 19
3.4 Symmetry of Crystals . . . . . . . . . . . . . . . . . . . . . . . . 21
3.4.1 Symmetry Groups . . . . . . . . . . . . . . . . . . . . . . 21
3.4.2 Group Multiplication Tables . . . . . . . . . . . . . . . . 22
3.4.3 Point Group Operations . . . . . . . . . . . . . . . . . . . 23
3.4.4 Limitations Imposed by Translational Symmetry . . . . . 24
3.4.5 Point Group Nomenclature . . . . . . . . . . . . . . . . . 24
3.5 Bravais Lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5.1 Exercise 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5.2 Cubic Bravais Lattices. . . . . . . . . . . . . . . . . . . . 28
3.5.3 Tetragonal Bravais Lattices. . . . . . . . . . . . . . . . . . 30
3.5.4 Orthorhombic Bravais Lattices. . . . . . . . . . . . . . . . 31
3.5.5 Monoclinic Bravais Lattice. . . . . . . . . . . . . . . . . . 32
3.5.6 Triclinic Bravais Lattice. . . . . . . . . . . . . . . . . . . . 32
3.5.7 Trigonal Bravais Lattice. . . . . . . . . . . . . . . . . . . . 33
3.5.8 Hexagonal Bravais Lattice. . . . . . . . . . . . . . . . . . 33
3.5.9 Exercise 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.6 Point Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.7 Space Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.8 Crystal Structures with Bases. . . . . . . . . . . . . . . . . . . . 39
3.8.1 Diamond Structure . . . . . . . . . . . . . . . . . . . . . . 39
3.8.2 Exercise 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.8.3 Graphite Structure . . . . . . . . . . . . . . . . . . . . . . 40

1
3.8.4 Exercise 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.8.5 Hexagonal Close-Packed Structure . . . . . . . . . . . . . 41
3.8.6 Exercise 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.8.7 Exercise 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.8.8 Other Close-Packed Structures . . . . . . . . . . . . . . . 43
3.8.9 Sodium Chloride Structure . . . . . . . . . . . . . . . . . 45
3.8.10 Cesium Chloride Structure . . . . . . . . . . . . . . . . . 45
3.8.11 Fluorite Structure . . . . . . . . . . . . . . . . . . . . . . 47
3.8.12 The Copper Three Gold Structure . . . . . . . . . . . . . 47
3.8.13 Rutile Structure . . . . . . . . . . . . . . . . . . . . . . . 48
3.8.14 Zinc Blende Structure . . . . . . . . . . . . . . . . . . . . 48
3.8.15 Zincite Structure . . . . . . . . . . . . . . . . . . . . . . . 49
3.8.16 The Perovskite Structure . . . . . . . . . . . . . . . . . . 50
3.8.17 Exercise 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.9 Lattice Planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
3.9.1 Exercise 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.9.2 Exercise 9 . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
3.10 Quasi-Crystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4 Structure Determination 56
4.1 X Ray Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.1.1 The Bragg conditions . . . . . . . . . . . . . . . . . . . . 56
4.1.2 The Laue conditions . . . . . . . . . . . . . . . . . . . . . 57
4.1.3 Equivalence of the Bragg and Laue conditions . . . . . . . 59
4.1.4 The Ewald Construction . . . . . . . . . . . . . . . . . . . 60
4.1.5 X-ray Techniques . . . . . . . . . . . . . . . . . . . . . . . 61
4.1.6 Exercise 10 . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.1.7 The Structure and Form Factors . . . . . . . . . . . . . . 62
4.1.8 Exercise 11 . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.1.9 Exercise 12 . . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.1.10 Exercise 13 . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.1.11 Exercise 14 . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.1.12 Exercise 15 . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.1.13 Exercise 16 . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.1.14 Exercise 17 . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.2 Neutron Diffraction . . . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3 Theory of the Differential Scattering Cross-section . . . . . . . . 76
4.3.1 Time Dependent Perturbation Theory . . . . . . . . . . . 77
4.3.2 The Fermi-Golden Rule . . . . . . . . . . . . . . . . . . . 78
4.3.3 The Elastic Scattering Cross-Section . . . . . . . . . . . . 80
4.3.4 The Condition for Coherent Scattering . . . . . . . . . . . 83
4.3.5 Exercise 18 . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.3.6 Exercise 19 . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.3.7 Exercise 20 . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.3.8 Anti-Domain Phase Boundaries . . . . . . . . . . . . . . . 86
4.3.9 Exercise 21 . . . . . . . . . . . . . . . . . . . . . . . . . . 87

2
4.4 Elastic Scattering from Quasi-Crystals . . . . . . . . . . . . . . . 88
4.5 Elastic Scattering from a Fluid . . . . . . . . . . . . . . . . . . . 90

5 The Reciprocal Lattice 93


5.0.1 Exercise 22 . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.1 The Reciprocal Lattice as a Dual Lattice . . . . . . . . . . . . . . 94
5.1.1 Exercise 23 . . . . . . . . . . . . . . . . . . . . . . . . . . 97
5.2 Examples of Reciprocal Lattices . . . . . . . . . . . . . . . . . . . 97
5.2.1 The Simple Cubic Reciprocal Lattice . . . . . . . . . . . . 97
5.2.2 The Body Centered Cubic Reciprocal Lattice . . . . . . . 98
5.2.3 The Face Centered Cubic Reciprocal Lattice . . . . . . . 98
5.2.4 The Hexagonal Reciprocal Lattice . . . . . . . . . . . . . 99
5.2.5 Exercise 24 . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.3 The Brillouin Zones . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.3.1 The Simple Cubic Brillouin Zone . . . . . . . . . . . . . . 100
5.3.2 The Body Centered Cubic Brillouin Zone . . . . . . . . . 101
5.3.3 The Face Centered Cubic Brillouin Zone . . . . . . . . . . 101
5.3.4 The Hexagonal Brillouin Zone . . . . . . . . . . . . . . . . 102

6 Electrons 103

7 Electronic States 103


7.1 Many Electron Wave Functions . . . . . . . . . . . . . . . . . . . 104
7.1.1 Exercise 25 . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.2 Bloch’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
7.3 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . 114
7.4 Plane Wave Expansion of Bloch Functions . . . . . . . . . . . . . 116
7.5 The Bloch Wave Vector . . . . . . . . . . . . . . . . . . . . . . . 118
7.6 The Density of States . . . . . . . . . . . . . . . . . . . . . . . . 120
7.6.1 Exercise 26 . . . . . . . . . . . . . . . . . . . . . . . . . . 122
7.7 The Fermi-Surface . . . . . . . . . . . . . . . . . . . . . . . . . . 123

8 Approximate Models 125


8.1 The Nearly Free Electron Model . . . . . . . . . . . . . . . . . . 125
8.1.1 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . 125
8.1.2 Non-Degenerate Perturbation Theory . . . . . . . . . . . 126
8.1.3 Degenerate Perturbation Theory . . . . . . . . . . . . . . 129
8.1.4 Empty Lattice Approximation Band Structure . . . . . . 132
8.1.5 Exercise 27 . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.1.6 Degeneracies of the Bloch States . . . . . . . . . . . . . . 137
8.1.7 Exercise 28 . . . . . . . . . . . . . . . . . . . . . . . . . . 145
8.1.8 Brillouin Zone Boundaries . . . . . . . . . . . . . . . . . . 146
8.1.9 The Geometric Structure Factor . . . . . . . . . . . . . . 148
8.1.10 Exercise 29 . . . . . . . . . . . . . . . . . . . . . . . . . . 150
8.1.11 Exercise 30 . . . . . . . . . . . . . . . . . . . . . . . . . . 152
8.1.12 Exercise 31 . . . . . . . . . . . . . . . . . . . . . . . . . . 153

3
8.1.13 Exercise 32 . . . . . . . . . . . . . . . . . . . . . . . . . . 154
8.1.14 Exercise 33 . . . . . . . . . . . . . . . . . . . . . . . . . . 154
8.2 The Pseudo-Potential Method . . . . . . . . . . . . . . . . . . . . 155
8.2.1 The Scattering Approach . . . . . . . . . . . . . . . . . . 159
8.2.2 The Ziman-Lloyd Pseudo-potential . . . . . . . . . . . . . 160
8.2.3 Exercise 34 . . . . . . . . . . . . . . . . . . . . . . . . . . 161
8.3 The Tight-Binding Model . . . . . . . . . . . . . . . . . . . . . . 162
8.3.1 Tight-Binding s Band Metal . . . . . . . . . . . . . . . . . 167
8.3.2 Exercise 35 . . . . . . . . . . . . . . . . . . . . . . . . . . 170
8.3.3 Exercise 36 . . . . . . . . . . . . . . . . . . . . . . . . . . 170
8.3.4 Exercise 37 . . . . . . . . . . . . . . . . . . . . . . . . . . 171
8.3.5 Exercise 38 . . . . . . . . . . . . . . . . . . . . . . . . . . 172
8.3.6 Exercise 39 . . . . . . . . . . . . . . . . . . . . . . . . . . 172
8.3.7 Exercise 40 . . . . . . . . . . . . . . . . . . . . . . . . . . 173
8.3.8 Wannier Functions . . . . . . . . . . . . . . . . . . . . . . 173
8.3.9 Exercise 41 . . . . . . . . . . . . . . . . . . . . . . . . . . 175

9 Electron-Electron Interactions 176


9.1 The Landau Fermi Liquid . . . . . . . . . . . . . . . . . . . . . . 176
9.1.1 The Scattering Rate . . . . . . . . . . . . . . . . . . . . . 177
9.1.2 The Quasi-Particle Energy . . . . . . . . . . . . . . . . . 177
9.1.3 Exercise 42 . . . . . . . . . . . . . . . . . . . . . . . . . . 180
9.2 The Hartree-Fock Approximation . . . . . . . . . . . . . . . . . . 180
9.2.1 The Free Electron Gas. . . . . . . . . . . . . . . . . . . . 184
9.2.2 Exercise 43 . . . . . . . . . . . . . . . . . . . . . . . . . . 192
9.3 The Density Functional Method . . . . . . . . . . . . . . . . . . . 192
9.3.1 Hohenberg-Kohn Theorem . . . . . . . . . . . . . . . . . . 193
9.3.2 Functionals and Functional Derivatives . . . . . . . . . . 195
9.3.3 The Variational Principle . . . . . . . . . . . . . . . . . . 198
9.3.4 The Electrostatic Terms . . . . . . . . . . . . . . . . . . . 200
9.3.5 The Kohn-Sham Equations . . . . . . . . . . . . . . . . . 202
9.3.6 The Local Density Approximation . . . . . . . . . . . . . 204
9.4 Static Screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
9.4.1 The Thomas-Fermi Approximation . . . . . . . . . . . . . 208
9.4.2 Linear Response Theory . . . . . . . . . . . . . . . . . . . 211
9.4.3 Density Functional Response Function . . . . . . . . . . . 213
9.4.4 Exercise 44 . . . . . . . . . . . . . . . . . . . . . . . . . . 215
9.4.5 Exercise 45 . . . . . . . . . . . . . . . . . . . . . . . . . . 215

10 Stability of Structures 217


10.1 Momentum Space Representation . . . . . . . . . . . . . . . . . . 217
10.2 Real Space Representation . . . . . . . . . . . . . . . . . . . . . . 222

4
11 Metals 228
11.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
11.1.1 The Sommerfeld Expansion . . . . . . . . . . . . . . . . . 229
11.1.2 The Specific Heat Capacity . . . . . . . . . . . . . . . . . 231
11.1.3 Exercise 46 . . . . . . . . . . . . . . . . . . . . . . . . . . 234
11.1.4 Exercise 47 . . . . . . . . . . . . . . . . . . . . . . . . . . 234
11.1.5 Pauli Paramagnetism . . . . . . . . . . . . . . . . . . . . 234
11.1.6 Exercise 48 . . . . . . . . . . . . . . . . . . . . . . . . . . 237
11.1.7 Exercise 49 . . . . . . . . . . . . . . . . . . . . . . . . . . 237
11.1.8 Landau Diamagnetism . . . . . . . . . . . . . . . . . . . . 238
11.1.9 Landau Level Quantization . . . . . . . . . . . . . . . . . 239
11.1.10 The Diamagnetic Susceptibility . . . . . . . . . . . . . . . 241
11.2 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . 244
11.2.1 Electrical Conductivity . . . . . . . . . . . . . . . . . . . 244
11.2.2 Scattering by Static Defects . . . . . . . . . . . . . . . . . 244
11.2.3 Exercise 50 . . . . . . . . . . . . . . . . . . . . . . . . . . 251
11.2.4 The Hall Effect and Magneto-resistance. . . . . . . . . . . 252
11.2.5 Multi-band Models . . . . . . . . . . . . . . . . . . . . . . 260
11.3 Electromagnetic Properties of Metals . . . . . . . . . . . . . . . . 263
11.3.1 The Longitudinal Response . . . . . . . . . . . . . . . . . 266
11.3.2 Electron Scattering Experiments . . . . . . . . . . . . . . 274
11.3.3 Exercise 51 . . . . . . . . . . . . . . . . . . . . . . . . . . 278
11.3.4 Exercise 52 . . . . . . . . . . . . . . . . . . . . . . . . . . 280
11.3.5 The Transverse Response . . . . . . . . . . . . . . . . . . 285
11.3.6 Optical Experiments . . . . . . . . . . . . . . . . . . . . . 288
11.3.7 Kramers-Kronig Relation . . . . . . . . . . . . . . . . . . 289
11.3.8 Exercise 53 . . . . . . . . . . . . . . . . . . . . . . . . . . 290
11.3.9 Exercise 54 . . . . . . . . . . . . . . . . . . . . . . . . . . 291
11.3.10 The Drude Conductivity . . . . . . . . . . . . . . . . . . . 291
11.3.11 Exercise 55 . . . . . . . . . . . . . . . . . . . . . . . . . . 296
11.3.12 Exercise 56 . . . . . . . . . . . . . . . . . . . . . . . . . . 296
11.3.13 The Anomalous Skin Effect . . . . . . . . . . . . . . . . . 297
11.3.14 Inter-Band Transitions . . . . . . . . . . . . . . . . . . . . 299
11.4 Measuring the Fermi-Surface . . . . . . . . . . . . . . . . . . . . 300
11.4.1 Semi-Classical Orbits . . . . . . . . . . . . . . . . . . . . 301
11.4.2 de Haas - van Alphen Oscillations . . . . . . . . . . . . . 305
11.4.3 Exercise 57 . . . . . . . . . . . . . . . . . . . . . . . . . . 307
11.4.4 The Lifshitz-Kosevich Formulae . . . . . . . . . . . . . . . 308
11.4.5 Other Fermi-Surface Probes . . . . . . . . . . . . . . . . . 313
11.4.6 Cyclotron Resonances . . . . . . . . . . . . . . . . . . . . 315
11.5 The Quantum Hall Effect . . . . . . . . . . . . . . . . . . . . . . 319
11.5.1 The Integer Quantum Hall Effect . . . . . . . . . . . . . . 319
11.5.2 Exercise 58 . . . . . . . . . . . . . . . . . . . . . . . . . . 325
11.5.3 The Fractional Quantum Hall Effect . . . . . . . . . . . . 326
11.5.4 Quasi-Particle Excitations . . . . . . . . . . . . . . . . . . 328
11.5.5 Skyrmions . . . . . . . . . . . . . . . . . . . . . . . . . . . 330

5
11.5.6 Composite Fermions . . . . . . . . . . . . . . . . . . . . . 338

12 Insulators 341
12.1 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
12.1.1 Holes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
12.1.2 Intrinsic Semiconductors . . . . . . . . . . . . . . . . . . . 347
12.1.3 Extrinsic Semiconductors . . . . . . . . . . . . . . . . . . 349
12.1.4 Exercise 59 . . . . . . . . . . . . . . . . . . . . . . . . . . 352
12.2 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . 353
12.3 Optical Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 353

13 Phonons 354

14 Harmonic Phonons 354


14.1 Lattice with a Basis . . . . . . . . . . . . . . . . . . . . . . . . . 360
14.2 A Sum Rule for the Dispersion Relations . . . . . . . . . . . . . . 360
14.2.1 Exercise 60 . . . . . . . . . . . . . . . . . . . . . . . . . . 363
14.3 The Nature of the Phonon Modes . . . . . . . . . . . . . . . . . . 363
14.3.1 Exercise 61 . . . . . . . . . . . . . . . . . . . . . . . . . . 364
14.3.2 Exercise 62 . . . . . . . . . . . . . . . . . . . . . . . . . . 364
14.3.3 Exercise 63 . . . . . . . . . . . . . . . . . . . . . . . . . . 365
14.3.4 Exercise 64 . . . . . . . . . . . . . . . . . . . . . . . . . . 365
14.3.5 Exercise 65 . . . . . . . . . . . . . . . . . . . . . . . . . . 366
14.3.6 Exercise 66 . . . . . . . . . . . . . . . . . . . . . . . . . . 366
14.3.7 Exercise 67 . . . . . . . . . . . . . . . . . . . . . . . . . . 367
14.4 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
14.4.1 The Specific Heat . . . . . . . . . . . . . . . . . . . . . . 370
14.4.2 The Einstein Model of a Solid . . . . . . . . . . . . . . . . 371
14.4.3 The Debye Model of a Solid . . . . . . . . . . . . . . . . . 372
14.4.4 Exercise 68 . . . . . . . . . . . . . . . . . . . . . . . . . . 374
14.4.5 Exercise 69 . . . . . . . . . . . . . . . . . . . . . . . . . . 374
14.4.6 Exercise 70 . . . . . . . . . . . . . . . . . . . . . . . . . . 374
14.4.7 Exercise 71 . . . . . . . . . . . . . . . . . . . . . . . . . . 375
14.4.8 Lindemann Theory of Melting . . . . . . . . . . . . . . . . 375
14.4.9 Thermal Expansion . . . . . . . . . . . . . . . . . . . . . 377
14.4.10 Thermal Expansion of Metals . . . . . . . . . . . . . . . . 379
14.5 Anharmonicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
14.5.1 Exercise 72 . . . . . . . . . . . . . . . . . . . . . . . . . . 380

15 Phonon Measurements 381


15.1 Inelastic Neutron Scattering . . . . . . . . . . . . . . . . . . . . . 381
15.1.1 The Scattering Cross-Section . . . . . . . . . . . . . . . . 382
15.2 The Debye-Waller Factor . . . . . . . . . . . . . . . . . . . . . . 387
15.3 Single Phonon Scattering . . . . . . . . . . . . . . . . . . . . . . 389
15.4 Multi-Phonon Scattering . . . . . . . . . . . . . . . . . . . . . . . 390
15.4.1 Exercise 73 . . . . . . . . . . . . . . . . . . . . . . . . . . 391

6
15.4.2 Exercise 74 . . . . . . . . . . . . . . . . . . . . . . . . . . 391
15.4.3 Exercise 75 . . . . . . . . . . . . . . . . . . . . . . . . . . 391
15.5 Raman and Brillouin Scattering of Light . . . . . . . . . . . . . . 391

16 Phonons in Metals 394


16.1 Screened Ionic Plasmons . . . . . . . . . . . . . . . . . . . . . . . 395
16.1.1 Kohn Anomalies . . . . . . . . . . . . . . . . . . . . . . . 396
16.2 Dielectric Constant of a Metal . . . . . . . . . . . . . . . . . . . . 396
16.3 The Retarded Electron-Electron Interaction . . . . . . . . . . . . 399
16.4 Phonon Renormalization of Quasi-Particles . . . . . . . . . . . . 400
16.5 Electron-Phonon Interactions . . . . . . . . . . . . . . . . . . . . 402
16.6 Electrical Resistivity due to Phonon Scattering . . . . . . . . . . 403
16.6.1 Umklapp Scattering . . . . . . . . . . . . . . . . . . . . . 408
16.6.2 Phonon Drag . . . . . . . . . . . . . . . . . . . . . . . . . 409

17 Phonons in Semiconductors 410


17.1 Resistivity due to Phonon Scattering . . . . . . . . . . . . . . . . 410
17.2 Polarons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
17.3 Indirect Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . 411

18 Impurities and Disorder 413


18.1 Scattering By Impurities . . . . . . . . . . . . . . . . . . . . . . . 416
18.2 Virtual Bound States . . . . . . . . . . . . . . . . . . . . . . . . . 418
18.3 Disorder . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421
18.4 Coherent Potential Approximation . . . . . . . . . . . . . . . . . 422
18.5 Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 423
18.5.1 Anderson Model of Localization . . . . . . . . . . . . . . . 424
18.5.2 Scaling Theories of Localization . . . . . . . . . . . . . . . 425

19 Magnetic Impurities 428


19.1 Localized Magnetic Impurities in Metals . . . . . . . . . . . . . . 428
19.2 Mean Field Approximation . . . . . . . . . . . . . . . . . . . . . 428
19.2.1 The Atomic Limit . . . . . . . . . . . . . . . . . . . . . . 431
19.3 The Schrieffer-Wolf Transformation . . . . . . . . . . . . . . . . . 431
19.3.1 The Kondo Hamiltonian . . . . . . . . . . . . . . . . . . . 434
19.4 The Resistance Minimum . . . . . . . . . . . . . . . . . . . . . . 435

20 Collective Phenomenon 440

21 Itinerant Magnetism 440


21.1 Stoner Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
21.1.1 Exercise 76 . . . . . . . . . . . . . . . . . . . . . . . . . . 442
21.1.2 Exercise 77 . . . . . . . . . . . . . . . . . . . . . . . . . . 442
21.2 Linear Response Theory . . . . . . . . . . . . . . . . . . . . . . . 442
21.3 Magnetic Instabilities . . . . . . . . . . . . . . . . . . . . . . . . 444
21.4 Spin Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 447

7
21.5 The Heisenberg Model . . . . . . . . . . . . . . . . . . . . . . . . 449

22 Localized Magnetism 450


22.1 Holstein - Primakoff Transformation . . . . . . . . . . . . . . . . 451
22.2 Spin Rotational Invariance . . . . . . . . . . . . . . . . . . . . . . 454
22.2.1 Exercise 78 . . . . . . . . . . . . . . . . . . . . . . . . . . 457
22.3 Anti-ferromagnetic Spinwaves . . . . . . . . . . . . . . . . . . . . 458
22.3.1 Exercise 79 . . . . . . . . . . . . . . . . . . . . . . . . . . 460

23 Spin Glasses 460


23.1 Mean Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . 462
23.2 The Sherrington-Kirkpatrick Solution. . . . . . . . . . . . . . . . 463

24 Magnetic Neutron Scattering 467


24.1 The Inelastic Scattering Cross-Section . . . . . . . . . . . . . . . 467
24.1.1 The Dipole-Dipole Interaction . . . . . . . . . . . . . . . . 467
24.1.2 The Inelastic Scattering Cross-Section . . . . . . . . . . . 467
24.2 Time Dependent Spin Correlation Functions . . . . . . . . . . . . 471
24.3 The Fluctuation Dissipation Theorem . . . . . . . . . . . . . . . 473
24.4 Magnetic Scattering . . . . . . . . . . . . . . . . . . . . . . . . . 475
24.4.1 Neutron Diffraction . . . . . . . . . . . . . . . . . . . . . 475
24.4.2 Exercise 80 . . . . . . . . . . . . . . . . . . . . . . . . . . 476
24.4.3 Exercise 81 . . . . . . . . . . . . . . . . . . . . . . . . . . 477
24.4.4 Spin Wave Scattering . . . . . . . . . . . . . . . . . . . . 477
24.4.5 Exercise 82 . . . . . . . . . . . . . . . . . . . . . . . . . . 478
24.4.6 Critical Scattering . . . . . . . . . . . . . . . . . . . . . . 478

25 Superconductivity 480
25.1 Experimental Manifestation . . . . . . . . . . . . . . . . . . . . . 480
25.1.1 The London Equations . . . . . . . . . . . . . . . . . . . . 481
25.1.2 Thermodynamics of the Superconducting State . . . . . . 483
25.2 The Cooper Problem . . . . . . . . . . . . . . . . . . . . . . . . . 485
25.3 Pairing Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 489
25.3.1 The Pairing Interaction . . . . . . . . . . . . . . . . . . . 489
25.3.2 The B.C.S. Variational State . . . . . . . . . . . . . . . . 491
25.3.3 The Gap Equation . . . . . . . . . . . . . . . . . . . . . . 493
25.3.4 The Ground State Energy . . . . . . . . . . . . . . . . . . 494
25.4 Quasi-Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 496
25.4.1 Exercise 83 . . . . . . . . . . . . . . . . . . . . . . . . . . 499
25.5 Thermodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
25.6 Perfect Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . 501
25.7 The Meissner Effect . . . . . . . . . . . . . . . . . . . . . . . . . 503

26 Landau-Ginsberg Theory 504

8
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 a faster time
scale that governs the electronic motion. The large difference between 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 con-
stituents of different mass into electrically neutral material. The slow moving
nuclear masses can be considered to be quasi-static, and are responsible for
defining the structure of matter. In this approximation, the fast moving elec-
trons 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 Approximation. In this approximation, the
electronic states are treated as if the nuclei were at rest at fixed positions. How-
ever, when treating the slow motions of the nuclei, the electrons are considered
as adapting instantaneously to the potential of the charged nuclei, thereby min-
imizing 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
where Z is the nuclear charge and a is the radius of the atomic orbital. The stan-

9
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 quantized total electronic energy
of the hydrogen atom
Z e2 me v 2
Ee = − +
a 2
Z e2
= −
2a
me Z 2 e4
= − (3)
2 n2 h̄2
which is a standard result 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 e2
−α r (4)
a3
where α is a dimensionless constant. Hence, the equation of motion for the
displacement of a nuclei of mass Mn is
Z e2 d2 r
−α 3
r = Mn (5)
a dt2
which shows that the nuclei undergo harmonic oscillations with frequency
Z e2
ω2 = α (6)
Mn a3
The semi-classical quantization condition
I
dr Mn v = n h̄ (7)

yields the energy for nuclear motion as

EN = n h̄ ω
 12
me Z 2 e4 1

me
= n α2 (8)
h̄2 Mn

10
Thus, the ratio of the energies of nuclear motion to electronic motion are given
by the factor
 1
EN me 2
∼ (9)
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 Migdal.

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 charac-


ters of the equilibrium structures of the dressed nuclei in matter are 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 parti-
cles 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 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 order 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 neg-
ligible. In other materials, the potential due to the ionic charges can produce
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 can be
thought of sharing the properties 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 are introduced. It shall be shown how the
electronic transport properties of metals are dominated by states with energies
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 dress or screen 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 exci-

11
tation 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, and through collective action, they spontaneously break the sym-
metry of the Hamiltonian. In many cases, the spontaneously broken symmetry
is accompanied by the formation of a new branch of low energy excitations.

12
2 Crystallography
Crystallography is the study of 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 inter-
actions between the nuclei (dressed by their accompanying clouds of electrons),
one may assume that the nuclear or ionic motion can be treated classically. 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 between ions,
d, is of the order of angstroms, the uncertainty of the momentum is given by h̄d
and the kinetic energy EK for this quantum zero point motion is given by

h̄2
EK = (10)
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 temperatures. For these reasons, He behaves
like a quantum fluid. However, for the heavier nuclei, the quantum nature of
the particles manifest themselves in more subtle ways.

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

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 mat-
ter. Due to the near equality of these two length scales, the electron density 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 significant
variation that can be interpreted in terms of the electron density of isolated
atoms, subject to significant modifications when brought together. As the elec-
tron density for isolated atoms is usually spherically symmetric, the structure

13
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 macroscopic characteristics of fluids are
that they are spatially uniform and isotropic, which means that the average en-
vironment 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 ) (11)
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
to the time averaged positions of the atoms.

In particular, for a fluid, spatial homogeneity ensures that the average den-
sity ρ(r) at position r is equal to the average density at a displaced position
r + R,
ρ(r) = ρ(r + R) (12)
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 short-ranged structure which is exemplified by


locating one atom and then examining the positions of the neighboring atoms.
The local spatial correlations are expressed by the density - density correlation
function which is expressed as an average

C(r, r0 ) = ρ(r) ρ(r0 )

14
X
= δ 3 ( r − ri ) δ 3 ( r0 − rj )
i,j
(13)

Since fluids are homogeneous, the correlation function is only a function of the
difference of the positions r − r0 . Furthermore, since the 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 ) (14)

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

= δ 3 ( r − r0 ) ρ(r) (15)

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) (16)

For a system which possesses continuous translational invariance, the pair dis-
tribution function can be evaluated as
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 ) (17)
V
i6=j

Since the sum over i runs over each of the inter-atomic separations rj − ri for

15
each value of j, spatial homogeneity demands that the contribution from each
different j value is identical. There are N such terms, and this leads to the
expression for the pair distribution function involving an atom at the central
site r0 , and the others at sites i
X
g(r) = ρ(0) δ 3 ( r − ri + r0 ) (18)
i

where
N
ρ(0) = (19)
V
As this only depends on the radial distance | r | this is called the radial distri-
bution function g(r). For large r, the pair distribution function, like C(r, 0),
2
approaches ρ(0) , or
2
lim g(r) → ρ(0) (20)
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, unlike most other materials. In the liquid phase the density is
higher, 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 − (21)
r r
The potential has a short ranged repulsion between the atoms caused by the
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 . Another model potential that is
of use is the hard sphere potential which excludes the center of another atom
from the region of radius 2 a centered on the central atom.

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 liquid. On randomly packing spheres, one finds a packing
fraction of atoms given by 0.638. The packing fraction is defined as the total
volume of the hard spheres divided by the (minimum) volume that contains all
the spheres. Random packings of hard spheres can be used to calculate the
radial distribution function g(r). The model shows that there are strong short-
ranged correlations between the closest atoms and that there are 12 in three
dimensions at radial distances 2 a. These correlations show up as a strong peak

16
in the radial distribution function at 2 a, and there are other peaks correspond-
ing to the next few shells of neighboring atoms.

17
3.2 Crystalline Solids
A perfect crystal consists of a space filling periodic array of atoms. It can
be partitioned into identical individual structural units that can be repeatedly
stacked together to form the crystal. The structural unit is called the unit cell.
There are many alternate ways of performing the partitioning and, therefore,
there are many alternate forms for the unit cells. The unit cells which have the
smallest possible volume are called primitive unit cells. A unit cell may contain
one or more atoms. The positions of the atoms, when referenced to a specific
point in the unit cell, composes the basis of the lattice.

3.3 The Direct Lattice


Equivalent points 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 (22)

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 (n1 , n2 , n3 ) runs through all the positive and negative integers. The set of
translations R is 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 through a lattice vector Rm , 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 (23)

and the lattice points in the two frames are related via

n0i = ni + mi (24)

and as the numbers ni and n0i take on all possible integer values, the set of all
lattices 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

18
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 (25)

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
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.

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 × 81 = 1.

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


vectors via
Vc = | a1 . ( a2 ∧ a3 ) | (26)
The primitive unit cell is a unit cell of minimum volume.

3.3.2 The Wigner-Seitz Unit Cell


An alternate method of constructing a unit cell is due to Wigner and Seitz.
The Wigner-Seitz cell has the important property that there are no arbitrary
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 symmetry
as the lattice. The Wigner-Seitz unit cell is constructed by forming a set of

19
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 equations of the planes bisecting the vector from the central point to
the i-th lattice point is given by
 
1
r − R . Ri = 0 (27)
2 i

where Ri is the lattice vector. The sections of planes closest to the origin form
the surface of the Wigner Seitz-cell.

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 point than any other.

20
3.4 Symmetry of Crystals
A symmetry operation acts on a crystal producing a new crystal, shifting the
atoms to new positions such that the new crystal is identical in appearance to
the original crystal. That is, the positions of the atoms in the new crystal coin-
cide with the positions of similar atoms in the original crystal. 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 product of any two symmetry operators from the set, say A and B,
defined as A B = C then C 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
and third operators are combined before they are combined with the first.

A(BC ) = (AB)C (28)

(III) There exists a symmetry operator which leaves all the atoms in their
original places, called the identity operator E. The product of an arbitrarily
chosen symmetry operator of the group with the identity gives back the arbi-
trarily chosen operator.
AE = EA = A (29)

(IV) For each operator in the group, there exists a unique inverse operator
such that when the operator is combined with the inverse operator, they produce
the identity.
A A−1 = A−1 A = E (30)

21
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. Un-
der a translation which is not the identity, no point remains invariant. The
point group of the lattice consists of the set of symmetry operations in 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
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 element B.
E . . B .
. . . . .
A . . C . (31)
. . . . .
. . . . .

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 the dual role as the list of group
elements and also are the elements 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
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


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
between the two H atoms. This is a two-fold axis since a second rotation by π is
equivalent to the identity. The two-fold rotation 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

22
is symmetric under reflection in a mirror plane passing 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, 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. A general trans-
lation which is not the identity, leaves no point unchanged by the translations.
A second sub-group is formed by the set of all transformations which leave the
same 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.

The possible symmetry elements of the point group are:-

2 π
Rotations through integer multiples of n around an axis. The n-fold rota-
tions are denoted as Cn .

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

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

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

23
for odd n, (Sn )n = σ.

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 point is invariant under all the transformations of the point
group, the rotation axes and mirror planes must all intersect at these points.

3.4.4 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 = (32)
n 2
Thus, cos 2nπ must be integer or half odd integer which is in the set {±1, 0, ± 21 }.
This restriction limits the possible n-fold rotation axis to be of order n =
1 , 2 , 3 , 4 , 6 and allows no others. Thus, a crystalline lattice can only
contain a two, three-fold, four-fold or six-fold axis of rotation. However, 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 invariance.

3.4.5 Point Group Nomenclature


The point groups are referred to by using two different notation schemes, the
Schoenflies and the International notation. In the following examples, first the
groups are labelled by their Schoenflies designation and then their International
designation is given.

The point groups are:

24
Cn The groups Cn only contain an n-fold rotation axis. The group contains
as many elements as the order of the axis. The international symbol is n.

Cn,v The groups Cn,v contain the n-fold rotation axis and have vertical mir-
ror planes which contains the axis of rotation. The effect of the n-fold axis, if n
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 equivalent
mirror planes. The other n2 rotations merely bring the mirror plane into coinci-
dence 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 equivalent to a
mirror reflection in the bisecting (vertical) mirror plane. The effect of Cn is to
transform this bisecting mirror planes into a set of n2 equivalent bisecting 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 n2 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 vertical the 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 contain a
set of bisecting mirror planes.

Cn,h The groups Cn,h contain the n-fold rotation, and have a horizontal mir-
ror plane which is perpendicular to the axis of rotation. These groups contain
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 these 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 reflections which are, in
general, designated by n.

Sn The groups Sn only contain the n-fold rotation - reflection axis. For even
n, the group contains only n elements as (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

25
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
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 coordinate of point P 0 is − z. Construct the line joining P and P 0 . The
mid-point 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 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 contains all the elements of Dn and also contain a hor-
izontal mirror plane perpendicular to the n-fold axis. The effect of a rotation
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 the two
sets of two-fold axes also have their 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 contains all the elements of Dn and mirror planes
which contain the n-fold axis and bisect the two-fold axes. The effect 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 remaining n elements are rotation reflections about the principle
2k+1
axis of the form 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.

26
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 the center of the opposite faces. The compound
action of two of the three-fold rotations yields a rotation about a two-fold axis.
There are three such two-fold axes passing through the midpoints of opposite
edges of the tetrahedron. The tetrahedral group has 12 elements. The symme-
try operations can also be found in a cube, if the three four-fold rotation axes
present in the the cube are discarded. The group has the International symbol
of 23.

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


flection plane passing through one edge and the mid point of the opposite edge
of the tetrahedron. The reflection planes bisects a pair of two-fold axes of T .
There are six mirror planes σd . For the cube, these mirror planes are the diago-
nal 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 24 elements. The group Td has the International designation as 43m

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


which bisects the angle between the three-fold axes. For the tetrahedron, this
group is equivalent to Td , but for the cube, the mirror plane is parallel to op-
posite faces. There are three such horizontal planes. The planes bisect the
angles between the three-fold axes, and, therefore, converts them into six-fold
rotation reflection axes. Since the group contains S6 , it also contains I. Hence,
Th = Ti × CI . The group has 24 elements. The International designation for
2
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 24 elements. 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 equiva-
lent to a reflection in a diagonal mirror plane. The C3 axes becomes S6 axes,
just as in the case of Th . The group contains 48 elements. The International
4 2
designation is either m 3m or m3m.

27
3.5 Bravais Lattices
There are an infinite number of choices for the primitive lattice vectors, however,
only a few special lattices are invariant under point group operations. These
are the Bravais lattices. In three dimensions there are 14 Bravais lattices types.
The 14 Bravais lattices are organized according to seven crystal systems. The
Bravais lattices can be categorized in terms of the number of symmetry opera-
tions.

The 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 ).

——————————————————————————————————

3.5.1 Exercise 1
Show that the volume of a primitive unit cell, Vc is given by
  21
Vc = a b c 1 + 2 cos α cos β cos γ − cos2 α − cos2 β − cos2 γ (33)

——————————————————————————————————

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.

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 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 corners
of the cube. A rotation of any multiple of 23π will bring the cube into coinci-
dence 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
centers of opposite edges. The highest symmetry group when mirror symmetry

28
is not included is the octahedral group O. The octahedral group O contains 24
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 48 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
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 vertices
of the cube and also one at the central point which is 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 vectors from any lattice point joining
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
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

29
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 consecutive
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
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.

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 symme-
try tetragonal point group which is D4h or 4/mmm with 16 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

30
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
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. Namely, the c-axis
height is a 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 to 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) only has 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 4 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 base centered orthorhombic
lattice (C).

The base centered orthorhombic lattice (C) can be constructed from the
tetragonal lattice in the following manner. 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

31
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 = γ = π2 6= β. The monoclinic systems is 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.

There are two types of monoclinic Bravais lattices: the simple monoclinic
(P) and the body centered monoclinic Bravais lattice (I). The two monoclinic
Bravais lattices correspond to the two tetragonal Bravais lattices. The four or-
thorhombic 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

32
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 system. 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
later 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 stretch-
ing 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
Bravais lattice. The point symmetry group is D3h or 62m with 12 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
for the b.c.c. structure and 60 degrees for the f.c.c. structure.

The trigonal unit cell contains two equilateral triangles. In the trigonal lat-
tice the equilateral triangles form 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 24 sym-
metry elements.

33
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 from 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 (34)

34
In summary the following structures were found:

π
Cubic. 3 a = b = c α = β = γ = 2

π
Tetragonal. 2 a = b 6= c α = β = γ = 2

π
Orthorhombic. 4 a 6= b 6= c α = β = γ = 2

π
Monoclinic. 2 a 6= b 6= c α = β = 2 6= γ

Triclinic. 1 a 6= b 6= c α 6= β 6= γ

2 π π
Trigonal. 1 a = b = c α = β = γ < 3 6= 2

π 2 π
Hexagonal. 1 a = b 6= c α = β = 2 , γ = 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 2
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
(b.c.c.) and simple cubic (s.c.) lattices, for n = 1, n = 2 and n = 3.

——————————————————————————————————

35
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.

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
Bravais Lattice types and by the Schoenflies designation followed by the appro-
priate (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 (4mmm), 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 which only consists of the identity and
the inversion operation.

The triclinic system only contains C1 (1) and Cs (m).

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).

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

This completes the enumeration of the 32 point groups.

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

36
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 basis with the various point group symmetries on the various Bra-
vais Lattices. For example, the 5 cubic point groups can be placed on the three
cubic Bravais Lattices, yielding 15 cubic space groups. Likewise, 7 tetragonal
groups can be placed on the two tetragonal Bravais Lattices, yielding 14 tetrag-
onal 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 Bravais lattice. These new
symmetry elements include :-

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


Bravais lattice, 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 trans-
lations by lattice vectors which produce one complete rotation by 2 π. Thus, n
screw operations, each producing a rotation of 2nπ , produce a translation of m
lattice spacings.

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 a, b, c (according to whether
the translation is along the a, b and c axis), or n and d (the diagonal or diamond
glide which are special cases involving translations along more than one axis).

The hexagonal close-packed lattice structure has both of these types of non-
symmorphic symmetry operations. The hexagonal close-packed structure 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 dis-
tance 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 possible 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 between two
neighboring B lattice sites and is equidistant to the two A lattice sites that form

37
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 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 lat-
tice along the c axis. A line can be constructed which connects any two of the
three B sites inside the hexagonal unit cell. Form a parallel line connecting 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 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 operation is a translation by
c
2 along the c axis followed by a reflection in the plane.

There are two different systems of nomenclature for space groups, one 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 essen-
tial symmetry elements that are present. Thus, for example, the space group
P 63 /mmc has a primitive (P ) hexagonal Bravais lattice with point group sym-
metry 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).

38
3.8 Crystal Structures with Bases.
Crystal structures are specified by giving the basis and the 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 polyhedra around


each inequivalent site in the lattice. This provides information about the local
environment of the atom which is important for bonding. Small deformations
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. There-
fore, 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. Jensen, (The Structures of Binary Compounds, North
Holland publishers (1988)) and by Villars and Daams (Journal of Alloys and
Compounds, 197, 177 (1993)).

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
lattice is the face centered cubic lattice, and has a two atom basis. In the dia-
mond structure, both atoms are identical. They are located at the sites of the
Bravais lattice (0, 0, 0) and at a second site displaced 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 structure consists of
two interpenetrating face centered cubic lattices with atoms on each lattice site.
Diamond possesses a center of inversion located half way between the origins
of the two 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 four neighbor-
ing atoms form a tetrahedron centered on each atom. The tetrahedra centered
on the two inequivalent lattice sites have different orientations. 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 structure. 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.

——————————————————————————————————

39
3.8.2 Exercise 3
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 (35)

where a is the length of the side of the hexagon. The atoms are located at [0, 0, z]
and [0, 0, 12 + z] where z ≈ 0, and the coordinates are given in terms of the
primitive lattice vectors. Another two atoms are located at the positions [ 23 , 23 , z]
and [ 13 , 13 , 12 + z], where z ≈ 0. The structure is formed in layers, in which
each atom is bonded to three other atoms, thereby forming a two-dimensional
hexagonal network. The central site of the two-dimensional hexagonal ring is
open. The stacking sequence of the layers just corresponds to a translation of
one layer by [ 31 , 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 lay-
ers. 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
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.

——————————————————————————————————

40
3.8.4 Exercise 4
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
based on an f.c.c. form with 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?

——————————————————————————————————

3.8.5 Hexagonal Close-Packed Structure


The hexagonal close-packed structure has hexagonal symmetry, the space group
is P 63 /mmc. It is composed of the hexagonal Bravais lattice, and has a basis
composed of two atoms. The two identical atoms are positioned at [0, 0, 0]
which is at the vertex of the primitive lattice cell, and has the other atom
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 (36)

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 (37)
2 2 3 2

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 12 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 of radius and forming a close-packed hexagonal
layer. The second layer is formed by stacking a second hexagonal layer of atoms
above the first. However, the center of the second layer of atoms are positioned
above the dimples in the first layer. There are two sets dimples of dimples
between the atoms, so there are two different choices for placing the second

41
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 (38)
3 1 3 2 2 3
or [ 13 , 13 , 12 ].

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 12 nearest neighbor atoms.

On assuming a radius of the atomic



spheres to be r, 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 (39)
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 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.

——————————————————————————————————

3.8.6 Exercise 5
c
Show that the a ratio for an ideal hexagonal close-packed lattice structure is
  12
c 8
= (40)
a 3

——————————————————————————————————

42
3.8.7 Exercise 6
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 have
triangular arrays of atoms, where the sides of the triangle side has length √a2 .
The normal to the planes are in the direction [1, 1, 1] i.e.
 
1
n̂ = √ êx + êy + êz (41)
3
where ê are the orthogonal unit vectors of the conventional cell. The equations
of the planes are  
r − m a êx . n̂ = 0 (42)

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 through
a
s = m √ (43)
3
for integer m.

43
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 (44)
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 (45)
2
which corresponds to the face diagonal of the conventional unit cell that lies in
the triangular plane and
 
1
ê2 = √ êx + êy − 2 êz (46)
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 (47)
3 2
This can be re-written as √
2 3 a
√ ê2 (48)
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
4
3 and 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.

44
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 cu-
bic 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
neighbors 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 with vertices at (0, 0, 0) and
the Cl− ions are located on a face centered cubic lattice with vertices 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 ) (49)

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− ) (50)
2
Combining the above two equations yields an inequality for the ratio of the ionic
radii of the ions
r(Cl− ) √
+
≤ 1 + 2 (51)
r(N a )
If this inequality is not obeyed, the Pauli forces render the structure unstable.

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 next 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

45
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
3
of opposite charge located at a distance 2 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)
≤ (52)
r(Cs+ ) 2

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 site of the
conventional f.c.c. unit cell is surrounded by an octahedron and, therefore, has
a coordination number of 6. There are also tetrahedral holes with coordination
number 4. 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. The tetrahedral holes are best 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 octants
corners. The are 12 trigonal holes which are located near the 8 vertices of the
conventional unit cell. The trigonal sites lie in the plane 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 = (53)
r(R+ )
For the tetragonal sites, by considering the body diagonal of the octant, one
expects that   √
− + 3
2 r(X ) + r(R ) = a (54)
2
and by considering the face diagonal
a
2 r(X − ) < √ (55)
2

46
Hence, we find the tetragonal hole has the limiting radius ratio of
r
r(X − )

3
> 2 + 1 (56)
r(R+ ) 2
In particular, the radius ratio rules suggest that the range of radii ratios where
the various configurations are stable are given by

6.45 > rr > 4.45 trigonal 3


4.45 > rr > 2.41 tetrahedral 4 (57)
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 is 8 hole and the hole size is larger
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)
≤ (58)
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 form 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 ). 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 revered. The O anions are in the f.c.c. positions and the
Li cations form a simple cubic array.

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 3 Cu atoms and one
Au per unit cell. All the atoms are located on the sites of a face centered cubic

47
unit cell. The Au atom can be envisaged as being positioned on the corners of
the cube, whereas the 3 Cu atoms sit on the centers of the faces of the cube,
forming octahedra. Thus, the basis of the structure consists of the position of
the Au atom
r0 = 0 (59)
and the three Cu atoms are located at
a
r1 = ( êy + êz )
2
a
r2 = ( êx + êz )
2
a
r3 = ( êx + êy ) (60)
2
The Au atoms have 12 Cu nearest neighbors located at a distance √a , whereas
2
the Cu atoms only have 4 Au nearest neighbors.

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 numer-
ous 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 lie on lines which are oriented along one set of face diagonals
of the base. The atoms are also located on horizontal lines through the body
centers, and are orthogonal to the lines in the base. The titanium ion is sur-
rounded by six O atoms which form a slightly distorted octahedron.

3.8.14 Zinc Blende Structure


Zinc Blende structure or ZnS 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 the face centers of the cube. The S 2− are positioned on 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 related to the diamond structure,
except that Zinc Blende involves two different types of atoms. 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 ratio rules
suggest that this structure will be adopted whenever

r 
3
2 + 1 > rr > 1 + 2 (61)
2

48
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.

3.8.15 Zincite Structure


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

a1 = a êx
a √
a2 = ( êx + 3 êy )
2
a3 = z êz (62)

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
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.

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 con-
tain only Zn or only S ions. The set of planes form layers consisting of a pair
of planes. In 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 tetrahe-
dral bonds join the atoms in the successive layers. Due to the orientation of the
inter-layer tetrahedral bonds, successive pairs of planes are displaced horizon-
tally. 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
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.

49
3.8.16 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, 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 lies on the face centers. The T iO2 form a set of parallel planes sepa-
rated by planes of BaO. Each T i atom is surrounded by an octahedron of O
atoms, which have corners which are shared with the octahedron surrounding
the neighboring T i atoms.

——————————————————————————————————

3.8.17 Exercise 7
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.

——————————————————————————————————

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 axis given by
the primitive lattice vectors a1 , a2 and a3 . The planes need not intersect any

50
particular axes at a lattice point, however, every lattice point on the three axes
will have one member of the family pass through it. In particular, one plane
must pass through the origin O.

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 intercepts x1 , x2 and x3 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 between one lattice plane with the three
primitive 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 mem-
bers 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 (63)
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 lay in the plane, the three vectors
1 1
a − a
h1 1 h2 2
1 1
a2 − a
h2 h3 3
1 1
a − a (64)
h3 3 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
   
2 1 1 1 1
n̂ ∝ κ a − a ∧ a − a
h1 1 h2 2 h2 2 h3 3

51
κ2
 
= h3 a1 ∧ a2 + h2 a3 ∧ a1 + h1 a2 ∧ a3
h1 h2 h3
(65)

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
same directions as the primitive ”reciprocal lattice vectors”.

The primitive reciprocal lattice vectors are defined by


a2 ∧ a3
b1 = 2 π (66)
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 (67)

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 b 3
 2

= (68)
dh
This is seen through the following consideration: 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 (69)

The minimum distance, s, between the origin and the plane is given by
Bh
s = r.
| Bh |
 
dh
= r . Bh

= κ dh (70)

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

52
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 (71)

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 8
Consider the planes (1, 0, 0) and (0, 0, 1) for a 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 9
Show that 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 between the three primitive lattice vectors of the reciprocal lattice,
bi , β1 ( 6 b2 , b3 ), β2 ( 6 b3 , b1 ) and β3 ( 6 b1 , b2 ) via

cos β2 cos β3 − cos β1


cos α1 = (72)
| sin β2 sin β3 |

and also find the inverse relation.

——————————————————————————————————

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. 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 restriction on the symmetry of the lattice

53
but puts a restriction on the vectors that describe the ”unit cells”. For example,
an Al − M n quasi-crystal (Schechtman, Blech, Gratais and Cahn, Phys. Rev.
Lett. 53, 1951 (1984)) has icosahedral symmetry, with two, three and five-fold
axes. The structure is made from blocks consisting of a central M n atom sur-
rounded by 12 Al atoms arranged at the corners of an icosahedron. This type
of icosahedral structure often the arrangement of 13 atoms which has the lowest
energy (F.C. Frank, Proc. Roy. Soc. London, 215, 43 (1952)). The icosahedra
are stacked together with the same orientation. The voids are formed with 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 re-
striction 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
Bragg reflections (Levine and Steinhart, Phys. Rev. Lett. 53, 2477 (1984)).

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


vais lattice structure in higher dimensions (six or more) onto three dimensions
(P. Kramer and R. Neri, Acta. Crystallogr. Sec. A 40, 580 (1984)). To il-
lustrate this, consider a square two-dimensional lattice, with lattice constant 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 (73)

and the upper line is determined by

y = a + ( x + a ) tan θ (74)

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 satisfy
the inequality
1 + ( n + 1 ) tan θ > m > n tan θ (75)
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 (76)

where X
m = m0 Θ(1 + (n + 1) tan θ − m0 ) Θ(m0 − n tan θ) (77)
m0

For irrational values of the slope, the resulting array of points is a quasi-periodic
array. The spacing between consecutive points of the quasi-periodic array is ei-
ther given by cos θ or sin θ. The spacings are not distributed periodically, but

54
nevertheless, are distributed according to some√irregular or more complex pat-
tern. If the slopes of the line are equal to 12 ( 5 − 1 ) the array of projected
points is 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 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
1
2 ( 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 adjoined
so that the markings match (Gardner, Scientific American, 236, 110 (1977)).
The result is a tiling without long-ranged 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 orien-
tational order, as can be seen by decorating each tile with lines. The lines on
the tiles join up to form five sets 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 set form a Fibonacci series.

55
4 Structure Determination
Structure can be determined by experiments in which beams of particles are
scattered from the structure. Elastic scattering experiments are usually pre-
ferred as the underlying lattices are not dynamically deformed by the process.
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 should have
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 conditions


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 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 (78)
2

56
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 (79)
λ 2
and constructive interference occurs whenever
θ
n λ = 2 d sin (80)
2
This is Bragg’s law. 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,
but one must also consider scattering from the different families of planes in the
solid. Different families of planes of atoms in a solid have different orientations.
Since a plane of every family passes through each lattice point, the different
orientations may have different spacings between members of the families of
planes, so 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 condition 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 wave vector of the incident beam is expressed as k, where



k = ê (81)
λ
and the reflected wave has wave vector k 0
2π 0
k0 = ê (82)
λ
where ê and ê0 are two unit vectors. Let us consider two scattering centers
separated by a vector displacement d. Then, the difference in optical path length

57
for x-rays scattered from one atom is composed of two non-equal segments
θ
d cos = d . ê
2
θ0
d cos = d . ê0 (83)
2
The optical path difference between the two waves is given by the difference
θ θ0
d cos − d cos = d . ( ê − ê0 ) (84)
2 2
Thus, constructive interference of the scattered waves from two unit cells occurs
whenever  
d. ê − ê0 = mλ (85)

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 (86)

If this condition is fulfilled for the set of vectors d that are all the Bravais Lattice
vectors R, one finds the Laue condition for coherent scattering
 
0
R. k − k = 2πm (87)

or alternatively    
0
exp i k − k .R = 1 (88)

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
wave will be extremely intense.

If the scattering vector q is defined as

q = k − k0 (89)

the Laue condition is satisfied for the special set of q values, Q which satisfy
 
exp i Q . R = 1 ∀ R (90)

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

k0 = k − Q (91)

58
which can be squared to yield

k 02 = k 2 − 2 k . Q + Q2 (92)

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 (93)
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
(94)

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,
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 can be
used to construct the general Q.

4.1.3 Equivalence of the Bragg and Laue conditions


Since a plane belonging to each family of planes passes through each lattice
point, it is obvious 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. As k and k 0 have the same magnitude, they make the

59
θ
same angle 2 with the Bragg plane.

Due to the elastic scattering condition one has |Q| = 2 k sin θ2 and if the
scattering is coherent then the magnitude of Q can be written as |Q| = 2 πd n ,
where n is the order of the reflection and d is a distance characteristic of the
lattice. Combining the elastic and Laue conditions, one has
θ πn
k sin =
2 d
θ
2 d sin = nλ (95)
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 of a family of planes.

4.1.4 The Ewald Construction


Since the Laue condition is very restrictive, the vectors k which produce coher-
ent scattering are relatively few and far between. The Ewald construction (P.P.
Ewald, Z. Krist. 56, 129 (1921)) provides a convenient way of visualizing how
the Laue condition may be fulfilled.

The incident wave vector wave 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
represented by vectors k 0 directed from points on the sphere’s surface directed
to the center of the sphere. The scattering wave vectors q = k − k 0 are
directed from the origin towards the points on the surface of the sphere.

Since the wave vectors Q which are solutions of

Q.R = 2πm (96)

form a lattice of points (the reciprocal lattice) including Q = 0, a lattice


point has to be centered on the origin. This lattice is indexed by three integers
(m1 , m2 , m3 ) corresponding to the components along three primitive (recipro-
cal) lattice vectors. When a second point of the lattice of Q points resides on the
surface of the Ewald sphere, say at − k 0 , it produces a Bragg reflected beam. In
this case, the Laue condition is satisfied and the incident beam will be Bragg re-
flected 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

60
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
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 Friedel.

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
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 (97)

Furthermore, since k and a1 are perpendicular

k 0 . a1 = − Q . a1 (98)

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

61
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 10
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 the paper they indicate an f.c.c. structure for the solid.
Indicate the conventional axes on their unit cell.

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


x-rays of wavelength λ = 0.9 A, for the dopings 3, 4 and 6 in the paper, what
are the angles 2 θ for the first 5 diffraction peaks for the observed structures?

——————————————————————————————————

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 den-
sity of each atom can be expressed in terms of the form factor. The form factors
for an atom in a solid differ only slightly from the form factors of isolated atoms,
and are mainly determined by the atomic charge number Z. Although there

62
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) (99)
j

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

| S(Q) |2 (100)

The Q dependence of the intensity can be used to determine the basis of the
crystal. Unfortunately, since only the modulus of S(Q) can be found from ex-
periment and not its phase, 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 there is another atom of the same type at − rj
and S(Q) is purely real. The phase problem just simplifies to the question as
to whether S(Q) is positive or negative.

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 (101)
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 (102)

For each Q that satisfies the Laue condition, there is a vector −Q which cor-
responds to the negative integer − m. The structure factor S(−Q) is just the

63
complex conjugate of S(Q)
S(−Q) = A − i B (103)
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 Lattice.

The face centered cubic lattice can be represented in terms of a simple cubic
lattice with a four atom basis. The scattering from this lattice 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 (conven-
tional) unit cell of the face centered cubic lattice consists of the atomic positions
r1 = 0
 
a
r2 = êx + êy
2
 
a
r3 = êz + êx
2
 
a
r4 = êz + êy (104)
2
The Bragg vectors for the conventional simple cubic cell are easily found to be
 

bx = êx
a
 

by = êy
a
 

bz = êz (105)
a
so a general simple cubic Bragg scattering vector is given by
  

Q = m1 êx + m2 êy + m3 êz (106)
a
The geometric structure factor for the conventional f.c.c. unit cell is found to
be
X  
SG (Q) = exp i Q . rj
j
"  
= 1 + exp + i π ( m1 + m2 ) +

64
   #
+ exp + i π ( m1 + m3 ) + exp + i π ( m2 + m3 )

(107)

When evaluated at the Bragg vectors, the geometric structure factor adds co-
herently
SG (Q) = 4 (108)
if the integers (m1 , m2 , m3 ) are either all even or are all odd. The geometric
structure factor interferes destructively

SG (Q) = 0 (109)

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.

The Body Centered Cubic Lattice.

The body centered cubic lattice can be viewed as a simple cubic lattice with
a two atom basis

r0 = 0
 
a
r1 = êx + êy + êz (110)
2
Then, the geometric structure factor for the conventional b.c.c. unit cell is just
 
a
SG (Q) = 1 + exp i Q . ( êx + êy + êz )
2
 
a
= 1 + exp i ( Qx + Qy + Qz ) (111)
2
Now the Bragg vectors for the simple cubic structure are just

Q = ( m1 êx + m2 êy + m3 êz ) (112)
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

65
= 2 f or ( m1 + m2 + m3 ) even

= 0 f or ( m1 + m2 + m3 ) odd
(113)

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 Lattice.

The diamond lattice is an f.c.c. lattice with a two atom basis

r0 = 0
 
a
r1 = êx + êy + êz (114)
4
where the conventional f.c.c. unit cell has linear dimension a.

From the discussion of scattering from an f.c.c. lattice, one finds that the
Q vectors of the Bragg spots can be expressed in terms of the set of primitive
vectors for the b.c.c. lattice
X
Q = mi bi (115)
i

The primitive vectors are given by


 

b1 = êy + êz − êx
a
 

b2 = êz + êx − êy
a
 

b3 = êx + êy − êz (116)
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 ) (117)
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 f or ( m1 + m2 + m3 ) 2 × even

66
SG (Q) = 0 f or ( m1 + m2 + m3 ) 2 × odd

SG (Q) = 1 ± i f or ( m1 + m2 + m3 ) odd
(118)

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
(119)

The rule for the modulation of intensities is expressed directly in terms of the
quantity
X Qi a 1
= ( m1 + m2 + m3 ) (120)
i
4 π 2
Thus, one can describe the system of Bragg spots as residing on a b.c.c. lattice
with cubic cell of side 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 cubic lattices of side 4aπ .
The length 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 are
vanishing. Thus, the non-zero intensities on this simple 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

67
simple cubic lattice all have the same intensities.

Extinctions due to Glide Planes and Screw Axes.

Consider a solid with a glide plane, along the êz axis perpendicular to 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
(121)
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
  
m3
= exp 2 π i ( x m1 + z m3 ) 1 + (−1)

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

(122)

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
structure factor for m1 = 0 and m3 = 0 is made up of contributions with 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

(123)

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

68
——————————————————————————————————

4.1.8 Exercise 11
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.
Fleming et. al. Nature 352, 701 (1991). Calculate the structure factor for these
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 12
The Hendriks-Teller model for x-ray diffraction from a disordered system con-
siders 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 average geometric structure factor for this model, and
show that
 
p ( 1 − p ) 1 − cos Q da
SG (Q) =
1 − p(1 − p) − p cos Qa − (1 − p) cos[ Q(a + da) ] + p(1 − p) cos Qda
(124)
In a scattering measurement on a random system, one measures the average of
| SG (Q) |2 . Determine the relation between SG (Q) and | SG (Q) |2 and describe
the results of a scattering measurement on this one-dimensional system.

——————————————————————————————————

Polyatomic Crystals.

For a polyatomic crystal the structure factor has both the geometric contri-
bution and the contribution from the atomic form factors of the basis atoms
X  
S(Q) = exp i Q . rj Fj (Q) (125)
j

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

69
The atomic form factor is normalized to the electronic charge of the atom.
For a single atom, the form factor is given by
Z  
3
F (Q) = d r ρ(r) exp − i Q . r (126)

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  
2
F (Q) = 2 π dr r 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 Q r
= 4π dr r2 ρ(r) (127)
0 Qr
For forward scattering, Q = 0, the form factor reduces to
Z ∞
F (0) = 4 π dr r2 ρ(r)
0
= Z (128)
where Z is the atomic number. Typically F (Q) decreases monotonically with
increasing Q, falling off as a power of Q12 for large Q.

——————————————————————————————————

4.1.10 Exercise 13
Calculate the x-ray scattering intensities for the following close-packed structure
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 14
Find the atomic form factor for the hydrogen atom, using the electron density
 
1 2r
ρ(r) = exp − (129)
π a3 a

70
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), ( 21 , 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 )

(130)

As 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

(131)

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 15
Potassium Chloride has the same structure as N aCl. However, K + and Cl− are
iso-electronic and so have very similar structure factors. Determine the indices
(m1 , m2 , m3 ) of the allowed Bragg reflections.

71
——————————————————————————————————

4.1.13 Exercise 16
Calculate the structure factor for the zincblende structure. The zincblende
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 one
atom on each site of an f.c.c. lattice. However, there is a transition between 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 statistically averaged value
3 1
Fav (Q) = FCu (Q) + FAu (Q) (132)
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 )

(133)

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 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 )

(134)

72
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)
∼ (135)
I(2, 0, 0) FAu (0) + 3 FCu (0)
which leads to a relative intensity of about 0.09. 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 is easily observable by neutron diffraction.

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 [ 21 , 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. At temperatures above the
order-disorder transition temperature, 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 (136)

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 (137)

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, Tc ≈ 741 K, the order parameter has a
non-zero magnitude 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 (138)

73
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 17
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
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 the change in the
momentum of the neutron from the initial value h̄ k to the final value h̄ k 0 of

q = k − k0 (139)

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 ) (140)
mn

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


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

= | b |2 (141)
dΩ
Hence, the total cross-section is given by

σ = 4 π | b |2 (142)

74
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 ) b∗i bj (143)
dΩ i,j

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

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
b∗i bj = b∗i bj = | b |2 (144)
while, if i = j, one has the average of the squared amplitude

b∗i bi = | b |2 (145)

In general, the average has the form


 
b∗i bj = | b |2 + δi,j | b |2 − | b |2 (146)

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 (147)
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 (148)

for all lattice vectors R and m is any integer. When this condition is satisfied,
the scattering produces Bragg reflections similar to those observed in x-ray scat-
tering. When the Bragg scattering condition is satisfied the coherent scattering
has an intensity proportional to N 2 .

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


and is given by  
dσ 2 2
= N |b| − |b| (149)
dΩ

75
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.

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 (150)
dΩ
Consider a beam of particles collimated to have a momentum k that falls
incident on a crystal. The particles are assumed to interact with either the
electrons or nuclei of the solid. An example is x-ray diffraction, in which the
beam of photons interacts elastically with the electron density, or alternatively
in neutron diffraction experiments 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 ) (151)
j

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(rj , t) . A(rj , t)
j
2mc 2 m c2
(152)
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

76
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. Thus, in this P
approximation,
the x-rays couple to the density of the charged particles, ρ(r) = j δ( r − r j ).
For electrons, the coupling constant is proportional to the length

e2 e2 h̄
2
= ∼ 10−15 m (153)
2 me c 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


The incident beam has the asymptotic form of a momentum eigenstate with
eigenvalue h̄ k
  12    
1
Ψk (r, t) = exp + ik.r exp − i ωk t (154)
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    
0 1 X 3 0
< k | Ĥint | k > = d r exp − i k . 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

(155)

where R0 = r − rj . The integration over R0 yields the Fourier transform of


the interaction potential between the scattered particle and the j-th atom
Z  
Vj (q) = d3 R0 exp − i q . R0 Vj (R0 )
V
Z  
≈ d3 R0 exp − i q . R0 Vj (R0 ) (156)

77
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) > (157)
∂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) > (158)

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 (159)
∂t
where the new interaction operator is time dependent and is given by
   
ˆ i i
H̃ int (t) = exp + Ĥ0 t Ĥint exp − Ĥ0 t (160)
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 (161)
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 (162)
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

78
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) > (163)
However, it is more convenient to calculate the probability based on the matrix
elements evaluated in the interaction representation
< k 0 | Ψ̃k (t) > (164)
These two quantities are equivalent, as they are simply related via
 
0 i 0
< k | Ψk (t) > = exp − E(k ) t < k 0 | Ψ̃k (t) > (165)

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
(166)
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
(167)
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
(168)
The transition rate is evaluated as
 
2 η t
 exp h̄


P (k → k 0 , t) = | < k 0 | Ĥint | k > |2
∂t ( E(k 0 ) − E(k) )2 + η 2
 
2 2ηt η
= | < k 0 | Ĥint | k > |2 exp
h̄ h̄ ( E(k 0 ) − E(k) )2 + η 2
(169)

79
Then, in the limit η → 0, the transition rate becomes time independent and
energy dependent terms reduce to π times an energy conserving delta function
since
η
lim = π δ( E(k 0 ) − E(k) ) (170)
η → 0 ( E(k 0 ) − 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
(171)

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 dis-
tribution 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 dΩ.

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 ) (172)

where ρdω (E, k 0 ) is the density of final scattering states per unit energy range,
defined as
Z ∞
0 V
ρdω (E, k ) = dΩ dk 0 k 02 δ( E − E(k 0 ) ) (173)
( 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 direction dΩ.

4.3.3 The Elastic Scattering Cross-Section


The scattering cross-section is defined by
 

dΩ = P (k → dΩ) / F (174)
dΩ

80
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 h̄mk .
Thus, for particles of mass mn , the flux is given by
h̄ k
F = (175)
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 ∞
0 V dk 0 k 02
ρdΩ (E, k ) = dE 0
dΩ δ( E − E(k 0 ) )
( 2 π )3 0 dE 0
V dk 0 k 02
= dΩ (176)
( 2 π )3 dE 0

where the magnitude of k 0 is determined by the 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 (177)
mn
and so, the density of final states can be written as
V mn k 0
ρdΩ (E, k 0 ) = dΩ (178)
(2π)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 ) (179)

the final density of states ρdΩ (E, k) and the flux F into the expression eqn(174)
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
(180)

where q is the scattering vector

k − k0 = q (181)

The magnitude of the scattering vector is related to the scattering angle θ via
θ
q = 2 k sin (182)
2

81
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 (183)
mn
Substituting for Vj (q), in the above expression for the cross-section, yields the
formulae for the elastic neutron scattering cross-section
  
dσ X

= bj bj 0 exp − i q . Rj − Rj 0 (184)
dΩ 0
j,j

previously discussed.

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


c
F = (185)
V
if the incident vector potential is
r  

A(r, t) = êα c exp i ( k . r − ω t ) + c.c. (186)
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 ) = 3
dΩ (187)
( 2 π ) 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 Ak0 (r) . ρ̂(r) Ak (r)
dΩ h̄ c ( 2 π c ) 2 m e c
V
2  2
2 X   
e ∗
= êα . êβ S(q) S (q) exp − i q . R j − R j 0
4 π me c2
j,j 0
(188)

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

2
re = (189)
4 π me c2

82
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 as the basis
consists of one atom, the solid contains a total of N = N1 N2 N3 atoms.

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
(190)
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
 
X1 −1
n1 =N   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
(191)

This function exhibits the effect of the constructive and destructive interference
between the scattered waves emanating from the various atoms forming the

83
solid. The numerator of the function falls to zero at
2mπ
q . a1 = (192)
N1
for general integer values of m. The numerator has maximum magnitude at

(2m + 1)π
q . a1 = (193)
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

= re2 | F (q) |2 ×
dΩ
!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
(194)

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
occur when the three conditions

q . a1 = 2 π m1
q . a2 = 2 π m2
q . a3 = 2 π m3
(195)

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

∼ re2 | F (Q) |2 N 2 (196)
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
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

84
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 (197)

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 18
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 19
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 occurs for large Q values.

——————————————————————————————————

4.3.7 Exercise 20
Evaluate the effect of a significant number of thermally induced vacancies (miss-
ing atoms) in the elastic scattering cross-section from a crystal.

——————————————————————————————————

85
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 boundary. 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) (198)

where  
A2 (q) = exp i q . δR A1 (q) (199)

δR is the vector displacements of the origins of the two domains. The scattering
amplitude A1 (q) is given by
! N q a
! !
sin N1 2qx a sin 2 2 y sin N3 2qz a
A1 (q) ∝ q a
sin qx2 a sin y2 sin qz2 a
(200)

For a domain wall in the y − z plane, the displacement between the two Cu
sub-lattices is given by
1 a a
δR = ( N1 + ) a êx + êy + êz (201)
2 2 2
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 (202)

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 (203)

86
Thus, if m1 + m2 + m3 is even the intensity is modulated by the factor
N 1 qx a
4 cos2 (204)
2
whereas, if m1 + m2 + m3 is odd the intensity is modulated by the factor
N1 qx a
4 sin2 (205)
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 ∼ (206)
N1 a
the scattered intensity is finite and large. That is, the single anti-domain phase
boundary between identically 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 21
Consider the scattering produced by a CuCl type material, with anti-domain
walls. For simplicity, only consider the component of the scattering amplitude
associated with a single primitive lattice vector. Let p be the probability of not
crossing a domain wall on traversing one step a along a primitive lattice vector,
and q is the probability 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
m1 =1

2 N 1 p q + ( p2 − q 2 ) ( p2 − q 2 ) q 2
= 2 qx a −
2 2 2
2 ( q + ( p − q ) sin 2 ) ( q + ( p2 − q 2 ) sin2 qx2a )2
2
 
+ O ( p − q )N1 +1 (207)

87
Hence, show that the intensity of the ”super-lattice” Bragg peaks are dimin-
ished 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 12
vertices of the icosahedron, which is responsible for the five-fold symmetry.

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

sn = n a cos θ + m0 a sin θ (208)

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
(209)

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
2
A(q) = d 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)
(210)

88
which is to be evaluated on the one-dimensional line

q = q ( cos θ , sin θ ) (211)

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) (212)

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


X
B(r) = δ(x − na) δ(y − ma) (213)
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) (214)

This can be evaluated using the convolution theorem as the convolution of the
product of Fourier Transforms
d2 q 0
Z
A(q) = B(q − q 0 ) C(q 0 ) (215)
( 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 ) (216)
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 qy0
tan θ ) x
i qy0
!
0
exp[ i q y a (1 + tan θ) ] − 1
= ( 2 π ) δ( qx0 + qy0 tan θ )
i qy0
(217)

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) (218)
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

89
2π X
= δ( qx − Qx + ( qy − Qy ) tan θ ) ×
a2
Q
!
exp[ i ( qy − Qy ) a (1 + tan θ) ] − 1
×
i ( qy − Qy )
(219)
Evaluating this on the line in q space yields the amplitude for scattering from
the one-dimensional quasi-crystal
X qa
A(q) = 2 π δ( − Qx a − Qy a tan θ ) ×
cos θ
Q
!
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 )
(220)
This has delta function like peaks at the wave vectors given by
q a = 2 π ( m1 cos θ + m2 sin θ ) (221)
where m1 and m2 are integers. The intensities of the peaks are proportional to
 
sin2 π ( m1 sin θ − m2 cos θ ) ( cos θ + sin θ )
| A(q) |2 ∝ (222)
( 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 tan θ.

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

90
Z   X
= d3 r exp iq.r δ 3 (r − rj )
j
(223)
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
Z Z   X
3 0 0
= 3
d r d r exp iq.(r − r ) δ 3 (r − ri ) δ 3 (r0 − rj )
i,j
(224)
On considering the long time average of the atomic positions, one obtains
Z Z   X
I(q) ∝ d3 r d3 r0 exp i q . ( r − r0 ) δ 3 (r − ri ) δ 3 (r0 − rj )
i,j
Z Z  
= d3 r d3 r0 exp i q . ( r − r0 ) C(r, r0 )

(225)
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 ) (226)
Hence, the
Z Z Z  
I(q) ∝ d3 r ρ(0) + d3 r d3 r0 exp i q . ( r − r0 ) g(r − r0 )
Z  
= N + V d3 r exp i q . r g(r)

(227)
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
3 Z ∞
2 ( 2 π ) 3 4π 2
= N + N δ (q) + V dr r sin q r ( g(r) − ρ(0) )
V q 0
(228)

91
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.

92
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, the physical property only needs 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 (229)

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.

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 )
(230)

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) (231)
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.

——————————————————————————————————

93
5.0.1 Exercise 22
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) (232)
2π i

The quantity gi,j is given by the metric, since


1 X
aj . ak = gj,i b(i) . ak (233)
2π i

and since
b(i) . ak = 2 π δki (234)
one has
gj,k = aj . ak (235)
Hence, gj,k is the metric tensor. The metric tensor expresses the length s of a
vector r in terms of its components xi along the basis vectors ai . That is, if
X
r = xi ai (236)
i

then, for a constant metric, the length is given in terms of the components via
X
s2 = gi,j xi xj (237)
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  (238)
a1 a3 cos α2 a2 a3 cos α1 a23

The inverse transform is given by


X
b(i) = 2 π g i,k ak (239)
k

94
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 (240)
i k

and as X
aj = δjk ak (241)
k

one infers that X


δjk = gj,i g i,k (242)
i

Hence, the metric tensor is the inverse of the metric tensor for the dual vectors.

The volume of the unit cell, Vc , is given by

Vc2 = det ( gi,j ) (243)

or
 
Vc2 = a21 a22 a23 2 2 2
1 − cos α1 − cos α2 − cos α3 + 2 cos α1 cos α2 cos α3

(244)

The dual metric tensor is given by the inverse of the metric tensor, this is
evaluated 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
(245)

This dual metric is also defined as

bi . bj = ( 2 π )2 g i,j (246)

From this, one can immediately find that the length of the reciprocal lattice
vectors are given by
a2 a3
b1 = 2 π | sin α1 | (247)
Vc
etc., and the angle β3 between b(1) and b(2) is given by

( cos α1 cos α2 − cos α3 )


cos β3 = (248)
| sin α1 sin α2 |

95
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
(249)

These expressions are equivalent to the expression in terms of the vector product,
and they also satisfy the definitions of the primitive reciprocal lattice vectors

ai . b(j) = 2 π δij (250)

Any vector of the direct Bravais Lattice can be expressed as

R = n1 a1 + n2 a2 + n3 a3 (251)

A reciprocal lattice vector Q can also be written as

Q = m1 b(1) + m2 b(2) + m3 b(3) (252)

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) (253)

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 (254)

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 (255)

As the sum of the products of integers is still an integer ( say M ), the Laue
condition can be expressed as

Q.R = 2πM (256)

96
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 (257)

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.

——————————————————————————————————

5.1.1 Exercise 23
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 lattice vectors of the simple cubic direct
lattice are

a1 = a êx
a2 = a êy
a3 = a êz (258)

The reciprocal lattice vectors are determined to be



b(1) = êx
a

b(2) = êy
a

b(3) = êz (259)
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

97
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 (260)
2
a3
The volume of the unit cell is Vc = | a1 . ( a2 ∧ a3 ) | = 2 .

The reciprocal lattice vectors are determined to be


 
(1) 2π
b = êx + êy
a
 

b(2) = êy + êz
a
 

b(3) = êx + êz (261)
a

The three reciprocal lattice vectors span the three-dimensional reciprocal lattice,
but have different orientations from the direct lattice vectors. The reciprocal
lattice has cubic symmetry as can be seen by combining the three reciprocal
lattice vectors ( adding any two and subtracting the third ) to yield three or-
thogonal 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 (262)
2

98
The reciprocal lattice vectors are determined to be
 
(1) 2π
b = êx + êy − êz
a
 
(2) 2π
b = êx − êy + êz
a
 
(3) 2π
b = − êx + êy + êz (263)
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 reciprocal lattice vectors, which yields three orthogo-
nal 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 .

5.2.4 The Hexagonal Reciprocal Lattice


The hexagonal lattice has lattice vectors

 
a
a1 = 3 êx + êy
2

 
a
a2 = − 3 êx + êy
2
a3 = c êz (264)

The volume of the primitive unit cell is



3 2
Vc = a c (265)
2

The primitive reciprocal lattice vectors are


 
2π 1
b(1) = √ êx + êy
a 3
 
(2) 2 π 1
b = − √ êx + êy
a 3
2 π
b(3) = êz (266)
c

Thus, the reciprocal lattice of the hexagonal lattice is its own reciprocal lattice,
but is rotated about the z axis.

——————————————————————————————————

99
5.2.5 Exercise 24
A trigonal lattice is defined by three primitive lattice vectors a1 , a2 and a3 , all
of equal length a and an angle θ between any pair of these lattice vectors is
a constant. Show that the three vectors a1 = [m, n, p], a2 = [p, m, n] and
a3 = [n, p, m], referenced to an orthonormal basis represent a trigonal lattice.
Prove that the reciprocal lattice of a 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 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 . Secondly,
these connecting lines are bisected by planes. The equations for the set of these
planes are given by  
1
k − Q . Qi = 0 (267)
2 i
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(267) 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 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 lat-
tice 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. The sides of the cube are 2aπ and the Brillouin zone has
a volume of ( 2aπ )3 which, when given in terms of the volume of the unit cell of
3
the direct lattice, is equal to 8Vπc .

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

100
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-
cated at 2aπ ( 12 , 0, 0) is called X, and the mid-points of the edges at 2aπ ( 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 dodecahedral
rhombohedron. The cell is centered at the origin Γ = (0, 0, 0). The vertices are
located either on the 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).

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 the face centered cubic has twenty four vertices at
W = 2aπ (1, 12 , 0) and has six square faces which have centers on the Carte-
sian axes. The centers of the square faces are denoted by X and are located at
2 π
a (0, 1, 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
2 π 3 3
hexagonal faces are at a ( 4 , 4 , 0), and are denoted by K. The mid-points of
the edges between the square and hexagonal faces are denoted by U .

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.

101
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 and L = 2aπ ( √13 , 0, 2ac ).

Some of the interior high symmetry points are: Γ the zone center, Σ which
are located on the high symmetry lines Γ M , ∆ are the points on the lines Γ A.

102
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. The density functional method in principle
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 interac-
tions is that of disturbing the electron density around any excited electron. On
assuming 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 en-
ergy excited electrons merely involve the dressing of the single excited electron
thereby, forming a quasi-particle excitation. That is, the effect of the electron
excitation of the surrounding gas of electrons can be absorbed as renormaliza-
tions of the properties of the single-electron excitations. 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.

103
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
i=N
Xe h̄2 e2
 
1 X
Ĥ = − ∇2 + Vions (ri ) + (268)
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 (269)

where
i=N
Xe h̄2
 
Ĥ0 = − ∇2 + Vions (ri ) (270)
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 = (271)
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 )

for 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 )


(272)

104
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 (273)
Likewise, the permutation operators must also commute with any physical op-
erator Â
[ P̂i,j , Â ] = 0 (274)
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 = EΨ
P̂i,j Ψ = pi,j Ψ (275)

where pi,j are the eigenvalues of the permutation operators P̂i,j . As permutating
the same pair of particle indices twice always reproduces the initial wavefunc-
tion, one has
P̂i,j 2 = Iˆ (276)
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 (277)

Hence, the eigenvalues of the permutation operators must satisfy

pi,j 2 = 1 (278)

or
pi,j = ± 1 (279)
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 )


(280)

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

105
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 ) (281)

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 )
(282)

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
(283)
where we have assumed that the sets of single-electron eigenvalues have been
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
(284)
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 (285)
i=1

106
where the one-particle non-interacting Hamiltonian is given by

h̄2
Ĥi = − ∇2 + Vions (ri ) (286)
2m i
This one-particle Hamiltonian has eigenstates, φβ (ri ), which satisfy the eigen-
value equation
Ĥi φβ (ri ) = Eβ φβ (ri ) (287)
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 ) (288)

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 (289)
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

107
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 (290)

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± (291)
±

where the normalization condition is given by


X
| γ± |2 = 1 (292)
±

The two basis spinors χ± are usually denoted by the two component column
vectors  
1
χ+ = (293)
0
corresponding to an eigenstate of the Pauli matrix σz with spin up and
 
0
χ− = (294)
1

corresponding to the eigenstate with spin down. Thus, the arbitrary state can
be written as  
γ+
χσ = (295)
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) (296)

108
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− (297)
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 ) (298)
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
(299)
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 (300)
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 ) (301)
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
(302)
This two-particle density correlation function is normalized to twice the number
of pairs of electrons, Ne ( Ne − 1 ).

——————————————————————————————————

109
7.1.1 Exercise 25
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
translation through any Bravais lattice vector Ri

Vions (r − Ri ) = Vions (r) (303)

General properties of the solution of the Schrodinger equation for a single


electron in a solid can be found from 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
 
2
Ĥ φα (r) = − ∇ + Vions (r) φα (r) = Eα φα (r) (304)
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
Kramers (H.A. Kramers, Physica 2, 483 (1935)).

Bloch’s theorem applies to the eigenstates of the one-particle Hamiltonian,

h̄2
 
2
Ĥ = − ∇ + Vions (r) (305)
2m

in which the potential has the symmetry

Vion (r − Ri ) = Vions (r) (306)

110
for all lattice vectors Ri in the Bravais Lattice. Bloch theorem states that the
eigenfunctions can be found in the form
 
φn,k (r) = exp i k . r un,k (r) (307)

where the function un,k is invariant under the translation through any Bravais
Lattice vector
un,k (r − Ri ) = un,k (r) (308)

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) (309)

Thus, a translation of the wave function through a reciprocal lattice vector only
shows up through the presence of an exponential factor. Furthermore, if the
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) (310)

This translation operator can be applied to the wave function Ĥ φ(r) which
yields
T̂R Ĥ φ(r) = Ĥ(r − R) φ(r − R)
= Ĥ(r) φ(r − R)
= Ĥ T̂R φ(r) (311)
Thus, the Hamiltonian commutes with the translation operator which produces
a translation through a Bravais lattice vector,
[ Ĥ , T̂R ] = 0 (312)

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) (313)

111
Thus, the translation operators commute

[ T̂R , T̂R0 ] = 0 (314)

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) (315)

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) (316)

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) (317)

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 ) (318)

Since a general Bravais lattice vector can be expressed as the sum

R = n1 a1 + n2 a2 + n3 a3 (319)

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 (320)
Hence, on defining
 
c(a1 ) = exp − i 2 π x1
 
c(a2 ) = exp − i 2 π x2
 
c(a3 ) = exp − i 2 π x3 (321)

112
one can define a vector k via
 
k = x1 b(1) + x2 b(2) + x3 b(3) (322)

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 (323)

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) (324)

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 functions. 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) (325)

then it should be clear that as


 
exp iQ.R = 1 (326)

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) (327)
is adopted, then the eigenvalues must be related through
En,k+Q = En,k (328)
Thus, if k is real, any k value can be restricted to lie within one unit cell of
reciprocal space (F. Bloch, Zeit. für Physik, 52, 555 (1928)).

113
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 N = N1 N2 N3 lattice points.

Born-von Karman or periodic boundary conditions are imposed on the wave


function

φn,k (r − Ni ai ) = φn,k (r) f or i = 1 , 2 or 3 . (329)

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.
Application of Bloch’s theorem yields the condition

φn,k (r) = φn,k (r − Ni ai )


 
= exp − i Ni k . ai φn,k (r) f or i = 1 , 2 or 3

(330)

Thus, the periodic boundary conditions are fulfilled if the wave vectors k satisfy
the conditions  
exp − i Ni k . ai = 1 (331)

Since k can be written in terms of the primitive reciprocal lattice vectors, b(i) ,
via
i=3
X
k = xi b(i) (332)
i=1

and as ai . b(j) = 2 π δij , then the periodic boundary conditions require that
 
exp − i 2 π Ni xi = 1 f or i = 1 , 2 or 3

(333)

Thus, the components xi must be in the form of ratios


mi
xi = (334)
Ni

114
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 (335)
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) (336)
N
As the volume of the Brillouin zone is given by
 
b(1) . b(2) ∧ b(3) (337)

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
= (338)
N Vc
Now, since the volume of the solid V is N times the volume of the cell Vc ,
V = N Vc (339)
then the volume of k space associated with each Bloch state is
( 2 π )3
∆3 k = (340)
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 (341)
( 2 π )3

115
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
φn,k (r) = Cq exp i q . r (342)
q

where the wave vectors q are to be related to k. From Bloch’s theorem, the
Bloch functions can also expressed as
 
φn,k (r) = exp i k . r un,k (r) (343)

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 iQ.r (344)
Q

and the inverse transform is given by the integral


Z  
1 3
un,k (Q) = d r un,k (r) exp − i Q . r (345)
V V

On comparing the above two forms for the Bloch functions, one has
X  
φn,k (r) = Cq exp i q . r
q
X  
= un,k (Q) exp i(k + Q).r
Q

(346)

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 (347)
Q

116
and the inverse transform is given by the integral
Z  
1 3
Vions (Q) = d r Vions (r) exp − i Q . r (348)
V V

Furthermore, since Vions (r) is real, the Fourier transform of the potential has
the symmetry

Vions (−Q) = Vions (Q) (349)
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) (350)

and this implies that the Fourier transform of the potential has the property

Vions (Q) = Vions (−Q) = Vions (Q) (351)

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  
= Cq exp i q . r (352)
q
2m

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    
Vions (r) φn,k (r) = Vions (Q0 ) Cq0 exp i q 0 + Q0 . r (353)
q0 Q0

The form of the energy eigenvalue equation is simplified if q 0 is expressed as


q 0 = q − Q0 , so
X X  
Vions (r) φn,k (r) = Vions (Q0 ) Cq−Q0 exp i q . r (354)
q Q0

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 (355)
q
2m 0 Q

117
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 (356)
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”

(357)

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 (357) 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  
φn,k (r) = Ck−Q exp i ( k − Q ) . r
Q
  X  
= exp + ik.r Ck−Q exp − iQ.r
Q

(358)

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 (359)
Q

In order to make this approach tractable, it is necessary to truncate the infinite


set of coupled equations (357) 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.

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

118
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,
 
p̂ φn,k (r) = h̄ k φn,k (r) − i h̄ exp ik.r ∇ un,k (r) (360)

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 (361)

On substituting this relation, the Bloch function is re-written as


      
exp i k . r un,k (r) = exp i k 0 . r exp i Q . r un,k (r)
 
= exp i k 0 . r ũn,k (r) (362)

where  
ũn,k (r) = exp iQ.r un,k (r) (363)

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) since it has the periodicity of
the Bravais Lattice as  
exp i Q . R = 1 (364)

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
number 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. 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  
φn,k (r) = exp i k . r un,k (r) (365)

119
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) (366)
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 the 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 (367)
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 = (368)
V
The quantity A is expressed as the integral
V X Z
A = 2 d3 k An,k (369)
( 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 EF , 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 E, by introducing the one-electron density of states ρ(E). The density
of states per spin is defined by the integration over the dirac delta function
X Z d3 k
ρ(E) = V δ( E − En,k ) (370)
n
( 2 π )3

If the quantity An,k only depends on (n, k) through En,k , then


An,k = A(E) (371)

120
so the quantity A can be represented as an integral over the density of states
Z EF
A = 2 dE ρ(E) A(E) (372)
−∞

The density of states ρ(E) can be calculated by noting that the infinitesimal
R E+∆E
integral E ρ(E) dE ∼ ρ(E) ∆E is the number of states in the energy
range between E and E + ∆E, or the allowed number of k values between E
and E + ∆E in each of the energy bands. Thus, on integrating over an energy
range ∆E and using the definition of the density of states in terms of the dirac
delta function, one finds
E+∆E
X Z d3 k
Z
ρ(E) ∆E ∼ V dE δ( E − En,k )
n E ( 2 π )3
X Z  
V 3
= d k Θ(E + ∆E − E n,k ) − Θ(E − E n,k )
( 2 π )3 n
(373)

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 E and E + ∆E. Furthermore, since ∆E is an infinitesimal quantity,
∆E can be expressed in terms of the perpendicular distance between the two
surfaces of constant energy.

Let Sn (E) be the surface En,k = E lying within the primitive cell and let
δk(k) be the perpendicular distance between the surfaces Sn (E) and Sn (E +
dE) at point k. Then, as Sn (E) is a surface of constant E and ∇ En,k is
perpendicular to that surface

E + ∆E = E + | ∇ En,k | δk(k)

∆E
δk(k) = (374)
| ∇ En,k |

Hence, the density of states can be expressed as an integral over a surface of


constant energy
X Z d2 S 1
ρ(E) = V 3
(375)
n Sn (E) ( 2 π ) | ∇ En,k |

This gives an explicit relation between the density of states and the band struc-
ture.

121
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 (376)
The band energy En,k must have at least one maximum and one minimum in
the Brillouin zone. At each of these k points, the integrand in ρ(E) diverges.
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 theorem (L. van Hove, Phys. Rev. 89, 1189 (1953), also
see the discussion by H.P. Rosenstock, Phys. Rev. 97, 290 (1955)).

In three dimensions these singularities are integrable. That is, the integra-
tion over the surface area yield a finite value for ρ(E). In the three-dimensional
case the divergences show up in the slopes of the density of states ∂ρ(E)∂E , and
are the van Hove singularities. The van Hove singularities at the density of
states occur at the values of E where ∇ En,k vanishes at some points of the
surface Sn (E). Typical van Hove p singularities occur at the band edges where
the density of states varies as | E | . Although the density of states ρ(E)
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 26
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
Ek = E0 + E1 αi ki2 a2i (377)
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 (378)

Characterize the different types of van Hove singularities in the density of states
and sketch the energy dependence in the vicinity of the singularity for d = 1, 2

122
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
two in the successively 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.

Two different types of ground states result.

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
an energy interval which separates the lowest unoccupied band state and the
highest occupied band state. The density of states must be zero in this energy
interval. The width of the interval, where ρ(E) = 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-
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.

Metals.

A number of bands may be partially filled. In this case, the highest occu-
pied Bloch states have an energy EF 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 EF is non-zero, ρ(EF ) 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

123
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
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 = EF (379)
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. If the full reciprocal lattice is used, the Fermi-surface is
represented in the extended zone scheme. If the Fermi-surface is represented
within a single primitive unit cell of the reciprocal lattice, it is represented in a
reduced zone scheme.

124
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.

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 model works for two main 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.

8.1.1 Perturbation Theory


The wave function for an electron in a Bloch state with wave vector k is given
by  
X
φk (r) = Ck−Q exp i ( k − Q ) . r (380)
Q

125
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
( k − Q )2 − E Ck−Q + Vions (Q − Q0 ) Ck−Q0 = 0 (381)
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 solutions of this equation for fixed k are
labelled by 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 (382)

where the zero-th order energy eigenvalues are given by

(0) h̄2
Ek − Q = ( k − Q )2 (383)
2m
and the zero-th order energy eigenfunctions are
 
(0) 1
φk (r) = √ exp i ( k − Q ) . r (384)
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 (385)
1 2 m

(0)
then φk (r) can be made of any linear combination of the functions 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.

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) | (386)
1

126
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 (387)
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 (388)
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 (389)
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) (390)
1

This only yields a constant shift in the zero-th order energy eigenvalues which
can be absorbed into the definition of the reference energy. It is also seen
from eqn(389) that, to first order, the change in the coefficient Ck−Q remains
1
undetermined, so we may set
(1) (0)
Ck−Q = Ck−Q (391)
1 1

(0)
This is seen by substituting the first-order expression for E − Ek−Q into
1
eqn(389). In the following discussion, we shall neglect the effect of the average
potential V (0)ions .

The coefficients of the other plane wave components of the Bloch function
satisfy
 
(0)
X
Ek − Q − E Ck−Q + Vions (Q − Q0 ) Ck−Q0 = 0 (392)
Q0

127
(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 (393)
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(393), into the secular equation which
determines Ck−Q , eqn(387), one finds
1

(0)
 X | Vions (Q − Q) |2 (0)
1
E − Ek − Q Ck−Q = (0)
Ck−Q (394)
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)
(395)
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)
(396)
1
Q ( Ek − Q − Ek − Q )
1

This relation shows that weakly perturbed non-degenerate bands repel each
other. For example, if
(0) (0)
E k − Q > Ek − Q (397)
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 (398)
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.

128
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 C
coefficients. These two coefficients satisfy the pair of equations
(0)
( E − Ek − Q ) Ck−Q = Vions (Q2 − Q1 ) Ck−Q (399)
1 1 2

and
(0)
( E − Ek − Q ) Ck−Q = Vions (Q1 − Q2 ) Ck−Q (400)
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 (401)
1 2

This quadratic equation has the solution for the energy eigenvalue
v
 E (0) (0) u (0) (0)
k − Q + Ek − Q k − Q − Ek − Q
 u E 2
1 2
± 1 2
+ | Vions (Q1 − Q2 ) |2
t
E =
2 2
(402)
Whenever the Bloch wave vector k takes on special values such that unperturbed
bands cross
(0) (0)
Ek − Q = E k − Q (403)
1 2

the energy bands simplify to yield the two branches


(0)
E = Ek − Q ± | Vions (Q1 − Q2 ) | (404)
1

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 (405)
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” (406)

for some reciprocal lattice vector Q” 6= 0. This requires that vector q lies on
the Bragg plane bisecting Q”, as this condition reduces to

Q”2 = 2 q . Q” (407)

129
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”) | (408)
The coefficients corresponding to these energies are found from the two coupled
equations. In this case, where the unperturbed bands cross, the coefficients are
related via  
Cq = ± sign Vions (Q”) Cq−Q” (409)

which produces two standing wave solutions. If Vions (Q”) > 0, then the pair
of states are the anti-bonding state
Q” . r
 
| φ+
q (r) |2
∝ cos 2
2

E+ = Eq(0) + | Vions (Q”) | (410)

and the bonding state


Q” . r
 
| φ−
q (r) |
2
∝ sin2
2

E− = Eq(0) − | Vions (Q”) | (411)

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
| φq (r) | ∝ sin
2

E+ = Eq(0) + | Vions (Q”) | (412)

while the bonding state is given by the other form


Q” . r
 
− 2 2
| φq (r) | ∝ cos
2

E− = Eq(0) − | Vions (Q”) | (413)

In this context, the wave function


Q” . r
 
φpq (r) ∝ sin (414)
2

130
is called p-like as it vanishes at the lattice points, whereas
Q” . r
 
φsq (r) ∝ cos (415)
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
(416)
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), ( when Vions (Q”) < 0 ).

The Bragg planes have other significance as can be inferred from the gradient
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
(417)
which is found as
 
(0) (0)
" Eq − Eq − Q” #
h̄2 Q” Q”

∇ q E± = q− ± s
m 2 2 2
(0) (0)
Eq − Eq − Q” + 4 | Vions (Q”) |2
(418)
On the Bragg plane, one has
(0)
Eq(0) = Eq − Q” (419)

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 − (420)
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.

131
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 (421)
δ → 0 2δ
However, since the point k is equivalent to the point k − Q, one has

Ek+δQ = E−Q+k−δQ (422)

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 (423)
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 (424)

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.

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.

132
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 correspond to

Γ ≡ (0, 0, 0) T he zone center

K ≡ ( 34 , 34 , 0) T he hexagonal edge center

X ≡ (1, 0, 0) T he diamond f ace center

W ≡ (1, 21 , 0) T he corner

L ≡ ( 12 , 12 , 12 ) T he hexagonal f ace center

The electron bands are usually plotted against k along the high symmetry
directions

Γ → X → W → L → Γ → K → X

2 π
The length of these linear segments ( in units of a ) are given by
√ √ √
1 √1 3 3 2 10
1 2 2 2 4 4

The band energies in the empty lattice approximation can be plotted along
these axes in units of E0 where
h̄2 4 π2
 
E0 = (425)
2m a2

and the components of the reduced wave vectors k̃i where


ki a
k̃i = (426)

0
The energy of the various bands can be constructed from the various Ek−Q .
The first band that is considered is simply Ek0 , where Q = (0, 0, 0) thus,

Ek0
 
= k̃x2 + k̃y2 + k̃z2 (427)
E0

133
which for Γ → X is just

= k̃x2 f or 0 ≤ k̃x ≤ 1 (428)

For X → W this band dispersion is given


1
= 1 + k̃y2 f or 0 ≤ k̃y ≤ (429)
2
For W → L this band is described by
1 1
= + k̃x2 + ( 1 − k̃x )2 f or ≤ k̃x ≤ 1 (430)
4 2
For L → Γ this dispersion is given as
1
= 3 k̃x2 f or 0 ≤ k̃x ≤ (431)
2
For Γ → K this band takes the form
3
= 2 k̃x2 f or 0 ≤ k̃x ≤ (432)
4
The last segment is given by K → X in which the band takes the form
3
= k̃x2 + 9 ( 1 − k̃x )2 f or ≤ k̃x ≤ 1 (433)
4

0 4 π
The next band to be considered is simply Ek−Q , where Q = a (1, 0, 0)
thus,
0
Ek−Q  
2 2 2
= ( k̃x − 2 ) + k̃y + k̃z (434)
E0
which for Γ → X is just

= ( k̃x − 2 )2 f or 0 ≤ k̃x ≤ 1 (435)

For X → W this band dispersion is given


1
= 1 + k̃y2 f or 0 ≤ k̃y ≤ (436)
2
For W → L this band is described by
1 1
= + ( k̃x − 2 )2 + ( 1 − k̃x )2 f or ≤ k̃x ≤ 1 (437)
4 2
For L → Γ this dispersion is given as
1
= ( k̃x − 2 )2 + 2 k̃x2 f or 0 ≤ k̃x ≤ (438)
2

134
For Γ → K this band takes the form
3
= ( k̃x − 2 )2 + k̃x2 f or 0 ≤ k̃x ≤ (439)
4
The last segment is given by K → X in which the band takes the form
3
= ( k̃x − 2 )2 + 9 ( 1 − k̃x )2 f or ≤ k̃x ≤ 1 (440)
4

4 π
0
The next band is Ek−Q , where Q = a ( 12 , 12 , 12 ) thus,

0
Ek−Q  
= ( k̃x − 1 )2 + ( k̃y − 1 )2 + ( k̃z − 1 )2 (441)
E0

which for Γ → X is just

= 2 + ( k̃x − 1 )2 f or 0 ≤ k̃x ≤ 1 (442)

For X → W this band dispersion is given


1
= 1 + ( k̃y − 1 )2 f or 0 ≤ k̃y ≤ (443)
2
For W → L this band is described by
1 1
= + ( k̃x − 1 )2 + k̃x2 f or ≤ k̃x ≤ 1 (444)
4 2
For L → Γ this dispersion is given as
1
= 3 ( k̃x − 1 )2 f or 0 ≤ k̃x ≤ (445)
2
For Γ → K this band takes the form
3
= 1 + 2 ( k̃x − 1 )2 f or 0 ≤ k̃x ≤ (446)
4
The last segment is given by K → X in which the band takes the form
3
= 1 + ( k̃x − 1 )2 + ( 2 − 3 k̃x )2 f or ≤ k̃x ≤ 1 (447)
4

4 π
0
The next band is Ek−Q , where Q = a ( 12 , − 12 , 12 ) thus,

0
Ek−Q  
2 2 2
= ( k̃x − 1 ) + ( k̃y + 1 ) + ( k̃z − 1 ) (448)
E0

135
which for Γ → X is just

= 2 + ( k̃x − 1 )2 f or 0 ≤ k̃x ≤ 1 (449)

For X → W this band dispersion is given


1
= 1 + ( k̃y + 1 )2 f or 0 ≤ k̃y ≤ (450)
2
For W → L this band is described by
9 1
= + ( k̃x − 1 )2 + k̃x2 f or ≤ k̃x ≤ 1 (451)
4 2
For L → Γ this dispersion is given as
1
= 2 ( k̃x − 1 )2 + ( k̃x + 1 )2 f or 0 ≤ k̃x ≤ (452)
2
For Γ → K this band takes the form
3
= 1 + ( k̃x − 1 )2 + ( k̃x + 1 )2 f or 0 ≤ k̃x ≤ (453)
4
The last segment is given by K → X in which the band takes the form
3
= 1 + ( k̃x − 1 )2 + ( 4 − 3 k̃x )2 f or ≤ k̃x ≤ 1 (454)
4

It is seen that some branches of these bands are highly degenerate. When
Vions 6= 0, the degeneracy of the various branches may be lifted. Group theory
can be used to determine whether or not the potential lifts the degeneracy of
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. Segall, Physical Review 124, 1797 (1961).

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 ) (455)
2 a

136
The Cartesian coordinates of the high symmetry points are

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 27
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 ).

——————————————————————————————————

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 theory (L.P. Bouckaert, R. Smoluchowski and E. Wigner,
Phys. Rev. 50, 58 (1936)).

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) (456)

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

T̂ (R) Ô(Aj ) φn,k (r) = T̂ (R) φn,k (A−1


j r)
 
= exp − i k . A−1 j R φn,k (A−1
j r) (457)

137
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 Rj−1 R )
= ( Aj k ) . R (458)

we find that
 
T̂ (R) Âj φn,k (r) = exp − i ( Aj k ) . R φn,k (A−1
j r)
 
= exp − i ( Aj k ) . R Ô(Aj ) φn,k (r)

(459)

Since the quantity  


exp − i ( Aj k ) . R (460)

is the eigenvalue of the translation operator T̂ (R), the Bloch wave vector of
the function Ô(Aj ) φn,k (r) is Aj k. As this is an energy eigenfunction, the
transformed function is a Bloch function. That is,

φn,Aj k (r) = Ô(Aj ) φn,k (r) (461)

Since the point group symmetry operations commute with the Hamiltonian,

[ Ĥ , Ô(Aj ) ] = 0 (462)

the Bloch states Ô(Aj ) φn,k (r) all have the same energy En,k .

A basis set can be constructed by repeated application of the point group


symmetry operators on the Bloch functions. The same vector k cannot appear
in distinct bases created from a Bloch function 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 (463)

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

138
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 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 (464)

where Q is a reciprocal lattice vector. As an example, the groups of the k vec-


tors for the points Γ, Z, M, A of the simple tetragonal lattice coincide with the
D4h point group of the tetragonal lattice itself. The groups of the k 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.
The group of the k vector has irreducible representations, and these are called
the small representations.

The basis functions 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 vector in the star may be
different, the small representation of any one can be chosen for the symmetriza-
tion 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 the
equivalent wave vectors. The basis functions corresponding to the different ir-
reducible 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
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

139
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 nearly 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.

At a general point along the eight [1, 1, 1] directions, the group of the k
vector is C3v and contains six elements in three classes. The are the identity
E, a class consisting of the rotation C3 about the [1, 1, 1] axis and its inverse
C3−1 , and 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 axis are either singly or doubly degenerate,
when the crystalline potential is introduced. The branch which emanates from

140
2
k = 0 with energy E = 2h̄m k 2 belongs to the Λ1 representation as this is
compatible with the Γ1 representation.

At the end point L where k = πa (1, 1, 1), the symmetry operations are iden-
tical to those of Λ. In the free electron approximation, the state at L is doubly
degenerate (ignoring spin) since the wave vectors πa (1, 1, 1) and − πa (1, 1, 1)
differ by a reciprocal lattice vector Q = 2aπ (1, 1, 1). Using the compatibility
relations, one can show that the next highest band has Λ1 symmetry. These two
levels 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 V (Q).

On continuing along the band with Λ1 symmetry, one reaches the point
k = 2aπ (1, 1, 1). Since the primitive 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) (465)
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
 

φkj (r) ∼ exp i ( ± x ± y ±z ) (466)
a
are degenerate. They form the basis of an eight-dimensional representation
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 (467)
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

141
This eight-dimensional representation, Γ, is reduced into the irreducible repre-
sentations, Γµ , via X
Γ = aµ Γµ (468)
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) (469)
µ

The multiplicity aµ can be found from the orthogonality relation


X
gi χ(Ai ) χµ (Ai ) = g aµ (470)
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 ) (471)

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

nµ X µ
P̂ µ = χ (A) Ô(A) (472)
g
A

In this, nµ is the dimension of the µ-th irreducible representation, i.e., nµ =


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

Ôµ φk (r) = φµk (r) (473)

In this way, one can construct the set of symmetrized basis functions:

142
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 levels may be split by the
application of a potential, however, the degeneracies cannot be completely lifted.

Along the X direction, 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

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

143
about the y and z axis, and 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
potential.

On continuing along the X direction, one reaches the point (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
 

φk (r) = exp ± i x
a
 

φk (r) = exp ± i y
a
 

φk (r) = exp ± i z
a
(474)
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 (475)
µ

The multiplicities aµ are calculated from


X
gi χ(Ai ) χµ (Ai ) = g aµ (476)
i

144
which leads to the decomposition

Γ = Γ1 + Γ3 + Γ4 (477)

into a one-dimensional, a two-dimensional and a three-dimensional irreducible


representation. The basis functions can be symmetrized 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 cos4πx
a − cos 4πy
a − cos 4πz
a

Γ4
sin 4πx
a
sin 4πy
a
sin 4πz
a

h̄2
Hence, the six-fold degenerate energy level E 0 = 2m ( 4aπ )2 may have the
degeneracy lifted by V (Q).

——————————————————————————————————

8.1.7 Exercise 28
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
a3 a a a
r
8 2πx 2πy 2πz
φΓ4 (x) = sin cos cos
a3 a a a
r
8 2πx 2πy 2πz
φΓ4 (y) = cos sin cos
a3 a a a
r
8 2πx 2πy 2πz
φΓ4 (z) = cos cos sin
a3 a a a
(478)

145
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
(479)
while all other matrix elements are zero.

——————————————————————————————————

8.1.8 Brillouin Zone Boundaries


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 should be per-
pendicular 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 energy
near the Bragg plane, and also due to the accompanying band splitting, the cir-
cles 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 of the empty lattice approxima-
tion.

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

146
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 repeated zone scheme.

The Hume-Rothery rules provide a correlation of crystal structure with the


number of electrons per unit cell, or band filling. 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. If it is assumed that the noble metals have one
electron outside the closed d shell, then the alloys have an f.c.c. phase for
an average number of electrons per atom up to 1.38, while the b.c.c. phase
is stable for band-fillings between 1.38 and 1.48. In the f.c.c. structure, the
smallest vectors from the zone center to each face of the Brillouin zone have the
form 21 2aπ (1, 1, 0), whereas for the f.c.c. lattice these vectors are of the form
1 2 π
2 a (1, 1, 1). Therefore, the radius of the Fermi-sphere, kF , at which it first
makes contact with the Brillouin zone boundary is given by
√ π
kF = 3 f or f.c.c.
a
√ π
kF = 2 f or b.c.c. (480)
a
When the Fermi-sphere first makes contact with the zone boundary, the occu-
pied band is depressed by V (Q) resulting in an energy lowering which stabilizes
the structure. In the free electron approximation, the number of electrons per
primitive unit cell, n, is given by
V 4π 3
n = 2 k (481)
N ( 2 π )3 3 F
where
V a3
= f or f.c.c.
N 4
V a3
= f or b.c.c. (482)
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.

147
8.1.9 The Geometric Structure Factor
The potential Vions (r) is a periodic function and can be defined in terms of the
ionic potentials, Vatom , the lattice vectors R, and the basis vectors rj , via
X X
Vions (r) = Vatom (r − R − rj ) (483)
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

(484)

where we have used the Laue condition


 
exp i Q . R = 1 (485)

and the transformation r0 = r − R. Furthermore, since the Bravais Lat-


tice vectors do not explicitly appear in the summand, the sum over R merely
produces a factor of N X
≡ N (486)
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) (487)
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) (488)
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

148
Bragg plane also vanishes. An example of this is given by the hexagonal close-
packed lattice.

The unit cell of the reciprocal lattice of the (direct space) hexagonal closed
packed lattice is a hexagonal 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 all q values on the hexagonal top and
bottom of the prism. The structure factor can be evaluated as
 
2 4
S(Q) = 1 + exp i π ( m1 + m2 + m3 ) (489)
3 3
which vanishes when m1 = m2 = 0 and m3 = ± 1, corresponding to q laying
on the 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
coupling (C. Herring, Phys. Rev. 52, 361 (1937)).

Since the gaps vanish on some faces of the Brillouin zones, it is sometimes
helpful to define a set of zones, the Jones zones, which are separated by planes
in which gaps do occur.

The spin-orbit interaction can lead to the re-occurrence of small gaps in the
bands (M.H. Cohen and L. Falicov, Phys. Rev. Letts. 5, 544 (1960)). The spin-
orbit interaction is a relativistic effect, which appears as 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 4 q h̄ q
− 3
p + 2 2
σ . ∇ φ ∧ ( p − A ) + ∇2 φ
8m c 4m c c 8 m2 c2
(490)
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 (491)
2m c
which, together with the identity
      
σ.a σ.b = σ0 a . b + iσ. a ∧ b (492)

leads to
  2
1 q
ĤP = σ0 − i h̄ ∇ − A + σ0 q φ
2m c

149
 
h̄ q
− σ. ∇ ∧ A + A ∧ ∇ (493)
2mc
which, since
 
∇ ∧ A Ψ(r) = Ψ(r) ∇ ∧ A − A ∧ ∇ Ψ(r) (494)

and as B = ∇ ∧ A, yields the non-relativistic Pauli Hamiltonian including


the anomalous Zeeman interaction
  2 
1 q h̄ q
ĤP = σ0 p − A − σ . B + σ0 q φ (495)
2m c 2mc

Thus, all the terms in the first line of equation (490) 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
(496)

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) (497)
4 m c2
Due to its reduced symmetry, the spin-orbit interaction raises the degeneracy
of the bands at high symmetry points in k space (R.J. Elliott, Phys. Rev. 96,
280 (1954)), such as those on the hexagonal faces of the h.c.p. Brillouin zone.

——————————————————————————————————

8.1.10 Exercise 29
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 (498)

150
The energy of the two bands can be written as
h̄2 2
Ek,± = k + ∆E± (kk ) (499)
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) | (500)
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± (E) = d3 k δ( E − Ek,± ) (501)
( 2 π )3

Show that the density of states is given by


   
2m V
kmaxk (E) − kmink (E) (502)
h̄2 4 π2
where E = ∆E± (kmk ) defines the maximum and minimum value of kk .

Show that, if the constant energy surface cuts the zone, i.e.,
E 0Q − | V (Q) | ≤ E ≤ E 0Q + | V (Q) | (503)
2 2

then for the lower band one has


Q
kmaxk (E) = (504)
2
and r
2mE
kmink (E) = − + O(|V (Q)|2 ) (505)
h̄2
for E > 0.

Show that
  
V m Q
ρ+ (E) = kmaxk (E) − f or E ≥ E 0Q + | V (Q) | (506)
4 π 2 h̄2 2 2

∂ρ
Show that the energy derivative of the density of states, ∂E , is singular at
the energies
E = E 0Q ± | V (Q) | (507)
2
——————————————————————————————————

151
8.1.11 Exercise 30
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) (508)
a 2

The three planes are the (2, 0, 0), (1, 1, 1) and (1, 1, 1) planes. The four free
electron energies are

h̄2 2
E10 = k
2m
2
h̄2


E20 = k − (1, 1, 1)
2m a
2
h̄2


E30 = k − (1, 1, 1)
2m a
2
h̄2


E40 = k − (2, 0, 0) (509)
2m a
h̄2
0
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

E10 − E V1 V1 V2


0


V 1 E 2 − E V 2 V 1


= 0
0


V 1 V 2 E 3 − E V 1




V2 V1 V1 E40 − E
where V2 = V (2, 0, 0) and V1 = V (1, 1, 1) = V (1, 1, 1), and that at W
the roots are
0
E = EW − V2 doubly degenerate

0
E = EW + V2 ± 2 V1 singly degenerate (510)

Two Bragg planes meet at the point U, which corresponds to the k value
 
2π 1 1
kU = (1, , ) (511)
a 4 4

152
Show that at the U point the band energies are given by

E = EU0 − V2

V2 1
q
E = EU0 + ± V22 + 8 V12 (512)
2 2

where
h̄2 2
EU0 = k (513)
2m U
is the free electron energy at point U.

——————————————————————————————————

8.1.12 Exercise 31
Consider a nearly free electron band structure near a Bragg plane. Let
Q
k = + q (514)
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 (515)
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

µ = E 0Q − | V (Q) | + ∆ (516)
2

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 = (517)
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) | (518)
h̄2

153
This area is measurable through de Haas - van Alphen experiments.

——————————————————————————————————

8.1.13 Exercise 32
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) (519)
h̄2
where
| V (Q) |2
m E
δ(θ) = √ (520)
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 (521)

where
1 | V (Q) |2
 
2 kF Q + 2 kF
∆µ = − ln (522)
8 µ0 Q Q − 2 kF

——————————————————————————————————

8.1.14 Exercise 33
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 (523)
2

(i) Find an expression for the imaginary part of k for E within the gap.

154
(ii) Show that for E at the center of the gap, the imaginary part of k satisfies
2 s 
2
Q2 Q2

2m
Im k = − ± + | 2 V (Q) |2 (524)
2 2 h̄
Q
Thus, on solving for k given E, there is a range of Im k when Re 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.

——————————————————————————————————

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 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-
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
the nuclear potential by the ion core electrons.

155
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 (525)

The valence band Bloch function can be expressed in terms of a smooth function
v
ψk,n (r) (526)

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
> (527)
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 components (J.C. Phillips and L. Kleinman, Phys. Rev. 116, 287 (1959)).

Since the Bloch state, | φvk,n >, satisfies the one-particle Schrodinger equa-
tion
Ĥ | φvk,n > = Ek,n v
| φvk,n > (528)
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
(529)

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 > (530)

where
X  
v v c
V̂ (Ek,n ) = Ek,n − Ek,α | φck,α > < φck,α | (531)
α

156
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 (532)
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 (533)
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 − Ek,αc
d3 r0 φ∗c 0 0 c
k,α (r ) Ψ(r ) φk,α (r) (534)
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
 
− ∇2 + Vions (r) φvk,n (r) = Ek,n
v
φvk,n (r) (535)
2m
v
then the Schrodinger equation for the smooth function ψk,n (r) has the form

h̄2
 
− v
∇2 + Vions (r) + V̂ (Ek,n ) v
ψk,n v
(r) = Ek,n v
ψk,n (r) (536)
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 ) (537)

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
X  
v
ψk,n (r) = Ck−Q exp i ( k − Q ) . r (538)
Q

157
Thus, the pseudo-potential may be treated as a weak perturbation and gives
results very similar to those of the nearly free electron model.

There are many different forms that the pseudo-potential can take (B.J.
Austin, V. Heine and L.J. Sham, Phys. Rev. 127, 276 (1962)). The non-local
pseudo-potential can be approximated by a local potential and, as its energy
dependence is weak, E v can be set to zero in the pseudo-potential. In this
approximation, the pseudo-potential is almost zero within the core. This is
a result of the so-called cancellation theorem (M. Cohen and V. Heine, Phys.
Rev. 122, 1821 (1961)). The cancellation theorem can be found from classical
considerations. 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 approximation to the atomic pseudo-potential (N.W. Ashcroft,
Phys. Letts. 23, 48 (1966)) is given by
Z e2
Vpseudo (r) = − f or r > Rc
r

Vpseudo (r) = 0 f or r < Rc


(539)
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 local pseudo-potential is a smooth function of


the wave vector q.
4 π Z e2
Vpseudo (q) = − cos q Rc (540)
q2
Only the values of Vpseudo (q) at the reciprocal lattice vectors Q are physi-
cally important and most of these are small. 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 f or r > Rc
r

Vpseudo (r) = 0 f or r < Rc (541)


The Fourier transform of the screened pseudo-potential is given by
4 π Z e2
Vpseudo (q) = − cos q Rc (542)
q 2 + kT2 F

158
which is weakened with respect to the original potential.

8.2.1 The Scattering Approach


The pseudo-potential is a potential that gives the same eigenvalues as 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) (543)

where
h̄2 k 2
E = (544)
2m
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) = (545)
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) = (546)
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) (547)
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 (548)

where nl is an integer chosen such that the


π
| ∆l | < (549)
2

159
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.

8.2.2 The Ziman-Lloyd Pseudo-potential


Ziman and Lloyd independently 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 (550)
l

where P̂l projects onto the states with angular momentum l (J.M. Ziman, Proc.
Phys. Soc. (London) 86, 337 (1965), P. Lloyd, Proc. Phys. Soc. (London),
86, 825 (1965)). 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 R̃l + + V (r) R̃l (r) = E R̃l (r)
2 m r2 ∂r ∂r 2 m r2
(551)
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 (552)
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).

160
Thus, the coefficient Bl (E) is related to Ll (E) via

jl0 (kR) 2m
Ll (E) − k = Bl (E) (553)
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 | (554)
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 angu-
lar momentum l if there are core states with angular momentum l otherwise, the
electrons experience the full potential. Thus, in C the 2s electron experience 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. 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.

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.

——————————————————————————————————

8.2.3 Exercise 34
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.

——————————————————————————————————

161
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 atoms (J.C. Slater and G.F. Koster, Phys. Rev 94, 1498
(1954)). 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 > (555)
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 (556)
Thus, the Hamiltonian is written as the sum of a single ion Hamiltonian and
the potential due to the rest of the ions
Ĥ = Ĥ0 + ∆V (557)

In the tight-binding method it is convenient to define Wannier functions, φen , as


a transform of the Bloch functions
X  
φk,n (r) = exp i k . R φen ( r − R ) (558)
R

Thus, the Wannier functions are centered around the different lattice points R.
The Wannier states are almost localized states and are composed of a linear
superposition of the atomic states
X
| φen > = bn,m | φm > (559)
m

The band structure is found from the energy eigenvalue equation for the Bloch
wave functions
Ĥ | φk,n > = Ek,n | φk,n > (560)
or  
Ĥ0 + ∆V | φk,n > = Ek,n | φk,n > (561)

This energy eigenvalue equation is projected onto the atomic wave function
| φm > located at O leading to
 
< φm | Ĥ | φk,n > = < φm | Ĥ0 + ∆V | φk,n >

= Ek,n < φm | φk,n > (562)

162
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 >


(563)

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 >


(564)

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
X  
φk,n (r) = exp i k . R φen ( r − R )
R
X  
= bn,m0 exp ik.R φm0 ( r − R )
R,m0

(565)

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
< φm | φk,n > = δm,m0 bn,m0 +
m0
X X   Z
+ bn,m0 exp ik.R d3 r φ∗m (r) φm0 (r − R)
m0 R6=0

(566)

Substituting this into the energy eigenvalue equation, one obtains the equation
X  
Ek,n − Em δm,m0 bn,m0 +
m0
  X   Z
+ Ek,n − Em bn,m0 exp ik.R d3 r φ∗m (r) φm0 (r − R)
m0 ,R6=0

X Z
= bn,m0 d3 r φ∗m (r) ∆V (r) φm0 (r) +
m0
X   Z
+ bn,m0 exp ik.R d3 r φ∗m (r) ∆V (r) φm0 (r − R)
m0 ,R6=0

(567)

163
The first term on the left side involves the overlap of two atomic wave function
both centered at site 0. These atomic wave functions are part of an orthonormal
set of eigenfunctions. The second term on the left hand side involves the overlap
of atomic wave functions at site 0 and site R, and may be expected to be
exponentially smaller than the first term.
Z
3 ∗

1  d r φm (r) φm0 (r − R) (568)

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 re-
maining term represents the delocalization of the electrons. The magnitudes of
the coefficients bn,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 .
Generally, this 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 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 .

On keeping the two center and three center integrals in which R is limited
to a few neighbor sites to O, and to atomic states with a few energies close to
Em , the set of equations truncate into a finite set. These 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) (569)

in which φj are atomic wave functions and R represent atomic positions relative
to the central atom 0. 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 to have wave functions with small spatial extents
form narrow bands while the higher energy bands have broader band widths.

164
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.
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 approximation 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) (570)
The overlap between hydrogen-like 1s wave functions
r
κ3
 
φ1s (r) = exp − κ r (571)
π
can be evaluated from the Fourier transformed wave function
r
κ3 8πκ
φ1s (q) = (572)
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 (573)
( 2 π )3
with the result that
   
1 2 2
t1s,1s,σ = − 1 + κR + κ R exp − κR (574)
3
On using the hydrogenic-like 2s and 2p wave functions,
r
κ3
   
φ2s (r) = 1 − κr exp − κ r
π

165
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

(575)

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 ) (576)
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
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
   
13 3 3
t2s,2p,σ = κ R exp − κ R
30
   
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
(577)

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
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 (578)

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 differ-
ent 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,

166
the tight-binding overlap integrals between these sets of states are denoted,
respectively, by td,d,δ , td,d,π and td,d,σ . The matrix elements for arbitrary orien-
tations are tabulated in the article of Slater and Koster (1954). 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 (579)

In general, the tight-binding bands obtained by considering d bands alone is


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 strongly (V. Heine,
Phys. Rev. 153, 673 (1967)).

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, Re 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 (580)

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


X X
φk (r) = exp[ i k . R ] aj,m φm (r − rj − R) (581)
R j,m

where rj are the positions of the basis atoms and aj,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
molecules. These 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 > (582)

or bs = 1. All other coefficients are set to zero, corresponding to the assumption


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

167
regarded as forming part of the core of the ions. The energy eigenvalue equation
truncates to
  X  Z 
3 ∗
Es,k − Es 1 + exp i k . R d r φs (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

(583)

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) (584)

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 (585)

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) (586)

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 −   (587)
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 − (588)
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
(589)

168
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
(590)
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 > (591)
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 ) (592)

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 (593)
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 (594)
∂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 = = (595)
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
(596)

169
This only vanishes along the lines joining L ( 12 , 12 , 12 ) to the vertices W (1, 12 , 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.

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 equation
by a factor of 2 (J. Friedel, P. Lenghart and G. Leman, J. Phys Chem. Solids
25, 781 (1964)).

——————————————————————————————————

8.3.2 Exercise 35
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
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) (597)

Show that the overlap parameters are given by


 
2 2
γx,x = − tppσ cos θx + tppπ sin θx
 
γx,y = − tppσ − tppπ cos θx cos θy (598)

Thus, the tight-binding parameters not only depend on the distance, R, but
also depend on the direction.

——————————————————————————————————

8.3.3 Exercise 36
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) (599)
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 (600)

170
and  
X
γi,j (k) = exp ik.R γi,j (R) (601)
R

and Z
γi,j (R) = − d3 r φ∗i (r) ∆V (r) φj (r − R) (602)

and
βi,j = γi,j (0) (603)

Show that, using cubic symmetry,

βx,x = βy,y = βz,z = β (604)

and all other overlap matrix elements are zero

βx,y = βy,z = βx,z = 0 (605)

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 + 2 tppσ cos kx a + 2 tppπ ( cos ky a + cos kz a )


Ey,k = Ep + 2 tppσ cos ky a + 2 tppπ ( cos kx a + cos kz a )
Ez,k = Ep + 2 tppσ cos kz a + 2 tppπ ( cos kx a + cos ky a )
(606)

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 .

——————————————————————————————————

8.3.4 Exercise 37
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 = − Mz E − Ek + My − Mx
(607)
1 1 0 0
− My − Mx E − E k + Mz

171
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 (608)
2 2
and cyclic permutations. The energy Ek0 is given by
 
ky a kz a kx a kz a kx a ky a
E0,k = Ep − β − 4 γ2 cos cos + cos cos + cos cos
2 2 2 2 2 2
(609)
Evaluate the integrals in terms of the overlap of atomic wave functions by us-
ing 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.

——————————————————————————————————

8.3.5 Exercise 38
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


  
X
d p a p a
φk = exp i k . R a φx2 −y2 (r) + bx φx (r − êx ) + by φy (r − êy )
2 2
R
(610)
find the energy bands for the CuO2 planes.

——————————————————————————————————

8.3.6 Exercise 39
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

172
density of states when d → ∞.

——————————————————————————————————

8.3.7 Exercise 40
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.

——————————————————————————————————

8.3.8 Wannier Functions


Consider the position r to have a fixed value. The Bloch functions can be
written as  
X
φk,n (r) = exp i k . R fn (r, R) (611)
R

The Bloch function φk for fixed r is periodic in k, with periodicity given by the
primitive reciprocal lattice vectors Q. Clearly
X  
φk+Q,n = exp i(k + Q).R fn (r, R)
R
X  
= exp ik.R fn (r, R)
R

= φk,n (612)

since Q and R satisfies 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
Z  
1 3 0 0
fn (r, R) = d k exp − i k . R φk0 ,n (r) (613)
Ωc Ω c

where the integration volume Ωc is the volume of one cell of the reciprocal lattice.

The simultaneous transformations r → r − R0 and R → R − R0 leave


fn (r, R) unchanged

fn (r, R) = fn (r − R0 , R − R0 ) (614)

173
This is proved by considering the effect of the transformation r → r − R0 on
the definition of the functions fn (r, R)
X  
φk,n (r) = exp i k . R fn (r, R0 )
0
(615)
R0

Applying the transformation on the Bloch function yields


X  
φk,n (r − R0 ) = exp i k . R fn (r − R0 , R0 )
0
(616)
R0

and then, on transforming the sum over R0 as R0 = R − R0 , one has


X  
φk,n (r − R0 ) = exp i k . (R − R0 ) fn (r − R0 , R − R0 ) (617)
R

On comparing the above expression with the result of Bloch’s theorem


 X  
φk,n (r − R0 ) = exp − i k . R0 exp i k . R fn (r, R) (618)
R

one recovers the symmetry relation

fn (r, R) = fn (r − R0 , R − R0 ) (619)

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) (620)

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
X  
φk,n (r) = exp i k . R φen (r − R) (621)
R

where φen (r) are the Wannier functions (G. Wannier, Phys. Rev. 52, 191 (1947)).
The Wannier functions at different sites are orthogonal. Thus, as they are lin-
early related to the Bloch wave functions 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
Z  
1
φen (r − R) = d3 k exp − i k . R φk,n (r) (622)
Ωc Ω c

174
The Wannier functions are localized around the site R, as can be seen by sub-
stituting the expression for the Bloch functions in the above equation
Z  
1
φen (r − R) = d3 k exp + i k . ( r − R ) un,k (r) (623)
Ωc Ω c

The phase factor in the integral over d3 k has the effect of localizing the Bloch
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
π x   sin [ π y π z

sin [ ax ] ay ]   sin [ az ]

φen (r) = π x π y π z (624)
ax ay az

which have amplitudes that decay algebraically outside the unit cell. This alge-
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 width (W. Kohn Phys. Rev. 115 (1959), E.I.
Blount, Solid State Physics, Vol 13, Acad. Press, (1962)).

——————————————————————————————————

8.3.9 Exercise 41
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 (625)

Also show that the Wannier functions are normalized to unity


Z
d3 r | φen (r) |2 = 1 (626)

——————————————————————————————————

175
9 Electron-Electron Interactions
In the last chapter, the effects of interactions between electrons were neglected
in the calculation of the energies of single-electron excitations and the single-
electron wave functions. The neglect of the effects of electron-electron interac-
tions is certainly not justifiable from considerations of the relative strength of
the effect of the Coulomb interactions with the potential due to the lattice of
nuclei compared with the electron-electron interactions. However, due to the
Pauli exclusion principle, the lowest energy excitations of an interacting elec-
tron gas can be put into a one to one correspondence 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. A important result of this blocking principle is
that the low energy excitations of an electron gas behave very similarly to that
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-
tinger, who showed that electrons with energies close to the Fermi-energy have
scattering rates that vanish as the energy approaches the Fermi-energy, to all
orders in the electron-electron interaction. 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. Although, a rigorous derivation of Fermi Liquid
theory must consider processes of all order in the electron-electron interaction,
we shall only consider the lowest 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 processes 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
− = − (627)
2m 2m 2m 2m
or
( k − k0 ) . q = q2 (628)
For fixed k and k 0 this is an equation of a sphere of diameter | k − k 0 |,
( k − k0 )
centered on 2 . Thus, q ranges from 0 to k − k 0 , and conservation of
( k − k0 )
energy ensures that k − q lies on a sphere of radius | 2 |, centered

176
( k + k0 )
at 2 , passing through k and k 0 . However, since k − q must be above
the Fermi-surface there are additional restrictions due to the Pauli exclusion
principle, namely
| k − q | ≥ kF (629)
and
| k 0 + q | ≥ kF (630)
Thus, only a segment of the surface of this sphere represents final states of the
possible processes. This segment 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 vanishes (J.J. Quinn and R.A. Ferrell, Phys. Rev. 112,
812 (1958)).

9.1.1 The Scattering Rate


The scattering rate can be evaluated from Fermi’s Golden rule
2 X
4 π e2

1 2π X m 0 2
τk
=
h̄ q 2 + k2 2 δ( ( k − k ) . q − q )
q TF k0 <k

F

(631)
The sum over k 0 is performed where k 0 lies within the Fermi-sphere.

Thus, the quasi-particle scattering rate vanishes as Ek → µ at zero tem-


perature. At finite temperatures the quasi-particle scattering rate at the Fermi-
energy varies as ( kB T )2 (E. Abrahams, Phys. Rev. 95, 834 (1954)). The
quasi-particle concept remains valid in the limit Ek → µ and T → 0.

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 system can be estimated from perturbation theory. To second order in
the perturbation, the energy of the state with an additional electron in 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

(632)

177
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
|kn |<kF k n | Ĥint | k m0 − q k m + q
<
|kn |<kF ,n6=m,m0 k n >

X X
+
Ek0 + Ek0 − Ek0 −q − Ek0 +q
q m,m0 m0 m m0 m

(633)

The excitation energy for adding an electron to state k is defined by

Ekexc = Ek+ − Egs (634)

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 m + q >
X X
+
Ek0 + Ek0 0
− Ek−q − Ek0 +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
|k+q|<kF |km |<kF m m

(635)

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 sec-
ond order terms. The first of this pair represents the scattering of the electron
from the state k from an electron k m in the Fermi-sea, to final states k − q and
k m + q above the Fermi-sea. The last term represents a subtraction, as this
represents a scattering process for a pair of electrons that initially are below the
Fermi-surface which is forbidden by the Pauli exclusion principle as the state k
is occupied by an electron. The k independent terms are absorbed into a shift
of the Fermi-energy.

This excitation represents the excitation energy for adding an electron to


state k. However, the many-body state consists of a linear superposition of
single-electron states and states where the added electron is dressed by electron-
hole pairs. The quasi-particle weight Z −1 (k) is defined as the fraction of the
initial bare electron contained in the quasi-particle. To lowest order in Ĥint , the

178
quasi-particle weight or wave function renormalization is calculated as
2

< k k | Ĥint | k − q k + q >
X X m m
Z(k) = 1 +
( Ek0 + Ek0 − Ek−q0 − Ek0 +q )2
|k−q|>kF |km |<kF m m

2

< k + q k m | Ĥint | k k m + q >
X X
+ 0
( Ek+q + Ek0 − Ek0 − Ek0 +q )2
|k+q|<kF |km |<kF m m

(636)

which is greater than unity. Thus, the fraction of the bare electron in the quasi-
particle state is always less than unity. This conclusion remains valid to all or-
ders 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 dis-
tribution 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 contained in the Fermi-function. Thus,
the concept of a Fermi-surface remains well defined for interacting electron sys-
tems.

The quasi-particle weight has the effect that the excitation energy for a single
quasi-particle is given by the expression
Ekexc
Eqp (k) = (637)
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 an reduction
of the dispersion. The reduced dispersion is interpreted in terms of an increase in
the effective mass of the quasi-particle. The density of single-electron excitations
is given by the quasi-particle contribution
X  
−1
ρqp (E) = Z(k) δ E − Eqp (k) (638)
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

179
to a specific microscopic Hamiltonian. Since the perturbation due to electron-
electron interaction is long-ranged, there are divergent terms in the perturbation
expansion. The divergent terms first appear in the expansion 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 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 asso-
ciated with any term can be eliminated 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 devel-
oped. These approximations involve the summation of infinite subsets of the
terms that appear in the perturbation expansion. One such approximation is the
Hartree-Fock approximation. The Hartree-Fock approximation is self-consistent
first order perturbation theory 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 42
Using a perturbation expansion, find the energy of a free electron gas to first
order in the electron-electron interaction.

——————————————————————————————————

9.2 The Hartree-Fock Approximation


The Hartree-Fock approximation consists of writing the many-electron wave
function as a single Slater determinant, much the same way 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 account (D.R. Hartree, Proc. Camb. Phil. Soc., 24, 89,1928)).

The Hartree-Fock approximation can be expressed in terms of the Rayleigh-


Ritz variational principle (V.A. Fock, Zeit. für Physik, 61, 126, (1930)), in which
the many-particle wave function is written as a single Slater determinant (J.C.

180
Slater, Phys. Rev. 35, 210 (1930)). The Hamiltonian operator is expressed as
X  p̂2 
1 X e2
i
Ĥ = + Vions (ri ) + (639)
2m 2 | ri − rj |
i i 6= j

The expectation value of the Hamiltonian in terms of a Slater determinant Φ of


a complete set unspecified single-electron wave functions φα,σ (r) is given by
i=N
Ye  Z 
Ĥ = d ri Φ∗α1 , . . . αNe (r1 , . . . rNe ) Ĥ Φα1 , . . . αNe (r1 , . . . rNe )
3

i=1
(640)
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
+ 3
d 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
(641)
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
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 and φα (r) → φα (r) χσ . Therefore, in the matrix elements there is
not only an integration over r, but also the matrix elements of the spin states
has 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 ri Φ∗α1 , . . . αNe (r1 , . . . rNe ) Ĥ Φα1 , . . . αNe (r1 , . . . rNe )
3

i=1

181
i=N
Xe  Z 
− λαi d3 ri φ∗αi (ri ) φαi (ri ) − 1 (642)
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) (643)
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  
δΩ = d3 r δφ∗α (r) − ∇2 + Vions (r) − λα φHF,α (r)
α
2m
e2
X Z Z
+ d3 r d3 r0 δφ∗α (r) φ∗HF,β (r0 ) φHF,β (r0 ) φHF,α (r)
| r − r0 |
α,β
XZ e2
Z
− 3
d r d3 r0 φ∗HF,β (r) δφ∗α (r0 ) φHF,β (r0 ) φHF,α (r)
| r − r0 |
α,β
(644)

The expression for δΩ must vanish identically for any of the independent and
arbitrary variations δφ∗α (r), if the Hartree-Fock wave functions φHF,α (r) mini-
mize the average energy. In order for this to be true, for each value of α, the
coefficient of δφ∗α (r) must vanish identically. After interchanging the variables
r and r0 in the last term, one finds that the normalized Hartree-Fock wave
functions must satisfy the set of equations

h̄2
 
0 = − ∇2 + 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  
3 0 ∗ 0 0
− d r φHF,β (r ) φHF,α (r ) φHF,β (r)
| r − r0 |
β
(645)

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 (646)

182
This notation recognizes that the spatial component of the wave function, ψα (r),
depends on all the quantum numbers represented by α, including the spin quan-
tum 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  
+ d3 r0 χTσ0 ψβ∗ (r0 ) ψ β (r 0
) χσ 0 ψα (r) χσ
| r − r0 |
β

e2
X Z  
− d3 r0 χTσ0 ψβ∗ (r0 ) ψ α (r 0
) χ σ ψβ (r) χσ0
| r − r0 |
β
(647)

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 (648)

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

e2
X Z  
3 0 ∗ 0
− δσ0 ,σ d r ψβ (r ) ψβ (r) ψα (r0 )
| r − r0 |
β
(649)

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 (650)
| r − r0 |
β

represents a contribution to the potential from the average electrostatic potential


due to the all electrons in the system. This potential includes the contribution
from an the electron in 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 is non-local, as it relates the unknown eigenfunction

183
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) (651)
| 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.

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
(652)
These sets of equation 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 ex-
change potentials. The procedure is repeated until self-consistency is achieved.
The contributions to the direct and exchange potentials, arising from the state
where β = α, exactly cancel in the non-local operator. Therefore, there are no
self interaction terms in the Hartree-Fock approximation. The cancellation of
the self interaction has the effect that the linear potential operator is the same
for all the single-electron wave functions.

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) 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 determining
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
ψk,σ (r) = √ exp i k . r χσ (653)
V

184
where V is the volume of the crystal. It can be seen that this is true 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 (654)

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 assume that
the Hartree-Fock solution does not break translational invariance. Magnetic
solutions which also break translation invariance have been found by Overhauser
(A.W. Overhauser, Phys. Rev. Letts. 4, 462 (1960), Phys. Rev. 128, 1437
(1962)) and also by Kohn and Nettel (W. Kohn and S.J. Nettel, Phys. Rev.
Letts. 5, 8 (1960)). 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 energy. 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

(655)

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 from
Z 2π Z π
1
Vexch (r, r0 ) = − dϕ dθ sin θ
( 2 π )3 0 0
Z kF
e2
 
× dk 0 k 02 exp i k 0 . ( r − r0 )
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
(656)

The integration over k 0 can be performed with the aid of an identity obtained

185
by differentiating the expression
Z 1
sin α
dx cos α x = (657)
0 α
with respect to α. That is,
Z 1  
sin α cos α
dx x sin α x = − (658)
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
(659)

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.

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
(660)

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
(661)
which can be evaluated directly. An alternate method involves using the con-
volution theorem, in which case the expression
e2
Z
V
Vexch (k) = − d3 r 0
(2π) 3 | r − r0 |
Z  
× d3 k 0 ψk∗0 (r0 ) ψk0 (r) exp i k . ( r0 − r )
|k0 |≤kF
(662)

can be used. The plane wave nature of the eigenfunctions can be utilized to
write the expression as
e2
Z
V 3 0
Vexch (k) = − d r ×
( 2 π )3 | r − r0 |

186
Z  
× d3 k 0 | ψk0 (r0 ) |2 exp i ( k0 − k ) . ( r − r0 )
|k0 |≤kF
(663)

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 |
(664)
Hence, the expression for the exchange contribution to the eigenvalue λk is given
by
4 π e2
Z
1 3 0
Vexch (k) = − d k
( 2 π )3 |k0 |≤kF | k − k 0 |2
Z kF
e2 | k + k0 |
= − dk 0 k 0 ln (665)
πk 0 | k − k0 |
The integral can be evaluated as
2 e2
 
k
Vexch (k) = − kF F (666)
π kF
where
1 1 − x2 |1 + x|
F (x) = + ln (667)
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 (668)
2m π kF

The total energy of the electron system is given by the sum of the kinetic
energy and the exchange energy
X h̄2 k 2
EHF = 2
2m
k
XZ e2
Z
− d3 r d3 r0 ψk∗ (r) ψk∗0 (r0 ) ψk (r0 ) ψk0 (r)
| r − r0 |
k,k0

187
X  h̄2 k 2 
= + λk (669)
2m
k

where the summations are restricted to the values of k and k 0 which are within
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
(670)
The summations over k can be evaluated by transforming them into integrals
Z 2 2
4πV 2 h̄ k
EHF = 2 dk k
( 2 π )3 k ≤ kF 2m
2
 2
kF − k 2
  
| kF + k |
Z
2e 4πV 2 1
− kF dk k + ln
π ( 2 π )3 k ≤ kF 2 4 k kF | kF − k |
2 5 2
 
V h̄ kF V e 4 1 1
= 2
− 3
kF − (671)
π 10 m π 3 12

Ne
The number of electrons, per spin, 2 is given by
Ne V 4π 3
= k (672)
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 (673)
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
= (674)
kF3
Thus, the magnitude of the Fermi-wave vector kF is given by
 1
9π 3 1
kF = (675)
4 rs a0

188
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
1  1
e2
    
9π 3 3 9π 3 1 3 1
= −
2 a0 4 5 4 rs2 2 π rs
2.21 0.9163
= − Rydbergs (676)
rs2 rs
2
where 1 Rydberg = 2ea0 . The Hartree-Fock energy has a minimum at the
rs value given by rs ∼ 4.8 and has a cohesive energy of 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 − (677)
π a30 a0
Therefore, typical values of rs are given by the density at the nuclear position
1
( 34 ) 3
rs =
Z
0.9086
= (678)
Z
and at the first Bohr radius r = Z a0
1 2
( 34 ) 3 e 3
rs =
Z
1.7696
= (679)
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 lattice potential is
zero, the Hartree-Fock approximation coincides with second order perturbation
theory. If higher order terms are included (M. Gell-Mann and K. Brueckner,
Phys. Rev. 106, 347, (1957), W.J. Carr and A.A. Maradudin, Phys. Rev. A
133, 371 (1964)), one obtains the expression for the energy per electron
 
E 2.21 0.9163
= − + 0.06218 ln r s − 0.094 + O(r s ) (680)
Ne rs2 rs
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

189
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 phase (E.P. Wigner, Phys. Rev. 46,
1002 (1934)). 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 elec-
trostatic interaction and the energies of the vibrations of the electronic lattice
(W.J. Carr, R.A. Coldwell-Horsfall, and A.E. Fein, Phys. Rev. 124, 747 (1961)).
The energy of the Wigner crystalline phase is given by

e2
 
E 1.792 2.65 0.73
= − + 3 − + ... (681)
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
(682)

On making the spin dependence explicit, by writing

φα (r) = ψα (r) χσ
φβ (r) = ψβ (r) χσ0 (683)

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
(684)
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) (685)
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

190
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
(686)

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 potential has a similar form and can be thought of arising from a
deficiency in the density of parallel spins 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 spins.

In the Hartree-Fock approximation, the energies of the excited states are


given by Koopmans’ theorem (T.A. Koopmans, Physica 1, 104 (1933)). 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 approximation, the quasi-particles
energies are given by
Eqp (k) = λk (687)
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

= (688)
2 π2 dk
k(E)

where k(E) is the value of k that satisfies the equation

Eqp (k) = E (689)

From the above, one sees that at the Fermi-energy defined by

EF = Eqp (kF ) (690)

191
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 (691)
π 2 k2 | kF − k |
which diverges logarithmically at k(EF ) = 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 caused by the neglect
of screening, results in the one-electron density of states falling to zero just at
the Fermi-energy.

——————————————————————————————————

9.2.2 Exercise 43
Show, using perturbation theory, that the second order correction to the energy
of a free electron gas is given by
2
4 π e2

(2) m X 1
∆E = − 2 (692)
h̄ 0
q 2 V q . ( k − k0 + q )
k,k ,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 value of k ∼ kF and k 0 ∼ kF .
Show that the contribution to ∆E (2) is proportional to
d3 q
Z Z
dq
3
= 4 π
q q
= 4 π ln q (693)

and, thus, diverges for q → 0.

Simple second order perturbation theory does not work for the free electron
gas. None the less, perturbation theory can be applied by using more elaborate
techniques which take into account the screening of the Coulomb interaction.

——————————————————————————————————

9.3 The Density Functional Method


The density functional method provides an exact method for calculating the
electron density and ground state energy for interacting electrons in the pres-
ence of a crystalline potential. As such, it can be used to determine the stability

192
of various lattice structures. It can also be used to determine ground state prop-
erties or static properties of the electronic systems such as those provided by
elastic scattering experiments. It is based on the Hohenberg and Kohn Theorem
(P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964)).

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
Vext (r) varies between one crystal structure and the next. The Hamiltonian is
written as the sum of the kinetic energy of the electrons and the interaction and
external potentials acting on each electron.

The Hohenberg-Kohn theorem associates every non-degenerate many-body


ground state wave function with a unique external potential. A second map
exists between the many-body ground state wave function and the ground state
electron density. Therefore, the expectation value of any ground state property
can be expressed in terms of a unique functional of the ground state density.
Having established this, the ground state properties and electron density can
then be evaluated from 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 unique energy func-
tional which contains the external potential due to the lattice potential, then
one can find the ground state energy and electron density. This functional is
not known, however, it is customary to make the local density approximation.
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 interpreted in terms of the energies of independent Bloch
electrons. The energy dispersion relations generated this way do show a marked
similarity with the experimentally determined bands of simple metals.

The basis for density functional theory is provided by a theorem proved


by Hohenberg and Kohn (P. Hohenberg and W. Kohn, Phys. Rev. 136 B864
(1964)).

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. Thus, the density ρ(r) can be used as the
basic variable. Furthermore, the energy can be expressed as a unique functional
of the energy density involving the potential due to the ionic lattice. This estab-
lishes a variational principle which can be used to calculate the electron density,
ρ(r), and the total energy of the electronic system.

193
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 ) (694)

and
Ĥ 0 Ψ0 (r1 , . . . rNe ) = E 0 Ψ0 (r1 , . . . rNe ) (695)
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 ) (696)
i=1

However, as the primed and unprimed Hamiltonian are related through

Ĥ = Ĥ 0 + V̂ − V̂ 0 (697)

and as Ψ0 is the ground state of Ĥ 0 with energy eigenvalue E 0 , the energies


satisfy an inequality
Z
E < E0 + d3 r ρ(r) ( V (r) − V 0 (r) ) (698)

However, by similar reasoning, it can also be shown that the energies also satisfy
the inequality
Z
E0 < E + d3 r ρ(r) ( V 0 (r) − V (r) ) (699)

where the prime and unprimed quantities are interchanged. The assumption
that the ground state densities of the primed and unprimed Hamiltonian are
equal has been used. Adding these two inequalities leads to an inconsistency

E + E0 < E + E0 (700)

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). This means that the electron density, ρ(r),
can be taken to be the principal variable.

194
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) (701)

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 pairwise sum of interactions

e2 ρ(r) ρ(r0 )
Z Z
ECoul [ρ] = d3 r d3 r 0 (702)
2 | r − r0 |
This yields a number which is the value of the energy, and this number depends
on the 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) (703)

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 at

195
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] + . . . (704)
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) (705)
δρ(r)

then the quantity


δF [ρ]
(706)
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) (707)

In this case, one can define a functional derivative in terms of the partial deriva-
tives,
∂f
(708)
∂ρ
and the vector quantity
∂f
(709)
∂∇ρ
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 . (710)
∂ρ ∂∇ρ

If the functions ρ satisfy appropriate conditions at the boundaries of the inte-


gration, the equation can be integrated by parts to eliminate the term
Z
∂f
d3 r ( ∇δρ ) . (711)
∂∇ρ
In this case, the first order functional derivative is evaluated as

δF [ρ] ∂f ∂f
= − ∇. (712)
δρ(r) ∂ρ ∂∇ρ

The extension to functionals containing higher order derivatives is quite straight-


forward.

196
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 − r0 ) (713)
since, for this particular choice, the value of δF 1 [ρ, δ 3 (r − r0 )] is given by

δF [ρ]
δ 1 F [ρ, δ 3 (r − r0 )] = (714)
δρ(r0 )

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)
= d3 r + d3 r 0 0
δρ(r1 ) 2 | r − r1 | 2 | r − r1 |
Z
ρ(r)
= e2 d3 r (715)
| 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 (716)

is simply evaluated as
δFn [ρ]
= n ρ(r1 )n−1 (717)
δρ(r1 )
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 (718)
∂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 (719)
δρ(r)

197
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 F [ρ] is an extremal. In order that the functional F [ρ] is minimized, we
require that
δ 2 F [ρ, δρ] > 0 (720)
for every δρ.

The second order functional derivative is defined via


δ 2 F [ρ]
Z Z
δ 2 F [ρ, δρ] = d3 r d3 r0 δρ(r) δρ(r0 ) (721)
δρ(r) δρ(r0 )

On using the choice


δρ(r) = δ 3 (r − r1 ) (722)
0
for the deviation centered at r in the first derivative and the choice

δρ0 (r0 ) = δ 3 (r0 − r2 ) (723)

when differentiating the second time, one obtains

δ 2 F [ρ]
δ 2 F [ρ, δρ, δρ0 ] = (724)
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
= (725)
δρ(r1 ) δρ(r2 ) | r1 − r2 |
A second example is provided by the functional derivative of the mono-nomial
Z
Fn [ρ] = d3 r ρ(r)n (726)

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 (727)
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) (728)

198
in which the energy functional F [ρ] depends on the kinetic energy T̂ given by
Ne
h̄2 X
T̂ = − ∇2 (729)
2 m i=1 i

and the electron-electron interaction energy, V̂int , given by

1 X e2
V̂int = (730)
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
(731)

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.

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
(732)

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 
≤ d ri Ψ0∗ (r1 , . . . rNe ) Ĥ Ψ0 (r1 , . . . rNe )
3
(733)
i=1

and so
E[ρ] ≤ E[ρ0 ] (734)

199
This establishes the minimum principle for the energy functional
δE[ρ]
= 0 (735)
δρ(r)
subject to the constraint that the total number of electrons are fixed
Z
d3 r ρ(r) = Ne (736)

The condition that ρ is V -representable may be replaced by a less stringent


condition of N representable (M. Levy, Proc. Nat. Acad. Sci. 76, 6062 (1979)),
which only requires
ρ(r) > 0
Z
d3 r ρ(r) = Ne
Z
1
d3 r | ∇ ρ 2 (r) |2 < ∞ (737)

Having established the existence of the variational function, the precise form of
the functional remains to be determined.

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 with the electrons in the system and, therefore,
represents the Hartree terms. That is, the energy functional representing the
kinetic and electron-electron interaction energies is written as
ρ(r) ρ(r0 )
Z Z
1
F [ρ] = 3
d r d3 r 0 + G[ρ] (738)
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 |
(739)
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) + e 2
d3 r 0 (740)
| 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) (741)

200
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 
3
Ω[ρ] = E[ρ] − µ d r ρ(r) − Ne (742)

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)
2 Z
δ 2 Ω[ρ]
Z
λ
+ d3 r d3 r0 δρ(r) δρ(r0 ) + . . . (743)
2 δρ(r) δρ(r0 )
The extremal condition
δΩ[ρ]
= 0 (744)
δρ(r)
becomes
δE[ρ]
= µ (745)
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
1
δE = 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) (746)
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)
(747)
Since the first two terms are identified with the electrostatic potential, the func-
tional derivative is given by
ρ(r0 )
Z
δE[ρ]
= Vions (r) + e 2
d3 r 0
δρ(r) | r − r0 |
δG[ρ]
= − | e | φes (r) + (748)
δρ(r)

201
Hence, Ω is minimized if ρ satisfies the equation

δG[ρ]
− | e | φes (r) + = µ (749)
δρ(r)

For large Ne , µ is equal to the chemical potential given by


∂E
µ = (750)
∂Ne

9.3.5 The Kohn-Sham Equations


The Kohn-Sham equations provide a formal correspondence between the many-
body problem and an effective (non-interacting) one-body problem (W. Kohn,
L.J. Sham, Phys. Rev. 140, A1133 (1965)). This 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] = d ri Ψ∗ (r1 , . . . rNe ) T̂ Ψ(r1 , . . . rNe )
3
(751)
i=1

so the non-electrostatic contribution to the energy functional may be written as


the sum
G[ρ] = T [ρ] + Exc [ρ] (752)
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 (753)
δρ(r) δρ(r)

Thus, the quantity


δExc [ρ]
− | e | φes (r) + (754)
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) + (755)
δρ(r)

Thus, minimizing the energy functional entails solving the equation

δT [ρ]
Vef f [ρ, r] + = µ (756)
δρ(r)

202
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) (757)

in which the electron-electron interaction terms are absent. The variational


procedure leads to
δT [ρ]
Vs (r) + = µ (758)
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) (759)
2m
and then the electron density is given by
i=N
Xe
ρ(r) = | φs,αi (r) |2 (760)
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) (761)
2m
and the electron density is given by
i=N
Xe
ρ(r) = | φef f,αi (r) |2 (762)
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
Ne
X Z
T [ρ] = λef f,αi − d3 r Vef f [ρ, r] ρ(r) (763)
i=1

Thus, one also has to minimize the sum of the effective one-body eigenvalues
Ne
X
λef f,αi (764)
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

203
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, or equivalently, that the non-local terms in the density functional
can be expanded in powers of the gradient. This expansion could 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) | + . . . (765)

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 expressions for the uniform electron gas is replaced by similar
expressions 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
 
E = Ne − kF − 0.0311 ln kF + O(1) (766)
5 2m 4 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 Brueckner
(M. Gell-Mann and K. Brueckner, Phys. Rev. 106, 364 (1957)). For the free
electron gas, the electrostatic interaction energy between the electrons and the
smeared out lattice of ions cancels identically with the Hartree term. To obtain

204
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) (767)
4 π h̄2
Combining this together with the two relations
  13
kF = 3 π2 ρ (768)

and
Ne = V ρ (769)
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) (770)
4 π h̄2
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
(771)

Since the effective potential is given by the sum

δG[ρ]
Vef f [ρ, r] = − e φ(r) + (772)
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) (773)
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 coefficient (J.C. Slater, Phys. Rev. 81, 385 (1951)). 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 interpolates between the high density limit and
the low density limit. As the density is reduced, the electrons are expected to
undergo a phase transition and form a Wigner crystal. Since the energy is ex-
pected 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
calculations for the correlation energy of the homogenous electron gas (D.M.

205
Ceperley and B.J. Alder, Phys. Rev. Lett. 45, 566 (1980)).

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. Basically,
the electron density varies too rapidly for the gradient expansion to be useful.

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 static external potential
is 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) (774)

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 the
external potential which is defined by the Poisson equation

− ∇2 φext (r) = 4 π | e | ρext (r) (775)

The total potential φes (r) satisfies the Poisson equation


 
− ∇2 φes (r) = 4 π | e | ρext (r) − ρind (r) (776)

where ρext is assumed to have a positive charge, and the induced electron density
ρind is associated with a negative charge. 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 ) (777)
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 ) (778)
V

206
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 (779)
V V
and the Fourier transform of the dielectric constant is defined by
Z  
3
ε(q) = d r ε(r) exp − i q . r (780)
V

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) (781)

Hence, the total potential is reduced by the dielectric constant


φext (q)
φes (q) = (782)
ε(q)

The Fourier Transform of the Poisson equations yield

q 2 φext (q) = 4 π | e | ρext (q) (783)

and  
2
q φes (q) = 4 π | e | ρext (q) − ρind (q) (784)

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) (785)

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) (786)
The definition of the dielectric constant, (q), can be used to yield the relation

4 π | e | ρind (q)
ε(q) = 1 + (787)
q2 φes (q)

On expressing the total scalar potential as a potential energy term acting on


the electrons
V (q) = − | e | φes (q) (788)
and defining the response function χ(q) as the ratio of the induced density to
the potential
ρind (q)
χ(q) = (789)
V (q)

207
one finds that the expression for the dielectric constant reduces to
4 π | e |2
ε(q) = 1 − χ(q) (790)
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
techniques. However, in making approximations, it is imperative that only the
response to the total field is approximated and not the response to the external
field. In a metal, it is all the electrons that take part in screening an exter-
nal 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 ap-
proximate 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. This neglect produces simple algebraic coupled equations. The
inclusion of Umklapp scattering produces an infinite set of coupled equations
which has no known analytic solution.

9.4.1 The Thomas-Fermi Approximation


The Thomas-Fermi approximation is based on the assumption that the potential
is slowly varying. The energy of a Bloch state is given by
h̄2 k 2
− | e | φes (r) (791)
2m
The momentum of the highest occupied energy is r dependent kF (r) and is
given by
h̄2 kF2 (r)
− | e | φes (r) = µ (792)
2m
Thus, the electron density at position r is expressed in terms of a local Fermi-
wave vector
1 4 π kF3 (r)
ρ(r) = 2 (793)
8 π3 3
On expressing the Fermi-wave vector in terms of the chemical potential and the
electrostatic potential, the total density becomes
 3   32
1 2m 2
ρ(r) = µ + | e | φes (r) (794)
3 π2 h̄2

208
The induced density is given in terms of the electrostatic potential via
 3 "   32   23 #
1 2m 2
ρind (r) = µ + | e | φes (r) − µ (795)
3 π2 h̄2

This is the basis of the Thomas-Fermi Theory. On assuming that φes (r) is small
compared with µ, the equation can be linearized yielding
 
∂ρ0
ρind (r) = | e | φes (r) (796)
∂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 (797)
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 (798)
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
= (799)
π a0
and by the alternate expression
r
kT F 4
=
kF π kF a0
  13
16 1
= 2
rs2

1
= 0.8145 rs2 (800)

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.

209
This can be most clearly seen by applying 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) = (801)
q2
The total potential is given by

φext (q)
φes (q) =
ε(q)
4πZ |e|
= (802)
q 2 + kT2 F

which no longer possesses the long-ranged divergence as k → 0. On perform-


ing the inverse Fourier transform, thereby transforming the potential back into
direct space, one has
 
Z |e|
φes (r) = exp − kT F r (803)
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 (804)
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.

The Thomas-Fermi approximation is deficient. 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 (805)
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 molecules (N.L. Balazs, Phys. Rev. 156, 42
(1967)). The Thomas-Fermi method is deficient as it assumes that the poten-
tial 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

210
local approximation for the kinetic energy 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.

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 approximately linear in the perturba-
tion, so perturbation theory may be used. The effect of a perturbing potential
δV (r) on the electronic system is considered. The effect of this one-body po-
tential on the one-body 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
given by
X Mn0 ,k0 ;n,k
ψn,k (r) = φn,k (r) + φn0 ,k0 (r) (806)
En,k − En0 ,k0
n0 ,k0 6=n,k

where the Mn0 ,k0 ;n,k is the matrix element of the perturbing potential between
two Bloch functions,
Z
Mn ,k ;n,k =
0 0 d3 r0 φ∗n0 ,k0 (r0 ) δV (r0 ) φn,k (r0 ) (807)

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. (808)
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 response is not local in the perturbation
but is non-local. The response is expressed in the form
Z
ρind (r) = d3 r0 χ(r, r0 ) δV (r0 ) (809)

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. (810)
0 0
En,k − En0 ,k0
n,k,σ n ,k 6=n,k

where the summation over n, k, σ runs over all one-electron states that were oc-
cupied before the perturbation was turned on. Due to the Pauli exclusion prin-
ciple, the summation over n0 , k 0 is restricted to the unoccupied states. The ex-
pression for the response is expected to be modified by the presence of electron-
electron interactions.

211
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
"
3 3 0
Z Z  
0 4 m d k d k 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
(811)

where the integration over k 0 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 range of integration over
the magnitude of k 0 can be extended between − ∞ and + ∞ and evaluated by
means of contour integration which leads to
kF
sin 2 k | r − r0 |
Z
0 2m 2
χ(r, r ) = − 2 dk k
h̄ ( 2 π )3 0 | r − r0 |2
(812)

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
(813)

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
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 en-
ergy 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 oscillations.

It is more convenient to consider the Fourier transform of the response func-


tion Z  
χ(q) = d3 r χ(r) exp − iq.r (814)
V

212
The response function χ(q) is evaluated from
 
2m 1 X 1 1
χ(q) = 2 2 +
h̄ V k<k k 2 − ( k + q )2 k2 − ( k − q )2
F

(815)

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 dk k ln (816)
h̄ q 4 π 2 0 |q − 2k|

The response is given explicitly by

4 kF2 − q 2
  
m kF 1 | 2 kF + q |
χ(q) = − + ln (817)
h̄2 π 2 2 8 q kF | 2 kF − q |

This is the Lindhard function for the free electron gas (J. Lindhard, Kgl. Danske
Videnskab. Selskab. Mat. Fys. Medd. 28, 8 (1954)). 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 (818)
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
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.

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
functional theory. The relation between the induced density and the external
potential is given by the screened response function
Z
ρind (r) = − d3 r0 χs (r − r0 ) | e | φext (r0 ) (819)

213
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 )
(820)
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 ) (821)

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) (822)

and the non-interacting response function is given by

ρind (q) = − χ0 (q) | e | φef f (q) (823)

The relationship between the effective and external potential is given by


 
4π|e| π|e|
φef f (q) = φext (q) + − + Γ xc (q) ρind (q) (824)
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) =   (825)
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) (826)
ε(q) q2

The exchange contribution to Γxc (q) is given in the limit q → 0 by


  2  4 
5 q 73 q
Γxc (q) = 1 + + + ... (827)
9 2 kF 225 2 kF

214
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 + (828)
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 44
h̄2 k2
Calculate the Lindhard function for a free electron gas Ek0 = 2 m in d =
1, d = 2 and d = 3 dimensions, at zero temperature.

——————————————————————————————————

9.4.5 Exercise 45
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 (829)
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

215
per site.

——————————————————————————————————

216
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) + (830)
2m h̄2 k6=k0 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 (831)
V

which is just the average potential when k = k 0 . The sum over the energies
of all the occupied Bloch states, (k, σ), contribute 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. It can be evaluated
in the approximation that the Coulomb interactions between different Wigner-
Seitz cells are totally screened (E. Wigner and F. Seitz, Phys. Rev. 43, 804
(1933), Phys. Rev. 45, 509 (1934)). This means that the ion-ion interactions

217
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 (832)
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 by
3Z
ρ = 3 (833)
4 π RW S

For the uniform density, the electron-electron repulsion term is evaluated as


3 Z 2 e2
Z
Ees = d3 r Vions (r) ρ(r) + (834)
5 RW S
For the free-electron approximation for the kinetic energy to be valid, the elec-
trostatic 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) = − f or r ≥ Rc
r
= 0 f or r ≤ Rc (835)

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 − + (836)
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.

For a solid with structure, the electrostatic energy can be expressed as the
sum
E = EM + Ec (837)
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 (838)
RW S

218
where α is the structure-dependent Madelung constant. The Madelung con-
stants 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 = (839)
2 RW S RW S

and, as it is the electrostatic energy associated with the spherical pseudo-


potential core, 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 = (840)
h̄2 k6=k0 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 = (841)
h̄2 |k|<kF ,σ k6=k0
k 2 − k 02

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 (842)
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 ) (843)
i,j

219
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 3
= d r exp − i q . Ri Vj (r − Ri − rj ) exp − i q . ( r − Ri )
V i,j
(844)

This can be expressed as


X   X  
Vions (q) = exp − i q . Ri exp − i q . rj Vj (q)
i j
(845)

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 (846)

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) (847)
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 = (848)
V2
k<kF ,σ
h̄2 Q6=0 k 2 − ( k + Q )2

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 N X 2 X fk − fk+Q
= 2
| S(Q) |2 | V0 (Q) |2
2 V Ek − Ek+Q
Q6=0 k,σ
2
1 N X
= | S(Q) |2 | V0 (Q) |2 χ(Q) (849)
2 V
Q6=0

220
The summation over q is limited to the reciprocal lattice vectors Q. Therefore,
it depends on the lattice structure through 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 (850)
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 as 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) (851)
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) (852)

one can combine this with the contribution from the Bloch energies. The factor
4 π
q 2 χ(q) is related to the dielectric constant ε(q) through

4 π e2
ε(q) = 1 − χ(q) (853)
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) (854)
2 V
Q6=0

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

221
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
nodes at the wave vectors q0 = n 2Rcπ . The structural energy is composed of
negative contributions, but the contributions from the reciprocal lattice vectors
which are close to the nodes, contribute little to the stability of the structure.
In fact, reciprocal lattice vectors at the nodes would correspond to the special
case in which the band gap at the appropriate Brillouin zone boundary is 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 Q vector for (2, 0, 0) approaches q0 in In, there is a loss in structural


stability and the series undergoes a transition from the f.c.c. to a tetragonal
structure (V. Heine and D.L. Weaire, Solid State Physics, 24, 1 (1970)). When
this transition occurs, the set of equivalent f.c.c. reciprocal lattice vectors that
have qQ0 ∼ 1, 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
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.

10.2 Real Space Representation


The dominant electronic structural energy is given by a sum over all Q of

1 N2 X
Estructural = | S(Q) |2 | V0 (Q) |2 χ(Q) ε(Q) (855)
2 V
Q6=0

222
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  
N2 δq,Q = exp i q . ( R − R0 ) (856)
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) (857)
2V 0
q6=0 R6=R

where θ(q) is defined to be

1
θ(q) = | V0 (q) |2 χ(q) ε(q) (858)
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  
| S(q) |2 = exp i q . ( ri − rj ) (859)
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) (860)
2
i6=j q6=0

The Fourier transform of θ(q) is defined as


X  
θ(Ri,j ) = θ(q) exp i q . Ri,j (861)
q

where the vector Ri,j denotes the relative position of the two atoms. On chang-
ing the sum to an integration, θ(Ri,j ) is evaluated as
Z  
V 3
θ(Ri,j ) = d q θ(q) exp i q . R i,j
( 2 π )3
Z ∞ Z 1  
2πV 2
= dq q d cos θ θ(q) exp i q R i,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 sin q Ri,j
= dq q 2 θ(q) (862)
2 π2 0 q Ri,j

223
Thus, the electronic contribution to the structural energy has the real space
representation
1 X
Estructural = θ(Ri,j ) (863)
2
i6=j

The Madelung energy, which is the sum over the interaction energies of the ions

1 X Z 2 e2
EM adelung = (864)
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) (865)
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) (866)
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) (867)
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)
(868)

The integral in the first term 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

224
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) | (869)
R π 0 q ε(q)

The long-ranged nature of the Coulomb interaction between the bare ions has
been completely eliminated due to the screening. The very weak logarithmic
singularity at q = 2 kF leads to Friedel oscillations in the potential at asymp-
totically large distances R
cos 2 kF R
Θ(R) = A (870)
R3
However, at intermediate distances, the pair potential can be approximately
expressed as the sum of three (damped) oscillatory terms (D.G. Pettifor and
M.A. Ward, Solid. State. Commun. 49, 291 (1984))
3
Z 2 e2 X
   
Θ(R) = Bn cos αn 2 kF R + φn exp − βn kF R (871)
R n=1

where the phase shift depends on the ionic core radius Rc and the electron den-
sity. This form is obtained as a result of approximating the Lindhard function
χ(q) by a ratio of polynomials (Padé approximation). The integration over q
can be performed via contour integration. The pairs of complex poles in the
integrand produce the terms which have a damped oscillatory dependence 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

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

225
Al
n 1 2 3
αn 0.156 0.644 0.958
βn 0.793 0.698 0.279
Bn 7.954 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 po-
tential, although it only has a magnitude of about 10−2 eV, dominates the
structural energy.

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

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 stability of the h.c.p. and f.c.c structures of N a, M g and
Al (A.K. McMahan and J.R. Moriarty, Phys. Rev. B 27, 3235 (1983)). At
ambient pressure, N a and M g are h.c.p. and Al is f.c.c.. 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 near-
est 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 interaction decreases. This means that the oscillations in the pair
potential move out to larger distances. This causes the changes in the pair po-
tential 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
increasing Rrsc , which corresponds to increasing pressure. The change in the

226
phase shifts moves the oscillations in the pair potentials to distances larger dis-
tances than the neighbor distances. This shows that as the pressure is increased,
one may expect 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 neighbors of the close-packed structures. The separations
1
of the neighbors in the b.c.c. structure should be scaled by a factor of 2− 3 to
yield the same electron density as the close-packed structures. After this scal-
ing, 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 ex-
pect 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 ex-
perience 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 structures, which are hard to extract from other methods, as the struc-
tural 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. To obtain a more accurate
description of structural stability, it is necessary to utilize density functional
calculations.

227
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 (872)


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 () =   (873)
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 (874)
−∞

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 Sommerfeld (A. Sommerfeld, Zeit. für Physik,
47, 1 (1928)). The low-temperature Sommerfeld expansion of the electronic
specific heat, for non-interacting electrons, shall be examined.

228
11.1.1 The Sommerfeld Expansion
The total energy of the solid can be expressed as an integral
Z +∞
E = 2 d ρ()  f () (875)
−∞

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 () (876)
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 (877)
The Fermi-function becomes
1
f (µ + kB T x) = (878)
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
(879)
The integral over the negative range of x is re-expressed in terms of the new
variable y where
y = −x (880)
Thus, the energy is expressed as
Z µ
E = 2 d ρ() 
−∞
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
(881)

229
However, the Fermi-function satisfies the relation

1 − f ( µ − kB T y ) = f ( µ + kB T y ) (882)

or equivalently
1 1
1 − = (883)
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 )

(884)

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
(885)

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! (886)
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) = (887)
lm
l=1

Using this, one finds that



( − 1 )l+1
 2n+1 
X 2 − 1
= ζ(2n + 2) (888)
l2n+2 22n+1
l=1

230
The Riemann zeta functions have special values

π2
ζ(2) =
6
π4
ζ(4) = (889)
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)
X 2 − 1 ∂
+ 4 ( kB T )2 ζ(2n + 2) ( kB T )2n µ ρ(µ)
n=0
22n+1 ∂µ
(890)

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 () (891)
−∞

which also can be expanded in powers of T 2 as


Z µ
Ne = 2 d ρ()
−∞
∞  2n+1
(  " #)
 (2n+1)
2
X 2 − 1 2n ∂
+ 4 ( kB T ) ζ(2n + 2) ( kB T ) ρ(µ)
n=0
22n+1 ∂µ
(892)

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
 
∂E
= (893)
∂T Ne

231
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 ∂
CNe = 4 kB T (n + 1) 2n
ζ(2n + 2) ( kB T )2n µ ρ(µ)
n=0
2 ∂µ
  "
∂µ
+ 2 µ ρ(µ)
∂T Ne
∞  2n+1
(  ) #
 (2n+2) 
2 2
X 2 − 1 ∂
+ 4 kB T 2n+1
ζ(2n + 2) (kB T )2n µ ρ(µ)
n=0
2 ∂µ
(894)

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 ∂
0 = 4 kB T (n + 1) 2n
ζ(2n + 2) ( kB T )2n ρ(µ)
n=0
2 ∂µ
  "
∂µ
+ 2 ρ(µ)
∂T Ne
∞  2n+1
(  )#
 (2n+2)
2 2
X 2 − 1 ∂
+ 4 kB T 2n+1
ζ(2n + 2) ( kB T )2n ρ(µ)
n=0
2 ∂µ
(895)

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 2π 4
= kB T ρ(µ) + O(kB T 3) (896)
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
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

232
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) (897)
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) (898)

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) )
−∞
(899)

Thus, in this approximation, the coefficient of the linear T term is given by


CN
γ = lim
T → 0 T
 
∂S
=
∂T Ne
2 2 π2
= kB Z ρ(µ) (900)
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 like
CeCu6 , CeCu2 Si2 , CeAl3 and U Be13 , the value of the γ coefficients are ex-
tremely large, of the order of 1 J / mole of f ion / K2 , which is 1000 times larger

233
than Cu. The quasi-particle mass enhancements are inferred by comparison to
L.D.A. electronic density of states calculations and are about 10 to 30. The
enhancement is assumed to be due to the strong electron-electron interactions,
which the L.D.A. fails to take into account.

——————————————————————————————————

11.1.3 Exercise 46
Calculate the next to leading order term in the low-temperature electronic-
specific heat.

——————————————————————————————————

11.1.4 Exercise 47
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  f or  > 0
ρ() = − α1  f or  < 0 (901)
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.

The magnetization due to the electronic spins can be calculated from


 
∂Ω
Mz = − (902)
∂Hz
where the grand canonical potential is given by
X  !
Ω = − kB T ln 1 + exp − β ( Eα − µ ) (903)
α

234
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 = − H z Sz
2 me c
= − µB Hz σ (904)
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 = (905)
2 me c
The energy of a particle can then be written as
Eσ (k) = E(k) − µB σ Hz (906)
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(  − E(k) ) (907)
k

Thus, in the presence of a field, one has the spin dependent density of states
ρσ () = ρ( + µB σHz ) (908)
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 dE ρ(E) ln 1 + exp − β ( E − µB σHz − µ )
−∞ σ
(909)
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 dE σ ρ(E)  
−∞ σ 1 + exp β ( E − µB σHz − µ )
Z ∞ X
= µB dE σ ρ(E) f ( E − µB σHz )
−∞ σ
 
= µB Ne (σ = 1) − Ne (σ = − 1) (910)

235
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 ) = (911)
∂Hz

and is given by
Z ∞ X ∂
χp (T, Hz ) = − µ2B dE σ 2 ρ(E) f ( E − µB σHz )
−∞ σ
∂E
(912)

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( E − F ) (913)
∂E
so that the zero temperature value of the Pauli-susceptibility is evaluated as

χp (0) = 2 µ2B ρ(F ) (914)

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 susceptibil-
ity can be evaluated by integration by parts, to obtain
Z ∞
2 ∂
χp (T ) = 2 µB dE f (E) ρ(E) (915)
−∞ ∂E

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
(916)
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 ) (917)
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

236
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 ) 2 
" #
2 ( )
 2
π ∂ ρ(F ) ∂
χp (T ) = 2 µ2B ρ(F ) + k2 T 2 − F 4 4
+ O( kB T )
6 B ∂2F ρ(F )
(918)
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 = (919)
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 (920)
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. 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 elec-
trons.

——————————————————————————————————

11.1.6 Exercise 48
Determine the field dependence of the low-temperature Pauli susceptibility.

——————————————————————————————————

11.1.7 Exercise 49
Determine the high temperature form of the Pauli susceptibility.

——————————————————————————————————

237
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 (921)
2m 2
where
| e | Hz
ωc = (922)
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 (923)
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 = (924)
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 r = h̄ ωc (925)
2 c c 2
Thus, the degeneracy is given by

L2
D = m ωc
2 π h̄
| e | L2
D = Hz (926)
hc

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
(927)
h̄ ωc Hz
m c

The numerical constant has the value


| e | h̄
∼ 1.16 × 10−8 eV / G (928)
mc

238
so, with µ ∼ 1 eV and Hz ∼ 104 G, one finds that the number of occupied
Landau levels is approximately given by
µ
∼ 104 (929)
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) (930)
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
 
| e | Hz ∂φ
− ∇2 φ − i x + x φ = Eφ (931)
2m mc ∂y 2 m c2
This can be solved by the substitution
 
φ(r) = f (x) exp i ( ky y + kz z ) (932)

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
(933)
which is recognized as the equation for the harmonic oscillator with energy
eigenvalue
h̄2 kz2
 
1
E − = n + h̄ ωc (934)
2m 2
where
e2 Hz2
ωc2 = (935)
m2 c2
That is, the motion in the plane perpendicular to the field, Hz , is quantized
into Landau levels (L.D. Landau, Zeit. für Physik, 64, 629 (1930)). The energy
spacing between the levels is given by h̄ ωc , where
| e | Hz
ωc = (936)
mc
and the orbit is centered around the position
h̄ ky c
x0 = − (937)
| e | Hz

239
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 (938)

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 (939)
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) (940)
The periodic boundary conditions imply that
 
exp i ky Ly = 1 (941)

or

ky = ny (942)
Ly

The x dependent factor of the wave function f (x) is centered at x0 where

h̄ ky c
x0 = − (943)
| e | Hz

For a sample of width Lx , one must have Lx > x0 > 0, so one has the equality

| e | Hz Lx
> − ky > 0 (944)
h̄ c
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
= (945)
2 π h̄ c
independent of n. Thus, the degeneracy D of every Landau is given by

| e | Hz Lx Ly
D = (946)
2 π h̄ c

240
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 = (947)
m ωc
The degeneracy of the Landau levels can also be expressed as
Lx Ly
D = (948)
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
Rauh (A. Rauh, Phys. Stat. Solidi, B 65, K131 (1974), A. Rauh, Phys. Stat.
Solidi, B 69, K9 (1975)) and by Harper (P.G. Harper, Ph.D. Thesis, Univer-
sity of Birmingham (1954), P.G. Harper, Proc. Phys. Soc. London, A 68, 874
(1955)). These authors have shown that the periodic potential causes the Lan-
dau 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
(949)
On integrating over kz , one finds
µ
 12 −1
h̄ωc  32
8 D Lz

2m X2  1
Ω = − µ − h̄ ωc ( n + ) (950)
3 2π h̄2 n=0
2

or
µ
 12 −1
h̄ωc  32
2 V

2m X2  1
Ω = − m ωc µ − h̄ ωc ( n + ) (951)
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
(952)

241
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 µ2 − h̄ ωc µ 2 + . . .
3 ( π 2 h̄2 ) h̄2 5 16
(953)
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 (954)
3 π 2 h̄2 h̄2
where we have expressed the orbital magnetic moment in terms of the (orbital)
Bohr magneton
| e | h̄
µB = (955)
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
V m 2m 2 1
= µ2B 2 2 µ2 (956)
π h̄ h̄2
Hence, the spin and orbital susceptibilities are related via
1
χd = − χp (957)
3
Thus, the Landau diamagnetic susceptibility is negative and has a magnitude
which, for free electrons, is just one third of the Pauli paramagnetic susceptibility
(L.D. Landau, Zeit. für Physik, 64, 629 (1930)). The diamagnetism results from
the quantized orbital angular momentum 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
= − (958)
χp 3 m∗
and the diamagnetic susceptibility can be larger than the Pauli susceptibility.
The susceptibility of Bismuth is negative.

242
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 factor.

243
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 (959)

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 ) (960)
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

244
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
(961)

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 }) (962)
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 ) )
0

0 == 2
(963)
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 (964)

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
(965)

245
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)

(966)
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
(967)
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
(968)
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 (969)

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 (970)

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 =

246
    
h̄ k |e|E
= − .∇ f (k, r, t) − . ∇k f (k, r, t) d3 k d3 r dt
m m h̄
(971)

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 (972)
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
(973)
 
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 )
(974)

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) (975)
dt
Since
1 1
= (976)
τ (k → k 0 ) τ (k 0 → k)

247
the collision integral can be simplified to yield
  "  #
X 1
I f (k, r, t) = f (k 0 , r, t) − f (k, r, t) (977)
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) (978)
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) (979)
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) (980)
∂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.

248
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)
(981)

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 ) (982)
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 (983)

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 ) (984)
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 ) (985)

on integrating over the azimuthal angle φ0

dΩ0 = dθ0 sin θ0 dφ0 (986)

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 (987)
h̄ −1

249
This yields the result
Z 1
2m
= 4π k(k.E)2 d cos θ0 cos3 θ0 | V ( 2 k cos θ0 ) |2 (988)
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 (989)
h̄ −1 2

On identifying the non-equilibrium part of the distribution function with

∂f0 (k) ∂f0 (k)


Φ(k) = ( k . E ) A(E) (990)
∂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) (991)

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 (992)
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
(993)

250
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-volume
which has been displaced from the equilibrium position in the direction of the
applied field. The displacement of the Fermi-volume produces the average cur-
rent 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 (994)
h̄ ∂E(k)
k

On recognizing the zero temperature property of the Fermi-function


 
∂f0 (k)
− = δ( E(k) − µ ) (995)
∂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 integral
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) | (996)
3 h̄2 k ∂E(k)

For free electron bands, the conductivity tensor is evaluated as

ρ e2 τtr
σα,β = δα,β (997)
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 50
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

251
field with a Fourier amplitude Eβ (q, ω) via Ohm’s law
X
jα (q, ω) = σα,β (q, ω) Eβ (q, ω) (998)
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 (999)

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
(1000)

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

252
is embodied in the definition of the Hall constant, RH
Ey
RH = (1001)
jx Hz
which for semi-classical free carriers of charge q and density ρ is evaluated as
1
RH = (1002)
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) (1003)

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) (1004)

The substitution of the zero field equilibrium distribution function f0 (p) in the
term proportional to f0 (p) without any magnetic field corrections is consistent
with the equilibrium 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) (1005)

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)
(1006)

253
since due to the vector identity

(A ∧ B).A = 0 (1007)

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 (1008)
∂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 = (1009)
m∗
one finds
|e| 1
|e|E.v + (v ∧ B).C = (v.C ) (1010)
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 (1011)

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 (1012)
m∗ c τtr

To solve the above algebraic equation it is convenient to change variables


 
|e|
ωc = B (1013)
m∗ c

This shall be solved by finding the components parallel and transverse to B.

254
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 (1014)

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 (1015)
τtr c
but
ω c ∧ ( ω c ∧ C ) = ω c ( ω c . C ) − ωc2 C (1016)
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 (1017)
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 (1018)
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 (1019)
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 ) (1020)

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)
(1021)

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

255
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) (1022)
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

(1023)

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 )

(1024)

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
(1025)
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 (1026)
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,

256
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 (1027)
m
The non-zero off-diagonal terms are found as
ρ e2 ωc τ 2
σy,x = − σx,y = (1028)
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 (1029)
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 (1030)
2 2 1 + i ωc τ
and the modulus is just given by
σ0 σ0
|z − Ex | = Ex (1031)
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 = (1032)
Jx
and from the Boltzmann equation analysis of the magneto-conductivity
ΨH = tan−1 ωc τ (1033)

257
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 (1034)
j

to obtain X
Ei = ρi,j Jj (1035)
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 = (1036)
ρ|e|
Thus, the Hall resistivity is
Ey
ρyx =
Jx
Hz
= (1037)
ρ|e|
Thus, the Hall constant RH is given by
EY 1
RH = = (1038)
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 afield 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:-

(i) for low fields, Hz such that ωc τ < 1 then


∆ρx,x = ρx,x (Hz ) − ρx,x (0) ∝ Hz2 (1039)

258
(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 (1040)
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 (1041)

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
The five unknown quantities may be determined by five experiments.

(1) When J and H are parallel to the x axis one has the longitudinal
magneto-resistance given by

ρx,x = ρ0 + ( b + c + d ) H 2 (1042)

(2) When J k x , H k y then

ρx,x = ρ0 + b H 2 (1043)

which is the transverse magneto-resistance.


(3) With J k (1, 1, 0) i.e.
J
J = √ (1, 1, 0) (1044)
2
and
H
H = √ (1, 1, 0) (1045)
2

259
one has a different longitudinal magneto-resistance
d
∆ρ = ( b + c + ) H2 (1046)
2

(4) When H = H (0, 0, 1) then a second transverse magneto-resistance is found


as
∆ ρ = b H2 (1047)
H
but when H = √
2
(1, −1, 0) then the magneto-resistance is found as

d
∆ρ = (b + ) H2 (1048)
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 metals.

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 ) (1049)
2
and  
σ+ − σ −
Jy = Ex sin 2 ΨH (1050)
2
Thus,

σx,x = σ0 cos2 ΨH
 
ρ+ − ρ−
σy,x = σ0 sin ΨH cos ΨH
ρ+ + ρ−
(1051)

260
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 (1052)
σ0

ρ+ − ρ−
1 + ρ+ + ρ− ωc2 τ2

since tan Ψh = ωc τ .

Furthermore, as
sec2 ΨH = 1 + tan2 ΨH
= 1 + ωc2 τ 2 (1053)
then
1 ( 1 + ωc2 τ 2 )
ρx,x = 2 (1054)
σ0

− ρ−
1 + ρρ++ + 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
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 = (1055)
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

261
Jy
sin θ = (1056)
J
Thus
E sin θ E Jy
RH = = 2
H J  J
H 
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

(1057)

The Hall coefficient saturates to


1
RH → (1058)
| e | ( ρ+ − ρ− )

for large magnetic fields.

262
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) = − (1059)
c ∂t
where E and B represent the microscopic electric and magnetic fields, and ρe and
j are the microscopic charge and current densities. These 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 (1060)

The solutions for the potentials are not unique, as the gauge transformations

A → A0 = A + ∇ Λ (1061)

and
1 ∂Λ
φ → φ0 = φ − (1062)
c ∂t
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
 
1 ∂
∇2 φ + ∇.A = − 4 π ρe (1063)
c ∂t

and
1 ∂2A
 
1 ∂φ 4π
∇2 A − − ∇ ∇.A + = − j (1064)
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 (1065)

263
or the Lorentz Gauge
1 ∂φ
∇.A + = 0 (1066)
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 out the effect 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) (1067)
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
(1068)

In the wave-vector and frequency domain, the Coulomb gauge condition is ex-
pressed as
q . A(q, ω) = 0 (1069)
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
(1070)

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) (1071)

and a transverse term


 
j T (q, ω) = j(q, ω) − q̂ q̂ . j(q, ω) (1072)

Thus, the second non-trivial Maxwell equation separates into the transverse
equation

ω2
 
2 4π
− q + 2 A(q, ω) = − j (q, ω) (1073)
c c T

264
and the longitudinal equation
ω 4π
−q φ(q, ω) = − j (q, ω) (1074)
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) (1075)
This is just Poisson’s equation, and it has the solution

φ(q, ω) = ρe (q, ω) (1076)
q2
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. Although
this is an instantaneous relation, the signals transmitted by the electromagnetic
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 equation, as can
be seen by examining the continuity equation which expresses conservation of
charge
∂ρe
+ ∇.j = 0 (1077)
∂t
The continuity equation can be Fourier transformed to yield
− ω ρe (q, ω) + q . j(q, ω) = 0 (1078)
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) (1079)
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) (1080)
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) (1081)

so the longitudinal current can be viewed as being produced either by the charge
density or by the scalar potential.

265
11.3.1 The Longitudinal Response
The currents and charge densities are usually broken down into the external
contributions and the induced contribution, via

j(q, ω) = j ind (q, ω) + j ext (q, ω)


ρe (q, ω) = ρe ind (q, ω) + ρe ext (q, ω) (1082)

The external scalar potential is given in terms of the external charge density via
Poisson’s equation

− q 2 φext (q, ω) = − 4 π ρe ext (q, ω) (1083)

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) = (1084)
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 combining the relation

j L (q, ω) = φ(q, ω) (1085)

with the expression for the induced component of the longitudinal current

 
j L (q, ω)ind = φ(q, ω) − φ(q, ω)ext (1086)

Hence, on using the definition of the frequency dependent dielectric constant,
one obtains

 
j L (q, ω)ind = 1 − ε(q, ω) φ(q, ω) (1087)

The total scalar potential φ(q, ω) can be related to the 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) (1088)
c
If the longitudinal part of the electric field is identified as

E L (q, ω) = − i q φ(q, ω) (1089)

then one obtains the relation between the longitudinal current and the longitu-
dinal electric field
 

j L (q, ω)ind = 1 − ε(q, ω) E L (q, ω) (1090)

266
Hence, as the longitudinal conductivity σL is defined by the relation

j L (q, ω)ind = σL (q, ω) E L (q, ω)


(1091)

one finds that the conductivity and the dielectric constant are related through
 

σL (q, ω) = 1 − ε(q, ω) (1092)

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 − (1093)
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) (1094)
and the scalar potential acting on the electrons produces the potential δV (q, ω)
where
δV (q, ω) = − | e | φ(q, ω) (1095)
Thus, the frequency dependent dielectric constant may be written as

4 π e2 ρind (q, ω)
ε(q, ω) = 1 −
q2 δV (q, ω)
4 π e2
= 1 − χ(q, ω) (1096)
q2

(H. Ehrenreich and M.H. Cohen, Phys. Rev. 115, 786 (1959)) 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) = (1097)
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) (1098)

267
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 ) (1099)
−∞

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 (1100)
i

and the time dependent perturbation is


Z
Ĥint (t) = d3 r0 ρ̂(r0 , t) δV (r0 , t) (1101)

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) > (1102)
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 (1103)
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 > (1104)
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 >
(1105)

268
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 ) (1106)
−∞

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

(1107)
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 (1108)
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 )

(1109)

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
(1110)

269
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 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 (1111)

and
αj = βj ∀ j 6= i (1112)
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) (1113)

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 unoccupied
in the final state | Ψn > and 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 (1114)

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
(1115)

270
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 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

(1116)

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) (1117)
−∞

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 δ. The factor of exp [ − δ t ] damps out the oscillations
in the integrand as t → ∞. Thus, one finds that in the (q, ω) domain the
response function is complex and is given by the expression
" #
2 X 1
χ(q, ω + iδ) =
V h̄ ω + i δ + Ek − Ek+q
|k|<kF |k+q|>kF
" #
2 X 1

V h̄ ω + i δ + Ek − Ek+q
|k|>kF |k+q|<kF

(1118)

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 f or Ek < EF (1119)

and
fk = 0 f or Ek > EF (1120)

271
The response function can then be written as the sum over all k as
" #
2 X fk ( 1 − fk+q )
χ(q, ω + iδ) =
V h̄ ω + i δ + Ek − Ek+q
k
" #
2 X fk+q ( 1 − fk )

V h̄ ω + i δ + Ek − Ek+q
k
" #
2 X fk − fk+q
= (1121)
V h̄ ω + i δ + 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 δ → 0,
the imaginary part of the response function is found as
 
2π X
Im χ(q, ω + iδ) = − δ h̄ ω + Ek − Ek+q
V
|k|<kF
 
2π X
+ δ h̄ ω + Ek − Ek+q
V
|k+q|<kF

(1122)
From this analysis, one can see that for positive ω the imaginary part of χ(q, ω)
is non-zero in for the region of (ω, q) phase space, where
h̄ h̄
( − 2 kF q + q 2 ) < ω < ( + 2 kF q + q 2 ) (1123)
2m 2m
It is only in this region that the argument of the first delta function in Im χ(q, ω)
has a solution
h̄ h̄
k.q = ω − q2 (1124)
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 given (J. Lindhard,
Kgl. Danske Videnskab. Selskab. Mat. Fys. Medd. 28, 8 (1954)). The real
part is given by
" (
kT2 F (2 m ω − h̄ q 2 )2
  
kF
Re 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

(1125)

272
and the imaginary part is given by
π kT2 F m ω
 
Im ε(q, ω + iδ) = 2 m ω < 2 h̄ q kF − h̄ q 2
2 q 2 h̄ q kF
(1126)
" #
π kT2 F kF (2 m ω − h̄ q 2 )2
 
Im ε(q, ω + iδ) = 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 (1127)
and
 
Im ε(q, ω + iδ) = 0 2 h̄ q kF + h̄ q 2 < 2 m ω (1128)

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, as 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
Re χ(q, ω) = + 1 + + ... (1129)
3 π2 m ω2 5 mω

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 + + ... (1130)
ω 5 mω

where the expression for the electron density ρ


kF3
ρ = 2 (1131)
6 π2
has been used, and the plasmon frequency ωp has been defined via
4 π ρ e2
ωp2 = (1132)
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 + + ... (1133)
5 m ωp (q)

273
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 (1134)
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 enquire as to the nature of the excitations at larger q values, such
that the phase velocity of the plasmons becomes greater than the Fermi-velocity
vF = h̄ mkF . 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 (1135)
2
for k < kF . The edges of the continuum stretch from 2h̄m ( 2 kF q + q 2 )
2
to 2h̄m ( − 2 kF q + q 2 ). When the plasmon merges into the continuum it
undergoes significant broadening. This sort of damping is called Landau damp-
ing. Landau damping can also be viewed classically, in terms of electrons surf
riding the waves in the potential field. Imagine a wave with phase velocity ωq is
propagating through an electron gas, and consider the electrons with velocity
is 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 scat-
tering of a beam of charged particles or fast electrons. The coupling takes place
via the Coulomb interaction
X e2
Ĥint = (1136)
i
| r − ri |

274
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


with energy E(k) to state k 0 with energy E(k 0 ) is given by
2  
1 2 π
< k 0 Ψn | Ĥint | k Ψ0 > δ En + E(k 0 ) − E0 − E(k)

=
τ (k → k 0 ) h̄
(1137)
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 ) (1138)

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

(1139)

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
(1140)

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 ∞
4 π e2

1 2π dt X
0 = exp[ i ω t ] < Ψn | exp[ i q . ri ] | Ψ0 > ×
τ (k → k ) h̄2 q2 V −∞ 2π i

275
X Ĥ0 t Ĥ0 t
< Ψ0 | exp[ i ] exp[ − i q . ri ] exp[ − i ] | Ψn >
i
h̄ h̄
(1141)
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 (1142)
Thus, on combining the above expressions, the inelastic scattering rate is found
as
2 Z ∞
2 π 4 π e2

1 dt
= exp[ i ω t ] < Ψn | ρ̂−q | Ψ0 > ×
τ (k → k 0 ) h̄ 2 q 2
−∞ 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
(1143)
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 >


(1144)
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
0 = 2 2
exp[ i ω t ] < Ψ0 | ρ̂q (t) ρ̂−q (0) | Ψ0 >
τ (k → k ) h̄ q −∞ 2 π
(1145)
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 ∞
dt
S(q, ω) = V 2 exp[ i ω t ] < Ψ0 | ρ̂q (t) ρ̂−q (0) | Ψ0 > (1146)
−∞ 2 π h̄

276
On substituting this relation into the scattering rate we obtain the result
2
4 π e2

1 2π
= S(q, ω) (1147)
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 (1148)
|k|<kF |k+q|>kF

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) (1149)
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 scattering (L. van Hove, Phys. Rev. 95, 249 (1954)).

The Fluctuation-Dissipation Theorem (H.B. Callen and T.A. Welton, Phys.


Rev. 83, 34 (1951)) 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) = Im (1150)
4 π 2 e2 exp[ − β h̄ ω ] − 1 ε(q, ω + iδ)

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. (1151)
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) = − ∇ − (1152)
| r − r(t) |

277
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) = (1153)
ε(q, ω + iδ)

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) = − Im D(q, ω)
8π ε(q, ω + iδ)
" # 2
ω Im ε(q, ω + iδ)
= 2
D(q, ω)
8π | ε(q, ω + iδ) |
" # 2
ω Im ε(q, ω + iδ)
= D(q, ω)
8π ( Re ε(q, ω) )2 + ( Im ε(q, ω + iδ) )2

(1154)

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 51
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 theorem (P. Noziêres and D. Pines, Nuovo Cimento, 9,
470 (1958)).

——————————————————————————————————

Solution

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) (1155)

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) (1156)

278
The dielectric constant is defined as
1 φ(q, ω)
= (1157)
ε(q, ω) φext (q, ω)

which can be expressed as

1 ρind (q, ω)
= 1 + (1158)
ε(q, ω) ρext (q, ω)

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) (1159)
ε(q, ω) 4 π e2

However, the interaction operator is given by


Z
Ĥint (r, t) = d3 r ρ̂(r) δV (r, t) (1160)

where the electron density is given by


X
ρ̂(r) = δ 3 (r − ri ) (1161)
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 
0 0
| Φ0 (t) > = 1 − dt Ĥint (t ) | Ψ0 > (1162)
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
= − dt0 d3 r0 < Ψ0 | [ ρ̂(r, t) , ρ̂(r0 , t0 ) ] | Ψ0 > δV (r0 , t0 )
h̄ −∞
Z ∞ Z
= dt0 d3 r0 χ(r − r0 , t − t0 ) δV (r0 , t0 ) (1163)
−∞

279
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) (1164)

where
X  | < Ψ0 | ρ̂q | Ψn > |2
χ(q, ω) = V
n
h̄ ω + i δ + En − E0
| < Ψ0 | ρ̂q | Ψn > |2

+ (1165)
− h̄ ω − i δ + En − E0
The imaginary part of the response function is found as
X  
2
Im χ(q, ω) = − π V | < Ψ0 | ρ̂q | Ψn > | δ( h̄ ω + En − E0 ) − δ( h̄ ω − En + E0 )
n
(1166)
Thus, the zero temperature limit of the fluctuation-dissipation theorem has the
form
4 π 2 e2 V X
 
1 2
Im = | < 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) (1167)
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 52
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) = − Im ln (1168)
πv ε(ω + iδ) ω
Assume that the dielectric constant is independent of q, for q < q0 where h̄ q0
is the maximum momentum transfer.

——————————————————————————————————

Solution 52

280
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 3 t
= dt d r P (r, t) exp −
τ 0 τ
Z ∞ Z  
1 3 ∂ t
= dt d r E(r, t) D(r, t) exp −
4πτ 0 ∂t τ
(1169)
where we have inserted an exponential convergence factor. The convergence
factor will be absorbed in the displacement and electric fields. The Fourier
transform is expressed as
Z Z ∞  
1
D(q, ω) = d3 r dt D(r, t) exp − i ( q . r − ω t ) (1170)
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 ) (1171)
( 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(1172)
2 τ ( 2 π ) −∞ 2 π ( 2 π )3 ε(q, ω + 2τi )
in the limit τ → ∞. The Fourier transform of the displacement field is given
by
Z ∞  
|e|
Z
1 3 t
D(q, ω) = d r dt exp − i ( q . r − ω t ) − ∇
V 0 2 τ | r − vt|
Z ∞
4πiq|e|
 
t
= − 2
dt exp i ( ω − q . v ) t −
V q 0 2τ
4πiq|e| 2τ
= − (1173)
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 −∞ q 2 ε(q, ω + 2τ ) 4 τ 2
+ ( ω − q . v )2
(1174)

281
which in the limit τ → ∞ reduces to
Z ∞
2 e2 d3 q
Z

P = 2
dω 2
δ( ω − q . v ) (1175)
(2π) −∞ q ε(q, ω + iδ)

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 + iδ)
(1176)
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 (1177)
π v −∞ ε(ω + iδ) ω

The power loss spectrum, P (ω), is defined in terms of an integral over positive
frequencies Z ∞
P = dω P (ω) (1178)
0
On using the symmetry properties of the dielectric constant under the transfor-
mation ω → − ω, one finds that the contribution from real part of the inverse
dielectric constant vanishes. Hence, one obtains the final result for P (ω)

2 e2
 
ω q0 v
P (ω) = − Im ln (1179)
πv ε(ω + iδ) 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 analyzed (L. Marton, J.A. Simpson, H.A. Fowler and N.
Swanson, Phys. Rev. 126, 182 (1962)). Experimentally, it is found that the fast
electron 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 (1180)
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

282
theoretically derived inelastic scattering cross-section.

The mean free path can be estimated from the scattering cross-section. One
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 Im (1181)
dΩ dω k π h̄2 q ε(q, ω + iδ)

For the frequency range

2 m ω > 2 h̄ q kF + q 2 (1182)

the imaginary part of χ(q, ω + iδ) vanishes as δ → 0. Therefore, in this limit,


the imaginary part of the dielectric constant also vanishes

Im ε(q, ω + iδ) → κ δ (1183)

for some value finite of κ. Hence, in this limit one has


" #
1
Im = − π δ( Re ε(q, ω) ) (1184)
ε(q, ω + iδ)

Furthermore, in this region of (ω, q) space one has the approximate expression

ωp2 (q)
ε(q, ω) = 1 − (1185)
ω2
where the plasmon dispersion relation is given by
 
3 2 2
ωp2 (q) = ωp2 + q vF + . . . (1186)
5

and the plasmon frequency by

4 π ρ e2
ωp2 = (1187)
m
Thus, one finds the single (plasmon) pole approximation for the inverse dielectric
constant
" #  2
ω − ωp2 (q)

1
Im = −πδ
ε(q, ω + iδ) ω2
ω2
= −π δ( ω − ωp (q) )
ω + ωp (q)
π
= − ωp (q) δ( ω − ωp (q) ) (1188)
2

283
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
k 0 2 m2q ρ V e2


= (1189)
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 (1190)
2
and
2 mq
( k − k0 ) ( k + k0 ) = ωp

2 mq ω p
( k − k0 ) = (1191)
h̄ ( k + k 0 )

Foe small scattering angles, θ  1, these can be combined to yield

h̄2 ωp2
 
θ
q2 ≈ k2 4 sin2 +
2 4 E(k)2
 
≈ k2 θ2 + θ02 (1192)

Hence, one has

dσ k 0 2 m2q ρ V e4
≈ 2
dΩ k m h̄ ωp h̄ k ( θ2 + θ02 )
2

mq e4 ρ V
≈ (1193)
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 (1194)
h̄ ωp E(k) m θ0

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, which must

284
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 (1195)
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) (1196)
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) (1197)
c c T ind
Ohm’s law can be expressed in the form

jT ind
(q, ω) = σT (q, ω) ET (q, ω) (1198)

where σT is the transverse conductivity and the total transverse electric field is
given by
ω
ET (q, ω) = i A(q, ω) (1199)
c
This leads to
ω2
 
4πω
− q2 + 2 + i σ T (q, ω) A(q, ω) = 0 (1200)
c c2

The transverse dielectric function is identified in terms of the optical conduc-


tivity
4πi
εT (q, ω) = 1 + σT (q, ω) (1201)
ω
The photon dispersion relation can be re-written as
 2
ω
εT (q, ω) = q 2 (1202)
c

If εT (q, ω) > 0, then there are undamped transverse electromagnetic waves.


Otherwise, q would be complex which implies that E T (r; t) is attenuated as it
enters into the sample. In other words, if Im ε(q, ω) = 0 and Re ε(q, ω) > 0,

285
the material is transparent to transverse electromagnetic waves. The dispersion
relation is given by
 2
cq
εT (q, ω) = (1203)
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) = (1204)
m 1 − iωτ
which leads to the dielectric constant being given by

4 π e2 ρ 1
ε(0, ω) = 1 − (1205)
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 (1206)

which is given by

ω 2 − ωp2 = c2 q 2 (1207)

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 length will be reflected from the
metal. Hence, as εL (0, ω) = εT (0, ω) which implies that optical experiments
that measure εT (q, ω) produce similar information to characteristic 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
(1208)

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)

286
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) (1209)

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 ) (1210)
2m i

and the diamagnetic current is given by

| e |2 X 3
j d (r) = − δ (r − ri ) A(ri ) (1211)
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) (1212)
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 ) (1213)
−∞ c
β

where the paramagnetic response function is given by the ground state expec-
tation value
 
α,β i
Rj,j (r, r0 , t − t0 ) = + < Ψ0 | jpα (r, t) , jpβ (r0 , t0 ) | Ψ0 > Θ( t − t0 )

(1214)

This is known as the Kubo formula for the conductivity (R. Kubo, J. Phys. Soc.
Jpn. 12, 570 (1957)). 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 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) (1215)
c j,j
β

287
Hence, to linear order in the vector potential, the total transverse current is
given by
" #
α
X 1 α,β | e |2
jind (q, ω) = R (q, ω) − δα,β ρ0 Aβ (q, ω) (1216)
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) (1217)
c
as " #
1 | e |2
σTα,β (q, ω) = α,β
Rj,j (q, ω) − δα,β ρ0 (1218)
iω m
The conductivity should be evaluated using a microscopic theory, and has a real
part and an imaginary part that are connected by causality. The conductivity
determines the material’s properties and how transverse electromagnetic radia-
tion 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
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) = Im ε(q, ω + iδ) | E T (q, ω) |2

ω3
= Im ε(q, ω + iδ) | A(q, ω) |2 (1219)
4 π c2
As the imaginary part of the dielectric constant is related to the real part of the
conductivity,  

Im ε(q, ω + iδ) = Re σ(q, ω) (1220)
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 (1221)
c

288
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 − (1222)
c c
Experiments measure the absorption coefficient η which is the fraction of light
absorbed in passing through unit thickness of the material

Re j . E κω
η = = 2 (1223)
n | E |2 c
Another, experimental method (ellipsometry) measures the reflectance of light.
This involves measuring the ratio of the reflected to the incident intensities, 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 po-
larized light has polarization parallel to the plane of incidence. The reflectances
are given in terms of the complex refractive index ñ = n + i κ, via the Fresnel
formulas
cos θ − ( ñ2 − sin2 θ ) 12

Rs (θ) = 1
(1224)
cos θ + ( ñ2 − sin2 θ ) 2
and
ñ cos θ − ( ñ2 − sin2 θ ) 12
2
Rp (θ) = 1
(1225)
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 relation (H.A. Kramers, Nature 117, 775 (1926), R. de
L. Kronig, J. Opt. Soc. Am., 12 547 (1926)).

11.3.7 Kramers-Kronig Relation


Causality requires that the frequency be continued in the upper half complex
plane ω + i δ in the response functions. This has the consequence that the
response function is analytic in the upper half complex plane. Also, it is required
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 + iδ) − 1 ) = dz (1226)
2πi c z − ω − iδ
where the contour is taken around the point z = ω + iδ. 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 + iδ) − 1 = Pr dz (1227)
πi −∞ z − ω

289
in which the contribution of the small semi-circle around the pole at z = ω + i δ
has been cancelled out. On writing

ε(q, z) = Re ε(q, z) + i Im ε(q, z) (1228)

one finds
Z +∞ Im ε(q, z)
1
Re ε(q, ω + iδ) − 1 = Pr dz (1229)
π −∞ z − ω
and Z +∞ Re ε(q, z) − 1
1
Im ε(q, ω + iδ) = − Pr dz (1230)
π −∞ z − ω
These relations can be recast in the form
Z ∞ z Im ε(q, z)
2
Re ε(q, ω + iδ) − 1 = Pr dz (1231)
π 0 z2 − ω2
and Z ∞ Re ε(q, z)

Im ε(q, ω + iδ) = − Pr dz (1232)
π 0 z2 − ω2
These are the Kramers-Kronig relations (H.A. Kramers, Nature 117, 775 (1926),
R. de L. Kronig, J. Opt. Soc. Am., 12 547 (1926)). They can be used to analyze
experimental data or as consistency checks.

——————————————————————————————————

11.3.8 Exercise 53
Derive the form of the Kramers-Kronig relation for the imaginary part of the
dielectric constant ε(q, ω)
Z ∞
4 π σ(q, 0) 2ω Re ε(q, z)
Im ε(q, ω) = − Pr dz 2 (1233)
ω π 0 z − ω2
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 Im εT (q, ω + iδ) = ω (1234)
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 − (1235)
ω → ∞ ω2

290
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) = (1236)
0 2 m

where ρ is the electron density. Kramers-Kronig relations and sum rules can be
established for a variety of response functions (P.C. Martin, Phys. Rev. 161,
143 (1967)). 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 54
The n-th moment of the imaginary part of the dielectric constant is defined by
Mn Z ∞
Mn = dω ω n Im ε(q, ω + iδ) (1237)
0
Show that M1 is given by
e2 ρ
M1 = 2 π 2 (1238)
m
and that M−1 is given by
 
π
M−1 = ε(q, 0) − 1 (1239)
2

——————————————————————————————————

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 (1240)
m 1 − iωτ

291
The Drude formula for the conductivity can be obtained directly from Kubo
formulae, in the case of a 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 > (1241)
for t > 0, one finds
 
i X t
Rα,β (r, r0 , t) = < Ψ0 | jpα (r) | Ψn > < Ψn | j β (r0 ) | 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
(1242)
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
(1243)
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
= − Im φγ 0 (r) ∇ φγ (r) (1244)
m
where the electron in the state labelled by the one-electron quantum number γ is
the 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 − E0 = E γ 0 − Eγ (1245)
Thus, one has
   
X Im φ∗γ (r) ∇α φγ 0 (r) Im 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

292
   
X Im φ∗γ (r0 ) ∇β φγ 0 (r0 ) Im φ∗γ 0 (r) ∇α φγ (r)
e2 h̄2
+
m2 h̄ ω + i η + Eγ 0 − Eγ
γ,γ 0
(1246)
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 − h̄ (k − k ) . (r − r )
e2 h̄2 X
Rα,β (r, r0 , ω) = − (k α
+ k 0α
) (k β
+ k 0β
)
4 m2 V 2 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

(1247)
α β
where k and k 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

(1248)
In this expression, the effect of the Pauli-exclusion principle is automatically
accounted for.

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) (1249)
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

(1250)

293
On Taylor expanding the Fermi-function f (Ek+q ) in powers of (Ek+q − Ek )
one has

2 e2 X ( Ek − Ek+q )
 
α,β α,β ∂f
R (q, ω) = − δ Ek
3mV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

2 e2 X
   
∂f h̄ ω + i η
= − δ α,β Ek 1 −
3mV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

(1251)

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
(1252)

since the boundary terms vanish. Furthermore, since 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 (1253)

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
 
α,α ∂f h̄ ω + i η
R (q, ω) = + Ek
m 3mV ∂Ek h̄ ω + i η + Ek − Ek+q
k,σ

(1254)

On substituting this expression into the conductivity, one finds the first term
cancels with the diamagnetic current. This cancellation is responsible for pro-
hibiting current 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 cancel-
lation 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

(1255)

294
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( E − EF ) (1256)
∂E
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 (1257)
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
E k
e2 τ X 2 ∂Ek
σ α,β (q, ω) = − δ α,β (1258)
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 (1259)
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 (1260)
m
and at finite frequencies has a real part that decays like ω −2
ρ e2 τ 1
Re σ α,β (ω) = δ α,β (1261)
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 Re σ α,β (ω) = δ α,β (1262)
0 2m
Hence, the intensity of the Drude peak provides a measure of the number of
conduction electrons in a system of non-interacting electrons. For a metal with

295
interacting electrons, the Drude peak, when integrated over a low frequency
range yields an estimate of the quasi-particle weight.

——————————————————————————————————

11.3.11 Exercise 55
Show that microwaves, with low frequency frequency ω, satisfy the equation
4 π i σ(ω) ω
− ∇2 E(r, ω) = E(r, ω) (1263)
c2
where σ(ω) is the diagonal component of the conductivity tensor. Solve this
equation for the electric field and hence calculate the classical skin depth δ.
The classical skin depth is defined as the distance δ that an electric field pene-
trates into a metal before being attenuated.

——————————————————————————————————

The analysis of Exercise 55 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
in space. This regime corresponds to the anomalous skin effect (A.B. Pippard,
Proc. Roy. Soc. A, 191, 385 (1947), A.B. Pippard, Proc. Roy. Soc. A, 224
273 (1954)). 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 compo-
nents of the fields E(q, ω) and B(q, ω) and by using an approximate expression
for the conductivity tensor in which both the wave vector and frequency depen-
dence are kept (D.C. Mattis and G. Dresselhaus, Phys. Rev. 111, 403 (1958)).
This procedure is crucial for the discussion of the anomalous skin effect.

——————————————————————————————————

11.3.12 Exercise 56
The conductivity tensor can be expressed as an integral over the Fermi-surface,
d2 S τ v α (k) v β (k)
Z
σ α,β (q, ω) = (1264)
| 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 expression
3π σ0
σ x,x (qz êz , 0) = (1265)
4 | qz | λ

296
(G.E.H. Reuter and E.H. Sondheimer, Proc. Roy. Soc. A195, 336 (1948))

——————————————————————————————————

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 electric field Ex (z, ω) is governed by
∂ 2 Ex ω2 4πiω
2
+ 2
Ex = − jx (1266)
∂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
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
= − (1267)
∂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
+ 2 Ex = − jx + 2 δ(z) (1268)
∂z 2 c c2 ∂z z=0
Hence, on Fourier transforming with respect to z and using Ohm’s law
jx (qz , ω) = σ x,x (qz ) Ex (qz , ω) (1269)
one obtains the solution
 
∂Ex

2 ∂z


z=0
Ex (qz , ω) = ω2 4 π i ω
(1270)
c2 − qz2 + c2 σ x,x (qz , ω)
In the limit of extremely long mean free path λ → ∞, the conductivity
simplifies to
3 π σ(0, 0)
σ x,x (qz , 0) = (1271)
4 | qz | λ
(G.E.H. Reuter and E.H. Sondheimer, Proc. Roy. Soc. A195, 336 (1948)). The
spatial dependence of the electric field is given by the inverse Fourier transform,
 
  Z ∞ exp − i q z z
∂Ex dqz
Ex (z, ω) = 2 2 (1272)
∂z z=0 −∞ π ωc2 − qz2 + c23 λπ | iq ω | σ(0, 0)
z

297
On defining δ via
 13
3 π 2 ω σ(0, 0)

δ −1 = (1273)
c2 λ
one has
 

 Z ∞ | qz | exp − i qz 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
(1274)
For low frequencies, the decay of the electric field is governed by δ the anoma-
lous 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
 
2 δ ∂Ex i
≈ − (1 + √ ) (1275)
3 ∂z z=0 3
Far from the surface, the field has an exponential decay
 2  
δ z
Ex (z, ω) ∼ exp − (1276)
z δ
which decays over a distance δ.

This result for the anomalous skin depth δ was first obtained by Pippard,
upto a numerical factor (A.B. Pippard, Proc. Roy. Soc. A, 191, 385 (1947),
A.B. Pippard, Proc. Roy. Soc. A, 224 273 (1954)). Pippard noted that only the
fraction of the electrons λδ close to the surface may participate in the screening
process. That is, only the electrons moving parallel to the surface are strongly
effected by the electric field. The electrons that remain within the penetration
depth δ before being scattered, subtend an angle of
δ
dθ ≈ (1277)
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 (1278)
π
since θ ≈ 2. Hence, the effective electron density, ρef f is given by
δ
ρef f ≈ ρ (1279)
λ

298
where ρ is the uniform electron density. The 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 ∼ 2
σ(0) (1280)
c λ
Hence, one has Pippard’s relation
  13
−1 4 π ω σ(0)
δ ∼ (1281)
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
due to inter-band transitions can be written as
2 Z 2
d3 k
  
e
Im ε(ω+iδ) = 8 π 2 h̄2

êα . M k δ( Ec,k − Ev,k − h̄ ω )

mω (2π) 3
(1282)
The matrix elements for the inter-band transitions are given by
Z    
3
êα . M k = êα . d r exp − i k . r uv,k (r) ∇ exp + i k . r uc,k (r)

(1283)

where un,k are the periodic functions of r in the Bloch functions. The sum over
k can transformed into an integral
2

2 Z êα . M k
d2 S
  
e
Im ε(ω+iδ) = 8 π 2 h̄2
mω ( 2 π )3
∇ ( E c,k − E v,k )

Ec −Ev =h̄ω
(1284)

299
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) = (1285)
( 2 π )3
∇k ( Ec,k − Ev,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, at which


∇k ( Ec,k − Ev,k ) = 0 (1286)
Ec −Ev =h̄ω

The inter-band transitions produces 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
of the photon, the energy of the absorbed photon is approximately given by the
energy difference of the electron’s initial and final states

h̄ ω ≈ Ec,k − Ev,k0 (1287)

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 Measuring the Fermi-Surface


The Fermi-surface determines most of the thermodynamic, transport and optical
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 magnetization
(W.J. de Haas and P.M. van Alphen, Proc. Amsterdam, Acad. 33, 1106 (1930)).

300
1
The magnetization is periodic in the inverse of the applied magnetic field H .
1
Onsager pointed out that the period in H is given by the expression
 
1 |e|
∆ = 2π Ae (1288)
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 (L. Onsager, Phil. Mag. 45,
1006 (1952)). 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 reconstruct the three-dimensional Fermi-surface
(D. Schoenberg, Proc. Roy. Soc. A 170, 341 (1939)). 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 (1289)
h̄ k
and
|e|
h̄ k̇ = − v(k) ∧ H (1290)
c
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 (1291)
c

The component of the velocity perpendicular to the applied field is given by


 
ṙ⊥ = ṙ − Ĥ ṙ . Ĥ
 
= Ĥ ∧ ṙ ∧ Ĥ (1292)

301
where Ĥ is the unit vector in the direction of H. Hence,
c h̄
ṙ⊥ = − Ĥ ∧ k̇ (1293)
| e | Hz
Thus, on integrating this one finds that the displacement ∆r⊥ is given in terms
of the displacement in k space through
c h̄
∆r⊥ = − Ĥ ∧ ∆k (1294)
| 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 = (1295)

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⊥ (1296)
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 = (1297)
∇ E⊥ | e | H z

k

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
(1298)

302
This relation can be substituted into the expression for the period to yield

h̄2 c
I
1
T = ∆k dk (1299)
| e | Hz ∆E
However, the area between the two successive orbits is given by the integral
I
∆A = ∆k dk (1300)

Thus, the period can be expressed as

h̄2 c ∆A
T = (1301)
| e | Hz ∆E
The orbits can be quantized through the energy uncertainty relation

En+1 − En =
T
| e | Hz ∆E
= (1302)
h̄ c ∆A

Furthermore, as
∆E En+1 − En
= (1303)
∆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 (1304)
h̄ c
This difference equation can be solved to yield the area enclosed by the n-th
Landau orbital as  
2π|e|
An = n + λ Hz (1305)
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 equation
(L. Onsager, Phil. Mag. 43, 1006, (1952)).

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 + (1306)
2
The mechanical momentum is given by
|e|
p = h̄ k − A (1307)
c

303
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 (1308)
c
The equation of motion can be integrated with respect to time, to yield

|e|
h̄ k = − r ∧ H (1309)
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 (1310)
c

However, the integral I


dr ∧ r = 2 Ar êz (1311)

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 (1312)

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 + (1313)
2 | e | Hz

Since the real space and momentum space orbitals are related via
h̄ c
∆r = ∆k (1314)
| 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 (1315)
2 h̄ c

304
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
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 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 + (1316)
2 2m
where
| e | Hz
ωc = (1317)
mc
The orbit maps out a circle of area
 
2π|e| 1
An = Hz n + (1318)
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 contained
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

305
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 ) (1319)
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
(1320)
Thus, the extremal orbit crosses the Fermi-surface when Hz is increased by
n ∆H ≈ Hz (1321)
∆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 (1322)
∆H h̄ c
Thus, n decreases by unity at fields given by
∆H 2π|e| 1
− = − (1323)
H2 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(E) = D Lz δ( E − n h̄ ωc − )
n −∞ 2 π 2m
Lx Ly Lz X Z ∞ h̄2 kz2
= m ω c dk z δ( E − n h̄ ω c − )
4 π 2 h̄ n 0 2m
Lx Ly Lz 2 X θ( E − n h̄ ωc )
= 2 2 m ωc
p
4 π h̄ n 2 m ( E − n h̄ ωc )
(1324)
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 (1325)
2 π h̄

306
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
coupling, the average of A is given by
X Z +∞
A = dE f (E) ρ( E − µB σ Hz ) Aσ (E) (1326)
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 (1327)

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 (1328)
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 57
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
(1329)
∂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|
2
< (1330)
Hz h̄ c A
and A is the area of the extremal orbit.

307
——————————————————————————————————

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 ) (1331)
α

where the sum over α runs over all the one-electron states.

The Lifshitz-Kosevich formulae describes the oscillatory parts of M (I.M.


Lifshitz and A.M. Kosevich, Sov. Phys. J.E.T.P. 2, 636 (1956)). This shall be
examined in the T → 0 limit. In the limit T → 0 one has
X  
lim Ω = Eα − µ Θ( µ − Eα ) (1332)
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 ) (1333)
α
The dispersion relation for free electrons in an applied field is given by
h̄2 kz2
 
1
Eα = + n + h̄ ωc − µB Hz σ (1334)
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
(1335)
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 (1336)
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
(1337)

308
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+ = − (1338)
h̄ ωc 2
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
(1339)

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 + ) (1340)
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

(1341)

where the upper limit of integration is given by


µ mc
x+ = + µB σ (1342)
h̄ ωc | e | h̄

However, as
| e | h̄
µB = (1343)
2mc
the upper limit of integration becomes
µ σ
x+ = + (1344)
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.

309
On Fourier analyzing β(x) one has

X 1
β(x) = 1 + 2 cos 2πp ( x − ) (1345)
p=1
2

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
! 32 #
x+
X 1
+2 dx cos 2πp ( x − ) µ − x h̄ ωc + µB Hz σ
p=1 0 2
(1346)
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+    32
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
(1347)
since the boundary term vanishes. Integrating by parts once again
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+ cos πp −
2
dx cos 2πp ( x − ) x+ − x
8 π 2 p2 2 0 2
(1348)
Changing variables from x to u, where
u2
 
2 p x+ − x = (1349)
2
so
1
dx = − u du (1350)
2 p
Then, 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
(1351)

310
The cosine term can be decomposed as

π u2 π u2 π u2
cos ( + φ ) = cos cos φ − sin sin φ (1352)
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 (1353)
0 2
which for large x have the limits
1
C(∞) = S(∞) = (1354)
2
Thus, the integral 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
(1355)

Thus, the oscillatory part of the grand canonical potential ∆Ω is


  32  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
(1356)

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

311
  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
(1357)

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 = − (1358)
∂Hz
Thus, the magnetization also has oscillations that are periodic in H1z . Further-
more, for a free electron gas the extremal area of the Fermi-surface is just 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 experiments. 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. This can be used to map out the Fermi-surface.

The Lifshitz-Kosevich formulae, valid at finite temperatures, is


 
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
(1359)

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
(1360)

The splitting between the up-spin and down-spin bands has modified the rel-
ative phase of the higher harmonics in the oscillations. This can be used to

312
extract the ratio of the band mass of the electron to the electron mass in vac-
uum. For systems which are on the verge of ferromagnetism, the spin splitting
factor should be enhanced 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 of the impurities τ1 . Both these effects reduce the amplitude
of the de Haas - van Alphen oscillations (R.B. Dingle, Proc. Roy. Soc. A 211,
257 (1952)). The oscillations can only be seen at low temperatures T < 1 K
and for samples of high purity, as indicated by small residual resistances. 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 accompanied 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 experiments.

11.4.5 Other Fermi-Surface Probes


There are many other probes of the Fermi-surface, these include the attenuation
of sound waves. Consider sound waves in a crystal propagating perpendicular to
the direction of the applied magnetic field and having a transverse polarization
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 length of
the mean free path is sufficiently long, 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 electrons
motion. This strongly depends on the mismatch between q −1 and the diameter
of the orbit. When the radius of rq the orbit is such that
λ
2 rq = (1361)
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

313
the coupling is strong. When
2 rq = λ (1362)
the electron is sequentially accelerated and decelerated by the field. The cou-
pling 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 rq = ( n + )λ (1363)
2
and weak coupling occurs when the interference is destructive
2 rq = n λ (1364)
The period differs slightly from the asymptotic large n variation just described.
Assume that the projection of the trajectory on the plane perpendicular to the
applied field Hz is circular. The energy transfer between the electron and the
electronic wave in one orbit is given by
Z ω2π
c
dt E(r, t) . v(t) (1365)
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) = vF sin ωc t (1366)
and
x(t) = rq sin ωc t (1367)
Thus, the energy transfer in a period is evaluated as
Z ω2π  
c
Ey vF dt exp i q rq sin ωc t sin ωc t
0
 

= Ey vF J1 ( q rq ) (1368)
ωc
Thus, 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. The Pauli exclusion principle forbids electrons in
other states to undergo low energy excitations, since the slightly higher energy
states are already occupied.

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
∆ = (1369)
λ 2 rq

314
By variation of q, and H one can map out the Fermi-surface.

11.4.6 Cyclotron Resonances


This method requires the application of an microwave electric field at the sur-
face 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 of the surface.

A static (d.c.) magnetic field is applied parallel to the surface say in the x
direction, so that the electrons undergo spiral orbits in real space. The veloc-
ity 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 these electron couple that to the
microwave 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 fre-
quency ωc is large, and for microwave frequencies ω where the anomalous skin
depth phenomenon occurs. The condition of a long mean free path and large
cyclotron 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 = (1370)
ωc
which is the period of the cyclotron motion. In general, the period is given by

h̄2 c
 
∂A
TH = (1371)
| e | Hx ∂E

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. That is

TH = n TE = n (1372)
ω
Hence, this requires that the frequency of the cyclotron orbit match with the
frequency of the a.c. electric field

ω = n ωc (1373)

315
so that the a.c. field resonates with the electronic motion in the uniform field.
This condition can be written as
1 n
= 2 π | E | h̄2 c ω   (1374)
Hx ∂A
∂E

The factor
h̄2
 
∂A
mc = (1375)
2π ∂E
h̄2 k2
is known as the cyclotron mass. For free electrons, A = π k 2 and E = 2 m
so the cyclotron mass coincides with the electron mass

mc = m (1376)
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.

The geometry shall be considered where a surface has a normal in the z


direction and a d.c. magnetic field is applied parallel to the surface in the x
direction
H = êx Hx (1377)
and the a.c. electric field is in the y direction
 
E = êy Ey exp i ( q z − ω t ) (1378)

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 velocity lies in the plane z = const. .

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 (1379)
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

316
evaluated at the electron’s position
Z t+TH   
v
− |e| dt0 Ey exp i q z0 + q sin ωc t0 − ω t0
t ωc
    
2π qv
∼ − | e | Ey exp i q z0 − ω t I ωωc
ωc ωc
(1380)
q v
where In (x) is the modified Bessel function of order n. For large ωc , this has
the asymptotic form
  12   
2π π
∼ − | e | Ey exp i q z0 − ωt − (1381)
q v ωc 4

Due to the phase differences around the orbit, only a fraction


  12
2 π ωc
(1382)
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 − (1383)
q ωc 4

The previous traversal caused a similar displacement, but with t → t − 2ωcπ .


However, only the fraction  

exp − (1384)
ωc τ
of electrons survive traversing one cyclotron orbit without scattering. As these
all have a similar form, one can obtain the average energy displacement experi-
enced by an electron between one scattering and the next
  12   
2πv π
v h̄ ∆ky = − | e | Ey exp i q z0 − ωt − F (1385)
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
∞  
X 2πn
F = exp − (1 + iωτ )
n=0
ωc τ

1
=   (1386)
2 π
1 − exp − ωc τ (1 + iωτ )

317
It is the imaginary part of the probability for survival that causes the resonances
in the surface impedance.

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 is, thus, 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 in which the a.c. field has been turned off.

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 | dkx kx exp i ∆ky v
8 π3 qv 4
e2
 
1
= dk x kx F Ey (1387)
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 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 .

The surface impedance is defined as


4 π i ω Ex
Z(ω) = ∂Ex
∂x
  12
2πω
∼ (1 − i)
σ
1
∼ F−2 (1388)
The surface impedance has oscillations with varying field Hx . It should be ob-
vious from the above discussion that the oscillations provide information on ωωc ,
but the dominant contribution 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-surface.

318
11.5 The Quantum Hall Effect
The Quantum Hall Effect is found in two-dimensional electron systems, in which
an electric field is applied perpendicular to the plane where the electrons are
confined. Experimentally, the electrons can be confined to a two-dimensional
sheet in a metal oxide semiconductor field effect transistor. The application
of a strong electric field to the surface of a semiconductor may pull down the
conduction band at the surface of the semiconductor. If the energy of these field
induced surface states is less than the Fermi-energy of the metal, electrons will
tunnel across the insulating oxide barrier and occupy them. After equilibrium
has been established, the electrons at the surface of the semiconductor will form
a two-dimensional electron gas.

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 simple harmonic motion.
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 (1389)
the Hamiltonian for the two-dimensional motion in the x − y plane is given by
2
p̂2x m ωc2

Ĥ = + x̂ − X̂ (1390)
2m 2

where the cyclotron frequency ωc is given by

| e | Bz
ωc = (1391)
mc

319
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 + (1392)
| e | Bz
The energy eigenvalues of the shifted Harmonic oscillator are given by
1
Eν,0 = 0 + h̄ ωc ( ν + ) (1393)
2
where ν is the quantum number for the Landau levels. The Landau levels are
independent of ky and, therefore, are degenerate. Since the values of ky are
quantized via
2 π ny
ky = (1394)
Ly
for a surface of length Ly , then the possible values of ky are limited by the
restriction
Lx > X > 0 (1395)
which yields the total number of degenerate states as
Bz | e |
D = Lx Ly
2 π h̄ c
Φ|e|
= (1396)
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 = (1397)
|e|
The density of states can be approximately expressed as a discrete set of delta
functions
 
| e | Bz Lx Ly X 1
ρ(E) = δ E − 0 − h̄ ωc ( ν + )
hc ν
2
 
m ωc Lx Ly X 1
= δ E − 0 − h̄ ωc ( ν + )
2 π h̄ ν
2
(1398)

The weight associated with each delta function corresponds to the degeneracy
of each Landau level. On defining the cyclotron radius rc as
s
h̄ c
rc = (1399)
| e | Bz

320
one finds that the relative position operator can be expressed as
 
rc †
x̂ − X̂ = √ aky + aky (1400)
2
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 (1401)
2
as p̂y can be taken to be diagonal. 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 ) (1402)

= − | ie h̄| Bc z . On substituting for the x


 
since the commutator ŷ , X̂
component of the displacement from the center of the orbit
rc ω c
v̂y = √ ( a†ky + aky ) (1403)
2
Thus, as the velocity operators are non-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. On using the form of the Kubo formula for the conductivity
tensor, per unit area, appropriate for single particle excitations
"
i e2 1 X
σα,β (ω) = < ν, ky | v̂α | ν 0 , ky > < ν 0 , ky | v̂β | ν, ky > ×
ω + i η Lx Ly 0
ν,ν ,ky
#
f (Eν ) − f (Eν 0 ) X f (Eν )
× − δα,β
h̄ ω + i η + h̄ ωc ( ν − ν 0 ) m
ν,ky

(1404)

one finds that the diagonal component is zero

Re σx,x (0) = 0 (1405)

but, nevertheless, the off-diagonal term is finite and quantized


e2
Re σx,y (0) = (n + 1) (1406)
2 π h̄

321
where n is the quantum number for the highest occupied Landau orbital. As
the field is changed, the peaks in the density of states associated with the Lan-
dau levels sweep through the Fermi-level. The Hall resistivity should, therefore,
show a set of steps as the applied field is increased. This phenomenon is the
integer quantum Hall effect.

Experimentally, it is found that the steps in σx,y (0) are not discontinuous,
but instead show a finite slope in the transition region. Furthermore, the di-
agonal component of the resistivity is non-zero, but shows spikes for fields in
the region where the transition between successive plateaus occur. This phe-
nomenon is associated with impurity scattering. The effect of impurities is to
broaden the set of delta function peaks in the density of states into a set of
Gaussians. This allows the transition between the plateaus to be continuous.
In fact, if all the states contributed to the conductivity, the steps of the stair-
case would be smeared out into a straight line, just like in the Drude theory for
three-dimensional metals. Fortunately, the states in the tails of each Gaussian
are localized, as the deviation from the ideal Landau level energy indicates that
these states experience a larger impurity potential than average. The large po-
tential acts to localize the electrons in the states with energies in the Gaussian
tail and so do not contribute to the Hall conductivity. In fact, in two dimen-
sions with zero field, one can show that all the states are localized in an infinite
sample. However, the samples are finite and have edges. The edges have ex-
tended states that carry current. The edge states can be understood in analogy
with the classical motion where there are skipping orbits, in which 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 is given by
p̂y q
v̂y = − Ay
me me c
= ωc ( x̂ − X̂ ) (1407)
and the probability density for the shifted harmonic oscillator is symmetrically
peaked about X. Since the current carried by the state is given by an integral
which is almost anti-symmetric
Z L
q ωc
jky ,ν = dx | φν ( x − X ) |2 ( x − X ) (1408)
me 0
it vanishes. 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 L
q ωc
jky ,ν = dx | φν ( x ) |2 x (1409)
me 0
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

322
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 level 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, Laughlin proposed a gauge theoretic argument which overcomes
these shortcomings (R.B. Laughlin, Phys. Rev. B 23 5632, (1981)). Laughlin
envisaged an experiment in which the two-dimensional sample is in the form of
the surface of a hollow cylinder of radius R. 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 confined. 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 Φ does lead to a finite vector potential AΦ
which satisfies
∇ ∧ AΦ = 0 (1410)
which, therefore, can be written in the form
AΦ = ∇ Λ (1411)
This vector potential does cause the electronic wave functions to acquire an
Aharonov-Bohm phase factor of
 
qc
exp i Λ (1412)

For the vector potential
Φ
AΦ = êϕ (1413)
2πr
one finds that
Φ
Λ = ϕ (1414)

and, hence, the phase factor is given by
 
qcΦ
exp i ϕ (1415)
2 π h̄
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
 
qcΦ
exp i 2π (1416)
2 π h̄

323
Since the fundamental flux quantum Φ0 is defined by
2 π h̄
Φ0 = (1417)
qc
this Aharonov-Bohm factor can be written as
 
Φ
exp i 2π (1418)
Φ0
Thus, if Φ is an integer multiple of Φ0 , i.e.

Φ = µ Φ0 (1419)

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
extended throughout the sample. The vector potential at position z is given by
Φ
Br z + = µ Φ0 (1420)
2πR
which, according to the flux quantization condition, must be an integer multiple
of Φ0 for the phase of an extended wave function to be single valued. If Φ
is adiabatically increased by Φ0 , then the maximum the extended states in a
Landau level must be translated along the z axis by amounts ∆z
Φ0
∆z = −
2 π R Br
Φ0
= −L (1421)
Φ
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.
Since, in the pure systems, the fully occupied Landau level contains a number
| ∆z | Φ
m = = (1422)
L Φ0
of electrons, the adiabatic change of Φ results in the transfer of electrons be-
tween neighboring extended states. Hence, in the dirty system, all the delocal-
ized electrons in the Landau level are translated along the z direction by one
spacing, skipping over the localized states. The net result is that one electron is
translated across the entire 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, adding 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 L (1423)

324
The current around the cylinder Iϕ , from all the electrons in a single Landau
level, is given by
∂E
Iϕ = − c (1424)
∂Φ
which leads to a current density
c q Ez L
jϕ = −
L Φ0
q2
= − Ez (1425)
2 π h̄
Hence, on summing over all occupied Landau levels, one has

q2
jϕ = − n Ez (1426)
2 π h̄
In this n is the number of completely occupied Landau levels with extended
states and the Fermi-energy is in a mobility gap. The Hall conductivity is given
by
jy q2
σH = − = n (1427)
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 Klitz-
ing in 1980. (K. von Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Letts.
494 (1980)). The steps can only be discerned in very clean samples. At much
higher fields, where only the lowest Landau level should be occupied, Gossard,
Stormer, and Tsuei discovered a similar type of effect which is known as the
fractional quantum Hall effect (D.C. Tsuei, H.L. Stormer and A.C. Gossard,
Phys. Rev. Letts. 48, 1559 (1982)). This phenomenon involves the effect of the
Coulomb repulsion between electrons in the Landau levels. Laughlin showed
that the energy of the interacting electron states can be minimized by allowing
the electrons to form a ground state with a different symmetry from the bulk
(R.B. Laughlin, Phys. Rev. Letts. 50, 1395 (1983)).

——————————————————————————————————

11.5.2 Exercise 58
Evaluate the Kubo formula for the real part of the diagonal and off diagonal
quantum Hall conductivities by first taking the limit ω → 0 and then taking
the limit  → 0. Also estimate the effects of introducing scattering due to
random impurities. The effect of the scattering lifetime can be introduced by
including imaginary parts of the energies of the occupied and unoccupied single
particle states of the form ± i 2h̄τ . Choose the signs to ensure that the wave

325
functions for the excited states (with an electron - hole pair) will decay to the
ground state after a time τ . Compare your result with the conductivities ob-
tained for the three-dimensional Drude model.

——————————————————————————————————

11.5.3 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 can be written as
m
x2j + yj2
r   
2 xj + i yj
φm (rj ) = exp − (1428)
π m! 2ξ 4 ξ2
where the length ξ is given by
s
h̄ c
ξ = (1429)
q Bz

The probability density for finding a particle has a peak which form a circle
around the origin. The radius of the circle depends on m.

The many-particle ground state wave function corresponding to the 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 N particle wave function is
Ne
Y  x2k + yk2
 Y  
Ψ ∼ ( xi − xj ) + i ( yi − yj ) exp − (1430)
i>j
4 ξ2
k=1

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 = . (1431)

..
..

i>j .

z0
Ne z1
Ne z2
Ne . . . z Ne −1 Ne

Hence, the Ne electrons occupy the zero-th Landau levels single particle states
with all the angular momentum quantum numbers in the range between m =

326
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 ( Ne − 1 )
2 h̄. Since each m value is occupied, this state corresponds to a
uniform particle density of
1
ρ = (1432)
2 π ξ2
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 function
Ne
p Y
Y  x2k + yk2
 
Ψp ∼ ( xi − xj ) + i ( yi − yj ) exp − (1433)
i>j
4 ξ2
k=1

for odd integers p. The higher power of p 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 linear superposition contains 1 particle for
every p values of m, this state corresponds to a uniform particle density of
1
ρ = (1434)
2 π ξ2 p
particles per unit area. The filling factor, ν is defined as
Φ0 ρ Φ0
ν = Ne = (1435)
Φ Bz
This state corresponds to a state with the fractional filling, ν = p1 , of the
lowest Landau level. The energy of the Laughlin ground state is given 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 as
  2
Eg 0.78213 0.211 0.012 e
= − √ 1 − 0.74 + (1436)
Ne p p p1.7 rc
where r

rc = (1437)
me ω c
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.

327
11.5.4 Quasi-Particle Excitations
The quasi-particle excitations of the Laughlin state, like the quasi-particle ex-
citations 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 at zero, it does add an Aharonov-Bohm phase to the system.
First we shall consider the filled nr = 0 Landau level.

The single particle wave function experiences a vector potential of the form
 
Bz r Φ
A = + êϕ (1438)
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
(1439)
Then, with the ansatz
 
1
φ(r, ϕ) = √ exp + iµϕ R(r) (1440)

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 c 2 π h̄ 2 h̄ c
(1441)
Hence, the solutions for the lowest Landau level are of the form

r2
     
qΦ ν
φ(r, ϕ) ∼ exp i ϕ exp i ν ϕ r exp − (1442)
c 2 π h̄ 4 ξ2

where

ν = µ − (1443)
c 2 π h̄
Since the wave function is single valued µ must be an integer, say m. Thus, on
increasing ν the particles move away from the origin. On subtracting one flux
quantum, Φ0 , through the center of the loop, where
c 2 π h̄
Φ0 = (1444)
q
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

328
is the quasi-hole excitation in the filled Landau level.

The wave function of the Laughlin state, when a quasi-hole has been added 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 x2 + y 2
Y  
Ψ+
p ∼ ( xi + i yi ) ( xi − xj ) + i ( yi − yj ) exp − k 2 k
i i>j

k=1
(1445)
and the wave function with a quasi-hole at r0 is given by
Ne  Ne
p Y
x2 + y 2
Y Y  
Ψ+
p ∼ xi − x0 + i ( yi −y0 ) ( xi − xj ) + i ( yi − yj ) exp − k 2 k
i=1 i>j

k=1
(1446)
where one flux quanta has also been removed from r0 . This state has angular
momentum of
Ne ( Ne − 1 )
Mz = p h̄ + Ne h̄ (1447)
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
charge − pq . Alternatively, one may notice that by adding p quasi-holes at the
same point and then add an electron there, one just obtains the Laughlin wave
function 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 ) (1448)
i=1

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 − (1449)
i=1
∂xi ∂yi ξ2

329
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 par-
ticle 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 excitation spectra
has a gap.

Since each quasi-particle of charge pq is attached to one flux quantum, the


statistics are neither fermion nor boson. Two quasi-particles are exchanged in
the 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 states. 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 = (1450)
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 is only possible in two or less


dimensions. If the permutation of two particles produces a phase difference of πp
then the reverse permutation process must yield a phase change of − πp . In two
dimensions, the permutation process and the reversed process are distinguish-
able. However, if the two-dimensional process is embedded in three dimensions
the processes are no longer distinct. On rotating the plane by in which the par-
ticles are contained in by π, the interchange becomes equivalent to the reverse
interchange. Hence,
   
π π
exp + i = exp − i (1451)
p p

which yields p = 1.

11.5.5 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 is determined by the quantity
q
(1452)
2 m∗ c

330
where m∗ is the band mass, which in GaAs has the value of m∗ ≈ 0.07 me
where me is the free electron mass. The strength of the spin coupling is given
by
gq
(1453)
me c
where g is the gyro-magnetic ratio g ≈ 2. This is further reduced by the
strong spin-orbit coupling in GaAs to a value given by the Lande gl factor,
gL ≈ 0.45. The spin directions are, therefore, determined via the exchange
parts of the Coulomb interaction. The magnitude of the Coulomb interaction
is given by
q2
(1454)

where  ∼ 12 and ξ is the magnetic length. For fields of order B = 10 Tesla
this is the same order of magnitude as h̄ ωc .

The spin magnetization M contributes to the effective magnetic field

B ef f = B + 4 π M (1455)

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 in the filling factor. The system will respond to
the change in the filling factor by transferring charge. Thus, spin and charge
excitations are coupled. The lowest energy coupled spin-charge excitations are
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
Ĥint = − g µB B ef f (r) . S (1456)
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 spins 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 ) (1457)

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

331
case, the Zeeman Hamiltonian is given by
ĤZ = − µB ( B . σ ) (1458)
which can be expressed as
 
cos θ sin θ exp[− i ϕ ]
ĤZ = − µB B (1459)
sin θ exp[ + i ϕ ] − cos θ
which for fixed (θ, ϕ) has an eigenstate of ĤZ
cos θ2
 
χ+ = (1460)
sin θ2 exp[ + i ϕ ]
which has the eigenvalue
E0 = − µB B (1461)
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 (1462)
sin θ2 exp[ + i ϕ ] h̄
where the time dependence is given purely by the exponential phase factor.

If the direction of the field (θ(t), ϕ(t)) is changed very slowly, one expects the
spin will adiabatically follow the field direction. That is, if the field is rotated
sufficiently slowly, one does not expect the spin to make a transition to the
state with energy E = + µB B where the spin is aligned anti-parallel to the
field. However, the wave function may acquire a phase which is different from
the time and energy dependent phase factor expected for a static field. This
extra phase is the Berry phase δ, 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)
(1463)
We shall assume that the wave function takes 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
(1464)
which instantaneously follows the direction of the field but is also modified by the
inclusion of the Berry phase. 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) ]
(1465)

332
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) ] (1466)

One finds that the derivative of θ w.r.t. t cancels and that the equation simplifies
to
∂δ ∂ϕ θ
− + sin2 = 0 (1467)
∂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 ) (1468)
2
On completing one orbit in spin space, the extra phase is given by

δ = (1469)
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 of
Ω Φ0
∆Φ = (1470)
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
Φ0
ν = ∆N (1471)
∆Φ
Hence, the contour encloses an extra charge

∆Q = q ∆N
∆Φ
= νq
Φ0

= νq (1472)

Thus, the extra charge is determined by the Berry phase and the filling fraction
ν, also the spin and charge excitations are coupled.

333
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 level. The ground state
wave function is written in second quantized form as
Y †
| Ψ0 > = am,↑ | 0 > (1473)
m

The creation of a charged spin-flip excitation at the origin requires adding a


down spin electron in m = 0. However, to allow for the spin excitation to have
a finite spatial extent and the charge density to re-adjust, the wave function
needs to be able to reduce the charge density and net spin at the origin by
redistributing them on neighboring shells. The state is also an eigenstate of
total angular momentum Jz . Thus, to a first approximation the excited state
wave function can be written as
  Ne
u0 a†1,↓ a†0,↓ a†m,↑ | 0 >
Y
| Ψ+ > ≈ v0 + a0,↑
m=0
  Ne
v0 a†0↑ + u0 a†1,↓ a†0,↓ a†m,↑ | 0 >
Y

m6=0
(1474)

where v0 and u0 are variational parameters which, since the wave function is
normalized, must satisfy

| u0 |2 + | v0 |2 = 1 (1475)

Iterating this process leads to the skyrmion wave function


Ne  
vm a†m↑ + um a†m+1,↓ a†0,↓ | 0 >
Y
| Ψ+ > =
m=0
(1476)

where one expects that as m → N then | 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 minimizing the expectation value of Ĥ.

The Hamiltonian can be approximated by


1 X
Vm,m0 a†m,↑ a†m0 +1,↓ am0 ,↑ am+1,↓ (1477)
X
Ĥ = m,σ a†m,σ am,σ +
m,σ
2! 0m,m

which is a simplified version of the skyrmion Hamiltonian. On using the relations

< Ψ+ | a†m,↑ am,↑ | Ψ+ > = | vm |2

334
< 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
(1478)

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
2! 0m,m
(1479)

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 (1480)
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 (1481)
2 m 0
m

and
1 ∗ X
( m+1,↓ − λm ) u∗m + v Vm,m0 vm0 u∗m0 = 0 (1482)
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 (1483)
2 2

where we have defined the parameter


1 X ∗
∆m = Vm,m0 vm 0 um0 (1484)
2! 0 m

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
 
2 1 m,↑ − m+1,↓
| um | = 1 ± p
2 ( m↑ − m+1,↓ )2 + 4 | ∆m |2
(1485)

335
Far from the center of the spin flip excitation one expects that the spins will be
polarized parallel to the field and the ground state will be recovered. Hence, we
shall use the upper signs. The equations (1484) and (1485) 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 (1486)
2 (  0
m↑ −  0 2
m +1,↓ ) + 4 | ∆m0 |
2
m0

The solution uniquely determines the wave function up to an undetermined


phase. We note that as the Zeeman splitting increases, the magnitude of ∆m
decreases.

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 . σ )(r) = Ψ† (r) ( η̂ . σ ) Ψ(r) (1487)

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
m0 ,β
(1488)

Hence, one finds the component of the spin density operator in the form
X X
( η̂ . σ )(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,↑ am0 ,↑ − a†m,↓ am0 ,↓
X
+ φ∗m (r) φm0 (r) cos θ0
m,m0
(1489)

On taking the expectation value of the spin density operator in the skyrmion
state one finds
X
< Ψ+ | ( η̂ . σ )(r) | Ψ+ > = sin θ0 exp[ − i ϕ0 ] φ∗m (r) φm+1 (r) um vm

m
X
+ sin θ0 exp[ + i ϕ0 ] φ∗m+1 (r) φm (r) vm u∗m
m

336
X  
+ cos θ0 | φm (r) |2 | vm |2 − | φm+1 (r) |2 | um |2
m
(1490)

The wave functions can be expressed in planar polar coordinates as


 
1
φm (r, ϕ) = √ exp + i m ϕ Rm (r) (1491)

where Rm (r) is real. Hence, we have


X
< Ψ+ | ( η̂ . σ )(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  
0 2 2 2 2
+ cos θ Rm (r) | vm | − Rm+1 (r) | um |
m
(1492)

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
(1493)
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 . σ )(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
(1494)

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)
tan θ0 =  m m
m  (1495)
2 2
P 2 2
m0 | vm0 | Rm0 (r) − | um0 | Rm0 +1 (r)

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 ξ 2 (1496)

337
In this case, one has Rm+1 (r) ∼ 2 √rm ξ Rm (r), hence, the out of plane angle
at the distance r from the origin is governed by

θ0

r um
tan ∼ √ (1497)
2 2 m ξ 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 = (1498)
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,↑
(1499)
∆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 gyromag-
netic 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.

11.5.6 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
effect is observed is given by the expressions
n
ν = (1500)
2nr ± 1
and
n
ν = 1 − (1501)
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.

338
These other states can be expressed in terms of composite fermions. 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

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

k=1
(1502)

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) will be that 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 (1503)
and, thus, the composite particle is a fermion.

Consider the state with the general filling factor


n
ν = (1504)
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 (1505)

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 fractional
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∗ = (1506)
Φ∗
in which Φ∗ and, therefore, ν ∗ are assumed to be positive. Then, substituting
Φ∗ and Φ into the relation given by eqn(1505) yields
1 1

= − 2r (1507)
ν ν

339
or
ν∗
ν = (1508)
2 ν∗ 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 with
2 r flux tubes attached to each electron. 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.

340
12 Insulators
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.

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. 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 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

341
valence band and conduction band states as
d
X
Ec (p) = Eg + ai p2i
i=1
d
X
Ev (p) = − bi p2i (1509)
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)
= (1510)
mα,β ∂pα ∂pβ

one finds a diagonal effective mass tensor mα,β , which is positive for the con-
duction band and negative for the valence band.

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 =
), which are separated by a band gap Eg . At temperatures comparable to the
band gap
Eg ∼ kB T (1511)
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 to 0.5 eV. Thus, the number of thermally
activated electron hole pairs is expected to be extremely small under ambient
conditions.

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.

342
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 conduction 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 semiconducting
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) = − (1512)
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 (1513)
me e2
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) = − (1514)
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

343
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
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.

344
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 E, Ph (E) is given by the
probability that an electron is not occupying that state

Ph (E) = 1 − f (E) (1515)

Thus, as
1
Ph (E) = 1 − f (E) =   (1516)
1 + exp − β (E − µ)

one finds that the probability of finding a hole in a state of energy E is also
given by the Fermi-function except that E → − E 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 (1517)

Thus, the momentum of a hole is opposite to that of the missing electron

kh = − ke (1518)

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 (1519)
dt
As the momentum of a hole is given by

kh = − ke (1520)

one finds that


dk h
h̄ = +|e|E (1521)
dt
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 (1522)

345
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 ) (1523)

where the number of electrons is now Ne − 1 = Z Nn − 1. Hence,


Qh = | e | (1524)
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 ) (1525)
Thus,
vh = ve (1526)
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 (1527)
dt
and since
ve = vh (1528)
and the charge of the hole is | e | one has
dv h
− me = |e|E (1529)
dt
Hence,
mh = − me (1530)
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.

346
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

∂2E 1
= < 0 (1531)
∂2k mh
The energy of the conduction electron is given by

h̄2 k 2
E(k) = Eg + (1532)
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(E) = ( E − Eg ) 2 f or E > Eg (1533)
2 π2 h̄2
The number of electrons in the conduction band, Nc is given by
Z ∞
Nc = dE ρ(E) f (E) (1534)
Eg

The Fermi-function can be expressed in terms of ( E − Eg ) as

1
f (E) =     (1535)
exp β ( E − Eg ) exp − β ( µ − Eg ) + 1

E
If it is assumed that µ = 2g then, as the lowest conduction band state is at
E = Eg , the Fermi-function for the conduction electrons is almost classical
since both exponents are positive. Thus, with the assumption
 
f (E) ∼ exp − β ( E − µ ) (1536)

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 E to the

347
dimensionless variable x = β ( E − Eg ). Then
 3   Z ∞  
V 2 me kB T 2 1
Nc ∼ exp − β ( E g − µ ) dx x 2 exp − x
2 π2 h̄2 0
 23  
2 π me kB T
= 2V exp − β ( Eg − µ )
h2
(1537)
which is usually expressed in terms of the thermally De Broglie wavelength of
the conduction electrons λe defined by
h
λe = √ (1538)
2 π me kB T
as  
V
Ne = 2 3 exp − β ( Eg − µ ) (1539)
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 (E) = 1 − f (E)
1
=  
1 + exp − β (E − µ)

which is the Fermi-function except that E → − E and µ → − µ. The density


of states for the holes in the valence band is given by
  32
V − 2 mh 1
ρ(E) = ( − E )2 f or E < 0 (1540)
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 = dE Ph (E) ρ(E)
−∞
  32 Z 0 1
V − 2 mh kB T ( − E )2
= dE
h̄2
 
2 π2 −∞
exp − β(E − µ) + 1
  32  
− 2 π mh kB T
∼ 2V exp − βµ
h2
(1541)

348
which also depends on µ.

Since in an intrinsic semiconductor the electrons and holes are created in


pairs one has
Nc = Nv (1542)
Therefore, the equation for the unknown variable µ is given by
    32  
− mh
exp 2βµ = exp β Eg (1543)
me

or on taking the logarithm one has



Eg 3 − mh
µ = + kB T ln (1544)
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 (1545)
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 (1546)

349
where the total energy is
E = − nd Ed (1547)
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 (1548)

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 = (1549)
nd ! ( 2 Nd − nd )!

Hence, the Free energy can be calculated in the thermodynamic limit as

F = − Ed n d
 
− kB T 2Nd ln 2Nd − nd ln nd − ( 2Nd − nd ) ln( 2Nd − nd )

(1550)

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 ) (1551)

Thus, on exponentiating
 
nd
exp β (µ + Ed ) = (1552)
2 N d − nd

which leads to the number of electrons occupying the donor orbitals as


2 Nd
nd =   (1553)
exp β ( − Ed − µ ) + 1

which is just governed by the Fermi-function f (−Ed ) and the number of orbitals
2 Nd .

350
The number of thermally excited conduction electrons is given by
Z
1
nc = 2 V d3 k  
exp β ( E(k) − µ ) + 1
  32  
2 π me kB T
= 2V exp βµ (1554)
h2

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 (1555)
nd h2

This is the law of mass action for the dissociation reaction

nd → nc + ( N d − nd ) (1556)

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 (1557)
nd λ

where λ is the thermal De-Broglie wavelength,


h
λ = √ (1558)
2 π me kB T
On using the condition of electrical neutrality

N d = nd + n c (1559)

one can eliminate nd to find the number of conduction electrons as


 
nc ( N d + nc ) 2V
= exp − β Ed (1560)
( 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 E d
2 λ λ3
(1561)

351
With this expression for the number of conduction electrons one can solve for µ
from
nc λ 3
 
µ = kB T ln (1562)
2V
At sufficiently low temperatures when

λ3 N d
 
 exp − β Ed (1563)
2V

one finds that


µ = − Ed (1564)
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 (1565)

and the chemical potential is found as

N d λ3
 
µ = kB T ln (1566)
2V
——————————————————————————————————

12.1.4 Exercise 59
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 (1567)
nd ! ( Nd − nd )!
Hence, show that
Nd
nd =   (1568)
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 (1569)
nd λ
——————————————————————————————————

352
12.2 Transport Properties
Transport in doped semiconductors is mainly due to scattering from donor impu-
rities. The potential due to the isolated donor impurities is a screened Coulomb
interaction
Z e2
 
V (r) = − exp − kT F r (1570)
εr
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 )
(1571)
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 ε ( 1 − cos θ + ( kTkF )2 )
(1572)
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  
7
∼ β exp β µ dk k exp − k
2 me
(1573)
On changing the variable of integration from k to the dimensionless variable
h̄2 k 2
x = (1574)
2 me kB T
one finds that the temperature dependence of the conductivity is governed by
 
3
σx,x ∼ ( kB T ) exp β µ (1575)

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

353
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 positions.

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 (1576)
where ui is the deviation from the equilibrium position Ri . It is assumed that
the total energy can be formulated as a constant plus pairwise 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 ) (1577)
2 i,j

The Hamiltonian governing the motion of the ions is just given by the sum of
the ionic kinetic energies of the atoms of mass M and the pair-wise interactions
X P̂ 2
i
Ĥ = + V̂ (1578)
i
2 M

The harmonic approximation assumes that 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
(1579)

354
However, since the Ri are equilibrium positions, and the total force on the atom
located at Ri is given by X
∇ Θ(Ri − Rj ) (1580)
j

the total force must vanish in equilibrium. Hence, the potential to second order
in u is just given by
V̂ = Veq + V̂harmonic (1581)
where the harmonic potential is given by
1 X X µ
V̂harmonic = (u − uµj ) Θνµ (Ri − Rj ) (uν,i − uν,j ) (1582)
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 ) (1583)
u≡0

The harmonic potential V̂harmonic is usually expressed directly in terms of the


displacements, and not their differences
1 X X µ ν
V̂harmonic = u D (R − Rj ) uν,j (1584)
2 i,j µ,ν i µ i

where the dynamical matrix is given by


X
Dµν (Ri − Rj ) = δRi ,Rj Θνµ (Ri − R”) − Θνµ (Ri − Rj ) (1585)
R”

In general, when the interactions are not just pairwise, the harmonic potential
can still be defined via the second derivative of the total energy
∂2E
Dµν (R − R0 ) = (1586)
∂uµi ∂uν,j

For a solid with a mono-atomic Bravais lattice, the dynamical matrix Dµν (R−R0 )
possesses the symmetry

Dµν (R − R0 ) = Dνµ (R0 − R) (1587)

due to the analyticity of the pair potential. Also, one has

Dµν (R − R0 ) = Dµν (R0 − R) (1588)

which follows as every Bravais Lattice has inversion symmetry. Due to transla-
tional invariance, one has the sum rule
X
Dµν (R) = 0 (1589)
R

355
This follows from consideration of the absence of energy change due to a uniform
displacement of the solid

r(Ri ) = Ri + ∆ ∀i (1590)

Under this displacement, the energy change of the solid is zero, and so
1 X X µ ν
0 = ∆ Dµ (Ri − Rj ) ∆ν (1591)
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 (1592)

and
[ ui,µ , uj,ν ] = [ P̂i,µ , P̂j,ν ] = 0 (1593)

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 (1594)
N q

and the momentum operator becomes


 
1 X
P̂ i = √ p̂q exp i q . Ri (1595)
N q

where N = N1 N2 N3 is the number of unit cells in the crystal. On assuming


that the lattice displacements satisfy Born-von Karman boundary conditions,
the momenta are quantized by
n1 n2 n3
q = b + b + b (1596)
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 + q 0 ) δµ,ν (1597)

and
[ uq,µ , uq0 ,ν ] = [ p̂q,µ , p̂q0 ,ν ] = 0 (1598)

356
The quantity ∆(q) is the Dirac delta function, modulo reciprocal lattice vectors,
and is defined via  
X
exp i q . R = N ∆(q) (1599)
Ri

This is non-zero if q = Q is a reciprocal lattice vector, 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 (1600)
q
2M 2

where D(q) is the Fourier Transform of the dynamical matrix


X  
D(q)
e = D(R)
e exp − iq.R (1601)
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(Ri − Rj ) exp − i q . ( Ri − Rj )
e
i,j
    !
1 X e
= D(R) exp − iq.R + exp + iq.R − 2
2
R
q.R
X  
= −2 D(R)
e sin2 (1602)
2
R

Thus, D(q)
e is a real symmetric matrix. Every real 3 × 3 symmetric matrix has
three real eigenvalues, thus, one may find three eigenfunctions

D(q)
e α (q) = M ωα2 (q) α (q) (1603)

where α (q) are the eigenvectors and the eigenvalues have been written as
M ωα2 (q). It is necessary that the eigenvalues are positive for the lattice to
be stable.

The eigenvectors are usually normalized, such that

α (q) . β (q) = δα,β (1604)

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) (1605)
α

357
and X
p̂q = P̂qα α (q) (1606)
α

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 (1607)
α

and also X
p̂†q . p̂q = P̂q† α
P̂qα (1608)
α

Hence, the Hamiltonian is diagonal in the polarization indices α,

P̂q† α
P̂qα
" #
X M ωα2 (q) † α
Ĥ = + Qq Qα
q (1609)
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 (1610)
2 h̄ 2 h̄ M ωα (q) q

and the creation operator is the Hermitean conjugate of the annihilation oper-
ator s s
M ω α (q) 1
a†q,α = Q†q α − i P† α (1611)
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 − q 0 ) δα,β (1612)

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 = aq,α aq,α + aq,α aq,α (1613)
q,α
2

358
Due to the commutation relations, one can show that the operator a†q,α when
acting on an energy eigenstate produces an energy eigenstate in which the en-
ergy eigenvalue is increased by an amount h̄ ωα (q). Likewise, the annihilation
operator acting on an energy eigenstate 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 (1614)

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 + (1615)
q,α
2

This implies that each normal mode (q, α) has quantized excitations. These
quantized lattice vibrations are known as phonons.

The completeness relation for the polarization vectors is just

α,µ (q) . α,ν (q) = δµ,ν (1616)

The original displacements and momenta can be expressed in terms of the


phonon creation and annihilation operators via
s    
1 X h̄
ui = √ α (q) aq,α + a†−q,α exp i q . Ri
N q,α 2 M ωα (q)

(1617)

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

(1618)

In the Heisenberg representation, the displacements and momentum operators


become time dependent. The time dependence occurs through factors of
 
exp ± i ωα (q) t (1619)

appearing along with the phonon creation and annihilation operators.

359
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
lattice becomes unstable. The discussion will now 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. The displacements are labelled by the lattice vector index
i and also by an index l 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 l
uli = √ Qq α (q) exp i q . Ri (1620)
N q α

The polarization vectors lα (q) satisfy the generalized ortho-normality condition

l=p
X
lβ (q)∗ . lα (q) Ml = δα,β (1621)
l=1

where Ml is the mass of the l-th atom in the basis.

14.2 A Sum Rule for the Dispersion Relations


The eigenvalue equation

D(q)
e α (q) = M ωα2 (q) α (q) (1622)

can be related to the Fourier transform of the pair potential. The Fourier
Transform of the dynamical matrix is given by
X  
D(q)
e = D(R)
e exp − i q . R (1623)
R

360
Hence, one has
X  
D(R)
e exp − iq.R α (q) = M ωα2 (q) α (q) (1624)
R

The dynamical operator is given in terms of the pair potential


X
Dµν (Ri − Rj ) = δRi ,Rj ∇µ ∇ν Θ(Ri − R”) − ∇µ ∇ν Θ(Ri − Rj )
R”

(1625)

and the Fourier Transform of the pair potential is given by


X  
Θ(R) = exp i k . R Θ(k) (1626)
k

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 − Rj )
Rj k
X  
− kµ k ν Θ(k) exp i k . ( Ri − R” ) (1627)
R”

Thus, the phonon frequencies satisfy the eigenvalue equation


X X  
M ωα2 (q) α (q) = k Θ(k) exp i ( k − q ) . ( Ri − Rj ) k . α (q)
j k
X X  
− k Θ(k) exp i k . ( Ri − Rj ) k . α (q)
j k

(1628)

On making use of conservation of momentum modulo the reciprocal lattice vec-


tors 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) ) (1629)
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

361
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) (1630)

which are the longitudinal sound modes. Hence, the transverse modes only ex-
its 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. This
provides an example of the Goldstone mode which occurs because the continu-
ous translational symmetry of the Hamiltonian is spontaneously broken at the
phase transition when the solid is formed. Goldstone’s theorem may be roughly
stated as ”When a continuous symmetry of the Hamiltonian is spontaneously
broken in a phase transition, a continuous branch of normal modes appears
which extend to zero energy that dynamically restore the broken symmetry”.
Goldstone’s theorem also depends on the condition that long range forces are
not present. If long range forces are present the symmetry restoring mode may
acquire a finite frequency at q = 0 through the Kibble-Higgs mechanism. In
this case the resulting modes are called Higgs bosons.

The Sum Rule.

The Sum Rule is obtained by using the orthogonality relations for the po-
larization vectors. On taking the scalar product of the eigenvalue equation for
the dispersion relation with an eigenvector and summing over the polarizations,
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

(1631)

Utilizing
α,µ (q) . α,ν (q) = δµ,ν (1632)
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 (1633)
Q

for the phonon modes.

——————————————————————————————————

362
14.2.1 Exercise 60
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) = (1634)
α
M V

which defines the plasmon frequency for the ions. In the long wave length 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  
2
D(q) = − 2
e D(R) sin
e (1635)
2
R

as  2
1 X e
D(q)
e = − D(R) q . R (1636)
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 termi-
nology of longitudinal and transverse polarizations in the vicinity of the high
symmetry points.

The solid 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.

363
The phonon density of states is given by an integral over the first Brillouin
zone, and a summation over the polarization index α.
X Z d3 q
ρ(ω) = V δ( ω − ωα (q) ) (1637)
α
( 2 π )3

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
X Z d2 S 1
ρ(ω) = V 3
(1638)
α Sα (ω) ( 2 π ) | ∇ωα (q) |

The van Hove singularities occur when the group velocity vanishes,

∇ωα (q) = 0 (1639)

The van Hove singularities are usually integrable in three dimensions, but still
give rise to anomalous slopes or discontinuities in the derivatives of ρ(ω). An
example of the van Hove singularities in the phonon modes is given by Al which
has two transverse and one longitudinal contribution to the density of states.

——————————————————————————————————

14.3.1 Exercise 61
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.

——————————————————————————————————

14.3.2 Exercise 62
Consider a one-dimensional line of ions, with equal mass, but alternating charges,
such that the charge on the p-th ion is

ep = e ( − 1 )p (1640)

Assume that the inter-atomic potential has two contributions:-

(A) A short ranged force between nearest neighbors with a force constant
C1 = γ.

364
e2
(B) A Coulomb interaction between all the ions Cp = 2 ( − 1 )p p3 a3
where a is the atomic spacing.

(i) Show that



ω(q)2 ( − 1 )p
 
2 q a
X
= sin + σ 1 − cos q p a (1641)
ω02 2 p=1
p3

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 ξ(3).

1
(iii) Show that the speed of sound becomes imaginary if σ > 2 ln 2 .

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 63
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 64
(i) Show that the linear chain with nearest neighbor (harmonic) interactions has
a dispersion relation
qa
ω(q) = ω0 | sin | (1642)
2
and that the density of states is given by
2 1
ρ(ω) = p (1643)
πa 2
ω0 − ω 2
which has a van Hove singularity at ω = ω0 .

(ii) Show that in three dimensions the van Hove singularities near a maxi-
mum of ωα (q) gives rise to a term in the density of states that varies as
q
ρ(ω) ∝ ω02 − ω 2 (1644)

365
and, thus, has a singularity in the first derivative of the density of states with
respect to ω.

——————————————————————————————————

14.3.5 Exercise 65
(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.

(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 laying within the plane.

——————————————————————————————————

14.3.6 Exercise 66
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 (1645)
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 (1646)

and  
B = 2 Θ”(d) − Θ0 (d)/d (1647)

366
(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 (1648)
M 4
and r
8A + 2B qa
ωt (q) = sin (1649)
M 4

(iii) Find the frequencies when q = √q (1, 1, 1).


3

(iv) Show that when q = √q2 (1, 1, 0) that 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 √ (1650)
M 4 2 M 2 2

and the two transverse modes are


s
8A + 4B qa 2A qa
ωt1 (q) = sin2 √ + sin2 √ (1651)
M 4 2 M 2 2

and s
8A + 2B qa 2A qa
ωt2 (q) = sin2 √ + sin2 √ (1652)
M 4 2 M 2 2

——————————————————————————————————

14.3.7 Exercise 67
Consider a phonon with wave vector along the axis of a cubic crystal. Then
consider the sums in
q.R
X  
D(q)
e = D(R)
e sin2 (1653)
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 (1654)
2kF q̂ . R

367
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.

——————————————————————————————————

14.4 Thermodynamics
A harmonic lattice has an energy given by
 
X 1
Eharmonic = Ecl h̄ ωα (q) nq,α + (1655)
q,α
2

where Ecl is the ground state energy of the lattice in the classical approximation
and free energy F is defined in terms of the partition function as
 
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
q,α
1 − exp[ − β h̄ ωα (q) ]

(1656)

Thus, the free energy is given by


X h̄ ωα (q) X   
F = Ecl + + kB T ln 1 − exp − β h̄ ωα (q) (1657)
q,α
2 q,α

The pressure P is found from the thermodynamic relation

dF = dE − T dS − S dT
dF = − S dT − P dV (1658)

Hence, the pressure is given by


 
∂F
P = −
∂V T,N
1 X ∂h̄ωα (q)
   
∂Ecl
= − −
∂V T,N 2 q,α ∂V T,N

368
X  ∂h̄ωα (q)  1
−   (1659)
∂V T,N
q,α exp β h̄ ωα (q) − 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 = (1660)
V ∂T P
As the equation of state is a relation between pressure, temperature and volume
P = P (T, V ) (1661)
Thus, the infinitesimal derivatives are related by
   
∂P ∂P
dP = dT + dV (1662)
∂T V ∂V T
For a process at constant P , dP = 0, thus
 
∂P
∂T
 
∂V
= −  V (1663)
∂T P ∂P
∂V
T

The bulk modulus, B defined by


 
∂P
B = −V (1664)
∂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. This relation is found by relating the temperature derivative of
the entropy with respect to temperature, at constant volume to the temperature
derivative of the entropy with respect to temperature, at constant pressure. This
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 (1665)
∂T P ∂P T

369
Using the equation of state relating P and V
V = V (T, P ) (1666)
to find    
∂V ∂V
dV = dT + dP (1667)
∂T P ∂P T
Thus, on combining the expression for dS in terms of dV and dT and the
equation for dV results in the expression
          
∂S ∂S ∂V ∂S ∂V
dS = + dT + dP
∂T V ∂V T ∂T P ∂V T ∂P T
(1668)
Thus, one has the relation
       
∂S ∂S ∂S ∂V
= + (1669)
∂T P ∂T V ∂V T ∂T P
∂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
= (1670)
∂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 (1671)
∂P
∂V
T

Thus, if there is to be a difference between CP and CV the phonon frequencies


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
(1672)

370
The specific heat is given by
 
∂S
CV = T
∂T V
!
X ∂N (ωα (q)) N (ωα (q)) + 1
= kB ln
q,α
∂T N (ωα (q))
X ∂N (ωα (q))
= h̄ ωα (q)
q,α
∂T

X  2  
= kB β h̄ ωα (q) N (ωα (q)) N (ωα (q)) + 1
q,α

(1673)

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 (1674)
0

Now some approximate models of the specific heat of the lattice vibrations shall
be examined.

14.4.2 The Einstein Model of a Solid


The Einstein model of a solid considers the phonons to have a constant frequency
ω0 , and is an approximate representation of the optic phonons. The phonon
density of states is given by

ρ(ω) = 3 N δ(ω − ω0 ) (1675)

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 (1676)

This vanishes exponentially at low temperatures, kB T  h̄ ω0 where


 
N (ω0 ) ≈ exp − β h̄ ω0 (1677)

and at high temperatures kB T  h̄ ω0


kB T
N (ω0 ) ≈ (1678)
h̄ ω0
so the specific heat saturates to yield the classical result

lim CV → 3 N kB (1679)
T → ∞

371
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
temperature variation. The Debye model of a solid provides an approximate
description of the low-temperature specific heat of a solid.

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 a an isotropic
solid. The dispersion relations of the phonon modes are represented by the
two-fold degenerate transverse mode

ωT (q) = vT q (1680)

and the singly degenerate longitudinal mode

ωL (q) = vL q (1681)

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) (1682)
( 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) = 4πq (1683)
( 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 ) (1684)
( 2 π 2 ) α vα3

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 ) (1685)
( 2 π ) α vα3
2

372
The value of the Debye frequency is determined by the condition
Z ωD
dω ρD (ω) = 3 N
0
V X ω3
D
= (1686)
( 6 π ) α vα3
2

Using this condition, the Debye density of states is written as


ω2
ρD (ω) = 9 N 3 Θ( ωD − ω )
ωD
(1687)

Thus, the Debye 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 (1688)
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 (1689)
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 (1690)
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
temperature specific heat, 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 (1691)
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 (1692)

373
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 68
Evaluate the integral
Z ∞
exp[ x ]
dx x4  2 (1693)
0
exp[ x ] − 1

needed in the low-temperature limit of the specific heat for the Debye model.

——————————————————————————————————

14.4.5 Exercise 69
Generalize the Debye model to a d-dimensional solid. Determine the high tem-
perature and leading low-temperature variation of the specific heat due to lattice
vibrations.

——————————————————————————————————

14.4.6 Exercise 70
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 (1694)
CV 12 dω ρ(ω)

Also evaluate the moment of the phonon density of states


Z
dω ω 2 ρ(ω) (1695)

in terms of the pair potentials between the ions.

——————————————————————————————————

374
14.4.7 Exercise 71
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 (1696)

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


Lindemann assumed that a lattice melts when the displacements due to lattice
vibrations becomes comparable to the lattice constants. Although this theory
does not address the appropriate mechanism it does give the right order of
magnitude for simple metals and transition metals. It is shall assumed that
melting occurs at a critical value of the ratio
u2i
γc = (1697)
a2i
This determines the estimated melting temperature through
h̄ X 2 nq,α + 1
γc = 2
(1698)
2 M N a q,α ωα (q)

The right hand side can easily be evaluated in two limits, the zero temperature
limit and the high temperature limit. At zero temperature, the relative mean
squared displacement is given by
h̄ X 1
γ = 2
2 M N a q,α vα q
Z
3 h̄ V 1
= d3 q
2 M N v a2 ( 2 π )3 q
3 h̄ V 2
= 2 π qD (1699)
2 M N v a2 ( 2 π )3
Using the relation
3
V qD 3V
N = = (1700)
6 π2 4 π a3
one obtains
  13
1 9 h̄ qD
γ ∼
2 2 π2 M v
h̄ qD kB TD
∼ 0.4 = 0.4 (1701)
M v M v2

375
At high temperatures, the relative mean squared displacement is given by
Z qD
h̄ V 6 kB T
γ = dq q 2
2 M N a2 2 π 2 0 h̄ v 2 q 2
3 V kB T qD
= 2 2
M a 2π 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 (1702)
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 of 16 , independent of the metal.
Of course, one expects that anharmonic effects may become important for large
displacements of the ions from their equilibrium positions.

Mermin-Wagner Theorem.

The Lindemann theory of melting may be extended to provide an example of


the Mermin-Wagner theorem. The Mermin-Wagner theorem states that finite
temperature phase transitions in which a continuous symmetry is spontaneously
broken cannot occur in lower than three dimensions (N.D. Mermin and H. Wag-
ner, Phys. Rev. Letts. 17, 1133 (1966)). Basically, if such a transition occurs
then there should be a branch of Goldstone modes that dynamically restores
the spontaneously broken symmetry. These normal modes produce fluctuations
in the order parameter. In a periodic solid where continuous translational in-
variance 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
u2i ∼ dd q
0 2 M ω(q) h̄ ω(q)
Z qD
∼ dq q d−3 (1703)
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. Likewise,
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 tem-
perature 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

376
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 .

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 (1704)
e
2 i,j i

Now consider the expanded lattice, in which the displacements are given by

ui =  Ri + ũi (1705)

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
(1706)

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
= − (1707)
∂T P ∂T V ∂V T
 
∂F
where the pressure is found from P = − ∂V . Using the expression for the
T
free energy of the lattice one finds that the coefficient of thermal expansion can

377
be written as
!
∂ h̄ ωα (q) ∂N (ωα (q))

1 X
α = − (1708)
B q,α ∂V ∂T

where B is the bulk modulus and N (ω) is the Bose-Einstein distribution func-
tion.

The specific heat can be written as

∂N (ωα (q))
X  
CV = h̄ ωα (q) (1709)
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) = − (1710)
ωα (q) ∂V
α B V
which is a dimensionless ratio of CV . Thus,
!
∂ ln ωα (q)
γα (q) = − (1711)
∂ 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 (1712)
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 = (1713)
CV
For most models γα (q) is roughly independent of T and is a constant.
!
∂ ln ωD
γα (q) ∼ γ = − (1714)
∂ 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 .

378
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 = (1715)
3 V
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 (1716)
∂T 3 V

Hence, the coefficient of thermal expansion for a metal is determined from


 
∂P
∂T
 
∂V
= −  V (1717)
∂T P ∂P
∂V
T

Hence, !
1 2 el
α = γ CVlatt + C (1718)
B 3 V
2
where 3 is the electronic Gruneisen parameter.

14.5 Anharmonicity
The anharmonic interactions give rise to the lifetime of phonons, provide tem-
perature dependent corrections to the phonon dispersion relations. These 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
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,

379
like solitons, and play an extremely important role in determining the actual
mechanical properties of any real solid.

——————————————————————————————————

14.5.1 Exercise 72
The full ionic potential of a mono-atomic Bravais Lattice has the form
1 X X
Veq + uµ (R) Dµ ν (R − R0 ) uν (R0 )
2! 0 µ,ν
R,R
1 X X
+ uµ (R) uν (R0 ) uλ (R”) Dµ ν λ
(R, R0 , R”)
3!
R,R0 ,R” µ,ν,λ

(1719)

where u(R) gives the displacement from the equilibrium position R.

(i) Show that if an expansion is made about the expanded lattice positions
defined by
R = (1 + )R (1720)
then the dynamical matrix is changed to

Dµ,ν (R − R0 ) = Dµ,ν (R − R0 ) +  δDµ,ν (R − R0 ) (1721)

where the change in the dynamical matrix is given by


X
δDµ,ν (R − R0 ) = Dµ ν λ (R, R0 , R”) Rλ ” (1722)
λ,R”

(ii) Show that the Gruneisen parameter is given by

α (q) δ D(q)
e α (q)
γα (q) = 2
(1723)
6 M ωα (q)

——————————————————————————————————

380
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 ) ) (1724)
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 the all the excited states of the lattice. The various contribu-
tions 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 = = (1725)
2 mn 2 mn
and the final energy is given by

P 02 h̄2 k 02
E0 = = (1726)
2 mn 2 mn
The energy transfer from the neutron to the sample is given by

h̄ ω = E − E 0 (1727)

This energy is the energy given to the excited phonon modes


X
h̄ ω = h̄ ωα (q) ( n0q,α − nq,α ) (1728)
q,α

as found from conservation of energy.

Due to the periodic translational invariance of the crystal and the short
range of the interaction, the momentum change of the neutron is given by
X  
0 0 0
p − p = h̄ k − h̄ k = h̄ q nq,α − nq,α + h̄ Q (1729)
q,α

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) to be determined.

381
15.1.1 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 is the angle subtended by the detector to
the target. The inelastic neutron scattering cross-section is given by the Fermi-
Golden rule expression. The Fermi-Golden rule is derived in the interaction
representation, and involves an integral over time of the expression with the
matrix elements
 Z +∞
1 Y Y
2 Re dt0 < nα,q | < k | Ĥint | k 0 > | n0α,q > ×
h̄ −∞ q,α q,α

i t0 i t0
Y     Y 
< n0α,q | < k 0 | exp + Ĥ0 Ĥint exp − Ĥ0 | k > | nα,q >
q,α
h̄ h̄ q,α

(1730)

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 represented byQ the number of quantaQ in each normal
mode and respectively are written as | q,α nα,q > and | q,α n0α,q >. The
dependence on the states of the lattice are suppressed in the following. The in-
tegration over t0 gives rise to an energy conserving delta function. This involves
the matrix elements of the interaction between the neutron and the nuclei in
the solid, but unlike the elastic scattering cross-section, previously derived, the
nuclei may be displaced from their equilibrium positions by ui according to

r(Ri ) = Ri + ui (1731)

Hence, the interaction Hamiltonian is given by


X 2 π h̄2
Ĥint = b δ 3 ( r − R i − ui ) (1732)
i
m n

and the matrix elements between the initial and final states of the neutron,
respectively, labelled by momentum k and k 0 , u,
X 2 π h̄2  
0 0
< k | Ĥint | k > = b exp i ( k − k ) . ( Ri + ui )
i
mn
(1733)

If the displacements ui have sufficiently small magnitudes, compared with the


neutron wave length, or the lattice spacing, the exponential term can be ex-
panded as a series in powers of ui ,
X 2 π h̄2  
0 0
< k | Ĥint | k > ≈ b exp i ( k − k ) . Ri ×
i
mn

382
!
0
× 1 + i ( k − k ) . ui + ....

(1734)

The first term in the expansion has been previously considered in the discussion
of neutron diffraction by a crystalline lattice, and gives rise to Bragg scattering.
The next term gives rise to single phonon scattering, while the higher order
terms represent scattering from multi-phonon excitations. In the interaction
representation, the terms involving the lattice displacements depend on time,
via
i Ĥ0 t0 i Ĥ0 t0
   
0
Ĥint (t ) = exp + Ĥint (0) exp − (1735)
h̄ h̄

Thus, on expressing the lattice displacements in terms of the phonon modes as


s  
0 1 X h̄
ui (t ) = √ α (q) exp i q . Ri ×
N q,α 2 M ωα (q)
 
0 † 0
× aq,α exp[ − i ωα (q) t ] + a−q,α exp[ + i ωα (q) t ]

(1736)

where ωα (q) are the phonon frequencies. The integrals over t0 can be expressed
in terms of energy conserving delta functions
Z ∞
dt0
 0 
it
exp ( E 0 − E ± h̄ ω ) = δ( E 0 − E ± h̄ ω )
−∞ 2 π h̄ h̄
(1737)

On evaluating the matrix elements between the initial and final states of the
lattice, the scattering cross-section is found as the sum of terms
2 " X  2
d2 σ 2 π h̄2
 
k 0 δ( E 0 − E )

∝ 0
b exp i ( k − k ) . R i
dΩ dω k mn
i

 
1 X h̄
+ n−q,α + 1 δ( E 0 − E + h̄ ωα (q) ) ×
N q,α 2 M ωα (q)
X   2
0
( k − k 0 ) . α (q)

×
exp i ( k + q − k ) . R i


i
1 X h̄
+ nq,α δ( E 0 − E − h̄ ωα (q) ) ×
N q,α 2 M ωα (q)
  2
X
exp i ( k + q − k ) . Ri ( k − k 0 ) . α (q)
0

×



i

383
#
+ .....

(1738)

The displacements have been expressed in terms of the normal modes, and nq,α
is just the number of phonons with momentum q 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    
0 0
exp i ( k − k − q ) . Ri = N ∆ k − k − q (1739)
i

modulo Q. Thus, the summation over q can be performed, leading to the second
and third terms involves the absorption or emission of a phonon of wave vector
q = (k − k 0 + Q) where Q is a reciprocal lattice vector. These terms are smaller
than the coherent Bragg terms by a factor of N1 .
2 "
d2 σ 2 π h̄2

k
∝ b N 2 ∆( k − k 0 ) δ( E 0 − E )
dΩ dω k0 mn
X
+ N ∆( k − k 0 − q ) δ( E 0 − E + h̄ ωα (q) ) ×
q,α
  2
h̄ ( k − k 0 ) .  (q)

× nq,α + 1 α

2 M ωα (q)
X
+ N ∆( k − k 0 + q ) δ( E 0 − E − h̄ ωα (q) ) ×
q,α
2
h̄ ( k − k 0 ) .  (q)

× nq,α α

2 M ωα (q)
#
+ .....

(1740)

The inelastic one phonon contributions are coherent as it involves the conserva-
tion of momentum, but has an intensity that is only proportional to N .

The energy of the phonon is given by h̄ ωq,α . The thermal average of the
number of phonons nq,α is given by the Bose Einstein distribution function,
which is a Boltzmann weighted average

N (ωα (q)) = nq,α


n=∞
1 X
= n exp[ − β n h̄ ωα (q) ]
Zq,α n=0

384
1 exp[ − β h̄ ωα (q) ]
= 2
Zq,α

1 − exp[ − β h̄ ωα (q) ]

(1741)

However, the partition function for a single phonon mode Zq,α is given by
n=∞
X
Zq,α = exp[ − β n h̄ ωα (q) ]
n=0
1
=
1 − exp[ − β h̄ ωα (q) ]
(1742)

Hence, the Bose-Einstein distribution function is found to be


1
N (ωα (q)) = (1743)
exp[ β h̄ ωα (q) ] − 1

At low temperatures, the number of thermally activated bosons is small, there-


fore, 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 involves the energy loss by the neutron
beam due to creation of individual phonons has an intensity governed by the
1 + N (ωα (q)) which is almost unity at low temperatures. The rate for inelas-
tic 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)) δ( E − E 0 − h̄ ωα (q) ) + N (ωα (q)) δ( E − E 0 + h̄ ωα (q) )
(1744)
The first term represents a processes in which the neutron loses energy due to
the emission of a phonon, whereas the second term represents a 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 + h̄ ω) N (ω)
= = exp − β h̄ ω (1745)
W (E + h̄ ω → 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 (1746)
Z

385
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 + h̄ ω) = P (E + h̄ ω) W (E + h̄ ω → E) (1747)

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
length through the sample.

In addition to the inelastic one phonon scattering cross-section, there are


two second order terms in u2 , which are cross-terms involving the term second
order in u(t0 ) from the expansion of one factor

2 π h̄2
 
0 0
b exp i ( k − k ) . ( Ri + ui (t ) ) (1748)
mn
and the leading zero-th order term from the other factor. This cross term is
proportional to
 2
1
− ( k − k 0 ) . ui (t0 ) .1
2!
 
1 1 X h̄
= − q ( k − k 0 ) . α (q) ×
2! N ωα (q) ωα0 (q 0 )
q,q 0 ,α,α0 2 M
   
× ( k − k 0 ) . α0 (q 0 ) exp i ( q + q 0 ) . Ri ×
 
× aq,α exp[ − i ωα (q) t0 ] + a†−q,α exp[ + i ωα (q) t0 ] ×
 
× aq0 ,α0 exp[ − i ωα0 (q 0 ) t0 ] + a†−q0 ,α0 exp[ + i ωα0 (−q 0 ) t0 ]

(1749)

The expectation value of both the cross-terms


  2    2
1 0 0 1 0
− ( k − k ) . ui (t ) .1 − 1 . ( k − k ) . ui (0) (1750)
2! 2!
is time independent and is given by
 2  
1 X h̄ 0
− ( k − k ) . α (q) 2 nq,α + 1
N q,α 2 M ωα (q)

(1751)

Hence, it is seen that the second order contribution can be decomposed into
an elastic and an inelastic one phonon contribution. The elastic contribution,

386
involves the scattering intensity from the nuclei, when they are displaced from
their equilibrium positions, either through the zero point fluctuations or through
the effect of a thermal activated population of lattice vibrations. This second
order contribution to the elastic scattering has a form similar to the intensity
of the coherent Bragg peak, and cannot be distinguished from it through ex-
periments. It is expected that inspection of all the even order terms in the
expansion should provide similar contributions, which will modify the intensity
of the observed Bragg scattering peak. These contributions, due to the fluc-
tuations of the nuclei from their equilibrium positions, give rise to a reduction
intensity which is governed by the Debye Waller factor.

15.2 The Debye-Waller Factor


The above second order contribution can be combined with the leading order
term, to give the first two terms of the expansion of the elastic scattering cross-
section, W ,
 2  
1 X h̄ 0
W = 1 − ( k − k ) . α (q) 2 N (ωα (q)) + 1 + ...
N q,α 2 M ωα (q)

(1752)

which can be exponentiated to yield the Debye-Waller factor


2
β h̄ ωα (q)
  
1 X h̄ 0
W = exp − ( k − k ) . α (q) coth
N q,α 2 M ωα (q) 2
(1753)
The Debye-Waller factor reduces the intensity of the Bragg peak. The Debye-
Waller factor also modifies the intensity of the Bragg peak in x-ray scattering.

The effects of the multi-phonon processes, of all order, can be ascertained by


examining the expectation value of the factor in the neutron scattering cross-
section given by
Z ∞
dt0 < k | Ĥint (0) | k 0 > < k 0 | Ĥint (t0 ) | k >
−∞
2 X
2 π h̄2
  
0
= b exp − i ( k − k ) . ( Ri − Rj ) ×
mn i,j
Z ∞    
× dt0 exp − i ( k − k 0 ) . ui (0) exp + i ( k − k 0 ) . uj (t0 )
−∞
(1754)

The energy conserving delta functions have been expressed as an integrals over

387
t0 via,

dt0 i t0
Z  
δ( E 0 − E ± h̄ ω ) = exp ( E 0 − E ± h̄ ω )
−∞ 2 π h̄ h̄
(1755)

and the energies in the exponentials can be expressed in terms of the non-
interacting Hamiltonian. Using this identity, the scattering cross-section can be
represented as a Fourier Transform of the thermal average two time correlation
function
Z ∞  0     
it
dt0 exp ( E − E 0 ) < | exp − i ( k − k 0 ) . ui (0) exp + i ( k − k 0 ) . uj (t0 ) | >
−∞ h̄
(1756)
For harmonic phonons, one can express this correlation function as
Z ∞      
0 0 0 0 0
= dt exp i t ω < | exp − i ( k − k ) . ui (0) exp + i ( k − k ) . uj (t ) | >
−∞
Z ∞     2 
1
= dt0 exp i t0 ω < | ( k − k 0 ) . ( ui (0) − uj (t0 ) )
exp − | >
−∞ 2
Z ∞       
= dt0 exp i t0 ω exp − < | ( k − k 0 ) . ui (0) ( k − k 0 ) . uj (t0 ) | > ×
−∞
  2    2 
1 0 1 0 0
× exp − < | ( k − k ) . ui (0) | > exp − < | ( k − k ) . uj (t ) | >
2 2
(1757)

where h̄ ω = E − E 0 is the energy loss experienced by the neutron. The last


two factors are identified with the Debye-Waller factor, which is given by
  2 
0
W = exp − < | ( k − k ) . ui (0) | >
2
β h̄ ωα (q)
  
1 X h̄
= exp − ( k − k 0 ) . α (q) coth
N q,α 2 M ωα (q) 2

(1758)

The frequency dependent factor can be expanded in terms of the number of


phonons
Z ∞       
dt0 exp i t0 ω exp − < | ( k − k 0 ) . ui (0) ( k − k 0 ) . uj (t0 ) | >
−∞
!
Z ∞     
= dt0 exp i t0 ω 1− < | ( k − k 0 ) . ui (0) ( k − k 0 ) . uj (t0 ) | > +...
−∞

(1759)

388
Thus, all the contributions to the scattering cross-section are reduced in inten-
sity 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 Single Phonon Scattering


The phonon dispersion relation can be inferred from a measurement of the single
phonon scattering peak. 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) (1760)
2 mn 2 mn
and
k = k0 + q + Q (1761)
since ω(q) is periodic with a periodicity of the reciprocal lattice vectors

ωα (q) = ωα (q + Q) (1762)

One can combine the equations as

h̄2 k 2 h̄2 k 02
= + h̄ ωα (k − k 0 ) (1763)
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 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 yield
sharp peaks in the spectrum. With knowledge of the magnitude of the final
momentum k 0 , one can construct k 0 − k, and also E 0 − E and hence find h̄ ωα (q)
for the normal mode. By varying the direction of k 0 and the magnitude of E,
one can map out successive surfaces and hence obtain the dispersion relation.

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)

(1764)

The width of the single phonon peak obtained in experiments have two ori-
gins, one is the experimental resolution and another component is not resolution

389
limited. The second component is due to the lifetime of the phonon τ , which ac-
cording to the energy time uncertainty principle gives rise to an energy width of

τ . 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.

15.4 Multi-Phonon Scattering


Processes in which two phonons are absorption or emitted satisfy the two con-
servation laws
h̄2 k 2 h̄2 k 02
= ± h̄ ωα (q) ± h̄ ωα0 (q 0 ) (1765)
2 mn 2 mn
and
k = k0 ± q ± q0 + Q (1766)
0
Conservation of momentum can be used to express q in terms of q, this gives
rise to the restriction
h̄2 k 2 h̄2 k 02
= ± h̄ ωα (q) ± h̄ ωα0 (k − k 0 ± q) (1767)
2 mn 2 mn

Since there are six quantities k 0 and q, which are undetermined. Even if the
direction of k 0 is fixed there still remains three unknown quantities q, which
produces a continuously varying final neutron energy. Hence, one obtains a
continuous spectrum. A similar analysis of the higher order multi-phonon pro-
cesses also yields a continuous spectrum. Only the one phonon spectrum gives
rise to a single peak.

Thus, in a general scattering experiment, with a specific scattering direction,


the analysis of the scattered neutrons energy provides a spectrum which con-
tains a continuous portion on superimposed with sharp peaks. The spectrum
may show an elastic Bragg peak depending on the magnitude of k and θ, or if
there are different isotopes one may observe incoherent nuclear scattering at zero
energy transfers. 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 function N (ωα (q)), whereas the
one phonon emission process has intensity proportional to [ N (ωα (q)) + 1 ].
Thus, it is usual to measure phonon emission at low temperatures.

——————————————————————————————————

390
15.4.1 Exercise 73
(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
emission.

——————————————————————————————————

15.4.2 Exercise 74
(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 (1768)
−∞ k dω dΩ
——————————————————————————————————

15.4.3 Exercise 75
Consider inelastic neutron scattering from a perfect fluid, described by the
Hamiltonian
X P̂ 2
i
Ĥ0 = (1769)
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 − (1770)
dω dΩ 2 π h̄2 q 2 2 h̄2 q 2 2M

——————————————————————————————————

15.5 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 the
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

391
processes proceed via the same mechanism, the scattering from optical phonons
is called Raman scattering and scattering from acoustic phonons is called Bril-
louin scattering.

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 process
h̄ ω 0 = h̄ ω ± h̄ ωα (q) (1771)
and one phonon absorption or emission process
h̄ k 0 n = h̄ k n ± h̄ q + h̄ Q (1772)
0
In these expressions (k, ω) and (k , 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. It reflects the change in the wavelength of the light
as it enters the solid. The phonon absorption (+) gives rise to the Stoke’s sifted
line, which has an intensity proportional to the number of activated phonons
∝ N (ωα (q)) (1773)
The phonon emission (−) gives rise to the anti-Stoke’s line with an intensity
proportional to
∝ [ 1 + N (ωα (q)) ] (1774)
which has contributions from spontaneous and stimulated emission.

Since the 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 small. Thus, as far as the scattering is
concerned the scattering triangle is almost isosceles. The momentum transfer q
is given by
θ
|q| = 2 n k sin
2
ω θ
= 2nk sin
c 2
(1775)
0
Since the direction of k and k are known from the experimental geometry, the
direction of q can be inferred, 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 (1776)
where vα is the velocity of sound. The magnitude of the phonon’s momentum
is given by
ωα (q) ωn θ
q = = 2 sin (1777)
vα (q) c 2

392
However, the energy of the acoustic phonon energy is equal to the change in
photon energy, ∆ ω,
ωα (q) = ∆ω (1778)
Thus, the velocity of the acoustic phonon is found as
∆ω c θ
vα (q) = csc (1779)
2ω n 2
The experimentally determined spectra has the form of a strong un-scattered
laser line, surrounded by a small Stoke’s line at higher frequencies, and a slightly
more intense anti-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.

393
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) = (1780)
ε(q, ω)

and if φext (q, ω) = 0 one must have ε(q, ω) = 0 for φ(q, ω) 6= 0. In this case,
one has an spontaneous density fluctuations and induced longitudinal current
" #

jL (q, ω) = φ(q, ω) − φext (q, ω)

 

= 1 − ε(q, ω) φ(q, ω)

(1781)

where Poisson’s equation and the continuity condition on the charge density have
been used. Using the definition of the longitudinal conductivity one recovers the
relation
4 π σ(ω)
ε(q, ω) = 1 − (1782)

which together with the Drude expression for the conductivity of a gas of ions
of charge Z | e | and mass M

Z 2 e2 ρions τ 1
σ(ω) = (1783)
M 1 − iωτ
This yields the Drude model for the dielectric constant of the ions, which for
ω  τ1 becomes
4 π Z e2 ρ
ε(q, ω) = 1 − (1784)
M ω2
where the density of ions is given in terms of the electron density ρ via
ρ
(1785)
Z

The condition for plasmon oscillations is given by

ε(q, ω) = 0 (1786)

394
The ionic plasmon frequency man be written in terms of the plasmon frequency
Z m 2
Ω2p = ωp (1787)
M
This corresponds to an unscreened phonon frequency. Since the factor ZMm ∼
1
4000 and h̄ ωp ∼ 10 eV, the unscreened phonon frequency is approximately
1
∼ 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
(1788)
q2
with the dielectric constant of the electron gas. In the Thomas-Fermi approxi-
mation this is given by
k2
εeg (q, ω) = 1 + T2F (1789)
q
Thus, within the Born-Oppenheimer approximation, one obtains the dielectric
constant as
4 π Z e2 ρ
ε(q, ω) = 1 − k2
(1790)
M ( 1 + qT2F ) ω 2
The screened ionic plasmons have frequencies given by

4 π Z e2 ρ
ε(q, ω) = 1 − 2
kT
= 0 (1791)
M (1 + q2
F
) ω2

Thus,
Z m 2 q2
ω2 = ωp 2 (1792)
M q + kT2 F
This is the Bohm-Staver model of the phonon frequency. This model results in
a linear dispersion relation ω(q) ≈ v q, where the velocity v is given by

Z m ωp2
v2 = (1793)
M kT2 F

As the Thomas-Fermi wave vector is given in terms of the Fermi-wave vector by

4 π e2 h̄2 π
≈ (1794)
kT2 F m kF

395
and the electron density is given by

kF3
ρ = (1795)
3 π2

the velocity of sound is related to the Fermi-velocity vF = m kF via

1 Z m 2
v2 = v (1796)
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 constant with the Lindhard expression

4 π e2 d3 k f (E(k + q)) − f (E(k))


Z
εeg (q, ω) = 1 − 2
(1797)
q 4 π 3 E(k + q) − E(k) + h̄ ω

where f (x) is the Fermi-Dirac distribution function. This 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 spectrum by producing kinks or infinities in the
derivative  
∂ω
(1798)
∂q q=2kF

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) (1799)

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) (1800)

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) (1801)

396
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) (1802)

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) (1803)

Then with the definition of the total dielectric constant one has the relation
 
ε(q, ω) = εions (q, ω) + εel (q, ω) − 1 (1804)

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. As the dielectric constant due
to the ions alone is given by
Ω2p
εions (q, ω) = 1 − (1805)
ω2
and the electronic dielectric constant ( at low frequencies ) is given by the
Thomas-Fermi approximation
kT2 F
εel (q, ω) = 1 + (1806)
q2
Hence, the low frequency dielectric constant is given by

kT2 F Ω2p
ε(q, ω) = 1 + − (1807)
q2 ω2
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) = + (1808)
εel (q, ω) εel (q, ω)

one considers only the dressed external potential


φext (q, ω)
φdressed (q, ω) = (1809)
εel (q, ω)

397
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)
(1810)
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) (1811)
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 (1812)

the dressed ionic dielectric constant can be defined by


 
1
εdressed
ions (q, ω) = ε ions (q, ω) + ε el (q, ω) − 1
εel (q, ω)
 
1
= 1 + εions (q, ω) − 1
εel (q, ω)
(1813)
The dressed dielectric constant is calculated as
 
1
εdressed
ions (q, ω) = 1 + 2
kT
εions (q, ω) − 1
1 + q2
F

Ω2p
 
1
= 1 − 2
1 +
kT F ω2
q2
(1814)
This can be written in terms of the phonon dispersion relation ω(q)2

ω(q)2
εdressed
ions (q, ω) = 1 − (1815)
ω2
since the phonon oscillations occur when the dielectric constant vanishes
εdressed
ions (q, ω(q)) = 0 (1816)
By inspection of the dressed dielectric constant the phonon frequency is found
as
q2
ω(q)2 = 2 Ω2 (1817)
q + kT2 F p

398
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 ω2 − ω(q)2
q2
(1818)

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
 

= 2 1 +
kT F + q 2 ω − ω(q)2
2

(1819)

Thus, there is an additional contribution in the effective interaction due to the


screening by the ions. The ω dependence of the interaction is representative that
the effective interaction is not instantaneous but instead is a retarded interac-
tion. The effective interaction between a pair of electrons involves a momentum
transfer q = k − k 0 and energy transfer h̄ ω = E(k) − E(k 0 ). 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.

399
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 screened electron-electron interaction is
X e2
∆E(k) = f (E(k 0 )) < k k 0 | 0 |
| k k0 >
| r − r
k0

e2
X Z Z   
0 3 3 0 0 0
= f (E(k )) d r d r 1 − exp i ( k − k ) . ( r − r )
0
| r − r0 |
k
X 4 π e2
= ∆EH − f (E(k 0 ))
| k − k 0 |2
k0

(1820)

The first term is the Hartree term which is k independent and can be absorbed
into a shift of the chemical potential and the second term is the exchange term
which depends on k. The exchange term affects the quasi-particle dispersion
relation. If the effect of phonon screening is included the exchange term becomes
" #
X
0 4 π e2 h̄2 ω(k − k 0 )2
− f (E(k )) 1+
| k − k 0 |2 + kT2 F ( E(k) − E(k 0 ) )2 − h̄2 ω(k − k 0 )2
k0
(1821)
where the exchange interaction is Thomas-Fermi screened, and there is also a
phonon contribution.

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 µ − E(k) − h̄ω(k − k 0 )
Z
1 0
=− h̄ ω(k − k ) ln

8 π 3 h̄ v(k) | k − k 0 |2 + kT2 F µ − E(k) + h̄ω(k − k 0 )
(1822)

where k 0 lies on the Fermi-surface. Substitution of E(k) = µ immediately


demonstrates 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 approxima-
tion under consideration is given by the Thomas-Fermi quasi-particle theory.
Secondly, when the quasi-particle energy is within h̄ ωD of µ, | E(k) − µ | <
h̄ ωD , the logarithmic term can be expanded in inverse powers of h̄ ω. Then it
is seen that the phonon contribution to the screening produces a change in the
dispersion relation
E T F (k) − µ
E(k) − µ = (1823)
1 + λ

400
where λ is the wave function renormalization due to the phonons and is given
by
d2 S 0 4 π e2
Z
1
λ = (1824)
8π 3 h̄ v(k ) | k − k 0 |2 + kT2 F
0

This has the result that the quasi-particle velocity is given by


1
v(k) = ∇ E(k)

1 1
= ∇ E T F (k)
1 + λ h̄
(1825)

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) (1826)
The coupling can be estimated via

4 π e2 d2 S 0
Z
1
λ < (1827)
kT2 F 8 π h̄ v(k 0 )
3

but
4 π e2 ∂ρ 1
= =
kT2 F ∂µ ρ(µ)
−1
d2 S 0
 Z
= (1828)
4 π 3 h̄ v(k 0 )

Hence, the phonon renormalization factor is usually less than unity

λ < 1 (1829)

Finally, the phonon corrections are negligible for electron energies far from the
Fermi-energy. For example when

| E(k) − µ | > h̄ ωD (1830)

then the dispersion relation suffers only small corrections


 2
TF h̄ ωD
E(k) − µ = E (k) − µ + O (1831)
E(k) − µ

Thus, there has to be a kink in the quasi-particle dispersion relation near the
Fermi-energy.

401
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 = (1832)
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 displacements of the ions
X
Ĥint = ûi . ∇Ri Vions (r) (1833)
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 and in also a
phonon of wave vector q has been excited. Hence,

Em − E0 = E(k − q) + h̄ ω(q) − E(k) (1834)

Thus, one can express the second order correction to the ground state energy
in a phenomenological manner as

X | λq |2
∆2 E = − f (E(k)) ( 1 − f (E(k − q)) )
E(k − q) + h̄ ω(q) − E(k)
k,q
(1835)
where f (x) is the Fermi-function. One can identify an effective electron - elec-
tron 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) = (1836)
δf (E(k)) δf (E(k − q))

Hence,

| λq |2
Vef f (q) = −
E(k) − E(k − q) − h̄ ω(q)
| λq |2

E(k − q) − E(k) − h̄ ω(q)
" #
2
2 h̄ ω(q)
= | λq |
h̄2 ω(q)2 − ( E(k) − E(k − q) )2
(1837)

402
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 (1838)
V q + kT F 2

For small q the coupling constant vanishes linearly with q, since

4 π e2 2 µ
= (1839)
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
(1840)

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 and, thus, the
interactions with the electron gas reduces the current flow and hence increases
the resistivity. The electron-ion interaction is given by
X
Ĥions = V (r − R) (1841)
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 (1842)

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 ) + . . . (1843)
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 ) (1844)
i

403
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 ) (1845)
This is a 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)

(1846)
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)

(1847)
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

(1848)

404
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)

(1849)
Thus, the transport scattering rate can be expressed as
 
1
( k . E ) f (E(k)) 1 − f (E(k)) =
τ
(1850)
 
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)

(1851)
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)

(1852)

405
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)

(1853)

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
(1854)

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 ) (1855)

406
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
(1856)
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)
(1857)
On using    
∂f (E(k))
= − β f (E(k)) 1 − f (E(k)) (1858)
∂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
(1859)
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 (1860)
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 (1861)

407
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
≈ (1862)
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 ) ∝ (1863)
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
∝ (1864)
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)) = (1865)
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.

408
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.

409
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) )

(1866)
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 ) (1867)
∂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))
(1868)
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
(1869)
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)
(1870)
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

410
N V m
= k | V (0) |2 kB T (1871)
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) (1872)

Thus, the conductivity has a temperature dependence given by


 
σx,x ∼ exp β µ (1873)

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

411
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 > (1874)
∂t
The state is decomposed in terms of eigenstates of Ĥ0 , | φn > with energy En
 
X En
|Ψ > = Cn (t) exp − i t | φn > (1875)
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) (1876)
∂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 (1877)
To first order one has
Z t  
i E0 − En
Cn1 (t) = − dt0 < φn | Ĥint | φm > exp i t0 (1878)
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) (1879)
∂t h̄
m6=0

412
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). 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 > (1880)

If the wave function, in the presence of an impurity, is written as the superpo-


sition X
ψα (r) = Cα (k) φk (r) (1881)
k

the energy eigenvalue can be expressed as


X
( Eα − E(k) ) Cα (k) = Cα (k 0 ) < k | Vimp | k 0 > (1882)
k0

If the quantity k0 Cα (k 0 ) < k | Vimp | k 0 > is well defined and non zero
P
function of k, then there exist eigenvalues Eα between every consecutive pairs
of E(k). For a large system, where E(k) are very closely spaced, the eigen-
values form a continuum. These eigenstates correspond to weakly perturbed
Bloch states. On the other hand, if the potential is attractive, and there is a
minimum value of E(k), below which there can be bound state 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 . In this case
one can easily solve for the bound states. The above equations can be solved
writing X
γ = Cα (k 0 ) (1883)
k0

Therefore, one has


Vimp
Cα (k) = γ (1884)
Eα − E(k)
The above two equations leads to the self-consistency condition for the bound
state energy Eα
X Vimp
1 = (1885)
Eα − E(k)
k

which shows that, for an attractive potential, there may be a critical value of
Vimp needed for a bound state to form.

413
A more powerful way of solving the same problem involves use of the one-
particle resolvent Green’s function defined by the operator
Ĝ(z) = ( z − Ĥ )−1 (1886)
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 poles of the Hamiltonian are composed
of the discrete bound states at negative energies and a semi-continuous spectrum
at positive energies, the Green’s function has a branch cut across the real axis,
x = Re z,
  X
< Ψ | Ĝ(x−i) − Ĝ(x+i) | Φ > = 2 π i < Ψ | En > < En | Φ > δ( x − En )
n
(1887)
where | En > is the energy eigenstate corresponding to the energy eigenvalue
En .

The resolvent Green’s function can be obtained by expressing the Hamilto-


nian in terms of the unperturbed Hamiltonian Ĥ0 and the interaction Ĥint ,
Ĥ = Ĥ0 + Ĥint (1888)
Then, the Green’s function
 −1  −1
Ĝ(z) = z − Ĥ = z − Ĥ0 − Ĥint (1889)

can be re-written as
 −1  −1
Ĝ(z) = z − Ĥ0 + z − Ĥ0 Ĥint Ĝ(z) (1890)

which can be expressed in terms of the non-interacting resolvent Green’s func-


tion, Ĝ0 (z), as
Ĝ(z) = Ĝ0 (z) + Ĝ0 (z) Ĥint Ĝ(z) (1891)
The non-interacting resolvent Green’s function is easily evaluated in terms of
the matrix elements between the eigenstates of Ĥ0 .
1 1
< φn | | φm > = < φn | | φm >
z − Ĥ0 z − Em
< φn | φm >
=
z − Em
δn,m
= (1892)
z − Em
which is diagonal. For simultaneous momentum eigenstates the non-interacting
resolvent Green’s function only has the diagonal matrix elements
1 < k0 | k >
< k0 | |k > = (1893)
z − Ĥ0 z − E(k)

414
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) (1894)
where the T-matrix is defined as
 −1
T̂ (z) = Ĥint 1 − Ĝ0 (z) Ĥint (1895)

Thus, the poles of the T-matrix are related to the poles of the Green’s function.
As the matrix elements of Ĥint are independent of k, the matrix elements of the
T-matrix between the Bloch states can be evaluated as
 −1
X Vimp
T̂ (z) = Vimp 1 − (1896)
z − E(k)
k

Since the imaginary part of the trace of the Green’s function is related to the
density of state via
1 X
ρ0 (E) = − Im < k | G0 (E + i) |k > (1897)
π
k

one can express the real part of the Greens’s function as an integral
Z ∞
X Vimp Vimp ρ0 (E)
= dE (1898)
z − E(k) 0 z − E
k

Thus, the imaginary part of the T-matrix is non-zero for z on the positive real
axis. The T-matrix has isolated poles outside the continuum at the negative
energy z which are given by the solutions of
Z ∞
1 ρ0 (E)
= dE (1899)
Vimp 0 z − E

The minimum value of the attractive 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
z = 0 Z ∞
1 ρ0 (E)
= − dE (1900)
Vc 0 E
d−2
Since ρ0 (E) ∝ E 2 , the integral converges for three dimensions and higher,
but diverges for two and one dimensions. The bound states are exponentially
localized around the impurity site.

415
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 > (1901)

The can be re-expressed as an integral equation with an initial state given by


incident momentum eigenstate | p > as

( E − Ĥ0 + i ) | Ψ > = V̂imp | Ψ > (1902)

has general solutions which are the superposition of the solutions of the homo-
geneous equation and a particular solution of the inhomogeneous equation
1
|Ψ > = |k > + V̂imp | Ψ > (1903)
E − Ĥ0 + i
To ensure that | Ψ > − | k > is an outgoing wave  must be chosen as a
positive infinitesimal term. The asymptotic behavior of the outgoing equation
can be expressed as
 
1
Ψ(r) = 3 exp i k . r
( 2 π h̄ ) 2
 
0 0
exp i k . ( r − r )
X Z d3 r 0
+ 3 0 Vimp (r0 ) Ψ(r0 )
0
( 2 π h̄ ) E − E(k ) + i 
k

(1904)

which has the well known solution


 
1
Ψ(r) = 3 exp i k . r
( 2 π h̄ ) 2
 
0
exp i k | r − r |
d3 r 0
Z
m
− Vimp (r0 ) Ψ(r0 )
2π ( 2 π h̄ )3 | r − r0 |
(1905)

The asymptotic form can be expressed in terms of the phase shifts δl (k) via a
partial wave analysis.
∞  
X (2l + 1) l lπ
lim Ψ(r) ∼ i exp i δl Pl (cos θ) sin( k r − + δl )
r → ∞ kr 2
l=0
(1906)

Since the T-matrix has matrix elements which satisfy

< k 0 | T̂ | k > = < k 0 | V̂imp | Ψ > (1907)

416
This leads to
 
∞ exp i 2 δl − 1
m X
< k 0 | T̂ | k > = Pl (cos θ) (1908)
2π 2ik
l=0

and the differential scattering cross-section is given by


2
dσ m
< k 0 | T̂ | k >

= (1909)
dΩ 2π

In the limit k → 0 only the s-wave phase shift δ0 is significant and one finds
 
exp i 2 δ0 − 1
m
T = (1910)
2π 2ik
and the scattering cross-section is given by
dσ sin2 δ0
= (1911)
dΩ k2
and the total cross-section σ is given by
4 π sin2 δ0
σ = (1912)
k2
The density of states due to the impurity can be expressed in terms of the phase
shift δ0 (k). The number of states is evaluated in a volume of radius R, and the
wave function is required to vanish at r = R. Hence, the phase shift must
satisfy the condition
k R + δ0 (k) = n π (1913)
Since successive states satisfy this condition with consecutive integers n and
n + 1 then the change in k between two states is given by
∂δ0
∆k ( R + ) = π (1914)
∂k
Thus, the number of states per k interval is
 
1 1 dδ0 (k)
= R + (1915)
∆k π dk
On integrating this with respect to E one finds that the number of states due
to the impurity with energy less than E, N (E) is given by
1
N (E) = δ0 (E) (1916)
π
The impurity density of states, per spin is given by
 
1 ∂δ0
ρimp (E) = (1917)
π ∂E

417
The condition for electrical neutrality for a charge Z | e | determines the phase
shift at the Fermi-energy, through Freidel’s sum rule
2
Z = δ0 (µ) (1918)
π
For systems were the phase shift is rapidly varying at the Fermi-energy, there is
a large impurity density of states. For resonant scattering, as Friedel has shown,
there exists a virtual bound state at the Fermi-energy. This finite density of
states at the Fermi-energy gives rise to an impurity contribution to the specific
heat and the susceptibility.

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 3 d 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) = (1919)
2 m r2
which prevents the 3 d states from being filled until after the 4 s states are filled
or, in the case of the lanthanide elements, the 4 f states remain unfilled until
after the 6 s, 5 p and 5 d states are all occupied. When the nuclear potential is
strong enough, such that the 3 d or 4 f 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, although the 4 f wave function is localized, in
that it has a spatial extent of 0.7 a.u. which lies inside the core like 5 s and 5 p
orbitals, its’ ionization energy is small and comparable to the ionization energy
of the band like 6 s and 6 p orbitals. As the localized state is degenerate with
the conduction band states, there is a finite probability amplitude for tunnelling
through the barrier. The virtual bound state describes an extended state that,
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 (1920)
where Ĥc describes the electrons in th conduction band, the Hamiltonian Ĥd
represents the (isolated) localized d level on the impurity and the term ĤV
describes the coupling. The conduction band is expressed in terms of the number
of conduction electrons in the Bloch states (k, σ) with dispersion relation E(k)
through
X
Ĥc = E(k) n̂k,σ
k,σ

E(k) c†k,σ ck,σ


X
= (1921)
k,σ

418
where c†k,σ and ck,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 (1922)
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 term
 
1 X † ∗ †
ĤV = √ V (k) ck,σ dσ + V (k) dσ ck,σ (1923)
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 ant-commute with all the local operators.

The Resolvent Green’s function can be calculated from the expression


1 1
( z − Ĥ0 ) = ĤV (1924)
z − Ĥ z − Ĥ

Evaluating the matrix elements of this equation between the eigenstates of Ĥ0
yields the coupled equations
1 1 X 1
( z − Ed ) < d | |d > = √ V (k) < d | |k >
z − Ĥ N k z − Ĥ
(1925)

and
1 1 1
( z − E(k) ) < d | |k > = √ V (k)∗ < d | |d >
z − Ĥ N z − Ĥ
(1926)

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
= (1927)
z − Ed − Σ(z)

419
where the self-energy Σ(z) is given by

1 X | V (k) |2
Σ(z) = (1928)
N z − E(k)
k

the real part of the self energy can be interpreted as producing a renormalization
of the energy of the localized level Ed , and the imaginary part can be interpreted
as giving rise to an width or lifetime τ such that

= Im Σ(Ed ) (1929)

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
= +
z − E(k)
1 V (k) V (k 0 )∗
+ Gdd (z)
N z − E(k) z − E(k 0 )
(1930)

From these equations it can be seen that the density of states of the d level is
given in terms of the imaginary part of Σ(E) via
 
1
ρd (E) = Im Gd,d (E − i)
π
1 Im Σ(E − i)
= 2 2
π
 
E − Ed − Re Σ(E − i) + Im Σ(E − i)

(1931)

The impurity density of states is approximately in the form of a Lorentzian


centered on Ed with a width given by Im Σ(E). The width is given by
π X
ImΣ(E) = | V (k) |2 δ( E − E(k) )
N
k
1
≈ π | V |2 ρ(Ed ) (1932)
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 ρ(E). Thus,
the virtual bound state can be interpreted in terms of a narrow band density of
states which is connected to the extended conduction band states.

420
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 is expected to be representative
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 (1933)
in which H0 describes the tight-binding bands of a pure metal with a dispersion
relation
Xd
E(k) = − t cos ki ai (1934)
i=1
and the randomness appears as a shift of the binding energies of the atomic
orbitals X
V̂ = ER | ψ R > < ψ R | (1935)
R

where 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 (1936)

which can be expressed as


 −1
G(z) = z − Ĥ0 − Σ(z) (1937)

where the operator Σ(z) is complex and is known as a self-energy. Since the con-
figurational averaged Green’s function has translational invariance, then so does
the self-energy. It represents the effect of the randomly distributed impurities
on the eigenvalue spectrum. Due to the fluctuations in the random potential,
the energy eigenvalues may 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 av-
erage. For strongly fluctuating potentials the resulting power series may be

421
slowly convergent, or it may not even be convergent at all. Therefore, 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 potential due to the impurities and
the self energy can be expressed as
V̂ (z) = V̂ − Σ(z) (1938)
The resolvent Green’s function for this type of disordered impurity problem can
be expressed as
 −1
Ĝ(z) = z − Ĥ0 − Σ(z) − V̂ (z) (1939)

which can be expressed in terms of the T-matrix via


Ĝ(z) = G(z) + G(z) T̂ (z) G(z) (1940)
where the T-matrix is given by
 −1
T̂ (z) = V̂ (z) 1 − G(z) V̂ (z) (1941)

On taking the configurational average one finds that the averaged T-matrix
must be zero
 −1
T (z) = V̂ (z) 1 − G(z) V̂ (z) = 0 (1942)

This equation can be used to obtain the self-energy.

For the A − B alloy the effective potential is


X
V̂ (z) = ( ER − Σ(z) ) | ψR > < ψR | (1943)
R

It is assumed that the concentration of A atoms is c and the concentration of B


atoms is ( 1 − c ) and these are randomly distributed. 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 averaged 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 >

(1944)

422
The Coherent Potential Approximation (C.P.A.) sets

T (z) = 0 (1945)

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) − E(k)
k
= < R0 | G0 ( z − Σ(z) ) | R0 > (1946)

Nevertheless, this can be solved numerically or alternatively if the sum over


Bloch energies can be evaluated analytically, a analytic solution may be found.

The C.P.A. is expected to be valid in the limit of a dilute concentration of


impurities 1  c, weak scattering t  | EA − EB | and fortuitously 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 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,
reflecting the increasing separation between sites which decreases the tendency
to form bands. The effects of the impurity scattering is to produce a smearing,
which washes out any structure such as van Hove singularities. Transport prop-
erties crucially depend on the spatial extended nature of the energy eigenstates,
which may be destroyed by long-ranged correlations in the random potentials.
This type of phenomenon is completely absent in C.P.A., and can lead to the
energy eigenstates becoming localized.

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 below the density of
states. A distribution in the spatial separation between impurity atoms, will
smear these discrete bound states. The localized states manifest themselves
as low energy tails to the density of states. As the strength of the disorder is
increased, the number of localized states in the tails of the density of states will

423
increase. One surprising feature is that a sharp energy, the mobility edge, sep-
arates 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
are extra-ordinarily long, and cannot be treated by perturbation methods but
require renormalization group type of approaches.

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
transition. The effect of many-body interactions complicate the physics on the
metallic side of the Anderson transition, where weak localization occurs and
there effects are most marked. On the insulating side of the transition, con-
duction will be still possible but only due to the thermal excitation of electrons
to the itinerant states above the mobility edge, or by thermally assisted tun-
nelling 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 arbitrarily small strengths of disorder. However, this
localization need not show up in experiments if the length scale over which the
states are localized is smaller than the sample size.

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 = (1947)
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 W
1 W W
P (ε) = f or − <ε <
W 2 2
= 0 otherwise
(1948)

424
The degree of disorder is measured by the dimensionless parameter W t.

For sufficiently large W/t the states are expected to all be localized. The
critical value of (W/t)c is expected to be dependent on the dimensionality of
the lattice. For W/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 lo-
calized. There 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) (1949)
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 (1950)

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 density of states of the localized states
is expected to be continuous. Numerical studies show that the wave function
exhibits long-ranged fluctuations close to the critical value of W/t, and appears
to be self similar when viewed at all length scales.

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 Thouless
introduced a number g(L) which describes the sensitivity of energy eigenvalues
on the boundary conditions, for a system with linear dimension L. This number
is defined as the ratio
∆E
g(L) = (1951)
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

425
functions are exponentially localized, it is expected that
 
2L
g(L) ∝ exp − (1952)
ξ(E)

On the other hand, if the wave functions are extended the energy shift due to
the different boundary conditions should be of the order of

(1953)
τ
where τ is the time required for the electron to diffuse to the boundary of
the sample. This diffusion time is essentially independent of 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. The quantity g(L) is also found
to be equal to a conductance

G(L) h
g(L) = (1954)
e2
and is related to the conductivity via

g(L) ∝ Ld−1 σ (1955)

The scaling theory of localization is based upon the length dependence of


g(L), (Abrahams, Anderson, Licciardello and Ramakrishnan). In a scale change
of a d-dimensional system with linear dimension L, 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)] (1956)

where f (x) is a universal scaling function, which only depends on the dimen-
sionality d of the lattice. On considering an infinitesimal scale change
dL
b = 1 + (1957)
L
then one can introduce a scaling function

d ln g(L) ∂f (b, g)/∂b


= = β[g(L)] (1958)
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.

426
The asymptotic forms of β can be found in the limits g → 0 and g → ∞.
In the strongly localized regime g → 0 where the wave function is exponentially
localized one finds that as
 
L
g(L) ∝ exp − 2 (1959)
ξ

then
L
β(g) = − 2 = ln g (1960)
ξ
Thus, β(g) tends to − ∞ as g → 0. In the metallic limit g → ∞, then as σ
is finite and independent of L one has

β(g) = ( d − 2 ) (1961)

From this one finds that the system is localized for all spatial dimensionalities
less or equal to two, d < 2, as β(g) is always negative and, thus, on increasing
L g(L) scales to zero. Thus, in the limit of a large system, no matter how
weak the randomness is, the states are always localized in two dimensions. In
two dimensions the conductivity decreases with increasing L, logarithmically
at small values of g and exponentially at large values of g. Furthermore, for
d > 2, there is a critical value of gc such that for g > gc the system scales
to the metallic limit, β(g) = ( d − 2 ) since as L is increased and when β(g)
is positive then limL → ∞ g(L) → ∞. For g < gc , on increasing L, β(g)
is negative and g(L) scales to zero. From the scaling theory one can infer the
dependence of conductivity on the concentration of impurities, c > c0 , close
to the metal insulator transition

σ = σ0 ( c − c0 )1 (1962)

where the exponent of unity can be exactly obtained via perturbation theory.

427
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 (1963)

where the Hamiltonian Ĥ0 represents the non-interacting virtual bound state

E(k) 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

(1964)

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↑ (1965)

This is the Anderson impurity Hamiltonian is exactly soluble using numerical


renormalization group or Bethe-Ansatz techniques. The mean field solution will
be outlined below.

19.2 Mean Field Approximation


The interaction term Ĥint can be represented in terms of fluctuations about the
average value
∆n̂σ = d†σ dσ − < | d†σ dσ | > (1966)
and the average spin value

nσ = < | d†σ dσ | > (1967)

as X
Ĥint = U ∆n̂↑ ∆n̂↓ + U ∆n̂σ n−σ + U n↑ n↓ (1968)
σ

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↓ (1969)
σ

428
The localized electrons experience an effective spin dependent binding energy
given by X
Ĥd = ( Ed + U n−σ ) d†σ dσ (1970)
σ

where the average number of electrons in the localized level of spin sigma is
found as an integral over the density of states of the virtual bound state, which
is given by Z ∞
nσ = dε f (ε) ρσd (ε) (1971)
−∞

where  
1 1
ρσd (ε) = Im (1972)
π ε − Ed − U n−σ − Σ(ε)
The self energy can be represented 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 as
1 ∆
ρσd (ε) = 2 (1973)
π ε − Ed − U n−σ + ∆2

Thus, at T = 0 one finds that


 
1 Ed − µ + U n−σ
nσ = cot−1 (1974)
π ∆
where as cot θ is defined on the interval 0 to π and runs between ∞ and − ∞,
then cot−1 x runs over the range from π to 0. These two coupled equations have
to be solved self-consistently. This can be done by changing the variables
µ − Ed
x =

U
y = (1975)

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 )
(1976)

Thee is a non-magnetic solution

n↑ = n↓ = n (1977)

This has a unique solution for 0 < n < 1 given by the solution of

cot π n = ( y n − x ) (1978)

429
corresponding 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 magnetic solutions. These solutions only occur for suffi-
ciently large values of y as can be seen by linearizing the self-consistency equa-
tions in terms of the variable m defined by

nσ = n + σ m (1979)

On equating the first two terms in the expansion in m, one finds

cot π n = (yn − x)
π
2 = y (1980)
sin π n
These can be re-written as
x 1
= ( θ − sin θ )
y 2π
1 1
= ( 1 − cos θ ) (1981)
y 2π
where θ = 2 π n. This leads to the identification of line separating 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 solution occurs
when U is small or when the d level is either almost completely filled or almost
completely empty.

For large y, magnetic solutions are described by


1
nσ = 1 −
πy
1
n−σ = (1982)
π(y − x)

These 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 impurity spin. How-
ever, the spin rotational invariance has been specifically broken by the mean field
approximation through the choice of a specific quantization axis.

Thus, the mean field solution of the Anderson model contains magnetic and
non-magnetic solutions. 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 state.

430
19.2.1 The Atomic Limit
In the case, when the hybridization is set to zero, the d orbital is entirely
localized. 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. This has energy 2 Ed + U .
The excitation energy required to put an additional particle in the d shell is,
therefore, either Ed or Ed + U depending on whether the initial d state of the
impurity is unoccupied or singly occupied. The ladder of excitations between
these four states are described by the four operators

d†σ ( 1 − d†−σ d−σ )


d†σ d†−σ d−σ
(1983)

The first two take the system from the non-degenerate vacuum state to the dou-
bly degenerate singly occupied states, and the second pair of operators take the
system from the doubly degenerate singly occupied state to the non-degenerate
doubly occupied state.

19.3 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
model of a localized magnetic moment by the Schrieffer-Wolf transformation.
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 (1984)

where Ĥ0 describes the ionic d states and the conduction band states.

E(k) c†k,σ ck,σ +


X X
Ĥ0 = Ed d†σ dσ
k,σ σ

+ U d†↑ d†↓ d↓ d↑
(1985)

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 = (1986)
k,σ

431
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 (1987)

where Ŝ is an anti-Hermitean operator. That is the operator Ŝ satisfies

Ŝ † = − Ŝ (1988)

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 to
 
0
| Ψ > = exp + Ŝ | Ψ > (1989)

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 of 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 ] ] + . . .
2!
(1990)

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 (1991)

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 > = (1992)
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 E(k) − Ed or E(k) −
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

432
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 E(k) − Ed
k,σ

c†k,σ dσ
 
+ V (k) d†−σ d−σ
E(k) − Ed − U
d†σ ck,σ
 
∗ †
− V (k) 1 − d−σ d−σ
E(k) − Ed
#
d†σ ck,σ

∗ †
− V (k) d−σ d−σ
E(k) − Ed − U
(1993)

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 ] + . . .
2!
(1994)

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
X †
α
σ̂k,k 0 = ck,δ < δ | σ α | γ > ck0 ,γ (1995)
δ,γ

and likewise for the local operators


X †
α
Ŝk,k 0 = dδ < δ | σ α | γ > dγ (1996)
δ,γ

one finds that in addition to a potential scattering term there is also an in-
teraction between the components of the spin density operators. The spin-flip
contribution of the interaction is of the form
"   
0 1 X 1 1
Ĥspin−f lip = +
2 0
k,k ,σ
E(k 0 ) − Ed E(k 0 ) − Ed − U
 
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

(1997)

433
To describe scattering of electrons close to the Fermi-energy one may set E(k) =
E(k 0 ), then the effective exchange interaction has the strength
 " #
0 ∗ 1 1
Jk,k0 = Re V (k) V (k ) +
Ed − E(k) U + Ed − E(k)
(1998)

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. 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 whenever 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 J is positive, and the ferromagnetic
interaction attempts to polarize the conduction electron spin density to be par-
allel to the local spin density.

19.3.1 The Kondo Hamiltonian


The resulting Hamiltonian is the Kondo Hamiltonian, it contains an interac-
tion between the localized magnetic moment and the spins of the conduction
electrons. The Hamiltonian can be expressed as

Ĥ = Ĥ0 + Ĥint (1999)

where Ĥ0 represents the Hamiltonian for the conduction electrons

E(k) c†k,σ ck,σ


X
Ĥ0 = (2000)
k,σ

and the interaction is given by

Ĥint = − J S . σ(0) (2001)

where S is a local moment and σ(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
σ α (0) = (2002)
N
k,k0 ;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

434
spin density

Ŝ ± = Ŝ x ± i Ŝ y
σ̂ ± = σ̂ x ± i σ̂ y
(2003)

with the aid of the identity


1
Ŝ . σ̂ = Ŝ z σ̂ z + ( Ŝ + σ̂ − + Ŝ − σ̂ + ) (2004)
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,↓ ck ,↑ + S ck,↑ ck ,↓
0 0 (2005)

19.4 The Resistance Minimum


The Kondo effect 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 flips
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) (2006)

where c is the concentration of impurities. Then, the resistivity shows a mini-


mum at a concentration temperature

ρ1 J ρ(µ) S(S + 1) 1 1
Tmin = ( )5 c5 (2007)
5b
in agreement with experimental findings.

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 sin electron in second order. The T-matrix will be evaluated on the
energy shell E(k) = E(k 0 ), and E will be set to the ground state energy. To
lowest order, the non-spin flip scattering matrix elements are given by
J z
< k 0 ↑ | T (1) (E + i) |k ↑ > = S (2008)
N
whereas to second order one has four non-zero contributions, two contributions
from the spin flip part ( Ŝ ± ) of the interactions and two contributions from

435
the non spin flip part ( Ŝ z ). The non spin flip part gives rise to a term in
T (2) (E + i) of
 2  z 2 X
J S
0 (2)
< k ↑ | Tzz (E + i) |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 + i 2 2

(2009)
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 + i 2
k1 ,k2

(2010)
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 (E(k 2 ))
=
N 2 E(k) − E(k 2 ) + i
k2
 2  z
2 X
J S f (E(k 1 ))

N 2 E(k 1 ) − E(k) + i
k1
 2  z
2 X
J S 1
=
N 2 E(k) − E(k 1 ) + i
k1

(2011)
The singularity at E(k 1 ) = E(k) 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
< k 0 ↑ | T+− (E + i) | k ↑ > = < k 0 ↑ | Ŝ + c†k ,↓ ck01 ,↑ Ŝ − c†k ,↑ ck02 ,↓ | k ↑ >
2N 1
E − Ĥ 0 + i 2
k ,k 1 2

(2012
and
 2 X
(2) J 1
0
< k ↑ | T−+ (E + i) | k ↑ > = < k 0 ↑ | Ŝ − c†k ck01 ,↓ Ŝ + c†k ck02 ,↑ | k ↑ >
2N 1 ,↑ 2 ,↓
k1 ,k2
E − Ĥ0 + i
(2013

436
respectively. These terms are calculated to be
2 X
Ŝ + Ŝ −

(2) J
< k 0 ↑ | T+− (E + i) | k ↑ > = < k 0 ↑ | c†k ,↓ ck01 ,↑ c†k ,↑ ck02 ,↓ | k ↑ >
2N 1
E − Ĥ0 + i 2
k1 ,k2
 2
J X f (E(k 2 )
= Ŝ + Ŝ −
2N E(k) − E(k 2 ) + i
k2

(2014)

and
2 X
Ŝ − Ŝ +

(2) J
< k 0 ↑ | T−+ (E + i) | k ↑ > = < k 0 ↑ | c†k ,↑ ck1 ,↓
0 c†k ,↓ ck2 ,↑ |
0 k↑>
2N 1
E − Ĥ0 + i 2
k1 ,k2
 2
J X 1 − f (E(k 1 )
= Ŝ − Ŝ +
2N E(k) − E(k 1 ) + i
k1

(2015)

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 2 − ( Ŝ z )2 + Ŝ z

Ŝ − Ŝ + = S 2 − ( Ŝ z )2 − Ŝ z
(2016)

The terms proportional to S 2 − ( Ŝ z )2 combine to yield an analytic contribution


to the T-matrix of
 2
(2) J X 1
< k 0 ↑ | Tsf (E + i) | k ↑ > = ( S 2 − ( Ŝ z )2 )
2N E(k) − E(k 1 ) + i
k1

(2017)

whereas the remaining contribution is proportional to S z and the integration is


divergent at E(k) = E(k 1 ) but the integration is cut off by the Fermi-function.
 2
0 (2) J X 2 f (E(k 1 )) − 1
< k ↑ | Tsf (E + i) | k ↑ > = Ŝ z
2N E(k) − E(k 1 ) + i
k1

(2018)

At finite temperatures, either the Fermi-function acts as a cut-off for the sin-
gularity when the scattered particle is on the Fermi-surface E(k) = µ, or if
the scattered particle is off the Fermi-surface, the excitation energy acts as a

437
cut off. In the latter case, the second order contribution to the real part of the
T-matrix can be evaluated as
 2
0 (2) J z

< k ↑ | Tsf (E + i) | k ↑ > ∼ Ŝ ρ(µ) ln ( E(k) − µ ) ρ(µ)

2N
(2019)

which is divergent when E(k) 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 square of
the T-matrix and is found as
 
1 2π 2 S ( S + 1 )

= ρ(µ) J 1 − 4 J ρ(µ) ln kB T ρ(µ) + . . .

τ h̄ 3
(2020)

which gives the 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 and the T-matrix do not diverge at T = 0. The logarith-
mic dependence found in perturbation theory saturates when all the 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 cal-
culated by various means (Abrikosov). In the ferromagnetic case, where J > 0,
the saturation occurs at a characteristic Kondo energy or Kondo temperature
TK given by


2 J ρ(µ) ln kB TK ρ(µ) = − 1
(2021)

or  
1
kB TK = ρ(µ)−1 exp − (2022)
2 J ρ(µ)
and all the results are finite.

For the case of anti-ferromagnetic coupling, the physics scales to a strong


coupling fixed point (Anderson) so the solution must be obtained by other
means such as Bethe Ansatz (Andrei,Weigmann). 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 , and the specific heat originating from the impurity
changes from a constant value at high temperatures to a low-temperature form
having a linear T dependence. 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-
energy. In fact analysis shows that the magnetic moments are being screened
by a compensating polarization of conduction electrons, and that the cloud and

438
moment form a singlet bound state of binding energy 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 exhibits
properties of the free moments. From the perspective of the Anderson impu-
rity 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-Suhl 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 Nozières.

439
20 Collective Phenomenon
21 Itinerant Magnetism
21.1 Stoner Theory
The Stoner theory of itinerant magnetism examines the stability of a band
of electrons to Coulomb interactions (E.C. Stoner, Rep. Prog. in Phys. 11
43 (1948)). 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 (2023)

The Hamiltonian for the non-interacting electrons in the Bloch states is written
as X
Ĥ0 = E(k) nk,σ (2024)
k,σ

The interaction Hamiltonian is given by


U X
Ĥint = n̂i,σ n̂i,−σ (2025)
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 of 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 (2026)
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

(2027)

440
Thus, in the mean field approximation, the Hamiltonian is given by
X U X
ĤM F = ( E(k) + U n−σ ) nk,σ − N n−σ nσ (2028)
2 σ
k,σ

The single particles have the spin dependent energy eigenvalues

Eσ (k) = E(k) + U n−σ (2029)

The magnetization is given by


 
1
Mz = g µB n+ − n−
2
Z ∞  
= µB d f () ρ( − U n− ) − ρ( − U n+ )
−∞
(2030)

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
become stable when Mz ∼ 0, in which case the equations can be linearized to
yield
  Z ∞

( n+ − n− ) = U n+ − n− d f () ρ() (2031)
−∞ ∂

The ferromagnetic state has the lowest energy when the self-consistency equa-
tion is satisfied. The integral can be performed by integration by parts yielding
Z ∞

1 = U d f () ρ()
−∞ ∂
Z ∞

= −U d ρ() f ()
−∞ ∂
(2032)

At low temperatures, the derivative of the Fermi-function can be replaced by a


delta function at the Fermi-energy.

− f () = δ( − µ) (2033)
∂
This yields the Stoner criterion for ferromagnetism as

1 < U ρ(µ) (2034)

441
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 at T = 0.
The magnetization is given by the solution of the non-linear equation. The non-
linear equation shows that the magnetization 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 then one per atom. In systems
which have bands that are more than half filled, the saturation magnetic mo-
ment is equal to a Bohr magneton per unoccupied state. At finite temperatures,
the value of the magnetization is reduced and disappears at a critical tempera-
ture Tc . Unfortunately, Stoner theory does not predict reasonable values of the
critical temperatures.

In the paramagnetic state Stoner theory predicts that the susceptibility


should be exchange enhanced over the non-interacting susceptibility χ0p via

χ0p
χp = (2035)
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 76
Determine the critical temperature Tc predicted by Stoner theory.

——————————————————————————————————

21.1.2 Exercise 77
Determine the paramagnetic susceptibility by using Stoner theory.

——————————————————————————————————

21.2 Linear Response Theory


The spatially varying magnetization MZ (r) of a paramagnetic system to a spa-
tially varying applied magnetic field Hz (r) are related by the z − z component

442
of the magnetic tensor susceptibility, via the linear relationship
Z
Mz (r) = d3 r0 χz,z (r, r0 ) Hz (r0 ) (2036)

This is a special case of the more general relation


Z X
Mα (r) = d3 r 0 χα,β (r, r0 ) Hβ (r0 ) (2037)
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
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 ) (2038)
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) (2039)

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) (2040)
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)

(2041)
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) (2042)
2
Thus, the charge density is given by two coupled equations
 
g µB
ρσ (q) = χ0 (q) − Hz σ + U ρ−σ (q)
2
(2043)

443
for each spin polarization. In the above expression, χ0 (q) is the Lindhard
density-density response function, per spin. On combining these equations one
finds, the z component of the magnetization produced by a magnetic field ap-
plied along the z direction is
 
g µB 0
Mz (q) = − χ (q) g µB Hz (q) + U Mz (q) (2044)
2

Thus, it is found that the Pauli paramagnetic susceptibility


Mz (q)
χz,z
p (q) =
Hz (q)
g 2 µ2B 2 χ0 (q)
= − (2045)
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
0 (q) = − 2 χ (q) (2046)

instead of the density-density response function χ0 (q). This yields the result

χz,z
2
0 (q)

g µB
χz,z
p (q) = (2047)
2 1 − U
2 χz,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 may have maxima at certain values
of q, say Q, which are determined by the band structure and the occupancy of
the non-interacting bands, through the non-interacting susceptibility χ0 (q). If
the value 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 the anal-
ysis was performed with the z axis being the axis of quantization, a similar
analysis could have been performed in which 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

444
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.

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

Uc z,z
1 = χ (Q) (2048)
2 0
If a non-interacting system is considered which has a maximum at Q = 0 this
reduces to the Stoner criterion for ferromagnetism as

lim χz,z
0 (Q) → 2 ρ(µ) (2049)
Q → 0

where ρ(µ) is the density of states, per spin, at the Fermi-energy. Thus, it is
found the critical value of U is given by the criterion

1 = Uc ρ(µ) (2050)

For values of U larger than the critical value, the paramagnetic state is unstable
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 ( nσ − n−σ ) 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 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 can be achieved without
any finite energy excitations. The zero energy excitations that uniformly rotate
the magnetization in a ferromagnet are the q = 0 Goldstone modes associ-
ated 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. This sus-
ceptibility coupled by an interaction can produce a spin density wave in which

445
the magnetization is modulated with this wave vector.

For tight-binding bands which satisfy the perfect nesting condition

E(k + Q) = − E(k) (2051)

for some Q. The non-interacting susceptibility can be evaluated as an integral


over the density of states
X f (−E(k)) − f (E(k))
χz,z
0 (Q) = 2
2 E(k)
k
Z ∞
1 − 2 f ()
= 2 d ρ()
−∞ 2
(2052)

From this one can see that if the density of states ρ(0) is non-zero, the sus-
ceptibility χ(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 momen-
tum 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
for Q = πa (1, 1, 1) in non-degenerate tight-binding bands on a simple cubic
lattice, where
i=3
X
E(k) = − 2 t cos ki a (2053)
i=1

Since the bands are symmetric around  = 0, the non-interacting susceptibility


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 magnetiza-
tion 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 cos π (2054)
i
a
The staggered magnetization on the neighboring lattice sites is oppositely ori-
ented, and is anti-ferromagnetically ordered. Since anti-ferromagnetic ordering
was first proposed by Louis Neel to describe classical magnets (L. Neel, Ann. de
Physique, 17, 64 (1932), Ann. de Physique 5, 256 (1936)), 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 accompanied by a metal-insulator transition. This process was first
discussed by J.C. Slater. 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 travelling in the solid experience a periodic potential which contains a

446
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 magnetization shows up as a spin dependent contribution to
the potential, and may produce gaps in the electronic dispersion relations at the
surface of the new Brillouin zone. If the magnitude of the spin dependent po-
tential is large enough, a gap may occur all around the Brillouin zone resulting
in a gap in the density of states and hence the insulating state. Such insulat-
ing 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 Waves


The dynamical response of the magnetization to a time and spatially varying
applied magnetic field of wave-vector q and frequency ω is given by the dynam-
ical response χz,z
p (q; ω). The imaginary part of this response function yields the
spectrum of magnetic excitations. The imaginary part of the reduced suscepti-
bility can be measured indirectly by inelastic neutron scattering experiments, in
which the neutrons spin interacts with the electronic spin density via a dipole-
dipole interaction. A simple extension of our previous analysis shows that, in
the mean field approximation,
χz,z
2
0 (q, ω)

g µB
χz,z
p (q; ω) = (2055)
2 1 − U2 χz,z
0 (q; ω)

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,
Im χz,z
2
0 (q, ω)
  
z,z g µB
Im χp (q; ω) = 2 2
2
 
U z,z U z,z
1 − 2 Re χ0 (q; ω) + 2 Im χ0 (q; ω)

(2056)
which on using the approximation
Re χz,z
0 (q; ω) = 2 ρ(µ)
π ω
Im χz,z
0 (q; ω) = ρ(µ)
2 q vF
(2057)
for the Lindhard susceptibility, shows that the system exhibits a continuum of
quasi-elastic magnetic excitations. As the instability is approached the spec-
trum is enhanced at low frequencies. These magnetic excitations are known as

447
paramagnons. The lifetime of the paramagnon excitations and the frequency
of the excitations soften as the value of U is increased to 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 equilibrium (paramagnetic)
state. The existence of large amplitude paramagnon fluctuations not only mani-
fest themselves in the inelastic neutron scattering cross-section which is directly
proportional to Im [ χα,β (q; ω) ] , an enhanced susceptibility but also lead to
logarithmic enhancement of the linear T term in the electronic specific heat
(Berk and Schrieffer, Doniach and Engelsberg), and enhancement in the T 2
term in the electrical resistivity (Lederer and Mills). These characteristics have
been observed in metallic P d (Schindler and Coles).

For values of U greater than Uc , the q = 0 response shows a sharp zero


energy mode that represents the Goldstone mode of the system. In the ferro-
magnetically ordered state, these excitations 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) (2058)

The transverse response is 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) (2059)

The spin flip response functions are

χ+,− (q; ω) = χ−,+ (q; ω) = 2 χx,x (q; ω) (2060)

In the limit q → 0, the non-interacting transverse response is given by



lim χ+,−
0 (q, ω) = U
(2061)
q → 0 h̄ ω + ∆
so the full transverse response function is given by

χ+,−
0 (q, ω)
lim χ+,− (q, ω) = lim
q → 0 q → 0 1 − U χ+,−
0 (q, ω)

=
U h̄ ω
(2062)

Thus, in accordance with Goldstone’s theorem the instability has produced a


dynamic response with a zero-frequency pole. For small q and low frequencies
ω < ∆, the spin waves have a dispersion relation of the form

ω = D q2 (2063)

448
The spin waves of the ferromagnetic state become broader for larger (ω, q) values
where the spin wave branch enters and coexists with the continuum of (Stoner)
spin-flip electron-hole excitations.

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.

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.
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 occurs at T = 0. The critical
fluctuations are not thermally excited and cannot be treated classically but are
zero point fluctuations associated with the existence of a quantum critical point.

21.5 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 at i via a local exchange interaction
X
Ĥint = − J S i . σi (2064)
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 ) (2065)
g µB

This polarization then interacts with the spin at site j via the 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

449
X  2 2
= − J 2 χz,z (Ri − Rj ) S i . S i
i,j
g µB

(2066)

The oscillations are produced by the oscillations of the response function of the
conduction electrons. The Fourier transform of J(R) shows oscillations at 2 kF .
This interaction was discovered independently, by Ruderman and Kittel, Ka-
suya and Yosida.

22 Localized Magnetism
The nearest neighbor Heisenberg exchange interaction couples spins localized
on adjacent lattice sites.
J X
Ĥ = − S(R + δ) . S(R) (2067)
2
R,δ

This interaction Hamiltonian can be derived from the model of itinerant mag-
netism, 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 Anderson. The exchange
constant is given in terms of the tight-binding matrix element t and the Coulomb
repulsion via
t2
J = − (2068)
U

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
(2069)
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 ) > (2070)
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

450
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
(2071)
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 > (2072)
N R

It corresponds to a state with total spin N S − 1 as one spin is flipped over.


This state is a coherent superposition of all states with one spin flipped and has
total momentum q. The excitation energy is found from
 
( Ĥ − Eg ) | q > = 2 J S 3 − cos qx a − cos qy a − cos qz a | q > (2073)

At long wave lengths, the spin wave excitation energy h̄ω(q) is found as
 
h̄ ω(q) = 2 J S 3 − cos qx a − cos qy a − cos qz a

∼ J S q 2 a2
(2074)
which is the branch of collective Goldstone modes which restore the sponta-
neously broken spin rotational invariance of the ferromagnet. The vanishing
of the frequency at q = 0 is a general consequence that the total spin is
a constant of motion, and in this limit the spin wave
 state just
 corresponds to
a reduced value of the total magnetization R exp − i q . R S − (R) → S − .
P

The above excitations of the spin system are small amplitude excitations that
have a close resemblance to the harmonic phonons of the crystalline lattice. The
effects of the interactions could be expected to produce small anharmonic cor-
rections to these excitations, providing them with a lifetime and a renormalized
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

451
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 transformation of the spins represents the effect


of the spin operators by a function of bosons operators (T. Holstein and H.
Primakoff, Phys. Rev. 58, 1098 (1940)). 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 > (2075)

In particular the spin raising and lowering operators

Ŝi± = Six ± i Siy (2076)

have the commutation relations with Ŝ x


 
±
z
Ŝi , Ŝi = ± h̄ Ŝi± (2077)

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 > (2078)

The Holstein Primakoff transformation represents the (2 S + 1 ) basis states


| mi > by the infinite number of basis states of a boson number operator a†i ai .
The boson basis | ni > states are defined through

a†i ai | ni > = ni | ni > (2079)

The relation between the basis spin states and boson states is provided by the
boson representation of the z component of the spin operator

Ŝiz = S − a†i ai (2080)

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 boson
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 effect 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

452
the Hilbert space of the spin system. This is achieved by the representation of
the spin lowering operator
q 
Ŝi− = a†i 2 S − a†i ai

(2081)
which projects out states with more than 2 S bosons present. The raising
operator is given by the Hermitean conjugate
q 
+ †
Ŝi = 2 S − ai ai ai

(2082)
This transformation respects the spin commutation relations.

Given the nearest neighbor ferromagnetic ( J > 0 ) Heisenberg Hamiltonian


J X
Ĥ = − Ŝ(R) . Ŝ(R + δ) (2083)
2
R,δ

this 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

(2084)
On representing this in terms of the boson operators, and expanding 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

(2085)
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 (2086)
N q

and the creation operator is the Hermitean conjugate


 
† 1 X
aR = √ exp − i q . R a†q (2087)
N q

453
Substitution of these, and performing the sums over the spatial index, yields an
approximate expression for the Hamiltonian. The 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

J X X  
2 †
Ĥ = − N Z S + J S aq aq 1 − cos q . δ (2088)
2 q δ

The terms of higher order in S1 yield quantum corrections to the ground state
energy, the spin wave energies and also produce anharmonic interactions be-
tween the spin waves.

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)) (2089)
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) ∼ (2090)
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
J S
(2091)

This should result in the low-temperature specific heat being proportional to


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 = N 12 has the same energy
as the state where all the spins are oriented along the direction (θ, ϕ). The

454
ferromagnetic state where the spins are fully polarized along the z axis is given
by
Y 1
|0 >= | mj = > (2092)
j
2

The state where the polarization is rotated through (θ, ϕ) is given by


Y  θ
 
θ −

| θ, ϕ > = cos + exp i ϕ sin Ŝj |0 > (2093)
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 (2094)

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 operators for the spins at site j is found as
   
 cos θ sin θ exp − i ϕ 
 
 
   
 
 sin θ exp + i ϕ
 − cos θ 

This spin operator has two eigenstates


 
θ θ
| θ, ϕ + > = cos | + > + exp + iϕ sin |− > (2095)
2 2
and
 
θ θ
| θ, ϕ − > = cos | − > − exp − iϕ sin |+ > (2096)
2 2
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) and are in the | θ, ϕ + > spinor state. The rotated ferromagnetic state
has all the spins aligned in the same direction. This classical ferromagnetic state
has an infinitesimal overlap with the un-rotated ferromagnetic state, as
 N  
θ θ
< 0 | θ, ϕ > = cos = exp N ln cos (2097)
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!

455

( ST−ot )n
 
X
(N −n) θ θ
= cos sinn exp inϕ |0 >
n=0
2 2 n!
(2098)

where the total spin operator is given by


X
ST−ot = Ŝj− (2099)
j

since ( Sj− )2 ≡ 0. The number of q = 0 spin waves, n, are distributed with


the binomial probability
 
N θ θ
P (n) = C cos2(N −n) sin2n (2100)
n 2 2

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 un-normalized states with n spin
wave present are found as

( Stot )n
|n > = |0 > (2101)
n!
These states have the normalization
 
N
< n|n >= C (2102)
n

The matrix elements of the total spin lowering operator between states with
total numbers of q = 0 spin waves close to the maximum of the wave packet
are found first by noting that
   
− N
< n + 1 | ŜT ot | n > = N − n C
n
(2103)

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) A(n + 1) < n + 1 | ŜT−ot | n >
n

456
  X  
1 N
∼ N sin θ exp − iϕ A(n)2 C
2 n
n
 
1
= N sin θ exp − iϕ
2
(2104)
and likewise, the matrix elements of the spin raising operator are given by the
complex conjugate. Likewise, the matrix elements of the ŜTz ot is given by
 
z N N
< n | ŜT ot | n > = ( − n) C (2105)
2 n
which for n ∼ nmax = N sin2 θ
2 yields
 
N
< n | ŜTz ot | n > ∼ N cos θ C (2106)
n
Thus, one has
X
< θ, ϕ | STz ot | θ, ϕ > = A(n)2 < n | ŜTz ot | n >
n
 
X
2 N
∼ N cos θ A(n) C
n
n
= N cos θ
(2107)
Thus, the total spin operator has matrix elements between the coherent state
that exactly corresponds to the classical vector.

Thus, the different classical ferromagnetic states are represented by coherent


states which are superpositions of states with arbitrary numbers of Goldstone
modes excited. 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 arbitrary
small magnetic field. In this case the different classical states have zero over-
lap 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 78
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?

457
——————————————————————————————————

22.3 Anti-ferromagnetic Spinwaves


One can perform a similar analysis for an anti-ferromagnet ( J < 0 ) in a
Neel state. Neel ordering 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 (2108)

and the spin raising operators for the B spins are


q 
Ŝi+ = b†i 2 S − b†i bi

(2109)

which projects out states with more than 2 S bosons present. The lowering
operator is given by the Hermitean conjugate
q 
− †
Ŝi = 2 S − bi bi bi

(2110)

for the B sub-lattice.

The Hamiltonian can be written as


X  
Ĥ = − J ( S − a†i ai ) ( S − b†j bj ) − S ( a†i b†j + ai bj ) (2111)
i,j

On defining the Fourier transformed operators


r  
2 X
aq = exp − i q . Ri ai
N i
r  
2 X
bq = exp − i q . Rj bj
N j
(2112)

458
then one has
X 
Ĥ = −N zJ S 2
+ J S z ( a†q aq + b†−q b−q )
q
  
( a†q b†−q + aq b−q )
X
+ exp − iq.δ
δ
(2113)
This form is still not diagonal, it is necessary to use the Bogoliubov canonical
transformation
   
αq = exp + Ŝ aq exp − Ŝ
   
β−q = exp + Ŝ b−q exp − Ŝ

(2114)
where the operator is given by
X θq  
Ŝ = b†−q a†q − aq b−q (2115)
q
2

and θq has still 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
2 2 q
(2116)
in which θq is chosen so that the Hamiltonian is diagonal. The inverse transfor-
mation is given by
θ θ †
aq = cosh αq + sinh β
2 2 −q
θ θ †
b−q = cosh β−q + sinh α
2 2 q
(2117)
This value of θq is found as
 
1 X
tanh θq = − exp − i q . δ (2118)
z
δ

The resulting approximate Hamiltonian again can be interpreted in terms of a


zero point energy and a sum of harmonic normal modes.

——————————————————————————————————

459
22.3.1 Exercise 79
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) (2119)
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 ) (2120)
S
where v
u  Xd 2
2 X u 2
Id = t d − cos qi a (2121)
N d q i=1

The amplitude of the q = 0 spin wave is divergent, but can be neglected in


the limit N → ∞. As the q = 0 spin wave has zero frequency, one can
compose a state which is a superposition of the q = 0 spin wave excitations.
The dynamics of the finite frequency spin wave excitations can be examined for
finite time scales before the zero energy amplitude wave packet of q = 0 spin
waves diverges re-orienting the sub-lattice magnetization.

23 Spin Glasses
Spin glasses are found when magnetic impurities are randomly distributed in a
metal, such as F e in a 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 ferromag-
netic and anti-ferromagnetic signs. The distribution of interactions prevents the
local magnetic moments from forming a long-range ordered phase at low temper-
atures. Nevertheless, the random spin system may freeze into a spin glass state
below a critical temperature. At high temperatures the spins are disordered,

460
and as the temperature is reduced, the spins which are most strongly interacting
progressively build up there 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 have to be large scale
reorientation 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 of the spin under consideration with 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
susceptibility and a rounded peak in the specific heat. The susceptibility follows
a Curie-Weiss law at high temperatures

µ2B S ( S + 1 )
χ(T ) = c (2122)
3 kB ( T − Θ )
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
to c. At lower temperatures 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) (2123)

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 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 percent 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 temperature. The cusp becomes rounded and the temperature of
the peak diminishes as the a.c. frequency is lowered. The susceptibility satu-
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

461
memory effect, in that, in field cooled samples ( H 6= 0 ) the susceptibility sat-
urates at Tf , and the curve χ(T ) is reversible as it is also followed for increasing
temperature at fixed field. However, the zero field cooled sample ( H = 0 )
shows a cusp at Tf and as the susceptibility is zero until the field is applied,
by definition the temperature is only allowed to increase. However, the value
of the susceptibility increases with increasing measurement time. The magne-
tization 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,
by Edwards and Anderson, the formulation of a model which is exactly soluble
mean field theory by Sherrington and Kirkpatrick. The Sherrington-Kirkpatrick
model consists of an Ising interaction
X
Ĥ = − Ji−j Siz . Sjz (2124)
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 (2125)
and the average value of the square is given by

2 J2
< Ji−j >= (2126)
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
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
(2127)

462
where the exchange interactions jij 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 (2128)
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 (2129)
j

as 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 > (2130)
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 law
(Edwards and Jones)
1
q
ρ(Jλ ) = 4 J 2 − Jλ2 (2131)
2 π J2
where, obviously, 2 J is the largest eigenvalue. The spin glass freezing temper-
ature is determined as
kB Tf = 4 J (2132)
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 re-
action term. Namely, the reaction term includes the effect of the central spin on
the neighbors back on itself, before the thermal averaging is performed (Thou-
less, Anderson and Palmer).

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 (2133)
n→0 n

463
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 J S
i−j i S j
i−j
2 π J2 2 J2 α

( β J )2 X X α α β β
 
n
Z = T race exp Si Sj Si Sj
2N i,j α β
(2134)

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 | > (2135)

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 Si 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
α,β
(2136)

In this 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

464
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, and leads to
N
− β F = N ln 2 + ( β J )2 (2137)
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
integral by the saddle point value

( β J )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
α,β

( β J )2
  
= exp nN 1 − 2 q − q 2 (n − 1) ×
2
" #
z2
Z    X
dz p
× exp N ln T race √ exp − exp ( β J ) z 2 q Sα
2π 2 α
( β J )2
  
2
= exp n N 1 − 2 q − q (n − 1) ×
2
" #
z2
Z   
dz n
p
× exp N ln √ exp − 2n cosh (βJ )z 2q
2π 2
(2138)

The saddle point value of q is found by differentiating 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
z2
Z      
dz
coshn ( β J ) z 2 q 1 + (n − 1) tanh2 ( β J ) z 2 q
p p
= √ exp −
2π 2
(2139)

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 (2140)
−∞ 2π 2

465
The temperature variation of the order parameter is given by
  2 
1 T
q(T ) = 1 − f or T < Tf
2 Tf
  12
2 T
lim q(T ) = 1 − (2141)
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 ) (2142)
3 kB T

Although the long-ranged model is exactly soluble in the mean field approxi-
mation, it is only soluble for all temperatures below the freezing temperature if
the symmetry between the different replicas are broken. Replica symmetry is
specific to interacting random systems (Almeida and Thouless), and the exact
solution of the mean field model involves repeated replica symmetry breaking
(Parisi). This repeated replica symmetry breaking has the consequence that
the dynamics of the low-temperature system are frozen and no longer consis-
tent with the ergodic hypothesis.

466
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 (2143)

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 is a dipole field given by
 
ge µB σ e ∧ r |e| v ∧ r
H = ∇ ∧ 3
− (2144)
|r| 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
"  
σe ∧ r
= gn µn σ n . ∇ ∧ ge µB
| r |3
 #
|e| σn ∧ r σn ∧ r
− p. + .p (2145)
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

467
| φn > to the final state | φn0 > is given by
2 X
d2 σ k 0 V mn
  
< φn0 ; k 0 , σN
0 σe ∧ r
= ( ge gn µn µB )2

P (n) | σ N . ∇ ∧
dω dΩ k 2 π h̄2 | r |3
n,n0
  2
1 σn ∧ r σn ∧ r
− p. 3
+ 3
. p | φn ; k, σn > δ( h̄ ω + En − En0 )
2 h̄ |r| |r|
(2146)

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 are defined via

h̄ ω = E(k) − E(k 0 )

h̄ q = h̄ k − h̄ k 0
(2147)

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
V q2
(2148)

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
     
σe ∧ r 4πi
< k 0 | pe ∧ | k > = − σ e ∧ ( q ∧ p ) exp + i q . r e
| r |3 V q2 e

(2149)
 
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 X
d2 σ k 0 2 mn

= ( ge gn µn µB )2 P (n) δ( h̄ ω + En − En0 )
dω dΩ k h̄2 n,n0

468
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 |
(2150)

Since the nuclear Bohr magneton has the value


| e | h̄
µn = (2151)
2 mp c

the coupling constant can be simplified

2 mn gn e2
2 gn ge µn µB = = re (2152)
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) (2153)
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
(2154)

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 (2155)
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

469
X  i q ∧ pe
  
× < φn0 | σe − exp + i q . re | φn >
e
h̄ | q |2 β
(2156)

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 . re (2157)
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
(2158)

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 . r e
e
h̄ | q |2 α

i q ∧ pj
X   X    
= exp − i q . R σj + exp − i q . r j
j
h̄ | q |2 α
R

(2159)

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
X  
Ŝα (q) = exp − i q . R F (q) ŜR
R

(2160)

470
The form factor F (q) is defined as the Fourier transform of the normalized spin
density for the ion. By definition

F (0) = 1 (2161)

In this case, the inelastic neutron scattering spectrum can be expressed as

d2 σ k0
= re2 | F (q) |2 S(q; ω) (2162)
dω dΩ k
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 components of the correlation function measured in a scattering ex-
periment can be defined through the function S α,β (q; ω) where
X
S α,β (q; ω) = P (n) δ( h̄ ω + En − En0 ) ×
n,n0
" #
× < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >

(2163)

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
(2164)

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 − En0 ) 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 0

n,n

471
" #
× < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >

(2165)

On expressing the product of the phase factor and the matrix elements of the
Ŝ(q) as an operator in the interaction representation
 
i
exp ( E n − En0 ) t < φn | Ŝα (q) | φn0 >

   
i i
= < φn | exp + Ĥ0 t Ŝα (q) exp − Ĥ0 t | φn0 >
h̄ h̄
= < φn | Ŝα (q; t) | φn0 >
(2166)

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
(2167)

The final states are a complete set of states, therefore, on using the completeness
relation, one finds
Z ∞  " #
dt 1
P (n) < φn | Ŝα (q; t) Ŝβ† (q; 0) | φn >
X
S α,β (q; ω) = exp i ω t
−∞ 2 π h̄ n
Z ∞  
dt 1
= exp i ω t < | Ŝα (q; t) Ŝβ† (q; 0) | >
−∞ 2 π h̄
(2168)

This is the Fourier Transform with respect to time of the thermal averaged spin -
spin correlation function. Furthermore, as the inverse spatial Fourier transform
of the spin density operator and the Hermitean conjugate are given by
Z  
1
Ŝα (q) = d3 r exp − i q . r Ŝα (r)
V
Z  
1
Ŝα† (q) = d3 r0 exp + i q . r0 Ŝα (r0 )
V
(2169)

Using this, and the spatial homogeneity of the system, one finds that the inelas-
tic neutron scattering spectrum is related to the spatial and temporal Fourier

472
transform of the spin - spin correlation function
Z ∞ Z  
1 dt 1
S α,β (q; ω) = d3 r exp i ( ω t − q . r ) < | Ŝα (r; t) Ŝβ (0; 0) | >
V −∞ 2 π h̄
(2170)

Thus, the inelastic neutron scattering probes the Fourier transform of the equi-
librium spin correlation functions.

24.3 The Fluctuation Dissipation Theorem


The spin - spin correlation function satisfies the principle of detailed balance

P (n) δ( h̄ ω + En − En0 ) < φn | Ŝα (q) | φn0 > < φn0 | Ŝβ† (q) | φn >
X
S α,β (q; ω) =
n,n0
(2171)

If the equilibrium probability P (n) is given by the Boltzmann expression


 
1
P (n) = exp − β En (2172)
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)
n,n0
  X
= exp β h̄ ω P (n) δ( En − En0 − h̄ ω ) < φn | Ŝβ† (q) | φn0 > < φn0 | Ŝα (q)
n,n0
 
= exp β h̄ ω S β,α (−q; −ω)

which is a statement of the principle of detailed balance. The correlation func-


tion 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


 
i
χα,β (r, r0 ; t − t0 ) = − < | Ŝα (r, t) , Ŝβ (r0 , t0 ) | > Θ( t − t0 ) (2174)

473
which can be expressed as
 
α,β i X i
χ (r; t) = − P (n) exp ( En − E n0 ) t < φn | Ŝα (r) | φn0 > < φn0 | Ŝβ (0) | φn
h̄ h̄
n,n0
 
i X i
+ P (n) exp ( En0 − E n ) t < φn | Ŝβ (0) | φn0 > < φn0 | Ŝα (r) | φn
h̄ 0

n,n

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
(2176)

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 >

h̄ ω + En0 − En + i δ
(2177)

The imaginary part of the dynamic susceptibility is given by


 
α,β 1 X
Im χ (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 >

(2178)

which can be written as


  "  #
1
Im χα,β (q; ω + iδ) = − S α,β (q; ω) 1 − exp − β h̄ ω
2
(2179)

or
   
S α,β (q; ω) = 2 1 + N (ω) Im χα,β (q; ω − iδ)

(2180)

474
This is the fluctuation dissipation theorem, it relates the dynamical response of
the system to an external perturbation to the naturally occurring excitations in
the system as measured by neutron scattering experiments.

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
limit of
Z ∞  
dt 1
α,β
S (q; ω) = exp i ω t < | Ŝα (q; t) Ŝβ† (q; 0) | >
−∞ 2 π h̄
(2181)

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) | > (2182)
t → ∞ h̄
produces a Bragg peak with finite intensity as
Z ∞  
dt
δ(ω) = exp i ω t (2183)
−∞ 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) | >
t → ∞ h̄
(2184)

and for a static system for which

< | Ŝα (q; t) | > = < | Ŝα (q; 0) | > (2185)

475
then the Bragg peak has an intensity given by
1
= lim < | Ŝα (q; 0) | > < | Ŝβ† (q; 0) | > (2186)
t → ∞ h̄
For a paramagnetic system the (quasi-stationary) average value of the spin
vector is zero
< | Ŝα (q; 0) | > = 0 (2187)
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 along
certain directions ( say α ), then

< | Ŝα (Q; 0) | > 6= 0 (2188)

then 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 | > (2189)

dΩ Bragg V
Q

For a small single domain single crystal the magnetic elastic scattering is ex-
tremely anisotropic, 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
leads to the determination of the preferred directions of the magnetic moments.

——————————————————————————————————

24.4.2 Exercise 80
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.

——————————————————————————————————

476
24.4.3 Exercise 81
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?

——————————————————————————————————

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, the energy difference between the initial state En and the final state
En0 so that energy conservation leads to

En0 − En = h̄ ωq (2190)

The matrix elements in the spin - spin correlation function can be evaluated as
Y Y
< φn | Sα (q) | φn0 > = < nq0 | Sα (q) | n0q” > (2191)
q0 q”

where the number of spin waves in the initial state are related to the number
in the final state via

n0q0 = nq0 f or q 0 6= q (2192)

and

n0q = nq + 1 (2193)

For a ferromagnet the matrix elements are evaluated as

< nq | Sz (q) | nq + 1 > = 0 (2194)

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
2
(2195)

and
1
< nq | Sy (q) | nq + 1 > = < nq | ( S + (q) − S − (q) ) | nq + 1 >
2i
1 q √
= + nq + 1 2S
2i
(2196)

477
Thus, the inelastic neutron scattering from the single spin wave excitations 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

d2 σ ( 2 π)2 N S
  X  
= re2 1 + q̂z2 δ( h̄ ω − h̄ ω(q 0 ) ) δ( q − q 0 −Q ) 1 + N (ω(q 0 ))
dω dΩ emit V 2 0 q ,Q
(2197)
Likewise, the absorption process has a scattering cross-section given by

d2 σ ( 2 π)2 N S
  X
= 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
(2198)
These satisfy the principle of detailed balance, and give rise to a Stokes and
anti-Stokes line in the spectrum of the scattered neutrons.

——————————————————————————————————

24.4.5 Exercise 82
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

478
susceptibility. For q values close to the Bragg peak, the susceptibility varies as

1
∝ (2199)
( q − Q )2 + ξ −2

where ξ the correlation length, in the mean field approximation, is given by


s
−1 Tc
ξ = a (2200)
6 ( T − Tc )

Thus, the critical scattering diverges as ( T − Tc )−1 as the transition temper-


ature is approached.

479
25 Superconductivity
The electrical resistivity ρ(T ) of metals at low temperatures is expected to be
described by the Drude model
m
ρ(T ) ∝ 2 (2201)
e τ
The resistivity vary with temperature according to
ρ(T ) ∼ ρ(0) + A T 2 + B T 5 (2202)
as 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. How-
ever, it was discovered by Kammerlingh Onnes that the resistivity of a metal
may become so small as to effectively vanish for all temperatures below a critical
temperature Tc (H. Kammerlingh Onnes, Comm. Phys. Lab. Univ. Leiden,
Nos. 119, 120, 122 (1911)). 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.

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-
feld, which is flux exclusion (W. Meissner and R. Ochsenfeld, Naturwiss. 21,
787 (1933)). A superconductor excludes the magnetic induction 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 superconductor. 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
(2203)

480
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 (2204)
so B = 0 implies that
1
M = − H (2205)

so that perfect diamagnetism implies that the susceptibility is given by
1
χ = − (2206)

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 (2207)
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 ).

25.1.1 The London Equations


A phenomenological description of superconductivity was developed by the Lon-
don brothers. Basically, this description is based on two phenomenological con-
stitutive 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) (2208)
mc

481
which expresses the perfect conductivity of a superconductor. It relates the
microscopic current in the superconductor which screens the applied magnetic
field. Here, ns is density of superfluid electrons. This is slightly different from
the condition of perfect conductivity in a metal which would be a relation be-
tween the time derivative of the current that flows in response to an electric
field. Here, it has been assumed that the electric field is transverse and the con-
dition for perfect conductivity has been integrated with respect to time, thereby
allowing constant currents to screen the static applied magnetic field. In order
that the continuity equation be satisfied in a steady state, the condition that
∇ . A(r, t) = 0, which defines the London gauge must be imposed.

The second London equation comes from Maxwell’s equations


j(r, t) 1 ∂
∇ ∧ B(r, t) = + E(r, t) (2209)
c c ∂t
and with the definitions

B(r, t) = ∇ ∧ A(r, t)
1 ∂
E(r, t) = − A(r, t)
c ∂t
(2210)

one finds
1 ∂2 ns e2
 
∇ ∧ ∇ ∧ + A(r, t) = − A(r, t) (2211)
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
λL , via

ns e2 1
= (2212)
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

482
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 − (2213)
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) = − (2214)
∂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 − f or z > 0 (2215)
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
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 (2216)

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 )) (2217)

483
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 ) (2218)

The magnetization in the normal state is negligibly small, but the supercon-
ducting state is perfectly diamagnetic, so
1
Mn − M s = + Hc (T ) (2219)

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 (2220)
4π ∂T
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
= − (2221)
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 = (2222)
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 − (2223)
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

484
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 identified
through the insight of Fröhlich (H. Fröhlich, Phys. Rev. 79, 845 (1950)), who
predicted that the superconducting transition temperature Tc should be pro-
portional to the phonon frequency. Furthermore, as the square of the phonon
frequency is inversely proportional to the mass of the ions M , the supercon-
ducting transition temperature should depend upon the isotopic mass through
1
Tc ∝ M − 2 (2224)

This isotope effect was confirmed in later experiments by Maxwell (E. Maxwell,
Phys. Rev. 78, 447 (1950), Phys. Rev. 79, 173 (1950)) and Reynolds et. al.
(C.A. Reynolds, B. Serin, W.H. Wright and L.B. Nesbitt, Phys. Rev. 78, 487
(1950)) on simple metals. However, in transition metals the exponent of the
isotope effect is reduced and may become zero, and in α − U the exponent is
positive. The occurrence of a positive isotope effect does not necessarily 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 is.

Consider a pair of electrons of spin σ and σ 0 , excited above the Fermi-


energy. 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 > (2225)
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
such that X
| C(k) |2 = 1 (2226)
k

485
The wave function must be an energy eigenstate of the Hamiltonian Ĥ and,
thus, satisfies
Ĥ | Ψq > = Eq | Ψq > (2227)
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
Eq − E(k) − E(k + q) C(k) = − V (k, k 0 ) C(k 0 ) (2228)
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) = (2229)
Eq − E(k) − E(k + q)

where α is given by
1 X
α(k) = − V (k, k 0 ) C(k 0 )
N 0 k

(2230)

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
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 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 =
N 0
E(k 0 ) + E(k 0 + q) − Eq
k

(2231)

Thus, the energy eigenvalue is determined from the equation

1 X V (k, k 0 )
1 =
N 0
E(k 0 ) + E(k 0 + q) − Eq
k

(2232)

For Cooper pairs with zero total momentum q = 0, this equation reduces to
Z µ + h̄ωD
ρ()
1 = V d (2233)
µ 2 − E

486
The density of states ρ() can be approximated by a constant ρ(µ), and the
integral can be inverted to give the energy eigenvalue as
2 h̄ ωD
E = 2µ −  2
 (2234)
exp V ρ(µ) − 1

This eigenvalue is less than the energy of the two independent electrons, thus,
the electrons are bound together. It is concluded that, due to the sharp cut off
of the integral at the Fermi-energy, the electron 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 cat-
egorized 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

(2235)

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

(2236)

Thus, for singlet pairing one must have

CS=0 (k) = CS=0 (−k) (2237)

which requires that, when expanded in spherical harmonics, the expansion only
contains even components of orbital angular momentum. For triplet pairing one

487
has
CS=1 (k) = − CS=1 (−k) (2238)
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 = 2. The high Tc superconductor such as Sr doped La2 CuO4 found by Bed-
norz and Muller in 1986 (Tc = 35 K) or Y Ba2 Cu3 O7 (Tc = 90 K) are slightly
exceptional to this rule. These materials evolve from an anti-ferromagnetic
insulator phase at zero doping, but as the doping increases they lose the an-
tiferromagnetism and become metallic paramagnets. A superconducting phase
appear above a small critical doping 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 compo-
nent 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 variations. This and the multiple superconducting transitions found
in U P t3 and T h doped U Be13 indicate that the gap is dominated by compo-
nents with non-zero angular momentum.

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) (2239)

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 (2240)

but  
X
ψ(r) = C(k) exp ik.r (2241)
k

Thus,
Z X  
ξ2 = d3 r r 2 C(k) C ∗ (k 0 ) exp i ( k − k0 ) . r
k,k0

488
X
= | ∇k C(k) |2
k
 2
4 h̄ vF
=
3 2µ − E
 2  
4 vF 4
= exp
3 2 ωD V ρ(µ)
(2242)
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 standard weak coupling theory of superconductivity due to Bardeen,


Cooper and Schrieffer, (B.C.S.), 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
electron phonon system can be expressed as the sum
Ĥ = Ĥ0 + Ĥint (2243)
where the non-interacting Hamiltonian is given by
ε(k) c†k,σ ck,σ +
X X
Ĥ0 = h̄ ωα (q) a†q,α aq,α (2244)
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 = (2245)
k,σ q,α

The Hamiltonian will be transformed via


   
0
Ĥ = exp + Ŝ Ĥ exp − Ŝ (2246)

where Ŝ is chosen in a way that will eliminate the interaction term ( at least to
in first order ). The operator S can be thought of as being of the same order as
Ĥint . That is on expanding the transformed Hamiltonian in powers of Ŝ
 
Ĥ 0 = Ĥ0 + Ĥint + Ŝ , Ĥ0
      
1 3
+ Ŝ , Ŝ , Ĥ0 + Ŝ , Ĥint + O Ĥint
2
(2247)

489
and then Ŝ is chosen such that
 
Ĥint = Ĥ0 , Ŝ (2248)

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 > (2249)
The complete set of energy eigenstates are energy eigenstates of the non inter-
acting electron and phonon Hamiltonian. The non-zero matrix elements only
occur between states which involve a difference of unity in the occupation num-
ber 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 +
ε(k + q) − ε(k) − h̄ ω(q) ε(k + q) − ε(k) + h̄ ω(q)
k,σ q,α
(2250)
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
(2251)
On evaluating the commutation relation one finds a renormalization of the elec-
tron 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
( ε(k + q) − ε(k) )2 − h̄2 ω(q)2
(2252)
Thus, for electrons within h̄ ω(q) of the Fermi-energy there is an attractive inter-
action between the electrons. As 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 interac-
tion. As the interaction is only attractive at sufficiently small frequencies the
interaction is only attractive after long time delays. In the B.C.S. theory this
interaction is further approximated. The approximation consists of only retain-
ing scattering between electrons of opposite momentum, as this maximizes the
phase space of allowed final states. That is the momenta and spin are restricted
such that
k = − k0 (2253)

490
and also
σ = − σ0 (2254)
This 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

ε(k) c†k,σ ck,σ − V (k, k 0 ) c†−k0 ,↓ c†k0 ,↑ ck,↑ c−k,↓


X X
Ĥ = (2255)
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 product over wave vector k. For each wave vector
k the Cooper pair ((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) ck,↑ c−k,↓ | 0 > (2256)
k

The amplitudes satisfy the constraint

| u(k) |2 + | v(k) |2 = 1 (2257)

The normal state for non-interacting electrons just corresponds to the special
case,
| u(k) |2 = Θ( µ − ε(k) ) (2258)
The functions u(k) and v(k) are variational parameters that are found be min-
imizing the expectation value of the Hamiltonian, which includes the pairing
interaction.

The expectation value for the appropriate energy, in the B.C.S. state, is
given by
X
< ΨBCS | ( Ĥ − µ N ) | ΨBCS > = 2 ( ε(k) − µ ) | u(k) |2
k

491
X
− V (k, k 0 ) v(k)∗ u(k) u(k 0 )∗ v(k 0 )
k,k0

(2259)
The term involving the double sum is eliminated by introducing a quantity
X
∆(k) = V (k, k 0 ) u(k 0 )∗ v(k 0 ) (2260)
k0

On minimizing the energy, subject to the constraint of conservation of proba-


bility, with respect to u(k) and v(k)∗ one finds
 
0 = 2 ( ε(k) − µ ) + λ u∗ (k) − ∆(k) v ∗ (k)

0 = λ v(k) − ∆(k) u(k)


(2261)
where λ is the Lagrange undetermined parameter. These equations can be
solved to yield  
2 1 ε(k) − µ
| u(k) | = 1 − (2262)
2 Eqp (k)
and  
2 1 ε(k) − µ
| v(k) | = 1 + (2263)
2 Eqp (k)
where we have defined
∆(k)
u(k) v ∗ (k) = (2264)
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) (2265)

The resulting expression can be solved for Eqp (k) to yield


p
Eqp (k) = + ( ε(k) − µ )2 + | ∆(k) |2 (2266)
The factor | u(k) |2 , is the probability of finding an electron of momentum k
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 1/Zk , in the
superconductor the distribution drops smoothly to zero as k increases above kF .
Thus, the concept of Fermi-surface is not well defined in the 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 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 ) .

492
25.3.3 The Gap Equation
The ”energy gap” parameter satisfies the non-linear integral equation
X ∆∗ (k 0 )
∆(k) = V (k, k 0 ) (2267)
2 E(k 0 )
k0

where V (k, k 0 ) is the attractive pairing interaction mediated by the phonons.


The interaction can be approximated by the attractive s-wave potential

V (k, k 0 ) = V f or | ε(k) − µ | < h̄ ωD

V (k, k 0 ) = 0 f or | ε(k) − µ | > h̄ ωD


(2268)

In this case one finds

∆(k) = ∆(0) f or | ε(k) − µ | < h̄ ωD

∆(k) = 0 f or | ε(k) − µ | > h̄ ωD


(2269)

where 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) dε p
0 ε + | ∆(0) |2
2

h̄ ωD
= V ρ(µ) sinh−1 (2270)
| ∆(0) |
which is solved as
h̄ ωD
| ∆(0) | = 1 (2271)
sinh V ρ(µ)
This gap 2 ∆(0) just corresponds to the energy required to break a 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
2 Eqp (k ) 2
k0

(2272)

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

493
with increasing temperature. This can only happen if | ∆(T ) | decreases with
increasing temperature. For sufficiently high temperatures, the equation can
reduces to
∆(T ) = ∆(T ) V ρ(µ) β h̄ωD (2273)
which only has the trivial solution ∆(0) = 0. The critical temperature where
the gap first vanishes ∆(Tc ) = 0 is given by
h̄ωD
tanh βc2 ε
Z
1 = ρ(µ) V dε
0 2ε
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
2 4 exp γ
(2274)

The superconducting gap decreases with increasing temperature and vanishes


at a critical temperature Tc given by
 
1
kB Tc = 1.14 h̄ωD exp − (2275)
V ρ(µ)
1
The critical temperature is proportional to M − 2 as expected from the isotope
effect. 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 ( Tc − T ) T → Tc (2276)
7 ξ(3) B

Thus, the order parameter has a typical mean field variation with an exponent
of β = 12 close to Tc .

25.3.4 The Ground State Energy


The normal state is unstable to the B.C.S. state only if it has a higher energy.
At T = 0 the stability can be seen be examining the energy
" #
X ( ε(k) − µ )2
< ΨBCS | ( Ĥ − µ N ) | ΨBCS > = ( ε(k) − µ ) − p
k
( ε(k) − µ )2 + | ∆(k) |2
| ∆(0) |2

V
(2277)

494
The condensation energy is defined as the difference between the energy of the
superconducting state and the normal state
Z 0
∆E = < ΨBCS | ( Ĥ − µ N ) | ΨBCS > − 2 dε ε ρ(µ + ε)
−∞
(2278)

The condensation energy is evaluated by writing the sum over k as an integral


over the density of states.
Z h̄ω " #
ε2
∆E = dε ρ(µ + ε) ε − p
0 ε2 + | ∆(0) |2
Z 0 " #
ε2
+ dε ρ(µ + ε) − ε − p
−h̄ωD ε2 + | ∆(0) |2
| ∆(0) |2

V
(2279)

Then the integral over states below the Fermi-energy, ε < 0, is transformed to
an integral over positive ε. This leads to
Z h̄ωD " #
ε2
∆E = 2 dε ρ(µ + ε) ε − p
0 ε2 + | ∆(0) |2
| ∆(0) |2

V

" #
Z h̄ωD p
= 2 ρ(µ) dε ε − ε2 + | ∆(0) |2
0
h̄ωD
| ∆(0) |2 | ∆(0) |2
Z
+ 2 ρ(µ) dε p −
0 ε2 + | ∆(0) |2 V
(2280)

The integrals are evaluated as

| ∆(0) |2 | ∆(0) |2 | ∆(0) |2


∆E = − ρ(µ) + −
2 V V
| ∆(0) |2
= − ρ(µ)
2
(2281)

which shows that the condensation energy comes from the attractive potential
that lowers the energy of the pair more than the increase in the potential energy

495
caused by the confinement 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
| ∆(0) |. Therefore, the B.C.S. state has lower energy than the normal state
when the gap is non-zero.

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 ) (2282)

The linearized mean field Hamiltonian is given by


X  
ĤM F − µ N = ( ε(k) − µ ) c†k,↑ ck,↑ + ( ε(−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

(2283)
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 ) (2284)
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 = ( ε(k) − µ ) c†k,↑ ck,↑ + ( ε(−k) − µ ) c†−k,↓ c−k,↓
k

496
X  | ∆(0) |2

− ∆(k) ck,↑ c−k,↓ + c†−k,↓ c†k,↑ ∆(k)∗ +
V
k

(2285)

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
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 (2286)

and    
βk† = exp + Ŝ c†−k,↓ exp − Ŝ (2287)

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 ĤM F exp − Ŝ (2288)

as they have the same eigenvalues and the eigenstates are related via
 
| φ0n > = exp + Ŝ | φn > (2289)

The operator Ŝ is chosen to be of the form


 
θk c†k,↑ c†−k,↓ − c−k,↓ ck,↑
X
Ŝ = (2290)
k

Explicitly, the transformation yields

αk = ck,↑ cos θk − c†−k,↓ sin θk

βk† = c†−k,↓ cos θk + ck,↑ sin θk


(2291)

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

497
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


(2292)

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
number operators vanish. The normal terms in the Hamiltonian are found as
X  
( ε(k) − µ ) c†k,↑ ck,↑ + ( ε(−k) − µ ) c†−k,↓ c−k,↓
k
"    #
αk αk† βk βk† αk† βk†
X
2 2
= ( ε(k) − µ ) sin θk + + cos θk αk + βk
k
 
αk† βk†
X
+ ( ε(k) − µ ) sin 2θk + βk αk
k

(2293)

The anomalous terms are evaluated as


X  † †

− ∆(k) ck,↑ c−k,↓ + c−k,↓ ck,↑ ∆(k)
k
   
βk† βk − αk αk†
X
= − Re ∆(k) sin 2θk
k
   
αk† βk† + βk αk
X
+ Re ∆(k) cos 2θk
k

(2294)

The off diagonal terms can be made to vanish by choosing


 
Re ∆(k)
tan 2θk = −  (2295)
ε(k) − 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. 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
 
Eqp (k) αk† αk + βk† βk
X
ĤM F = E0 + (2296)
k

498
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
p
Eqp (k) = + ( ε(k) − µ )2 + | ∆(k) |2 (2297)
The canonical transformation shows that the quasi-particles are part electron
and part hole like. Basically, this is a consequence that the quasi-particle ex-
citation consists of a single unpaired electron (k, σ), in the presence of the
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 83
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
(2298)

——————————————————————————————————

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
(2299)

499
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
Z +∞
T ∂∆(T )2
  
2 ∂f
Cqp (T ) = − dE ρqp (E) E 2 − (2300)
T −∞ 2 ∂T ∂E

which involves the average of the temperature derivative of the square of the
quasi-particle energy at µ, and the quasi-particle density of states
X  
ρqp (E) = δ E − Eqp (k) (2301)
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 ) (2302)
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
7 ξ(3)
(2303)

This ratio is a measure of the quantity


2
∂∆2 (T )
  
1 ∆(0)
2 T
∼ (2304)
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.

500
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 (2305)
T −∞ ∂E

The B.C.S. quasi-particle density of states is evaluated as


X  
ρqp (E) = δ E − Eqp (k)
k
Z ∞ 
p

= dε ρ(ε) δ E − ( ε − µ )2 + ∆(T )2
−∞
|E|
∼ ρ(µ)
|ε − µ|
|E|
= ρ(µ) p f or | E | > ∆(T )
E 2 − ∆(T )2
(2306)

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 − (2307)
kB T
in B.C.S. theory.

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.
The paramagnetic current is expressed as

e2 h̄ 1 X
 
j p (q; ω) = ( 2 k − q ) ( 2 k − q ) . A ×
4 m2 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
(2308)

501
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 h̄ 1 X
 
∂f (E(k))
j p (0; 0) = 2 2 k(k.A) − (2309)
m c V ∂E
k

The total current is found by combining the paramagnetic current with the
diamagnetic current

e2 h̄ 1 X ρ e2
 
∂f (E(k))
j(0; 0) = 2 2 k(k.A) − − A
m c V ∂E mc
k

e2 h̄ 1 ρ e2
Z  
4 ∂f (E(k))
= 2 2 dk k A − − A
m c 6 π2 ∂E mc
 2
ρ e2
Z   
e h̄ 1 4 ∂f (E(k))
= dk k − − A
m2 c 3 π 2 ∂E mc
(2310)

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 ) (2311)
∂E

which leads to the vanishing response as


2m
µ kF4 = 1 (2312)
h̄2 kF6

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 − (2313)
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 (2314)
mc
This shows that a current will flow in a superconductor in response to a static
vector potential, that is the current will screen an applied magnetic field. This
leads to the Meissner effect.

502
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 exter-
nal field. The Pauli spin susceptibility will also be modified by the superconduc-
tivity, and provide information about the pairing. The zero field susceptibility
is defined as a derivative of the magnetization, χs (T ) = ( ∂M ∂H ). The magneti-
zation, 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)) (2315)
2
k

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 bro-
ken 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) − (2316)
2

and so the spin susceptibility takes the usual form


 2 Z +∞  
g µB ∂f
χs (T ) = − 2 dE ρqp (E) (2317)
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 − (2318)
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
single particle eigenstates, and the spin up and spin down states are mixed. In

503
→−
− →
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 − (2319)
χn π ξ0
for strong spin-orbit scattering and for weak spin-orbit scattering one has

χs (0) π ξ0
= (2320)
χn 6 lso
Hence, a partial Meissner effect at T = 0 can be found in a conventional
superconductor.

26 Landau-Ginsberg 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. The latter length is given by the spatial extent of The Cooper pair
wave function or coherence length ξ
h̄ vf
ξ = (2321)
π ∆(T )

Type I Superconductors.

In simple (non-transition) metals the penetration depth is small, e.g., λ ∼


300 A for Al while the coherence length is large ξ ∼ 1 × 104 A as vF is large.
Materials where
λ 1
κ = < √ (2322)
ξ 2
are type I superconductors.

Type II Superconductors.

504
For transition metals, rare earths and intermetallic compounds, where the
band mass is very large, λL is very large ( ∼ 2000 A for V3 Ga ) and as the
Fermi-velocity is small ( vF ∼ 104 m /s ) then as Tc and ∆(0) are high xi is
small ( ∼ 50 A ). Materials where
λ 1
κ = > √ (2323)
ξ 2
are type II superconductors.

Since λL and κ diverge the same way at Tc the dimensionless ratio κ is ap-
proximately temperature independent.

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
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 (2324)
which are normal and regions where
B = 0 (2325)
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 (2326)
A 8π 8π
Hc2
where the term ξ 8 π is the energetic cost of setting the order parameter to
H2
zero, and the diamagnetic energy is reduced by λ 8 πc . Because of this pos-
itive 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 Φ0 , the flux is quantized in units of Φ0 .
hc
Φ0 = = 2.07 × 10−7 Gauss cm2 (2327)
2e

505
The vortices first enter the superconductor at a critical field Hc1 . The vortices
form a triangular lattice of vortices. 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 .

506

Potrebbero piacerti anche