Vous êtes sur la page 1sur 374

Quantum field theory: Lecture Notes

Rodolfo Alexander Diaz Sanchez


Universidad Nacional de Colombia
Departamento de Fsica
Bogota, Colombia

August 23, 2015


Contents

1 Relativistic quantum mechanics 7


1.1 Survey on quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.1 Vector subspaces generated by eigenvalues . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.2 Symmetries in quantum mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Irreducible inequivalent representations of groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Connected Lie groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.5 Lorentz transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.6 The inhomogeneous Lorentz Group (or Poincare group) . . . . . . . . . . . . . . . . . . . . . . . . 19
1.6.1 Four-vectors and tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
1.7 Some subgroups of the Poincare group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.7.1 Proper orthochronous Lorentz group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
1.7.2 Discrete transformations in the Lorentz group . . . . . . . . . . . . . . . . . . . . . . . . . . 24
1.7.3 Infinitesimal transformations within the proper orthochronus Lorentz group . . . . . . . . . 25
1.8 Quantum Lorentz Transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
1.8.1 Four-vector and tensor operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.8.2 Infinitesimal quantum Lorentz transformations . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.8.3 Lorentz transformations of the generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
1.8.4 Lie algebra of the Poincare generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
1.8.5 Physical interpretation of Poincares generators . . . . . . . . . . . . . . . . . . . . . . . . . 30
1.9 One-particle states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1.9.1 One-particle states under pure translations . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1.9.2 One-particle states under homogeneous Lorentz transformations . . . . . . . . . . . . . . . 32
1.9.3 Physical little groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.9.4 Normalization of one-particle states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
1.10 One-particle states with non-null mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
1.10.1 Wigner rotation and standard boost . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
1.11 One-particle states with null mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
1.11.1 Determination of the little group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
1.11.2 Lie algebra of the little group ISO (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
1.11.3 Massless states in terms of eigenvalues of the generators of ISO (2) . . . . . . . . . . . . . . 50
1.11.4 Lorentz transformations of massless states . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
1.12 Space inversion and time-reversal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
1.13 Parity and time-reversal for one-particle states with M > 0. . . . . . . . . . . . . . . . . . . . . . . 56
1.13.1 Parity for M > 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
1.13.2 Time reversal for M > 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
1.13.3 Parity for null mass particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
1.13.4 Time-reversal for null mass particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
1.14 Action of T 2 and Kramers degeneracy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

2
CONTENTS 3

2 Scattering theory 65
2.1 Construction of in and out states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
2.2 The Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
2.3 Symmetries of the Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2.3.1 Lorentz invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2.3.2 Internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
2.3.3 Parity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.3.4 Time-reversal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
2.3.5 PT symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
2.3.6 Charge-conjugation C, CP and CPT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
2.4 Rates and cross-sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
2.4.1 One-particle initial states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
2.4.2 Two-particles initial states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.4.3 Multi-particle initial states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.4.4 Lorentz transformations of rates and cross-sections . . . . . . . . . . . . . . . . . . . . . . . 95
2.5 Physical interpretation of the Diracs phase space factor 4 (p p ) d . . . . . . . . . . . . . . . 98
2.5.1 The case of N = 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.5.2 The case with N = 3 and Dalitz plots . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
2.6 Perturbation theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.6.1 Distorted-wave Born approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
2.7 Implications of unitarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
2.7.1 Generalized optical theorem and CPT invariance . . . . . . . . . . . . . . . . . . . . . . . . 111
2.7.2 Unitarity condition and Boltzmann H-theorem . . . . . . . . . . . . . . . . . . . . . . . . . 112

3 The cluster decomposition principle 115


3.1 Physical states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
3.1.1 Interchange of identical particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
3.1.2 Interchange of non-identical particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
3.1.3 Normalization of multi-particle states . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
3.2 Creation and annihilation operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
3.2.1 Commutation and anti-commutation relations of a (q) and a (q) . . . . . . . . . . . . . . . 119
3.3 Arbitrary operators in terms of creation and annihilation operators . . . . . . . . . . . . . . . . . 120
3.4 Transformation properties of the creation and annihilation operators . . . . . . . . . . . . . . . . . 121
3.5 Cluster decomposition principle and connected amplitudes . . . . . . . . . . . . . . . . . . . . . . . 122
3.5.1 Some examples of partitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
3.6 Structure of the interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
3.6.1 A simple example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
3.6.2 Connected and disconnected parts of the interaction . . . . . . . . . . . . . . . . . . . . . . 131
3.6.3 Some examples of the diagrammatic properties . . . . . . . . . . . . . . . . . . . . . . . . . 133
3.6.4 Implications of the theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

4 Relativistic quantum field theory 137


4.1 Free fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
4.2 Lorentz transformations for massive fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
4.2.1 Translations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
4.2.2 Boosts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.2.3 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
4.3 Implementation of the cluster decomposition principle . . . . . . . . . . . . . . . . . . . . . . . . . 145
4.4 Lorentz invariance of the Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
4 CONTENTS

4.5 Internal symmetries and antiparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147


4.6 Lorentz irreducible fields and Klein-Gordon equation . . . . . . . . . . . . . . . . . . . . . . . . . . 149

5 Causal scalar fields for massive particles 151


5.1 Scalar fields without internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151
5.2 Scalar fields with internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
5.3 Scalar fields and discrete symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

6 Causal vector fields for massive particles 162


6.1 Vector fields without internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
6.2 Spin zero vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
6.3 Spin one vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
6.4 Spin one vector fields with internal symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
6.4.1 Field equations for spin one particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
6.5 Inversion symmetries for spin-one fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176

7 Causal Dirac fields for massive particles 178


7.1 Spinor representations of the Lorentz group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
7.2 Some additional properties of the Dirac matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
7.3 The chiral representation for the Dirac matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
7.4 Causal Dirac fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
7.5 Dirac coefficients and parity conservation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193
7.6 Charge-conjugation properties of Dirac fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
7.7 Time-reversal properties of Dirac fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
7.8 Majorana fermions and fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
7.9 Scalar interaction densities from Dirac fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
7.10 The CPT theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209

8 Massless particle fields 211

9 The Feynman rules 223


9.1 General framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
9.2 Rules for the calculation of the Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
9.3 Diagrammatic rules for the Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
9.4 Calculation of the Smatrix from the factors and diagrams . . . . . . . . . . . . . . . . . . . . . . 229
9.5 A fermion-boson theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
9.5.1 Fermion-boson scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
9.5.2 Fermion fermion scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
9.5.3 Boson-boson scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
9.6 A boson-boson theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
9.7 Calculation of the propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
9.7.1 Other definitions of the propagator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
9.8 Feynman rules as integrations over momenta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
9.9 Examples of application for the Feynman rules with integration over four-momenta variables . . . 252
9.9.1 Fermion-boson scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
9.9.2 Fermion-fermion scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
9.9.3 Boson-boson scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
9.10 Examples of Feynman rules as integrations over momenta . . . . . . . . . . . . . . . . . . . . . . . 258
9.10.1 Fermion-boson scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 259
9.10.2 Fermion-fermion and Boson-boson scattering . . . . . . . . . . . . . . . . . . . . . . . . . . 260
CONTENTS 5

9.11 Topological structure of the lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260


9.12 Off-shell and on-shell four-momenta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
9.12.1 The rth derivative theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264

10 Canonical quantization 268


10.1 Canonical variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
10.1.1 Canonical variables for scalar fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
10.1.2 Canonical variables for vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
10.1.3 Canonical variables for Dirac fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
10.2 Functional derivatives for canonical variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
10.3 Free Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
10.3.1 Free Hamiltonian for scalar fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
10.4 Interacting Hamiltonians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
10.5 The Lagrangian formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
10.6 From Lagrangian to Hamiltonian formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
10.6.1 Setting the Hamiltonian for the use of perturbation theory . . . . . . . . . . . . . . . . . . 286
10.7 Gauges of the Lagrangian formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
10.8 Global symmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290
10.9 Conserved quantities in quantum field theories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
10.9.1 Space-time translation symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 294
10.9.2 Conserved currents and Lagrangian densities for space-time symmetries . . . . . . . . . . . 295
10.9.3 Additional symmetry principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
10.9.4 Conserved current for a two scalars Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . 298
10.10Lorentz invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
10.10.1 Currents and time-independent operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
10.10.2 Generators and Lie algebra between the homogeneous and inhomogeneous generators . . . 303
10.10.3 Invariance of the Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
10.10.4 Lie algebra within the homogeneous generators . . . . . . . . . . . . . . . . . . . . . . . . . 305
10.11The transition to the interaction picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
10.11.1 Scalar field with derivative coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
10.11.2 Spin one massive vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
10.11.3 Dirac Fields of spin 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
10.12Constraints and Dirac Brackets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316

11 Quantum electrodynamics 323


11.1 Gauge invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
11.1.1 Currents and their coupling with A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325
11.1.2 Action for the photons (radiation) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326
11.1.3 General overview of gauge invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
11.2 Constraints and gauge conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
11.3 Quantization in Coulomb Gauge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
11.3.1 Canonical quantization of the constrained variables . . . . . . . . . . . . . . . . . . . . . . . 332
11.3.2 Quantization with the solenoidal part of ~ . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
11.3.3 Constructing the Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
11.4 Formulation of QED in the interaction picture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
11.5 The propagator of the photon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
11.6 Feynman rules in spinor electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
11.6.1 Drawing the Feynman diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
11.6.2 Factors associated with vertices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
6 CONTENTS

11.6.3 Factors associated with external lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344


11.6.4 Factors associated with internal lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
11.6.5 Construction of the Smatrix process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
11.7 General features of the Feynman rules for spinor QED . . . . . . . . . . . . . . . . . . . . . . . . . 346
11.7.1 Photon polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
11.7.2 Electron and positron polarization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
11.8 Example of application: Feynman diagrams for electron-photon (Compton) scattering . . . . . . . 350
11.9 Calculation of the cross-section for Compton scattering . . . . . . . . . . . . . . . . . . . . . . . . . 352
11.9.1 Feynman amplitude for Compton scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
11.9.2 Feynman amplitude for the case of linear polarization . . . . . . . . . . . . . . . . . . . . . 356
11.9.3 Differential cross-section for Compton scattering . . . . . . . . . . . . . . . . . . . . . . . . 358
11.9.4 Differential cross-section in the laboratory frame . . . . . . . . . . . . . . . . . . . . . . . . 360
11.10Traces of Dirac gamma matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
11.11Some properties of slash momenta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365

12 Path integral approach for bosons in quantum field theory 366


12.1 The general path-integral formula for bosonic operators . . . . . . . . . . . . . . . . . . . . . . . . 367
12.1.1 Probability amplitude for infinitesimal time-intervals . . . . . . . . . . . . . . . . . . . . . . 369
12.1.2 Probability amplitude for finite time intervals . . . . . . . . . . . . . . . . . . . . . . . . . . 369
12.1.3 Calculation of matrix elements of operators through the path-integral formalism . . . . . . 371
12.2 Path formalism for the Smatrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
Chapter 1

Relativistic quantum mechanics

1.1 Survey on quantum mechanics


A Hilbert space E, is a complete vector space with inner product. Given two vectors |i , and |i in such a space,
there is a complex number h |i that satisfies the following axioms

h |i = h |i
h |1 + 2 i = h |1 i + h |2 i
h1 + 2 |i = h1 |i + h2 |i
h |i 0 ; and h |i = 0 |i = 0 (1.1)

where we define the norm of a vector |i as

k|ik2 h |i ; k|ik2 = 0 |i = 0

which is positive definite i.e. it is positive for any non-zero vector and zero for the null vector. A vector is
normalized if its norm is equal to unity. Physical states in quantum mechanics are described by normalized
vectors (or kets) on the Hilbert space E. Physical observables in quantum mechanics are eigenvalues of hermitian
operators with a complete spectrum (that is, the eigenvectors of each one of these operators form a basis on the
Hilbert space). We recall that the adjoint A of a linear operator A, is another linear operator on a Hilbert space
that satisfies the condition E

hA |i = h A ; |i , |i E

further, it also happens that two linearly dependent (normalized) vectors |i describe the same physical state.
This fact induces the following definition

Definition 1.1 Given a normalized state |i E, A ray induced by a ket |i is the set of all normalized vectors
that are linearly dependent with |i n o
R|i ei |i : [0, 2)

Two vectors belonging to the same ray describe the same physical states. On the other hand, two vectors
belonging to different rays are linearly independent so that they represent different physical states. Therefore, if
we think on each ray as a single object, we could say that a given physical state is represented by a single ray
and that a given ray represents a unique physical state. This one-to-one relation between rays and physical states
justify the introduction of such a concept.
For a given observable A, a given ray R posseses a unique eigenvalue if the vectors of R are eigenvectors of
A
A |i = |i |i R

7
8 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

the eigenvalues of an hermitian operator are real (this fact is necessary to interpret such eigenvalues as physical
observables), and eigenvectors associated with different eigenvalues are orthogonal.
When the spectrum of a given observable is degenerate, a given eigenvalue n could be associated with several
linearly independent eigenvectors as follows

A |nm i = n |nm i ; m = 1, . . . , gn

where gn is the degree of degeneracy. Therefore, there will be gn rays Rm


n associated with a given eigenvalue n .
If a system is in a state described by the ray R, and we measure an observable A, the probability of finding
the eigenvalue k of A is given by
gk
X
P (k ) = |hkm |i|2 ; |i R and |km i Rm
k
m=1

since observables (hermitian complete linear operators on E), have eigenvectors that form a basis of E the sum of
these probabilities is equal to the unity
gn gn
" gn
#
X XX 2
XX XX
m m m m m
P (k ) = |hk |i| = h |k i hk |i = h| |k i hk | |i = h| I |i
k k m=1 k m=1 k m=1
X
P (k ) = 1
k

thus, the feature for the observables of having a complete spectrum, is essential to keep the conservation of
probability.

1.1.1 Vector subspaces generated by eigenvalues


Let {E1 , E2 , . . . , Eq } be a set of subspaces of a given vector space E. We say that E is the direct sum of such a
set of subspaces and denote it as
E = E1 E2 . . . Eq
if any given arbitrary vector x E is expresible in a unique way in the form

x = x1 + x2 + . . . + xq such that xk Ek

in words, given an arbitrary x E, it can be expressed by a sum of qvectors {xk }, where each xk belongs to a
subspace Ek in the set. In addition, there is one and only one vector xk belonging to Ek , that can be part of this
decomposition. We say that xk is the projection of x into the subspace Ek .
Let A be a linear operator of a vector space E into itself. We say that a vector subspace Ek is invariant under
the action of A if for any x Ek , we have that Ax = x Ek . In words, a subspace Ek is invariant under A if by
restricting the domain of A to Ek , the resultant range is also contained in Ek .
Let A be an observable (hermitian complete linear operator on E). For simplicity, we shall assume that its
spectrum is discrete. Its eigenvalues and eigenvectors are given by

A |nm i = an |nm i ; m = 1, ..., gn

with gn the degeneracy of the eigenvalue an . By taking all linearly independent vectors of the form
 1 2
, , | gn i
n n n

and forming all possible linear combinations (including the null linear combination) we obtain the set of all
eigenvectors of A, with eigenvalue an (plus the null vector). This set forms a subspace of E, called the vector
1.2. SYMMETRIES IN QUANTUM MECHANICS 9

subspace induced by the eigenvalue an of A, and is denoted by Ean . The dimensionality of such a subspace is
clearly the degree gn of degeneracy of an . On the other hand, by virtue of the completeness of A, the set of all its
linearly independent eigenvectors forms a basis of E. Consequently, given the complete set {|nm i} an arbitrary
vector x E, can be written in a unique way, as a linear combination of the form
gn
X
X
x = n,m |nm i (1.2)
n=1 m=1
Xg1 g2
X gn
X
x = 1,m |1m i + 2,m |2m i + ... + n,m |nm i + . . .
m=1 m=1 m=1
gk
X
x = x1 + x2 + . . . + xn + . . . ; xk k,m |km i (1.3)
m=1

several observations are in order at this step: (1) a given vector xk as defined in (1.3) belongs to Eak . (2) Since
the complete set of scalars that define the linear combination in (1.2) is unique (for a given order of the basis),
each vector xk defined in (1.3) is also unique for a given x. In conclusion, the Hilbert space E, can be decomposed
in a direct sum of subspaces generated by the eigenvalues {an } of a given observable A
E = Ea1 Ea2 . . . Ean . . . (1.4)
where the dimension of each Ean is the degree of degeneracy of the eigenvalue an associated. In particular, if the
eigenvalue is non-degenerate, the associated subspace is one-dimensional. Further, it is quite obvious that each
subspace Ean is invariant under the observable A.
We should keep in mind that the decomposition (1.4) depends on the observable chosen. By choicing another
observable B, we should take its eigenvalues bm and construct the associated subspaces Ebm , in order to construct
the decomposition of E induced by B.

1.2 Symmetries in quantum mechanics


A (passive) symmetry transformation is a change in the point of view that does not change the results of possible
experiments. If an observer O sees the system in a state described for the ray R, an equivalent observer O could
see the same physical system in another state described by the ray R . Nevertheless, both observers must find the
same physics, for example they should find the same probabilities

P (R, k ) = P R , k
indeed this is only a necessary condition. Additional conditions are necessary for the transformation to be a
symmetry transformation1 . We shall establish without proof a theorem owe to Wigner concerning with the
characterization of possible structures for symmetry operators

Theorem 1.1 The symmetry representation theorem: Any symmetry transformation R R can be characterized
by an operator U on the Hilbert space, such that if |i R then U |i R with U being either linear unitary
U ( + ) = U () + U () ; |i , |i E , , C
hU |U i = h |i ; |i , |i E
or antilinear antiunitary
U ( + ) = U () + U () ; |i , |i E , , C

hU |U i = h |i ; |i , |i E
1
In this context, the rays R and R are associated with the same physical state because both are seen from different observers. If
they are seen by the same observer and R =6 R , they must be associated with different physical states.
10 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

The adjoint of a linear operator is defined as


E

hU |i = h U ; |i , |i E (1.5)

We shall see that the relation (1.5) is not consistent for antilinear operators. To prove it, let us consider an
arbitrary complex linear combination of the form

= 1 1 + 2 2 (1.6)

substituting (1.6) on the LHS of Eq. (1.5), using the antilinearity of U and the axioms (1.1) we have

hU (1 1 + 2 2 ) |i = h1 U 1 + 2 U 2 |i
hU (1 1 + 2 2 ) |i = 1 hU 1 |i + 2 hU 2 |i (1.7)

on the other hand, using Eq. (1.5), the antilinearity of U and the axioms (1.1), we can write the same expression
as
E E E

hU (1 1 + 2 2 ) |i = h1 1 + 2 2 U = 1 h1 U + 2 h2 U
hU (1 1 + 2 2 ) |i = 1 hU 1 |i + 2 hU 2 |i (1.8)

equating equations (1.7, 1.8) we obtain that i = i which cannot be hold by arbitrary complex values of 1 , 2 .
In other words2 , the condition (1.5) cannot be satisfied by an antilinear operator because Eq. (1.7) says that the
left-hand-side (LHS) of Eq. (1.5) is linear in , while Eq. (1.8) says that the same expression is antilinear in .
Therefore, we shall define the adjoint of an antilinear operator as
E

hU |i h U = h |U i ; |i , |i E (1.9)

with this definition, the conditions for both unitarity or antiunitarity take the form

U = U 1 (1.10)

the identity is a trivial symmetry transformation which is linear and unitary. Many symmetries in Physics are
continuous in the sense that we can connect the associated operator U with the identity by means of a continuous
change in some parameters. This is the case in rotations, translations and Lorentz transformations. In that case,
the requirement of continuity demands for the symmetry to be represented by a unitary linear transformation. To
see it, we observe that we cannot pass continuously from a linear unitary operator (the identity) to an antilinear
antiunitary operator, such a transition requires at least one discrete transformation. Symmetries represented by
antilinear antiunitary operators involve a reversal in the direction of times flow.
If a symmetry transformation is infinitesimally closed to the identity, it can be represented by a linear unitary
operator written as
U = I + iT
with a real infinitesimal quantity. For U to be unitary and linear T must be linear hermitian. Most of
observables in quantum mechanics such as the angular momentum, momentum, Hamiltonian etc, arise from
symmetry transformations in this way.
The set of symmetry transformations satisfy the axioms of a group. If Ti is a transformation that takes Rn
into Rn , we see that (a) the identity is a symmetry transformation, (b) composition of symmetry transformations
yields another symmetry transformation, (c) transformations are associative, (d) each symmetry transformation
has an inverse that is also a symmetry transformation
2
We should take into account that the Hilbert space of quantum mechanics is a complex (rather than real) vector space.
1.2. SYMMETRIES IN QUANTUM MECHANICS 11

The unitary or antiunitary operators {U (Ti )} associated with these symmetry transformations carry the group
properties of the set {Ti }. However, the operators U (T ) act on vectors of the Hilbert space instead of rays. If
T1 takes Rn into Rn , then when U (T1 ) acts on a vector |n i Rn , it yields a vector U (T1 ) |n i Rn . Further
if T2 takes Rn into Rn then U (T2 ) acting on U (T1 ) |n i must yield a vector in the ray Rn . On the other hand
T2 T1 also takes Rn into Rn . Consequently U (T2 T1 ) |n i is also in the ray Rn , so these vectors can only differ by
a phase n (T2 , T1 )
U (T2 ) U (T1 ) |n i = ein (T2 ,T1 ) U (T2 T1 ) |n i (1.11)

we shall see that with one important exception, the linearity or antilinearity of U (T ) tells us that these phases
are independent of the state |n i. Consider two linearly independent states |A i and |B i, applying Eq. (1.11)
to the states |A i and |B i as well as to the state |AB i |A i + |B i, we have

U (T2 ) U (T1 ) |AB i = eiAB (T2 ,T1 ) U (T2 T1 ) |AB i


eiAB (T2 ,T1 ) U (T2 T1 ) [|A i + |B i] = U (T2 ) U (T1 ) [|A i + |B i] (1.12)
iAB (T2 ,T1 ) iAB (T2 ,T1 )
e U (T2 T1 ) |A i + e U (T2 T1 ) |B i = U (T2 ) U (T1 ) |A i + U (T2 ) U (T1 ) |B i
iAB (T2 ,T1 ) iAB (T2 ,T1 )
e U (T2 T1 ) |A i + e U (T2 T1 ) |B i = eiA U (T2 T1 ) |A i + eiB U (T2 T1 ) |B i (1.13)

any linear unitary or antilinear antiunitary operator has an inverse (its adjoint) which is also linear unitary or
antilinear antiunitary respectively. Multiplying (1.13) by U 1 (T2 T1 ) we find
n o n
U 1 (T2 T1 ) eiAB (T2 ,T1 ) U (T2 T1 ) |A i + eiAB (T2 ,T1 ) U (T2 T1 ) |B i = U 1 (T2 T1 ) eiA U (T2 T1 ) |A i
o
+eiB U (T2 T1 ) |B i

when the operator U 1 (T2 T1 ) jumps over the complex numbers, the latter become their complex conjugates when
the operator is antilinear thus

eiAB (T2 ,T1 ) |A i + eiAB (T2 ,T1 ) |B i = eiA |A i + eiB |B i (1.14)

where the minus sign in the phases occurs when the operator is antilinear. Since |A i and |B i are linearly
independent, we equate coefficients in (1.14) to obtain

eiAB = eiA = eiB

hence the phases are independent of the states and Eq. (1.11) can be written as an equation of operators

U (T2 ) U (T1 ) = ei(T2 ,T1 ) U (T2 T1 ) (1.15)

in the case = 0, U (T ) provides a representation of the group of symmetry transformations. For non-null phases
we obtain a projective representation or a representation up to a phase. The structure of the Lie group
cannot tell us whether physical state vectors furnish an ordinary or a projective representation, but it can tell us
whether the group has any intrinsically projective representation.
The exception for the preceding argument has to do with the fact that it may not be possible to prepare the
system in a state represented by |A i + |B i. For instance, it is widely believed that we cannot prepare a system
in a superposition of two states whose total angular momentum are integer and half-integer respectively. In that
case there is a superselection rule between different classes of states and the phases (T2 , T1 ) could depend on
which of these classes of states are acting the operators U (T2 ) U (T1 ) and U (T2 T1 ).
It could be shown that any symmetry group with projective representations can be enlarged in such a way
that its representations can all be defined as ordinary i.e. with = 0, without changing the physical contents.
12 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

1.3 Irreducible inequivalent representations of groups


For the sake of simplicity, we shall restrict our discussion to finite-dimensional vector spaces but most of our
results are applied to infinite-dimensional vector spaces.
Let G = {Ti } be a group. Each element of the group Tn can be mapped into a linear operator U (Tn ) of a
vector space V onto itself, in such a way that

U (Tk Tm ) = U (Tk ) U (Tm ) (1.16)

relation (1.16) guarantees that the group structure of G is preserved in the mapping. The set of linear operators
{U (Ti )} is called a representation of the group G, in the vector space V . If the mapping Ti U (Ti ) is one-to-one,
we say that the representation is faithful and that there is an isomorphism between G and {U (Ti )}. In that case,
both sets are totally identical as groups. However, in some cases, the mapping Ti U (Ti ) is not one-to-one, in
that case the representation is degenerate and we say that the mapping is a homomorphism. In that case, some
information about G is not carried by {U (Ti )}.
Let {U (Ti )} be a representation of a group G in a vector space V . If we take  an arbitrary non-singular
operator Q of V onto itself, it is obvious that the set of operators {W (Ti )} Q U (Ti ) Q1 also forms a
representation. But equally obvious is the fact that this representation does not contain any new information.
Conversely, suppose that we have two representations {U (Ti )} and {W (Ti )} of the same group in the same
vector space V , if it exists an operator Q such that W (Ti ) = Q U (Ti ) Q1 , for all W (Ti ) and for all U (Ti ) of
each representation, then we say that {U (Ti )} and {W (Ti )} are equivalent representations, and we take them
essentially as a single representation.
However, it happens in some cases that there is not a non-singular operator Q that connects the two repre-
sentations {U (Ti )} and {W (Ti )} (defined on the same vector space) in the way described above. In that case we
say that we have two (or more) inequivalent representations of G in the vector space V .
Let {U (Ti )} be a representation of a group G in a vector space V . Suppose that exists a proper subspace Vk of
V that is invariant under all linear operators in the set {U (Ti )}. It means that we can restrict to Vk the domain of
each U (Ti ), because the range of each U (Ti ) under such a restriction will be contained in Vk . As a consequence,
we can form a representation {U (Ti )} of G in the vector space Vk V . We say that the representation {U (Ti )}
of G in V , is reducible because such a representation can be restricted to a proper subspace of V .
Even more, suppose that V can be decomposed in non-null vector subspaces {Vp } such that

V = V1 V2 . . . Vm (1.17)

and that each subspace Vp is invariant under all operators U (Ti ). In that case we say that the representation
defined in V , is fully reducible into representations defined on each proper subspace Vk . Of course, it could
happen that a given subspace Vp could be further reduced in smaller non-null vector subspaces invariant under
all U (Ti ), so that the representation on Vp is in turn reducible. The idea is to find a decomposition like (1.17)
such that none of the subspaces Vp can be decomposed into smaller non-null subspaces in which we can form
representations of {U (Ti )}. In that case, we say that our representation is irreducible and that each Vp is a
minimal invariant subspace under {U (Ti )}. In addition, for most of the cases of interest, the subspaces of the
decomposition (1.17) are orthogonal each other, i.e. for any given vector xi Vi and any given vector xk Vk we
have that hxk |xi i = 0 if k 6= i. We say that (1.17) is an orthogonal decomposition and we denote it as Vi Vk .
We shall assume from now on that we are dealing with orthogonal decompositions unless otherwise indicated.
When we have a representation {U (Ti )} of G in V , we can form the matrix representations of each U (Ti ) by
taking any orthonormal basis
V {|va i} (1.18)
of V (we shall call it the original basis). Now, suppose that {U (Ti )} is reducible in V , and that such a
reduction can be carried out as in Eq. (1.17). In that case, it is more convenient to choose the basis in the
following way: we take an orthonormal basis of the subspace V1 of dimension d1 , that is a set of vectors {|w1,r1 i}
1.3. IRREDUCIBLE INEQUIVALENT REPRESENTATIONS OF GROUPS 13

where r1 runs over d1 linearly independent vectors within V1 . We proceed in the same way with V2 and so on,
then we form a basis for the whole space V as follows

{|wi,ri i ; i = 1, 2, . . . , m ; ri = 1, 2, . . . , di } (1.19)

it is easy to see that in this basis ordered as

{|w1,1 i , |w1,2 i , . . . , |w1,d1 i , |w2,1 i , |w2,2 i , . . . , |w2,d2 i , . . . , |wm,1 i , |wm,2 i , . . . , |wm,dm i} (1.20)

the matrix representatives of each U (Tp ) in V are all block-diagonal. To see it, we observe that each Vi is invariant
under all U (Tp ). Hence, |wi,ri i Vi implies that U (Tp ) |wi,ri i Vi and taking into account that |wk,rk i Vk and
that Vi Vk for i 6= k, we have
hwk,rk | U (Tp ) |wi,ri i = 0 If i 6= k
Therefore, the matrix representation of each U (Tp ) in the ordered basis (1.20) does not connect two vectors
associated with different subspaces Vi and Vk . Thus, we form submatrices associated with each Vk , with zeros in
the other entries. Since the basis given by (1.20) simplifies considerably the texture of the matrix representation
of the operators U (Tp ), we call it the canonical basis associated with the representation {U (Tp )} in V .
Let us illustrate these facts with an example. Assume that {U (Ti )} is a representation of a group G on a
seven dimensional vector space V , and that V can be decomposed in three minimal orthogonal invariant subspaces
under {U (Ti )}, as
V = V1 V2 V3
where V1 is 2-dimensional, V2 is 3-dimensional and V3 is 2-dimensional, let us take an orthonormal basis on each
subspace as follows

V1 {|w1,1 i , |w1,2 i} ; V2 {|w2,1 i , |w2,2 i , |w2,3 i} ; V3 {|w3,1 i , |w3,2 i}

so that we shall use the following orthonormal ordered basis in V

{|w1,1 i , |w1,2 i , |w2,1 i , |w2,2 i , |w2,3 i , |w3,1 i , |w3,2 i}

under this ordered basis, the matrix representative of each U (Ti ) will have the following texture

0 0 0 0 0
0 0 0 0 0

0 0 0 0 A22 (U (Ti )) 0 0

D (U (Ti )) 0 0 0 0 =
0 B33 (U (Ti )) 0 (1.21)
0 0 0 0 0 0 C22 (U (Ti ))

0 0 0 0 0
0 0 0 0 0
D (U (Ti )) = A22 (U (Ti )) B33 (U (Ti )) C22 (U (Ti )) (1.22)

where the symbol denotes elements that could be non-null. The matrices A22 , B33 , C22 are the matrix
representatives of each U (Ti ) in the subspaces V1 , V2 , V3 respectively. So the block-diagonal texture of Eq. (1.21),
shows that the representation in V can be expressed as a direct sum of representations in V1 , V2 , V3 . Conversely, if
we have two (or more) representations in spaces V1 , V2 , it is clear that we can form a new representation by taking
the direct sum of them which is a representation in V1 V2 . However, it is equally clear that the new representation
formed that way does not contain any new information with respect to the component representations.
Notwithstanding, we should keep in mind that even if {U (Ti )} in V is reducible, by choicing an arbitrary basis
such as (1.18), the matrices will not have the texture (1.21). To exhibit such a texture, an apropriate ordered basis
such as the canonical basis given by (1.20) must be chosen. Therefore, changing from our original basis to a
canonical basis in which the reduction is apparent, is one of the main challenges of group representation theory.
14 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

The previous discussion, shows us that in characterizing representations of a given group, we should take over
two types of redundancies: (a) Given two representations in the same vector space, we consider them different
only if they are inequivalent. Equivalent representations are consider as a single one. (b) Given a representation
on a vector space V , we should reduce it (if possible) in order to find the irreducible representations. The direct
sum of these irreducible representations has no more information with respect to the irreducible ones. Thus, only
irreducible representations are considered.
Therefore, in characterizing the representations of a given group, we intend to classify all (or as many as
possible) irreducible inequivalent representations. The technics and criteria for this classification are out of the
scope of the present treatment. By now and for future purposes, we only mention a couple of Lemmas that are
crucial in the theory of irreducible representations of groups.

Lemma 1 (Schurs lemma 1) Let U (G) and U (G) be two irreducible representations of a group G in V and V
respectively. Let A be a linear transformation from V to V which satisfies A U (g) = U (g) A for all g G. It
follows that either (i) A = 0, or (ii) A is an isomorphism from V onto V (i.e. V and V are isomorphic) and
U (G) is equivalent to U (G).

Lemma 2 (Schurs lemma 2) Let U (G) be an irreducible representation of a group G on the finite-dimensional
vector space V . Let A be an arbitrary operator in V . If A commutes with all the operators in the representation,
that is if A U (g) = U (g) A, g G, then A must be a multiple of the identity operator.

By now we shall only discuss an important consequence of Schurs lemma 2

Theorem 1.2 All irreducible representations of any abelian group must be 1-dimensional.

Proof : Let U (G) be an irreducible representation of an abelian group G. Let p be a fixed element of the
group. Now, U (p) U (g) = U (g) U (p) g G, because of the abelianity of the group. Hence U (p) is an operator
that commutes with all the U (g) s, we conclude from Schurs lemma 2 that U (p) = p E. Since p is arbitrary,
the representation {U (g)} is equivalent to the set of operators {p E}. But this representation is reducible in
contradiction with our hypothesis, unless E is the identity in one dimension. Therefore, U (G) is equivalent to
the representation p p C for all p G. QED.

1.4 Connected Lie groups


These are groups of transformations T () that are described by a finite set of continuous parameters

{} 1 , 2 , ..., r

in such a way that each element of the group is continuosly connected with the identity by a path within the
group. The group multiplication rule takes the form
       
T T () = T f , ; f , f 1 , , f 2 , , . . . , f r , = 1 , 2 , . . . , r (1.23)
  
where f , is a set of r functions such that for a given function f a , an a given couple of sets ,
we a a
 have f , = with another set of rparameters. According with (1.23), the set of r functions
f , provides the law of combination for the two sets of parameters and which in turn provides the law
of combination of the group.
By convention, it is customary to choose a = 0 as the coordinates of the identity, in that case we have

T () = T (0) T () = T (f (0, )) ; T () = T () T (0) = T (f (, 0))


T () = T (f (0, )) = T (f (, 0))
1.4. CONNECTED LIE GROUPS 15

consequently
f a (, 0) = f a (0, ) = a (1.24)
since these transformations are continuously connected with the identity, they must be represented on the Physical
Hilbert space by unitary (rather than antiunitary) operators U (T ()). Operators U (T ()) can be represented
by a power series, at least in a finite neighborhood of the identity

1 
U (T ()) = I + i a Ta + b c Tbc + O 3 (1.25)
2
where Ta and Tbc = Tcb are operators independent of the s, with Ta hermitian. Suppose that U (T ()) provides
an ordinary (non-projective) representation of this group of transformations, thus
  
U T U (T ()) = U T T ()

and using (1.23) we find


 
U T U (T ()) = U T f , (1.26)

by expanding condition (1.26) in powers of and , we shall obtain a condition. The expansion of f a , around
the identity (i.e. around = = 0) up to second order gives
  
a
 a f a , b f a , b 1 2 f a ,
f , = f (0, 0) + + + b c
b b 2 b c
==0 ==0 ==0
2 a
 2 a

1 f , 1 f ,
+ b c bc + b c + O (3)
2 2 b c
==0 ==0
a
 2 f a ,

 f a (0, ) f , 0 1
f a , = f a (0, 0) + b + b + b c
b b 2 b c
=0 =0 ==0
2 a
 2 a

1 f , 1 f ,
+ b c bc + b c + O (3)
2 2 b c
==0 ==0

where O (3) denotes terms of third order i.e. proportional to 3 , 3 , 2 , 2 . From Eq. (1.24), such an expansion
becomes

f a , = 0 + a b b + a b b + f a bc b c + ga bc b c + ha bc b c + O (3)

f a , = a + a + f a bc b c + ga bc b c + ha bc b c + O (3) (1.27)
  
1 2 f a , 1 2 f a ,
1 2 f a ,
f a bc ; ga bc ; ha bc
2 b c 2 b c 2 b c
==0 ==0 ==0

setting = 0 in (1.27) we obtain


f a (0, ) = a + ga bc b c + O (3)
thus, in order to be consistent with the condition (1.24) for arbitrary values of , we require that ga bc = 0.
Similarly, setting = 0 we observe that we require that ha bc = 0. In other words, the second order terms of
the type b c or b c would violate the condition (1.24), but second order terms of the type b c are in principle
allowed. Therefore, Eq. (1.27) becomes

2 f a ,
 1
f a , = a + a + f a bc b c + O (3) ; f a bc b c
(1.28)
2
==0
16 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

From Eqs. (1.25, 1.28) the RHS of Eq. (1.26) becomes


  1  
U T f , = I + if a , Ta + f b , f c , Tbc + . . .
2
 h i 1 b  
U T f , = I + i a + a + f a bc b c + . . . Ta + + b + . . . c + c + . . . Tbc + . . .
2
 h i 1 b  
a a a b c
U T f , = I + i + + f bc Ta + + b c + c Tbc + O (3) (1.29)
2
and substituting (1.25) on the LHS of Eq. (1.26) we have
  
 a 1 b c d 1 e f
U T U (T ()) = I + i Ta + Tbc I + i Td + Tef + O (3) (1.30)
2 2

Substituting (1.29) and (1.30) in Eq. (1.26) yields

   h i
1 1 e f
I + i Ta + b c Tbc
a
I + i Td + Tef + O (3) = I + i a + a + f a bc b c Ta
d
2 2
1  
+ b + b c + c Tbc + O (3) (1.31)
2

1 1
I + i d Td + e f Tef + i a Ta a d Ta Td + b c Tbc + O (3) = I + i a Ta + i a Ta + if a bc b c Ta
2 2
1 1 1
+ b c Tbc + b c Tbc + b c Tbc
2 2 2
1 b c
+ Tbc + O (3)
2

  1h b c i   1h b c i
I + i a + a Ta + + b c Tbc b c Tb Tc = I + i a + a Ta + + b c Tbc
2 2
a b c
+ [if bc Ta + Tbc ] + O (3) (1.32)

where we have used the fact that indices of sum are dummy and that Tbc = Tcb . The terms of order 1, , , 2
and 2 match automatically in Eq. (1.32). However, by matching coefficients of in such an equation we find a
non-trivial condition

Tb Tc = [if a bc Ta + Tbc ]
Tbc = Tb Tc if a bc Ta (1.33)

therefore given the structure of the group (1.26), i.e. the functions f a , , we have its quadratic coefficient f a bc
as can be seen in (1.28). From them, the second order terms in U (T ()) [Eq. (1.25)], can be calculated from the
generators Ta that appear in the first-order terms. Moreover there is a consistency condition: the operator Tbc
must be symmetric in b and c since it is the second derivative of U (T ()) with respect to b and c, as can be
seen from Eq. (1.25). Consequently, Eq. (1.33) demands that

Tbc = Tcb = Tc Tb if a cb Ta (1.34)

substracting (1.34) from (1.33) we obtain

0 = Tb Tc if a bc Ta + Tc Tb + if a cb Ta
Tb Tc Tc Tb = i [f a bc + f a cb ] Ta
1.5. LORENTZ TRANSFORMATIONS 17

and we obtain finally


[Tb , Tc ] = iC a bc Ta ; C a bc f abc + f a cb (1.35)
the set of commutation relations in (1.35) defines a Lie algebra. It can also be proved that condition (1.35) is
the only condition required to ensure that the process can be continued. In other words, the whole power series in
(1.25) for U (T ()) can be calculated from an infinite sequence of relations like (1.33), as long as we know the first
order terms, the generators Ta . It does not necessarily mean that the operators U (T ()) are uniquely determined
for all a if we know the generators T a , but it does mean that the operators U (T ()) are uniquely determined
in at least a finite neighborhood
 of the coordinates a = 0 associated with the identity, such that Eq. (1.26) is
satisfied if , and f , are in this neighborhood. 
In some cases, it happens that the function f , satisfies the condition (at least for some subset of the
coordinates a )  
f a , = f a , = a + a (1.36)
as it is the case in space-time translations or for rotations about a given fixed axis. In that case the coefficients
f a bc in Eq. (1.28) vanish and so do the structure constants in (1.35). Hence, all generators commute

[Tb , Tc ] = 0

from (1.36) the laws of combination (1.23, 1.26) for the group representation yield
   
U T U (T ()) = U T f , = U T + = U T +
 
= U T f , = U (T ()) U T
 
U T U (T ()) = U (T ()) U T (1.37)

hence the elements of the group commute each other. Consequently, when the function f () is given by (1.36),
the connected Lie group becomes abelian. In that case, we can calculate U (T ()) for all a and not only a
neighborhood of the identity. From Eqs. (1.26, 1.36), we find

U (T (2 )) U (T (1 )) = U (T (1 + 2 ))
U (T (N )) . . . U (T (2 )) U (T (1 )) = U (T (1 + 2 + . . . + N ))

by defining i /N , we see that for any positive integer N we have


   N

U (T ()) = U T (1.38)
N

setting N , the angle /N becomes infinitesimal. Thus, we can use expansion (1.25) for U (T (/N )) keeping
only the terms at first-order in . In this way we get
 N
i
U (T ()) = lim 1 + a Ta
N N
a
U (T ()) = exp [iTa ] (1.39)

1.5 Lorentz transformations


Special relativity establishes the existence of certain special reference frames called inertial frames which are
in constant relative motion among them. Special relativity is based on two basic postulates: (a) The laws of
nature are the same in all inertial reference frames. (b) The speed of light in vacuum, measured in any inertial
reference frame is the same regardless of the motion of the light source relative to that reference frame. The
second postulate represents a significant deviation with respect to Galilean and Newtonian mechanics. It leads in
18 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

turn to different transformations connecting coordinate systems in different inertial frames. We denote as x the
coordinates in one inertial frame S, with x1 , x2 , x3 cartesian space coordinates while x0 = t is a time coordinate,
the speed of light will be settled equal to unity. We shall use latin indices such as i, j, k when running over the
three space components, and greek indices such as , , when running over the four space-time indices.
Quantities for which we obtain the same value in any inertial frame are called Lorentz invariants. A well-known
invariant in special relativity is the quantity
2 2 2 2 2 2 2
(d )2 dx1 + dx2 + dx3 (dt)2 = dx1 + dx2 + dx3 dx0 (1.40)

where the invariance of this proper time is related with the invariance of c, the speed of light in vacuum. This
Lorentz invariant can be written as

1 0 0 0 dx1
2  0 1
0 0 dx2
(d ) dx1 dx2 dx3 dx0 0
= dx g dx
0 1 0 dx3
0 0 0 1 dx0

1 0 0 0
0 1 0 0
g = 0 0 1 0
(1.41)
0 0 0 1

where g defined in (1.41) is called the metric tensor. We use from now on a convention of sum over repeated
upper and lower indices. By using this invariant, we can relate the coordinates x of S, with the coordinates x in
any other inertial frame S , in the following way

g dx dx = g dx dx (1.42)

this equation leads to


x x x x
g = g = g
x x x x
x x
g = g (1.43)
x x
it is also easy to arrive to (1.42) from (1.43). Hence Eqs. (1.43) and (1.42) are equivalent. A light wave travelling
at unit speed satisfies
dx
= 1 g dx dx = (dx)2 (dt)2 = 0
dt

and the same holds for S . Any coordinate transformation x x that satisfies Eq. (1.43) is linear

x = x + a (1.44)

with a arbitrary constants, and a constant matrix3 .


Restricting for a while to a homogeneous transformation i.e. with a = 0, the Jacobian of such a transformation
gives
x x
x = x ;

x x

further, from (1.43) we see that satisfies the condition

g = g (1.45)
The matrix depends on the velocity of S with respect to S. However, since both frames are inertial, such a velocity is constant
3

and so is.
1.6. THE INHOMOGENEOUS LORENTZ GROUP (OR POINCARE GROUP) 19

it is convenient for some purposes to write the Lorentz transformation condition (1.45) in a different way. It is
easy to check that the matrix g in (1.41) coincides with its inverse (that we denote as g ). Multiplying Eq.
(1.45) by g we find
g (g ) = g (g ) g ( g ) = (g g ) =
(g ) ( g ) = = = g g
(g ) ( g ) = g g
the relativity principle demands for the Lorentz transformations to have an inverse. Defining M g , and
multiplying with the inverse of this matrix we find4
 
M ( g ) = M g M 1 M ( g ) = M 1 M g
( g ) = g g = g
g = g (1.46)
the condition (1.45) or equivalently (1.46) is usually called the generalized orthogonality condition in the Minkowski
metric space.

1.6 The inhomogeneous Lorentz Group (or Poincare group)


The set of all Lorentz (inhomogeneous) transformations {(, a)} form a group. If we perform a Lorentz transfor-
mation (1.44) and then a second Lorentz transformation x x , then the resultant transformation x x
is described by
x = x + a = ( x + a ) + a
 
x = x + a + a (1.47)
The bar in is used to distinguish one Lorentz transformation from the other, and same for a, with respect to
a. We should show that the effect is the same as a Lorentz transformation x x . In other words, we have to
verify that (1.47) defines a Lorentz transformation. To do this, let us define
Z
we should check that Z obeys the relation (1.45). For this, we take into account that both and must obey
such a relation
   
g Z Z = g = g
= g = g

g Z Z = g
Hence Z defines a Lorentz transformation. Equation (1.47) shows us how is the composition rule for the
transformations T (, a) induced on physical states
 
T , a T (, a) = T , a + a (1.48)
each admits an inverse. To find it, we start from the definition of inverse and use again the condition (1.45) to
write

1 = = g g = g g

1 = (g g )
4
The matrix A g = g = (g) is non-singular since both g and are non-singular. Therefore its transpose
e is also invertible.
MA
20 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

Thus, condition (1.45) says that the inverse of takes the form

1 = g g (1.49)

The reader can also check that


(det )2 = 1 (1.50)
from the composition law (1.48) it is easy to see what is the identity for the transformations T (, a)

T (1, 0) T (, a) = T (1, 1a + 0) = T (, a) ; T (, a) T (1, 0) = T (1, 0 + a) = T (, a)


I = T (1, 0) (1.51)

where 1 is the identity matrix 44, and 0 the null 4-components vector. The inverse of a given T (, a) can also
be obtained from
 
T 1 , 1 a T (, a) = T 1 , 1 a 1 a = T (1, 0)
  
T (, a) T 1 , 1 a = T 1 , 1 a + a = T (1, 0)

T 1 (, a) = T 1 , 1 a (1.52)

1.6.1 Four-vectors and tensors


The Lorentz invariant (1.40) suggests to define a Lorentz invariant norm for the 4-components vectors x as follows
2 2 2 2 
kxk2 (x, x) = x0 + x1 + x2 + x3 = x , x (1.53)
2
strictly speaking this is a pseudonorm because it is not positive-definite since in some cases x0 > xi xi . Taking
into account (1.41) this relation can be written as

(x, x) = x g x (1.54)

it is easy to see the invariance of this norm under a homogeneous Lorentz transformation, by using condition
(1.45) we find
  
x , x = x g x = ( x ) g x = (g ) x x = g x x

x , x = x g x = x g x = (x,x) (1.55)

a more convenient way to write this norm is the following

(x,x) = x g x x x ; x g x (1.56)
 
x = x1 , x2 , x3 , x0 ; x = (x1 , x2 , x3 , x0 ) = x1 , x2 , x3 , x0 (1.57)

we define a four-vector as any arrangement of four components that under a homogeneous Lorentz transforma-
tion, changes under the same prescription of x . That is V is a four-vector if under a homogeneous Lorentz
transformation we have
V = V (1.58)
for any contravariant four-vector V we can define a covariant four-vector as in Eq. (1.56)

V g V

multiplying by the inverse of g we obtain the inverse relation

g V = g g V = V = V
V = g V
1.6. THE INHOMOGENEOUS LORENTZ GROUP (OR POINCARE GROUP) 21

we can define the inner product between two (contravariant) four vectors V ,W as

(V, W ) V g W = V W

with the same procedure used to prove the Lorentz invariance of (x, x), we see that (V, W ) is also Lorentz invariant.
The summation of upper and lowered indices is called a contraction.
The Lorentz transformation of a covariant four-vector gives
 
V = g V = g V = g g V = g g V

where we have used the symmetrical nature of g . Applying Eq. (1.49) we obtain

V = V = 1 V (1.59)

so covariant four-vectors transform with the inverse transformation with respect to contravariant four-vectors. It
justifies the names covariant and contravariant.
Two adjacent four-vectors transform as
 
V W = ( V ) W = V W (1.60)

an arrangement of numbers characterized by two indices of the form T is called a second-rank Lorentz tensor if
under a homogeneous Lorentz transformation, it changes in a way similar to the two adjacent four-vectors in Eq.
(1.60), that is
T = T (1.61)
taking into account the expression (1.49) for the inverse of , we can write this transformation as

T = T = T 1 (1.62)

in matrix from it yields


Tcont = Tcont 1
which is a similarity transformation of Tcont under . We can define in an analogous way second-rank covariant
tensors T based on covariant four-vectors. By using Eq. (1.59) we have
 1   h 1  i 
1

1
V W = V V = V V


 
T = 1 1 T (1.63)

which can also be written as



  
T = 1 T 1 = T 1 (1.64)
Tcov = Tcov 1
(1.65)

from Eqs. (1.62, 1.64), it is easy to show that the contraction of two second rank tensors (one contravariant and
other covariant) is a Lorentz invariant (also known as a Lorentz scalar or a zero-rank Lorentz tensor)
h  ih  i  
T H

= T 1 H 1 = 1 1 T H
 1   h 1  i
= T H
= T H = T H = T H
22 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

thus we can denote this contraction with a quantity without Lorentz indices, and it is equal in any inertial reference
frame
T H = T H

=C

in a similar way, we can show that the contraction of a second-rank tensor T with a four-vector V , gives a
four-vector W
T V = W (1.66)

the previous developments justify the convention that indices may be lowered or raised by contraction with g
or g . For instance
T g T ; T g T (1.67)

a very important four-vector in special relativity is the four-momentum


 
p = p1 , p2 , p3 , p0 = p, p0 (p, E) , p = (p, E) (1.68)

hence p is the three momentum and the energy is the zeroth component. Taking into account the fundamental
relation
p2 + m2 = E 2 (1.69)

the pseudo-norm of the four momentum gives


2
p2 = p g p = p p = p0 + p2 = E 2 + p2
p2 = m2 (1.70)

note that four-momenta with positive pseudonorm leads to m2 > 0, i.e. to a non-physical mass. Thus, physical
states are related with four-momenta with non-positive pseudonorm. It is important to keep in mind that if the
metric tensor is chosen as = (1, 1, 1, 1), we obtain p2 = m2 and the positive values of p2 are the physical
ones. Hence, it is extremely important to know the conventions used in each case.

1.7 Some subgroups of the Poincare group


The whole group of transformations T (, a) is called the inhomogeneous Lorentz group, or the Poincare
group. It has several important subgroups. First the set of transformations with a = 0, {T (, 0)} clearly forms a
subgroup. To see it we observe first that the identity T (1, 0) of {T (, a)} is also contained in {T (, 0)}. Further
Eq. (1.48) gives us the composition law, on this subset
  
T , 0 T (, 0) = T , 0 + 0 = T , 0

and the composition law is closed within the subset {T (, 0)}. Finally, the inverse of any given element in
{T (, 0)}, also belongs to such a subset as can be seen from (1.52)
 
T 1 (, 0) = T 1 , 1 0 = T 1 , 0

this subgroup of the inhomogeneous Lorentz group is called the homogeneous Lorentz group. When we work
on the homogeneous Lorentz group, we usually simplify the notation and write

T (, 0) T () ; I T (1)
  
T T () = T ; T 1 (, 0) = T 1
1.7. SOME SUBGROUPS OF THE POINCARE GROUP 23

further, we note that Eq. (1.50) gives two possibilities (a) det = +1 (b) det = 1. Those transformations
with det = +1 obviously form a subgroup of either the homogeneous or inhomogeneous Lorentz groups5 . On
the other hand, by taking the 00 components of Eq. (1.45) we have

g 0 0 = g00 (1.71)

expanding the LHS explicitly and taking into account that g is diagonal, we obtain

g 0 0 = g 0 0 = g00 0 0 0 0 + gii i 0 i 0
2
g 0 0 = i 0 i 0 0 0 (1.72)

where we are using sum over repeated (upper) indices i. Substituting (1.72) into (1.71) and using g00 = 1, we
find 2 2
i 0 i 0 0 0 = 1 0 0 = 1 + i 0 i 0
2
similarly from Eq. (1.46) we can find that 0 0 = 1 + 0 i 0 i . We obtain finally
2
0 0 = 1 + i 0 i 0 = 1 + 0 i 0 i (1.73)

since the matrices are real, we have 0 i 0 i 0. Hence, we see from (1.73) that 0 0 1. Consequently, we
have that either 0 0 1 or 0 0 1.

1.7.1 Proper orthochronous Lorentz group


Those transformations with 0 0 1 form a subgroup. To show it, we assume by hypothesis that both and
satisfy the conditions
0 0 1, and 0 0 1 (1.74)
we have 0
0 = 0 0 = 0 0 0 0 + 0 i i 0 (1.75)
 0
if 0 i i 0 0, Eqs. (1.74, 1.75) yield immediately that 0 1. Now we should examine the case in which
0 i i 0 < 0. First we prove the inequality
p p
a + b + ab a (a + 2) b (b + 2) 0 ; if a 0 and b 0 (1.76)

we prove it as follows

(a b)2 0 (a + b + ab)2 a (a + 2) b (b + 2) 0
(a + b + ab)2 a (a + 2) b (b + 2)

if a, b are non-negative, we can take positive square roots on both sides and preserve the order relation in the
inequality, thus p p
a + b + ab a (a + 2) b (b + 2) ; if a 0 and b 0
then we obtain Eq. (1.76).
Defining the three vectors
 
v 1 0 , 2 0 , 3 0 ; v 0 1 , 0 2 , 0 3 (1.77)
p p
kvk = v v = i 0 i 0 ; kvk = v v = 0 k 0 k (1.78)
5
It is also clear that the set of Lorentz transformations with det = 1, does not form a subgroup of the Lorentz group. For
instance, it does not contain the identity.
24 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

Eq. (1.73) shows that the lengths of these two three-vectors are given by
q q 2
kvk = (0 0 )2 1 ; kvk = 0 0 1 (1.79)

and using the inequality


|v v| kvk kvk (1.80)
0 i
substituting (1.77) and (1.79) in (1.80) and taking into account that i 0 0, we have
0   q q 2
1 , 0 2 , 0 3 1 0 , 2 0 , 3 0 0 2
( 0 ) 1 0 0 1
0 i q q 2
i 0 0 2
( 0 ) 1 0 0 1
0 i q q 2
0 2
i 0 ( 0 ) 1 0 0 1
q q 2
0 i 0 2
i 0 ( 0 ) 1 0 0 1 (1.81)

substituting (1.81) in (1.75), we find


0
0 = 0 0 0 0 + 0 i i 0
0 q q 2
0 0 0 0 0 (0 0 )2 1 0 0 1 (1.82)

since 0 0 1 and 0 0 1, we can express them as

0 0 = 1 + a , 0 0 = 1 + b ; a0, b0 (1.83)

using (1.83) on the RHS of (1.82) we have


q q 2 q q
2
0 0 0
0 0 ( 0 ) 1 0 1 = (1 + a) (1 + b) (1 + a) 1 (1 + b)2 1
0 2
q q 2 p p
0 0 0 2
0 0 ( 0 ) 1 0 0 1 = 1 + a + b + ab a (a + 2) b (b + 2)

and using Eq. (1.76) we find q q 2


0 0 0 0 (0 0 )2 1 0 0 1 1 (1.84)
0
and combining Eqs. (1.82, 1.84) we find 0 1, and the subset of Lorentz transformations with 0 0 1
is closed under composition. It is left to the reader to show that the inverse of an element in this subset, also
belongs to the subset [homework!!(2)]. Hence such a subset forms a subgroup.
The subgroup with det = +1 and 0 0 +1, is known as the proper orthochronus Lorentz group
(either homogeneous or inhomogeneous). It is not possible by a continuous change of parameters to jump from
det = +1 to det = 1 (or vice versa), neither from 0 0 +1 to 0 0 1. Consequently, any Lorentz
transformation that can be obtained through a continuous change of parameters from the identity, must have the
same sign of det and 0 0 as the identity. Therefore, those elements connected continuously with the identity
must belong to the proper orthochronus Lorentz group.

1.7.2 Discrete transformations in the Lorentz group


Any Lorentz transformation is either proper orthochronus, or can be written as an element of the proper or-
thochronus subgroup with one of the discrete transformations P, T or PT where P is the parity or space inversion
operator, which is diagonal with elements

P 0 0 = 1 , P i i = 1 (1.85)
1.8. QUANTUM LORENTZ TRANSFORMATIONS 25

while T is the time-reversal matrix, also diagonal with elements

T 0 0 = 1 , T i i = 1 (1.86)

consequently, the study of the whole Lorentz group reduces to the study of the (connected) proper orthochronus
Lorentz group, along with the study of the discrete space inversion and time reversal operators. By now, we shall
study the proper orthochronus Lorentz group, which is a connected Lie group.

1.7.3 Infinitesimal transformations within the proper orthochronus Lorentz group


Much of the information about connected Lie groups can be extracted by examining the behavior of the group
elements in a neighborhood of the identity. For the inhomogeneous Lorentz group, the identity is characterized
by the transformations
= , a = 0 (1.87)
therefore a neighborhood of the identity can be written as

= + , a = (1.88)

where and are infinitesimals. The Lorentz condition (1.45) taken over the infinitesimal transformations
(1.88) gives

g = g ( + ) ( + ) = g + g + g + g

= g + g + g + O 2

g = g + + + O 2

therefore, keeping terms up to first order, the Lorentz condition lead to

= (1.89)

hence is a second-rank antisymmetric tensor, whose degrees of freedom are given by N (N 1) /2. For N = 4
(dimensions) we obtain 6 independent components. By adding the 4 degrees of freedom associated with the
components of , we obtain a total of 10 parameters for an inhomogeneous Lorentz transformation6 .

1.8 Quantum Lorentz Transformations


By now we shall restrict ourselves to proper orthochronus Lorentz transformations for which all transformations
are continuous. As we have already seen, in the framework of quantum mechanics the elements T (, a) of the
Lorentz group are operators applied on rays. However, in practice we transform kets of the Hilbert space and
not rays. Thus we should construct a representation U (, a) consisting of sets of operators acting on the Hilbert
space. From the Wigners representation theorem and the connected nature of the proper orthochronus Lorentz
group, we see that the transformations T (, a) induce a linear unitary (rather than an antilinear antiunitary)
transformation on vectors of the Hilbert space of states

|i U (, a) |i

in order to carry the group structure, the operators U (, a) must respect the composition law (1.48)
 
U , a U (, a) = U , a + a (1.90)
6
In a more physical point of view, we require three parameters for a boost (the components of a velocity), three for a rotation
(e.g. the Euler angles), and four to perform a space-time translation. Hence, six parameters belong to the homogeneous Lorentz
transformations (boosts and rotations), and four parameters belong to the inhomogeneous part.
26 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

strictly speaking we could also form projective representations or representations up to a phase as shown in Eq.
(1.15). In general, it is necessary to enlarge the Lorentz group to avoid the appearance of phase factors on the
RHS of Eq. (1.90). The inverse of U (, a) is indicated by Eq. (1.52)

U 1 (, a) = U 1 , 1 a (1.91)

Since U (1, 0) carries any ray into itself, it must be proportional to the unit operator. By an apropriate choice
of phase, it can be settled as equal to the identity.

1.8.1 Four-vector and tensor operators


In quantum Mechanics, observables are eigenvalues of a complete hermitian operator. For example the three
components of the classical linear momentum is replaced by a corresponding set of three-Hermitian operators

p (p1 , p2 , p3 ) P (P1 , P2 , P3 )

for the energy we have the Hamiltonian. On the other hand, we have seen that the four components consisting of
the three momentum components plus the energy forms a four-vector in special relativity. Thus, the corresponding
arrangement in quantum mechanics

p (p1 , p2 , p3 , E) P (P1 , P2 , P3 , H)

could be taken as the prototype of a four-vector operator. Thus, we shall study the transformation of P
under a quantum Lorentz transformation in order to define other four-vector operators as the ones that under a
Lorentz transformation change in the way prescribed by P . We then start with four-momentum eigenstates

P |pi = p |pi (1.92)

Then we apply a quantum homogeneous Lorentz transformation U () on both sides of (1.92)

U () P |pi = p U () |pi (1.93)

it is important to emphasize that U () acts on vectors of the Hilbert space and not on four-vectors p of the
Minkowski space. Owing to it, U () passes by the eigenvalue p . Inserting an identity in Equation (1.93) we
have   
U () P U 1 () U () |pi = p U () |pi U () P U 1 () {U () |pi} = p U () |pi
which can be rewritten as

U () p = p p ; p U () |pi , U () U () P U 1 () (1.94)

it is convenient to introduce the eigenvalues p associated with the transformed state |p i. Obviously, the trans-
formation p p must be carried out with the transformation in the Minkowski space associated with the
quantum Lorentz transformation U (). Therefore

p = p p = 1 p = p (1.95)

substituting (1.95) in (1.94) we find



U () p = p p = P p

since this is valid for |p i arbitrary we obtain

U () = P
1
U () P U () = P

this induces the following definition


1.8. QUANTUM LORENTZ TRANSFORMATIONS 27


Definition 1.2 (four-vector operators) An arrangement of four component operators O O1 , O2 , O3 , O0 is
called a (contravariant) four-vector operator if under a quantum homogeneous Lorentz transformation, it changes
under the formula
U () O U 1 () = O (1.96)

By superposing four-vector operators we can define second-rank (or higher order) Lorentz tensors, in a similar
way as we did in section 1.6.1. For instance, a contravariant second-rank Lorentz tensor transform under a
homogeneous Lorentz transformations as
T = T (1.97)
If we compare Eq. (1.58) that defines a contravariant four-vector in Minkowski space with Eq. (1.96) that
defines a contravariant
 four-vector operator on a Hilbert space, we realize that the RHS of these equations differ
since = 1 . Something similar can be seen by comparing (1.61) with (1.97) for second-rank contravariant
tensors.

1.8.2 Infinitesimal quantum Lorentz transformations


We shall use the formalism developed in Sec. 1.4 about connected Lie groups. In such a section, we saw that
much of the properties of the group structure can be developed from expansions of the elements of the group up
to first order in the parameters. For an infinitesimal Lorentz transformation (1.88) the corresponding operator
U (, a) acting on the Hilbert space, is given by

U (, a) = U (1 + , ) (1.98)

and must be equal to the identity plus terms linear in and , that we shall parameterize as in Eq. (1.25)7

1 
U (1 + , ) = 1 + i J i P + O 2 , 2 , (1.99)
2
where J , P are operators independent of the parameters and . For U (1 + , ) to be unitary, the operators
J , P must be hermitian
J = J ; P = P (1.100)
since is antisymmetric, we can take the operators J to be antisymmetric as well8

J = J (1.101)

on the other hand, the expansion of U 1 (1 + , ) up to first order is given by

1 
U 1 (1 + , ) = 1 i J + i P + O 2 , 2 , (1.102)
2
it can be seen by multiplying Eqs. (1.99, 1.102), and observing that we obtain the identity plus terms of second
order in and/or . In addition each parameter is accompanied by an associated generator J and
P . We have ten independent parameters and so ten independent generators (owing to the antisymmetry
of and J ).
7
Comparing Eqs. (1.25, 1.99) we see that we have chosen our parameters in the form a ( , ).
8
As a matter of example, for a given couple of numbers say 2, 3; what we have to fix is the quantity
1   23  23  23 (23)
i 23 J 23 + 32 J 32 = i J J 32 iJ
2 2 2
so that only J (23) is fixed. Thus we can choose J 23 J (23) /2, which leads to J 23 = J 32 .
28 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

1.8.3 Lorentz transformations of the generators


We now examine the Lorentz transformation properties of the generators J , P of the continuous Poincare
group. To do it, we examine the transformation properties of U (1 + , ) induced by another new (and in general
finite) transformation U (, a)
U (, a) U (1 + , ) U 1 (, a) (1.103)
of course, (, a) are totally independent of (, ). From Eqs. (1.90, 1.91) we can write this product as

U (, a) U (1 + , ) U 1 (, a) = [U (, a) U (1 + , )] U 1 , 1 a

= [U ( (1 + ) , + a)] U 1 , 1 a

= U (1 + ) 1 , (1 + ) 1 a + + a

= U 1 + 1 , a 1 a + + a

U (, a) U (1 + , ) U 1 (, a) = U 1 + 1 , 1 a

then we have
U (, a) U (1 + , ) U 1 (, a) = U (1 + , ) (1.104)
1 , 1 a (1.105)
on the other hand, by using the expansion (1.99) on the LHS of Eq. (1.104) we find
 
1
U (, a) U (1 + , ) U (, a) = U (, a) 1 + i J i P U 1 (, a)
1
2
 
1
U (, a) U (1 + , ) U (, a) = 1 + U (, a) i J i P U 1 (, a)
1
(1.106)
2
now using the expansion (1.99) on the RHS of Eq. (1.104) we obtain
1
U (1 + , ) = 1 + i J i P (1.107)
2
equating Eqs. (1.106, 1.107) and using the definitions (1.105) we find
 
1 1  
U (, a) J P U 1 (, a) =

1 J 1 a P (1.108)
2 2
now, equating coefficients of on both sides of Eq. (1.108), we have
 
1 1  
U (, a) J U 1 (, a) = 1 J + 1 a P
2 2
1  1  1 
= 1 J + 1 a P + 1 a P
2 2 2
  
U (, a) J U 1 (, a) = 1 J + 1 a P + 1 a P (1.109)
manipulating the RHS of Eq. (1.109), and using Eq. (1.49) as well as the antisymmetry of , we have
  
K 1 J + 1 a P + 1 a P
 h  i h  i
= 1 J + 1 a P + 1 a P
        
K = 1 J + 1 a P + 1 a P
= [ ] J + [ a ] P + [ a ] P
= J [ a ] P + a P
K = [J a P + a P ] (1.110)
1.8. QUANTUM LORENTZ TRANSFORMATIONS 29

substituting (1.110) in (1.109), and taking into account that is infinitesimal but otherwise arbitrary, we find
U (, a) J U 1 (, a) = [J a P + a P ]
now, equating coefficients of on both sides of Eq. (1.108), we have
U (, a) P U 1 (, a) = () P = ( ) P
U (, a) P U 1 (, a) = P
putting them together we have
U (, a) J U 1 (, a) = [J a P + a P ] (1.111)
1
U (, a) P U (, a) = P (1.112)
For homogeneous Lorentz transformations (a = 0), Eqs. (1.111, 1.112) give
U () J U 1 () = J ; U () P U 1 () = P (1.113)
comparing with Eqs. (1.96, 1.97), the transformations (1.113) say that J is a second-rank tensor operator and
P a four-vector operator. On the other hand, by using pure translations ( = 0, = ), Eqs. (1.111,
1.112) give
U (1, a) J U 1 (1, a) = [J a P + a P ] = J a P + a P
U (1, a) P U 1 (1, a) = P
hence P is translation-invariant but J is not. In particular, by applying Eqs. (1.111, 1.112) to the space-space
components of J (i.e. components of the type J ij ), we obtain the usual change of angular momentum under a
change of the origin relative to which we calculate such an angular momentum [homework!!(3)].

1.8.4 Lie algebra of the Poincare generators


Now we apply transformations (1.111, 1.112) to the case in which U (, a) is infinitesimal by itself, so that
= +
e , a = e (1.114)
where we use the notation (e , e), to emphasize that this infinitesimal parameters are totally independent of the
ones previously used. By applying the expansions (1.99, 1.102) on the LHS of Eq. (1.111) and keeping terms up
to first order in (e
, e) we have
   
1 1 1
J U (1 + e , e) J U (1 + e , e) = 1 + ie J ie P J 1 ie J + ie P
2 2
  
1 1
= J + ie J J ie P J 1 ie J + ie P
2 2
1 1 
= J ie J J + ie
J P + ie J J ie P J + O e 2 , e2 ,
e e
2 2
1 1

= J + ie J J ie J J + ie J P ie P J + O e 2 , e2 ,
e e
2 2
1 h i 
J = J + ie J , J ie [P , J ] + O e 2 , e2 ,
e e (1.115)
2
Further, using (1.114) on the RHS of Eq. (1.111), and keeping terms up to first order in (e , e) we have
J = ( +
e ) ( +
e ) [J e P + e P ]

= ( + e ) [J e P + e P ] + ( + e )
e J + O e 2 , e2 ,
e e

= ( + e ) [J e P + e P ] + e J + O e 2 , e2 ,
e e

J = [J e P + e P ] + e J + e J + O e 2 , e2 ,
e e (1.116)
30 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

and using the symmetry of g and the antisymmetry of we find



J = J e P + e P +
e g J +
e g J + O
e 2 , e2 ,
e e

= J g e P + g e P
e g J
e g J + O
e 2 , e2 ,
e e
  
J = J + e (g P g P ) e g J + g J + O e 2 , e2 ,
e e (1.117)

Alternatively, from (1.116) we can also write



J = J e P + e P +
e g J +
e J + O
e 2 , e2 ,
e e

= J + e (g P g P ) +
e g J +
e g J + O e 2 , e2 ,
e e

J = J + e (g P g P ) +
e (g J + g J ) + O
e 2 , e2 ,
e e (1.118)

equating Eqs. (1.115, 1.117) we have


1
[J , J ] + ie
ie [J , P ] = e (g P g P )
e (g J + g J )
2
since (e
, e) are infinitesimal but otherwise arbitrary we equate coefficients of
e , and e on both sides of this equation
1
i [J , J ] = (g J + g J ) (1.119)
2
i [J , P ] = (g P g P ) (1.120)

on the other hand, equating Eqs. (1.115, 1.118), we obtain the same condition with e, but equating the coefficients
of
e we find
1
i [J , J ] = (g J + g J ) (1.121)
2
Equations (1.119) and (1.121) are essentially identical. We can put such equations in a more symmetrical form
by adding them, from which we have

i [J , J ] = (g J + g J ) + (g J + g J )

performing a similar procedure from Eq. (1.112) we reproduce the result (1.120) and obtain the additional
condition
[P , P ] = 0 (1.122)
collecting all equations (1.121, 1.120, 1.122) and taking into account that g is symmetric, we obtain

i [J , J ] = g J g J g J + g J (1.123)

i [P , J ] = g P g P (1.124)

[P , P ] = 0 (1.125)

1.8.5 Physical interpretation of Poincares generators


In some senses the physical interpretation of the operators J and P is easier in a three-dimensional notation.
Conserved quantities in quantum mechanics are related with operators that commute with the Hamiltonian or
energy operator H = P 0 . We define then the momentum three-vector

P P 1, P 2, P 3 (1.126)

and the angular momentum three-vector



J J 23 , J 31 , J 12 {J1 , J2 , J3 } J km = kmn Jn (1.127)
1.8. QUANTUM LORENTZ TRANSFORMATIONS 31

the energy operator (Hamiltonian)


H = P0 (1.128)

and the remaining generators form what is called the boost three-vector

K J 01 , J 02 , J 03 {K1 , K2 , K3 } J 0i = Ki (1.129)

So we have all ten degrees of freedom (the remaining components of J are not independent because of the
antisymmetry of it). In a three-dimensional notation, the commutation relations (1.123, 1.124, 1.125) can be
written as [homework!!(4)]

[Ji , Jj ] = iijk Jk (1.130)


[Ji , Kj ] = iijk Kk (1.131)
[Ki , Kj ] = iijk Jk (1.132)
[Ji , Pj ] = iijk Pk (1.133)
[Ki , Pj ] = iHij (1.134)
[Ji , H] = [Pi , H] = [H, H] = 0 (1.135)
[Ki , H] = iPi (1.136)
[Pi , Pj ] = 0 (1.137)

from relations (1.135) we can observe that operators P, J commute with H, and so they are conserved. However
Eq. (1.136) shows that K is not conserved. Consequently, we shall not use the eigenvalues of K to label physical
states. The commutation relations (1.130) forms the well-known Lie algebra of angular momentum operators.
Moreover commutation relations (1.133, 1.137), coincides with the ones expected between a linear momentum
operator and an orbital angular momentum operator.
Commutation relations (1.130), show that the set of generators Ji forms a closed algebra, such an algebra in
turn generates a subgroup [the subgroup of three-dimensional rotations SO (3)]. In the same way, commutation
relations (1.137) define a closed algebra of the Pi generators, since these generators commute with each other, the
subgroup generated (the subgroup of space translations), is abelian. However, we usually work with the subgroup
of space-time translations instead of the group of space translations. Further, commutation relations (1.132) show
that boosts generators do not form a closed algebra and do not generate a subgroup. Indeed, Eq. (1.132) shows
the well-known feature that two succesive boosts might generate a rotation.
For = 1, we obtain pure space-time translations. The set T (1, a) of all pure translations forms an abelian
subgroup of the inhomogeneous Lorentz group with a group multiplication rule given by Eq. (1.48)

T (1, a) T (1, a) = T (1, a + a) (1.138)

considering that this subgroup is characterized by only four parameters a , the multiplication rule is given by a
relation of the type (1.36) as it corresponds to an additive abelian subgroup. Hence we can use Eq. (1.39) valid
for rules of multiplication of the type (1.36). Thus, any finite translation is represented on the physics Hilbert
space by
U (1, a) = exp (iP a ) (1.139)

in the same way, a rotation R by an angle || around the direction , is represented on the physical Hilbert space
by
U (R , 0) = exp (iJ) (1.140)

to obtain (1.140) the argument is the same as to obtain (1.139), if we take into account that rotations around a
fixed axis commute each other.
32 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

1.9 One-particle states


We shall classify one-particle states according with their transformation properties under inhomogeneous Lorentz
transformations.
Equation (1.125), shows that the components of the four-momentum commute with each other. Therefore,
they admit a complete set of common eigenstates that we denote as |p, i, where denotes the remaining degrees
of freedom. Then our starting point will be momentum eigenstates

P |p, i = p |p, i (1.141)

in the most general states (for instance those describing several unbound particles), the remaining labels could be
either discrete or continuous and even both. Nevertheless, we shall take as part of the definition of a one-particle
state, that the label is purely discrete, and we shall restrict ourselves to study that case. In particular, specific
bound states of two or more particles, such as the lowest state of the hydrogen atom, is to be considered as a
one-particle state though this is not an elementary particle. By now we shall not distinguish between composite
or elementary particles.
On the other hand, all eigenstates of P with a given eigenvalue p0 (plus the null vector) form a subpace Ep0 of
the Hilbert space (see section 1.1.1). The dimensionality of such a subspace is equal to the degree of degeneracy
of the eigenvalue p0 , that is the number of linearly independent vectors of the form |p0 , i for a fixed p0 .

1.9.1 One-particle states under pure translations


We shall start the characterization of states of the type |p, i, under inhomogeneous Lorentz transformations.
From Eqs. (1.141, 1.139) we see the action of pure translations over those states

U (1, a) |p, i = exp [ia P ] |p, i = eipa |p, i (1.142)

therefore, |p, i is an eigenstate of pure translation operators U (1, a). In addition, we observe that Ep is the
subspace induced by each eigenvalue eipa . Even more, the one-dimensional subspaces Ep, are invariant9 (and
of course minimal) under the representation {U (1, a)} of the subgroup of pure translations. Since any given
minimal subspace invariant under {U (1, a)} is one-dimensional, we say that the vectors |p, i are singlets with
respect to {U (1, a)}. This is consistent with the fact that all irreducible representations of any abelian group are
one-dimensional (see theorem 1.2 page 1.2).

1.9.2 One-particle states under homogeneous Lorentz transformations


As for quantum homogeneous Lorentz transformations U (, 0) U (), equation (1.112) shows that they trans-
form a state |p, i in another momentum eigenstate but with eigenvalue p
   
P U () |p, i = U () U 1 () P U () |p, i = U () U 1 () P U () |p, i
   h  i
= U () U 1 P U 1 1 |p, i = U () 1 P |p, i
= p [U () |p, i]
P [U () |p, i] = (p) [U () |p, i]

hence U () |p, i must be an eigenstate of P with eigenvalue p. From the previous facts, we could say that
U () |p, i belongs to the subspace Ep . Hence it is a linear combination of a basis within this subspace
X
U () |p, i = C , (, p) p, (1.143)

9
The subspace Ep, consists of all vectors of the form |p, i, with (p, ) fixed and with running over all complex scalars.
1.9. ONE-PARTICLE STATES 33

equation (1.143) shows that, by using the basis {|p, i} we can find a matrix representation of U (), i.e. a
matrix representation of the Lorentz group in the Hilbert space E. The idea now is to characterize the irreducible
representations and also the minimal invariant subspaces of E under {U (, a)}. Therefore, we should look for the
apropriate canonical basis in which the reduction is apparent. It is natural to associate states of a specific particle
type with the components of a irreducible representation of the inhomogeneous Lorentz group. It could happen
that different particle species may correspond to isomorphic representations10 .
First we note that a subspace of the type Ep is not invariant under {U ()}, since a given element |p, i Ep is
mapped through U () into an element that belongs to another subspace Ep . Notice however that this mapping
preserves the norm (or pseudonorm) of the four-vector p i.e. the quantity p2 g p p . In addition, if p2 0
then the sign of p0 is also preserved [homework!!(5)]. It is then convenient for each value of p2 and (for p2 0)
each sign of p0 to choose a standard four-momentum k . Hence we write any p of this class11 as

p = L (p) k (1.144)

where L is a Lorentz transformation that connects two four-vectors within the same class. It is clear that such
a transformation must depend on p, but also implicitly on the chosen standard vector k. Equation (1.143) shows
that a quantum operator U () takes p into another element of the same class. Then we can define the states
|p, i of momentum p, as a transformation from the reference point |k, i such that

|p, i N (p) U (L (p)) |k, i (1.145)

where N (p) is a constant of normalization that we shall choose later. By comparing Eqs. (1.144, 1.145) we see
that if we associate p of the Minkowski space with |p, i of the Hilbert space, and the standard vector k is
associated with |k, i, it is logical that the transformation L (p) on the Minkowski space that connects k with p
should have an associated operator U (L (p)) that connects |k, i with |p, i, except for a possible normalization
constant. Note that the defined operator U (L (p)) only transforms the momenta degrees of freedom but not the
other degrees of freedom . Indeed, equation (1.145), says how the labels are related for different momenta.
Applying a homogeneous Lorentz transformation U () on Eq. (1.145) we find

U () |p, i N (p) U () U (L (p)) |k, i = N (p) U (L (p)) |k, i


 
= N (p) U (L (p)) U 1 (L (p)) U (L (p)) |k, i
  
= N (p) U (L (p)) U L1 (p) U (L (p)) |k, i

U () |p, i = N (p) U (L (p)) U L1 (p) L (p) |k, i (1.146)

observe that the Lorentz transformation L1 (p) L (p) transforms k into itself as can be seen by applying Eq.
(1.144)
 1   
L (p) L (p) k = L1 (p) L (p) k = L1 (p) p
= L1 (p) (p) = k

so it belongs to the set of all Lorentz transformations that leave k invariant

W k = k (1.147)
10
Sometimes it is even convenient to define particle types as irreducible representations of a group that contains the proper or-
thochronus Lorentz group as a proper subgroup. For example, we shall see later that for massless particles with space inversion
symmetry, it is usual to associate a given irreducible representation of the proper orthochronus inhomogeneous Lorentz group includ-
ing space inversion, with a single-particle type.
11
We are forming a partition of the set of all four-vectors {p} (that is, a disjoint collection of subsets whose union equals the set).
For p2 > 0, a subset Spb2 (or class) of this collection consist of all vectors with a fixed (positive) norm p2 . For p2 0, we define
2
partitions of the form Spa+
2 consisting of all vectors with a fixed (negative) norm p and p0 > 0. Finally a class Spa2 consists of all

vectors with a fixed (negative) norm p and p < 0. Of course, to obtain the set of all {p} we must run over all values of p2 within
2 0

each subset.
34 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

this set forms a group [homework!!(6)] known as the little group. For any W belonging to the Little group i.e.
satisfying (1.147), equation (1.143) becomes
X
U (W ) |k, i = D (W ) k, (1.148)

We have said that subspaces of the type Ep are not invariant under homogeneous Lorentz transformations.
Equation (1.148), shows that by restricting to the little group W (in the Minkowski space), we are generating
invariant subspaces Ek under the representation U (W ) of the little group in the Hilbert space.
It can be seen that the coefficients D (W ) provides a representation of the little group, we see it as follows
X   
D W W k, = U W W |k, i = U W U (W ) |k, i

X X   
= U W D (W ) k, = D (W ) U W k,

" # (" # )
X X  X X   
= D (W ) D W k, = D W D (W ) k,

X  X    
D W W k, = D W D (W ) k,

and resorting to the linear independence of states |k, i we finally obtain


   
D W W = D W D (W )

we have already seen that the transformation

W (, p) L1 (p) L (p) (1.149)

belongs to the little group. Substituting (1.149) in Eq. (1.146), and using (1.148) we have
" #
X
U () |p, i = N (p) U (L (p)) [U (W (, p)) |k, i] = N (p) U (L (p)) D (W (, p)) k,

" #
X
U () |p, i = N (p) D (W (, p)) U (L (p)) k,

and from the definition (1.145) we obtain

X X  



 |p, i
U () |p, i = N (p) D (W (, p)) U (L (p)) k, = N (p) D (W (, p))
N (p)

N (p) X
U () |p, i = D (W (, p)) p, (1.150)
N (p)

comparing Eq. (1.150) with Eq. (1.143) we observe that apart from the problem of the normalization, we have
replaced the problem of determining the coefficients C, in the transformation rule (1.143), to the (reduced)
problem of determining the coefficients D, which are the matrix elements associated with the representation of
the little group [see Eq. (1.148)]. This approach of deriving representations of a group (the homogeneous Lorentz
group in our case) from the representations of a little group, is called the induced representation method.
1.9. ONE-PARTICLE STATES 35

Type of p Standard k Little group


(a) p2 = M 2 < 0, p0 > 0 (0, 0, 0, M ) SO (3)
2 2 0
(b) p = M < 0, p < 0 (0, 0, 0, M ) SO (3)
(c) p2 = 0, p0 > 0 (0, 0, , ) ISO (2)
2 0
(d) p = 0, p < 0 (0, 0, , ) ISO (2)
(e) p2 = N 2 > 0 (0, 0, N, 0) SO (2, 1)
(f) p = 0 (0, 0, 0, 0) SO (3, 1)
Table 1.1: This table displays the six types of classes of four-momenta. We also display a convenient choice for
the standard four-vector k , and the most general subgroup of the proper orthochronus Lorentz group that leave
k invariant (little group). The quantities M, N, , are all positive.

1.9.3 Physical little groups


It is clear that the partition we have defined consists of six types of classes12 : (a) classes of the type Spa+
2 in which

p = M < 0, and p > 0. (b) classes of the type Sp2 in which p = M < 0, and p < 0. (c) classes S0a+ with
2 2 0 a 2 2 0

p2 = 0, and p0 > 0. (d) classes S0a with p2 = 0, and p0 < 0. (e) Classes Spb2 with p2 = N 2 > 0, finally (f) the
class S00 with p = 0. The numbers M and N are positive.
Now, it is well-known that the case p2 > 0 lead to a non-physical mass [see Eq. (1.70) and discussion below],
and p2 0 with p0 < 0 is a non-physical configuration. Therefore, only the cases (a), (c), (f) leads to an
interpretation in terms of physical states (as far as we know). For each class we want to establish a suitable
standard vector k and the associated little groups.
For the class (a) with p2 = M 2 < 0, and p0 > 0 a natural standard vector is k (0, 0, 0, M ). The little
group is the set of all homogeneous proper orthochronus Lorentz transformations that leave the vector (0, 0, 0, M )
invariant. This k describes a particle with non-null mass at rest. It is a fact that a proper orthochronus Lorentz
transformation can always be expressed as a composition of a boost followed by a rotation in the three space
coordinates. From physical grounds it is pretty obvious that a transformation containing a boost does not leave
this four-vector invariant, since in that case we are changing to another reference frame with relative velocity with
respect to the first, in this new reference frame the particle is not at rest any more. By contrast, a rotation of
the three space coordinates keeps the particle at rest and does not alter the rest mass or the rest energy of the
particle13 . Consequently, the little group will be SO (3) which is the group of all continuous rotations in three
dimensions.
For the case (c) with p2 = 0, p0 > 0, a good choice is k = (0, 0, , ) that describes a particle of null rest mass,
which travels to the speed of light along the axis x3 , the vector k is invariant under rotations around the axis
x3 (two dimensional rotations). We shall see later that its little group is ISO (2) which is the group of euclidean
geometry consisting of rotations and translations in two dimensions.
The case (f) is trivial, the condition p = 0 indicates the vacuum (no particles at all) which is left invariant by
any U (), hence the little group is the whole proper orthochronus Lorentz group in 3 + 1 dimensions (three space
+ 1 time coordinates). We symbolize it as SO (3, 1). Table 1.1, shows the six types of classes with the apropriate
standard vectors and little groups.
The little groups SO (2, 1) and SO (3, 1) for p2 > 0 and p = 0 have no non-trivial finite dimensional unitary
representations, so if we had states with a given momentum p with p2 > 0 or p = 0 that transform non-trivially
under the little group, there would have to be an infinite number of them. However, in what follows we only
consider classes of the type (a) and (c) (with p2 0 and p 6= 0), which are the only ones that posseses a
non-trivial Physical interpretation.
12
The number of classes is infinite. For instance, we have as much classes of the type Spb2 , as the number of different positive values
2
of p . The number of classes is then continuous.
13
The invariance of the value of p2 and the sign of p0 are valid for arbitrary proper orthochronus Lorentz transformations. Therefore,
they are valid for the little group in particular.
36 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

1.9.4 Normalization of one-particle states


Now we should take over the problem of the normalization of the states. A convenient starting point is to choose
the states with standard momentum k to be orthonormal in the extended sense, that is14


k , k, i = 3 k k (1.151)

from condition (1.151), the representation of the little group in Eqs. (1.148, 1.150) must be unitary

D (W ) = D 1 (W )

now we wonder about the scalar products between states of arbitrary momenta. We can calculate that inner
product by using the unitarity of U () in Eqs. (1.145, 1.150)



p , |p, i = N (p) p , U (L (p)) |k, i = N (p) hk, | U (L (p)) p , = N (p) hk, | U 1 (L (p)) p ,
   
= N (p) hk, | U L1 (p) p ,
( " #)
N (p ) X  1
= N (p) hk, | 1
D W L (p) , p L (p) p ,

N (L1 (p) p )

N (p ) X 
= N (p) 1
D W L1 (p) , p hk, | k ,
N (L (p) p )


N (p) N (p ) X 

p , |p, i = D W L1 (p) , p k , k, i ; k L1 (p) p
N (k )

and then using the orthonormalization condition (1.151) we find



N (p) N (p ) X  
p , |p, i =
D W L1 (p) , p 3 k k
N (k )

N (p) N (p )  
p , |p, i =
D W L1 (p) , p 3 k k ; k L1 (p) p (1.152)
N (k )

finally by the definition (1.145), we have N (k ) = 1. Further, since also k = L1 (p) p, then k = k if and only
if p = p . Therefore, the delta function 3 k k is proportional to 3 (p p ). As a consequence, theinner
product (1.152) is non-null only for p = p . Now, for p = p , the little group transformation W L1 (p) , p here
is trivial. To see it we use definition (1.149) to write

W (, p) L1 (p) L (p)
  
W L1 (p) , p = L1 L1 (p) p L1 (p) L (p) = L1 L1 (p) p = L1 (k)

but definition (1.144) with p = k, shows that L (k) must be the identity and so L1 (k), therefore

W L1 (p) , p = 1

from these facts, the scalar product in (1.152) becomes




p , |p, i = |N (p)|2 3 k k ; k L1 (p) p (1.153)

nevertheless, it is desirable to write the inner product in terms of (p p ). Thus, we should find the constant of
proportionality that relates 3 (k k) with 3 (p p). Note that the Lorentz-invariant integral of an arbitrary
14
We are writing hk , | k, i as proportional to (k k ) instead of (k k ). It is because given the three spatial components,
the constraint k2 = M 2 , along with the condition k0 > 0 (or k0 < 0), provides us the zeroth temporal component. Hence k = k is
equivalent to k = k .
1.9. ONE-PARTICLE STATES 37

scalar (Lorentz-invariant) function f (p) over four-momenta with p2 = M 2 0 and p0 > 0, which corresponds
to cases (a) and (c) in table 1.1, can be written as
Z
 
I = d4 p p2 + M 2 p0 f (p)
Z  
2   
I = d3 p dp0 p0 p2 + M 2 p0 f p, p0 (1.154)

where the delta function ensures the satisfaction of the condition p2 = M 2 , and the step function guarantees
that only the terms with p0 > 0 contribute in the integral. We can rewrite the integral by using the property
 1
x2 e2 = [ (x + e) + (x e)] (1.155)
2 |e|

hence performing the integral over p0 in (1.154) by using (1.155), we obtain


h  p   p i
Z p0 p2 + M 2 + p0 + p2 + M 2  
I = d3 p dp0 p p0 f p, p0
2 p2 + M 2
  p 
since p0 forbids negative values of p0 , the term p0 + p2 + M 2 does not contribute, hence
 p 
Z p0 p2 + M 2  
I = d3 p dp0 p p0 f p, p0
2 p2 + M 2
 p 
Z f p, p2 + M 2
I = d3 p p (1.156)
2 p2 + M 2

since we have taken the positive square root for p0 , the step function is not necessary anymore. When integrating
f p, p0 on the mass shell p2 + M 2 = 0, the integral (1.156) shows that the invariant volume element is

d3 p
p (1.157)
p2 + M 2

the delta function is defined as


Z Z
   hp i d3 p
F (p) = F p 3 p p d3 p = F p p2 + M 2 3 p p p (1.158)
p2 + M 2

hence combining Eqs. (1.158, 1.157), we see that the invariant delta function is
p  
p2 + M 2 3 p p = p0 3 p p (1.159)

thus the form of the new delta function defined in (1.159), must be preserved under a Lorentz transformation.
As a consequence, since p and p are related with k and k, by the same Lorentz transformation L (p), we have
 
p0 3 p p = k0 3 k k (1.160)

which is the relation between 3 (p p) and 3 (k k) that we were looking for. Substituting (1.160) in (1.153),
the inner product becomes  0

2 p 
p , |p, i = |N (p)| 0
3 p p
k
38 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

sometimes the normalization is chosen such that N (p) = 1. However, in that case we should keep track of the
factor p0 /k0 in scalar products. The most usual convention (which is the one we adopt here) is that
s
k0
N (p) = (1.161)
p0
with this choice, the final form of the inner product yields


p , |p, i = 3 p p (1.162)
we shall study now the cases (a) and (c) of table 1.1, that is the case of a particle with non-null mass M > 0, and
the case of particles with zero mass p2 = 0.

1.10 One-particle states with non-null mass


When p2 < 0, table 1.1 shows that the little group is SO (3). For this group, all irreducible representations are
finite-dimensional and can be settled as unitary15 . Further, a unitary representation can be broken up into a
direct sum of irreducible unitary representations. The irreducible representations characterized by the matrices
(j)
D (R) have dimensionality 2j + 1, with j = 0, 1/2, 1, 3/2, 2, . . .. Infinitesimal rotations can be written in the
form (1.88)
Rik = ik + ik ; ik = ki
the antisymmetry comes from Eq. (1.89), and the fact that SO (3) is a subgroup of SO (3, 1). Considered as a
subgroup of SO (3, 1), we can write the infinitesimal rotation as in Eq. (1.99), but taking into account that for
pure rotations, the only generators that appear in that expansion are the angular momentum operators defined
in Eq. (1.127)
1 1 1 1
U (1 + , 0) = 1 + iik J ik = 1 + i23 J 23 + i31 J 31 + i12 J 12
2 2 2 2
1 1 1
U (1 + , 0) = 1 + i1 J1 + i2 J2 + i3 J3 (1.163)
 232 31 12 2 2
J J ,J ,J {J1 , J2 , J3 } ; {23 , 31 , 12 } {1 , 2 , 3 }
hence, the matrix representative associated with the (j)representation reads
i   i  
(j) (j)
D , (1 + ) = , + ik Jik = , + p Jp(j) (1.164)
2 2

the canonical basis of 2j + 1 orthonormal vectors for the (j) representation is denoted by
{|j, i ; = j, j 1, j 2, . . . , j}
and the matrix representations of the generators in the canonical basis are well-known from the theory of angular
momenta
  1 hp p i
(j)
J1 = j (j + 1) ( + 1)
,+1 + j (j + 1) ( 1)
,1 (1.165)
2
  1 hp p i
(j)
J2 = j (j + 1) ( + 1) ,+1 j (j + 1) ( 1) ,1 (1.166)
2i

     p
(j) (j) (j) (j) (j)
J J23 iJ31 = J1 iJ2 = ,1 (j ) (j + 1) (1.167)

   
(j) (j)
J12 = J3
= , (1.168)
,
= j, j 1, j 2, . . . , j
15
It means that, if a given irreducible representation of SO (3) is non-unitary, it could become unitary through a similarity transfor-
mation.
1.10. ONE-PARTICLE STATES WITH NON-NULL MASS 39

For future purposes we shall construct a rotation R (b


p) that takes the three-axis into the direction of the three-
vector p. The determination of a direction requires two angles, in spherical coordinates we use the azimuthal and
polar angles (both are measured with respect to the three-axis). The direction of a unitary vector determined
by these angles is specified as
p
b
p = sin cos u1 + sin sin u2 + cos u3 (1.169)
|p|

it is clear on geometrical grounds that the rotation R (b p) can be carried out as follows: (a) a rotation R2 () around
X2 by the angle [it takes (0, 0, 1) into (sin , 0, cos )], and then (b) a rotation R3 () around X3 by an angle
[it takes the vector (sin , 0, cos ) into the direction defined by (1.169)].
Let us do this process explicitly (for simplicity, we shall only use the three-space coordinates), the first rotation
R2 () clearly gives
cos 0 sin 0 sin
R2 () u3 = 0 1 0 0 = 0
sin 0 cos 1 cos
the second rotation R3 () is described by

sin cos sin 0 sin cos sin
R3 () 0 = sin cos 0 0 = sin sin = p b
cos 0 0 1 cos cos

the complete process is then


R (b b
p) u3 = R3 () R2 () u3 = p (1.170)
the complete matrix of rotation yields

cos sin 0 cos 0 sin
p) = R3 () R2 () = sin cos
R (b 0 0 1 0
0 0 1 sin 0 cos

cos cos sin cos sin
p) = cos sin cos sin sin
R (b (1.171)
sin 0 cos

it is easy to check that its inverse corresponds to the transpose as it must be for three-dimensional rotations

cos cos cos sin sin
R1 (bp) = Re (b
p) = sin cos 0 (1.172)
cos sin sin sin cos

it is convenient to write R (b b . According with


p) in terms of the cartesian components pbi of the unitary vector p
Eq. (1.169) such components yield

pb1 = sin cos , pb2 = sin sin , pb3 = cos (1.173)

from Eq. (1.173) the matrix in (1.171) becomes


pb3 pb1 pb2
sin sin pb1
p) =
R (b pb3 pb2
sin
pb1
sin pb2 (1.174)
sin 0 pb3

using (1.173) again, we have


pb21 + pb22 = sin2 = pb2 pb23 = 1 pb23
40 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS


since 0 then sin 0, so that sin = sin2 . Therefore
q
sin = 1 pb23 (1.175)
substituting (1.175) in (1.174) we have

pb1 pb3 2 pb2 pb1
1b
p3 1bp23
p) =
R (b pb2 pb3 2 pb1
pb2
(1.176)
p1bp3 1b p23
1 pb23 0 pb3
Finally, taking into account that the transformation is embedded in the four-dimensional Minkowski space
Eqs. (1.171, 1.172, 1.176) must be written as

pb1 pb3 2 pb2 2 pb1 0
cos cos sin cos sin 0 1bp3 1bp3
cos sin cos sin sin 0 pb2 pb3 pb1
pb2 0
R (b
p) = = 2 2 (1.177)
sin 0 cos 0 p1bp3 1b p3
1 pb 2 0 pb3 0
0 0 0 1 3
0 0 0 1
pb pb p
1 3 2 pb2 pb3 2 1 pb23 0
cos cos cos sin sin 0 1b p3 1b p3
sin cos 0 0 pb2 pb1
0 0
1
R (b p) = = 1b p32 1b 2
p3 (1.178)
cos sin sin sin cos 0

pb1 pb2 pb3 0
0 0 0 1
0 0 0 1
The rotation R (bp) is carried out at the Minkowski four-dimensional space, though it acts non-trivially only
within the three-space coordinates. According with Eq. (1.170), the corresponding representation on the Hilbert
space becomes
U (R (b
p)) = U (R3 () R2 ()) = U (R3 ()) U (R2 ())
which in terms of the generators is written as
p)) = exp (iJ3 ) exp (iJ2 ) ; 0 , 0 < 2
U (R (b (1.179)
there is an important difference between the rotation itself and their representations: in representations associated
with integer values of (j) we obtain the same element by shifting or by 2; by contrast, in representations
associated with half-integer values of (j), we obtain the element with opposite sign by shifting or by 2.
Thus the only representations that can describe geometrical objects are integer representations, while half-integer
representations can only describe intrinsic variables16 . Representations in quantum mechanics such as Eq. (1.179),
could be associated either to integer or half-integer representations.
Since the rotation R (b
p) described by (1.177) has the role of taking the three-axis into the direction (1.169), it
is clear that we could have added an initial rotation around the three-axis without affecting that result. Thus, any
other choice on R (b p) in the Minkowski space would differ from this one at most in an initial rotation around the
three-axis. For the quantum operator U (R (b p)) in Eq. (1.179), such a difference leads just to a mere redefinition
of the phase of the one-particle states.

1.10.1 Wigner rotation and standard boost


For a particle of mass M > 0, and spin j, Eq. (1.150) combined with normalization (1.161), becomes17
s
(p)0 X (j)
p,

U () |p, i = D
(W (, p)) (1.180)
p0
16
It is because of these reasons that the spherical harmonics Yl,m (, ) only admit integer values of l and m.
17
By now, we say spin referring to the label (j) associated with a given irreducible representation of SO (3).
1.10. ONE-PARTICLE STATES WITH NON-NULL MASS 41

now we shall calculate the little-group element (Wigner rotation) W (, p) defined by Eq. (1.149)

W (, p) = L1 (p) L (p) (1.181)

to calculate this Wigner rotation, we should choose  a standard boost L (p) that carries the standard four
momentum k = (0, 0, 0, M ) to p = p1 , p2 , p3 , p0 [as required by Eq. (1.144)]. Indeed, it is easier to construct
the inverse transformation
k = L1 (p) p
in Eq. (1.177) we have characterized the three-dimensional rotation R (b p) that takes the three-axis into the
direction of the unitary vector p 1
b . Thus, R (bp) applied on p preserves the norm of the three-vector and put it
along with the three-axis, while the time-component remains invariant

R1 (b
p) p = 0, 0, |p| , p0

next we construct a boost B 1 (|p|) that eliminates the third component of the new four-vector. It is clear that the
first and second components do not require modification. Since the pseudonorm must be preserved and the final
vector has no space components, the time-component must be the rest mass M of the particle (otherwise it would
not be a Lorentz transformation). It is clearly carried out by a pure boost B 1 (|p|) along the three-component
to a new reference frame in which the particle is at rest, so we find
0

0 1 0 0 0 0 0
0 0 1 0 0 0
0
B 1 (|p|) = p 0 |p| 0 = p0

|p| 0 =
|p| 0 0 M |p| M |p| M p 0
M 2
p0 p0 p0 (p0 )
0 0 |p| M
2
M M p +M M

then we have already arrived to k by using the sequence

k = B 1 (|p|) R1 (b
p) p (1.182)

Moreover, since k contains a null three-vector, it is clear that we can add an arbitrary three-dimensional rotation,
and we shall not alter k . For reasons to be understood later, we shall add the rotation R (b p) to the previous
sequence of transformations. Then we find

k = L1 (p) p R (b
p) B 1 (|p|) R1 (b
p) p (1.183)

on the other hand, we write B 1 as a function of |p| because p0 is not independent of |p|. Such a fact is more
apparent if we define
p r
p2 + M 2 p0 p p2 + M 2 M 2 |p|
= 2 1 = 2
2 = (1.184)
M M M M M
so that B 1 (|p|) can be written as a function of the single parameter

1 0 0 0
0 1 0
B 1 (|p|) = p0
0 0 2 1
p
0 0 2 1

and B (|p|) is easily obtain



1 0 0 0
0 1 0
B (|p|) = p 0 (1.185)
0 0 p 1
2

0 0 2 1
42 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

from Eq. (1.183) we finally obtain the standard boost L (p)

p) B (|p|) R1 (b
L (p) = R (b p) (1.186)

whose explicit form is obtained by picking up Eqs. (1.177, 1.178, 1.185)



pb1 pb3 2 pb2 pb1 0
1b p3 1bp23
pb2 pb3 pb2 0 pb1
L (p) =
p1bp3
2 1b p23

1 pb2 0 pb3 0
3
0 0 0 1
p
pb1 pb3 2 pb2 pb3 2 1 pb23 0
1 0 0 0 1b p3 1b
p3

0 1 0 pb2 pb1 0
p 0 1bp23 p23
1b
0
0 0 2 1
p pb1 pb2 pb3 0
0 0 2 1 0 0 0 1

p
pb1 pb3 2 pb2 pb1 0 pb1 pb3 2 pb2 pb3 2 1 pb23 0
1b p3 1b p23 1bp3 1b p3

pb2 pb3 p p
pb2 0
pb1 b b
2 2
0 0 21
L (p) =
p1bp3
2 1b p23
1b p3 p 1b p3

1 pb2 pb3 0 p pb1 2 1
3 0 p pb2 p pb3
0 0 0 1 pb1 2 1 pb2 2 1 pb3 2 1

p
(pb22 +bp21 pb23 + pb21 pb21 pb23 )
p23
1b
pb1 pb2 ( 1) pb1 pb3 ( 1) pb1 2
1
(pb21 +bp22 pb23 + pb22 pb22 pb23 ) p
pb1 pb2 ( 1) pb2 pb3 ( 1) pb2 2 1
L (p) = p23
1b p (1.187)

3 ( 1)
pb1 pbp 3 ( 1)
pb2 pbp 1 +p ( 1) pb23 pb3 2 1
pb1 2 1 pb2 2 1 pb3 2 1

the first two diagonal elements L1 1 (p) and L2 2 (p) can be simplified as follows

1   1  2 
L1 1 (p) = 1 pb23 pb22 + pb21 pb23 + pb21 pb21 pb23 = 1 pb23 pb2 + pb21 pb23 + pb21 1 pb23
1  2  
= 1 pb23 pb2 + pb21 pb21 + pb21 pb23 + pb21 1 pb23 (1.188)
1      
= 1 pb23 1 pb23 pb21 1 pb23 + pb21 1 pb23 = 1 pb21 + pb21
1
L 1 (p) = 1 + pb21 ( 1) (1.189)

similarly
L2 2 (p) = 1 + pb22 ( 1) (1.190)

substituting (1.189) and (1.190) in (1.187) the standard boost L (p) becomes
p
1 + ( 1) pb21 ( 1) pb1 pb2 ( 1) pb1 pb3 pb1 p 2 1

( 1) pb1 pb2 1 + ( 1) pb22 ( 1) pb2 pb3 pb2 p 2 1
L (p) = (1.191)
( p 1) pb1 pb3 ( p 1) pb2 pb3 1 + p( 1) pb23 pb3 2 1
pb1 2 1 pb2 2 1 pb3 2 1
1.10. ONE-PARTICLE STATES WITH NON-NULL MASS 43

as a matter of consistency, it can be checked explicitly that L (p) applied on k = (0, 0, 0, M ) yields p =
p1 , p2 , p3 , p0 .

(1)p21 (1)p1 p2 (1)p1 p3 p1 2 1 M p1 2 1
1 + p2 2 2
p p |p| 0 |p|

(1)p1 p2 (1)p22 (1)p2 p3 p2 2 1 M p2 2 1
1 + 0
L (p) k = p 2 p2 p 2
|p| =
|p|



(1)p1 p3 (1)p p (1)p23 p3 2 1 0 2 1
2 3
1 + M p 3

p2 p2 p2 |p| M |p|
p1 2 1 p2 2 1 p3 2 1
M
|p| |p| |p|
1
p1 p
p2 p2
L (p) k =
p3 = p3 = p

p0 p0

where we have used Eqs. (1.184).


In summary,  the standard boost L (p) that carries the standard four momentum k = (0, 0, 0, M ) to p =
p1 , p2 , p3 , p0 is given by Eq. (1.191) which can be expressed in the form

Li k (p) = ik + ( 1) pbi pbk


p
Li 0 (p) = L0 i (p) = pbi 2 1 ; L0 0 =
p
p2 + M 2 p0
pbi pi / |p| , = (1.192)
M M
It is important that when is an arbitrary three-dimensional rotation R, the Wigner rotation W (R, p) is the
same as R for all p. In other words, if belongs to the little group, then the element W (, p) of the little group
induced by must coincide with itself18 . We see it by noticing that the boost (1.192) can be expressed as in
Eq. (1.186)
p) B (|p|) R1 (b
L (p) = R (b p) (1.193)
where R (b
p) is the rotation (1.177) that takes the three-axis into the direction of p, and

1 0 0 0
0 1 0
B (|p|) = p 0 (1.194)
0 0 21
p
0 0 2 1

is a pure boost along the three-axis. Then for an arbitrary rotation R we have from (1.181, 1.193) that
 1  
W (R, p) = L1 (Rp) R L (p) = R (Rb p) B (|Rp|) R1 (Rbp) R R (bp) B (|p|) R1 (b
p)
p) B 1 (|p|) R1 (Rb
W (R, p) = R (Rb p) B (|p|) R1 (b
p) R R (b p) (1.195)

but the rotation R1 (Rbp) R R (b b , and then


p) takes the three-axis (i.e. the unit vector u3 ) into the direction p
into the direction Rb
p, and then back to the three-axis. Therefore, the whole operation can only change the unit
vectors u1 and u2 . Hence the whole operation is a rotation by some angle around the three-axis (or around the
unit vector u3 )
cos sin 0 0
sin cos 0 0
R1 (Rbp) R R (bp) = R ()
(1.196)
0 0 1 0
0 0 0 1
18
In order to obtain such a property we require the structure (1.193), which explains the convenience of the (a priori unnecessary)
introduction of R (b
p) in the step from Eq. (1.182) to Eq. (1.183).
44 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

since matrices (1.194, 1.196) act non-trivially on different subspaces of the Minkowski space, they commute with
each other
[R () , B (|p|)] = 0 (1.197)
from Eqs. (1.196, 1.197), the Wigner rotation (1.195) becomes
 
W (R, p) = R (Rb p) B 1 (|p|) R1 (Rbp) R R (bp) B (|p|) R1 (b p) B 1 (|p|) R () B (|p|) R1 (b
p) = R (Rb p)
p) B 1 (|p|) B (|p|) R () R1 (b
= R (Rb p) R () R1 (b
p) = R (Rb p)

and using (1.196) again, we finally obtain


 
p) R () R1 (b
W (R, p) = R (Rb p) = R (Rb
p) R1 (Rb p) R1 (b
p) R R (b p)
W (R, p) = R

then states of a moving massive particle (and by extension multi-particle states), transform in the same way
under rotations as in non-relativistic quantum mechanics. Consequently, all tools developed for rotations in non-
relativistic quantum mechanics, such as the spherical harmonics, Clebsch-Gordan coefficients etc. can be utilized
in relativistic quantum mechanics.

1.11 One-particle states with null mass


1.11.1 Determination of the little group
We will first characterize the little group associated to case (c) in table 1.1. We shall choose the standard four-
vector as
k = (0, 0, 1, 1) (1.198)
and we should find the set of Lorentz transformations W , that satisfy

W k = k

since W is unitary and W k = k, we have the following properties for an arbitrary four vector t

hW t |W ti = ht |ti ; hW t |ki = hW t |W ki = ht |ki (1.199)

applying (1.199) to a time-like four-vector of the form t (0, 0, 0, 1), we obtain

(W t) (W t) = t t = 1 (1.200)

(W t) k = t k = 1 (1.201)

by writing (W t) = (, , , ) we can rewrite Eq. (1.201) as

(, , , ) (0, 0, 1, 1) = 1 = 1 = 1 +

consequently, any four-vector that satisfies the condition (1.201) can be parameterized as

(W t) = (, , , 1 + ) (1.202)

in addition, the first condition (1.200) combined with (1.202) yields

(W t) (W t) = 2 + 2 + 2 (1 + )2 = 1
2 + 2
= (1.203)
2
1.11. ONE-PARTICLE STATES WITH NULL MASS 45

therefore, , are the independent parameters. However, we shall use the parameter to simplify some expressions.
The effect of W on t is the same as that of the following Lorentz transformation


1 0
0 1
S (, )

(1.204)
1
1 +

it can be checked out explicitly


1 0 0
0 1
0
S t =

=
= W t
1 0
1 + 1 1+

where we have used equation (1.202). It does not mean that W = S (, ), but it means that

 
S 1 (, ) W t = S 1 (, ) [W t] = S 1 (, ) [S (, ) t] = t

hence, S 1 (, ) W is a Lorentz transformation that leaves the time-like four-vector (0, 0, 0, 1) invariant. There-
fore, S 1 (, ) W is a pure rotation19 . It could also be seen that


1 0 0 0
0 1 0 0
S k =

= = k
1 1 1
1 + 1 1

hence S like W , leaves the reference light-like vector k invariant20 . Since S and W belong to the little group
then S 1 (, ) W also does. Further since S 1 (, ) W leaves t = (0, 0, 0, 1) and k = (0, 0, 1, 1) invariant, we
conclude that S 1 (, ) W leaves the three-axis invariant. Thus, such an operator must be a rotation around
the three-axis

cos sin 0 0
sin cos 0 0
S 1 (, ) W = R () ; R ()
(1.205)
0 0 1 0
0 0 0 1

from Eq. (1.205) we obtain

W (, , ) = S (, ) R () (1.206)

19
Since a Lorentz transformation leaves the pseudonorm invariant, and this transformation leaves the zeroth component invariant,
the three-vector norm remains invariant.
20
In that sense, S is something like a standard element of the little group.
46 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

since W is arbitrary, we conclude that the most general element W of the little group can be parameterized as in
Eq. (1.206). To characterize this little group, we first establish the multiplication rule for S (, )

1 0 1 0
 0 1 0 1
S , S (, ) =





1 1
1 + 1 +

1 0 +
0 1 +
=
+ + 1

+ + +
+ + 1 + + + +

1 0 ( + ) +
 0 1 + +
S , S (, ) =
+ + 1 + + +

(1.207)
 + + + 
+ + + + + 1 + + + +

and applying Eq. (1.203), we see that

2 + 2 2 + 2
+ + + = + + +
2 2
2
2 2 2 2
+ + 2 + + + 2 ( + )2 + +
= =
2 2
+ + + = + , + (1.208)

substituting (1.208) in (1.207) we find



1 0 ( + ) +
 0 1 + +
S , S (, )
+ + 1 + , +
 
 + , + 
+ + + , + 1 + + , +
 
S , S (, ) = S + , + (1.209)

the rotations R () are obviously abelian since they are around the three-axis only
 
R R () = R + (1.210)

and it is easy to check explicitly that

S 1 (, ) = S (, ) ; R1 () = R () (1.211)

by reasons to be understood later, we call the little group defined in (1.206) as ISO (2). From Eqs. (1.209,
1.210, 1.211) we see that the sets {S (, )} and {R ()} form subgroups of the little group ISO (2). Besides, the
subgroup formed by {S (, )} [or equivalently by {W (, , = 0)}], is invariant in ISO (2). It means that for
any given element g ISO (2) and any given element h S (, ) we have

ghg1 S (, ) ; g ISO (2) ; h S (, ) (1.212)


1.11. ONE-PARTICLE STATES WITH NULL MASS 47

to see it, we observe first that



cos sin 0 0 1 0
sin cos 0 0 0 1
R () S (, ) R1 () =


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

cos sin 0 0
sin cos 0 0
0

0 1 0
0 0 0 1

1 0 cos sin cos + sin
0 1 sin cos sin + cos
R () S (, ) R1 () =
cos + sin sin + cos


1 (, ) (, )
cos + sin sin + cos (, ) (, ) + 1

by defining
e cos + sin ; e sin + cos
(1.213)

we obtain
1 0 e e
0 1 e e

R () S (, ) R1 () = (1.214)

e e 1 (, ) (, )

e e
(, ) (, ) + 1
and
  e2 + e2
( cos + sin )2 + ( sin + cos )2
e, e = =
2  2 
2 cos2 + sin2 + 2 cos2 + sin2
=
2
  2 + 2
e, e = = (, ) (1.215)
2

substituting (1.213, 1.215) in (1.214) we find



1 0 e
e
0 1 e
  e
  
e

R () S (, ) R1 () =

e e 1 e, e e, e

=S
e ,
   

e e e, e 1+ e, e
R () S (, ) R1 () = S ( cos + sin , sin + cos ) (1.216)

using Eqs. (1.206, 1.216) we have


       1
W , , S (, ) W 1 , , = S , R () S (, ) S , R ()
   
= S , R () S (, ) R1 () S 1 ,
       
W , , S (, ) W 1 , , = S , S e, e S 1 , = S + e , + e
   
W , , S (, ) W 1 , , = S e, e ; e cos + sin , e sin + cos (1.217)
48 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

since the RHS of Eq. (1.217) clearly belongs to the subgroup {S (, )}, we observe by comparing with Eq.
(1.212), that {S (, )} is invariant in the little group ISO (2). The multiplication rule for the whole little group
ISO (2) is obtained by combining Eqs. (1.206, 1.209, 1.210, 1.216)
        
W , , W (, , ) = S , R [S (, ) R ()] = S , R S (, ) R1 R R ()
  
= S , S cos + sin , sin + cos R +
 
= S + cos + sin , sin + cos + R +

= W + , + cos + sin , sin + cos +

the composition law is then given by


  
b, b ;
W , , W (, , ) = W + , b + cos + sin , b = sin + cos + (1.218)

the multiplication rule (1.218) is the one associated with the group of euclidean geometry consisting of translations
in two dimensions (by a vector |, i) and rotations in two dimensions by an angle . However, it is important to
clarify that we call this little group ISO (2), because it is isomorphic with the euclidean group of rotations and
translations in two dimensions, but our little group is a proper subgroup of the homogeneous Lorentz group, hence
translations in the Minkowski space (or in the Hilbert space of quantum mechanics) are not contained in this little
group but only rotations and boosts. In other words, the transformations contained in our little group (understood
as transformations in the Minkowski or Hilbert space) are not rotations and translations in two dimensions, they
are just isomorphic with the group of bidimensional translations and rotations in the euclidean space.
When a group does not have invariant abelian groups it is called semi-simple, showing that it has certain
simple properties. We have just seen that the little group ISO (2) is not semi-simple [according with Eq. (1.217),
it contains the translations in two dimensions S (, ) as an abelian invariant subgroup]. In the same way the
inhomogeneous Lorentz group is not semi-simple either.

1.11.2 Lie algebra of the little group ISO (2)


To obtain more information about the way in which ISO (2) is embedded in SO (3, 1), we shall study the Lie
algebra of ISO (2), and its relation with the Lie algebra of SO (3, 1). If , , become infinitesimal parameters we
obtain an infinitesimal transformation of the little group. Taking Eqs. (1.204, 1.205) with infinitesimal parameters
up to first order in , and/or , we have

0
1 1 0 0 1

1 1 0
W (, , ) = S (, ) R () =
0 0 = 1
1 0 0 0
1 0 + 1 0
0 1 0 0 0 1 + 0 1

1 0 0 0 0
0 1 0 0 0 
W (, , ) =
0 0 1 0 + 0 0 +O
2 , 2 , (1.219)
0 0 0 1 0 0

Parameterizing the transformation (1.219) as in Eq. (1.88) page 25, we find



0
0
W (, , ) = + ; =
0 0

0 0
1.11. ONE-PARTICLE STATES WITH NULL MASS 49

and the totally covariant transformation becomes



1 0 0 0 0
0 1
0 0 0
= g = 0 0

1 0 0 0
0 0 0 1 0 0

0
0
=
(1.220)
0 0
0 0

observe that is antisymmetric, in consistence with Eq. (1.89), page 25, and the fact that W (, , ) is a
particular case of a Lorentz transformation (notice however that is not antisymmetric).
Substituting Eq. (1.220) in Eq. (1.99) page 27, we obtain the generators J (which have also be taken as
antisymmetric) that contributes in the expansion of the associated operator U (W (, , )) in the Hilbert space
(for , , infinitesimal). Equation (1.220) shows that 03 = 30 = 0. Hence, the associated generators J 03 and
J 30 do not contribute in such an expansion21
1 
U (W (, , )) = 1 + i J + O 2
2
1  1  1 
= 1 + i 12 J 12 + 21 J 21 + i 13 J 13 + 31 J 31 + i 10 J 10 + 01 J 01
2 2 2
1  1 
+ i 23 J 23 + 32 J 32 + i 20 J 20 + 02 J 02
2 2
since both and J are antisymmetric, the products J are symmetric. Using this fact and the explicit
form of the elements given by Eq. (1.220) we find

U (W (, , )) = 1 + i12 J 12 + i13 J 13 + i10 J 10 + i23 J 23 + i20 J 20


= 1 + iJ 12 iJ 13 + iJ 10 iJ 23 + iJ 20
 
U (W (, , )) = 1 + i J 10 J 13 + i J 20 J 23 + iJ 12
U (W (, , )) = 1 + iA + iB + iJ 12 ; A J 10 J 13 , B J 20 J 23

using Eqs. (1.127, 1.129) page 30, we obtain

U (W (, , )) = 1 + iA + iB + iJ3 (1.221)
31 01 02 23
A J J = J2 K1 ; B J J = K2 J1 (1.222)

the Lie algebra of ISO (2), i.e. between the generators A, B and J3 can be obtained from the commutation
relations (1.130, 1.131, 1.132) page 31. For instance, using Eqs. (1.130, 1.131) we find

[J3 , A] = [J3 , J2 K1 ] = [J3 , J2 ] [J3 , K1 ] = iJ1 iK2 = iB

proceeding in the same way with the other generators, the Lie algebra of ISO (2) (written as a subgroup of the
Lorentz group) becomes
[J3 , A] = iB , [J3 , B] = iA , [A, B] = 0 (1.223)
these commutation relations can also be obtained from Eqs. (1.209, 1.210, 1.216) [homework!!(7)]. In the frame-
work of the euclidean space in two dimensions, J3 is the generator of rotations and A, B are generators of
translations (owing to it, A and B commute each other). However, definitions (1.222) show that in the framework
21
Of course, the generators J are null even in the general case because of the antisymmetry.
50 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

of the Minkowski space (or in the Hilbert space in quantum mechanics), A and B are generators that combine
rotations with boosts.
For future purposes we shall find the transformation properties of the generators A and B under rotations
U [R ()] within the little group. To do it, we shall use an infinitesimal transformation U [S (, )], such that
and are infinitesimals, but is arbitrary. From Eq. (1.216) we find

 
U [R ()] U [S (, )] U 1 [R ()] = U R () S (, ) R1 ()
= U [S ( cos + sin , sin + cos )]
h  i
U [R ()] U [S (, )] U 1 [R ()] = U S e, e e cos + sin , e sin + cos (1.224)
;

now using the expansion (1.221) we can write

U [R ()] U [S (, )] U 1 [R ()] = U [R ()] {1 + iA + iB} U 1 [R ()]


U [R ()] U [S (, )] U 1 [R ()] = 1 + iU [R ()] A U 1 [R ()] + iU [R ()] B U 1 [R ()] (1.225)

using again the expansion (1.221) on the RHS of Eq. (1.224) we have
h  i
U [R ()] U [S (, )] U 1 [R ()] = U S e, e = 1 + ie e
A + iB
= 1 + i ( cos + sin ) A + i ( sin + cos ) B
1
U [R ()] U [S (, )] U [R ()] = 1 + i (A cos B sin ) + i (A sin + B cos ) (1.226)

equating Eqs. (1.225, 1.226) yields

U [R ()] A U 1 [R ()] + U [R ()] B U 1 [R ()] = (A cos B sin ) + (A sin + B cos )

since and are infinitesimal but otherwise arbitrary, we can equate the coefficients of them to obtain

U [R ()] A U 1 [R ()] = A cos B sin (1.227)


1
U [R ()] B U [R ()] = A sin + B cos (1.228)

1.11.3 Massless states in terms of eigenvalues of the generators of ISO (2)


Since A and B are hermitian commuting operators, they admit a common basis of eigenvectors

A |k, a, b, i = a |k, a, b, i , B |k, a, b, i = b |k, a, b, i (1.229)

where denotes any additional quantum numbers required to define physical states. Since {|k, a, b, i} forms a
basis, the physical states are superpositions of such vectors
X
|k, i = cab |k, a, b, i (1.230)
a,b,

However, we shall see that if we find a set of non-zero eigenvalues of A, B they would belong to a continuous
spectrum. From Eqs. (1.229) and utilizing (1.227) we obtain

AU 1 [R ()] |k, a, b, i = U 1 [R ()] U [R ()] AU 1 [R ()] |k, a, b, i
= U 1 [R ()] {A cos B sin } |k, a, b, i
= U 1 [R ()] {cos A |k, a, b, i sin B |k, a, b, i}
= U 1 [R ()] {cos a |k, a, b, i sin b |k, a, b, i}
 1 
A U [R ()] |k, a, b, i = {a cos b sin } U 1 [R ()] |k, a, b, i
1.11. ONE-PARTICLE STATES WITH NULL MASS 51

a similar procedure can be done from Eqs. (1.229) and (1.228), so that we obtain
A |k, a, b, i = (a cos b sin ) |k, a, b, i (1.231)
1
B |k, a, b, i = (a sin + b cos ) |k, a, b, i ; |k, a, b, i U [R ()] |k, a, b, i (1.232)
note that is a continuous parameter, hence Eqs. (1.231, 1.232) show that if a and/or b are different from zero,
the spectra of A and B are continuous. Such a continuous degree of freedom like is not observed in massless
particles22 . Consequently, to avoid this additional continuous degree of freedom, we should demand that the only
possible eigenvectors involved in the superposition (1.230) that forms physical states |k, i, are eigenvectors with
eigenvalues a = b = 0. X X
|k, i = c00 |k, 0, 0, i c |k, i (1.233)

applying operators A or B on the physical states (1.233) we find
A |k, i = B |k, i = 0 (1.234)
hence particle states cannot be characterized by eigenvalues of A or B. These states are then distinguished by
their eigenvalues associated with the remaining generator J3 of ISO (2)
J3 |k, i = |k, i (1.235)
since the momentum k [the three-momentum associated with the standard four-momentum (1.198)] is in the
three-direction u3 , we have that gives the component of angular momentum in the direction of motion, called
the helicity.

1.11.4 Lorentz transformations of massless states


From the previous properties, we are able to find the Lorentz transformations of general states of masless particles.
By using the general arguments described in Sec. 1.4 [see Eq. (1.39), Page 17], we obtain that Eqs. (1.221) can
be generalized to the case of finite values of the parameters , and , since the associated subgroups are abelian
U (S (, )) = exp (iA) exp (iB) = exp (iA + iB) (1.236)
U (R ()) = exp (iJ3 ) (1.237)
where we have taken into account that A and B commute. According with Eqs. (1.206, 1.234), an arbitrary
element W of the little group acts on a massless state |k, i as
U (W ) |k, i = U (S (, )) U (R ()) |k, i = exp (iA + iB) exp (iJ3 ) |k, i
= exp (i) exp (iA + iB) |k, i
U (W ) |k, i = exp (i) |k, i
hence the representation D (W ) of the little group in Eq. (1.148) yields
D (W ) = exp (i) (1.238)
from this we find the general Lorentz transformation rule (i.e. under U ()), for a general massless particle state
|p, i of arbitrary helicity, by using Eqs. (1.150, 1.161)
p
N (p) X k0 /p0 X
U () |p, i = D (W (, p)) p, = q

exp [i (, p)] p,
N (p)
(k)0 / (p)0
s
(p)0
U () |p, i = exp [i (, p)] |p, i (1.239)
p0
22
All particles observed so far, are such that the only continuous degrees of freedom are related with the orbital Hilbert space Er .
Thus the continuous degrees of freedom are determined by the commuting operators {R1 , R2 , R3 } or by the commuting observables
{K1 , K2 , K3 }. The remaining degrees of freedom (such as spin or similar variables) are always discrete.
52 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

from Eqs. (1.149) and (1.206) we see that (, p) is defined by

W (, p) L1 (p) L (p) S ( (, p) , (, p)) R ( (, p)) (1.240)

we shall see later that electromagnetic gauge invariance arise from the part of the little group parameterized by
and .
Note that we have not restricted the spectrum of . Indeed, by contrast to the case of massive particles,
the spectrum of for massless particles is not restricted by the algebra. To see it, we observe that for massive
particles it is the Lie algebra of angular momentum (with all its three generators) associated with the little group
SO (3), that predicts the fact that is only integer or half-integer. The litte group ISO (2) of massless particles
only contain one generator of angular momentum (J3 by convention) that is not enough to restrict the spectrum
of .
Despite is not restricted on algebraic grounds, there are topological considerations that restrict the allowed
values of to be integer of half-integer, like in the case of massive particles [see section 2.7 of Ref. [1]].
In order to calculate the little group element (1.240) for a given and p, we proceed similarly as in the case
of massive particles: we have to fix a convention for the standard Lorentz transformation that takes us from the
stardard four-momentum k (0, 0, , ) to p . We can choose such a stardard Lorentz transformation to have
the form
L (p) = R (bp) B (|p| /) (1.241)
where B (u) is a pure boost along the u3 direction

1 0 0 0
0 1 0 0

B (u) = (u2 +1) (u2 1) (1.242)
0 0 2u 2u
(u2 1) (u2 +1)
0 0 2u 2u

and R (b p) is a pure rotation [see Eq. (1.177), page 40] that carries the three-axis into the direction of the
momentum determined by the unitary vector p b . Note that in passing from k to p , it is enough to determine
the change from k to p because of the conservation of the pseudonorm. Hence, the boost in (1.241) accomplishes
the role of changing the norm of the three-vector from |k| to |p|, while the rotation changes the direction to its
final one. The arguments to build up the standard boost (1.241) for massless particle states, are similar to the
ones we followed to obtain the standard boost (1.193) for massive particles.
We can see from Eq. (1.239) that helicity is Lorentz-invariant; a massless particle of a given helicity looks the
same (aside from its momentum) in all inertial reference frames23 . Therefore, based on SO (3, 1) transformations,
we could a priori think of massless particles of different helicities as different species of particles (in the same way as
particles with different electric charge are considered different species). Nevertheless, this is not the case when the
discrete Lorentz transformations are included. We shall see in the next section that particles of opposite helicity
are related by parity or space inversion. For electromagnetic and gravitational forces (that obey space inversion
symmetry), the massless particles of opposite helicity are called the same. The massless particles of helicity 1,
associated with electromagnetic phenomena are both called photons. In the same way, the (hypothetical) massless
particles of helicity 2, that are supposed to mediate the gravitational interactions are both called gravitons.
On the other hand, the almost massless particles of helicity 1/2 that are emitted in nuclear beta decay24
have no interaction (apart from gravitation which is negligible for many purposes) that respect the symmetry of
space inversion (weak interactions do not respect space inversion symmetry). Therefore, these particles receive
different names: neutrinos for helicity 1/2, and antineutrinos for helicity +1/2.
23
Observe that for a massless particle, a positive helicity cannot become a negative one (or vice versa). For such a projection to
change its sign, it is necessary that the momentum changes its sense. However, for a massless particle (which travels at the speed of
light) there is no any boost to another reference frame that can reverse the direction of motion i.e. the momentum of the particle.
24
For a long time, it was supposed that such particles were massless (neutrinos and antineutrinos). Nowadays we know these particles
are massive (though with masses much smaller that electron masses) so their helicities are not Lorentz-invariants anymore.
1.12. SPACE INVERSION AND TIME-REVERSAL 53

Though helicity of massless particles is Lorentz-invariant, Eq. (1.239) shows that the state itself is not. For
instance, owing to the helicity-dependent phase exp (i) in Eq. (1.239), a state formed by a linear superposi-
tion of one-particle states with opposite helicities will be changed by a Lorentz transformation into a different
superposition. As an example, a general one-photon state of four-momentum p might be written as

|p; i = + |p, +1i + |p, 1i ; |+ |2 + | |2 = 1

if |+ | =
6 | | and both are non-zero, we are in the case of elliptic polarization. Circular polarization occurs when
one of the coefficients vanishes, in the opposite extreme we have the linear polarization when |+ | = | |. The
overall phase of + and has no physical significance, and for linear polarization they could be adjusted such
that
= + (1.243)
but the relative phase is still important. For linear polarization with the choice (1.243), the phase of + can be
identified as the angle between the plane of polarization and a given fixed reference direction perpendicular to p.
Equation (1.239) shows that under a Lorentz transformation , this angle rotates by an amount (, p).

1.12 Space inversion and time-reversal


We have seen that the complete Lorentz group also contains some improper (discrete) transformations that leave
the pseudonorm of four-vectors invariant. Any non-continuous homogeneous Lorentz transformation can be written
as the product of a proper orthochronus transformation (with det = +1, and 0 0 1) with a discrete operation
that can be either P (space inversion), T (time reversal) or PT .

1 0 0 0 1 0 0 0
0 1 0 0
P = ; T = 0 1 0 0 (1.244)
0 0 1 0 0 0 1 0
0 0 0 1 0 0 0 1

We should study how to extend these discrete operations to the representations U (, a) of the inhomogeneous
Lorentz group in the Hilbert space. We shall first mount what we could call a scenario with conservation of time
reversal and parity. In this scenario, the fundamental multiplication rule (1.90) of the Poincare group
 
U , a U (, a) = U , a + a (1.245)

would be valid even if and/or involves a discrete transformation, and there are discrete operators P and T
on the Hilbert space associated with the discrete operators P and T defined on the Minkowski space

P U (P, 0) ; T U (T , 0) (1.246)

to induce these discrete operators on the Hilbert space, we note that the parameters (, a) which correspond to
an operator and a four-vector on the Minkowski space respectively, transform under the discrete symmetries as
follows
P T P T
PP 1 , T T 1 ; a Pa , a T a
therefore, it is natural to define the transformation of a continuous operator U (, a) on the Hilbert space through
the discrete operators P and T , as follows
 
P U (, a) P 1 = U PP 1 , Pa ; T U (, a) T 1 = U T T 1 , T a (1.247)

for any proper orthochronus transformation and for any translation a.


However, a series of experiments have shown that weak interactions such as those produced in nuclear beta
decay, violates the conservation of these discrete symmetries. In 1956-57 it was shown that parity is violated in
54 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

decays involving weak interactions. In 1964, indirect evidence appeared of violation of the time reversal symmetry.
Therefore, the present scenario will be only approximately true, and should be corrected when parity-violating
theories (theories involving the nuclear weak interaction) are treated.
By now, we shall use a generic operator W to indicate either P or T . An infinitesimal transformation associated
with Eqs. (1.247) is given by

W U (1 + , ) W 1 = U W (1 + ) W 1 , W

W U (1 + , ) W 1 = U 1 + WW 1 , W ; W = P, T and W P, T (1.248)

where we have taken into account that there is no infinitesimal parameter associated with W. Applying the
expansion (1.99) page 27, on both sides of Eq. (1.248) we find
 
1 1 
W 1 + i J i P W 1 = 1 + i WW 1 J i (W) P

2 2
1 1  
W [iJ ] W 1 W [iP ] W 1 = i W W 1 J iW P (1.249)
2 2
now we should take into account that the fact that prevented us to consider antilinear antiunitary operators in
representations of SO (3, 1) (according to Wigners representation theorem) was the continuous connection of all
its elements with the identity. However, discrete symmetries have no these properties and the door is open to
consider the possibility (and even the necessity) of antilinear antiunitary operators in the representations of the
extended Lorentz group. Therefore, we shall not take the i factor out of our W operators by now. Equating
coefficients of and in Eq. (1.249) we obtain

W [iJ ] W 1 = iW W J (1.250)
1
W [iP ] W = iW P ; W = P, T and W P, T (1.251)

where we have used the fact that W 1 = W for both P and T as can be seen from Eq. (1.244) [and using
a definition like (1.49)]. Equations (1.250, 1.251) are quite similar to Eqs. (1.111, 1.112) page 29, except from the
fact that we have not cancelled the i factor on both sides of Eqs. (1.250, 1.251), because of the possibility of
having antilinear antiunitary operators.
Setting = 0 and W = P in Eq. (1.251) we get

P iP 0 P 1 = iP 0 P = iP0 0 P 0
P (iH) P 1 = iH

where H is the energy operator. If P were antiunitary and antilinear then it would anticommute with i. Therefore,
we would obtain P HP 1 = H. From which we would find

H |i = E |i P HP 1 P |i = EP |i HP |i = EP |i
H {P |i} = E {P |i}

Then, for any eigenstate |i of the Hamiltonian with energy E > 0, there would be another state P |i of energy
E < 0. There are no states of negative energy (energy less than that of the vacuum)25 . Hence, it is necessary
for the operator P to be linear and unitary.
Now, setting = 0 and W = T in Eq. (1.251) we get

T (iH) T 1 = iT 0 P = iT0 0 P 0
T (iH) T 1 = iH
25
This also implies that the spectrum is not bounded from below, hence states would decay indefinitely toward lower and lower
energies. Stability would be impossible.
1.12. SPACE INVERSION AND TIME-REVERSAL 55

if we suppose that T is linear and unitary, we can cancel the i factor on both sides of this equation. Hence,
T HT 1 = H, leading again to the conclusion that for any state |i with energy E > 0, there would be a state
T |i of energy E < 0. To avoid this, we must conclude that the time reversal operator T is antilinear and
antiunitary.
Therefore, Eqs. (1.250, 1.251) for each discrete symmetry become

P J P 1 = P P J ; P P P 1 = P P (1.252)
1 1
T J T = T T J ; T P T = T P (1.253)

moreover, Eqs. (1.252, 1.253) can be rewritten in terms of the generators in three-dimensional notation given by
Eqs. (1.126, 1.127, 1.128, 1.129). For example, Eqs. (1.127, 1.85) says that Ji = J kl for some three-component
indices k, l and that Pi i = 1. Hence, we can use J = J kl in Eq. (1.252) and obtain

P Ji P 1 = P J kl P 1 = P k P l J = Pk k Pl l J kl = (1) (1) Ji = Ji

there is no sum over k, l. Similarly we have

T Ji T 1 = T J kl T 1 = Tk T l J = Tk k Tl l J kl = (+1) (+1) Ji = Ji

proceeding similarly with the other generators we obtain

P JP 1 = +J ; P KP 1 = K ; P PP 1 = P (1.254)
T JT 1 = J ; T KT 1 = +K ; T PT 1 = P (1.255)
1 1
P HP = T HT =H (1.256)

it is sensible that P must reverse sign under parity. Further, it is reasonable that parity preserves the sign of J,
at least for the orbital angular momentum since both the position and the linear momentum change signs, thus
the cross product must be invariant. On the other hand time reversal reverses J, once again it is sensible, since
under time reversal and observer would see the film backwards, hence the observer see that the spinning goes
on the opposite direction26 .
Observe for instance that the first of Eqs. (1.255) is consistent with the angular momentum commutation
relations
   
T [Ji , Jj ] T 1 = T Ji Jj T 1 T Jj Ji T 1 = T Ji T 1 T Jj T 1 T Jj T 1 T Ji T 1
= (Ji ) (Jj ) (Jj ) (Ji )
1
T [Ji , Jj ] T = [Ji , Jj ]

on the other hand


T (iijk Jk ) T 1 = iijk T Jk T 1 = iijk Jk

showing that
[Ji , Jj ] = iijk Jk T [Ji , Jj ] T 1 = T (iijk Jk ) T 1

it worths emphasizing that the antilinearity of T (i.e. the fact that T anticommutes with i) was essential in
obtaining this consistency. In general, it can be checked that Eqs. (1.254, 1.255, 1.256) are consistent with all the
commutation relations (1.130-1.137) [homework!!(8)].
Now we examine these discrete symmetries for massive particles and particles with null mass
26
We could think (classically) in time reversal as taking t t, and keeping positions unaltered r r. Since p = mdr/dt under
time reversal we obtain p = mdr/d (t) = p. Therefore J = r p = J.
56 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

1.13 Parity and time-reversal for one-particle states with M > 0.


1.13.1 Parity for M > 0
In Sec. 1.10, we took the reference standard vector k = (0, 0, 0, M ), the little group was SO (3) with generators
J1 , J2 , J3 . This k corresponds to a particle at rest with energy M (only self-energy). Therefore, the associated
eigenvectors |k, i in the Hilbert space are eigenvectors of P, H and J3 with eigenvalues p = 0, p0 = E = M and
respectively. From Eqs. (1.254) and (1.256) we see that P anticommutes with P, and it commutes with J and
H, thus

P [P |k, i] = P [P |k, i] = 0
H [P |k, i] = P [H |k, i] = M [P |k, i]
J3 [P |k, i] = P [J3 |k, i] = [P |k, i]

consequently the state P |k, i has the same quantum numbers 0, M, as |k, i. Therefore27 except for degeneracies
(arising from additional quantum numbers), both states can only differ by a phase

P |k, i = |k, i ; | | = 1 (1.257)

we shall see that the phase is independent of . To see it, we shall apply Eqs. (1.148, 1.163, 1.164) for an
infinitesimal transformation of the little group SO (3)

X
U (W ) |k, i = D (W ) k,

X (j)
U (j) (1 + p , 0) |k, i = D (1 + p ) k,

  X   
1 i
1 + ip Jp |k, i = + p Jp (j) k,
2 2


X 
Jp |k, i = Jp(j) k,

applying this for p = 1, 2 and using Eqs. (1.167) we find


X  (j)  X (j)

J1 |k, i = J1 k, ; iJ2 |k, i = iJ k,
2


X (j) (j)
 Xh p i
(J1 iJ2 ) |k, i = J1 iJ2 k, = ,1 (j ) (j + 1) k,


p
(J1 iJ2 ) |k, i = (j ) (j + 1) |k, 1i (1.258)

where j is the particles spin. Applying P on the LHS of Eq. (1.258), we find
p
P (J1 iJ2 ) |k, i = (J1 iJ2 ) P |k, i = (J1 iJ2 ) |k, i = (j ) (j + 1) |k, 1i (1.259)

and applying P on the RHS of Eq. (1.258), we find


p p
(j ) (j + 1)P |k, 1i = (j ) (j + 1) 1 |k, 1i (1.260)
27
Observe that the fact that P does not modify the quantum numbers 0, M (i.e. the eigenvalues of the four-momentum operator)
of the state |k, i, is related with the invariance under P of the associated classical four-momentum k (0, M ).
1.13. PARITY AND TIME-REVERSAL FOR ONE-PARTICLE STATES WITH M > 0. 57

and equating Eqs. (1.259, 1.260) we find


= 1
then is independent of , so we can rewrite Eq. (1.257) as

P |k, i = |k, i ; || = 1 (1.261)

the phase is called the intrinsic parity. It depends only on the species of particle on which P is acting.
Further, in Sec. 1.10, we saw that to pass from k to an arbitrary four-momentum p , we use the standard
boost L (p) defined by Eq. (1.192). Now, to obtain states |p, i in the Hilbert space with arbitrary momentum p,
we apply the associated unitary operator U (L (p)) according with definition (1.145)
s
k0
|p, i = N (p) U (L (p)) |k, i = U (L (p)) |k, i
p0
s
M
|p, i = U (L (p)) |k, i (1.262)
p0

By using the matrix representation of P, we obtain Pp = p,p0 . Now, we shall examine the transformation
properties of the standard boost L (p) given by Eqs. (1.191, 1.192) through the parity transformation. It is
expected that such a transformation PL (p) P 1 be related with L (Pp). To see this relation, we first evaluate
L (Pp). To do this, we first observe that according with Eq. (1.192) we find
 
(Pp) = p, p0 = p,p0

now, using the other Eqs. (1.192), we find



Li j (Pp) = Li j p,p0 = ij + [ (Pp) 1] (b pj ) = ij + [ (p) 1] pbi pbj = Li j (p)
pi ) (b
L0 0 (Pp) = (Pp) = (p) = L0 0 (Pp)
p p
Li 0 (Pp) = L0 i (Pp) = b pi 2 (p) 1 = Li 0 (p) = L0 i (p)
pi 2 (Pp) 1 = b

in summary we have

Li j (Pp) = Li j (p) ; L0 0 (Pp) = L0 0 (p)


Li 0 (Pp) = L0 i (Pp) = Li 0 (p) = L0 i (p) (1.263)

combining (1.191) with (1.263) the explicit form of L (Pp) yields


p
1 + ( 1) pb21 ( 1) pb1 pb2 p 1 p 2 1
( 1) pb1 pb3 b

( 1) pb1 pb2 1 + ( 1) pb22 ( 1) pb2 pb3 b p 2 p 2 1
L (Pp) = 2 (1.264)
( p1) pb1 pb3 ( p1) pb2 pb3 1 + (
p 1) pb3 b p3 2 1
p1 2 1 b
b p2 2 1 b p3 2 1

On the other hand, we can evaluate the transformation PL (p) P 1 explicitly by taking Eqs. (1.191, 1.244)
p
1 0 0 0 1 + ( 1) pb21 ( 1) pb1 pb2 ( 1) pb1 pb3 pb1 p 2 1
2
0 1 0 0 ( 1) pb p
1 2b 1 + ( 1) p
b ( 1) pb2 pb3 pb2 p 2 1
PL (p) P 1 = 0 2

0 1 0 ( p 1) pb1 pb3 ( p 1) pb2 pb3 1 + p ( 1) pb23 pb3 2 1
0 0 0 1 pb1 2 1 pb2 2 1 pb3 2 1

1 0 0 0
0 1 0 0
0

0 1 0
0 0 0 1
58 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

after multiplying these matrices we obtain the matrix in (1.264). Thus we find
  p 
PL (p) P 1 = L (Pp) ; Pp = p,p0 = p, p2 + M 2 (1.265)

applying P on Eq. (1.262) and using Eqs. (1.247, 1.265), we have


s s s
M M   M  
P |p, i = 0
P U (L (p)) |k, i = 0
P U (L (p)) P 1 P |k, i = 0
U PL (p) P 1 P |k, i
p p p
s
M
P |p, i = [U (L (Pp))] |k, i
p0

which according with Eq. (1.262), can also be written as

P |p, i = |Pp, i (1.266)

1.13.2 Time reversal for M > 0


Equations (1.255, 1.256), show that the effect of T on the zero three-momentum state of reference |k, i yields

P (T |k, i) = 0 ; H (T |k, i) = M (T |k, i) ; J3 (T |k, i) = (T |k, i)

therefore, the quantum numbers 0, M of P and H remain invariant while the quantum number of J3 changes
sign
T |k, i = |k, i (1.267)
where is a phase factor that could depend on . Applying T to (1.258) and recalling that T anticommutes with
J and i, we have
p
T (J1 iJ2 ) |k, i = (j ) (j + 1) T |k, 1i
p
(J1 iJ2 ) T |k, i = (j ) (j + 1) 1 |k, 1i
p
(J1 iJ2 ) |k, i = (j ) (j + 1) 1 |k, 1i (1.268)

and using again Eq. (1.258) in the LHS of Eq. (1.268), it becomes
p
(J1 iJ2 ) |k, i = [j ()] [j () + 1] |k, 1i
p
(J1 iJ2 ) |k, i = [j ] [j + 1] |k, 1i (1.269)

and equating Eqs. (1.268, 1.269) we find


= 1 (1.270)
we write the solution in the form
= (1)j (1.271)
where is another phase that depends only on the species of particle. Combining Eqs. (1.267, 1.271), we obtain

T |k, i = (1)j |k, i (1.272)

but unlike the intrinsic parity , the time-reversal phase has no physical significance because we can redefine
the one-particle states in such a way that this phase is removed

|k, i |k, i = ei 2 |k, i ; ei
1.13. PARITY AND TIME-REVERSAL FOR ONE-PARTICLE STATES WITH M > 0. 59

the new state under time-reversal gives



n o
T |k, i = T ei 2 |k, i = ei 2 T |k, i = (1)j ei 2 ei |k, i = (1)j ei 2 ei 2 ei 2 |k, i
T |k, i = (1)j |k, i

nevertheless, we shall continue using the phase in Eq. (1.272) in order to be free of choosing different phases. But
we should keep in mind that such a phase is physically irrelevant.
Once again, in order to study arbitrary states |p, i, we use the standard Lorentz boost defined by Eq. (1.192).
From Eqs. (1.191, 1.244) we can check explicitly that
 p 
T L (p) T 1 = L (Pp) ; Pp = T p = p, p2 + M 2 (1.273)

in words, changing the sign of each element of L with an odd number of time-indices is the same as changing
the signs of elements with an odd number of space-indices.
Applying P on Eq. (1.262) and using Eqs. (1.247, 1.273), we have
s s s
M M   M  
T |p, i = 0
T U (L (p)) |k, i = 0
T U (L (p)) T 1 T |k, i = 0
U T L (p) T 1 (1)j |k, i
p p p
s
M
T |p, i = (1)j [U (L (Pp))] |k, i
p0

which according with Eq. (1.145) page 33, can be rewritten as

T |p, i = (1)j |Pp, i (1.274)

It worths observing that time-reversal understood as putting the film backwards clearly reverse the sense of
the three-momentum and left the energy invariant. Consequently, it transforms the four-momentum in the same
way as space-inversion. By contrast, we observe that when applying time-reversal to the four-vector of position
x , the three space-components remain invariant while the time-coordinate obviously reverse sign. Consequently,
time-reversal applied on the position four-vector gives minus the space-inversion applied on such a four-vector.
Therefore, the matrix representation of T given by Eq. (1.244) works well on position four-vectors but not on
four-momenta. In turn it has to do with the fact that a matrix representation works well in all circumstances
only if the associated operator is linear. However, time-reversal is antilinear so that the matrix representation
(1.244) is not always adequate. From the previous discussion the statement Pp = T p given in Eq. (1.273) must
be interpreted in terms of the transformation of the operators but not in terms of the matrix representations
(1.244). On the other hand, we also note that if time-reversal acted on the four-momentum according with the
representation of T in Eq. (1.244), the sign of p0 would be reversed leading to a non-physical state.
Finally, we observe that parity is a linear operator, so that the matrix representation (1.244) for P, works well
for both the four-momentum and the position four-vector.

1.13.3 Parity for null mass particles


In section 1.11 we took as the reference four-momentum k = (0, 0, , ), which describes a particle of null mass
traveling at the speed of light along the three-axis. Thus, the associated quantum state |k, i is eigenvector of P
with eigenvalue k = (0, 0, , ), and it is also eigenvector of J3 with eigenvalue (helicity). In the Minkowski
space the parity operator P gives a state with four-momentum

Pk = (0, 0, , )
60 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

thus the associated quantum state must be associated with opposite three-momentum. It can also be seen by
taking into account that |k, i is eigenvector of P with eigenvalue k = (0, 0, ), as follows

P [P |k, i] = P [P |k, i] = P [k |k, i] = k [P |k, i]

as for J3 we see that


J3 [P |k, i] = P [J3 |k, i] = [P |k, i]
hence P does not change the eigenvalue of J3 . However, since J3 is the component of angular momentum along
u3 (i.e. the original direction of motion), and P |k, i has the opposite direction of motion k, we see that the
component of spin along the direction of motion of P |k, i has opposite sign with respect to |k, i (precisely
because did not change its value). Thus, parity inverts the direction of motion and helicity of the state.
This shows that the existence of a space inversion symmetry requires that any species of massless particle with
non-zero helicity must be accompanied with another of opposite helicity. Since P does not leave the standard
momentum invariant, it is convenient to use an operator that provides such an invariance. Hence we shall consider
an operator28 
U R21 P (1.275)
where R21 is a rotation that takes Pk to k (i.e. that reverses the transformation of P over |k, i). Consequently,
R2 is a rotation that takes k to Pk. Of course, the simplest transformation that takes (0, 0, , ) to (0, 0, , )
is a rotation of around the two-axis29
U (R2 ) = exp (iJ2 ) (1.276)

Since U R21 reverses the sign of J3 , we have

U R21 P |k, i = |k, i (1.277)

where is a phase factor that could depend on .


Now we characterize the action of P on an arbitrary one-particle state |p, i. We do it by applying P on Eq.
(1.145) with L (p) given by (1.241) and using Eq. (1.277) to obtain
r   
|p|  1
 1  1
 
P |p, i = N (p) P U (L (p)) |k, i = P U R (b
p ) B U R2 P U R2 P |k, i
p0
r     
|p|  
= 0
P U R (bp) B P 1 U (R2 ) U R21 P |k, i
p
r     
|p| 1
= U PR (bp) B P U (R2 ) |k, i
p0
r     
|p| 1
P |p, i = U PR (bp) B P R2 |k, i (1.278)
p0

Further, it can be checked that the Lorentz boost (1.242) commutes with R21 P (and so with its inverse
P 1 R2 )30 , and that P commutes with the rotation R (b
p) of Eq. (1.177) that takes the three-axis into the direction
28
Since we are considering the extended Poincare group (including discrete operations), we could think about operator (1.275), as
an extension of the Little group that includes discrete symmetries that leave k invariant. We shall also consider an operator of the
form U R21 T as another extension of the little group associated with the discrete time-reversal symmetry.
29
This is a convention since it could also be around the one-axis. An additional original rotation around the three-axis could be
considered, but it is not necessary.
30
If [A, B] = 0, then A1 also commutes with B:
 1   
A ,B = A1 B BA1 = A1 BA A1 A1 ABA1 = BA1 A1 B
 1   
A ,B = A1 , B = 0
1.13. PARITY AND TIME-REVERSAL FOR ONE-PARTICLE STATES WITH M > 0. 61

b . Therefore we have
of p
     
|p| 1
 1
 |p| 1 |p|
PR (b
p) B P R2 = PR (b p) P R2 B p) PP R2 B
= R (b

   
|p|  |p|
PR (b
p) B P 1 R2 = R (b
p) R2 B (1.279)

substituting (1.279) in (1.278) we obtain


r    
|p|
P |p, i = 0
U R (b
p) R2 B |k, i (1.280)
p

it worths emphasizing that R (b p) R2 is a rotation that takes the three-axis into the direction of b
p, however
U (R (b
p) R2 ) is quite different from U (R (bp)), as can be seen by applying Eq. (1.179)

p)) = exp [i ( ) J3 ] exp [i ( ) J2 ]


U (R (b (1.281)

the azimuthal angle is chosen as + if 0 < or as if < 2. It is in order that the azimuthal angle
remains within the interval [0, 2). On the other hand, the two-component of p is given by p u2 = p sin sin ,
since 0 we have sin 0 so that p u2 > 0 if 0 < < , and p u2 < 0 if < < 2. Then we could say
alternatively that the azimuthal angle in (1.281) is chosen as + or according whether the two-component
of p is positive or negative respectively31 .
From Eq. (1.281) we obtain

U 1 (R (b p) R2 ) = U 1 (R (b
p)) U (R (b p)) U (R (b
p)) U (R2 )
= {exp [i ( ) J2 ] exp [i ( ) J3 ]} {exp [iJ3 ] exp [iJ2 ]} exp (iJ2 )
1
U (R (b p) R2 ) = exp [i ( ) J2 ] exp [iJ3 ] exp [i ( ) J2 ]
p)) U (R (b

or equivalently
U (R (b p)) exp [i ( ) J2 ] exp [iJ3 ] exp [i ( ) J2 ]
p) R2 ) = U (R (b
but a rotation of around the threeaxis reverses the sign of J2 . Therefore

U (R (b p)) exp [iJ3 ] exp [i ( ) (J2 )] exp [i ( ) J2 ]


p) R2 ) = U (R (b
U (R (b
p) R2 ) = U (R (b
p)) exp [iJ3 ] (1.282)

Note that despite U (R (bp) R2 ) does not coincide with U (R (bp)) as anticipated, they differ by an initial rotation
around the three-component, and the associated initial rotation around the three-axis in the Minkowski space,
clearly does not alter the action of R (b
p) R2 over the three-axis. From the form of the standard boost (1.241), we
can obtain L (Pp) as follows
   
|p| |p| 
L (p) = R (b
p) B L (Pp) = R (b p) B ; Pp = p, p0 (1.283)

using (1.282, 1.283), we have


        
|p| |p| |p|
p) R2 ) U B
U (R (b p)) exp [iJ3 ] U B
= U (R (b = U (R (b
p)) U B exp [iJ3 ]

(1.284)
31
If the two-component of p is zero, we have = 0 (positive one-component) or = (negative one-component). For = 0 (positive
one-component) we choose + , and for = (negative one-component) we choose .
62 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

and substituting (1.284) in equation (1.280) we have


r   
|p|
P |p, i = 0
U (R (b
p)) U B exp [iJ3 ] |k, i
p
r   
|p|
= 0
U R (b p) B exp [i] |k, i
p
r

P |p, i = exp [i] U (L (Pp)) |k, i
p0
and according with (1.145) page 33, we finally obtain

P |p, i = exp [i] |Pp, i (1.285)

where the phase is or + if the two-component of p is positive or negative respectively. This produces a
change in sign only if is half-integer. Such a change in sign in the operation of parity for masless particles with
half-integer spin is due to the convention adopted in Eq. (1.179) for the rotation used to define massless particles
of arbitrary momentum. Such kind of discontinuities are unavoidable owing to the fact that SO (3) is not simply
connected32 .

1.13.4 Time-reversal for null mass particles


We already saw that the reference four-momentum is k = (0, 0, , ), and that the quantum state |k, i is
eigenvector of P , J3 with eigenvalues k = (0, 0, , ), and respectively. We also see that in the Minkowski
space, the time reversal operator T gives a state with four-momentum

T k = Pk = (0, 0, , ) = p, p0

Further, Eq. (1.255) shows that T anticommutes with P and J. Therefore, it reverses the sign of p and .

P [T |k, i] = T P |k, i = k [T |k, i] ; J3 [T |k, i] = T J3 |k, i = [T |k, i]


hence T does not change the helicity Jk, b because it reverses the sign of both quantities. Consequently, time-
reversal says nothing about whether massless particles of one helicity are accompanied with others of helicity
. Like in the case of P , time-reversal does not leave the stardard four-momentum invariant, from  which it is
more convenient to work with a related operator that does leave k invariant. We see that U R21 T does this
role where R2 is the rotation defined in (1.276), and takes k into Pk. This commutes with J3 , so

U R21 T |k, i = |k, i (1.286)

where is in principle a helicity-dependent phase.


Once again, we shall characterize the action of T on an arbitrary one-particle state |p, i. We do it by applying
T on the state (1.145) and using Eq. (1.286) we find
r   
|p|   1   
T |p, i = N (p) T U (L (p)) |k, i = 0
T U R (b p) B U R21 T U R21 T |k, i
p
r     
|p|  
= 0
T U R (b p) B T 1 U (R2 ) U R21 T |k, i
p
r     
|p| 1
= U T R (b
p ) B T U (R2 ) |k, i
p0
r     
|p| 1
T |p, i = U T R (b p) B T R2 |k, i (1.287)
p0
32
SO (3) is doubly-connected and it in turn leads to double-valued representations.
1.14. ACTION OF T 2 AND KRAMERS DEGENERACY 63

Since R21 T commutes with the boost B (|p| /), and T commutes with the rotation R (b
p), we obtain
       
|p| 1
 1
 |p| 1 |p| |p|
T R (b
p) B T R2 = T R (b p) T R2 B p) T T R2 B
= R (b = R (bp) R2 B

and Eq. (1.287) becomes r    


|p|
T |p, i = U R (b
p) R2 B |k, i
p0
and using Eq. (1.284) we obtain

r   
|p|
T |p, i = U (R (b
p)) U B exp [iJ3 ] |k, i
p0
r   
|p|
= 0
U R (b p) B exp [i] |k, i
p
r

= exp [i] U (L (Pp)) |k, i
p0

and finally, from (1.145) page 33 we find

T |p, i = exp [i] |Pp, i (1.288)

again the positive or negative sign appears according to whether the two-component of p is positive or negative
respectively. If the two-component of p is null, the positive or negative sign appears according to whether the
one-component of p is positive or negative respectively.

1.14 Action of T 2 and Kramers degeneracy


Using (1.274) for massive one-particle states, and taking into account that T is antilinear antiunitary, we have
h i h i
T 2 |p, i = T [T |p, i] = T (1)j |Pp, i = (1)j [T |Pp, i] = (1)j (1)j+ P 2 p,
T 2 |p, i = (1)2j |p, i (1.289)

we shall see that for massless particles the result is the same. For massless particles, if the two-component
of p is positive (negative), then the two-component of Pp is negative (positive). Therefore, since T |p, i =
exp [i] |Pp, i, we have

T |Pp, i = exp [i] P 2 p, = exp [i] |p, i (1.290)

From Eqs. (1.288, 1.290), we obtain

T 2 |p, i = T { exp [i] |Pp, i} = exp [i] T |Pp, i = exp [i] exp [i] |p, i
T 2 |p, i = exp [2i] |p, i

as long as is integer or half-integer, we can rewrite it as

T 2 |p, i = (1)2|| |p, i (1.291)

we usually define the spin of a massless particle as the absolute value of the helicity, from which Eq. (1.289) is
equivalent to (1.291).
64 CHAPTER 1. RELATIVISTIC QUANTUM MECHANICS

Suppose now that T 2 acts on a state associated with a system of non-interacting particles, either massive or
massless. Such an action yields a factor (1)2j or (1)2|| for each particle. Hence if the system consists of an
odd number of particles of half-integer spin or helicity (perhaps with some additional particles of integer spin or
helicity), we obtain an overall phase given by

T 2 |i = |i (1.292)

if we now turn on several interactions, all of which respect time-reversal invariance, this result will be preserved
even if rotational invariance is not respect. As an example, we could add an arbitrary static gravitational
and/or electric field.
Now suppose that |i is an eigenstate of the Hamiltonian. Since T commutes with the Hamiltonian, T |i is
also an eigenstate of H with the same eigenvalue (same energy). It is natural to ask whether T |i is the same
state or a different one. In the latter case we would have a degeneracy (two or more stationary states with the
same energy). Assuming that T |i is the same state as |i, both of them can only differ at most in a phase

T |i = |i

in that case
T 2 |i = T [ |i] = T |i = |i = ||2 |i = |i
in contradiction with Eq. (1.292). Therefore, any energy eigenstate |i that satisfies Eq. (1.292) must be
degenerate with (at least) another eigenstate of the same energy. This is known as Kramers degeneracy. This
conclusion is trivial if the system is in a rotationally invariant environment, because the total angular momentum
j of such an state would have to be half-integer, leading to 2j + 1 = 2, 4, 6, . . .degenerate states. The surprising
result is that even if the rotational invariance is broken (for instance by introducing external fields such as an
electrostatic field) a two-fold degeneracy persists as long as the fields introduced are time-reversal invariant.
As a particular case, if any particle had an electric or gravitational dipole moment then the degeneracy among
its 2j + 1 spin states would be entirely removed in a static electric or gravitational field. It implies that such
dipole moments are forbidden by time-reversal invariance.
Chapter 2

Scattering theory

We have constructed so far one-particle states in the framework of relativistic quantum mechanics. Nevertheless,
in nature and experiments we have sets of interacting particles. By now, we shall study predictions concerning
experiments in which a set of non-interacting particles (that are very far from each other) approach each other
from macroscopically large distances, and interact (or collide) in a microscopically small region, after which the
products of the interaction fly off in such a way that they become far from each other again. We use to say that
we prepare the system of non-interacting particles in a region far from the collision region (the incoming or simply
the in region) at a time t , they approach to the collision or interaction region at a time t 0, and
after the products travel out again they arrive to the outcoming region (the out region) at a time t , in
which the particles are non-interacting again. The in and out regions corresponding to t and t
respectively, are called asymptotic regions. In the asymptotic regions particles are effectively non-interacting,
hence they can be described as direct products of the one-particle states that we have constructed in section 1.9.
In such experiments, we detect particles and infer probability distributions or cross sections. Our aim in
this chapter is to develop a formalism that permits to calculate such probabilities and cross-sections.

2.1 Construction of in and out states


A state that describes several non-interacting particles can be considered as a state that transforms under the
inhomogeneous Lorentz group, as a direct product of one-particle states. We shall label the one-particle states
with the four-momenta p , the third component of spin (or the helicity for massless particles) and an additional
discrete label n that take into account that we could be dealing with several species of particles. The index n
could mean several indices that specifies the particle type by indicating its mass, charge, spin, etc.
The general transformation rule under the inhomogeneous Lorentz transformations for a one-particle state of
a massive particle is obtained by using the transformation rule (1.180) page 40, valid for homogeneous Lorentz
transformations and adding the inhomogeneous part (space-time translations) given by (1.142) page 32
s
(p)0 X (j)
U (, a) |p, i = exp [ia (p) ] 0
D (W (, p)) p, (2.1)
p

and for massless particles, we combine Eq. (1.239) page 51, with Eq. (1.142) page 32 to find
s
(p)0
U (, a) |p, i = exp [ia (p) ] exp [i (, p)] |p, i (2.2)
p0

where W (, p) is the Wigner rotation defined in Eq. (1.181) page 41, associated with the little group of massive
particles SO (3) or the one associated with masless particles ISO(2). The general transformation rule for a set of

65
66 CHAPTER 2. SCATTERING THEORY

non-interacting particles is then obtained by making direct products of the transformation rule (2.1) or (2.2) for
massive and massless particles respectively
s
(p1 )0 (p2 )0
U (, a) |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i = exp {ia [(p1 ) + (p2 ) + . . .]}
p01 p02
X
F1 1 (W (, p1 )) F2 2 (W (, p2 )) p1 , 1 , n1 ; p2 , 2 , n2 ; . . .
1 2
(2.3)

where (
(j )
Dii (W (, pi )) if p2 < 0
Fi i (W (, pi )) = i (2.4)
exp [ii (, pi )] if p2 = 0
(j)
where D (W (, p)) are the unitary matrices associated with the (j) irreducible representation of SO (3) with
dimension 2j + 1, and the (, p) angle is the one defined in Eq. (1.240) page 52. The states are normalized
according with Eq. (1.162) page 38

 
p1 , 1 , n1 ; p2 , 2 , n2 ; . . . |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i = 3 p1 p1 1 1 n1 n1 3 p2 p2 2 2 n2 n2
permutations (2.5)

the term permutations is added to take into account of the possibility that it is some permutation of the
particle types n1 , n2 , . . .that are of the same species as the particle types n1 , n2 , . . .; we shall see later that its sign
is 1 if this permutation includes an odd permutation of half-integer spin particles, and it is +1 otherwise.
We shall often abbreviate the notation by using a single symbol (a greek letter) for the whole collection of
quantum numbers
(p1 , 1 , n1 ; p2 , 2 , n2 ; . . .) (2.6)
then we have the following assigments
Z X X Z
|i |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i ; d d3 p1 d3 p2 (2.7)
n1 ,1 n2 ,2

in summing and integrating over states it is understood that we include only configurations that do not differ
simply by the exchange of identical particles. In this notation, orthonormalization (2.5) and completeness for
states normalized as in (2.5) are written as


|i = 3 (2.8)
Z
|i = d |i h |i (2.9)

we should keep in mind that the transformation rule (2.3) is only valid for non-interacting particles. Setting
= and a = (0, 0, 0, ), that is, using a pure time-translation U (, a) = exp (iH ), Eq. (2.3) requires
among other conditions, that |i be an energy eigenstate

H |i = E |i (2.10)

with energy equal to the sum of one-particle energies

E = p01 + p02 + . . . (2.11)

without interaction terms. That is, no terms involving two or more particles at a time.
2.1. CONSTRUCTION OF IN AND OUT STATES 67

In scattering processes, equation (2.3) applies to both asymptotic states that we call in states denoted as
|+ i (at t ) and out states denoted as | i (at t +), the asymptotic states |+ i and | i will
be found to contain the particles described by the label if observations are made at t or t +
respectively.
To keep manifest Lorentz invariance, it is more convenient to use the Heisenberg picture, in which state-
vectors do not change in time, so that a state-vector |i describes the whole space time history of a system of
particles. In this picture, operators carry all time-dependence. Therefore, we do not say that | i are the limits
at t of a time dependent state vector | (t)i.
By defining a state we are implicitly choicing an inertial reference frame. Different observers see equivalent
state-vectors but not the same state-vector. Suppose that a standard observer O sets its clock so that t = 0 is a
given time during the collision process, and another observer O (at rest with respect to O) sets its clock so that
t = 0 is at a time t = . The two observers time-coordinates are related by t = t . If O sees the system in a
state |i, O will see the system in a state

U (1, ) |i = exp [iH ] |i

hence, the appearance of the state long before or after the collision (asymptotic states in whatever the basis used
by O) is found by applying a time-translation operator exp [iH ] with or +, respectively. If
the state is an energy eigenstate, it cannot be localized in time (because of the time-energy uncertainty principle,
the characteristic time of evolution of the system is infinite). In that case, the operator exp [iH ] simply yields
an irrelevant phase exp [iE ]. However, the situation is different when the state |i consists of a wave-packet
of energy eigenstates Z
|i = d g () |i (2.12)

we shall assume that the amplitude g () is non-zero and varies smoothly over some finite range E of energies.
The in and out states are defined such that the superposition
Z Z

exp [iH ] d g () = d g () exp [iE ]

has the form of a corresponding superposition of free particle states for << (E)1 or >> (E)1 ,
respectively.
To construct this in and out states, suppose we can divide the time-translation generator (Hamiltonian)
into two terms, a free-particle Hamiltonian H0 and an interaction V

H = H0 + V

in such a way that H0 has eigenstates |i(0) with the same appearance as the eigenstates | i of the complete
Hamiltonian
E E

H0 (0) = E (0) (2.13)
D E 

(0) (0) = (2.14)

where we have assumed that H0 has the same spectrum as the full Hamiltonian H. It demands that the masses
appearing in H0 be the physical masses that are actually measured, which are not necessarily the bare mass terms
appearing in H. If there is any difference, it must be included in V and not in H0 . Further, any relevant bound
states in the spectrum of H should be introduced into H0 as if they were elementary particles. For instance,
in rearrangement collisions in which some bound states appear in the initial state but not the final state or
vice-versa, one must use a different split of H into H0 and V in the initial and final states. In non-relativistic
problems it is usual to include the binding potential in H0 .
68 CHAPTER 2. SCATTERING THEORY

The in and out states can now be defined as eigenstates of H, not H0



H = E (2.15)
(0)
since the form or appearance of is the same as the one of eigenstates | i, the in and out states satisfy
the condition Z Z E


d g () exp [iE ] d g () exp [iE ] (0) (2.16)

for and + respectively. Equation (2.16) can be rewritten as


Z Z E


lim exp [iH ] d g () lim exp [iH0 ] d g () (0) (2.17)

we can write Eq. (2.17) as


Z Z E

lim d g () lim exp [+iH ] exp [iH0 ] d g () (0) (2.18)

now, apart from being smooth, g () is arbitrary from which Eq. (2.18) leads to a formula for the in and out
states E
= () (0) ; ( ) exp [+iH ] exp [iH0 ] (2.19)
nevertheless, we should take into account that () in Eq. (2.19) gives meaningful results only when acting
on a smooth superposition of energy states.
As a consequence of definition (2.16), the in and out states are normalized like the free-particle states.
We can see it by observing that the LHS of Eq. (2.16) is obtained by applying the unitary operator exp [iH ]
to a time independent state (since we are in the Heisenberg picture)
Z Z


d g () exp [iE ] = exp [iH ] d g ()

therefore, its norm is time-independent, in particular equals the norm of its limit , which is the norm of
the RHS of Eq. (2.16). The equality of the square of the norms of the wave-packets on both sides of (2.16) can
be written as
Z Z D E



d d exp [i (E E ) ] g (a) g () = d d exp [i (E E ) ] g (a) g () (0) (0)

since this equality holds for all smooth functions g (), the scalar products must be equal, thus

D (0) (0) E
= = ( )

for future purposes we shall write an explicit formal solution of the energy eigenvalue equation (2.15) that satisfies
the conditions (2.16). To do it, we rewrite Eq. (2.15) as

(H0 + V ) = E

(E H0 ) = V

the operator E H0 contains at least one null eigenvalue [since any given E is part of the spectrum of H0
according with Eqs. (2.13, 2.15)], therefore it is not invertible. Such an operator
annihilates not only the free-

particle state (0) but also the continuum of other free particle states (0) of the same energy. In order to
permit the invertibility of the operator we shall shift it by a quantity i, where is a positive infinitesimal
number1
(E H0 i) = V (2.20)
1
The shift is made in the imaginary part to ensure that E i is not part of the spectrum of H0 (which must be real).
2.1. CONSTRUCTION OF IN AND OUT STATES 69

we shall write a temptative


(0) solution of Eq. (2.20), as the solution of the homogeneous equation (with V 0)

which is given by , plus a particular solution obtained by taking into account that the operator E H0 i
is invertible (0) E
= + (E H0 i)1 V (2.21)

using the completeness of the free-particle states (0) this solution becomes
E Z ED 
(0) (0) (0)
= + d (E H0 i)1 V
E Z E nD o

= (0) + d (0) (0) (E H0 i)1 V
E Z E nD o
(0) (0) (0) 1
= + d (E E i) V

where we have used the fact that H0 is assumed to have the same spectrum as the full Hamiltonian H. We then
have finally
Z
(0) E
T
(0)
E D
(0)
= + d ; T
V (2.22)
(E E i)

Expresions (2.21, 2.22) are known as the Lippmann-Schwinger equations. Now we should show that equations
(2.22) with +i or i in the denominator, satisfies the condition (2.16) for an in or an out state, respectively.
To show it, let us consider the wave-packets
Z

g (t) d eiE t g () (2.23)
E Z E
(0)
g (t) d eiE t g () (0) (2.24)

we want to show that |g+ (t)i and |g (t)i approach g(0) (t) when t and t +, respectively.
Substituting (2.22) in (2.23) we obtain
Z ( )
E Z T E
g (t) (0)
d eiE t g () (0) + d
(E E i)
Z E Z Z E
eiE t g () T (0)
g (t) = d eiE t g () (0) + d d
(E E i)

by using Eq. (2.24) and interchanging the order of integration we find

E Z EZ
eiE t g () T
g (t) = g(0) (t) + d (0) d
(E E i)
E Z E
g (t) = g(0) (t) + d (0) I (2.25)

Z
eiE t g () T
I d (2.26)
(E E i)

we examine first the case for t . By making the complex extension of E , the exponential in Eq. (2.26) for
t yields
eiE t = ei(ReE +iImE )|t| = ei(ReE )|t| e(ImE )|t| (2.27)
the integration in (2.26) is over all quantum numbers, in particular over the energy. As for the integration with
respect to energy, we can close the contour of integration for the energy variable in the upper half complex
70 CHAPTER 2. SCATTERING THEORY

plane with a large semi-circle. It is clear that at any point of the semi-circle (except over the real axis) we have
ImE > 0. Therefore, it is clear from (2.27) that the contribution from this semi-circle is killed by the factor
exp [iE t], which is exponentially small for t and ImE > 0. The integral is then given by a sum over
the singularities of the integral in the upper half plane.

The functions g () and T could in general have singularities at values of E with finite positive imaginary
parts, but their contribution is exponentially attenuated for t . For this fact to be satisfied, t must be
much greater than both the time uncertainty in the wave-packet g () and the duration of the collision, which

respectively govern the location of the singularities of g () and T in the complex E plane. This leaves the
1
singularity in (E E i) . In this factor, the singularity in the upper-half plane is at ImE = , since
0+ then |ImE | 0 and the exponential in Eq. (2.27) does not necessarily damp this contribution. However
such a singularity in the upper half-plane appears for I but not for I+ . We conclude then that I+ vanishes for
t .
In a similar way, for t + we must close the contour of integration in the lower half-plane, and we see
that
I vanishes in this limit. This reasoning along with Eq. (2.25) shows that |g+ (t)i approaches g(0) (t) when

t , and that |g (t)i approaches g(0) (t) when t +, showing that the condition (2.16) is satisfied.
For future purposes, we shall represent (E E i)1 in a more convenient way as

1 1 (E i) (E i) E
= = 2 2
= 2 2
i
E i E i (E i) E + E + (E + 2 )
2

Hence we write
P
(E i)1 = i (E) (2.28)
E
P E
2 2
; (E) (2.29)
E E + (E + 2 )
2

the function P/E in (2.29) approaches 1/E for |E| >> , and vanishes for E 0, so for 0 is behaves like the
principal value function P/E, that permits to integrate 1/E times any smooth function of E, by excluding an
infinitesimal interval around E = 0. On the other hand, the function
 (E) in (2.29) is of order for |E| >> , and
2
gives unity when integrated over all E. Indeed / x + 2 is one of the well-known functions that approaches
the Dirac delta function when approaches zero from the positive side. From this discussion, we can drop the
label in Eq. (2.28) to write
P
(E i)1 = i (E)
E

2.2 The Smatrix


In scattering experiments, the initial state (the in state |+ i) is usually prepared to have a definite particle
content (defined by the set of quantum numbers ) at t , and then it is measured how the state looks like
at t (the out state | i) with particle content . The probability amplitude for the transition is
thus the scalar product

S = + (2.30)
this array of complex amplitudes is called the Smatrix. If there were no interactions, the in and out states
would be the same so that
S = ( ) for non-interaction (2.31)
The rate for a reaction is proportional to

|S ( )|2 (2.32)
2.2. THE SMATRIX 71

It is important to take into account that in and out states belong to the same Hilbert space. They
only differ in the way they are labelled: by their appearance at t or at t +. Any in state can be
expanded as a superposition of out states by using the completeness of the latter
Z Z
+

= d + = d S (2.33)

thus the coefficients of the expansion are the Smatrix elements (2.30). Thus, the role of this matrix is to connect
two complete sets of orthonormal states, hence it must be unitary. This fact is more apparent by applying the
orthonormality and completeness of both in and out states
Z Z Z  

+ +
+
+

( ) = = d
= d S S = d S

S

Z Z Z  




( ) = = d + + = d S S
= d S S

so we have obtained
Z  
d S
S = ( ) SS = 1 (2.34)

Z  
d S S = ( ) SS = 1 (2.35)

it worths emphasizing that the conditions S S = 1 and SS = 1 are not equivalent for matrices in infinite
dimensions.
In many cases, instead of dealing with the S matrix, it is convenient to work with an operator S, defined
such as its matrix elements between free-particle states coincide with the corresponding elements of the Smatrix
D E

(0) S (0) S (2.36)

since the formula (2.19) connects in and out states with states of free particles (but with the same particle
content), it provides a formula for the Soperator

nD (0) on

Eo
S = + = (+) () (0)
D E

S = (0) (+) () (0) (2.37)

observe that (t) does not depend on the particle content. Comparing Eqs. (2.36, 2.37) and using the second of
Eqs. (2.19) we have

S = (+) () U (, ) (2.38)

U (, 0 ) ( ) (0 ) = exp [+iH0 ] exp [iH ( 0 )] exp [iH0 0 ] (2.39)

we shall derive an alternative formula for the Smatrix by returning to Eqs. (2.25, 2.26), for the in state
|g+ (t)i, and then taking t +
Z Z
+ (0) E E +
eiE t g () T
g (t) = g (t) + d (0) I + ; I+ d (2.40)
(E E + i)

We now close the contour of integration for E in the lower half-plane, running from E = to E = +,
and then back to E = through a large semi-circle in the lower half-plane. Once again, the singularities in
+
T and g () are damped by the exponential at t + in the lower half-plane, so they give no contribution.
72 CHAPTER 2. SCATTERING THEORY

But we now pick up a contribution coming from the factor (E E + i)1 . With our choice of circulation,
the singularity is circled in a clockwise sense. We have a pole of first order and by the method of residues, the
contribution to the integral over E is given by
+

eiE t g () T
+
2i (E E + i) = 2i eiE t g () T
(E E + i) E =E i
E E =i

i.e. the value of the integrand at E = E i times a factor 2i. Therefore, in the limit 0+ and t +,
the integral I+ over in (2.40) has the asymptotic behavior
Z
+ iE t +
I 2ie d (E E ) g () T (2.41)

hence for t +, and using (2.24, 2.41), equation (2.40) gives


Z E Z EZ
+ (0) iE t (0)
g (t +) = d eiE t
g () 2i d e +
d (E E ) g () T
Z E  Z 
+
g (t +) = d (0)
eiE t +
g () 2i d (E E ) g () T (2.42)

and expanding (2.23) in a complete set of out states, we have


Z Z Z
+ +

g (t) d e iE t
g () = d e iE t
g () d +
Z Z
+
g (t) = d e iE t
g () d S

and since S contains a factor (E E ), we can write


Z Z
+
g (t) = d eiE t d g () S

and using the property (2.16) that defines the out states, the asymptotic behavior of |g+ (t)i for t + yields
Z E Z
+
g (t +) = d (0) eiE t d g () S (2.43)

comparing Eqs. (2.42, 2.43), and using the linear independence of (0) , we have
Z Z
+
d g () S = g () 2i d (E E ) g () T
Z Z h i
+
d g () S = d g () ( ) 2i (E E ) T

+
using the arbitariness of g () and T we obtain
+
S = ( ) 2i (E E ) T (2.44)

For a weak interaction V , equation (2.44) suggests a simple aproximation for the Smatrix: for small V we can
+
neglect the difference between the in states and the free particles states in the definition of T given in Eq.
(2.22). In that case Eq. (2.44) becomes
D E

S ( ) 2i (E E ) (0) V (0) (2.45)
2.2. THE SMATRIX 73

equation (2.45) is known as the Born approximation. We shall study later how to calculate higher-order terms.
On the other hand, the Lippmann-Schwinger equations (2.21) for the in and out states can provide an
alternative proof of the orthonormality of these states, the unitarity of the Smatrix, and the validity of Eq.
(2.44) without dealing with limits at t .
First by applying the Lippmann-Schwinger equations (2.21) either on the LHS or the RHS of the matrix
element h | V | i we find

nD
o D


V = (0) + V (E H0 i)1 V = (0) V + V (E H0 i)1 V


n E o
E

V = V (0) + (E H0 i)1 V = V (0) + V (E H0 i)1 V

and equating both expressions, we have



(0) E
1
D

(0)


V
+ V (E H0 i) V = V + V (E H0 i)1 V




T + V (E H0 i)1 V = T
+ V (E H0 i)1 V




T
T = V (E H0 i)1 V V (E H0 i)1 V (2.46)
(0)
inserting an identity with free-states on the RHS of (2.46) we have
Z ED 

1 (0) (0)

D V (E H0 i) d V
Z ED 


V (E H0 i)1 d (0) (0) V
Z

n
Eo D

= d V (E H0 i)1 (0) (0) V
Z

n
Eo D

d V (E H0 i)1 (0) (0) V
Z

ED

= d (E E i)1 V (0) (0) V
Z

ED

d (E E i)1 V (0) (0) V
Z Z
D = d (E E i) T T d (E E i)1 T
1
T (2.47)

and equating Eqs. (2.46, 2.47) we obtain


Z h i
T T = d (E E i)1 (E E i)1 T

T (2.48)

By dividing Eq. (2.48) by E E 2i we get


Z
T T
= dF T T (2.49)
E E 2i E E 2i
1 1
F
(E E i) (E E 2i) (E E i) (E E 2i)
as for F we have
(E E i) (E E i) (E E 2i)
F = =
(E E i) (E E 2i) (E E i) (E E i) (E E 2i) (E E i)
1
F = (2.50)
(E E i) (E E i)
74 CHAPTER 2. SCATTERING THEORY

substituting (2.50) in (2.49) we find


Z
T T T T
= d
E E 2i E E 2i (E E i) (E E i)
Z
T T T T
+ = d
E E 2i E E 2i (E E i) (E E i)

! Z
!
T T T
T
+ = d
E E 2i E E 2i E E i (E E i)

the factors in the denominators on the integrals on the RHS can be replaced with 2, since the only important
thing is that these are positive infinitesimals. Hence

! Z
!
T T T
T
+ = d (2.51)
E E i E E i E E i (E E i)

By applying Eq. (2.51), we shall show now that the quantity ( ) + (E E i)1 T
+
defines a unitary
matrix.
+
T   +
T
K = ( ) + ; K
= K = ( ) + (2.52)
E E i E E i
" #" #
  +
T +
T
K K = ( ) + ( ) +
E E i E E i
+ +
T T
= ( ) ( ) + ( ) + ( )
E E i E E i
+ +
T T
+
E E i E E i

hence the matrix product gives


Z Z "
  +
T
d K K = d ( ) ( ) + ( )
E E i
+
#+
T +
T T
+ ( ) +
E E i E E i E E i
Z   + + Z + +
T T T T
d K K = ( ) + + + d
E E i E E i E E i E E i
Z  
d K K = ( ) + F (2.53)

+
! + Z ! +
T T +
T T
F + + d
E E i E E i E E i E E i

and using Eq. (2.51) we see that F = 0. Therefore, Eqs. (2.52, 2.53) become

+ Z  
T
K = ( ) + ; d K K = ( ) (2.54)
E E i
2.3. SYMMETRIES OF THE SMATRIX 75

on the other hand Eq. (2.22) can be rewritten as


" #
E Z (0)
T
Z
T E
(0) (0)
= + d = d ( ) +
E E i E E i
Z E

= d (0) K (2.55)

 
since (0) is an orthonormal basis and Eqs. (2.52, 2.55) show that (0) is connected with | i through a
unitary operation, we conclude that | i is also an orthonormal basis.
The unitarity of the Smatrix is proved similarly by multiplying (2.48) with (E E ).

2.3 Symmetries of the Smatrix


We shall study what is meant by the invariance of the Smatrix under various symmetries. Further, we shall
study the conditions on the Hamiltonian, to ensure such invariance properties.

2.3.1 Lorentz invariance

For any proper orthochronus Lorentz transformation in the Minkowski space x x + a, we cna define a unitary
operator on the Hilbert space U (, a) by specifying that it acts as in Eq. (2.3) on either the in or the out
states. We say that a theory is Lorentz invariant when the same operator U (, a) acts as in (2.3) on both in
and out states. Since the operator U (, a) is unitary, we may write



S = + = U (, a) U (, a) + (2.56)

and combining Eqs. (2.3, 2.56) we obtain the Lorentz invariance (indeed covariance) property of the Smatrix,
for arbitrary Lorentz transformations and four-translations a . We start rewriting Eq. (2.3) as
s
(p1 )0 (p2 )0
U (, a) |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i = exp {ia [p1 p2 . . .]}
p01 p02
X
F1 1 (W (, p1 )) F2 2 (W (, p2 )) |p1 , 1 , n1 ; p2 , 2 , n2(2.57)
; . . .i
1 2

and the adjoint of this equation is


s

   (p1 )0 (p2 )0
p1 , 1 , n1 ; p2 , 2 , n2 ; . . . U (, a) = exp ia p
1 + p2 + . . .
p0 0
1 p2
X  

F1 1 W , p1 F2 2 W , p2 p1 , 1 , n1 ; p2 , 2 , n(2.58)

2; . . .

1 2

Now if we assume that the in and out states are given by


+
= |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i ; = p1 , 1 , n1 ; p2 , 2 , n2 ; . . . (2.59)
76 CHAPTER 2. SCATTERING THEORY

then we can combine Eqs. (2.56, 2.59) with Eqs. (2.57, 2.58) to obtain
  
Sp1 ,1 ,n1 ;p2 ,2 ,n2 ;...;p1,1 ,n1 ;p2 ,2 ,n2 ;... = exp ia p
1 + p2 + . . . p1 p2 . . .
s
(p1 )0 (p2 )0 (p1 )0 (p2 )0

p01 p02 p0 0
1 p2
X
F1 1 (W (, p1 )) F2 2 (W (, p2 ))
1 2
X  
F1 1 W , p1 F2 2 W , p2
1 2
Sp1,1 ,n1 ;p2,2 ,n2 ;...;p1 ,1 ,n1 ;p2 ,2 ,n2 ;... (2.60)

where the factors Fi i (W (, pi )) are defined by Eq. (2.4). In particular, since the LHS of Eq. (2.60) is
independent of a , the RHS must also be. Consequently, the Smatrix vanishes unless the four-momentum is
conserved. We can then parameterize the part of the Smatrix that represents actual interactions among the
particles in the form
S ( ) = 2iM 4 (p p ) (2.61)

where we have used a structure similar to (2.44) but including the conservation of the four-momentum (not only
the conservation of the energy). We shall see later that M contains additional delta factors.
Indeed, equation (2.60) is not a theorem but a definition of what we mean by the Lorentz invariance of the
Smatrix. We shall see that only certain chosen Hamiltonians lead to the existence a unitary operator that acts
as in (2.3) [or equivalently as in (2.57)] on both in and out states. Hence, one of our main tasks is to find the
conditions on the Hamiltonian that ensures the Lorentz invariance of the Smatrix. To find them, it is useful to
work with the Soperator
D E
(0) (0)
S = S

the free particle states defined in Section 1.9 provides a representation of the inhomogeneous Lorentz group, such
that we can always define a unitary operator U0 (, a) that induces the transformation (2.3) [or equivalently the
transformation (2.57)] on these states
s
E (p1 )0 (p2 )0

U0 (, a) p1 , 1 , n1 ; p2 , 2 , n2 ; . . .(0) = exp {ia [p1 + p2 + . . .]}
p01 p02
X E

F1 1 (W (, p1 )) F2 2 (W (, p2 )) p1 , 1 , n1 ; p2 , 2 , n2 ; .(2.62)
. .(0)
1 2

Equation (2.62) will hold if the unitary operator U0 (, a) commutes with the Soperator

U0 (, a)1 SU0 (, a) = S (2.63)

as in section 1.7.3, the condition (2.63) can be expressed in terms of infinitesimal transformations, leading to
commutation relations with the generators. We shall denote the generators of these infinitesimal transformations as
a momentum P0 , angular momentum J0 , and a boost generator K0 , that along with H0 generates the infinitesimal
inhomogeneous Lorentz transformations, when acting on free-particle states. Hence Eq. (2.60) is equivalent to say
that the Smatrix is unaffected by these transformations [which is more clear in the equivalent equation (2.56)].
In turn, it is equivalent to the fact that the Soperator commutes with these generators

[H0 , S] = [P0 , S] = [J0 , S] = [K0 , S] = 0 (2.64)


2.3. SYMMETRIES OF THE SMATRIX 77


Since the operators H0 , P0 , J0 , K0 generate infinitesimal inhomogeneous Lorentz transformations on states (0) ,
they obey the commutation relations (1.130-1.137) page 31
h i
J0i , J0j = iijk J0k (2.65)
h i
J0i , K0j = iijk K0k (2.66)
h i
K0i , K0j = iijk J0k (2.67)
h i
J0i , P0j = iijk P0k (2.68)
h i
K0i , P0j = iH0 ij (2.69)
 i    h i
J0 , H0 = P0i , H0 = P0i , P0j = 0 (2.70)
 i 
K0 , H0 = iP0i (2.71)

in the same way, we can define a set of exact generators P, J, K, H (H is of course the full Hamiltonian),
such that they generate the transformations (2.3) [or equivalently (2.57)] on, say the in states. It is not obvious
however that the same operators generate the same transformations on the out states. The group structure
says that these exact generators satisfy the same commutation relations (1.130-1.137)
 
J i, J j = iijk J k (2.72)
 
J i, K j
= iijk K k (2.73)
 i j
 k
K ,K = iijk J (2.74)
 i j
J ,P = iijk P k (2.75)
 i j
K ,P = iHij (2.76)
 i     
J ,H = P i, H = P i, P j = 0 (2.77)
 i 
K ,H = iP i (2.78)

In almost all known field theories, the effect of interactions is to add a potential term V to the Hamiltonian, but
leaving the momentum and angular momentum unchanged

H = H0 + V , P = P0 , J = J0 (2.79)

the only known exceptions are theories with topologically twisted fields, for instance theories with magnetic
monoples, where the angular momentum of the states depends on the interactions. Eq. (2.79) implies that the
commutation relations (2.65, 2.68) become the commutation relations (2.72, 2.75). If in addition the interaction
commutes with the free-particle momentum and angular momentum i.e.

[V, P0 ] = [V, J0 ] = 0 (2.80)

and using (2.70) we obtain


     
J i, HJ0i , H0 + V = J0i , H0 = 0
=
 i     
P , H = P0i , H0 + V = P0i , H0 = 0

then Equation (2.70) also becomes Eq. (2.77). On the other hand, the Lippmann-Schwinger equation (2.21) or
equivalently Eq. (2.19) show that the operators that generate translations and rotations on the in and out
states are simply P0 and J0 . To see it, we observe that P0 is the generator of translations for the free-particle
78 CHAPTER 2. SCATTERING THEORY


states (0) and P0 commute with both H0 and H. Let U (1, a) be a pure translation operator on free-particle
states (hence its generators are P0 ). Using the previous facts and multiplying Eq. (2.19) by U (1, a) yields
E E

U (1, a) = U (1, a) () (0) = U (1, a) exp (+iH ) exp (iH0 ) (0) = exp (+iH ) exp (iH0 ) U (1, a) (
E

= exp (+iH ) exp (iH0 ) exp [ia (p1 + p2 + . . .)] (0)
n Eo

= exp [ia (p1 + p2 + . . .)] exp (+iH ) exp (iH0 ) (0)

U (1, a) = exp [ia (p1 + p2 + . . .)]

and we conclude that U (1, a) is also a pure translation operator on in and out states. A similar argument
shows that J0 is the generator of rotations for in and out states. From the fact that P0 and J0 commute with
H0 and H, we also see that P0 and J0 commute with the operator U (t, t0 ) defined in Eq. (2.39). Therefore, P0 and
J0 commute with the Soperator U (, ) defined by Eq. (2.38). Finally, since there are energy-conservation
delta functions in both terms of (2.44), we see that the Soperator commutes with H0 .
We still have to show that K0 commutes with the Soperator. Unlike the case of P and J, we cannot set the
boost generator K as equal to the free-particle counterpart K0 , because Eqs. (2.69) and (2.76) would lead to
   
H = i K i , P i = i K0i , P0i = H0

and we obtain H = H0 , in contradiction with the fact that the interaction V is added to the Hamiltonian. Then,
when we add a correction V to the time-translation generator H0 , we must also add a correction W to the
boost generator K0
K = K0 + W (2.81)
and we shall concentrate on the commutation relation (2.78). The LHS of Eq. (2.78) yields
 i       
K , H = K0i + W i , H = K0i , H + W i , H
         
= K0i , H0 + K0i , V + W i , H = iP0i + K0i , V + W i , H
 i     
K , H = iP i + K0i , V + W i , H (2.82)

where we have used (2.71). Substituting (2.82) in (2.78) we have


       
iP + K0i , V + W i , H = iP i K0i , V + W i , H

so we obtain the condition


[K0 , V ] = [W, H] (2.83)
The condition (2.83) is empty by itself, since for any V we could define W to satisfy such a condition: the
matrix representations of operators on both sides of Eq. (2.83) must coincide in any basis. In particular, by using
the basis of Heigenstates |i and |i equation (2.83) becomes

h| [K0 , V ] |i = h| [W, H] |i (2.84)

and the RHS of Eq. (2.84) yields

h| [W, H] |i = h| [WH HW] |i = h| [WE E W] |i = (E E ) h| W |i


such that Eq. (2.84) becomes
h| [K0 , V ] |i = (E E ) h| W |i
consequently, the condition (2.83) is satisfied for any V by defining the matrix elements of W between Heigenstates
|i and |i as
h| [K0 , V ] |i
W = h| W |i = (2.85)
E E
2.3. SYMMETRIES OF THE SMATRIX 79

We should observe that to guarantee the Lorentz invariance of a theory is not enough to show the existence of a set
of exact generators satisfying the Lie algebra (2.72-2.78), but also that these operators should act the same way
on the in and the out states. Therefore, it is not enough to find an operator K that satisfies Eq. (2.83), we
also require the remaining Lorentz invariance condition that K0 commutes with the Soperator [see Eq. (2.64)].
To show the conditions to obtain [K0 , S] = 0, we shall consider the commutator of K0 with the operator
U (t, t0 ) defined by Eq. (2.38). We shall apply now the following result

Theorem 2.1 Suppose we have two operators A and B such that B commutes with their commutator, that is

[B, C] = 0 ; C [A, B] (2.86)

if F (B) is a function of the operator B then we have

[A, F (B)] = [A, B] F (B) (2.87)

where F (B) is the derivative of F (B) with respect to B defined as



X
X
n
F (B) = fn B F (B) nfn B n1 (2.88)
n=0 n=0

in particular, it is easy to show that [exp (B)] = exp (B).

We start evaluating the commutator


[K0 , exp (iH0 t)]
with A = K0 and B = H0 Eq. (2.71) gives C = [K0 , H0 ] = iP0 , and since P0 commutes with H0 , we can apply
Eq. (2.87) to obtain

[K0 , exp (iH0 t)] = [K0 , H0 ] {it exp (iH0 t)} = iP0 {it exp (iH0 t)}
[K0 , exp (iH0 t)] = tP0 exp (iH0 t) (2.89)

further if Eq. (2.80) is satisfied, then P0 = P commutes with H and from Eqs. (2.87, 2.78) we obtain

[K, exp (iHt)] = [K, H] {it exp (iHt)} = tP exp (iHt)


[K, exp (iHt)] = tP0 exp (iHt) (2.90)

and using Eqs. (2.89, 2.90) along with (2.38) we have

[K0 , U (, 0 )] = [K0 , exp {+iH0 } exp {iH ( 0 )} exp {iH0 0 }]


= exp {+iH0 } [K0 , exp {iH ( 0 )} exp {iH0 0 }]
+ [K0 , exp {+iH0 }] exp {iH ( 0 )} exp {iH0 0 }
[K0 , U (, 0 )] = exp {+iH0 } exp {iH ( 0 )} [K0 , exp {iH0 0 }]
+ exp {+iH0 } [K0 , exp {iH ( 0 )}] exp {iH0 0 }
+ P0 exp (iH0 ) exp {iH ( 0 )} exp {iH0 0 }

[K0 , U (, 0 )] = exp {+iH0 } exp {iH ( 0 )} [0 P0 exp (iH0 0 )]


+ exp {+iH0 } [K0 , exp {iH ( 0 )}] exp {iH0 0 }
+ P0 exp (iH0 ) exp {iH ( 0 )} exp {iH0 0 }
80 CHAPTER 2. SCATTERING THEORY

[K0 , U (, 0 )] = ( 0 ) P0 exp {+iH0 } exp {iH ( 0 )} exp (iH0 0 )


+ exp {+iH0 } [K W, exp {iH ( 0 )}] exp {iH0 0 }
[K0 , U (, 0 )] = ( 0 ) P0 U (, 0 ) + exp {+iH0 } [K, exp {iH ( 0 )}] exp {iH0 0 }
exp {+iH0 } [W, exp {iH ( 0 )}] exp {iH0 0 }

[K0 , U (, 0 )] = ( 0 ) P0 U (, 0 ) + exp {+iH0 } { ( 0 ) P0 exp {iH ( 0 )}} exp {iH0 0 }


exp {+iH0 } [W, exp {iH ( 0 )}] exp {iH0 0 }
[K0 , U (, 0 )] = ( 0 ) P0 U (, 0 ) ( 0 ) P0 U (, 0 )
exp {+iH0 } [W, exp {iH ( 0 )}] exp {iH0 0 }
[K0 , U (, 0 )] = exp {+iH0 } [exp {iH ( 0 )} , W] exp {iH0 0 }
[K0 , U (, 0 )] = exp {+iH0 } exp {iH ( 0 )} W exp {iH0 0 }
exp {+iH0 } W exp {iH ( 0 )} exp {iH0 0 }

inserting an identity we find

[K0 , U (, 0 )] = exp [+iH0 ] exp [iH ( 0 )] {exp [iH0 0 ] exp [iH0 0 ]} W exp [iH0 0 ]
exp [+iH0 ] W {exp [iH0 ] exp [iH0 ]} exp [iH ( 0 )] exp [iH0 0 ]
[K0 , U (, 0 )] = U (, 0 ) {exp [iH0 0 ] W exp [iH0 0 ]} {exp [+iH0 ] W exp [iH0 ]} U (, 0 )

and we obtain finally

[K0 , U (, 0 )] = U (, 0 ) W (0 ) W ( ) U (, 0 ) (2.91)
W (t) exp [iH0 t] W exp [iH0 t] (2.92)

now taking the limits and 0 + we have

[K0 , U (, +)] = U (, +) W (+) W () U (, +)


[K0 , S] = S W (+) W () S (2.93)

then for [K0 , S] to vanish, a sufficient condition is that

W (t ) 0 (2.94)

It occurs if the matrix elements of W between H0 eigenstates are sufficiently smooth functions of energy, then
matrix elements of W (t) between smooth superpositions of energy eigenstates vanish for t .
In conclusion, the condition (2.83) along with the condition W (t ) 0, are sufficient to obtain

[K0 , S] = 0

in particular the condition W (t ) 0 is satisfied if the matrix elements of W between H0 eigenstates are
sufficiently smooth functions of energy. If we add the condition (2.80) we have a set of sufficient conditions for
the Lorentz invariance of the Smatrix [i.e. to satisfy Eqs. (2.64)].
Note that the smoothness condition on W is a natural one, since it is like the condition on the matrix elements
of V that is necessary to make V (t) 0 when t . The latter is required for the very idea of the Smatrix,
since it is the condition to obtain asymptotic free states.
On the other hand, by using = 0 and 0 = in Eq. (2.91) we obtain

[K0 , U (0, )] = U (0, ) W () W (0) U (0, )


[K0 , U (0, )] = W (0) U (0, ) (2.95)
2.3. SYMMETRIES OF THE SMATRIX 81

from definitions (2.19, 2.39, 2.92) we have

U (0, 0 ) = exp [iH0 ] exp [iH0 0 ] = (0 )


W (0) = W

so that equation (2.95) becomes

[K0 , ()] = W () K0 () () K0 + W () = 0

(K0 + W) () () K0 = 0
finally
K () = () K0 (2.96)
(0)
let us recall that according with Eq. (2.19), () is the operator that converts free particle states into

the corresponding in or out states | i.
Further, by applying Eqs. (2.79, 2.80) we see that P0 and J0 commutes with both H and H0 . Therefore P0
and J0 commutes with (), but P = P0 and J = J0 so this commutativity could be expressed as

P () = () P0 ; J () = () J0
(0)
finally, since all and | i are eigenstates of H0 and H respectively with the same eigenvalue E , we obtain

H () = () H0

in summary we can write

G () = () G0 ; G K, P, J, H
1
() G () = G0 ; G K, P, J, H (2.97)

showing that with our assumptions in and out states transform under inhomogeneous Lorentz transformations
just like the free-particle states. In addition, since (2.97) are similarity transformations, the exact generators
K, P, J, H satisfy the same commutation relations as K0 , P0 , J0 , H0 . Observe that in deriving these results
we have used the commutation relations (2.72, 2.75, 2.77, 2.78), while the remaining ones (2.73, 2.74, 2.76) are
automatically obtained.

2.3.2 Internal symmetries


Up to now we have only consider space-time symmetries (continuous or discrete). Notwithstanding, there are
other symmetries like the symmetry in nuclear physics under interchange of neutrons and protons, or the charge
conjugation symmetry between particles and antiparticles. Such kind of symmetries have nothing directly to do
with Lorentz invariance, and also appear the same in all inertial frames. We shall denote a symmetry transfor-
mation of this type as T , and acting on the Hilbert space of Physical states as a unitary operator U (T ), which
induces linear transformations on the indices ni that label the particle species.
X
U (T ) |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i = Dn1 ,n1 (T ) Dn2 ,n2 (T ) |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i (2.98)
n1 ,n2 ,...

if T is some other transformation, the operators U (T ) and U T must satisfy the group multiplication rule
 
U T U (T ) = U T T (2.99)

where T T is the transformation


 obtained by first performing the transformation T and then the transformation
T . By applying U T on Eq. (2.98) we see that the matrices D (T ) must satisfy the same rule
 
D T D (T ) = D T T (2.100)
82 CHAPTER 2. SCATTERING THEORY

on the other hand, by taking the scalar product of the states obtained by applying U (T ) on two different in
states (or two different out states), along with the normalization condition (2.5) page 66, we see that D (T )
must be unitary
D (T ) = D 1 (T ) (2.101)
Finally, by taking the scalar product of the states obtained by applying U (T ) on one out state and one in
state shows that D commutes with the Smatrix, in the sense that
X X

DN ,n (T ) DN ,n (T ) DN1 ,n1 (T ) DN2 ,n2 (T ) Sp , ,N ;p , ,N ,p1 ,1 ,N1 ;p2 ,2 ,N2 ;
1 1 1 2 2 2
1 1 2 2
N1 ,N2 N1 ,N2
= Sp1 ,1 ,n1 ;p2 ,2 ,n2 ,p1 ,1 ,n1 ;p2 ,2 ,n2 ; (2.102)

once again this is what we mean by a theory to be invariant under the internal symmetry T , since to derive (2.102)
we still have to show that the same unitary operator U (T ) will induce the transformation (2.98) on both in
and out states. This will be the case if there is an unperturbed transformation operator U0 (T ) that induces
these transformations on free-particle states
X
U0 (T ) |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i = DN1 ,n1 (T ) DN2 ,n2 (T ) p1 , 1 , N1 ; p2 , 2 , N2 ; . . .
N1 ,N2 ,...

and that commutes with both the free-particle and interaction parts of the Hamiltonian

U01 (T ) H0 U0 (T ) = H0
U01 (T ) V U0 (T ) = V

by using the Lippmann-Schwinger equations (2.22) or Eqs. (2.19), we obtain that U0 (T ) will induce the transfor-
mations (2.98) on in and out states as well as free-particle states. Therefore, we can derive Eq. (2.98) taking
U (T ) as U0 (T ).
A particularly important case is that of a one-parameter Lie group in which T is a function of one parameter
and  
T T () = T +
as we saw in Eq. (1.39) page 17, in that case the corresponding operators on the Hilbert space must have the form

U (T ()) = exp (iQ) (2.103)

with Q a Hermitian operator. The corresponding matrices D (T ) take the form

Dmn (T ()) = mn exp (iqn ) (2.104)

where qn are a set of real numbers that depend on the species of particles. In this case, Eq. (2.102) says that the
q s are conserved, because S vanishes unless

qm1 + qm2 + . . . = qn1 + qn2 + . . . (2.105)

perhaps the most classical example of this kind of conservation law is that of conservation of electric charge.
Further, all known processes conserve baryon number (the number of baryons, such as protons, neutrons, and
hyperons, minus the number of their antiparticles), and most of processes seem to conserve the lepton number2
(the number of leptons, such as electrons, muons, taos, and neutrinos minus the number of their antiparticles).
However, it is generally believed that this conservation laws are only very good aproximations. There are other
conservation laws that are definitely only approximate such as the conservation of strangeness and isotopic spin.
2
The recent observations of neutrino oscillations seem to violate the lepton number.
2.3. SYMMETRIES OF THE SMATRIX 83

2.3.3 Parity
As customary, our starting point is the transformation of single particle states under the symmetry involved. They
are given by Eq. (1.266) for massive particles and by (1.285) for massless particles.

P |p, i = |Pp, i massive particles


P |p, i = exp [i] |Pp, i null mass particles

As long as the symmetry under the transformation x x is valid, there must exist a unitary operator P under
which both in and out states transform as a direct product of single particle states

P p1 , 1 , n1 ; p2 , 2 , n2 ; . . . = a1 a2 Pp1 , 1 1 , n1 ; Pp2 , 2 2 , n2 ; . . . (2.106)

where we have defined


 
ni if p2 = M 2 < 0 +1 if p2 = M 2 < 0
ai = ; i =
ni exp [ini ] if p2 = 0 1 if p2 = 0

where P is the operator on the Minkowski space that reverses the space components of p and ai is the intrinsic
parity. By denoting the in and out states as
+
= |p1 , 1 , n1 ; p2 , 2 , n2 ; . . .i ; = p1 , 1 , n1 ; p2 , 2 , n2 ; . . .

The parity conservation condition for the Smatrix reads

Sp1 1 n1 ;p2 2 n2 ; ,p11 n1 ;p22 n2 ; = a a a1 a2 SPp1,1 1 ,n1 ;Pp2 ,22 ,n2 ; ,Pp1,1 1 ,n1 ;Pp2,2 2 ,n2 ; (2.107)
1 2

as in the case of internal symmetries, an operator P satisfying Eq. (2.106) will actually exists if the operator P0
defined to act this way on free-particle states commute with V and with H0

P0 H0 P01 = H0 ; P0 V P01 = V

We shall consider the case of massive particles from now on. The phases n could be inferred from either dynamical
models or experiments. However, neither can provide a unique determination of the s. It is owe to the fact that
we can always redefine P by combining it with any conserved symmetry operator. As an example, If B and L are
conserved (recall that we believe these symmetries are only good approximations), and if P is also conserved, the
following operator is conserved as well

P P exp (iB + iL + iQ) (2.108)

where B, L, Q are baryon number, lepton number and electric charge respectively, and , , are arbitrary real
phases. In other words, if B and L are conserved, then P is conserved whenever P is conserved, and any of
them can be called the parity. The neutron, proton and electron have different combinations of values of B, L, Q
and by apropriate choices of the phases , , we can define the intrinsic parities of all three particles to be +1.
Protons have Bp = 1, Lp = 0, Qp = e, neutrons have Bn = 1, Ln = 0, Qn = 0, and electrons have Be = 0,
Le = 1, Qe = e. We can settle the intrinsic parity of all of them to be equal by imposing that the total phase
of the exponential in (2.108) be equal for all three particles, that is

Bk + Lk + Qk = N0 with k = p, n, e

with N0 being an apropriate constant. These equations fixed the phases , , . Consequently, Once we have
settle the intrinsic parity of p, n, e to be equal (say to the value +1), the intrinsic parity of other particles are
not arbitrary anymore. For example, the intrinsic parity of charged pion (which can be emitted in the transition
n p + ) is no longer arbitrary. In addition, the intrinsic parity of any particle like the neutral pion 0 which
84 CHAPTER 2. SCATTERING THEORY

carries no conserved quantum numbers (such as B, L, Q) is always meaningful, though in that case P and P in
Eq. (2.108) coincide. Note in particular that the combination of parity with other symmetries, decreases the
number of different intrinsic parities assigned to the particles.
The space inversion P has the group multiplication P 2 = 1, hence its eigenvalues are 1. Hence, it is natural
to ask whether the intrinsic parities must always have the values 1. The fact is that sometimes the parity
operator that is conserved is not the space inversion operator, and could differ from the latter by a certain phase
transformation. Regardless P 2 = 1 or not, the operator P 2 behaves just like an internal symmetry transformation:

P 2 p1 , 1 , n1 ; p2 , 2 , n2 ; . . . = a21 a22 P 2 p1 , 21 1 , n1 ; P 2 p2 , 22 2 , n2 ; . . .

P 2 p1 , 1 , n1 ; p2 , 2 , n2 ; . . . = a21 a22 p1 , 1 , n1 ; p2 , 2 , n2 ; . . .

If this internal symmetry is part of a continuous symmetry group of phase transformations, such as the group of
multiplication by the phases exp(iB + iL + iQ) with arbitrary values of , , , then its inverse square root
IP must also be a member of this group3
IP2 P 2 = 1
it is also quite obvious that
[IP , P ] = 0
 
for instance if P 2 = exp [iB + . . .], then we take IP = exp 12 iB + . . . . As a consequence, we can define a
new parity operator
P P IP P 2 = 1
and P is conserved with the same extent as P . Hence, there is no reason why we should not call this the parity
operator. In that case, since P 2 = 1, then the intrinsic parities can only take the values 1.
However, we could have a theory in which there is some discrete internal symmetry which is not a member
of any continuous symmetry group of phase transformations. In that case, it is not necessarily possible to define
parity in such a way that intrinsic parities can only take the values 1.
For instance, it is a consequence of angular-momentum conservation that the total number F of all particles
of half-integer spin can only change by even numbers4 , so the internal symmetry operator (1)F is conserved. On
the other hand, all known half-integer spin particles have odd values of the sum B + L. Let us take as examples
of half-integer spin particles the proton, neutron and electron, we have

Bp + L p = Bn + L n = 1 + 0 = 1 ; Be + L e = 0 + 1 = 1

and B + L is odd for all of them. Combining these two facts we observe that if F is odd since each Bi + Li is odd
the sum of Bi + Li over all F particles is also odd. Similarly, if F is even the total value of B + L is also even. As
a consequence, as far as we know we have
(1)B+L = (1)F
if this is true, the discrete symmetry (1)F is part of a continuous symmetry group of internal symmetries,
consisting of the operators exp [i (B + L)] with arbitrary real . It has an inverse square root exp [i (B + L) /2].
In this case, if P 2 = (1)F then P can be redefined such that all intrinsic parities are given by 1 as discussed
above. Let us take as examples of half-integer spin particles the proton, neutron and electron, we have

Bp + L p = Bn + L n = 1 + 0 = 1 ; Be + L e = 0 + 1 = 1

and B + L is odd for all of them.


3
If P is part of the group hence P 2 also is. Further, any element of the group must have an inverse. In particular P 2 has an inverse
that we denote as IP2 . Now since P 2 is in general different from the identity, it is followed that P is not necessarily its own inverse.
Hence IP is not necessarily P .
4
If an odd number of half-integer spin particles appear or dissapear the total change of spin is half-integer, and it cannot be
compensated by an opposite change in the orbital angular momentum because such a momentum only takes integer values.
2.3. SYMMETRIES OF THE SMATRIX 85

However, if we discovered a particle of half-integer spin, and even value of B + L, then it would be possible
to have P 2 = (1)F without being able to redefine the parity operator to have eigenvalues 1. However, in this
case we have P 4 = 1, so all particles would have intrinsic parities either 1, or i. This would be the case of
the so-called Majorana neutrinos with j = 1/2 and B + L = 0, in which we expect that intrinsic parities take the
values i.
From Eq. (2.107) we see that if the product of intrinsic parities in the final state is equal to the product of
intrinsic parities in the initial state, the Smatrix must be even overall in the three-momenta. If the product of
intrinsic parities in the final state is minus the product of intrinsic parities in the initial state, the Smatrix must
be odd overall in the three-momenta.
For instance, in 1951 it was observed that a pion can be absorbed by a deuteron (a deuteron is a bound state
of a proton and neutron with even orbital angular-momentum, chiefly l = 0) from the l = 0 ground state of the
d atom, the reaction reads
+ d n + n (2.109)
the initial state have total angular momentum j = 1, where the pion has spin 0 and the deuteron has spin 1.
Therefore, the final state must have orbital angular momentum l = 1 and total neutron spin s = 1. It is because
the other possibilities

l = 0, s = 1
l = 1, s = 0
l = 2, s=1

which are allowed by angular momentum conservation (compatible with the addition rules of angular momenta)5 ,
are forbidden by the requirement that the final state is antisymmetric in the two neutrons (two identical fermions).
The allowed final state has l = 1 which corresponds to the vector representation of SO (3) for the orbital
variables. Thus, the matrix element is odd under reversal of the direction of all three momenta. Consequently,
the Smatrix is odd overall in the three-momenta. In turn, it implies that the product of intrinsic parities in
the final state is minus the product of intrinsic parities in the initial state. We then conclude that the intrinsic
parities of the particles in the reaction (2.109) must be related as

d = n2

on the other hand, we have seen that we are able to choose the neutron and proton to have the same intrinsic
parity. Further, since a deuteron is a bound state of a proton and a neutron we have d = n2 obtaining

= 1

hence, the negative pion is a speudoscalar particle. Similar analyses show that + and 0 has also negative parity,
as expected from the isospin invariance of the three particles.

Since pions have negative intrinsic parity, a spin zero particle that decays into three pions must have intrinsic
parity 3 = (1)3 = 1. To see it, we observe that in the Lorentz frame in which the decaying particle is at rest
we have l = 0. Consequently, l = s = j = 0 and we are in the scalar representation of SO (3). Thus, rotational
invariance only allows the matrix elements to depend on scalars or pseudoscalars. As for the scalar quantities
they can only be formed by scalar products of the pion three-momenta with each other, and all of them are even
under reversal of all momenta. Pseudoscalars can be formed by products of the type

pi (pj pk )
5
For instance, l = 2 and s = 1 gives three possible values of total angular momentum j = 3, 2, 1. Conservation of angular momentum
leads us to j = 1. Similarly the (allowed) case gives l = s = 1, which yields j = 2, 1, 0, hence we take j = 1 because of the conservation
of angular momentum.
86 CHAPTER 2. SCATTERING THEORY

Note that, since we are in the proper frame in which p1 + p2 + p3 = 0 the triple scalar product p1 (p2 p3 )
formed from the three pion momenta vanishes

p1 = (p2 + p3 ) p1 (p2 p3 ) = (p2 + p3 ) (p3 p2 ) = p2 (p3 p2 ) + p3 (p3 p2 ) = 0 + 0 = 0

and similarly with the other triple products. Therefore, the Smatrix does not depend on pseudoscalar quantities.
We then conclude that the Smatrix must be even overall in the three-momenta. Consequently, the product of the
intrinsic parities 3 in the final states must equal the product of intrinsic parities of the initial states (consisting
in this case of a single particle).
For the same reason, a spin zero particle that decays into two pions must have intrinsic parity 2 = +1.
Among the strange particles discovered in the late 1940s there were two particles of zero spin (inferred from the
angular distribution of their decay products): The particle was identified by its decay into three pions, and
hence was assigned to it the intrinsic parity 1. On the other hand, the particle was identified by its decays
into two pions and was given a parity +1. After more detailed studies the and particle seemed increasingly to
have identical masses and lifetimes. After many attempts of solution to the so-called puzzle, Lee and Yang
suggested in 1956 that both are the same particle (currently called K ) and that parity is not conserved in the
weak interactions involved in these decays.
We shall see later that the rate for a physical process (with 6= ) is proportional to |S |2 , where the
proportionality factors are invariant under reversal of all three-momenta. Hence, as long as the states and
contains a definite number of particles of each type, the phase factors in Eq. (2.107) clearly do not play any role
in the quantity |S |2 . Consequently, equation (2.107) implies that the rate is invariant under the reversal
of direction of all three-momenta. We have seen that for the decay of a K meson into two or three pions, it is a
trivial consequence of the rotational invariance, but it is a non-trivial restriction on rates associated with more
complicated processes.
As an example, following the suggestions of Lee and Yang concerning parity violation in weak interactions,
Wu and her collaborators measured the angular distribution of the electron in the final state of the decay

Co60 N i60 + e + (2.110)

the momentum of the antineutrino and nickel nucleus were not measured. The experiment focused on the angular
distribution of the electrons, finding that they are preferably emitted in a direction opposite to that of the spin of
the decaying nucleus. Recalling that parity reverses all three-momenta but not the spin, we see that by applying
parity to this process we would obtain that the parity transformed state is one in which electrons are preferably
emitted in the direction of the spin of the decaying nucleus. Therefore the parity transformed state is a physically
different one (and not observed in the laboratory), showing that the interaction (weak interaction) that governs
this decay violates parity conservation. A similar result was found for the decay of a positive muon + which is
polarized in its production through the process + + + with the subsequent decay of + e+ .

+ + e+ (2.111)

The latter experimental observations show that for the processes (2.110, 2.111), Eq. (2.107) is untenable and
parity is not conserved in the weak interactions responsible for these decays. We shall see however that parity is
conserved in strong and electromagnetic interactions, so it has still an important role in Physics.

2.3.4 Time-reversal
We saw that the time-reversal operator T transforms one-particle states according with Eqs. (1.274) and (1.288)
for massive and masless particles respectively

T |p, i = (1)j |Pp, i for p2 = M 2 < 0


T |p, i = exp [i] |Pp, i for p2 = 0
2.3. SYMMETRIES OF THE SMATRIX 87

And the multi-particle state transforms as customary, as the direct product of one-particle states, except that
under time-reversal transformation in which we watch the film in the opposite sense, we expect in and out
states to be interchanged

T p1 , 1 , n1 ; p2 , 2 , n2 ; . . . = a1 a2 Pp1 , 1 1 , n1 ; Pp2 , 2 2 , n2 ; . . . (2.112)
where  
ni (1)ji i if p2 = M 2 < 0 +1 if p2 = M 2 < 0
ai = ; i = (2.113)
ni exp [ini ] if p2 = 0 1 if p2 = 0
we abbreviate the assumption (2.112) in the form

T = T (2.114)
where T indicates a reversal of the three-momenta, multiplication of the spins ni by i and multiplication of
the state by the phase factors ni defined in Eq. (2.113). Since T is antiunitary, we have

+

= T T +

+

= T + T
therefore, time-reversal invariance of the Smatrix is expressed by
S, = ST ,T (2.115)
or more explicitly
Sp1 1 n1 ;p2 2 n2 ; ,p1 1 n1 ;p2 2 n2 ; = a1 a2 a1 a2 SPp1 ,11 ,n1 ;Pp2 ,22 ,n2 ; ;Pp1,1 1 ,n1 ;Pp2,2 2 ,n2
where we notice once again that besides the reversal of momenta and the transformation of spins, the role of the
initial and final states is interchanged.
The Smatrix will satisfy this transformation rule if the operator T0 defined on free-particle states
E E

T0 (0) T (0) (2.116)

commutes with the free-particle Hamiltonian [which is automatic, see Eq. (1.256)] and also with the interaction
T01 H0 T0 = H0 ; T01 V T0 = V (2.117)
in that case we can take T = T0 and use either (2.19) or (2.21) to show that time-reversal transformation acts
as stated in Eq. (2.114). For instance, applying T on the Lippmann-Schwinger equation (2.21) and using Eqs.
(2.116, 2.117) we obtain
E

T = T (0) + T (E H0 i)1 V
E


T = T (0) + (E H0 i)1 V T

observe that the sign of i is reversed because of the antilinearity of T . We can rewrite both sides as
(0) E
T = T + (E H0 i)1 V T

which is clearly the Lippmann-Schwinger equation for the state |T i, thus justifying Eq. (2.114)6 . Similarly
applying T on the factor ( ) defined in Eq. (2.19) and taking into account the antilinearity of T and the fact
that T commutes with H0 and H we find
T ( ) = T {exp (+iH ) exp (iH0 )} = {exp (iH ) exp (iH0 )} T = ( ) T
6
In the free-particle states we have not incident or outgoing states. Thus, there is nothing like that to interchange under
time-reversal.
88 CHAPTER 2. SCATTERING THEORY

and applying T on both sides of Eq. (2.19)


E E

T = T () (0) = () T (0)
E
T
= () T (0)

which is the equation (2.19) for the state |T i, again leading to Eq. (2.114).
In contrast to the case of parity conservation, time-reversal invariance condition (2.115) does not in general
imply that the rate for the process is the same as for the process T T due to the fact that time-
reversal also interchanges the in and out states. However, something like this is true when the Smatrix can
be expressed as
(0) (1)
S = S + S (2.118)

where matrix elements S (1) are much smaller that the ones of S (0) except for some particular process of interest for
which the matrix element of S (0) vanishes. Sometimes S (0) is simply the unit operator. Calculating the unitarity
condition up to first order in S (1) we have
h i h i  
1 = S S = S (0) + S (1) S (0) + S (1) = S (0) S (0) + S (0) S (1) + S (1) S (0) + O S (1)2

and using the zeroth-order relation S (0) S (0) = 1 we have to first order in S (1)

1 = 1 + S (0) S (1) + S (1) S (0) S (0) S (1) + S (1) S (0) = 0


S (1) = S (0) S (1) S (0) (2.119)

which yields a reality condition for S (1) . If S (1) as well as S (0) satisfies the time-reversal condition (2.115), the
matrix multiplication in (2.119) can be put in the form
Z Z Z Z
(1) (0) (1) (0) (0) (1) (0)
S = d d S S S = d d S S S
Z Z
(1) (0) (1)
S = d d S ST ,T S (0)
(2.120)

and recalling that S (0) does not contribute to the particular process by hypothesis, we see that Eq. (2.120)
represents the Smatrix for such a process. Since S (0) is unitary, the rates for the processes and T T
are then the same if summed over sets I and F of final and initial states that are complete with respect to S (0) .
(0)
In this context, completeness of I with respect to S (0) means that if S is non-zero, and either or are in
I, then both states are in I. Hence, there is no transition from a state of the set I to a state outside of such a
set. Completeness for F is defined in the same way. When restricted to the subsets of states I and F, equation
(2.120) becomes Z Z
(1) (0) (1)
S = d d S ST ,T S
(0)
F I
In the simplest case, we have complete sets I and F consisting of just one state each; that is both the initial
and final states are eigenvectors of S (0) with eigenvalues e2i and ei2i respectively, where and are both
real since S (0) is unitary and so its spectrum lies on the unit complex circle. In this case Eq. (2.120) simplifies to
Z Z n
(1) 2i
o (1) n 2i o
(1)
S = d d e ST ,T e ( ) = e2i ST ,T e2i

= e2i( + ) ST ,T
(1) (1)
S (2.121)

where and are called phase shifts. In this scenario, it is clear that the absolute value of the matrix for the
process is the same as the one for the process T T . For instance, if all particles in both the initial
2.3. SYMMETRIES OF THE SMATRIX 89

and final states are massive, and satified the conditions above and time-reversal invariance, the differential rate
for the specific process remains unchanged if we reverse both the momenta and the spin three-component of all
particles.
A good example is the nuclear beta decay N N + e + , with S (0) the Smatrix produced by the strong
nuclear and electromagnetic interactions alone, and S (1) the correction to the Smatrix generated by the weak
interactions. The process is the same as the one for the process T T as long as we neglect the weak
Coulomb interaction between the electron and the nucleus N in the final state. For this beta decay, the initial
and final states are eigenstates of the strong interaction Smatrix with = = 0. Under the assumption of
time-reversal invariance, the differential rate for a beta decay is unaltered if we reverse both the momenta and
the third componente of the spin of all particles7 . This prediction was not contradicted by the experiment in
1957 that led to the violation of parity. Time-reversal invariance is compatible with the observation that electrons
from the decay
Co60 N i60 + e +

are emmitted preferentially in a direction opposite to that of the spin of Co60 . To see it, we observe that by
reversing all three momenta and spins, the transformed state is such that electrons are preferentially emmitted
in the opposite direction of the spin of Co60 . We shall see later that indirect evidence against time-reversal
invariance was discovered in 1964. However, such a symmetry remains being a good aproximation in weak, strong
and electromagnetic interactions.
It is sometimes possible to use a basis of states for which T = and T = . In that case8 , Eq. (2.121)
simplifies even more
S = e2i( + ) S
(1) (1)
, (2.122)

(1)
which just says that iS has the phase + mod . This statement is known as Watsons theorem.
The phases in Eqs. (2.121, 2.122) can be measured in processes with interference between different final states.
An example is the decay of the spin 1/2 hyperon into a nucleon and a pion. The final state can only have
orbital angular momentum l = 0 or l = 1, and the angular distribution of the pion relative to the spin involves
the interference between these states. Therefore, according with Watsons theorem this interference depends on
the difference s p of their phase shifts.

2.3.5 PT symmetry
The 1957 experiment did not rule out the conservation of time-reversal, but showed that P T was not conserved.
If conserved this operator must be antiunitary for the same reasons as T . Hence, in processes like beta decay P T
conservation would lead to relations similar to (2.121)

S = e2i( + ) SPT ,PT


(1) (1)
(2.123)

for massive particles, it is clear that PT reverses the sign of (spin three-component), but not the three momenta.
In the beta decay
Co60 N i60 + e +

neglecting the Coulomb interaction between N i60 and e in the final state, it would lead to no preference for the
electron to be emitted in the same or opposite direction to the Co60 spin, in contradiction with the observations.
7
All particles involved in this interaction are massive (including the neutrino).
8
Imagine for instance two identical particles of three component of spin zero, and opposite momenta in the initial state, and a final
state with similar features (though with different species of particles for the initial and final states).
90 CHAPTER 2. SCATTERING THEORY

2.3.6 Charge-conjugation C, CP and CPT


Charge conjugation is a special case of internal symmetry that interchanges particles and antiparticles. It implies
the existence of a unitary operator C, with an effect on multi-particle states described by

C p1 1 n1 ; p2 2 n2 ; . . . = n1 n2 p1 1 nc1 ; p2 2 nc2 ; . . . (2.124)

where nc is the antiparticle of the particle of the type n, where n is a phase related with this transformation. If
this is the rule of transformation for both the in and out states, the Smatrix will satisfy the condition

Sp1 1 n1 ,p2 2 n2 , ;p1 1 n1 ,p2 2 n2 , = n n n1 n2 Sp1 1 nc c c c


1 ,p2 2 n2 , ;p1 1 n1 ,p2 2 n2 ,
(2.125)
1 2

as in any other internal symmetry, condition (2.125) is satisfied if the operator C0 that acts as in (2.124) on
free-particle states, commutes with H0 and V

C 1 H0 C = H0 ; C 1 V C = V

and in that case we take C = C0 .


The phase n is called a charge-conjugation parity. Like in the case of intrinsic parities n , the n are in
general not uniquely defined, because for any C operator defined to satisfy Eq. (2.124) we can define another
such operator with different n by multiplying C by a phase transformation such as exp [iB + iL + iQ]. The
only particles whose charge conjugation parities can be measured individually, are those completely neutral like
the photon or the neutral pion (that is particles carrying no conserved quantum numbers), and they coincide with
their own antiparticles.
In reactions involving only completely neutral particles, Eq. (2.125) says that the product of initial charge-
conjugation parities must be the same as the product of final states charge-conjugation parities. As an example,
we shall see later that quantum electrodynamics requires that photon has charge-conjugation parity = 1, so
the observation of the process
0 2
requires that 0 = +1. In turn, it implies that the process 0 3 should be forbidden as effectively occurs
as far as we know. For and 0 we have real charge-conjugation parities either +1 or 1. As in the case
of intrinsic parity, the charge-conjugation parity can always be defined as 1, if all internal charge-conjugate
phase transformations are part of continuous groups of phase transformations. In that case we can redefine C by
multiplying by the inverse square root of the internal symmetry C 2

IC2 C 2 = 1 ; [IC , C] = 0

and the new operator C CIC , satisfies the condition C 2 = 1.


For general reactions, the satisfaction of condition (2.125) requires that the rate for a process equals the rate
for the same process replacing particles by their antiparticles. Once again this was not contradicted directly by
the results of the 1957 experiment, since the analogous experiment with the corresponding antiparticles was not
carried out. However, these experiments showed that C is not conserved in the theory of weak interactions as
proposed by Lee and Yang to take into account on parity non-conservation. Indeed, we shall see later that the
observed violation of P T imply a violation of C conservation in any field theory of weak interactions. At the
present state of the art, we know that C and P are not conserved in the weak interactions responsible of processes
such as the beta decay, and the decay of a pion and muon. Notwithstanding, C and P are conserved in the strong
and electromagnetic interactions.
Further, the violation of parity and charge-conjugation symmetries, keep the door open for the conservation
of the CP symmetry. It had important implications on the properties of neutral K mesons. In 1954 Gellmann
and Pais pointed out that since K 0 does not coincide with its antiparticle (K 0 carries a non-zero strangeness
2.4. RATES AND CROSS-SECTIONS 91

which is an approximate symmetry) the particles with definite decay rates would be not K 0 or K 0 but the linear
combinations
K0 K0
this was originally explained in terms of C conservation but once the C violation was stated, it was explained in
the framework of CP conservation. If we arbitrarily define the phases in the CP operator and in the K 0 and K 0
states as E E

CP K 0 = K 0 ; CP K 0 = K 0

we can define self-charge-conjugate one-particle states


0 h Ei 0 h Ei
K1 1 K 0 + K 0 ; K2 1 K 0 K 0
2 2
which have CP eigenvalues +1 and 1 respectively. The fastest decay mode of these particles is into two pions but
CP conservation would allow it for K10 but not for K20 . The neutral Kmesons have spin zero, so the two-pion
final state has l = 0 so that P = +1. On the other hand, C = +1 for 2 neutral pions because the pion has
C = +1 (in addition, C = +1 for a state + with l = 0, because C interchanges the two pions).
Therefore, the K20 state is expected to decay in slower modes, into three pions or a pion, muon or electron and
neutrino, thus enlarging its mean lifetime. However, Fitch and Cronin in 1964 found that the long-lived neutral
K20 meson, has a small probability for decaying into two pions. They conclude that CP is slightly violated in
weak interactions, but seems to be more nearly conserved than C or P individually.
We shall see later that although neither C or CP are exactly conserved, we expect that CP T is exactly
conserved in all interactions at least in any quantum field theory. It is CP T the operator that provides the
correspondence between particles and antiparticles, and the fact that CP T commutes with the Hamiltonian tells
us that stable particles and antiparticles have exactly the same mass.
Further, since CP T is antiunitary, it relates the Smatrix for an arbitrary process to the Smatrix for the
inverse process with all spin three-components inverted and particles replaced by antiparticles (but three-momenta
are unchanged).
In cases in which the Smatrix can be divided as in Eq. (2.118) into a weak term S (1) that produces a given
reaction and a strong term S (0) that acts in the initial and final states, we can use an argument similar to the
one followed in the T conservation scenario to show that the rate of any process is equal to the rate of the same
process with particles replaced by antiparticles and spin three-components reversed, as long as we sum over sets
of initial and final states that are complete with respect to S (0) . In particular, despite the partial rates for decay
(0)
of the particle into a pair of final states 1 and 2 with S1 2 6= 0 may differ from the partial rates for the decay
of the antiparticle into the corresponding final states CPT 1 and CPT 2 , we shall see that the total decay rate
of any particle is exactly equal to the total decay rate of its antiparticle.
From the discussion above, we can understand that the 1957 experiment shows strong C, P and P T violation
but not CP violation. It is because such an experiment is consistent with T conservation and any quantum field
theory must preserve CP T . Thus such an experiment is consistent with CP conservation.
Similarly, the observation in 1964 of small CP violation in the weak interactions9 along with the CP T con-
servation assumption, leads to an indirect evidence of a tiny violation of time-reversal symmetry. This has been
verified by a more detailed studies of the K 0 K 0 system.

2.4 Rates and cross-sections


The Smatrix with elements S provides the amplitude of probability associated with the process .
However, we still have to relate such an amplitude with observables measured by experimentalists. Observables
are in some way related with the probability density given by |S |2 . On the other hand, Eq. (2.61) shows that
9
More recently, evidence of CP violation has been found in a B B system, where B is another neutral meson.
92 CHAPTER 2. SCATTERING THEORY

S contains a factor 4 (p p ) that ensures the conservation of three-momentum and energy. Hence, we should
 2
manage properly the factor 4 (p p ) in calculations.
From a fundamental point of view, we should manage with wave packets that represent particles localized far
from each other before the collision (which is the way the experiments are prepared), and then characterizing the
time evolution of these superpositions of multi-particle states. Our derivation will be however quite practical.
We consider that our whole system of physical particles is enclosed in a macroscopic volume V . This box
could be taken as a cube, but with points on opposite sides identified such that the single-valuedness of the spatial
wave-function requires the momenta to be quantized10
2
p= (n1 , n2 , n3 ) (2.126)
L
where ni are integers, and L3 = V . In that case, all three dimensional delta functions are redefined by an
integration over a bounded volume (instead of the whole space) of the same integrand that defines the original
delta function Z
3
 1
V p p 3 d3 x ei(pp )x (2.127)
(2) V
therefore, the new bounded delta function becomes discrete. We observe that if p = p, the RHS of Eq. (2.127)
becomes the (bounded) macroscopic volume of the box, hence
Z
3
 1 V
V p p 3 d3 x ei(pp )x = p,p (2.128)
(2) V (2)3

where p,p is an ordinary (discrete!) Kronecker delta. Moreover, substituting the Dirac delta functions in the
normalization condition (2.5) by the bounded Dirac delta functions in (2.128), we obtain that the states involved
in such a condition have scalar products in a box that are not only sums of products of Kronecker deltas, but also
h i h iN
contains a factor V / (2)3 for each particle in the state. Thus we also have a factor V / (2)3 , where N is
the number of particles in the state. Calculations of transition probabilities should use states of unit norm. We
shall introduce states that are normalized approximately for our box
v
u" #
Box u 3 N
t (2)
|i (2.129)
V

whose norm is given by



Box Box =
where is a product of Kronecker deltas, one for each three-momentum, spin, and species label, plus terms with
particles permuted as in the normalization condition (2.5). Thus, the Smatrix associated with states (2.129)
normalized in the box yields
s  N s   N D

+ V V E
Box(+) Box()
S = =
(2)3 (2)3
  (N +N )

+ V 2
S = = 3
Box
S (2.130)
(2)
10
It is a prominent feature on quantum mechanics that bounded systems posseses a discrete spectrum of momenta. Such a feature
depends on the particle-wave nature of physical systems, so we can extrapolate it to the case of relativistic quantum mechanics. For
macroscopic volumes a huge quantity of allowed discrete values is expected, and only in the limit in which the box becomes unbounded
in all its dimensions, we obtain a totally continuous spectrum of momenta.
2.4. RATES AND CROSS-SECTIONS 93

where S Box is calculated from states (2.129).

If we left the particles in the box forever, then every possible transition would occur again and again. A
meaningful transition probability occurs when we also put our system in a time box. That is, we assume that
the interaction is turned on for a time T . It is conventional to define that the particles arrive at the region
of collision at t 0. Thus, we define the box of time in which the interaction occurs as [T /2, T /2]. As a
consequence, the energy conservation delta function is replaced with11
Z T /2
1
T (E E ) = exp [i (E E ) t] dt (2.131)
2 T /2
If the multi-particle system is in the state before the interaction is turned on, the probability of finding it in
the state after the interaction is turned off, reads
" #
Box 2 3 (N +N )
(2)
P ( ) = S = |S |2 (2.132)
V

where we have used (2.130). It is the probability of transition into one specific box state . The number of
one-particle box states within a momentum space volume d3 p is the number of triplets of integers n1 , n2 , n3 for
which the momentum (2.126) lies within the volume d3 p centered at p. On the other hand, Eq. (2.128) says that
V / (2)3 is the density of such states in the three-momentum space. Therefore, the number of one-particle box
states within a momentum space volume d3 p is given by
d3 p
V
(2)3
we shall define the final state interval d as a product of d3 p for each final particle, hence the total number of
states in such a range is given by
  N
V
dN = d (2.133)
(2)3
and the total probability for the system initially in the |i state, to lie within a range d of final states is obtained
by combining Eqs. (2.132, 2.133) and gives
" #(N +N )   N
(2)3 2 V
dP ( ) = P ( ) dN = |S | d
V (2)3
" #N
(2)3
dP ( ) = P ( ) dN = |S |2 d (2.134)
V

In this section, we shall restrict our attention to final states that are different from the initial states , but that
also satisfy the condition that no subset of the particles in the state (other than the whole system itself) has
precisely the same four-momentum as some corresponding subset of the particles in the state . For such states,
we can define a delta-free matrix element M as follows
S 2iV3 (p p ) T (E E ) M (2.135)
the introduction of the box permits to interpret the squares of the delta functions in |S |2 for 6= . Since V3 (p p )
is zero for p 6= p we can write this square as
 3 2
V (p p ) = V3 (p p ) V3 (0)
[T (E E )]2 = T (E E ) T (0)
11
Observe that (2.128) along with (2.131) can be interpreted as the bounded Dirac delta function in the four-momentum, defined
in a bounded space-time box.
94 CHAPTER 2. SCATTERING THEORY

and V3 (0), T (0) can be obtained by using p = p and E = E in Eqs. (2.128, 2.131). We obtain finally
 2 V
V3 (p p ) = V3 (p p ) V3 (0) = V3 (p p ) (2.136)
(2)3
T
[T (E E )]2 = T (E E ) T (0) = T (E E ) (2.137)
2
from Eqs. (2.135, 2.136, 2.137) the differential of transition probability (2.134) becomes
" #N " #
3 N
(2)3 (2)
2iV3 (p p ) T (E E ) M 2 d
dP ( ) = |S |2 d =
V V
" #N
3  3 2
2 (2)
= (2) V (p p ) [T (E E )]2 | M |2 d
V
" #
3 N   
2 (2) V T
dP ( ) = (2) 3
V (p p ) 3 T (E E ) | M |2 d
V (2) 2

and we obtain finally


" #N 1  
2 (2)3 T
dP ( ) = (2) |M |2 V3 (p p ) T (E E ) d (2.138)
V 2

and letting V and T to be very large, the delta function product becomes an ordinary four-dimensional delta
function 4 (p p ). In this limit the transition probability is proportional to the time T during which the
interaction is acting, with a coefficient that can be interpreted as a differential transition rate (differential of
probability per unit time)
dP ( )
d ( ) = (2)3N 2 V 1N |M |2 4 (p p ) d (2.139)
T
where the structure (2.135) of S in this limit becomes
S 2i4 (p p ) M (2.140)
this is the master formula to connect the Smatrix elements with experimental measurements. There are two
cases of particular importance

2.4.1 One-particle initial states


When the initial state consists of a one-particle state then N = 1, and Eq. (2.139) becomes independent of the
volume V . It gives the transition rate for a single-particle state to decay into a general multi-particle state
d ( ) = 2 |M |2 4 (p p ) d (2.141)
however, we should take into account that many particles are unstable so that they decay spontaneously (that is,
they decay even in the absence of interaction). For a given (very big) sample of unstable particles, the number of
particles as a function of time is approximately given by a model like
N = N0 et
where N0 is the initial number of particles. From this model we can define a characteristic time of decay (mean
lifetime of the particle) as the time in which the number of particles in the sample has decayed by a factor of 1/e.
Thus it is clear that the mean lifetime of the particle described by the state is given by
= 1

2.4. RATES AND CROSS-SECTIONS 95

and is called the spontaneous decay width of the particle. In this case, we are characterizing not the spontaneous
decay but the decay induced by an interaction. Consequently, Eq. (2.141), is only valid if we can neglect the
spontaneous decay rate. In other words, such an equation only makes sense if the time T during which the
interaction acts is much less than the mean lifetime of the particle . However, it leads to the problem that
this condition could prevents us to pass to the limit T in T (E E ). There is an unremovable width

1 1
E &
T

in this delta function, such that Eq. (2.141) is only useful if the total decay rate 1 is much less than any of the
characteristic energies of the process.

2.4.2 Two-particles initial states


When N = 2, Eq. (2.139) is proportional to 1/V . In other words, it is proportional to the density of either
particle at the position of the other one. In experiments it is usually reported the transition rate per flux (instead
of the transition rate per density), also known as the cross section. The flux of either particle at the position of
the other one is defined as the product of the density 1/V and the relative velocity u
u
= (2.142)
V
where we define (by now) u as the velocity of one particle if the other is at rest. Therefore, the differential
cross-section is given by

d ( ) (2)4 1 u
u V |M |2 4 (p p ) d
d ( ) =
u /V
4
d ( ) (2)
d ( ) = |M |2 4 (p p ) d (2.143)
u

2.4.3 Multi-particle initial states


The cases N = 1, 2 are the most important, but transition rates with N 3 are also observable, and some of
them are important in branches such as Particle Physics, Chemistry, Astrophysics etc. As an example, in the
main reactions that release energy from the sun, two protons and an electron turn into a deuteron and a neutrino.
We shall see later some applications of the master transition rate formula (2.139), to the case of arbitrary number
of initial particle states.

2.4.4 Lorentz transformations of rates and cross-sections


The Lorentz transformation rule (2.60) for the Smatrix is complicated by the momentum-dependent matrices
associated with each particles spin. We can avoid such a complication by squaring the absolute value of (2.60)
after factoring out the Lorentz-invariant delta function in Eq. (2.140), and then sum over all spins. The unitarity
of the matrices defined in Eq. (2.4), page 66, shows that apart from the energy factors in (2.60), the sum is
Lorentz-invariant. It means that the quantity
X Y Y
|M |2 E E R (2.144)
spins
Q
is a scalar function of the four-momenta of the particles in states and . By E we mean the product of all

p
single-particle energies p0 = p2 + m2 for the particles in the state .
96 CHAPTER 2. SCATTERING THEORY

We can then write the spin-summed single-particle decay rate (2.141) in the form
P Q Q
spins |M |2 E E
X X
d ( ) = 2 |M |2 4 (p p ) d = 24 (p p ) d Q Q
E E
spins spins

taking into account that there is a single particle in the state we have
X R
d ( ) = 24 (p p ) d Q
E E
spins

X 2E1 R 4 (p p ) d
d ( ) = Q
E
spins

Q
the factor d/ E can be recognized as the product of the Lorentz-invariant momentum-space volume elements

defined in (1.157)12 . Since this product is Lorentz invariant, R and 4 (p p ) also are. The only non-invariant
factor is 1/E , where E is the energy of the single-particle initial state. We have found that the decay rate has
the same Lorentz transformation property of 1/E . This is a manifestation of the timedilation property of special
relativity, the faster the particle the slower it decays.
In the same way, after summing the spins, the cross-section (2.143) yields
P Q Q
spins |M |2 E E
X 4
(2) 4
d ( ) = (p p ) d Q Q
u E E
spins

but in this case there are two particles in the initial state , therefore
P 2Q Q
X 4 spins |M | E E
(2) 4
d ( ) = (p p ) d Q
u E1 E2 E
spins

X 4
4
(2) R (p p ) d
d ( ) = Q
u E1 E2 E
spins

where E1 and E2 are the energies of the two particles in the initial state . It is conventional to define the
cross-section (when summed over spins) to be a Lorentz invariant function of the four-momenta. Further, the
factors
d
R ; Q ; 4 (p p )
E

are already Lorentz invariant. Therefore, we must define the relative velocity u such that u E1 E2 is a scalar
(Lorentz invariant) function of four-momenta. We have said that in the Lorentz frame in which one of the particles
(say particle 1) is at rest, u is the velocity of the other particle. From the previous facts, the quantity u is
uniquely determined in an arbitrary Lorentz frame
q
(p1 p2 )2 m21 m22
u = (2.145)
E1 E2
12
Note that we are identifying d with d3 p. This is another way to say that the orbital variables are the only ones that can be
continuous.
2.4. RATES AND CROSS-SECTIONS 97

where p1 , p2 and m1 , m2 are the four-momenta and proper mass of the two particles in the state .
From Eq. (2.145) it is obvious that E1 E2 u is a Lorentz invariant function of the four-momenta (recall that
the contraction of two four-vectors is a scalar). Moreover, when the particle 1 is at rest, we have

p1 = (p1 , E1 ) = (0, m1 ) ; p2 = (p2 , E2 )


p1 p2 = p1 p2 = (p1 , E1 ) (p2 , E2 ) = m1 E2 (2.146)

so that in this frame, equation (2.145) gives


q q
p
(p1 p2 )2 m21 m22 (m1 E2 )2 m21 m22 m1 E22 m22
u = = =
E1 E2 E1 E2 m1 E2
p
2
E2 m2 2
u = (2.147)
E2
further, we also have
p22 = (p2 , E2 ) (p2 , E2 ) = |p2 |2 E22
and p22 = m22 , then

|p2 |2 E22 = m22


|p2 |2 = E22 m22 (2.148)

substituting (2.148) in (2.147) we obtain


|p2 |
u =
E2
which is precisely the velocity of particle 2, as required. Similarly, u becomes the velocity of particle 1 in the
reference frame in which the particle 2 is at rest13 .
It is also useful to see the form of u in the center of mass frame, in which the total three-momentum is
null. Hence, in such a frame we have

p1 = (p, E1 ) , p2 = (p, E2 ) (2.149)


2
p1 p2 = p E1 E2 (2.150)

on the other hand, we also have

E12 = p2 + m21 ; E22 = p2 + m22 (2.151)


m21 = E12 p 2
; m22 = E22 2
p (2.152)

using (2.150) and (2.152) we find

2  
(p1 p2 )2 m21 m22 = p2 E1 E2 E12 p2 E22 p2
= p4 + E12 E22 + 2E1 E2 p2 E12 E22 + E12 p2 + p2 E22 p4

(p1 p2 )2 m21 m22 = 2E1 E2 p2 + E12 p2 + p2 E22 = p2 2E1 E2 + E12 + E22
(p1 p2 )2 m21 m22 = p2 (E1 + E2 )2 (2.153)

and applying (2.153) in Eq. (2.145), we obtain



|p| (E1 + E2 ) p1 p2
u = = (2.154)
E1 E2 E1 E2
13
It is important to say that u has nothing to do with the four-velocity usually defined in special relativity.
98 CHAPTER 2. SCATTERING THEORY

as could be expected from a relative velocity. We should say however that in this frame u does not correspond
to a physical velocity. In particular, if we apply Eq. (2.154) to two ultrarelativistic particles we have14

p p

u = + 2
E1 E2
such that u = 2 which is greater than c (recall that c = 1, so that velocities are dimensionless in natural units).

2.5 Physical interpretation of the Diracs phase space factor 4 (p p ) d


The phase-space factor 4 (p p ) d appears in the general formula (2.139) and of course in the particular
cases (2.141, 2.143) for decay rates and cross-sections. We shall study the case in which we are in the center of
mass reference frame, where the total three-momentum of the initial state vanishes

p = 0

In particular if N = 1, it corresponds to the reference frame in which the initial particle is at rest. Returning
to the general case, if the thrre momenta of the final states are denoted by p1 , p2 , p3 , . . ., the phase-space factor
then becomes
  3 3 3
4 (p p ) d = 3 p1 + p2 + . . . 0 p0 0
1 + p2 + . . . p
0
d p d p d p3
4 3

 3 3 1 3 2
(p p ) d = p1 + p2 + . . . E1 + E2 + . . . E d p1 d p2 d p3 (2.155)

where E E is the total energy of the initial state. We can perform the integration over each pk , (say p1 ) just
by dropping the momentum delta function

4 (p p ) d E1 + E2 + . . . E d3 p2 d2 p3 (2.156)

with the undestanding that whenever p1 appears (as in E1 ), it must be replaced with

p1 = p2 p3 . . . (2.157)

similarly, we can use the remaining delta function to eliminate any one of the remaining integrals.

2.5.1 The case of N = 2


Let us now study the special case in which there are two particles in the final state . In that case Eq. (2.155)
yields 
4 (p p ) d E1 + E2 E d3 p1 d3 p2
after integrating over p1 we can write the delta function as in (2.156)

4 (p p ) d E1 + E2 E d3 p2 (2.158)

as long as we take into account that condition (2.157) must be satisfied, that is

p2 = p1 (2.159)

from Eqs. (2.151) we can write the phase-space form factor (2.158) as
q q 

4 (p p ) d p2 + m 2 + p2 + m2 E p2
1 1 1 2 1 d p1 d (2.160)

14
Recall that an ultrarelativistic particle is the one in which the kinetic energy is much greater than the self-energy of the particle.
Therefore, in such a limit we have E 2 p2 . In our case, if both particles are ultrarelativistic we have E12 E22 p2 .
2.5. PHYSICAL INTERPRETATION OF THE DIRACS PHASE SPACE FACTOR 4 (P P ) D 99

where d = sin d d is the solid angle differential for p1 . Equation (2.160) can be simplified by using the
property
(x x0 )
(f (x)) = (2.161)
|f (x0 )|
where f (x) is an arbitrary real function with a single simple zero at x = x0 . From Eq. (2.160) our function within
the Diracs delta function is given by
 q q
f p1 |p1 | + m1 + |p1 |2 + m2
2 2
2 E

it is convenient to redefine q q 2
f (x) x + m2
1 + x + m2
2 E ; x = p1

the argument y in which f (x) vanishes, i.e. f (y) = 0, yields


q q
y + m2
1 + y + m2
2 =E

we can obtain y by squaring two times as follows

  q  
y + m2
1 + y + m 2
2 + 2 y + m2
1 y + m2 2 = E2
q  
2y + m2
1 + m 2
2 E 2
= 2 y + m2
1 y + m2
2

2 2
 
2y + m2 2
1 + m2 E = 4 y + m2
1 y + m2
2

simplifying these equations we find


2  
2y + m2 2
1 + m2 E
2
4 y + m2
1 y + m2
2 = 0
E 4
2E m2
2 2 2 2 2 2
2 2E m1 4E y 2m1 m2 + m1 + m2
4 4
= 0

4E y + 2
m4
1 + m4 2 2 4 2 2 2 2
2 + 2m1 m2 + E 2E m1 2E m2 4m1 m2
2 2
= 0

2 2
4E 2 y + E 2 m2
1 m2 4m2 2
1 m2 = 0

from which the root gives



2 2 4m2 m2
2
E 2 m2
1 m2 1 2
y = p1 = 2
4E

since |p1 | is positive there is a simple zero at |p1 | = y k
q 2
E 2 m2 2
1 m2 4m2 2
1 m2
k =
2E
in addition
s
q q 
2 2 4m2 m2
E 2 m2
1 m2 1 2 4E 2 m2
E1 = p2
+1 m2
=
1 + k2
= m2
1 + 1
4E 2 4E 2
q 2 q 
2 2 2 2 2
E m1 m2 4m1 m2 + 4E m1 2 2 E 2 m2 2 2
2 + m1
E1 = =
 2E 2E
2 2
E m2 + m1 2
E1 =
2E
100 CHAPTER 2. SCATTERING THEORY

and similarly for E2 . Now the derivative evaluated at the simple zero gives

 q q  | |
 d |p |p
f k = |p1 |2 + m21 + |p1 |2 + m22 E = q 1
+q 1
d |p1 | |p1 |=k
2 2 2 2
|p1 | + m1 |p1 | + m2
|p1 |=k
k k
= p + p
k2 + m21 k2 + m22
 k k k (E1 + E2 ) k E
f k = + = =
E1 E2 E1 E2 E1 E2

In summary, the argument E1 + E2 E of the delta function in (2.160) has a unique zero at |p1 | = k , where
q 
E 2 m2
1 m2
2 4m2
1 m2
2

k = (2.162)
2E
q
E 2 m2
2 + m1
2
E1 = k2 + m2
1 = (2.163)
2E
q
E 2 m2
1 + m2
2
E2 = k2 + m2
2 = (2.164)
2E
with derivative
 q q 
 d k k k E

f k
= |p1 |2 + m21 + |p1 |2 + m22 E = + = (2.165)
d |p1 | |p1 |=k E1 E2 E1 E2

Therefore applying property (2.161) to our specific case Eq. (2.160) yields
q q 

(|p1 | k ) 2
4
(p p ) d p1 + m1 + p1 + m2 E p2
2 2 2 2
1 d p1 d = p1 d p1 d
|f (k )|
E1 E2 (|p1 | k ) 2
4 (p p ) d k d p1 d (2.166)
k E
Therefore, after performing the integration over d |p1 |, we drop the delta function and the differential d |p1 |
and replace |p1 | by k in the remaining terms on the RHS of Eq. (2.166)

k E1 E2
4 (p p ) d d (2.167)
E
and understanding that k , E1 and E2 are given everywhere by (2.162, 2.163, 2.164).
For the particular case N = 1, the differential decay rate (2.141) that describes the decay of one particle at
rest (zero three-momentum) and energy E into two particles is

d ( ) 2k E1 E2
= |M |2 (2.168)
d E
and the differential cross-section for a two-body scattering process 1 2 1 2 is given by Eq. (2.143) and yields

d ( ) (2)4 k E1 E2
= |M |2
d Eu
and using Eq. (2.154) for the relative velocity we find

d ( ) (2)4 k E1 E2 E1 E2
= |M |2
d E |p1 | (E1 + E2 )
2.5. PHYSICAL INTERPRETATION OF THE DIRACS PHASE SPACE FACTOR 4 (P P ) D 101

hence, we obtain finally

d ( ) (2)4 k E1 E2 (2)4 k E1 E2 E1 E2
= |M |2 = |M |2
d Eu E2k
k |p1 | = |p2 | . (2.169)

recall that all these formulas have been obtained in the CM reference frame.

2.5.2 The case with N = 3 and Dalitz plots


There are still many cases of interest in the case of N = 3 as it is the case of decyas of a single intitial particle
into three bodies. For the case with N = 3 equation (2.156) yields
q q q 
4
(p p ) d d3 p2 d3 p3 (p2 + p3 )2 + m2
1 + p2
2 + m2
2 + p2
3 + m2
3 E (2.170)

the momentum-space volume in spherical coordinates is given by


2
d3 p2 d3 p3 = p2
2 d p2 p3 d p3 d3 d23 d cos 23

where d3 is the differential element of solid angle for p3 and 23 , 23 are the polar and azimuthal angles of p2
reltive to the p3 direction. The orientation of the plane spanned by p2 and p3 is specified by 23 and the direction
of p3 , and the remaining angle 23 is fixed by the energy conservation condition
q q q
p2 2 2
2 + 2 |p2 | |p3 | cos 23 + p3 + m1 + p2 2
2 + m2 + p2 2
3 + m3 = E

the derivative of the argument of the delta function with respect to cos 23 is

E1 |p | |p |
= 2 3
cos 23 E1

so the integral over cos 23 can be done just by dropping the delta function and dividing by this derivative


4 (p p ) d p2 d p2 p3 d p3 E1 d3 d23

replacing mometa with energies , we obtain finally

4 (p p ) d E1 E2 E3 dE2 dE3 d3 d23 (2.171)

now we recall that the expression (2.144) obtained by summing |M |2 over spins and multiplying with the product
of energies. Such an expression is a scalar function of four-momenta. If we approximate this scalar as a constant,
Eq. (2.171) tells us that for a given initial state, the distribution of events plotted in the E2 , E3 plane is uniform.
Any deviation from this uniform distribution of events is a useful clue to the dynamics of the decay process,
including possible centrifugal barriers or resonant intermediate states. This are known as Dalitz plots, since it
was used by Dalitz in 1953 to analize the decay

K + + +
102 CHAPTER 2. SCATTERING THEORY

2.6 Perturbation theory


One of the most useful technique to calculate the Smatrix is perturbation theory. It consists of obtaining an
expansion in powers of the interaction term V when the Hamiltonian is given by H = H0 + V . We start with the
expressions (2.44) and (2.22) of the Smatrix
D

+
S = ( ) 2i (E E ) T +
; T = (0) V + (2.172)

where |+ i satisfies the Lippmann-Schwinger equations (2.22)


Z
+ (0) E +
T
(0)
E
= + d (2.173)
(E E + i)
+
from Eq. (2.172) it is clear that all non-trivial contents of the Smatrix is concentrated on the factor T . To

(0)
find an approximate value of such a factor, we apply V on both sides of Eq. (2.173), and we obtain
D D E Z + D E
T
(0) V + = (0) V (0) + d (0) V (0)
(E E + i)
Z + D E
+ T V
T = V + d ; V (0) V (0) (2.174)
(E E + i)
which is an integral equation for T + . The perturbation series is obtained by iterating Eq. (2.174). The first
iteration gives
Z Z  Z 
+ V1 + V1 V1 2 +
T = V + d1 T = V + d1 V1 + d2 T
(E E1 + i) 1 (E E1 + i) (E E2 + i) 2
Z Z Z
+ V1 V1 V1 V1 2
T = V + d1 + d1 d2 T+ (2.175)
(E E1 + i) (E E1 + i) (E E2 + i) 2
the second iteration yields
Z
+ V1 V1
T = V + d1
(E E1 + i)
Z Z " Z #
V1 V1 2 T+3 V2 3
+ d1 d2 V2 + d3
(E E1 + i) (E E2 + i) (E E3 + i)
Z Z Z
+ V1 V1 V1 V1 2 V2
T = V + d1 + d1 d2
(E E1 + i) (E E1 + i) (E E2 + i)
Z Z Z
V1 V1 2 V2 3
+ d1 d2 d3 T+ (2.176)
(E E1 + i) (E E2 + i) (E E3 + i) 3
Eqs. (2.175, 2.176), show that the RHS depends in turn on T + . However, it can also be observed that each term
(from left to right) increases the power of V and the last term on the RHS of these equations (which is the only
one that contains T + ) is of higher order (in powers of V ) than the other ones15 . Thus if V is sufficiently small,
higher order terms are smaller. Then, these expressions are calculable if we drop the latter term on the RHS on
either of these equations. If we drop the last term on the RHS of Eq. (2.175), we are calculating at second order
in V , while if we drop the last term in Eq. (2.176) we are calculating at third order in V . The expansion at third
order in V obtained from (2.176) reads
Z Z Z
+ V1 V1 V1 V1 2 V2
T = V + d1 + d1 d2 + ... (2.177)
(E E1 + i) (E E1 + i) (E E2 + i)
15
We should take into account that T + depends in turn on V according with the definition (2.172), thus the integral that contains
+
T is of order V n+1 in the potential, where n is the number of integrals in that term.
2.6. PERTURBATION THEORY 103

The method of calculation based on Eqs. (2.175, 2.176) is called old-fashioned perturbation theory. The
presence of the energy in the denominators obscure the underlying Lorentz invariance of the Smatrix. Notwith-
standing, it is still useful in clarifying the way that singularities of the Smatrix arise from several intermediate
states. We shall work mostly on a rewritten version of Eqs. (2.175, 2.176) known as time-dependent pertur-
bation theory, in which Lorentz invariance is more apparent, though it somewhat obscures the contribution of
individual intermediate states.
The simplest way to derive the time-ordered perturbation expansion is by starting with the Soperator defined
by Eqs. (2.38, 2.39)

S = U (, ) ; U (, 0 ) exp [iH0 ] exp [iH ( 0 )] exp [iH0 0 ] (2.178)

Differentiating this formula for U (, 0 ) with respect to gives a differential equation for U (, 0 )

d d
U (, 0 ) = {exp [iH0 ] exp [iH ( 0 )] exp [iH0 0 ]}
d d
= {exp [iH0 ]} (iH0 ) {exp [iH ( 0 )] exp [iH0 0 ]}
+ {exp [iH0 ]} (iH) {exp [iH ( 0 )] exp [iH0 0 ]}
d
U (, 0 ) = i {exp [iH0 ]} (H0 H) {exp [iH ( 0 )] exp [iH0 0 ]}
d
and inserting an identity operator we find

d
U (, 0 ) = i {exp [iH0 ]} V {exp [iH0 ] exp [+iH0 ]} {exp [iH ( 0 )] exp [iH0 0 ]}
d
d
U (, 0 ) = i {exp [iH0 ] V exp [iH0 ]} U (, 0 )
d
so we obtain finally

d
i U (, 0 ) = V ( ) U (, 0 ) (2.179)
d
V (t) exp [iH0 ] V exp [iH0 ] (2.180)

observe that even if V is time-independent the operator V (t) is time-dependent. Note that V (t) is the operator
associated with V through the interaction picture (not the Heisenberg picture)16 . Equation (2.179) along with
the initial condition
U (0 , 0 ) = 1 (2.181)
are satisfied by the following solution of the integral equation
Z
U (, 0 ) = 1 i dt V (t) U (t, 0 ) (2.182)
0

once again, we obtain an expansion of U (, 0 ) in powers of V by iteration of Eq. (2.182). The first iteration gives
Z Z  Z t1 
U (, 0 ) = 1 i dt1 V (t1 ) U (t1 , 0 ) = 1 i dt1 V (t1 ) 1i dt2 V (t2 ) U (t2 , 0 )
0 0 0
Z Z Z t1
U (, 0 ) = 1 i dt1 V (t1 ) + (i)2 dt1 dt2 V (t1 ) V (t2 ) U (t2 , 0 )
0 0 0

16
Remember that in the Heisenberg picture the operators are transformed in the same way as in equation (2.180), but with the
whole Hamiltonian H instead of H0 .
104 CHAPTER 2. SCATTERING THEORY

further iterations are given by


Z Z Z t1  Z t2 
2
U (, 0 ) = 1 i dt1 V (t1 ) + (i) dt1 dt2 V (t1 ) V (t2 ) 1i dt3 V (t3 ) U (t3 , 0 )
0 0 0 0
Z Z Z t1
U (, 0 ) = 1 i dt1 V (t1 ) + (i)2 dt1 dt2 V (t1 ) V (t2 )
0 0 0
Z Z t1 Z t2
+ (i)3 dt1 dt2 dt3 V (t1 ) V (t2 ) V (t3 ) U (t3 , 0 )
0 0 0
Z Z Z t1
U (, 0 ) = 1 i dt1 V (t1 ) + (i)2 dt2 V (t1 ) V (t2 ) dt1
0 0 0
Z Z t1 Z t2  Z t3 
3
+ (i) dt1 dt2 dt3 V (t1 ) V (t2 ) V (t3 ) 1 i dt4 V (t4 ) U (t4 , 0 )
0 0 0 0

the expansion yields


Z Z Z t1
2
U (, 0 ) = 1 i dt1 V (t1 ) + (i) dt1 dt2 V (t1 ) V (t2 )
0 0 0
Z Z t1 Z t2
3
+ (i) dt1 dt2 dt3 V (t1 ) V (t2 ) V (t3 )
0 0 0
Z Z t1 Z t2 Z t3
4
+ (i) dt1 dt2 dt3 dt4 V (t1 ) V (t2 ) V (t3 ) V (t4 ) U (t4 , 0 )
0 0 0 0

once again for V (t) sufficiently small, we can neglect the term containing U (t, 0 ) on the RHS, because it is the
term of highest order. We then have
Z Z Z t1
2
U (, 0 ) = 1 i dt1 V (t1 ) + (i) dt1 dt2 V (t1 ) V (t2 )
0 0 0
Z Z t1 Z t2
+ (i)3 dt1 dt2 dt3 V (t1 ) V (t2 ) V (t3 ) + . . . (2.183)
0 0 0

from Eq. (2.178) we see that we obtain the Soperator expansion by taking 0 = and = in Eq. (2.183)
Z Z Z t1
2
S = 1i dt1 V (t1 ) + (i) dt1 dt2 V (t1 ) V (t2 )

Z Z t1 Z t2
+ (i)3 dt1 dt2 dt3 V (t1 ) V (t2 ) V (t3 ) + . . . (2.184)

this equation can also be derived directly from the old-fashioned perturbation expansion (2.177), by using the
Fourier representation of the energy factors in such an equation
Z
1
(E E + i) = i d exp [i (E E ) ] (2.185)
0

with the understanding that the integrals must be evaluated by inserting a convergence factor e in the integrand
with 0+ .
We can rewrite Eq. (2.184) in a way useful to carry out calculations that are manifestly Lorentz invariant. We
define the time-ordered product of any product of operators that depend on time as follows

T {O (t1 ) O (t2 ) . . . O (tn1 ) O (tn )} = O (tp1 ) O (tp2 ) . . . O tpn1 O (tpn ) (2.186)
tpn < tpn1 < . . . < tp2 < tp1 (2.187)
2.6. PERTURBATION THEORY 105

so that the operators become causally ordered from right (past) to left (future). Observe that the time-ordered
product operator basically apply some permutation to the set of indices 1, 2, . . . , n; to reorder the operators
according with the indices p1 , p2 , . . . , pn that satisfies Eq. (2.187). For one and two operators it can be written as

T {V (t)} = V (t)
T {V (t1 ) V (t2 )} = (t1 t2 ) V (t1 ) V (t2 ) + (t2 t1 ) V (t2 ) V (t1 )

where ( ) is the step function. The time-ordered product of n operators is a sum over all n! permutations of such
products, and each one gives the same integral over all times t1 t2 tn . Therefore, Eq. (2.184) can be written as

X Z
(i)n
S =1+ dt1 dt2 dtn T {V (t1 ) V (t2 ) V (tn )} (2.188)
n!
n=1

which is sometimes called the Dyson series. If all V (ti )s commute with each other, this sum becomes
 Z 
S = exp i dt V (t)

However, this is not usually the case. Sometimes Eq. (2.188) even does not converge, and is at most an asymptotic
expansion in some coupling factors that appear in V . Nevertheless, Eq. (2.188) can be written as
 Z 
S = T exp i dt V (t)

where T indicates that the expression must be evaluated by time-ordering each term in the series expansion of
the exponential.
Now, we shall find some sufficient conditions (satisfied by a large class of theories) for which the Smatrix
is manifestly Lorentz invariant.
Recalling
that the Smatrix elements are the elements of the Soperator in
the basis of free particles (0) , (0) etc. We want to find the conditions for which the Soperator commute
with the operator U0 (, a) that produces Lorentz transformations on the free-particle states. This is equivalent
to find the conditions for which the Soperator commute with the generators H0 , P0 , J0 and K0 . To satisfy that
requirement we shall try by using the hypothesis that V (t) is an integral over three-space
Z Z
V (t) = d x H (x, t) = d3 x H (x)
3
(2.189)

where H (x) is a scalar in the sense that

U0 (, a) H (x) U01 (, a) = H (x + a) (2.190)

it can be checked that H (x) has a time-dependence consistent with Eq. (2.180). To see it, we write an infinitesimal
pure translation operator in the Lorentz group Eq. (1.99), page 27

U0 (1, ) = 1 i P0

and we specify it even more as a pure time-translation



U0 1, 0 = 1 i0 P00 = 1 + iH0 0 (2.191)

applying (2.191) in Eq. (2.190) and equating the coefficients of 0 , we obtain that the time dependence of H (x)
is consistent with Eq. (2.180) [homework!!(9)].
106 CHAPTER 2. SCATTERING THEORY

By using the hypothesis (2.189), the Soperator (2.188) may be written as a sum of four-dimensional integrals
X Z Z  Z  Z 
(i)n 3 3 3
S = 1+ dt1 dt2 dtn T d x1 H (x1 ) d x2 H (x2 ) d xn H (xn )
n!
n=1
X Z
(i)n 4
S = 1+ d x1 d4 x2 d4 xn T {H (x1 ) H (x2 ) H (xn )} (2.192)
n=1
n!

in Eq. (2.192) everything is manifestly Lorentz invariant except for the time-ordering of the operator product.
Now, recall that for two given events x1 and x2 , the time ordering of two causally connected events [time-like
events i.e. (x1 x2 )2 < 0] is a Lorentz invariant, but it is not a Lorentz invariant for causally disconnected events
[space-like events i.e. (x1 x2 )2 > 0]. Therefore, it is clear that a sufficient condition for Eq. (2.192) to be Lorentz
invariant, is that H (x) commute at space-like or light-like separations17
  2
H (x) , H x = 0 for x x 0 (2.193)
On the other hand, we shall use the results of section 2.3.1, to prove with a non-perturbative argument that an
interaction of the type (2.189) satisfying Eqs. (2.190) and (2.193) does lead to an Smatrix with the appropriate
Lorentz transformation properties. For an infinitesimal boost, Eq. (2.190) yields [homework!!(10), reproduce
equations (2.194), and (2.195)]

i [K0 , H (x, t)] = tH (x, t) + x H (x, t) (2.194)
t
hence, integrating over x and setting t = 0, we obtain
 Z 
[K0 , V ] = K0 , d3 x H (x, 0) = [H0 , W] (2.195)
Z
W d3 x x H (x, 0) (2.196)

If (as it is usually the case) the matrix elements of H (x, 0) between eigenstates of H0 are smooth functions of the
energy eigenvalues, then the same is true for V , as is necessary for the validity of the scattering theory, and also
true for W, which is necessary in the proof of Lorentz invariance [This smooth behavior leads to Eq. (2.94), page
80].
From Eq. (2.195) we see that the other condition for Lorentz invariance, given by commutation relation (2.83)
page 78, is also valid if and only if
[H0 , W] = [H, W] = [H0 + V, W]
or equivalently Z Z
0 = [W, V ] = d3 x d3 y x [H (x, 0) , H (y, 0)] (2.197)

It is clear that the causality condition (2.193) leads to (2.197). However, Eq. (2.197) provides a less restrictive
sufficient condition for the Lorentz invariance of the Smatrix.
Since the conditions we have found are sufficient but not necessary, it is clear that the class of theories that
satisfy these conditions are not the only ones that are Lorentz invariant. However, the most general Lorentz
invariant theories are not very different. There is always a commutation condition quite similar to (2.193) that
should be satisfied.
Note that the causality condition (2.193), has no counterpart in non-relativistic theories. It owes to the
fact that time-ordering is always Galilean invariant, it is obvious since Galilean transformations keep the time
coordinate invariant. Consequently, condition (2.193) is the one that makes the combination of Lorentz invariance
and quantum mechanics very restrictive.
17
We prefer to include the light-like condition because we shall see later that Lorentz invariance can be disturbed by troublesome
singularities at x = x .
2.6. PERTURBATION THEORY 107

2.6.1 Distorted-wave Born approximation


The methods described above are useful as long as the interaction term V is sufficiently small. A modified method
of approximation known as the distorted-wave Born approximation, is useful when the interaction contains
two terms
V = Vs + VW (2.198)
such that VW is weak but Vs is strong. We can define |s i as the in and out states when we consider Vs as
the whole interaction. In that case, we can write the Lippmann-Schwinger equation (2.21) associated with the
strong interaction.
E

s = (0) + (E H0 i)1 Vs s (2.199)

D

1
s = (0) +
s Vs (E H0 i) (2.200)

from (2.198, 2.200) the second of Eqs. (2.22) associated with the complete interaction yields18
D h

i

+
T = (0) V + = s s Vs (E H0 + i)1 (Vs + VW ) +


h i
= s VW + + s Vs Vs (E H0 + i)1 (Vs + VW ) +


n o
= s VW + + s Vs + (E H0 + i)1 (Vs + VW ) +


n o
+
T = s VW + + s Vs + (E H0 + i)1 V +

and using the Lippmann-Schwinger equations associated with the whole interaction V , we obtain


E
+
T = s VW + + s Vs (0) (2.201)

+
The second term on the RHS of Eq. (2.201) is the term T when it is associated with the strong interaction Vs
alone
D
E
(0)
s+
T (0) Vs +
s = s

Vs (2.202)

To prove Eq. (2.202) we just follow the same procedure that led to Eq. (2.201) but dropping VW everywhere.
D h

i
1
s+
T = (0) Vs +
s = s

s

Vs (E H 0 + i) Vs +
s

h i
= s Vs Vs (E H0 + i)1 Vs + s

n + o
= s Vs s (E H0 + i)1 Vs + s

E

+
T = s Vs (0)

Equation (2.201) is most useful when the second term on the RHS vanishes. That is, when the process
cannot be produced by the strong interaction alone. For such processes the matrix element (2.202) vanishes.
Consequently, Eq. (2.201) becomes

+
T = s VW + (2.203)
up to now the equation is exact, at least under our assumptions. Nevertheless, this equation is useful when VW
can be considered so weak to neglect its effect on the state |+ i in Eq. (2.203). In that case, we can replace the
E
18
Of course, there is no distinction between the free states (0) in the Lippmann-Schwinger equations associated with the strong
interaction alone, with respect to the free states in the Lippmann-Schwinger equations associated with the whole interaction.
108 CHAPTER 2. SCATTERING THEORY

state |+ i (associated with the whole interaction) by the state |+s i (due to the strong interaction alone), in Eq.
(2.203)

+
T s VW +s (2.204)
this is valid to first order in VW but to all orders in Vs .
This approximation is used in many scenarios of Physics. A good example is the nuclear beta or gamma decay
for which the Smatrix element is calculated by using Eq. (2.204) with Vs being the strong nuclear interaction
and VW is either the weak nuclear interaction or the electromagnetic interaction, and with |s i and |+
s i the final
and initial nuclear states. In particular, in nuclear beta decay we require a weak nuclear force to turn neutrons
into protons, even though we cannot ignore the presence of the strong nuclear force. Thus, the process cannot be
produced by the strong interaction alone, and the second term on the RHS of Eq. (2.201) vanishes, as required
in the distorted-wave Born approximation.

2.7 Implications of unitarity


Equation (2.61) shows that the non-trivial part of the Smatrix is contained in the amplitude M . We shall see
that the unitarity condition of the Smatrix imposes a useful condition relating the amplitude M for forward
scattering in an arbitrary multiparticle state , to the total rate for all reactions in that state. We start with the
parameterization (2.61) of the Smatrix for a general process

S = ( ) 2i4 (p p ) M (2.205)

the unitarity condition is


1 = S S (2.206)
writing the unitarity condition (2.206) explicitly and using the parameterization (2.205) we find
Z Z

( ) = d S S = d S S
Z

 
= d ( ) + 2i4 (p p ) M ( ) 2i4 (p p ) M
Z Z
( ) = d ( ) ( ) 2i d ( ) 4 (p p ) M
Z Z

+2i d ( ) 4 (p p ) M + 4 2 d 4 (p p ) 4 (p p ) M M

( ) = ( ) 2i 4 (p p ) M + 2i 4 (p p ) M
Z

+4 2 d 4 (p p ) 4 (p p ) M M

cancelling ( ) we obtain
Z
4 4 2
2i (p p ) M + 2i (p p ) M + 4 d 4 (p p ) 4 (p p ) M M = 0

and taking into account that

4 (p p ) 4 (p p ) = 4 (p p ) 4 (p p )

we have
Z
4 4 2 4
2i (p p ) M + 2i (p p ) M + 4 (p p ) d 4 (p p ) M M = 0 (2.207)
2.7. IMPLICATIONS OF UNITARITY 109

we can factor 24 (p p ) in (2.207). Thus, for p = p we obtain19


Z

iM + iM + 2 d 4 (p p ) M M = 0 (2.208)

if we further assume that = , the first two terms on the LHS of Eq. (2.208) become

iM + iM = iM + (iM ) = 2Re [iM ]
= 2Re [i (ReM + iImM )] = 2Re [iReM + ImM ]

iM + iM = 2ImM

therefore, Eq. (2.208) for = yields


Z

2ImM + 2 d 4 (p p ) M M = 0
Z
ImM = d 4 (p p ) |M |2 (2.209)

which is the most useful form of Eq. (2.208). We shall relate ImM with the total rate for all reactions produced
by an initial state in a volume V . To do it we integrate Eq. (2.139) over all final states
Z Z h i
d ( )
d = d (2)3N 2 V 1N |M |2 4 (p p )
d
Z
= (2)3N 2 V 1N d 4 (p p ) |M |2 (2.210)

substituting Eq. (2.209) in Eq. (2.210) we find


1
= (2)3N 2 V 1N ImM (2.211)

Let us examine the particular case in which is a two-particle state. In that case, using Eq. (2.143), we see
that the total cross-section in the state is given by
Z Z
d ( ) (2)4
d = |M |2 4 (p p ) d (2.212)
d u

once again substituting Eq. (2.209) in Eq. (2.212) we obtain

1 (2)4
= ImM
u
16 3
= ImM (2.213)
u

where u is the relative velocity given by (2.145).


This is usually expressed in terms of a scattering amplitude f ( ) coming from the cross-section in the
center-of-mass reference frame. The differential cross-section for two-body scattering in the center-of-mass frame
is given by Eq. (2.169)

d ( ) (2)4 k E1 E2 E1 E2 2

p1 = p2 .
= |M | ; k |p1 | = |p2 | ; k (2.214)
d kE 2
19
Of course, for p 6= p Eq. (2.207) becomes trivial.
110 CHAPTER 2. SCATTERING THEORY

so we define the scattering amplitude as


r
4 2 k E1 E2 E1 E2
f ( ) M (2.215)
E k
so that the differential cross-section is written as
d ( )
= |f ( )|2 (2.216)
d
it is clear that the phase of f ( ) is conventional, and is usually motivated by the wave mechanical interpre-
tation of f as the coefficient of the outgoing wave in the solution of the time-independent Schrodinger equation.
In particular for elastic two-body scattering (so that k = k and Ei = Ei for i = 1, 2), we have
r
4 2 kE1 E2 E1 E2
f ( ) = M
E k
4 2 E1 E2
f ( ) = M (2.217)
E
now taking the imaginary part on Eq. (2.217), setting = , and using Eq. (2.213) we have

4 2 E1 E2 4 2 E1 E2 u
Im f ( ) = Im M =
E E 16 3
E1 E2
Im f ( ) = u
E 4
and using the expression (2.154) for the relative velocity u in the center-of-mass frame, we have
E1 E2 k (E1 + E2 ) 1
Im f ( ) = = kE
E 4 E1 E2 E 4
k
Im f ( ) = (2.218)
4
This form of the unitarity condition (2.213) for elastic scattering of two-body states is known as the optical
theorem.
We can obtain information about the pattern of scattering at high energy from the optical theorem. The
scattering amplitude can be expected to be a smooth function of angles. Hence, there should be some solid angle
within which |f |2 has nearly the same value (say within a factor of 2) as in the forward direction ( = 0). In
that case, the total cross-section is bounded by
Z
1 1
|f |2 d |f ( )|2 |Im f ( )|2
2 2
combining these inequalities with Eq. (2.218), we find an upper bound on
 2 
1 2 1 k 2 32 2
|Im f ( )| =
2 2 16 2 k 2
32 2
(2.219)
k 2
total cross-section is expected to approach constants or grow slowly at high energies. Consequently, Eq. (2.219)
shows that the solid angle around the forward direction within which the differential cross-section is roughly
constant shrinks at least as fast as k2 for k . This increasingly narrow peak in the forward direction at high
energies is called a diffraction peak.
2.7. IMPLICATIONS OF UNITARITY 111

2.7.1 Generalized optical theorem and CPT invariance


Returning to the general case of reactions involving an arbitrary number of particles, we shall relate the total
interaction rates of particles and antiparticles by combining Eq. (2.209) with CPT invariance.
Since CPT is antiunitary, it does not imply a simple relation between the process , and the corresponding
process obtained by changing particles by their antiparticles. Due to the role of the time-reversal (that leads to the
antiunitarity of CP T ) it shall provide a relation between a process and the inverse process involving antiparticles.
By using the same arguments that led to Eq. (2.115) page 87 for time-reversal invariance, we can see that CPT
invariance requires the Smatrix to satisfy the condition

S, = SCPT , CPT (2.220)

we recall that CPT implies that we must reverse all spin x3 components, change all particles by their correspond-
ing antiparticles, and multiply the matrix element by several phase factors for the particles in the initial state and
by their complex conjugates for the particles in the final state (all three-momenta are left invariant). Since CP T
invariance requires that the mass of each particle is the same as the mass of the corresponding antiparticle, the
relation (2.220) must also be hold by the amplitude or coefficient of 4 (p p ) in S shown in Eq. (2.205)

M, = MCPT , CPT (2.221)

in the particular case in which initial and final states are the same ( = ), all phases cancel since their module
is the unity. Hence, when = Eq. (2.221) becomes

Mp1 1 n1 ;p2 2 n2 ; ,p11 n1 ;p22 n2 ; = Mp1 (1 )nc1 ;p2 (2 )nc2 ; ,p1 (1 )nc1 ;p2(2 )nc2 ; (2.222)

so that the amplitude M, must coincide with the amplitude associated with states of the corresponding an-
tiparticles (characterized by the quantum numbers nci ) with opposite spin x3 components20 . Consequently, the
generalized optical theorem (2.209) or equivalently (2.211), says that the total reaction rate from an initial state
consisting of some set of particles is the same as for an initial state consisting with the corresponding antiparticles
with spins reversed.
p1 1 n1 ;p2 2 n2 ; = p1 (1 )nc1 ;p2 (2 )nc2 ; (2.223)

In the particular case of one-particle states, the decay rate of any particle equals the decay rate of the antipar-
ticle with reversed spin. On the other hand, rotational invariance does not allowed particle decay rates to depend
on the spin x3 component of the decaying particle. Hence, as a special case of the general result (2.223) we see
that unstable particles and their corresponding antiparticles have the same lifetimes.
Equation (2.209) was obtained from the unitarity condition S S = 1. By using the unitarity condition SS = 1
very similar arguments lead to the relation (recall that both unitarity conditions are not equivalent in an infinite
dimensional Hilbert space) Z
ImM = d 4 (p p ) |M |2 (2.224)

such that Eqs. (2.209, 2.224) lead to a reciprocity relation


Z Z
d (p p ) |M | = d 4 (p p ) |M |2
4 2
(2.225)

it is important to say that this reciprocity relation is far-from trivial since in the general case there is not any
simple relation between M and M . By using Eq. (2.139), we can rewrite equation (2.225) in the following
20
Note that setting = simplifies the relation (2.221) because the processes and are the same, and also because the
phase factors are all cancelled.
112 CHAPTER 2. SCATTERING THEORY

way
Z Z
d ( ) V N 1 d ( ) V N 1
d = d
d (2)3N 2 d (2)3N 2
Z Z
V 1 d ( ) V N V 1 d ( ) V N
d = d
(2)2 d (2)3N (2) 2 d (2)3N

obtaining finally
Z Z   N
d ( ) d ( ) V
d c = d c ; c (2.226)
d d (2)3

2.7.2 Unitarity condition and Boltzmann H-theorem


Equation (2.226) can be used to derive the Boltzmann H-theorem which is crucial in kinetic theory. Let P d
be the probability of finding the system in a volume d of the space of multiparticle states |i. We start by
calculating the rate of decrease in P due to transitions to all other states. In other words, the probability per
unit time that flows outside the volume of multi-particle states. The probability per unit time for a transition of
a fixed state into a set of states within the volume d is given by

d ( )
d
d

and integrating over we obtain the probability per unit time to obtain the transition from a fixed into any
final state . However, we have not certainty that the initial state is precisely , hence we have to multiply this
expression by P . We then finally obtain
  Z
dP d ( )
= P d
dt in d

now we calcualte the rate of increase of P due to transitions from all other states into the state . In other
words, the probability per unit time that flows inside the volume of multiparticle states. It can be derived with a
similar argument as   Z
dP d ( )
= d P
dt out d
therefore, the rate of change of P reads
    Z Z
dP dP dP d ( ) d ( )
= = d P P d (2.227)
dt dt in dt out d d

first of all we integrate over states on both sides


Z Z Z Z Z
d d ( ) d ( )
P d = d d P d P d (2.228)
dt d d

interchanging the labelling of the integration variables in the integral of the second term in Eq. (2.228), we obtain
that
Z Z Z Z Z
d d ( ) d ( )
P d = d d P d P d
dt d d
Z Z Z Z Z
d d ( ) d ( )
P d = d d P d d P
dt d d
2.7. IMPLICATIONS OF UNITARITY 113

such that Z
d
P d = 0
dt
leads to the conservation of probability. On the other hand, we can calculate the rate of change of the entropy
Z   Z    Z       
dS d P d P dP P d P
d P ln = d P ln = d ln + P ln
dt dt c dt c dt c dt c
Z       Z     
dP P c d P dP P dP
= d ln + P = d ln +
dt c P dt c dt c dt
Z    
dS P dP
= d ln +1 (2.229)
dt c dt

Substituting (2.227) into Eq. (2.229) we find


Z Z     
dS P d ( ) d ( )
= d d ln + 1 P P (2.230)
dt c d d
Z Z     Z Z    
dS P d ( ) P d ( )
= d d ln + 1 P + d d ln + 1 P (2.231)
dt c d c d

interchanging the labelling of the integration variables in the second term, we write Eq. (2.231) as follows
Z Z     Z Z    
dS P d ( ) P d ( )
= d d ln + 1 P + d d ln + 1 P
dt c d c d
Z Z    Z Z
P d ( ) d ( )
= d d ln P d d P
c d d
Z Z    Z Z
P d ( ) d ( )
+ d d ln P + d d P
c d d
Z Z     
P P d ( )
= d d ln ln P
c c d
Z Z  
dS P c
d ( )
= d d P ln (2.232)
dt P c d

now for any two positive quantities x and y, the following inequality holds
y 
y ln yx (2.233)
x
setting x = P c and y = P c the inequality (2.233) becomes
   
P c P c P c
P c ln P c P c P ln P
P c P c c
   
P c P P
P ln c (2.234)
P c c c

substituting (2.234) into (2.232), we see that the rate of change of entropy is bounded by
Z Z  
dS P P d ( )
d d c
dt c c d
Z Z   Z Z  
dS P d ( ) P d ( )
d d c d d c
dt c d c d
114 CHAPTER 2. SCATTERING THEORY

and interchanging variables of integration in the second term we have


Z Z   Z Z  
dS P d ( ) P d ( )
d d c d d c
dt c d c d
Z Z  
dS P d ( ) d ( )
d d c c
dt c d d
Z Z  
dS P d ( ) d ( )
d d c c
dt c d d

Now, the unitarity relation (2.226) (with and interchanged) says that the integral over on the RHS of this
inequality vanishes. Therefore, we conclude that the entropy never decreases
Z  
dS d P
d P ln 0
dt dt c

which is the so-called Boltzmann H-theorem. In textbooks of statistical mechanics, such a theorem is deduced
using the Born approximation, for which |M |2 is symmetric in and so that

d ( ) d ( )
c = c
d d

or assuming time-reversal invariance, which would imply that |M |2 is unchanged by the interchange of and
, combined with the reversal of all momenta and spins. In that sense the present derivation is more general
since we have only used the unitarity result (2.226), to derive the Boltzmann H-theorem. By contrast the Born
approximation and time-reversal invariance are not exact.
We can also observe that the conditions to obtain a steady entropy also depends only on the unitarity relation
(2.226) and not of the Born approximation or time-reversal invariance. To see it we observe that the entropy
becomes constant when P becomes a function only of conserved quantities such as the total energy and charge,
times the factor c . In this case the conservation laws require

d ( ) P P
=0 or =
d c c

therefore we can replace


c
P P
c
in the first term of Eq. (2.227). Now using the unitarity relation (2.226) shows that in this case, the probability
P is time-independent.
Chapter 3

The cluster decomposition principle

Now we shall take up the problem of constructing the Hamiltonian operator in a suitable way. Such an operator
can be defined through its matrix elements between states of arbitrary numbers of particles. We shall see that
another more suitable way to construct the Hamiltonian is by expressing it as a function of operators that create
and destroy single particles. Historically, these operators appeared first after quantizing the electromagnetic field
as well as other fields. However, there is another motivation aside of the quantization of a classical field or the
question whether particles are created or destroyed. We shall see that if we express the Hamiltonian as a sum of
products or creation and annihilation operators with suitable non-singular coefficients, the Smatrix will satisfy
the cluster decomposition principle, which says that distant experiments produce uncorrelated results. It is for this
reason that the formalism of creation and annihilation operators is widely used in quantum statistical mechanics,
even if the number of particles is fixed. In relativistic quantum theories the cluster decomposition principle leads
inevitably to quantum field theory. Many attempts have been done to construct a relativistic quantum theory
that would not be a local field theory. Though it is possible to construct theories that are not field theories and yet
lead to a Lorentz-invariant Smatrix for two-particle scattering, such theories have always problems for systems
of more than two particles: Either the three-particle Smatrix is not Lorentz invariant, or else violates the cluster
decomposition principle.
We first construct the basis of states with arbitrary number of bosons and fermions, then define the creation
and annihilation operators, and show how we can construct Hamiltonians based on those operators, that yield
Smatrices that satisfy the cluster decomposition principle.

3.1 Physical states


The Hilbert space whose vectors describe the physical states must contain states describing arbitrary numbers
of particles. If Ei is the Hilbert space whose vectors describe physical one-particle states, then a Hilbert space
describing N -particles is given by
E (N ) = E1 . . . EN
since we have to describe physical states of 0, 1, 2, 3, . . . particles, the total Hilbert space must be

E = E (1) E (2) . . . E (N ) . . .

We shall consider physical states of 0, 1, 2, 3, . . . free particles. We could consider either free-particles states, or
in or out states. For definiteness we shall use free-particle states

|p1 1 n1 , p2 2 n2 , . . .i

but all our results will be valid for in or out states as well. As customary, denotes spin x3 components or
helicities for massless particles, and ni denotes particle species.

115
116 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

3.1.1 Interchange of identical particles


There is an essential point that has not been considered yet: the symmetrization postulate of quantum
mechanics. According with it, a set of identical particles (that is particles that possess the same quantum
number n that identifies the species) is described by physical states that are either symmetric or antisymmetric
under the interchange of two particles within the set of identical particles. When the set of identical particles
is described by a symmetric (antisymmetric) state we call the particles bosons (fermions). As far as we know,
all particles are either bosons or fermions. We describe this symmetry property of a set of identical bosons or
fermions as

|p1 1 n, p2 2 n, . . . , pi i n, . . . , pj j n, . . .i = n |p1 1 n, p2 2 n, . . . , pj j n, . . . , pi i n, . . .i (3.1)



+1 if the particles are bosons
n (3.2)
1 if the particles are fermions

these two cases are called the boson and fermion statistics. We shall see later that Bose and Fermi statistics are
only possible for particles with integer or half-integer spins respectively, but we shall not need this information by
now. We shall set up suitable normalization conditions for multi-boson or multi-fermion states.
We first notice that if two particles with momenta and spins pi , i and pj , j belong to identical species n, the
two state vectors

|p1 1 n, p2 2 n, . . . , pi i n, . . . , pj j n, . . .i ; |p1 1 n, p2 2 n, . . . , pj j n, . . . , pi i n, . . .i (3.3)

represent the same physical state. Otherwise, the particles would be distinguished by their order in the labelling of
the state-vector, and the first listed would not be identical to the second. Since both states in (3.3) are physically
indistinguishable, they must belong to the same ray, so that

|p1 1 n, p2 2 n, . . . , pi i n, . . . , pj j n, . . .i = n |p1 1 n, p2 2 n, . . . , pj j n, . . . , pi i n, . . .i (3.4)

where n is a complex number lying in the unitary complex circle. We could consider it, as part of the definition
of what we mean by identical particles.
Now the gist of the reasoning is to decide of what variables could n depends. If n only depends on the
species n, then interchanging the two particles in (3.4) again, we obtain

|p1 1 n, p2 2 n, . . . , pi i n, . . . , pj j n, . . .i = 2n |p1 1 n, p2 2 n, . . . , pi i n, . . . , pj j n, . . .i

such that 2n = 1, yielding Eq. (3.2) as the only two possibilities for the phase.
Suppose now that n could also depend on the numbers and species of the other particles in the state. It
would lead to the unpleasant conclusion that the symmetry of the state-vectors under interchange of particles in a
given place may depend on the presence of particles in any other place in the universe. These are the possibilities
that we discard from the cluster decomposition principle.
Another possibility is to assume that n could depend on the spins of the two particles. In that case, the
(j)
spin-dependent n phase factors would provide a one-dimensional representation of the rotation group in three-
dimensional space SO (3) (or in SU (2)). However, the only one-dimensional irreducible inequivalent representation
of SO (3), is the trivial one (consisting of the identity alone), there are not one-dimensional representations
consisting of various phase factors. Thus, n must be independent of the spin1 .
Moreover, let us examine the scenario in which n could depend on the three-momenta of the two particles
involved in the interchange. In this scenario, Lorentz-invariance would require that n would only depend on the
scalars
p1 p1 , p2 p2 , p3 p3
1
Note that for the group of rotations in two dimensions SO (2), there are non-trivial representations in one dimension consisting of
various phases. This fact opens the window for the possibility of having phases that depend on the spin in models constructed on a
two-dimensional space.
3.1. PHYSICAL STATES 117

which are symmetric under the interchange of the particles 1 and 2. Therefore, such dependence would not change
the argument, from which 2n = 1.
There is another possibility in which the states |p1 1 n, p2 2 n, . . .i could carry a phase factor that depends on
the path through the momentum space by which the momenta of the particles are brought to the values p1 , p2 etc.
In this case, the interchange of two particles twice might change the state by a phase factor for which 2n 6= 1. It
could be shown that this is a real possibility in two-dimensional space, but not for three or more spatial dimensions.

3.1.2 Interchange of non-identical particles


There is not any essential symmetry property under the interchange of particles of different species. However,
we could agree to choose certain symmetry patterns under these interchanges that will show to be convenient for
future purposes.
For instance we could agree to label the state-vector by listing all photon momenta and helicities first then all
electron momenta and spin x3 components, and so on by sweeping the elementary particle species. Alternatively,
we could allow the particle labels to appear in any order, and define the state-vectors with particle labels in an
arbitrary order as equal to the state-vectors with particle labels in some standard order times phase factors, whose
dependence on the interchange of particles of different species can be anything we like. For example, suppose that
we have an electron e a muon and a photon , we can describe this multiparticle state by listing the quantum
numbers of the electron followed by the ones of the muon and then the ones of the photon
|i |pe e ne , p n , p n i
but we can also list them (say) starting with the muon numbers, then the electron numbers and finally the photon
numbers
|p n , pe e ne , p n i
and since the quantum numbers of each particle are the saem as in the ket |i, it is clear that both kets must
describe the same physical state. Therefore both must belong to the same ray

|i = n
However, since particles are not identical, the phase do not have to follow the symmetrization postulate pattern.
We can accomodate any pattern to define the phases when non-identical particles are interchanged.
It is important to say that there are symmetries like the isospin invariance that relate particles of different
species. From this fact, it would be convenient to choose a convention to generalize the symmetry pattern in
(3.1, 3.2). We shall do it as follows: the state-vector will be chosen to be symmetric under the interchange of
any bosons with each other, or any bosons with any fermions, and antisymmetric with respect to interchange of
any two fermions with each other. This pattern is chosen regardless whether the particles interchanged are of the
same species or not. Of course, this pattern contains the symmetrization postulate as a special case.
Note that this reasoning also shows that the symmetry or antisymmetry of the state-vector under interchange of
particles of the same species but different helicities (or different x3 components of the spin), is purely conventional.
It is because we could agree from the beginning to list first the momenta of photons of helicity +1, then the
momenta of all photons with helicity 1, then the momenta of all electrons of = +1/2, then all three- momenta
of electrons with = 1/2 and so on. We shall choose the convention that the state-vector is symmetric or
antisymmetric under interchange of identical bosons or fermions of different helicities or spin u3 components, in
order to facilitate the use of rotational invariance.

3.1.3 Normalization of multi-particle states


The normalization of the multi-particle states must be defined in consistency with the previous symmetry con-
ventions. We shall use a label q to denote all quantum numbers of a single particle
q p, , n
118 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

thus, N particle states are denoted as


|q1 , q2 , . . . , qN i
the vacuum state with N = 0 is denoted as |0i. For N = 0, 1 the symmetry of interchange is not relevant, hence
we normalize these states as


h0 |0i = 1 ; q |qi = q q
 
q q 3 p p n n

On the other hand, for N = 2 the vectors |q1 , q2 i and |q2 , q1 i describe the same physical state, thus the normal-
ization is written as

   
q1 , q2 |q1 , q2 i = q1 q1 q2 q2 + q2 q1 q1 q2 (3.5)

1 if both particles are fermions

+1 otherwise

More generally
N!
X N
Y


q1 , q2 , . . . , qM

|q1 , q2 , . . . , qN i = N M P qi qP i (3.6)
P=1 i=1

where N M indicates that both states are orthogonal if M 6= N i.e. for states of different number of particles. The
sum is over all N ! permutations of the integers 1, . . . , N . Further

1 if P involves an odd permutation of fermions
P
+1 otherwise

thus P = 1 if the permutation implies an odd number of fermion interchanges and +1 otherwise. It is easy
to check that the orthonormalization condition (3.6) fullfills all symmetry and antisymmetry requirements under
interchange of the qi and also under the interchange of the qj .

3.2 Creation and annihilation operators


We shall define the creation and annihilation operators by their effect on normalized multi-particle states. We
denote the creation operator as a (q) (i.e. as the adjoint of certain operator a (q)) because of the analogy with
creation operators in the quantum harmonic oscillator. We define the creation operator a (p, , n) a (q), as
the operator that adds a particle with quantum numbers q at the front of the list of particles in the state2

a (q) |q1 q2 qN i |qq1 q2 qN i (3.7)

In particular, the N particle state can be obtained by applying N creation operators on the vacuum state

a (q1 ) a (q2 ) . . . a (qN 1 ) a (qN ) |0i = |q1 q2 qN 1 qN i (3.8)

on the other hand, the adjoint of the creation operator is denoted as a (q). We shall show that a (q) removes a
particle from any state on which it acts and then it is called the annihilation operator. We shall show it, for
the case in which the particles qq1 q2 qN are either all bosons or all fermions. We start by calculating the scalar
product of a (q) |q1 q2 qN i with an arbitrary state |q1 q2 qM i. Using Eq. (3.7) it yields




D 

q1 q2 qM a (q) |q1 q2 qN i a (q) q1 q2 qM q1 q2 qN i






q1 q2 qM a (q) |q1 q2 qN i = qq1 q2 qM q1 q2 qN i (3.9)
2
For the sake of phase conventions it is important to define in what place of the list the new particle is added.
3.2. CREATION AND ANNIHILATION OPERATORS 119

now we write the scalar product on the RHS of Eq. (3.9) according with Eq. (3.6).
N!
X M
Y +1







q1 q2 qM a (q) |q1 q2 qN i = qq1 q2 qM q1 q2 qN i = N,M +1 P qi qP i (3.10)
P=1 i=1
In equation (3.10), we shall separate the sum over all permutations P in a sum over the integer r that is permuted
into the first place Pr = 1, and a sum over mappings P from the remaining integers

r 1, 2, . . . , r 1, r + 1, . . . , N

into 1, 2, . . . , N 1. Moreover, since particles qq1 q2 qN are either all bosons or all fermions, the sign factor is

P = r1
P (3.11)

with equal to +1 for bosons and (1) for fermions [homework!!(11) fill the details to arrive to Eq. (3.12)]. By
separating the permutations as described above, and using (3.11), Eq. (3.10) becomes
N X
X M
Y




q1 q2 qM a (q) |q1 q2 qN i = N,M +1 r1
P (q q r ) qi qPi
r=1 P i=1

XN X M
Y




q1 q2 qM a (q) |q1 q2 qN i = N,M +1 r1
(q qr ) P qi qPi
r=1 P i=1

and using Eq. (3.6) again, we find


N
X






q1 q2 qM a (q) |q1 q2 qN i = N,M +1 r1
(q qr ) q1 q2 qM q1 q2 qr1 qr+1 qN i
r=1

finally, since the state |q1 q2 qM


i is arbitrary, we obtain

N
X
a (q) |q1 q2 qN i = r+1
(q qr ) |q1 q2 qr1 qr+1 qN i (3.12)
r=1

where we have used tha fact that r+1


= r1
. Equation (3.12) shows that the operator a (q) removes a particle
from any state in which it acts, as stated. As a special case of Eq. (3.12), we observe that for both bosons and
fermions, a (q) annihilates the vacuum

a (q) |0i = 0 h0| a (q) = 0 (3.13)

3.2.1 Commutation and anti-commutation relations of a (q) and a (q)


Applying a (q ) on Eq. (3.7) and using Eq. (3.12) we have
 
a q a (q) |q1 q2 qN i = a q |qq1 q2 qN i
N
X




a q a (q) |q1 q2 qN i = q q |q1 q2 qN i + r+2
q qr |qq1 q2 qr1 qr+1 qN i (3.14)
r=1

where the sign in the second term is r+2


because qr is in the (r + 1) th place in |qq1 q2 qN i.

On the other hand, applying a (q) in Eq. (3.12) we find
N
X



a (q) a q |q1 qN i = r+1
q qr |qq1 q2 qr1 qr+1 qN i (3.15)
r=1
120 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

substracting or adding Eqs. (3.14, 3.15) we obtain


n  o 
a q a (q) a (q) a q |q1 qN i = q q |q1 qN i (3.16)

now Eq. (3.16) holds for arbitrary states |q1 qN i containing either only bosons or only fermions (but could be
easily extended for states containing both bosons and fermions). Thus we obtain
  
a q a (q) a (q) a q = q q (3.17)

moreover, Eq. (3.17) leads to  


a q a (q) a (q) a q = 0 (3.18)
and taking the adjoint of (3.18) we also have
 
a (q) a q a q a (q) = 0 (3.19)

recall that in Eqs. (3.17, 3.18, 3.19) the top and bottom signs are for bosons and fermions respectively. Moreover,
according with the phase conventions discussed in section 3.1.2, the creation and/or annihilation operators for
particles of two different species commute if either particle is a boson, and anticommute if both are fermions.
In many textbooks this discussion is given in the opposite order. We could start with the commutation of
anticommutation relations (3.17, 3.18, 3.19) derived from the canonical quantization of some given filed theory.
Then multi-particle states are defined by constructing from the vacuum as in Eq. (3.8), and their scalar product
Eq. (3.6) is derived from the commutation or anticommutation relations. The latter order of reasoning is indeed
closer to the historical development. However, we have followed this order to show the necessity of the field
quantization later.

3.3 Arbitrary operators in terms of creation and annihilation operators


Let us define an operator O in the form of a sum of products of creation and annihilation operators as follows
Z
X
X   
O= dq1 dqN

dq1 dqM a q1 a qN

a (qM ) a (q1 ) CN M q1 qN

; q1 qM (3.20)
N =0 M =0

now we shall show that any linear operator can be expressed in the form prescribed by equation (3.20) by choicing
suitable values of the coefficients CN M . To do it, we shall prove that the CN M coefficients can be chosen such that
the matrix elements of this expression acquire any desired values. The proof will use mathematical induction.
First we start by proving that h0| O |0i can take any desired value by the apropriate choice of the CN M
coefficients. By using Eq. (3.20) we obtain
Z
X
X   
h0| O |0i = dq1 dqN

dq1 dqM CN M q1 qN

; q1 qM h0| a q1 a qN

a (qM ) a (q1 ) |0i
N =0 M =0

however Eqs. (3.13) say that only the terms with N = M = 0 contribute in this expansion

h0| O |0i = C00

thus to give the matrix element h0| O |0i any desired value, we only have to fix the value of C00 regardless the
values of the other CN M coefficients.
Now suppose that all matrix elements of O between N and M particle states with N < L, M K or
N L, M < K; have acquired any desire values by choicing appropriate values of the CN M coefficients. We must
3.4. TRANSFORMATION PROPERTIES OF THE CREATION AND ANNIHILATION OPERATORS 121

prove that the same is true for matrix elements of O between any Land Kparticle states. To do it we use Eq.
(3.20) to evaluate the matrix element in question
Z
X
X


q1 qL O |q1 qK i = dq1 dqN

dq1 dqM CN M q1 qN

; q1 qM
N =0 M =0

 
q1 qL a q1 a qN

a (qM ) a (q1 ) |q1 qK i



q1 qL O |q1 qK i = L!K!CLK q1 qL ; q1 qK
+terms involving CN M with N < L, M K or N L, M < K

whatever values we have given to CN M with N < L, M K or N L, M < K, there is some choice of CLK
that gives this matrix element any desired value.
The operator (3.20) could have been defined with any other order of the creation and annihilation operators.
The order followed in Eq. (3.20) in which all creation operators are put to the left of all annihilation operators is
usually called the normal order of the operators. For instance if we have defined an operator with the structure
of (3.20) but with the creation and annihilation operators in some other order, it is always possible to bring the
creation operators to the left of the annihilation operators by succesive use of the commutation or anticommutation
relations (3.17), from which it is clear that we pick up new terms from the delta function in Eqs. (3.17).
As an example, let us consider an additive operator F (like momentum, charge, etc.) for which

F |q1 qN i = [f (q1 ) + . . . + f (qN )] |q1 qN i (3.21)

such an operator can be written as in Eq. (3.20) using only the term with N = M = 1
Z
F = dq a (q) a (q) f (q) (3.22)

A very important particular case, is the free-particle Hamiltonian


Z p
H0 = dq a (q) a (q) E (q) ; E (q) = p2 + m2n (3.23)

where E (q) is clearly the single-particle energy3 .

3.4 Transformation properties of the creation and annihilation operators


We should characterize the transformation properties of the creation and annihilation operators under the various
symmetries we have already considered.
We shall start with the inhomogeneous proper orthochronus Lorentz transformations for massive particles. We
recall that the N particle states (of non-null mass) have the transformation property (2.57) under such a group
s
(p1 )0 (p2 )0
U0 (, a) |p1 1 n1 , p2 2 n2 , . . .i = exp [i (p1 ) a] exp [i (p2 ) a]
p01 p02
X (j ) (j )
D111 (W (, p1 )) D222 (W (, p2 ))
1 2
|p1 1 n1 , p2 2 n2 , . . .i (3.24)
3
Recall that for interacting multi-particle systems, energy cannot be assigned uniquely to each particle, but only to the whole
system. Thus, the energy operator (Hamiltonian) is not additive anymore.
122 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

(j)
where p is the three-vector part of the four-vector p, and D is the irreducible unitary spinj representation
of SO (3). Finally, W (, p) is the particular rotation

W (, p) L1 (p) L (p)

where L (p) is the standard boost that takes a particle of mass m 6= 0, from rest to four momentum p . It
is obvious that m and j depends on the species label n. We insist that this is for m 6= 0, the case of massless
particles will be studied later.
On the other hand, these states can be constructed from the vacuum as in Eq. (3.8)

|p1 1 n1 , p2 2 n2 , . . .i = a (p1 1 n1 ) a (p2 2 n2 ) |0i

where |0i is the vacuum state that we assume Lorentz-invariant

U0 (, a) |0i = |0i (3.25)

In order that the state (3.8) transforms properly [taking into account the transformation properties (3.24, 3.25)],
it is necessary and sufficient that the creation operator have the transformation rule
s
(p)0 X (j)
U0 (, a) a (pn) U01 (, a) = exp [i (p) a] D (W (, p)) a (p n) (3.26)
p0

Similarly, the discrete operators C, P and T , that produce charge-conjugation, space inversion, and time-reversal
on free-particle states, transform the creation operators as

C a (pn) C 1 = n a (pnc ) (3.27)


1
P a (pn) P = n a (p n) (3.28)
1 j
T a (pn) T = n (1) a (p n) (3.29)

as mentioned in Sec. 3.1 the whole formalism was constructed on free-particle states but can be extrapolated to
in or out states. We can then introduce operators ain and aout defined in the same way by their effects on
the in and out states. These operators satisfy the same Lorentz transformation rule described by Eq. (3.26),
but with the true Lorentz transformation operator U (, a) instead of the free-particle operator U0 (, a).

3.5 Cluster decomposition principle and connected amplitudes


It is a fundamental principle in Physics and indeed in all sciences, that phenomena that are sufficiently separated
in space are not correlated. Otherwise, any result of any experiment would depend on all other experiments in
the earth and even worse, on all other experiments and/or phenomena that occur in the earth and in the whole
universe. If this principle (known as the cluster decomposition principle) were not true, we were unable to make
any predictions on any experiment without knowing everything about the whole universe.
As usual the greek letters , will denote a collection of particles including for each particle a specification
of its momentum, spin and species. We shall also denote 1 + 2 + . . . + N to the state formed by combining
all particles in the states 1 , 2 , . . . , N . With this notation, we shall establish how could we express the cluster
decomposition principle in Smatrix theory.
Suppose we have a set of multi-particle processes 1 1 , 2 2 , . . . , N N and that each process is
very distant from each other. The fact that this set of processes produces uncorrelated results implies that the
Smatrix element for the overall process factorizes as follows

S1 +2 +...+N , 1 +2 +...+N S1 1 S2 2 SN N (3.30)


3.5. CLUSTER DECOMPOSITION PRINCIPLE AND CONNECTED AMPLITUDES 123

Thus, Eq. (3.30) holds if for all i 6= j all of the particles in states i and i are at a large spatial distance from all
of the particles in states j and j . The factorization of the Smatrix elements leads in turn to the factorization
of the corresponding transition probabilities, associated with uncorrelated experimental results. It says that the
probability of ocurrence for two (or more) independent or uncorrelated events, is the product of each probability
as it must be.
We shall rewrite Eq. (3.30) in a more transparent way by using a combinatoric trick. We define the connected
part of the Smatrix S C by the formula

X
S = SC1 1 SC2 2 (3.31)
partitions

where the sum is over all different ways of partitioning the particles in the state into clusters 1 , 2 , . . . , and
similarly a sum over all ways of partitioning the particles in the state into clusters 1 , 2 , . . . ,. We do not
count as different those partitions that merely arrange particles within a given cluster or permute whole clusters.
The factor is +1 or (1) according to whether the rearragements 1 2 and 1 2 involve
altogether an even or an odd number of fermion interchanges, respectively. We shall justify later the use of the
term connected.
Note that (3.31) is a recursive definition. For each and , the sum on the RHS of Eq. (3.31) consists of a
term SC plus a sum over products of two or more S C matrix elements, with a total number of particles in

each of the states j and j that is less than the number of particles in the states and 4 . In other words, we
can separate (3.31) in a connected term in which and are consider as single clusters, plus terms in which
and are separated into two or more clusters
X
C
S = S + SC1 1 SC2 2
partitions

Suppose that SC in this sum have been chosen such that Eq. (3.31) is satisfied for states , containing together

fewer than say N particles. Therefore, regardless the values found in this way for the Smatrix elements appearing
in the sum , we can always choose the remaining term S C in such a way that Eq. (3.31) is also satisfied by

states , containing a total of N particles. Thus Eq. (3.31) has no information by itself, it is simply a definition
of S C .
This argument works only if we consider that each j and j form non-empty sets. We must define the
connected vacuum-vacuum element S0,0 C to be zero. Equation (3.31) cannot be used to define the vacuum-vacuum

Smatrix S0,0 , which in the absence of time-varying external fields is simply defined to be the unity S0,0 = 1.
The following definition is useful to obtain the numbers of partitions in clusters for a given couple of multi-
particle states and .
Definition 3.1 A partition (n) {1 , 2 , . . . , r } of the positive integer
P n is a sequence of positive integers i ,
arranged in descending order, whose sum is n, that is i i+1 and ri=1 i = n.

It is clear that the number of partitions of a state with N particles into clusters, is the number of partitions
of the positive integer N (some permutations within a given partition must be included as we shall see later). The
partitions of the first four positive integers are given by
(1) = {1} ; (1) (2) = {2} ; (2) (2) = {1, 1} (3.32)
(1) (2) (3)
(3) = {3} ; (3) = {2, 1} ; (3) = {1, 1, 1} (3.33)
(1) (2) (3) (4)
(4) = {4} ; (4) = {3, 1} ; (4) = {2, 1, 1} ; (4) = {2, 1, 1} (3.34)
4
It is clear that if 1 , 2 , . . . , p defines a partition of the state with p 2, the number of particles Nk in each subset k must
satisfy the condition
N1 + N2 + . . . + Np = N
where N is the number of particles in the state . Thus Nk < N for each k . Similarly occurs for the state and any partition of it,
in two or more subsets.
124 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

3.5.1 Some examples of partitions


Thus, the simplest case arises when both and are one-particle states with quantum numbers q and q respec-
tively. In that case, the only term on the RHS of Eq. (3.31) is S C itself from which the connected Smatrix

becomes 
SqC q Sq q = q q (3.35)
apart from possible degeneracies, the proportionality of Sq q with (q q) comes from the conservation laws.
The absence of a proportionality factor in Eq. (3.35) comes from a suitable choice of relative phase between in
and out states. Here we are assuming that single-particle states are stable, such that there are no transitions
between single-particle states and any others, e.g. the vacuum.
The next simple case corresponds to transitions between two-particle states. In that case, Eq. (3.31) becomes

Sq1 q2 ,q1q2 = SqC q ,q1 q2 + SqC ,q1 SqC ,q2 + SqC ,q2 SqC ,q1
1 2 1 2 1
  2  
Sq1 q2 ,q1q2 = SqC q ,q1 q2 + q1 q1 q2 q2 + q1 q2 q2 q1 (3.36)
1 2

where is (1) if both particles in the interchange q1 q2 are fermions, and +1 otherwise. Observe that we
have used Eq. (3.35). We recognize that the two delta function terms in (3.36) just add up to the norm (3.5).
C is just (S 1) . However, the general case is more complicated.
Therefore, in this case S
For transitions between three-particle states, equation (3.33) shows that we can divide and in the following
three types of clusters (1) and as single clusters, (2) two clusters, (1 , 1 ) with one-particle and (2 , 2 ) with
two-particles, (3) Three-clusters, each one with a single particle. Thus, Eq. (3.31) gives

Sq1 q2 q3 ,q1q2 q3 = SqC q q ,q1 q2 q3


1 2 3

+SqC ,q1 SqC q ,q2q3 permutations


1 2 3

+SqC ,q1 SqC ,q2 SqC ,q3 permutations


1 2 3

using Eq. (3.35) we have

Sq1 q2 q3 ,q1 q2 q3 = SqC q q ,q1 q2 q3


1 2 3

+ q1 q1 SqC q ,q2q3 permutations
 2 3  
+ q1 q1 q2 q2 q3 q3 permutations

(3.37)

where the connected Smatrices for two-particle states are given by Eq. (3.36). Taking all permutations into
account the total number of terms in Eq. (3.37) is given by

1 + 9 + 6 = 16

as an example, the nine permutations associated with SqC q1 SqC q ,q2q3 in Eq. (3.37) are given by
1 2 3

SqC q1 SqC q ,q2q3 , SqC q2 SqC q ,q1q3 , SqC q3 SqC q ,q1 q2


1 2 3 1 2 3 1 2 3

SqC q1 SqC q ,q2q3 , SqC q2 SqC q ,q1q3 , SqC q3 SqC q ,q1 q2


2 1 3 2 1 3 2 1 3

SqC q1 SqC q ,q2q3 , SqC q2 SqC q ,q1q3 , SqC q3 SqC q ,q1 q2


3 1 2 3 1 2 3 1 2

In addition, we observe that we are not taking as different permutations within a given cluster or permutations
of complete clusters. For instance, we could express that the process 1 1 is uncorrelated with the process
23 2 3 by means of SqC q1 SqC q ,q2 q3 . However, it is clearly equivalent to say (for instance) that 1 1 is
1 2 3
uncorrelated with the process 32 2 3 , which is expressed by SqC q1 SqC q ,q3q2 (that implies a permutation within
1 2 3
3.5. CLUSTER DECOMPOSITION PRINCIPLE AND CONNECTED AMPLITUDES 125

a given cluster). It is also equivalent to say that the process 23 2 3 is uncorrelated with the process 1 1 ,
which is expressed by SqC q ,q3 q2 SqC q1 (that implies the permutation of complete clusters).
2 3 1
On the other hand, the transitions between four-particle states allow the following partitions (1) The four-
particles as a whole (2) Two clusters, each one with two-particles (3) two clusters the first with a single particle
and the second with three-particles. (4) Three-clusters two of them with a single particle and the third one with
two-particles. (5) Four clusters each one with a single-particle. Therefore, the Smatrix reads

Sq1 q2 q3 q4 ,q1 q2 q3 q4 = SqC q q q ,q1 q2 q3 q4


1 2 3 4

+SqC q ,q1 q2 SqC q ,q3 q4 permutations


1 2
3 4
+ q1 q1 SqC q q ,q2q3 q4 permutations
 234 
+ q1 q1 q2 q2 SqC q ,q3 q4 permutations
  34  
+ q1 q1 q2 q2 q3 q3 q4 q4 permutations (3.38)

and including all permutations the total number of terms in Eq. (3.38) yields

1 + 18 + 16 + 72 + 24 = 131

If we had not assumed that one-particle states are stable, there would be more terms in Eqs. (3.37, 3.38).
The process described above show that the definition of S C is recursive since we used Eq. (3.35) to define S C

for two-particle states, then use this definition in Eq. (3.37) when we define S C for three-particle states, then use

the previous definitions in Eq. (3.38) to obtain the definition of S C for four-particle states, and so on.

The gist of the definition of the connected part of the Smatrix is that the cluster decomposition principle is
equivalent to demand that S C must vanish when any one or more of the particles in the states and/or are

far away in space from the others.


We shall show the statement above by supposing that the states and are grouped into clusters 1 , 2 , . . .
and 1 , 2 , . . .. We also assume that all particles in the set i + i are far from all particles in the set j + j for
each j 6= i. On the other hand, suppose that SC vanishes if any particles in or are far from the others.
This in turn implies that SC vanishes if any particles in these states are in different clusters, so we have
X
SCk k (k)
SCk1 k1 SCk2 k2

so that the definition (3.31) gives


X X
(1)
S SC11 11 SC12 12 (2)
SC21 21 SC22 22

where (j) is a sum over all different ways of partitioning the clusters j and j into subclusters j1 , j2 , . . . and
j1 , j2 , . . .. But referred to Eq. (3.31) this is the desired factorization property (3.30), that is the manifestation
of the cluster decomposition principle in the Smatrix theory.
As a matter of example, let us take a four-particle reaction 1234 1 2 3 4 . Let the set of particles 1, 2, 1 , 2
be very far from the set 3, 4, 3 , 4 . In that case, if S
C vanishes when any particles in and/or are far from the

others, the only terms in Eq. (3.38) that survive, are given by (we shorten the notation qi i)

S1 2 3 4 ,1234 S1C 2 ,12 S3C 4 ,34


+ (1 1 2 2 1 2 2 1 ) S3C 4 ,34
+ (3 3 4 4 3 4 4 3 ) S1C 2 ,12
+ (1 1 2 2 1 2 2 1 ) (3 3 4 4 3 4 4 3 ) (3.39)
126 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

as a matter of example the terms



SqC q q q ,q1 q2 q3 q4 ; q1 q1 SqC q q ,q2 q3 q4 permutations (3.40)
1 2 3 4 2 3 4

must vanish when the set of particles 1, 2, 1 , 2 is very far from the set 3, 4, 3 , 4 , because in the first term all
particles are in the same cluster (hence they are not very far from each other) while in the second three of the four
particles are in the same cluster. Comparing (3.39) with (3.36) we observe that it is just the required factorization
condition (3.30)
S1 2 3 4 ,1234 S1 2 ,12 S3 4 ,34
It is clear that the cluster decomposition principle is formulated in spatial coordinates. The meaning of far in
such a principle refers to long distances in the coordinate space. We have in turn reestablished this principle as
the condition that S C vanishes if any particles in the states or are far from any others. In that sense, it

would be convenient to rexpress this in momentum space. The coordinate space matrix elements are defined as a
Fourier transform
Z

SxC x ...,x1 x2 ... d3 p1 d3 p2 d3 p1 d3 p2 SpC p ...,p1 p2 ... eip1 x1 eip2 x2 eip1 x1 eip2 x2 (3.41)
1 2 1 2

where we have temporarily dropped spin and species labels. By now we know the implications of the cluster
decomposition principle in SxC x ...,x1 x2 ... , but SpC p ...,p1 p2 ... is easier to construct and to connect with experiments
1 2 1 2
(experiments usually measure momenta but not position). Thus we want to know what conditions demands the
cluster decomposition principle for SpC p ...,p1 p2 ...
1 2

Let us start by assuming that SpC p ...,p1 p2 ... is well-behaved (Lebesgue-integrable). In that case, the Riemann-
1 2
Lebesgue theorem says that the integral (3.41) would vanish when any combination of spatial coordinates goes
to infinity. This is a very strong requirement, and as we shall see, much stronger than necessary to satisfy the
cluster decomposition principle. Translational invariance says that the connected Smatrix like the Smatrix
itself, can only depend on differences of coordinate vectors. Hence such a matrix should not change when all
xi and xj vary together, as long as their differences remain constant. This requires that the elements of S C in
a momentum basis must (like those of S) be proportional to a three-dimensional delta function that ensures the
momentum conservation5 , as well as the energy-conservation delta function required by scattering theory. Thus,
we can rewrite
 
SpC p ...,p1 p2 ... = 3 p1 + p2 + p1 p2 E1 + E2 + E1 E2 Cp1 p2 ,p1 p2 (3.42)
1 2

On the other hand, the presence of the delta function (which has singular behavior) spoils the Lebesgue-integrability
C
of Sp p ...,p1 p2 ... . However, this is not a problem, since the cluster decomposition principle only requires that
1 2
(3.41) vanishes when the differences among some of the xi and/or xi coordinates become large. However, if C
itself in Eq. (3.42) contained additional delta functions of linear combinations of three-momenta, this principle
would not be satisfied. For example, suppose that C contains a delta function that says that the sum of the
pi and pj for some subset of the particles vanished. In that case Eq. (3.41) would not change if all of the
xi and xj for the particles in that subset moved together (with constant differences) away from all the other xk
and xl , in contradiction with the cluster decomposition principle. Roughly speaking, the cluster decomposition
principle says that the connected part of the Smatrix unlike the Smatrix itself, contains only one momentum
conservation delta function.
In summary, we could say that the coefficient Cp1 p2 ,p1 p2 in Eq. (3.42) should be smooth when it is
considered as a function of its momentum labels. The simplest smoothness requirement would be to demand
that such a coefficient be analytic at p1 = p2 = = p1 = p2 = 0. Such condition would guarantee that
SxC x ,x1 x2 vanishes exponentially fast when any of the x and x is very distant from any of the other x and x .
1 2

5
Recall that translational invariance leads to conservation of the momentum of the whole system.
3.6. STRUCTURE OF THE INTERACTION 127

Nevertheless an exponential decrease of S C is not essential in the cluster decomposition principle. In fact, not all
theories satisfy this requirement of analiticity. For instance, in theories with massless particles, S C can have poles
at certain values of the momenta. In some cases, after fourier transforming, such poles give terms in SxC x ,x1 x2
1 2
that decay only as negative powers of coordinate differences. Indeed it is not necessary to rule out this behavior
to satisfy the cluster decomposition principle. The smoothness condition on S C could allow various poles and
branch-cuts at certain values of p and p . However, strong singularities such as Dirac delta functions should be
discarded.

3.6 Structure of the interaction


The next task is to construct a Hamiltonian that yields an Smatrix that satisfy the cluster decomposition
principle. It is in this point that the creation and annihilation operators become important.
The answer comes from the following theorem: The Smatrix satisfies the cluster decomposition principle if
the Hamiltonian can be expressed as in Eq. (3.20)

Z
X
X   
H= dq1 dqN

dq1 dqM a q1 a qN

a (qM ) a (q1 ) hN M q1 qN

, q1 qM (3.43)
N =0 M =0

in which the coefficients hN M contain just a single three-dimensional momentum-conservation delta function, that
is
 
hN M p1 1 n1 , pN N

nN ; p1 1 n1 , pM M nM = 3 p1 + . . . + pN p1 . . . pM

ehN M p n , p n ; p1 1 n1 , pM M n(3.44)
1 1 1 N N N M

where ehN M does not contain delta function factors. It worths emphasizing that the expression (3.43) is not
enough to guarantee that the Smatrix satisfies the cluster decomposition principle (CDP). Indeed, we have
seen in section 3.3 [see Eq. (3.20)], that any operator can be written according with the structure followed in
Eq. (3.43). It is essential the additional condition (3.44) that shows that the coefficients in the expansion (3.43)
contains only a global delta function.
To prove the theorem, we shall use the perturbation theory in its time-dependent form. An important advan-
tage of the time-dependent perturbation theory (with respect to the old-fashioned perturbation theory) is that the
combinatorics underlying the CDP is more apparent. For instance, if E = E1 + . . . + En is the sum of one-particle
energies, then eiEt is a product of functions of the individual energies while [E E + i]1 is not6 .
In the framework of time-dependent perturbation theory, the Soperator is given by Eq. (2.188)

X Z
(i)n
S= dt1 dt2 dtn T {V (t1 ) V (t2 ) V (tn )}
n!
n=0

where we adopt the convention that for n = 0, the time-ordered product is defined as the identity operator. From
the definition (2.36) of the Soperator, the Smatrix yields
D E X Z D E
(i)n
S : (0) S (0) = dt1 dt2 dtn (0) T {V (t1 ) V (t2 ) V (tn )} (0) (3.45)
n=0
n!

where V (t) is defined as in Eq. (2.180)

V (t) exp [iH0 t] V exp [iH0 t]


6
Recall that old-fashioned perturbation theory is based on iterations coming from expression (2.174), in which the energy appears
in the factor [E E + i]1 . On the other hand, we can pass from the old-fashioned to the time-dependent perturbation theory by
means of Eq. (2.185) in which the RHS depends on exponentials of the energy.
128 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

where H0 is the free-particle Hamiltonian and V is the interaction. It is convenient to express the free-particle
states by showing explicitly their particle content and the state of each particle
E E E
(0)
|q1 q2 qP i ; (0) q1 q2 qQ

(3.46)

On the other hand, according with Eq. (3.8), the free-particle states (0) and (0) can be expressed as a
product of creation operators acting on the vacuum |0i.
E
(0)
|q1 q2 qP i = a (q1 ) a (q2 ) . . . a (qP ) |0i (3.47)
E E      
(0)
q1 q2 qQ
= a q1 a q2 . . . a qQ
|0i

further V (t) as any other operator, can be written as a sum of products of creation and annihilation operators
Z
X
X    
V (tk ) = dq1k dqN
k
dq1k dqM
k
a q1k a qN
k

N =0 M =0
     
k
a qM a q1k CN M q1k qN
k k k
; q1 qM ;t (3.48)

we should emphasize that the coefficients CN M must be explicit functions of time. Thus the inner product in the
integrand of Eq. (3.45) can be written by using Eqs. (3.47, 3.48). The complete expression is very complicated
but the crucial point is that each term in the sum (3.45) could be written as a sum of vacuum expectation values of
products of creation and annihilation operators. Now, by using the commutation and anticommutation relations
(3.17) we can move each annihilation operator to the right passing the creation operators. Each time that we move
an annihilation operator to the right passing a creation operator we obtain two terms, as shown in Eq. (3.17) in
the form   
a q a (q) = a (q) a q + q q (3.49)
moving other creation operator past the annihilation operator in the first term generates yet more terms. However,
Eq. (3.13) shows that any annihilation operator that moves all the way to the right and acts on the vacuum |0i
gives zero, so at the end of the process we obtain only delta functions. In conclusion, the vacuum expectation
value (VEV) of a product of creation and annihilation operators can be written as a sum of different terms, each
term equal to the product of delta functions and signs from the commutators or anticommutators. In turn, it
means that each term in Eq. (3.45) can be expressed by a sum of terms, each term equal to a product of delta
functions and signs from the commutators or anticommutators and whatever factors are contributed by V (t),
integrated over all the times and integrated and sum over the momenta, spins and species in the arguments of the
delta functions.
Each of the terms generated in this way can be symbolized by a diagram. The diagram will be constructed
according with the following algorithm

1. We start by drawing n points called vertices, one for each V (t) operator.

2. For each delta function produced when an annihilation


operator in one of these V (t) operators moves past
a creation operator in the initial state (0) , we draw a line coming into a diagram from below that ends
at the corresponding vertex.

3. For each delta function produced when an annihilation operator in the adjoint of the final state (0) moves
past a creation operator in one of the V (t), draw a line from the corresponding vertex upwards out of the
diagram.

4. For each delta function produced when an annihilation operator in one V (t) moves past a creation in another
V (t), draw a line between the corresponding vertices.
3.6. STRUCTURE OF THE INTERACTION 129


5. For each delta function produced when an annihilation operator in the adjoint of the final state (0)

moves past a creation operator in the initial state (0) , draw a line from bottom to top, right through the
diagrams.

Each of the delta functions associated with one of these lines enforces the equality of the momentum arguments
of the pair of creation and annihilation operators represented by the line. In addition, there is at least one delta
function contributed by each of the vertices, which enforces the conservation of the total three-momentum at the
vertex.

3.6.1 A simple example


As a matter of example, let us assume that both the initial and final states consists of two-particle states. Hence,
equation (3.47) gives
E E E    
(0)
|q1 q2 i = a (q1 ) a (q2 ) |0i ; (0) q1 q2 = a q1 a q2 |0i (3.50)

so that by using equations (3.50) the inner product in (3.45) can be written as
D E    
(n) (0)
D T {V (t1 ) V (t2 ) V (tn )} (0) = h0| a q2 a q1 T {V (t1 ) V (t2 ) V (tn )} a (q1 ) a (q2 ) |0i

(2)
let us examine the term D . For simplicity we assume that t1 > t2
D E    
(2)
D (0) T {V (t1 ) V (t2 )} (0) = h0| a q2 a q1 V (t1 ) V (t2 ) a (q1 ) a (q2 ) |0i

in turn each V (tk ) can be expanded as in Eq. (3.48), then we have


X Z
X  


V (t1 ) V (t2 ) = dq11 dqN
1
1
dq 1
1 dq 1
M1 a q 1
1 a 1
qN 1
N1 =0 M1 =0
    
1 1 1 1 1 1
a qM 1
a q 1 C N1 M1 q 1 q ;
N1 1 q q M1 ; t

X X Z  


dq12 dqN2
2
dq 2
1 dq 2
M2 a q 2
1 a 2
qN 2
N2 =0 M2 =0
    
2 2 2 2 2 2
a qM 2
a q 1 C N2 M2 q 1 q ;
N2 1q q M2 ; t (3.51)

(2)
therefore the inner product D becomes

X
X
X Z
X
(2)
D = dq11 dqN
1
1
dq11 dqM
1
1
dq12 dqN
2
2
dq12 dqM
2
2
N1 =0 M1 =0 N2 =0 M2 =0
   h      i
h0| a q2 a q1 a q11 a qN 1
1
a q 1
M1 a q 1
1
h      i
a q12 a qN
2
2
2
a qM 2
a q12 a (q1 ) a (q2 ) |0i

for instance the term with N1 = M1 = N2 = M2 = 2 reads


Z
dq11 dq21 dq11 dq21 dq12 dq22 dq12 dq22
   h    i h 2  2   i
h0| a q2 a q1 a q11 a q21 a q21 a q11 a q1 a q2 a q22 a q12 a (q1 ) a (q2 ) |0i
130 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

or for N1 = M1 = N2 = M2 = 1 we have
Z    h  i h 2  i
dq11 dq11 dq12 dq12 h0| a q2 a q1 a q11 a q11 a q1 a q12 a (q1 ) a (q2 ) |0i (3.52)

let us concentrate on the vacuum expectation value (VEV) in the term (3.52)
       
h0| A |0i h0| a q2 a q1 a q11 a q11 a q12 a q12 a (q1 ) a (q2 ) |0i (3.53)

we shall pass the term a q12 to the right by succesive use of the commutation or anticommutation relations
       h  i
A a q2 a q1 a q11 a q11 a q12 a q12 a (q1 ) a (q2 )
       h  i
= a q2 a q1 a q11 a q11 a q12 q1 ,q12 a (q1 ) a q12 + q1 q12 a (q2 )
       h  i
= q1 ,q12 a q2 a q1 a q11 a q11 a q12 a (q1 ) a q12 a (q2 )
      h  i
+ q1 q12 a q2 a q1 a q11 a q11 a q12 a (q2 )
       h  i
= q1 ,q12 a q2 a q1 a q11 a q11 a q12 a (q1 ) q2 ,q12 a (q2 ) a q12 + q2 q12
2
     1  h 2
 1
 2 1
i
+ q1 q1 a q2 a q1 a q1 q12 ,q11 a q1 a q1 + q1 q1 a (q2 )

       
A = q1 ,q12 q2 ,q12 a q2 a q1 a q11 a q11 a q12 a (q1 ) a (q2 ) a q12
       
+q1 ,q12 q2 q12 a q2 a q1 a q11 a q11 a q12 a (q1 )
       
+ q1 q12 q12 ,q11 a q2 a q1 a q11 a q12 a q11 a (q2 )
      
+ q1 q12 q12 q11 a q2 a q1 a q11 a (q2 ) (3.54)

however, from Eq. (3.53) we recall that we are interested in the VEV of the operator A. Thus, the first line of
Eq. (3.54) does not contribute because there is an annihilation operator on the right-side. Then we can write
      h  i
A q1 ,q12 q2 q12 a q2 a q1 a q11 a q11 a q12 a (q1 )
      h  i
+ q1 q12 q12 ,q11 a q2 a q1 a q11 a q12 a q11 a (q2 )
    h   1 i
+ q1 q12 q12 q11 a q2 a q1 a q1 a (q2 )

      h   i
A q1 ,q12 q2 q12 a q2 a q1 a q11 q12 ,q11 a q12 a q11 + q12 q11 a (q1 )
      h  i
+ q1 q12 q12 ,q11 a q2 a q1 a q11 a q12 q2 ,q11 a (q2 ) a q11 + q2 q11
    h      i
+ q1 q12 q12 q11 a q2 q1 ,q a q11 a q1 + q1 q11 a (q2 )
1 1

once again we drop the terms with an annihilation operator on the right. Thus
      
A q1 ,q12 q2 q12 q12 q11 a q2 a q1 a q11 a (q1 )
       
+ q1 q12 q2 q11 q12 ,q11 a q2 a q1 a q11 a q12
      
+q1 ,q q1 q12 q12 q11 a q2 a q11 a q1 a (q2 )
1 1
     
+ q1 q12 q12 q11 q1 q11 a q2 a (q2 )
3.6. STRUCTURE OF THE INTERACTION 131

continuing the process, the effective operator that contributes to the VEV is a sum of products of delta functions.
With our notation, it is easy to track the origin of a given delta function. To show it, we remember what momenta
corresponds to the initial state, final state and interactions V (t1 ) and V (t2 ). They can be tracked from Eqs.
(3.50, 3.51)
E E E
(0) (0)

|q1 q2 i ; q1 q2
V (t1 ) q11 , q21 ; q11 , q21
V (t2 ) q12 , q22 ; q12 , q22

Therefore, q2 q12 is a delta function produced when  an annihilation
 operator in V (t2 ) moves past a creation
1
operator in the initial state . The delta factor q1 q1 is produced when an annihilation operator in the
adjoint of the final state moves past a creation operator in V (t1 ). Assuming that the time-ordered operator
gives V (t1 ) V (t2 ), the factor q12 q11 is produced when an annihilation operator in V (t1 ) moves past a creation
operator in V (t2 ).

3.6.2 Connected and disconnected parts of the interaction


A given diagram generated with the previous algorithm may be connected (i.e. every point connected to every
other by a set of lines), and if not connected, it breaks up into a number of connected pieces. The V (t) operator
associated with a vertex in one connected component effectively commutes with the V (t) associated with any
vertex in any other connected component, because for this diagram we are not including any terms in which an
annihilation operator in one vertex destroys a particle that is produced by a creation operator in the other vertex,
if we did, then the two vertices would be in the same connected component. Therefore, the matrix element in Eq.
(3.45) can be expressed as a sum over products of contributions, one from each connected component

D E X D
Y  (0) E
(0) 
(0) T {V (t1 ) V (t2 ) V (tn )} (0) = j T V (tj1 ) V tj,nj j (3.55)
C
clusterings j=1

where the sum is over all ways of splitting up the incoming and outgoing particlesand V (t) operators into
clusters (including a sum over from 1 to n) with nj operators V (tj1 ) V tj,nj and the subsets of initial
particles j and final particles j all in the jth cluster. It is clear that the product over the clusters on the
RHS of Eq. (3.55) must include all the n vertices on the LHS of that equation. It means that

n = n1 + . . . + n

and that is the union of all particles in the subsets 1 , 2 , . . . , , and similarly for the final states.
It may happen that some clusters in (3.55) have no vertices so that nj = 0. For these factors, we must take
the matrix element factor in Eq. (3.55) to vanish unless j and j are both one-particle states, because the only
connected diagrams without vertices are the ones of a single line running through the diagram from bottom to
top. In that case, the matrix element is just the delta function (j j ).
The subscript C in Eq. (3.55) means that we exclude contributions associated with disconnected diagrams,
i.e. contributions in which any V (t) operator or any initial or final particle is not connected to every other by a
sequence of particle creations and annihilations.
Now we substitute Eq. (3.55) into Eq. (3.45). We observe that each time variable is integrated from
to , so it makes no difference which of the ti s are sorted out into each cluster. Consequently, the sum over
clusterings yields a factor
n!
n1 !n2 ! n !
132 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

which is the number of ways of sorting out n vertices into clusters, each one containing n1 , n2 , . . .vertices:
Z D E

I dt1 dt2 dtn (0) T {V (t1 ) V (t2 ) V (tn )} (0)

Z
X X n! Y D
(0)   (0) E
I = dtj1 dtjnj j T V (tj1 ) V tj,nj j (3.56)
part n1 n
n1 !n2 ! n ! C
j=1
with n = n1 + . . . + n (3.57)

where the first sum is over all ways of partitioning the particles in the initial and final states into clusters 1
and 1 including a sum over the number of clusters.
As a matter of example, let us assume that we have an initial state and a final state each one with four
particles, and let us take the number of vertices as n = 4:
E E E
(0)
= |q1 q2 q3 q4 i ; (0) = q1 q2 q3 q4 V (tk ) k = 1, 2, 3, 4
E E
(0) (0)
and let us examine the case in which we have two partitions j and j with j = 1, 2 consisting of
E E E
(0) (0)
first partition 1 = |q1 q2 i , 1 = q1 q2 ; 12 1 2
E E E
(0) (0)
second partition 2 = |q3 q4 i , 2 = q3 q4 ; 34 3 4

in this case we have two clusters = 2 in which n1 = n2 = 2. Hence we have


n! 4!
= =6
n1 !n2 ! 2!2!
the six number of ways we can sort out these 4 vertices into these two clusters is given by
 
V (t1 ) V (t2 ) V (t3 ) V (t4 )
 
V (t1 ) V (t3 ) V (t2 ) V (t4 )
 
V (t1 ) V (t4 ) V (t2 ) V (t3 )
 
V (t2 ) V (t3 ) V (t1 ) V (t4 )
 
V (t2 ) V (t4 ) V (t1 ) V (t3 )
 
V (t3 ) V (t4 ) V (t1 ) V (t2 )

On the other hand, when I of Eq. (3.56) is substituted in Eq. (3.45) the n! term is cancel by the 1/n! term
in such an equation. Besides, owing to the constraint (3.57) the factor (i)n in Eq. (3.45) can be expressed as

(i)n = (i)n1 (i)n2 (i)n

from these facts instead of summing over n and then summing over n1 , n2 , . . . , n constrained by (3.57), we can
just sum independently over each n1 , n2 , . . . , n . It yields
Z
X X
Y
(i)nj D
(0)   (0) E
S = dtj1 dtjnj j T V (tj1 ) V tj,nj j (3.58)
n =0
nj ! C
partitions j=1 j

and comparing Eq. (3.58) with the definition (3.31) of the connected matrix elements S C , we see that such matrix

elements are given by the factors in the product of Eq. (3.58)


X Z D E
C (i)n
S = dt1 dtn (0) T {V (t1 ) V (tn )} (0)
n=0
n! C
3.6. STRUCTURE OF THE INTERACTION 133

in conclusion the matrix elements S C are calculated with a simple prescription: S C is the sum of all contributions

to the Smatrix that are connected, it means that we drop all terms in which any initial or final state or any
operator V (t) is not connected to all the others by a sequence of particle creations and annihilations. It explains
the adjective connected for S C .
We have seen that momentum is conserved at each vertex and along every line. Hence, the connected parts
of the Smatrix individually conserve momentum. Therefore, S C contains a factor 3 (p p ). We shall prove

that SC contains no other delta functions, such that the CDP holds.

Now we start with the hypothesis that the coefficients hN M in the expansion (3.43) of the Hamiltonian in terms
of creation and annihilation operators are proportional to a single three-dimensional delta function that provides
momenta conservation. This is automatically true for the free-particle Hamiltonian H0 and using our hypothesis,
it will be true separately for the interaction V . In our graphical interpretation of the matrix elements, it means
that each vertex contributes one three-dimensional delta function. The other delta functions in matrix elements
V just keep unchanged the momentum of any particle that is not created or annihilated at the corresponding
vertex.
Most of these delta functions just fix the momentum of intermediate particles. The only momenta that are
left unfixed by the delta functions are the ones that circulate in loops of internal lines. Observe that any line
that if cut leaves the diagram disconnected carries a momentum that is fixed by momentum conservation as some
linear combination of the momenta of the lines coming into or going out of the diagram. If the diagram contains
L lines that can all be cut at the same time without the diagram becoming disconnected, we say that it has
L independent loops, and there are L momenta that are not fixed by the neither the delta functions nor the
momentum conservation.
With V vertices, I internal lines, and L loops there are V delta functions: I L delta functions fix internal
momenta, leaving V I + L delta functions relating the momenta of incoming or outgoing particles. From a
well-known topological identity, we know that for every graph consisting of C connected pieces, the number of
vertices V , internal lines I, and loops L, we have the relation

V I +L=C (3.59)
C , that arises from graphs with C = 1, we have a single
in the particular case of a connected matrix element like S
three-dimensional delta function 3 (p p ), as we wanted to prove.
We can see the identity (3.59) by a heuristical argument as follows: A graph with a single vertex has VC =
1, LC = 0, C = 1. If we add V 1 vertices with just enough internal lines to keep the graph connected, we have
IC = VC 1, LC = 0 and C = 1. Any additional internal lines attached (without new vertices) to the same
connected graph produce an equal number of loops. Hence, IC = VC + LC 1 and C = 1. If a disconnected graph
consists of C connected parts, the sums of IC , VC , LC in each connected part will then satisfy the equation
X X X
IC = VC + LC C
C C C

where the sum is over each connected part. It leads to Eq. (3.59).
In the above argument it was not important that the time variables were integrated from to +. Conse-
quently, we can use the same argument to show that if the coefficients hN M in the Hamiltonian contain just single
delta functions, U (t, t0 ) can also be decomposed into connected parts, each one containing a single momentum-
conservation delta function factor. In addition, the connected part of the Smatrix also contains an energy-
conservation delta function, we shall see later that SC contains only a single energy-conservation delta fucntion

factor (E E ), while U (t, t0 ) does not contain any energy conservation delta function at all.

3.6.3 Some examples of the diagrammatic properties


In Fig. 3.1a,b we show how the lines entering into a vertex are showing a flux of momentum that must
be conserved. We could choose the convention of assuming that a momentum flux entering is positive and an
134 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

Figure 3.1: (a) Vertex with three incoming momenta. (b) Vertex with one incoming momentum and two outgoing
momenta. (c) A disconnected diagram consisting of two connected pieces.

Figure 3.2: (a) A diagram with one loop. (b) The previous diagram could be drawn with the same momentum in
the incoming and in the outgoing external lines. The momenta k within the loop is totally indetermined. (c) A
diagram with two loops.

outgoing flux of momentum is negative. Thus in Fig. 3.1a the conservation of momentum in the vertex gives

p1 + p2 + p3 = 0

while in the diagram of Fig. 3.1b, the conservation of momentum is expressed by

p1 p2 p3 = 0

in these figures the line with momentum p1 comes from a delta function produced when an annihilation operator
in one of the V (tk ) operators moves past a creation operator in the initial state . The line with p2 (and also
the line with p3 ), comes from a delta function produced when an annihilation operator in the adjoint of the final
state moves past a creation operator in one of the V (tk ).
In the diagram of Fig. 3.1c, the line carrying momentum q comes from a delta function produced when an
annihilation operator in one V (tk ) moves past a creation operator in another V (tm ). Moreover, the line carrying
momentum p5 comes from a delta function produced when an annihilation operator in the adjoint state moves
past a creation operator in the initial state . In Fig. 3.1c, the whole diagram is disconnected, and consists of
two connected pieces.
Figure 3.2a, shows a diagram that contains a loop, which is formed by the two internal lines with momenta
p1 + k and k. Observe that we can cut one of these internal lines, and the diagram remains connected. Since we
have two vertices, we have two delta functions associated with each vertex, they are given by

(p1 (p1 + k) + k) ; ((p1 + k) k p2 )

we can observe that the first delta is trivial, and says nothing about the momentum k. The another delta function
simply says that p1 = p2 , and says nothing about k. Therefore, in this diagram containing one loop, the
momentum k of one of the internal lines in the loop remains totally indetermined. It is easy to generalize the fact
3.6. STRUCTURE OF THE INTERACTION 135

Figure 3.3: (a) A diagram with two loops, there is one momenta within each loop (k1 and k2 ) that are totally
independent. (b) The previous diagram could be drawn with the same momentum for each internal line that is
not part of a loop.

that there are L totally indetermined momenta for a diagram with L loops. In practice, it is usual in this diagram
to put p1 in both external lines in order to simplify calculations, it is shown in Fig. 3.2b. Moreover, in Fig. 3.2b,
the diagram has V = 2, I = 2, L = 1 then applying Eq. (3.59) we have

V I +L = 22+1= 1

showing that C = 1, i.e. that it is a connected diagram.


In Fig. 3.2 we have V = 5, L = 2, I = 6. In particular, we observe that four internal lines are part of the
loops while two internal lines do not form a loop. By using Eq. (3.59), we clearly have

V I +L = 56+2= 1

so that C = 1 showing that it is a connected diagram.


As a last example, in the diagram of Fig. 3.3a, we have V = 6, I = 7, L = 2. We have six delta functions
owing to the six vertices, they are

(p1 + p2 q1 ) , (q1 (q1 + k1 ) + k1 ) , ((q1 + k1 ) k1 q2 )


(q2 + k2 (q2 + k2 )) , (q3 k2 + (q2 + k2 )) ; (q3 p3 p4 ) (3.60)

from the V = 6 delta functions, I L = 5 of them fix internal momenta since L = 2 momenta are totally
indetermined (k1 and k2 in this case). Thus, we have 5 delta functions that determine internal momenta. Finally,
V (I L) = 1 delta function is related with external momenta. Then, we only have one global delta function
that contains non-trivial information. Effectively, the six delta functions (3.60) says that

q1 = p1 + p2 , q1 = q2 , q2 = q3 , q3 = p3 + p4

or equivalently
p1 + p2 = q1 = q2 = q3 = p3 + p4
136 CHAPTER 3. THE CLUSTER DECOMPOSITION PRINCIPLE

thus all six deltas are equivalent to a single delta function


(p1 + p2 p3 p4 )
Owing to this fact, these kind of diagrams are usually written only in terms of the independent momenta as shown
in Fig. 3.3b.

3.6.4 Implications of the theorem


The fact that hN M in Eq. (3.43) should have only one three-dimensional momentum conservation delta function
factor is far from trivial and has deep implications. For example, let us assume that V has non-vanishing matrix
elements between two-particle states. Thus, Eq. (3.43) applied for the interaction
X X Z
  
V = dq1 dqN

dq1 dqM a q1 a qN

a (qM ) a (q1 ) vN M q1 qN

, q1 qM
N =0 M =0

must contain a term with N = M = 2, i.e. a term with coefficient v2,2 (p1 p2 , p1 p2 ) associated with two creation
and two annihilation operators. Therefore, such a coefficient is associated with two initial paritcle states and two
final particle states 
v2,2 p1 p2 , p1 p2 = Vp1 p2 ,p1 p2 (3.61)
where we have dropped the spin and species labels. But then the matrix element of the interaction between
three-particle states is7
  
Vp1 p2 p3 ,p1 p2 p3 = v3,3 p1 p2 p3 , p1 p2 p3 + v2,2 p1 p2 , p1 p2 3 p3 p3 permutations (3.62)
For instance, we could try to do a relativistic quantum theory that is not a field theory, by choosing v2,2 in
such a way that the two-body Smatrix is Lorentz invariant, and adjusting the rest of the Hamiltonian so that
there is no scattering in states containing three or more particles. We would then have to take v3,3 to cancel the
other terms in Eq. (3.62)
  
v3,3 p1 p2 p3 , p1 p2 p3 = v2,2 p1 p2 , p1 p2 3 p3 p3 permutations (3.63)
nevertheless, recalling that v2,2 (p1 p2 , p1 p2 ) has a factor 3 (p1 + p2 p1 p2 ), Eq. (3.63) means that each term
in v3,3 contains two delta function factors, violating the CDP. Therefore, in a theory that satisfies the CDP,
the existence of scattering processes involving two particles makes processes involving three or more particles
inevitable.
To solve problems involving three-bodies in quantum theories that satisfy the CDP, the term v3,3 in Eq.
(3.62) has no particular problems, but the extra delta function in the other terms make the Lippmann-Schwinger
equation difficult to solve directly. The problem is that these delta functions makes the kernel [E E + i]1 V
of this equation not square-integrable, even after factorizing out an overall momentum conservation delta function.
Hence, it cannot be approximated by a finite matrix, even of very large rank. Therefore, to solve problems with
three or more particles, we should replace the Lippmann-Schwinger equation with one that has a connected RHS.
Such equations have been developed and solved recursively but in a non-relativistic regime, while they have not
been succesful in relativistic theories.
The proof of the main theorem of this section is heavily relied on perturbation theory. However, it has been
shown that the reformulated Lippmann-Schwinger equations with non-perturbative dynamics are consistent with
the requirement that U C (t, t0 ) (and so S C ) should also contain only one momentum-conservation delta fucntion
factor, as required by the CDP, as long as the Hamiltonian satisfies the condition that the coefficient functions
hN M each contain a single momentum-conservation delta function.
7
In Eq. (3.62) we are omitting the possibility of having three clusters each of one-particle states, since they are in the free-particle
hamiltonian H0 and not in the interaction term V . Similarly, in Eq. (3.61) we do not include in the interaction term Vp1 p2 ,p1 p2
the case of two clusters each consisting of one-particle states, since such a term would be in the free-Hamiltonina H0 and not in the
interaction term V .
Chapter 4

Relativistic quantum field theory

In this chapter we shall show the necessity of the introduction of fields to unite succesfully the special relativity
with quantum mechanics while maintaining the Cluster Decomposition Principle (CDP). Further we shall discuss
some arising properties of the development of a relativistic quantum field theory: the connection between spin and
statistics, the existence of antiparticles, and several relationships between the particles and antiparticles being the
most remarkable the so-called CPT theorem.

4.1 Free fields


We saw in section 2.6 that the Smatrix is Lorentz invariant if the interaction can be written as
Z
V (t) = d3 x H (x, t) (4.1)

where H (x, t) is a scalar in the sense that

U0 (, a) H (x) U01 (, a) = H (x + a) (4.2)

and satisfying the conditions   2


H (x) , H x =0 for x x 0 (4.3)
we shall see later that the possibilities are more general than (4.3) but they are not very different from it. It is
also interesting to ask whether in Eq. (4.2) must be restricted to proper orthochronus Lorentz transformations,
or can also include discrete space inversions. In order to satisfy the CDP we shall express H (x) in terms of
creation and annihilation operators. Nevertheless, in doing this we shall confront the following problem: equation
(3.26) shows that under Lorentz transformations each creation or annihilation operator is multiplied by a matrix
that depends on the momentum carried by the operator. We should then find the way to couple such operators
together to form a scalar1 . The solution is to build H (x) based on fields. For this, we build annihilation fields
k+ (x) and creation fields k (x) as follows:
XZ Z
+
k (x) = d p uk (x; p, , n) a (p, , n) d3 q uk (x; q) a (q)
3
(4.4)
n
XZ Z
k (x) = 3
d p vk (x; p, , n) a (p, , n) d3 q vk (x; q) a (q) (4.5)
n

in other words, we avoid the problem of momentum dependence of the Lorentz transformation of annihilation and
creation operators, by integrating out all those momenta (and also the spin and species degrees of freedom). In
1
It is clear that condition (4.2) for H (x) to be a scalar, requires that the Lorentz transformation for H (x) be independent of any
momenta.

137
138 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

addition, the coefficients uk (x; p, , n) and vk (x; p, , n) are chosen such that under Lorentz transformations each
creation and annihilation field is multiplied by a position-independent matrix2 :
X 
U0 (, a) k+ (x) U01 (, a) = Dkk 1 k+ (x + a) (4.6)
k
X 
U0 (, a) k (x) U01 (, a) = Dkk 1 k (x + a) (4.7)
k

In principle, we could have different transformation matrices D for the annihilation and creation fields, but we
will see that it is always possible to choose the fields such that both matrices are equal.
By applying a second (inhomogeneous) Lorentz transformation , a, the total Lorentz transformation becomes
 
U0 , a U0 (, a) = U0 , a + a (4.8)

for the Lorentz transformation written as in the LHS of Eq. (4.8), we find from Eq. (4.6) that
     1
ek+ (x) U0 , a U0 (, a) k+ (x) U0 , a U0 (, a)
  
= U0 , a U0 (, a) k+ (x) U01 (, a) U01 , a (4.9)

 X  
= U0 , a Dkk 1 k+ (x + a) U01 , a
k
X  h  i
= Dkk 1 U0 , a k+ (x + a) U01 , a (4.10)
k
" #
X  X  + 
= Dkk 1 Dkm 1 m (x + a) + a
k m

X X   
= Dkk 1 Dkm 1 m
+
(x + a) + a (4.11)
m k
X   
ek+ (x) = D 1 D 1 km m
+
(x + a) + a (4.12)
m

and for the Lorentz transformation written as in the RHS of Eq. (4.8), we find from Eq. (4.6) that
  X  1  +  
ek+ (x) U0 , a + a k+ (x) U01 , a + a = Dkk k x + a + a
k
X  
ek+ (x) = Dkk 1 1
k+ (x + a) + a (4.13)
k

thus, equating equations (4.12, 4.13), we find that


   1 
D 1 D 1 = D (4.14)

and defining 1 1 and 2 1 we obtain

D (1 ) D (2 ) = D (1 2 ) (4.15)

so that Dmatrices provide a representation of the homogeneous Lorentz group. There are many representations
starting with the trivial representation D () = 1, the vector representation D () = , and a host of tensor
2
In other words, we demand that the Lorentz transformation of the creation and annihilation fields, be homogeneous in space-time.
4.2. LORENTZ TRANSFORMATIONS FOR MASSIVE FIELDS 139

and spinor representations. The representations mentioned above are irreducible. Notwithstanding, we do not
require now that the Dmatrix representation be irreducible. In general, in certain basis (the canonical basis)
it is a set of block-diagonal matrices, with an arbitrary array of irreducible representations in the blocks. The
index k here includes a label that runs over the types of particles described and the irreducible representations in
the different blocks, and also another that runs over the components of the individual irreducible representations.
We shall separate these fields later into irreducible fields each one describing a single particle species (and its
antiparticle), and transforming irreducibly under the Lorentz group.
Once we learn the way to construct fields satisfying the Lorentz transformation rules (4.6) and (4.7), we can
construct the interaction density as
X X X
H (x) = gk1 kN
, k k
1 M
k (x) k (x) k+1 (x) k+M (x) (4.16)
1 N
M N k1 kN
k k
1 M

where Eq. (4.16) is the analogous of expansion (3.43) but based on creation and annihilation fields instead of
creation and annihilation operators. The integrations over momenta, spins and species of Eq. (3.43) do not
appear explicitly in Eq. (4.16) because we have already carried out such an integration in defining the creation
and annihilation fields as can be seen in Eqs. (4.4).
The interaction density (4.16) will be a scalar in the sense of Eq. (4.2) if the constant coefficients gk1 kN
, k k
1 M
are chosen to be Lorentz covariant, in the sense that for all :
X X    
gk k , k1 kM = Dk k 1 Dk k 1 Dk1 k1 1 DkM kM 1 gk1 kN , k k
1 M
(4.17)
1 N 1 1 N N
k1 kN
k k
1 M

in which we have not included derivatives because we regard the derivatives of components of these fields as
simply additional sorts of field components. The task of finding coefficients gk1 kN
, k k
1 M
that satisfy Eq. (4.17)
is similar to the task of obtaining the Clebsch-Gordan coefficients to couple together several representations of
SO (3) to form rotational scalars.
Another important task is to set up the interaction density so that it satisfies (4.3). We shall later combine
creation and annihilation operators in such a way that this density commutes with itself at space-like and light-like
separations.

4.2 Lorentz transformations for massive fields


By now we shall study the case of massive particles. In order to obtain the coefficient functions uk (x; p, , n) and
vk (x; p, , n), we observe that Eq. (3.26), gives the transformation rules for the creation operators.
s
(p)0 X (jn )
U0 (, b) a (p,, n) U01 (, b) = exp [i (p) b] D (W (, p)) a (p n)
p0

where jn is the spin of particles of species n, and p is the three-vector part of p. Using the unitarity of the
(j )
rotation matrices Dn this equation becomes
s
(p)0 X (jn ) 
U0 (, b) a (p,, n) U01 (, b) = exp [i (p) b] D W 1
(, p) a (p n) (4.18)
p0

and taking the adjoint of (4.18) we obtain the transformation rule of the annihilation operators
s
(p)0 X (jn ) 
U0 (, b) a (p,, n) U01 (, b) = exp [i (p) b] 0
D W 1 (, p) a (p n) (4.19)
p
140 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

We also saw in Eq. (1.157), page 37 that the volume element d3 p/p0 is Lorentz invariant. Consequently, we have

d3 p d3 (p) d3 (p)
= ; d3 p = p0 (4.20)
p0 (p)0 (p)0

replacing d3 p by the expression (4.20) in Eqs. (4.4), we find

XZ d3 (p)
k+ (x) = p0 uk (x; p, , n) a (p, , n) (4.21)
n (p)0

from Eqs. (4.21, 4.19) the Lorentz transformation of k+ (x) gives

XZ
d3 (p)  
ek+ (x) U0 (, b) k+ (x) U01 (, b) = p0
0 uk (x; p, , n) U0 (, b) a (p, , n) U01 (, b)
n (p)
s
XZ d3 (p) (p) 0 X (j ) 
= p0 0 uk (x; p, , n)
exp [i (p) b]
0
Dn W 1 (, p) a (p n)
n (p) p

obtaining finally
X Z
U0 (, b) k+ (x) U01 (, b) = d3 (p) uk (x; p, , n) exp [i (p) b]
n
s )
(j )  p0
Dn W 1 (, p) a (p , , n) (4.22)
(p)0

a similar exercise for Eq. (4.5) yields


X Z
U0 (, b) k (x) U01 (, b) = d3 (p) vk (x; p, , n) exp [i (p) b]
n
s )
(j )  p0
Dn W 1 (, p) a (p , , n) (4.23)
(p)0

By comparing Eq. (4.6) with Eq. (4.22), we have


X  X Z
Dkk 1 k+ (x + b) = d3 (p) uk (x; p, , n) exp [i (p) b]
k n
s )
(j ) 1
 p0
Dn W (, p) a (p , , n) (4.24)
(p)0

and substituting (4.4) in (4.24) we have


" #
X  XZ X Z
1 3
Dkk d p uk (x + b; p , , n) a (p , , n) = d3 (p) uk (x; p, , n) exp [i (p) b]
k n n
s )
(j ) 1
 p0
Dn W (, p) a (p , , n)
(p)0
(4.25)
4.2. LORENTZ TRANSFORMATIONS FOR MASSIVE FIELDS 141

and using Eq. (4.20) on the LHS of Eq. (4.25) we have


" #
X  XZ d3 (p)
Dkk 1 p0 uk (x + b; p , , n) a (p , , n)
n (p)0
k
(Z s )
X (j )  p0
= d3 (p) uk (x; p, , n) exp [i (p) b] Dn W 1 (, p) a (p , , n) (4.26)
n (p)0
reorganizing the series and integrals in both sides of Eq. (4.26) we have
s s
XZ p 0 p 0 X 
3 1
d (p) a (p , , n) Dk k uk (x + b; p , , n)
n (p)0 (p)0
k
s ( )
XZ p0 3 X (j ) 
= 0 d (p) a (p , , n) uk (x; p, , n) exp [i (p) b] Dn W 1 (, p) (4.27)
n (p)
since this relation must hold for arbitrary x, and b, the terms in the brackets on both sides of Eq. (4.27) must
be equal, hence
s
p0 X  X (j ) 
0 Dkk 1 uk (x + b; p , , n) = uk (x; p, , n) exp [i (p) b] Dn W 1 (, p)
(p)
k

???therefore, we can see that in order that the field k+ (x) satisfies the Lorentz transformation rule (4.6), it is
necessary and sufficient that
s
X  p0 X (jn ) 
Dkk 1 uk (x + b; p , , n) = 0 D W 1 (, p) exp [+i (p) b] uk (x; p, , n) (4.28)
(p)
k

Similarly, by comparing Eqs. (4.7, 4.23), the necessary and sufficient condition for the field k (x) to satisfy the
Lorentz transformation rule (4.7), yields
s
X  p0 X (jn ) 
Dkk 1 vk (x + b; p , , n) = 0 D W 1 (, p) exp [i (p) b] vk (x; p, , n) (4.29)
(p)
k
 
we can put Eq. (4.28) in a slightly different form by using the fact that Dkk 1 = Dk1k
() and D W 1 =
1
D (W ). Multiplying Eq. (4.28) by Dk k () and summing over k we find
s
XX  X p0
Dk k () Dkk 1 uk (x + b; p , , n) = Dk k ()
k k
(p)0
k
X (j ) 
Dn W 1 (, p) exp [+i (p) b] uk (x; p, , n)

s
X X p0 X (jn ) 
1
k k uk (x + b; p , , n) = Dk k () D W (, p) exp [+i (p) b] uk (x; p, , n)
k
(p)0
k
X (j ) 
uk (x + b; p , , n) = Dn W 1 (, p)

s
X p0
Dk k () exp [+i (p) b] uk (x; p, , n) (4.30)
k
(p)0
142 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

(j )
similarly, when multiplying Eq. (4.30) by Dn (W (, p)) and summing over we have
X (j )
XX (j ) (j ) 
Dn (W (, p)) uk (x + b; p , , n) = Dn (W (, p)) Dn W 1 (, p)

s
X p0
Dkk () exp [+i (p) b] uk (x; p, , n)
k
(p)0

s
X (j )
X X p0
Dn (W (, p)) uk (x + b; p , , n) = Dkk () exp [+i (p) b] uk (x; p, , n)
k
(p)0
s
X (j )
X p0 

Dn (W (, p)) uk (x + b; p , , n) = Dkk () 0 exp [+i (p) b] uk x; p, , n
k
(p)

renaming and we obtain


s
X (j )
X p0
Dn (W (, p)) uk (x + b; p , , n) = Dkk () exp [+i (p) b] uk (x; p, , n)
k
(p)0

???therefore, in a slightly different form, Eq. (4.28) becomes


s
X (j ) p0 X
uk (x + b; p , , n) Dn (W (, p)) = Dkk () exp [i (p) b] uk (x; p, , n) (4.31)
(p)0 k

with a similar procedure Eq. (4.29), becomes


s
X (j ) p0 X
vk (x + b; p , , n) Dn (W (, p)) = Dkk () exp [i (p) b] vk (x; p, , n) (4.32)
(p)0 k

[homework!!(12) arrive to Eqs. (4.28), (4.32) with a correct procedure]The advantage of Eqs. (4.31, 4.32) with
respect to Eqs. (4.28, 4.29) is that the former are in terms of the Lorentz transformations instead of their inverses.
In summary Eqs. (4.31, 4.32), are the fundamental requirements that allow to calculate the coefficient functions
uk and vk , in terms of finite number of free parameters.
We shall use Eqs. (4.31) and (4.32) in three steps, considering in turn the three different types of proper
orthochronus Lorentz transformations.

4.2.1 Translations
We shall start by studying Eqs. (4.31, 4.32) for the case of U (1, b) corresponding with pure translations. By
setting = 1 and b arbitrary in Eq. (4.31) we find
s
X (jn ) p0 X
uk (x + b; p, , n) D (W (1, p)) = Dkk (1) exp [ip b] uk (x; p, , n) (4.33)

p0
k

from Eq. (1.181), page 41 we see that

W (, p) = L1 (p) L (p) W (1, p) = L1 (p) 1 L (p)


W (1, p) = 1 (4.34)
4.2. LORENTZ TRANSFORMATIONS FOR MASSIVE FIELDS 143

Then Eq. (4.33) becomes


uk (x + b; p, , n) = exp [ip b] uk (x; p, , n) (4.35)
observe that the traslation in x of the coefficient uk is carried out by a factor exp [ip b] which resembles the
translation operator with generator p and parameter b. Now, we define the standard coefficient uk (p, , n) in
the form
uk (x = 0; p, , n) (2)3/2 uk (p, , n) (4.36)
where the factor (2)3/2 is defined for convenience. From this definition, we obtain by setting x = 0 in Eq. (4.35)
that
uk (b; p, , n) = (2)3/2 exp [ip b] uk (p, , n)
the procedure for vk (x; p, , n) from Eq. (4.32) is similar. Since b is arbitrary, the coefficients uk (x; p, , n) and
vk (x; p, , n) acquire the form

uk (x; p, , n) = (2)3/2 eipx uk (p, , n) (4.37)


3/2 ipx
vk (x; p, , n) = (2) e vk (p, , n) (4.38)

From which we have obtained the xdependence of the coefficients uk and vk . Substituting (4.37, 4.38) in Eqs.
(4.4, 4.5) we see that the annihilation and creation fields are their Fourier transforms
X Z
+ 3/2
k (x) = (2) d3 p uk (p, , n) eipx a (p, , n) (4.39)
,n
X Z
3/2
k (x) = (2) d3 p vk (p, , n) eipx a (p, , n) (4.40)
,n

At this step we understand the convenience of defining the factor (2)3/2 in Eq. (4.36).
On the other hand, equations (4.37, 4.38) are valid for arbitrary values of p. Consequently, if it is valid for p
it is also valid for p , where p is the three-vector associated with p. In other words, Eqs. (4.37, 4.38) must
hold for an arbitrary Lorentz transformation. Thus, taking Eq. (4.31)
s
X (jn ) p0 X
uk (x + b; p , , n) D (W (, p)) = Dkk () exp [i (p) b] uk (x; p, , n) (4.41)
(p)0 k

and substituting (4.37) in (4.41) we obtain

s
X (j ) p0 X
ei(p)(x+b) uk (p , , n) Dn (W (, p)) = Dkk () ei[(p)b] eipx uk (p, , n)
(p)0 k
s
X (j ) p0 X
ei(p)(x+b) uk (p , , n) Dn (W (, p)) = Dkk () ei[(p)b+px] uk (p, , n) (4.42)
(p)0 k

now, since the dot product of two four-vectors is Lorentz invariant, we have

(p) b + p x = (p) b + (p) (x) = (p) (x + b)

therefore, the phases on both sides of Eq. (4.42) are equal and we obtain
s
X (jn ) p0 X
uk (p , , n) D (W (, p)) = Dkk () uk (p, , n)
(p)0 k
144 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

a similar result is obtained for the coefficients vk (p, , n) by substituting (4.38) in Eq. (4.32).
In conclusion, from Eqs. (4.37, 4.38) we see that Eqs. (4.31, 4.32) are satisfied if and only if
s
X (j ) p0 X
uk (p , , n) Dn (W (, p)) = Dkk () uk (p, , n) (4.43)
(p)0 k
s
X (jn ) p0 X
vk (p , , n) D (W (, p)) = Dkk () vk (p, , n) (4.44)
(p)0 k

Equations (4.43, 4.44) must be valid for arbitrary homogeneous Lorentz transformations. Note that with respect to
Eqs. (4.31, 4.32), equations (4.43, 4.44) are for the standard coefficients defined in (4.36), so that the xdependence
has dissapeared.

4.2.2 Boosts
Now, let us take p = 0 in Eqs. (4.43) and (4.44) and use as the
 standard boost L (q) that takes a particle of
0
mass m from rest to a given four-momentum q . For p = p, p = (0, m), it is clear that L (p) = 1, and from
definition (1.181), page 41 we have

W (, p) L1 (p) L (p) = L1 (q) L (q) 1 = L1 (q) L (q) = 1

for this special case, it is also clear that p = q. Using all these facts in Eq. (4.43) we find
X r X
(jn ) m
uk (q, , n) D (1) = Dkk (L (q)) uk (0, , n)

q0
k
X r X
m
uk (q, , n) = 0
Dkk (L (q)) uk (0, , n)

q
k

and the same replacements can be done in Eq. (4.44). Therefore, in this special case, Eqs. (4.43) and (4.44) yield
r
m X
uk (q, , n) = Dkk (L (q)) uk (0, , n) (4.45)
q0
k
r
m X
vk (q, , n) = Dkk (L (q)) vk (0, , n) (4.46)
q0
k

hence, if we know the quantities uk (0, , n) and vk (0, , n) for zero momentum, we can obtain the correspond-
ing quantities uk (p, , n) and vk (p, , n) for arbitrary momentum p, for a given representation D () of the
homogeneous Lorentz group.

4.2.3 Rotations
Now we take again p = 0, but with being a rotation R. Since rotations are Lorentz transformations that
preserve the norm of the three-momentum, we have p = 0. Here it is clear that

p = p Rp = (0, m)

In addition, since the standard boost L (p) takes a particle from rest to its final state also at rest, we have L (p) = 1.
Then we have

W (, p) W (R, p) L1 (Rp) R L (p) = L1 (p) R L (p)


W (, p) = R
4.3. IMPLEMENTATION OF THE CLUSTER DECOMPOSITION PRINCIPLE 145

From these considerations Eq. (4.43), yields


r
X (j ) mX
uk (0, , n) Dn (R) = Dkk (R) uk (0, , n)

m
k

and a similar procedure can be carried out for Eq. (4.44).


In summary, Eqs. (4.43) and (4.44) give
X (j )
X
uk (0, , n) Dn (R) = Dkk (R) uk (0, , n) (4.47)
k
X (j )
X
vk (0, , n) Dn (R) = Dkk (R) vk (0, , n) (4.48)
k

further, by applying Eqs. (4.47, 4.48) to infinitesimal rotations, such relations can be written in terms of the
generators of the rotation group (which are independent of the specific rotation R i.e. of the parameters)
X (j )
X
uk (0, , n) Jn = Jkk uk (0, , n) (4.49)
k
X (j )
X
vk (0, , n) Jn = Jkk vk (0, , n) (4.50)
k

where J(j) and J are the angular-momentum matrices in the representations D (j) (R) and D (R), respectively. It
is clear that any representation D () of the homogeneous Lorentz group, provides a representation of the three-
dimensional rotation group when is restricted to rotations R. Equations (4.49) and (4.50) say that if the fields
(x) described particles of a given spin j, then the representation D (R) must contain among its irreducible
components the spin -j representation D (j) (R), where the coefficients uk (0, , n) and vk (0, , n) describes how
the spinj representation of the rotation group is embedded in D (R). We shall see later that each irreducible
representation of the proper orthochronus Lorentz group contains any given irreducible representation of the
rotation group at least once. Therefore, if the fields k+ (x) and k (x) transform irreducibly (so that Dkk (R)
corresponds to an irreducible representation) then they are unique up to overall scale. More generally, the number
of parameters in the annihilation or creation fields (including their overall scales) is equal to the number of
irreducible representations in the field.
It can be shown that coefficient functions uk (p, , n) and vk (p, , n) given by Eqs. (4.45) and (4.46), with
uk (0, , n) and vk (0, , n) satisfying Eqs. (4.47) and (4.48), automatically satisfy the more general requirements
(4.43) and (4.44).

4.3 Implementation of the cluster decomposition principle


Substituting Eqs. (4.39, 4.40) in Eq. (4.16) and integrating over x, the interaction potential becomes [home-
work!!(13), prove Eqs. (4.51, 4.52, 4.53)]
XZ X X X X
V = d3 p1 d3 pN d3 p1 d3 pM
NM 1 M n n n1 nM
1 N 1 N
 
a p1 1 n1 a pN N

nN a (pM M nM ) a (p1 1 n1 )

VN M p1 1 n1 pN N

nN , p1 1 n1 pM M nM (4.51)

where the coefficient functions are given by


 
VN M p1 1 n1 pN N

nN , p1 1 n1 pM M nM = 3 p1 + . . . p1 . . . (4.52)

eN M p1 1 n1 pN N
V
nN , p1 1 n1 pM M nM
146 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

with 
eN M p1 1 n1 pN N
V
nN , p1 1 n1 pM M nM = (2)33N/23M/2
X X  
gk1 kN
,k k vk
1 M 1
p1 1 n1 vkN
pN N
nN uk1 (p1 1 n1 ) ukM (pM M nM ) (4.53)
k1 kN
k1 kM

the form of the interaction guarantees that the Cluster Decomposition Principle (CDP) is satisfied by the
Smatrix, since VN M has a single delta function factor, with a coefficient V eN M that at least for a finite number
of field types, has at most branch point singularities at zero particle momenta. We could say conversely that any
operator can be written as in Eq. (4.51) and the CDP demands that the coefficient VN M may be written as in
Eq. (4.52), i.e. as a product of a single momentum-conservation delta function times a smooth coefficient function
eN M . Any sufficiently smooth function (but not one that contains a delta function) can be expressed as in Eq.
V
(4.53). For general functions the indices k and k can have infinite range. However, we shall restrict k and k to a
finite range because of the principle of renormalizability to be discussed later.
We could summarize our results by saying that the CDP along with Lorentz invariance makes it natural that
the interaction density should be constructed out of the annihilation and creation operators.

4.4 Lorentz invariance of the Smatrix


By combining annihilation and creation operators in arbitrary polynomials (4.16), where the coupling coefficients
gk1 kN
,k k
1 M
are subjected only to the invariance condition (4.17), and a suitable reality condition (for H (x) to
be hermitian), we construct a scalar density that satisfies the CDP.
Now, for the Lorentz invariance of the Smatrix, it is also necessary that the interaction density satisfies the
commutation relation (4.3).
  2
H (x) , H x = 0 for x x 0 (4.54)
To check the conditions to obtain (4.54) we start by calculating the commutation and anti-commutation relations
between the creation and annihilation fields
h i
I k+ (x) , k (y)

where [. . .] and [. . .]+ denotes commutation and anticommutation respectively. We then use Eqs. (4.39, 4.40)
and the commutation and anti-comutation relations between creation and annihilation operators (3.17)
" Z Z #
X 3/2
X 3/2  ip y 
3 ipx 3
I = (2) d p uk (p, , n) e a (p, , n) , (2) d p vk p , , n e a p , , n
,n ,n
Z Z h
XX 3 3 ipx 3
 ip y
i
= (2) d p uk (p, , n) e d p vk p , , n e a (p, , n) , a p , , n

,n ,n
XX Z Z
 
I = (2)3 d3 p uk (p, , n) eipx d3 p vk p , , n eip y nn p p
,n ,n

Therefore, the commutation or anti-commutation relations between annihilation and creation fields yield
h i Z
1 X
k+ (x) , k (y) = 3 d3 p uk (p,, n) vk (p,, n) eip(xy) (4.55)
(2) ,n

where the sign indicates a commutator or anticommutator if the particles destroyed and created by the com-
ponents k+ and k are bosons or fermions respectively. We see from (4.55) that the commutation relation (4.54)
is not satisfied automatically by arbitrary functions of the annihilation and creation fields, because in general
the integral in (4.55) does not vanish even if (x y) is space-like. We cannot avoid this problem by making the
4.5. INTERNAL SYMMETRIES AND ANTIPARTICLES 147

interaction density out of annihilation or creation fields alone, since in that case the interaction would not be
hermitian. The only way to avoid such a difficulty is by making linear combinations of annihilation and creation
fields
k (x) k k+ (x) + k k (x) (4.56)
where the constants k and k and any other arbitrary constants in the fields, are adjusted such that
h i
[k (x) , k (y)] = k (x) , k (y) = 0 if (x y)2 0 (4.57)

note that by including explicit constants in Eq. (4.56) we are still leaving free the overall scale of the annihilation
and creation fields to choose it at our convenience. We shall see later how to choose the constants in the linear
combination (4.56) for several irreducibly transforming fields. The Hamiltonian density H (x) will satisfy the
commutation condition (4.54) if it is constructed out of such fields and their adjoints, with an even number of any
field components that destroy and create fermions.
The condition (4.57) is frequently described as a causality condition, because if xy is space-like, no signal can
reach y from x or vice versa. Therefore, the measurement of k at a space-time point x should not interfere with a
measurement of k or k at a point y. Such considerations of causality makes sense for the electromagnetic field,
since each of its components can be measured at a given space-time point. However, the fields we shall be dealing
with (such as the Dirac field of the electron) seem not to be measurable in any sense. Hence, we shall better
take Eq. (4.57) as a necessary condition for Lorentz invariance of the Smatrix but without any association with
measurability or causality.

4.5 Internal symmetries and antiparticles


We then construct fields (4.56) that satisfy Eq. (4.57). It could happen that particles that are destroyed and
created by these fields carry non-zero values of one or more conserved quantum numbers. We shall take as an
example the conservation of the electric charge. If particles of species n carry a value q (n) for the electric charge,
we can define the charge operator Q, as3
Q |q1 , . . . , qN i = {q (n1 ) + q (n2 ) + . . . + q (nN )} |q1 , . . . , qN i (4.58)
Q |q1 , . . . , qN i = Q |q1 , . . . , qN i (4.59)
where we have defined Q as the total charge of the state |q1 , . . . , qN i. That is
Q q (n1 ) + q (n2 ) + . . . + q (nN ) (4.60)
since q (nk ) is the charge of the species nk , and using the action of a (qr ) on an arbitrary multi-particle state [see
Eq. (3.12), Page 119] we obtain
N
X
Q a (q) |q1 , . . . , qN i = Q (1)r+1 (q qr ) |q1 . . . qr1 qr+1 . . . qN i
r=1
N
X
= (1)r+1 (q qr ) [q (n1 ) + . . . + q (nr1 ) + q (nr+1 ) + . . . + q (nN )]
r=1
|q1 . . . qr1 qr+1 . . . qN i
N
X
= (1)r+1 (q qr ) [Q q (nr )] |q1 . . . qr1 qr+1 . . . qN i
r=1
N
X
Q a (q) |q1 , . . . , qN i = [Q q (n)] (1)r+1 (q qr ) |q1 . . . qr1 qr+1 . . . qN i
r=1
3
Do not confuse the notation qr for the state of the particle qr pr , r , nr , with the notation q (nk ) that gives the charge of the nk
species.
148 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

where we have taken into account that because of the delta function, the term Q q (nr ) only contributes if q = qr
so that n = nr , where n denotes the species of particle destroyed by a (q). We obtain finally

Q a (q) |q1 , . . . , qN i = [Q q (n)] a (q) |q1 , . . . , qN i (4.61)

On the other hand,

a (q) Q |q1 , . . . , qN i = a (q) Q |q1 , . . . , qN i


a (q) Q |q1 , . . . , qN i = Q a (q) |q1 , . . . , qN i (4.62)

substracting Eqs. (4.61, 4.62) we obtain

{Q a (q) a (q) Q} |q1 , . . . , qN i = q (n) a (q) |q1 , . . . , qN i

since the state |q1 , . . . , qN i is arbitrary we obtain

[Q, a (p, , n)] = q (n) a (p, , n)

similarly by using the definition of the creation operator Eq. (3.7), page 118, we have

Q a (q) |q1 , . . . , qN i = Q |q, q1 , . . . , qN i = [Q + q (n)] |q, q1 , . . . , qN i = [Q + q (n)] a (q) |q1 , . . . , qN i


a (q) Q |q1 , . . . , qN i = Q a (q) |q1 , . . . , qN i

so that h i
Q, a (p, , n) = +q (n) a (p, , n)
In conclusion, if particles of species n carry a value q (n) for the electric charge, the commutator of the charge
operator with the creation and annihilation fields yield4

[Q, a (p, , n)] = q (n) a (p,, n) (4.63)


h i
Q, a (p, , n) = +q (n) a (p,, n) (4.64)

Note that for Eqs. (4.63, 4.64) to be fullfilled, we only require that the conserved quantum number be additive
i.e. we only require the property (4.58) for a generalized conserved charge.
Now, let us Q denotes the charge operator or any other symmetry operator. Since it must be a constant of
motion, it should commute with the Hamiltonian, and so with the interaction density H (x). In order that H (x)
should commute with the charge operator Q (or some other symmetry generator) it is necessary that it be formed
with fields that have simple commutation relations with the charge operator [similar to the relation given by Eq.
(4.63]:
[Q, k (x)] = qk k (x) (4.65)
since in that case, we can achieve that H (x) commute with Q, by constructing it as a sum of products of fields

k1 k2 and adjoints m 1 m2 such that

qk1 + qk2 + qm1 qm2 = 0 (4.66)

Now, according with (4.16), H (x) is written as a sum of products of the creation and annihilation fields. Thus,
we should examine the commutation relations of creation and annihilation fields with the charge operator, and
check for the conditions to satisfy Eq. (4.65). To do it, we use Eqs. (4.39, 4.63) to obtain
Z Z
  X X
Q, k+ (x) = (2)3/2 d3 p uk (p, , n) eipx [Q, a (p, , n)] = (2)3/2 d3 p uk (p, , n) eipx [q (n) a (p, , n)]
,n ,n
4
It is worth emphasizing that Eqs. (4.63, 4.64) are only commutator relations, and have no anti-commutator counter-parts. They
are valid regardless the species of particles that carry the conserved quantum number are bosons or fermions.
4.6. LORENTZ IRREDUCIBLE FIELDS AND KLEIN-GORDON EQUATION 149

and the satisfaction of Eq. (4.65) requires that we are able to factorize q (n) out of the sum such that
X Z
 +
 3/2
Q, k (x) = qk (2) d3 p uk (p, , n) eipx [ a (p, , n)] = qk k (x) (4.67)
,n

from which it is clear that Eq. (4.65) is satisfied for a given component of the annihilation field k+ (x) if and
only if all particle species n that are destroyed by the field carry the same charge q (n) = qk . Thus, the sum over
species in Eq. (4.67) is only over species of the same charge. Similarly, Eq. (4.65), is satisfied by one particular
component k (x) of the creation field if and only if all particle species n that are created by the field, carry the
charge q (n) = qk . We then conclude that in order that such a theory conserve quantum numbers like electric
charge, there must be a doubling of particle species carrying non-zero values of the conserved quantum numbers:
if a particular component of the annihilation field destroys a particle of species n, then the same component of the
creation field must create particles of a species n, called the antiparticles of the particles of species n, that have
opposite values of all conserved quantum numbers. These arguments then, predicts the existence of antiparticles.

4.6 Lorentz irreducible fields and Klein-Gordon equation


We have said that the Lorentz representations D () in Eq. (4.6, 4.7) do not have to be irreducible. By choicing
an apropriate basis (canonical basis) the matrix representation D () could acquire a block diagonal form, so that
fields that belong to different blocks cannot transform into each other under Lorentz transformations. On the
other hand, the Lorentz transformations do not have effect on the particle species. Consequently, instead of using
one big field, including many irreducible components and many particle species, we shall restrict our attention to
fields that destroy only a single type of particle (since in this case the label n is fixed, we shall omit it) and create
only the corresponding antiparticle, and that also transform irreducibly under the Lorentz group (as mentioned
above we could include or not include space inversion). It is understood that in general we shall have to consider
many different such fields, some perhaps formed by the derivatives of other fields. The next task is to finish
the determination of the coefficient functions uk (p, ) and vk (p, ), fix the values of the constants and of
Eqs. (4.56) and deduce relations between the properties of particles and antiparticles for fields that belong to the
simplest irreducible representations of the Lorentz group: the scalar, the vector, and Dirac spinor representations.
Let us then concentrate on all components of a field of definite mass m. For our purposes we first calculate
the quantities eipx . For them, we obtain

h
i h
i
eipx = g eip x = g ip eip x x = g ip eip x
h
i
= ip g eip x = p p g eipx = p p g eipx
eipx = p2 eipx = m2 eipx
now, from Eq. (4.39) and assuming that the sum is only over species of the same mass, we have
X Z
 
k+ (x) = (2)3/2 d3 p uk (p, , n) eipx a (p, , n)
,n
X Z
2 3/2
= m (2) d3 p uk (p, , n) eipx a (p, , n)
,n


k+ (x) = 2 +
m k (x) (4.68)
similarly, from Eq. (4.40) we obtain
X Z
3/2  

k (x) = (2) d3 p vk (p, , n) eipx a (p, , n)
,n


k (x) = m2 k (x) (4.69)
150 CHAPTER 4. RELATIVISTIC QUANTUM FIELD THEORY

and combining Eqs. (4.68, 4.69) with (4.56), we can obtain a field equation for k (x)
 
k (x) k k+ (x) + k k (x) = k k+ (x) + k k (x)
 
= m2 k k+ (x) + k k (x)
k (x) = m2 k (x) (4.70)

rewriting the second-order differential operator as

 (4.71)

we can rewrite the field equation (4.70) in the form



 m2 k (x) = 0 (4.72)

Expression (4.72), is called the Klein-Gordon equation. This will be one of our most important field equations
to work with. We recall that the validity of the equation requires that the expansion of the fields k (x) in Eqs.
(4.39, 4.40) be over particle species of the same mass. For example, Klein Gordon equation is valid if the sum is
over a particle and its corresponding antiparticle. Of course, such an equation is also valid if the expansion of the
fields is over a single particle species. In conclusion, the combination of Eqs. (4.39, 4.40) with Eq. (4.56) shows
that all components of a field of definite mass m satisfy the Klein-Gordon equation (4.72). Some fields satisfy
other field equations, depending on whether or not there are more field components than independent particle
states.
The most usual approach is to start with the field equation (e.g. the Klein-Gordon equation) or from the
Lagrangian density that generates it5 , and uses them to derive the expansion of the fields in terms of one-particle
annihilation and creation operators. In the present approach, the starting point are the particles, and we derive
the fields according with the dictates of CDP and Lorentz invariance, in this way the field equations arise as a
byproduct of this construction.
We have already proved in Sec. 3.6, that the condition that guarantees that a theory satisfies the CDP, is that
the interaction can be expressed as a sum of products of creation and annihilation operators, with all creation
operators to the left of all annihilation operators, and with coefficients that contain only a single momentum
conservation delta function. Owing to it, we should write the interaction in the normal ordered form
Z  
V = d3 x : F (x) , (x) :

the colons indicate that the enclosed expression is to be rewritten so that all creation operators stand to the left
of all annihilation operators. In doing this, we should ignore non-vanishing commutators or anticommutators, but
we should include minus signs for permutations of fermionic operators. Moreover, by using the commutation or
anticommutation relations of the fields, any such normal ordered function of the fields can be written as a sum
of ordinary products of the fields with complex number coefficients. Rewriting : F : in this way we can see that
if it is constructed out of fields that
 satisfy Eq. (4.57), with even numbers of any fermionic field components, we
shall obtain that : F (x) , (x) : will commute with : F (y) , (y) : when x y is space-like, despite the
normal ordering.

5
We can see the plausibility of the Klein-Gordon equation by a principle of correspondence for the relativistic free-particle relation
E p2 m2 = 0 or equivalently p p m2 = 0. Resorting to the principle of correspondence p i = i (, t ) we have
2

p p k (x) k (x) = k (x), and we obtain


 the Klein-Gordon equation. Note that the same principle of correspondence
along with the non-relativistic relation E = p2 /2m + V (x, t), lead us to the Schrodinger equation.
Chapter 5

Causal scalar fields for massive particles

We shall consider first the so-called scalar representation of the homogeneous Lorentz group in which D () = 1.
Since this irreducible representation is one-dimensional, we consider one-component annihilation and creation
fields + (x) and (x) that transform according with the scalar representation of the Lorentz group. When we
restrict us to rotations, it is the scalar representation of the SO (3) group, for which all matrix representatives of
rotations and/or generators collapse to the identity matrix in one dimension. The label of the identity (or scalar)
representation of SO (3) corresponds to j = 0 (zero spin particles), from which the labels , in Eqs. (4.49) and
(4.50) can only take a single value. Consequently, a scalar field can only describe particles of zero spin.

5.1 Scalar fields without internal symmetries


Let us assume by now that the field describes only a single species of particle, with no distinct anti-particle, the
quantities uk (0, , n) and vk (0, , n) can be written just as the numbers u (0) and v (0) (we omit the species label
n, the field component label k, and the spin label because each of these labels takes only one value). Thus, the
annihilation and creation fields (4.39) and (4.40) are written as
Z Z
1 1
+
(x) = q 3 ipx
d p u (p) e a (p) ; (x) = q d3 p v (p) eipx a (p) (5.1)
(2)3 (2)3

now since we are in the identity representation Dkk (L (p)) = kk , and Eqs. (4.45, 4.46) become
r r
m m
u (p) = 0
u (0) ; v (p) = v (0) (5.2)
p p0
so that the fields read
Z r Z r
1 m 1 m
+
(x) = q 3
d p 0
u (0) eipx a (p) ; (x) = q 3
d p v (0) eipx a (p)
p p0
(2)3 (2)3

now, the simplest way to settle + (x) and (x) as the adjoint of each other, is by choosing u (0) and v (0) to
be real and equal. Therefore, it is conventional to adjust the overall scales of creation and annihilation operators
such that these constants acquire the values
1
u (0) = v (0) = (5.3)
2m
then, combining Eqs. (5.2) with conventions (5.3) we find
1
u (p) = v (p) = p (5.4)
2p0

151
152 CHAPTER 5. CAUSAL SCALAR FIELDS FOR MASSIVE PARTICLES

Hence, in the scalar case without internal symmetries, the annihilation and creation fields (5.1) are written as
Z
1 1
+ (x) = q d3 p p a (p) eipx (5.5)
3 2p 0
(2)
Z
1 1
(x) = q d3 p p a (p) eipx = + (x) (5.6)
3 2p 0
(2)

A Hamiltonian density H (x) that is formed as a polynomial in + (x) and (x), automatically satisfy the
requirement (4.16) to transform as a scalar. We should also ensure that it also satisfies the condition of the
Lorentz-invariance for the Smatrix, expressed by the fact that H (x) must commute with H (y) at space-like
separations x y. This condition would be satisfied automatically if H (x) were a polynomial in + (x) alone,
because all annihilation operators commute (anticommute) if the particles are bosons (fermions). Then we have1
 + 
(x) , + (y) = 0 (5.7)

regardless the distance (x y) is space-like or not. We denote [. . .] to mean the commutator and [. . .]+ to mean
the anti-commutator. However, if the Hamiltonian density H (x) were a polynomial in + (x) alone, it would be
not hermitian. To be hermitian, H (x) must involve + (x) but also its adjoint + (x) = (x). The problem then
is that + (x) does not commute nor anti-commute with (y) for general space-like separations. To characterize
the commutation or anti-commutation algebra between + (x) and (y) we utilize the expressions (5.5, 5.6) for
such scalar fields
"Z Z #
 +
 1 3 1 ipx 3 1
 ip y
(x) , (y) = d p p a (p) e , d p p a p e
(2)3 2p0 2p0
Z Z ipx ip y h i
1 e e 
= 3 d3 p d3 p p p a (p) , a p
(2) 0
2p 2p 0

and using the commutation and anti-commutation relations for bosons and fermions respectively, Eqs. (3.17), we
obtain
Z
 +
 d3 p d3 p 
(x) , (y) = 3
p eipx eip y 3 p p
(2) 0 0
(2p 2p )
Z 3
 +  1 d p ip(xy)
(x) , (y) = 3 e
(2) 2p0

hence this commutator or anti-commutator can be expressed in terms of a single integral


 + 
(x) , (y) = + (x y) (5.8)
Z 3
1 d p ipx
+ (x) 3 e (5.9)
(2) 2p0

since d3 p/p0 is the Lorentz invariant volume [see Eq. (1.157), page 37)], the integral + (x) is manifestly Lorentz-
invariant. Consequently, for space-like x, it can depend only on the invariant square x2 > 0, and can be evaluated
in any convenient reference frame. We can evaluate + (x) for space-like x by choosing the coordinate system
such that2
x0 = 0, kxk = x2 p x = p0 x0 + p x = p x (5.10)
1
Equation (5.7) comes from the fact that [a (p) , a (p )] = 0, combined with equation (5.5)
2
It is important to recall that the choice (5.10) is only possible for space-like events. This choice means that in this reference frame,
the event (0, 0) is simultaneous with the event x,x0 with x 6= 0. Hence it is only possible for causal disconnected (space-like) events.
5.1. SCALAR FIELDS WITHOUT INTERNAL SYMMETRIES 153

from which Eq. (5.9) becomes Z


1 d3 p
+ (x) = p eipx (5.11)
(2)3 2
2 p +m 2

for a while we shall denote p = |p|, so by now p stands for the magnitude of the three-vector (instead of denoting
the four-vector). By using spherical coordinates in the three-momentum space we have d3 p = p2 dp sin d d.
By convention we choose3

x = kxk u3 = x2 u3 so that p x = kpk kxk cos = p x2 cos

picking up all these facts, Eq. (5.11) becomes


Z 2 Z Z  Z 2
1 p dp sin d d ipx2 cos 1 p2 dp ip x2 cos
+ (x) = p e = p e sin d d
(2)3 2 p 2 + m2 (2)3 0 2 p2 + m2 0 0
Z Z
2 p2 dp
+ (x) = 3
p I ; I eia cos sin d , a p x2 (5.12)
(2) 0 2 p2 + m2 0

let us first evaluate the integral I


Z  
ieia cos i ia  2 eia eia 2 sin a
I eia cos sin d = = e eia
= =
0 a 0 a a 2i a
 
2 sin a 2 sin p x2
I = = (5.13)
a p x2
substituting (5.13) in Eq. (5.12), we find
 
Z sin p x2
2 p2 dp
+ (x) = p
(2)3 0 p 2 + m2 p x2
and changing the variable of integration to
p dp
u ; du =
m m
it yields
   
Z 2 Z 2
1 m2 u2 m du sin mu x

1
m3 u du sin mu x
+ (x) = =
(2)2 0 m2 u2 + m2 mu x2 4 2 0 m u2 + 1 m x2
Z  
m u du
+ (x) = sin m x2 u (5.14)
4 2 x2 0 u2 + 1
which can be written in terms of a Hankel function
m  
+ (x) = K1 m x2 (5.15)
4 2 x2
we emphasize again that expression (5.15) is only valid for x2 > 0, i.e. for space-like x. This function is not zero.
However, we observe from (5.15) that for x2 > 0, + (x) is an even function in x . Instead of using only + (x),
we shall try to construct H (x) from a linear combination of + (x) and + (x) = (x)

(x) + (x) + (x) (5.16)


3
Since x is fixed (and so x), during the integration, choosing x along the three-axis is part of the definition of the reference frame.
154 CHAPTER 5. CAUSAL SCALAR FIELDS FOR MASSIVE PARTICLES

from (5.16) and using (5.7) and its adjoint we have

h i  
(x) , (y) = + (x) + (x) , (y) + + (y)

 +   
= (x) , (y) + + (y) + (x) , (y) + + (y)
 +       
= (x) , (y) + + (x) , + (y) + (x) , (y) + (x) , + (y)
       
= + (x) , (y) + + (x) , + (y) + (x) , (y) + (x) , + (y)
h i    
(x) , (y) = ||2 + (x) , (y) + ||2 (x) , + (y)

and using Eqs. (5.8), we have for x y being space-like, that


h i    
(x) , (y) = ||2 + (x) , (y) ||2 + (y) , (x) = ||2 + (x y) ||2 + (y x)
h i  
(x) , (y) = ||2 ||2 + (x y) (5.17)

where we have used the fact that the commutator (anticommutator) is antisymmetric (symmetric) under the
interchange of the fields. We also used the fact that + (x) is an even function of x. Similarly
    
[ (x) , (y)] = + (x) , (y) + (x) , + (y) = (1 1) + (x y) (5.18)

note that for the case of anticommutators, in which the + sign applies, expressions (5.17, 5.18) never vanishes.
Consequently, the particle cannot be a fermion. Thus, those expressions vanish if and only if the particle is a
boson and and are equal in magnitude
|| = || (5.19)
Let us redefine the phases of the states such that a (p) ei a (p), then a (p) ei a (p). According with
Eqs. (5.5, 5.6) this redefinition leads to + (x) ei + (x) and (x) ei (x). Finally, for (x) to be
invariant in Eq. (5.16) we should redefine

ei and ei

finally, by taking 
1
arg
2
we can make the phases of and to coincide. Therefore, under this convention = . Now, absorbing the
overall factor = in Eq. (5.16) we obtain

(x) + (x) + (x)

from which it is clear that (x) is self-adjoint

(x) + (x) + + (x) = (x) (5.20)

and using (5.5) and (5.6) we obtain its fourier expansion


Z
1 1 h i
+ +
(x) (x) + (x) = q d3 p p a (p) eipx + a (p) eipx (5.21)
(2)3 2p0

thus, the interaction density H (x) will commute with H (y) for space-like separations x y, if it is constructed as
a normal-ordered polynomial in the self-adjoint scalar field (x) defined by (5.20).
5.2. SCALAR FIELDS WITH INTERNAL SYMMETRIES 155

Despite the choice of relative phase in the two terms in Eq. (5.20) is conventional, it is also a convention that
once a given convention is taken, it will be applied wherever a scalar field for this particle appears in the interaction
Hamiltonian density. For example, suppose that besides the field (5.20) the interaction density contains another
scalar field e (x) for the same particle

e (x) = ei + (x) + ei + (x)

where is an arbitrary phase. This e like , would be causal in the sense that e (x) commutes with e (y) when
x y is space-like. However, e (x) would not commute with (y) for space-like separations of x and y. Therefore,
both fields cannot appear in the same theory.

5.2 Scalar fields with internal symmetries


In the previous treatment, we assumed that we have only one species of particles with no distinct antiparticles.
In other words, we have supposed that the particle described is totally neutral, i.e. that all intrinsic quantum
numbers of it are null. Nevertheless, if the particles that are destroyed and created by (x) carry some quantum
number like electric charge, we see that H (x) conserve this number if and only if each term in H (x) contains
equal numbers of operators a (p) and a (p). For instance, if a given term contains two operators of the type
a (p), then it would create two particles and so two units of charge, with the consequent increasing of the electric
charge of the multi-particle state. Therefore it is necessary that the same term contains two operators of the type
a (p) that destroy two particles and hence two units of charge, in order to obtain the same net charge at the end
of the process in the multi-particle state.
Nevertheless, it is impossible that each term in the interaction density H (x), contains the same number of
operators a (p) and a (p), if H (x) is constructed as a polynomial in (x) = + (x)++ (x). To see it, we observe
that according with Eqs. (5.5, 5.6) the field + (x) contains an operator a (p), while the field (x) contains an
operator a (p). Let us take for example a polynomial of second degree in (x), it would be of the form
   2
+ 2 = + (x) + (x) + + (x) + (x)
   2  2  
+ 2 = + (x) + (x) + + (x) + (x) + + (x) (x) + (x) + (x)

indeed only the last two terms on the RHS of this equation contain the same number of operators a (p) and a (p).
We can see the problem from another point of view. For H (x) to commute with the charge operator Q (or any
other symmetry operator) it is necessary that H (x) be constructed out of fields that have simple commutation
relations with Q such as Eq. (4.65). This is true for + (x) and its adjoint (x) as can be seen from Equations
(5.5, 4.63)
" Z # Z
 +
 1 d3 p ipx eipx d3 p
Q, (x) = q Q, p a (p) e =q p [Q, a (p)]
(2)3 2p0 (2)3 2p0
Z
eipx d3 p
= q q p a (p) = q+ (x)
2p 0
(2)3

and similarly for (x) = + (x), we then obtain


 
Q, + (x) = q+ (x)
h i  
Q, + (x) = Q, (x) = +q+ (x)

however, the self-adjoint field (x) = + (x) + + (x) defined in Eq. (5.20), does not satisfy a relation of the type
(4.65).
156 CHAPTER 5. CAUSAL SCALAR FIELDS FOR MASSIVE PARTICLES

We deal with this problem by assuming that there are two spinless bosons in which one of them is the charge
conjugate of the other. We shall see that the two bosons should be of equal mass and opposite generalized
charges. We denote + (x) and c+ (x) the annihilation fields for these two particles, where the label c denotes
charge conjugate. Then we have
   
Q, + (x) = q+ (x) ; Q, c+ (x) = qc c+ (x) (5.22)

of course the fields + (x) and c+ (x) are expanded as in Eq. (5.5) in terms of annihilation operators a (p) and
ac (p) respectively.
Now, we define the linear combination

(x) = + (x) + c+ (x) (5.23)

from Eqs. (5.22) the commutation relation of the operator Q with the field (x) defined by Eq. (5.23), yields
h i   h i
[Q, (x)] = Q, + (x) + c+ (x) = Q, + (x) + Q, c+ (x)

+ c c+
[Q, (x)] = q (x) + q (x) (5.24)

now, to find a relation of the form (4.65) for (x), we require to reconstruct (x) at the RHS of Eq. (5.24). To
do it we demand that qc = q, from which we obtain a relation of the form (4.65)
h i
[Q, (x)] = q + (x) + c+ (x)
[Q, (x)] = q (x)

we conclude that for the field (x) defined by Eq. (5.23), we obtain a relation of the form (4.65) [which is the
same commutation relation with Q of the operator + (x) alone], if we demand that qc = q. Therefore, Eqs.
(5.22) become
 
Q, + (x) = q+ (x) (5.25)
 
Q, c+ (x) = +qc+ (x) (5.26)

We should keep in mind that a particle that carries no conserved quantum numbers may or may not be its own
antiparticle, with a (p) = ac (p). By now we are assuming that the two spinless bosons have opposite charges (or
any other conserved quantum number), so that a (p) 6= ac (p).
On the other hand, since a (p) and ac (p) destroy different particles they must commute or anticommute. In a
similar way, the annihilation fields + (x) and c+ (x) destroy different particles. A similar argument follows for
the creation operators and fields. Thus, we have4
  h i
a (p) , ac p
= a (p) , ac p =0

   
+ (x) , c+ (y) = (x) , c (y) = 0

4
We can also see it in the following way
  
a (p, , n) , a p , , n
= p p nn (5.27)

and ac (p , , n ) a (p , , nc ). If the particle does not coincide with the antiparticle then n 6= nc such that (5.27) is clearly zero.
5.2. SCALAR FIELDS WITH INTERNAL SYMMETRIES 157

from which the commutator or anticommutator of (x) with its adjoint at a space-like separation is given by
h i h i
(x) , (y) = + (x) + c+ (x) , + (y) + c+ (y)

h i h i
+ c+
+
= (x) , (y) + (y) + c+
(x) , + (y) + c+ (y)

h i  
2
= || + (x) , + (y) + + (x) , c+ (y)

h i h i
+ (x) , (y) + ||2 c+ (x) , c+ (y)
c+ +

h i h i h i
2 + + 2 c+ c+
(x) , (y) = || (x) , (y) || (y) , (x) (5.28)

further it is clear that Eq. (5.8) could be applied for both pairs + (x) , + (y) and c+ (y) , c+ (x), since their
associated annihilation and creation operators a (p) , a (p) and ac (p) , ac (p) have the same commutation or
anticommutation algebra. Thus for space-like separations x y we obtain
h i
(x) , (y) = ||2 + (x y) ||2 c+ (x y)

where we have used the even property of this function for space-like separations. Now, for this commutator or
anticommutator to be null for all space-like separations x y, we require to factorize the functions + and c+ .
Then we demand that
+ (x y) = c+ (x y)
on the other hand, expression (5.11) says such a condition is satisfied for all space-like separations x y, if and
only if both functions are associated with the same mass. In other words, the causal condition leads us to the
condition that the particle and antiparticle have the same mass. Then, combining Eqs. (5.8) and (5.28) we have
h i  
(x) , (y) = ||2 ||2 + (x y) (5.29)

while (x) and (y) automatically commute or anticommute with each other for all x and y (regardless its
separation is space-like or not) because + and c+ destroy and create different particles. Once again, Eq.
(5.29) shows that Fermi statistics is ruled out here, because (x) cannot anticommute with (y) at space-like
separations unless = = 0, so that the fields (5.23) simply vanish. We conclude that a spinless particle must
be a boson5 .
Now restricting to Bose statistics, we see from Eq. (5.29), that in order that a complex (x) should commute
with (y) at space-like separation, it is necessary and sufficient that ||2 = ||2 , as well as for the particle and
antiparticle to have the same mass.
As before, we can redefine the relative phase of states of these two particles, and give and the same phase
such that = . Once again, we can eliminate the resultant common factor by redefinition of the field (x) in
(5.23) to obtain
(x) = + (x) + c+ (x) (5.30)
and expanding + (x) + c+ (x) as in (5.5) we obtain
Z h i
d3 p ipx c ipx
(x) = a (p) e + a (p) e (5.31)
(2)3/2 (2p0 )1/2
which is the essentially unique causal scalar field [notice however that this (x) is not hermitian anymore].
Expression (5.31) is valid both for purely neutral spinless particles6 [in that case we take ac (p) = a (p)], and for
particles with distinct antiparticles for which ac (p) 6= a (p).
5
We recall here that in the symmetrization postulate we have associated bosons with symmetric physical states and fermions with
antisymmetric physical states. We had not done any association between bosons and fermions with the spin of the particles.
6
By purely neutral we mean that the particle carries no conserved quantum numbers.
158 CHAPTER 5. CAUSAL SCALAR FIELDS FOR MASSIVE PARTICLES

We note for future purposes, that the commutator of the complex field (x) with its adjoint is

h i h i
(x) , (y) = + (x) + c+ (x) , + (y) + c+ (y)

h i h i
+ + c+ c+
= (x) , (y) + (x) , (y)
h i h i
+ + c+ c+
= (x) , (y) (y) , (x)

h i
(x) , (y) = + (x y) + (y x)

 
where we have used Eq. (5.8). The commutator (x) , (y) then becomes

h i
(x) , (y) = (x y)

Z h i
d3 p
(x y) + (x y) + (y x) = eip(xy) eip(xy) (5.32)
2p0 (2)3

this is a general relation that is for any kind of separation between x and y. We already saw that for x2 > 0, + (x)
is an even function of x. Hence, the function (x y) defined by Eq. (5.32) is null for space-like separations of
x and y.

5.3 Scalar fields and discrete symmetries

We shall consider now the effect of the inversion symmetries on the field (x) defined in (5.31). To do this, we take
the results obtained in section 3.4, in which we studied the transformation properties of the creation operators
a (p, , n) under C, P and T .
Equation (3.28), page 122 gives us the effect of the space-inversion operator on the creation operators, it is
clear that the effect on the annihilation operator is obtained by simply taking the adjoint of (3.28) and that the
operators ac (p) and ac (p) must obey identical rules of transformation, then

P a (p) P 1 = a (p) (5.33)


c 1 c c
P a (p) P = a (p) (5.34)

where and c are the intrinsic parities of the particle and antiparticle respectively. We can then apply the result
(5.33) to the annihilation field (5.5) to find

Z
1 1 1  
+
P (x) P = q d3 p p P a (p) P 1 eipx
2p 0
(2)3
Z
1 1
a (p) ei(pxp x )
0 0
= q d3 p p
(2)3 2p0
Z
1
a (p) ei(p(x)p x )
0 0
= q d3 p p
2p 0
(2)3
5.3. SCALAR FIELDS AND DISCRETE SYMMETRIES 159

and changing the variable of integration from p to p we have7


Z Z
1 i[p(x)p0 x0 ] 1
a (p) ei[(p,p )(x,x )]
0 0
+ 1 3
P (x) P = q d p p a (p) e =q d3 p p
(2)3 2p0 (2)3 2p0
Z
+ 1 1 
P (x) P = q d3 p p a (p) ei[p(Px)] ; Px x, x0
(2)3 2p0

P + (x) P 1 = + (Px) ; Px x, x0

similarly, applying the result (5.34) to the charge-conjugate of the creation field (5.6) we find

Z
1 1 1 h c i
P c+
(x) P = q d3 p p P a (p) P 1 eipx
(2)3 2p0
Z
1 1 h c c i
a (p) ei[pxp x ]
0 0
= q d3 p p
2p 0
(2)3

and changing the variable of integration from p to p we have


Z
c+ 1 c ac (p) i[pxp0 x0 ]
P (x) P = q d3 p p e = c c+ (Px)
3 2p 0
(2)

in summary, the fields + (x) and c+ (x) transform under space inversion according to the following prescription

P + (x) P 1 = + (Px) (5.35)


c+ 1 c c+ 0

P (x) P = (Px) ; Px = x, x (5.36)

It is worth emphasizing that when space-inversion is applied to the scalar field given by

(x) = + (x) + c+ (x) (5.37)

we obtain a different field P (x)


h i
P (x) P (x) P 1 = P + (x) + c+ (x) P 1 = P + (x) P 1 + P c+ (x) P 1
P (x) P (x) P 1 = + (Px) + c c+ (Px) (5.38)

Both fields (x) and P (x) are separately causal. However, in general and P do not commute at space-like
separations. In that case, the fields and P cannot appear in the same interaction. So the only way to preserve
Lorentz invariance, parity conservation and the hermiticity of the interaction is by demanding that P (x) be
proportional to (Px). In turn, it means that we should be able to factorize the phases on the RHS of Eq. (5.38)
and hence that
c = (5.39)
consequently, the intrinsic parity c = ||2 of a state that contains a spinless particle and its antiparticle is even.
Combining equations (5.38, 5.39) we then have
h i
1 + c+
P (x) P (x) P = (Px) + (Px)
7
In the change of variable p p, the differential of volume changes the sign but at the same time the three intervals of integration
(in cartesian coordinates) are also inverted. By inverting again the intervals of integration we reverse the sign again. Thus, the net
result is that the differential of volume does not change the sign.
160 CHAPTER 5. CAUSAL SCALAR FIELDS FOR MASSIVE PARTICLES

obtaining finally,
P (x) P 1 = (Px) (5.40)
the results are also valid when the spinless particle coincides with its antiparticle, since in this case = c , which
combined with Eq. (5.39), implies that the intrinsic parity of such a particle must be real: = 1.
Now we deal with time-reversal. From Eq. (3.29), page 122, we have

T a (p) T 1 = a (p)
T ac (p) T 1 = c ac (p)

and taking into account the antilinearity of T , we have


 
Z d3 p 1 Z
d 3p T a (p) T 1
T + (x) T 1 = T q p a (p) eipx T 1 = q p eipx
3 2p 0 3 2p 0
(2) (2)
Z
1
a (p) ei[px+p x ]
0 0
= q d3 p p
(2)3 2p0

with the change of variable p p, we obtain


Z Z
1 1
a (p) ei[px+p x ] = q a (p) ei[(p,p )(x,x )]
0 0 0 0
T + (x) T 1 = q d3 p p d3 p p
(2)3 2p0 (2)3 2p0
Z
1
= q d3 p p a (p) ei[p(Px)]
3 2p 0
(2)
T + (x) T 1 = + (Px)

and similarly for c+ (x). Thus we find

T + (x) T 1 = + (Px) (5.41)


c+ 1 c c+
T (x) T = (Px) (5.42)

applying time-reversal to the field (x) of Eq. (5.31), we obtain a new field T (x) T (x) T 1 . For this field to
be simply related with the field at the time-reversed point Px, we must have

c = (5.43)

obtaining
T (x) T 1 = (Px) (5.44)
Charge-conjugation can be managed similarly. By using the results of Sec. 3.4, Eq. (3.27), page 122, we have

Ca (p) C 1 = ac (p) (5.45)


c 1 c
Ca (p) C = a (p) (5.46)

where and c are the phases associated with the operation of charge conjugation on one-particle states. Using
Eqs. (5.45, 5.46) in the field expansion (5.5) and the charge-conjugate of (5.6) we find

C+ (x) C 1 = c+ (x) (5.47)


c+ 1 c +
C (x) C = (x) (5.48)
5.3. SCALAR FIELDS AND DISCRETE SYMMETRIES 161

again the charge-conjugation operation applied on the field (x) of Eq. (5.31) gives another field C (x)
C (x) C 1 . In order that C (x) commutes with (x) at space-like separations, C (x) must be proportional to
(x), for which it is necessary that
c = (5.49)
just as for the ordinary parity, the intrinsic charge-conjugation parity c of a state consisting of a spinless particle
and its antiparticle is even. We just have

C (x) C 1 = (x) (5.50)

again, these results also apply to the case in which the particle is its own antiparticle, so c = . In that case, the
charge-conjugation parity like the ordinary parity must be real = 1.
Finally, we point out that the results shown in this section are valid only if the discrete symmetry involved is
a good symmetry of the system.
Chapter 6

Causal vector fields for massive particles

The next simplest case arise when we deal with fields that transform as a four-vector, which is associated with
the so-called vector representation of the homogeneous Lorentz group. This is the four-dimensional representation
with which we constructed originally the Lorentz transformations in the Minkowski space, that is

D () (6.1)

It is important to say that there are massive particles W and Z 0 that at low energies are described by vector
fields, these particles are responsible for what we call the nuclear weak interaction. Further, one possible approach
to quantum electrodynamics (QED) is to describe the photon in terms of a massive vector field in the limit of
very small mass.

6.1 Vector fields without internal symmetries


By now, we shall assume that only one type of particle is described by this field, hence we drop the label n of
species. We shall consider later the possibility that the field describes both a particle and a distinct antiparticle.
The components of the annihilation and creation fields are obtained simply by taking (4.39, 4.40) by dropping
the label n. Further, we shall substitute the label k (component label) for a label (more usual for a four-vector
at the Minkowski space)

X Z
+ 3/2
(x) = (2) d3 p u (p, ) a (p, ) eipx (6.2)

X Z
3/2
(x) = (2) d3 p v (p, ) a (p, ) eipx (6.3)

In addition, the coefficient functions u (p, ) and v (p, ) for arbitrary momentum can be obtained from those
of zero momentum from Eqs. (4.45) and (4.46)
r r
m X m X

u (p, ) = 0

D (L (p)) u (0, ) ; v (p, ) = 0
D (L (p)) v (0, ) (6.4)
p
p

and using the convention of sum over repeated lower-upper indices as well as Eq. (6.1) we find
r
m

u (p, ) = L (p) u (0, ) (6.5)
p0
r
m

v (p, ) = 0
L (p) v (0, ) (6.6)
p

162
6.1. VECTOR FIELDS WITHOUT INTERNAL SYMMETRIES 163

in turn, the coefficient functions at zero momentum are subject to the conditions (4.49) and (4.50)
X (j)
u (0, ) J = J u (0, ) (6.7)

X (j)
v (0, ) J = J v (0, ) (6.8)

the rotation generators J in the four-vector representation (in cartesian coordinates) are given by [homework]

(Jk ) 0 = (Jk )0 = 0 (6.9)


i
(Jk ) j = iijk ; i, j, k = 1, 2, 3 (6.10)

it is clear from Eqs. (6.9) that the rotation generators act only on the three-dimensional space-coordinates. In
particular, we can calculate J 2

J 2 = J12 + J22 + J32


i
J 2 j = (J1 )i (J1 ) j + (J2 )i (J2 ) j + (J3 )i (J3 ) j
= (J1 )i m (J1 )m j + (J2 )i m (J2 )m j + (J3 )i m (J3 )m j
= im1 mj1 im2 mj2 im3 mj3
= i21 2j1 i31 3j1 i12 1j2 i32 3j2 i13 1j3 i23 2j3
= i3 j3 321 231 i2 j2 231 321 i3 j3 312 132
i1 j1 132 312 i2 j2 213 123 i1 j1 123 213

the product of Levi-Civita terms are of the form ijk jik which are equal to (1) because both terms have opposite
sign (they differ each other by an interchange), then we have
i
J2 j = i3 j3 + i2 j2 + i3 j3 + i1 j1 + i2 j2 + i1 j1
= 2 [i1 j1 + i2 j2 + i3 j3 ] = 2i m m j = 2i j

the other terms are easier to calculate


0
J2 = (J1 )0 (J1 ) + (J2 )0 (J2 ) + (J3 )0 (J3 ) = 0

we then obtain
0 
J2 = J2 0 =0 (6.11)

2 i i
J j = 2 j (6.12)

alternatively, we can obtain Eq. (6.12) by observing that J 2 is a casimir of the group SO (3), in the three-
dimensional coordinate space associated with the j = 1 irreducible representation1 . Thus, according with the
Schurs lemma, it must be proportional to the identity within the three-dimensional coordinate space, and specif-
ically of the form j (j + 1) I = 2I.
Setting = 0 in Eq. (6.7) and using (6.9) we obtain
X (j)
u0 (0, ) J = J 0 u (0, ) = 0 (6.13)

1
Note however that J 2 is not a Casimir of the Lorentz group, because it does not commute with all generators of such a group.
Hence, J 2 is not in general proportional to the identity at the Minkowski space.
164 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

2
it is more convenient to obtain an expression in terms of J(j) , because it is a Casimir operator of SO (3) and
(j)
thus proportional to the identity. To do this, we multiply Eq. (6.13) by J and sum over on both sides of this
equation, then we have
XX X  2
(j) (j)
u0 (0, ) J J = 0 u0 (0, ) J(j) =0 (6.14)

now setting = i in Eq. (6.7) and using (6.9) we have


X (j)
ui (0, ) J = J i u (0, )

X (j)
ui (0, ) J = J i m um (0, )

(j)
as before, we multiply by J and sum over on both sides of this equation, then we use equation (6.7) once
again, to obtain
XX (j) (j)
X (j)
ui (0, ) J J = J i m um (0, ) J

X  2
ui (0, ) J(j) = J i m J m u (0, )

and using Eqs. (6.12) we find


X  2 i
ui (0, ) J(j) = J i m J m n un (0, ) = J 2 n un (0, ) = 2i n un (0, ) (6.15)

thus, picking up Eqs. (6.14, 6.15) we obtain finally


X  2
u0 (0, ) J(j) = 0 (6.16)


X  2
ui (0, ) J(j) = 2ui (0, ) (6.17)

an analogous procedure can be done for the v coefficient from Eq. (6.8) to obtain
X  2
v 0 (0, ) J(j) = 0 (6.18)


X  2
v i (0, ) J(j) = 2v i (0, ) (6.19)


2
now we recall that J(j)
= j (j + 1) then Eqs. (6.16-6.19) become
j (j + 1) u0 (0, ) = 0 (6.20)
i i
j (j + 1) u (0, ) = 2u (0, ) (6.21)
0
j (j + 1) v (0, ) = 0 (6.22)
i i
j (j + 1) v (0, ) = 2v (0, ) (6.23)
a non-trivial solution requires that u0 (0, ) and/or ui (0, ) be non-null, and same for v 0 (0, ) and v i (0, ).
Equations (6.20, 6.22) show that u0 (0, ) and v 0 (0, ) can be non-null only if j = 0. In that case, Eqs. (6.21,
6.23) say that ui (0, ) and v i (0, ) must be zero. On the other hand, we see that Eqs. (6.21, 6.23) are consistent
for non-null values of ui (0, ) and v i (0, ) only if j (j + 1) = 2 or equivalently j = 1. In that case Eqs. (6.20,
6.22) show that u0 (0, ) and v 0 (0, ) must be null.
In conclusion consistent solutions for the coefficients at p = 0 can only be obtained in two cases
6.2. SPIN ZERO VECTOR FIELDS 165

1. If j = 0, the representation is one-dimensional and hence acquires a single value, therefore we omit the
label . In this case, ui (0) = v i (0) = 0, while u0 (0) and v 0 (0) could be non-zero.

2. If j = 1, the representation is three-dimensional and acquires three-different values. In this case, u0 (0) =
v 0 (0) = 0, while the ui (0) and v i (0) could be non-zero.

In other words, we have two possibilities for the spin of the particle described by the vector field: j = 0 or
j = 1. Let us now examine both possibilities in detail.

6.2 Spin zero vector fields


We shall study first the case of j = 0, in which

ui (0) = v i (0) = 0

By an appropriate normalization of the non-null components of u (0) and v (0), we can take the non-vanishing
components of these coefficients to have the values (recall that takes a single value so we drop it)
r r
0 m 0 m
u (0) i ; v (0) = i (6.24)
2 2

where we choose u0 (0) = v 0 (0) for the annihilation and creation fields to be the adjoint of each other, the
convenience of introducing the i factor in the choice of phase (6.24) will be clear later. The coefficient u (p)
for p 6= 0 can be obtained from Eq. (6.5) so that
r r r 
m m m
u (p) = L (p) u (0) = L (p) i 0
p0 p0 2
r
1

u (p) = im 0
L (p) 0 (6.25)
2p
similarly we obtain r
1

v (p) = im L (p) 0 (6.26)
2p0
we recall that L (p) is the standard boost that carries the standard four momentum k = (0, 0, 0, m) for a
massive particle to the four-momentum p . Such a boost is given by Eq. (1.192) page 43

Li k (p) = ik + ( 1) pbi pbk


p
Li 0 (p) = L0 i (p) = pbi 2 1 ; L0 0 =
p
p2 + m2
pbi pi / |p| , (6.27)
m
then we have
r r
i pi p 2 pi p2 + m2 pi p2 pi
L 0 (p) = 1= 1 = =
|p| |p| m2 |p| m2 m
q
p
0 p2 + m2 (p0 )2 p0
L 0 = = = =
m m m
in the last step we have used the fact that p0 > 0. We finally obtain
p
L 0 = (6.28)
m
166 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

substituting (6.28) in (6.25) we find


r
1
u (p) = ip (6.29)
2p0

similarly, from Eq. (6.26) we obtain


r
1
v (p) = ip (6.30)
2p0

substituting (6.29, 6.30) in the expressions (6.2, 6.3) for the annihilation and creation fields (in which there is no
sum over ) we find
Z r Z r
3/2 1 1
+
(x) = (2) d p (ip ) 3
0
a (p) eipx = (2)3/2
3
d p a (p) eipx (6.31)
2p 2p0
 Z r 
+ 3/2 3 1 ipx
(x) = (2) d p a (p) e (6.32)
2p0

Similarly we obtain2
 Z r 
3/2 3 1 ipx
(x) = (2) d p a (p) e
2p0

Taking into account Eqs. (5.5, 5.6) we obtain

+ (x) = + (x) ; (x) = (x) (6.33)

where (x) are the scalar annihilation and creation fields that we obtained for a spinless particle. So the vector
annihilation and creation fields here are nothing but the derivatives of the scalar annihilation and creation fields
for a spinless particle. It is immediate that the causal vector field for a spinless particle is also simply the derivative
of the causal scalar field
(x) = + (x) + (x) = (x) (6.34)

6.3 Spin one vector fields


We have already seen that for the case j = 1, we have [see Eqs. (6.20, 6.22)]

u0 (0, ) = v 0 (0, ) = 0 (6.35)

Before continuing, it is important to take into account that in Eqs. (6.7, 6.8) the matrix representations of the
generators on the LHS correspond to the irreducible representation j in the canonical basis |j, i, while the
RHS provides the matrix representations of the generators in the cartesian basis described by Eqs. (6.9, 6.10).
Therefore, for our present work we need besides the matrix representations of the generators in the cartesian
basis Eqs. (6.9, 6.10), the matrix representations for the generators associated with the irreducible representation
j = 1 in the canonical basis. Ordering the canonical basis as

|j = 1, = 1i , |j = 1, = 0i , |j = 1, = 1i

2
At this step we understand the introduction of the i factor in the choice of phase (6.24).
6.3. SPIN ONE VECTOR FIELDS 167

the matrix representation of the generators are given by



0 1 0 0 i 0
1 1
(J1 )(j=1) = 1 0 1 ; (J2 )(j=1) = i 0 i
2 0 1 0 2 0 i 0

1 0 0 (j=1) 1 0 0
(J3 )(j=1) = 0 0 0 ; J2 = 2 0 1 0
0 0 1 0 0 1

0 2 0 0 0 0
(J+ )(j=1) = 0 0 2 ; (J )(j=1) = 2 0 0 (6.36)
0 0 0 0 2 0
By setting = 0 in Eq. (6.7), and using Eq. (6.9) and (6.35), we obtain

X (j)
u (0, ) (Jk ) = (Jk ) u (0, )

X (j)
u (0, ) (Jk )0 = (Jk ) u (0, 0)

X (j)
(Jk )i n un (0, 0) = ui (0, ) (Jk )0

using (6.10), and taking into account that = 1, 0, 1; we have


(j=1) (j=1) (j=1)
iink un (0, 0) = ui (0, +1) (Jk )1,0 + ui (0, 0) (Jk )0,0 + ui (0, 1) (Jk )1,0 (6.37)
it is convenient to redefine as a label of matrix elements, hence 1, 0, 1 1, 2, 3
(j=1) (j=1) (j=1)
iikn un (0, 0) = ui (0, +1) (Jk )1,2 + ui (0, 0) (Jk )2,2 + ui (0, 1) (Jk )3,2 (6.38)
using (6.36) and setting k = 3, equation (6.38) becomes
(j=1) (j=1) (j=1)
ii3n un (0, 0) = ui (0, +1) (J3 )1,2 + ui (0, 0) (J3 )2,2 + ui (0, 1) (J3 )3,2
ii3n un (0, 0) = 0 (6.39)
for i = 1, 2 in (6.39) we find
i132 u2 (0, 0) = 0 ; i231 u1 (0, 0) = 0
u1 (0, 0) = u2 (0, 0) = 0 (6.40)
for i = 3 Eq. (6.39) gives no information. In addition for n = 3 in Eq. (6.39) we have no information either.
Therefore, u3 (0, 0) is arbitrary. On the other hand, Eqs. (6.35, 6.40) say that u (0, 0) = 0 for 6= 3. Ordering
the components as 1, 2, 3, 0; we shall normalize the fields so that the vector u (0, 0) takes the value

0
1 0


u (0, 0) = (6.41)
2m 1
0
now setting k = 1 in Eq. (6.38)
(j=1) (j=1) (j=1)
ii1n un (0, 0) = ui (0, +1) (J1 )1,2 + ui (0, 0) (J1 )2,2 + ui (0, 1) (J1 )3,2
ui (0, +1) + ui (0, 1)
ii1n un (0, 0) = (6.42)
2
168 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

taking i = 1 in (6.42) we have

0 = u1 (0, +1) + u1 (0, 1) (6.43)

with i = 2 in (6.42) and taking into account the normalization (6.41), we find


i 2213 u3 (0, 0) = u2 (0, +1) + u2 (0, 1)
1
i 2 = u2 (0, +1) + u2 (0, 1)
2m
i
= u2 (0, +1) + u2 (0, 1) (6.44)
m

setting i = 3 in (6.42), and using (6.40) we obtain

u3 (0, +1) + u3 (0, 1)


i312 u2 (0, 0) =
2
0 = u3 (0, +1) + u3 (0, 1) (6.45)

Finally, by setting k = 2 in Eq. (6.38), and using (6.36) we find

(j=1) (j=1) (j=1)


ii2n un (0, 0) = ui (0, +1) (J2 )1,2 + ui (0, 0) (J2 )2,2 + ui (0, 1) (J2 )3,2
i ui (0, +1) + i ui (0, 1)
ii2n un (0, 0) =
2
n i i
2i2n u (0, 0) = u (0, 1) u (0, +1) (6.46)

and for i = 1 in Eq. (6.46) and using the normalization (6.41), we get


2123 u3 (0, 0) = u1 (0, 1) u1 (0, +1)
1
2 = u1 (0, 1) u1 (0, +1)
2m
1
= u1 (0, 1) u1 (0, +1) (6.47)
m

with i = 2 in Eq. (6.46)

0 = u2 (0, 1) u2 (0, +1) (6.48)

for i = 3 in Eq. (6.46)


2321 u1 (0, 0) = u3 (0, 1) u3 (0, +1)
0 = u3 (0, 1) u3 (0, +1) (6.49)
6.3. SPIN ONE VECTOR FIELDS 169

picking up all relations obtained so far [Eqs. (6.35, 6.41, 6.43, 6.44, 6.45, 6.47, 6.48, 6.49)] we write

u0 (0, ) = 0 (6.50)

0
1
0
u (0, 0) = (6.51)
2m 1
0
u1 (0, +1) + u1 (0, 1) = 0 (6.52)
i
u2 (0, +1) + u2 (0, 1) = (6.53)
m
3 3
u (0, +1) + u (0, 1) = 0 (6.54)
1
u1 (0, 1) u1 (0, +1) = (6.55)
m
2 2
u (0, 1) u (0, +1) = 0 (6.56)
3 3
u (0, 1) u (0, +1) = 0 (6.57)

where we have ordered the components as 1, 2, 3, 0. These relations are enough to determine completely the
vectors u (0, 0) , u (0, +1) and u (0, 1), so we have obtained all vectors3 of the form u (0, ). For example,
by substracting equations (6.52, 6.55) and substracting equations (6.53, 6.56) we obtain

1 i
2u1 (0, +1) = ; 2u2 (0, +1) = (6.58)
m m

further by substracting equations (6.54, 6.57) and taking Eqs. (6.50) we find

2u3 (0, +1) = 0 ; u0 (0, +1) = 0 (6.59)

from Eqs. (6.58, 6.59) we have the complete vector u (0, +1). Ordering the components as 1, 2, 3, 0; it becomes

1
1 1 +i
u (0, +1) =
2 m 0
0

by adding the equations that we substract previously we obtain the complete vector u (0, 1). Ordering the
components as 1, 2, 3, 0; we obtain
1
1 1 i
u (0, 1) =
2 m 0
0

A totally similar exercise can be done with the coefficients v k from Eqs. (6.8) to obtain all vectors of the form
v (0, ).
In summary, by setting = 0 in Eqs. (6.7, 6.8) and examining the behavior of J3 we find

u0 (0, 0) = 0 ; ui (0, 0) = u3 (0, 0) 3 k uk (6.60)


0 i 3 3
v (0, 0) = 0 ; v (0, 0) = v (0, 0) k uk (6.61)
3
Notice however that we can determine such vectors only after fixing the value of u3 (0, 0).
170 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

so that the vectors u (0, 0) and v (0, 0) for = 0 are in the three-direction. A suitable normalization of the
fields, permits us to take these vectors as

0
1
0
u (0, 0) = v (0, 0) = (6.62)
2m 1

0

where the four components are ordered as 1, 2, 3, 0. Equations (6.7, 6.8) [or equivalently Eqs. (6.50)-(6.57)] also
provide the values of u (0, 1)4 . We then obtain

1
1 1 +i

u (0, +1) = v (0, 1) =
2 2m 0
0

1
1 1 i

u (0, 1) = v (0, +1) = (6.63)
2 2m 0
0

and applying Eq. (6.5) we obtain


r r h i
2m 1
u (p, ) = L (p) u (0, ) = L (p) 2m u (0, )
2p0 2p0
r h i
1
u (p, ) = L (p) e (0, ) ; e (0, ) 2m u (0, )
2p0

a similar exercise can be done from Eq. (6.6) for vk , we finally obtain
1/2
u (p, ) = v (p, ) = 2p0 e (p, ) (6.64)

e (p, ) L (p) e (0, ) ; e (0, ) 2m u (0, ) (6.65)

from the definition (6.65)5 and expressions (6.62, 6.63) we obtain the explicit form of e (0, )

0 1 1
0 +i
e (0, 0) = , e (0, +1) = 1 1
, e (0, 1) = i (6.66)
1 2 0 2 0
0 0 0

substituting (6.64) into (6.2, 6.3) we obtain the annihilation and creation fields

1 XZ d3 p
+ (x) = (x) = q p e (p, ) a (p, ) eipx (6.67)
3 2p 0
(2)

4 (1) (1)
Alternatively, the other components can be found by calculating the effect of the raising and lowering operators J1 J2 on u
and v.
5
Note that definition (6.65) simply establishes that the vector e at an arbitrary three-momentum p is connected to e at zero
three-momentum, by means of the standard boost that passes from zero three-momentum to the three-momentum p.
6.3. SPIN ONE VECTOR FIELDS 171

it is clear that the fields + (x) and + (y) commute or anticommute for all x and y. However, + (x) and
(y) do not. Their commutator for bosons or anticommutator for fermions yields
" #
 +  1 X Z d3 p X Z d3 p  
(x) , (y) = p e (p, ) a (p, ) eipx , p e p , a p , eip y
(2)3 2p0 2p 0

X Z 3 X Z 3 h i
1 d p d p  
= 3
p e (p, ) eipx p e p , eip y a (p, ) , a p ,
(2) 2p 0 2p 0

Z Z
1 X d3 p X d3 p  ip y 
= 3
p e (p, ) eipx
p e p , e p p
(2) 2p0 2p0
Z 3
 +  1 d p ip(xy) X
(x) , (y) = e e (p, ) e (p, )
(2)3 2p0

we write this commutation or anticommutation relation as


Z
 +  d3 p
(x) , (y) = 3 eip(xy) (p) (6.68)
(2) 2p 0
X

(p) e (p, ) e (p, ) (6.69)

based on Eq. (6.66) we can obtain (0). For example we have

3
X
12 (0) = e1 (0, ) e2 (0, ) = e1 (0, 0) e2 (0, 0) + e1 (0, +1) e2 (0, +1) + e1 (0, 1) e2 (0, 1)
=1

 0 
12 (0) = 0 12 12 i
2
i2
       
12 1 1 1 1
(0) = 0 0 + i + i = 0
2 2 2 2

with that procedure for each component we obtain



1 0 0 0
0 1 0 0
(0) =
0
(6.70)
0 1 0
0 0 0 0

applied on an arbitrary four vector, we have


1
1 0 0 0 x x1
0 1 0
0 x2 x2
(0) x =
0
=
0 1 0 x3 x3
0 0 0 0 x0 0
(0) x = (x, 0)

we can see that (0) is the projection matrix on the space orthogonal to the time direction.
By using Eqs. (6.65, 6.66, 6.69) we can obtain (p)
172 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

X X h i
(p) e (p, ) e (p, ) [L (p) e (0, )] L (p) e (0, )

h i h i
= [L (p) e (0, +1)] L (p) e (0, +1) + [L (p) e (0, 0)] L (p) e (0, 0)
h i
+ [L (p) e (0, 1)] L (p) e (0, 1)
  
= L 1 (p) e1 (0, +1) + L 2 (p) e2 (0, +1) L 1 (p) e1 (0, +1) + L 2 (p) e2 (0, +1)
  
+ L 3 (p) e3 (0, 0) L 3 (p) e3 (0, 0)
  
+ L 1 (p) e1 (0, 1) + L 2 (p) e2 (0, 1) L 1 (p) e1 (0, 1) + L 2 (p) e2 (0, 1)
  
1 i 1 i
= L 1 (p) L 2 (p) L 1 (p) + L 2 (p) + L 3 (p) L 3 (p)
2 2 2 2
  
1 i 1 i
+ L 1 (p) L 2 (p) L 1 (p) + L 2 (p)
2 2 2 2

1
(p) = [L 1 (p) + iL 2 (p)] [L 1 (p) + iL 2 (p)] + L 3 (p) L 3 (p)
2
1
+ [L 1 (p) iL 2 (p)] [L 1 (p) + iL 2 (p)]
2
1 i i 1
= L 1 (p) L 1 (p) L 1 (p) L 2 (p) + L 2 (p) L 1 (p) + L 2 (p) L 2 (p) + L 3 (p) L 3 (p)
2 2 2 2
1 i i 1
+ L 1 (p) L 1 (p) + L 1 (p) L 2 (p) L 2 (p) L 1 (p) + L 2 (p) L 2 (p)

2 2 2 2
(p) = L 1 (p) L 1 (p) + L 2 (p) L 2 (p) + L 3 (p) L 3 (p) (6.71)

and applying (6.27) in Eq. (6.71) we have

ij (p) = Li 1 (p) Lj 1 (p) + Li 2 (p) Lj 2 (p) + Li 3 (p) Lj 3 (p)


= [i1 + ( 1) pbi pb1 ] [j1 + ( 1) pbj pb1 ] + [i2 + ( 1) pbi pb2 ] [j2 + ( 1) pbj pb2 ]
+ [i3 + ( 1) pbi pb3 ] [j3 + ( 1) pbj pb3 ]
= i1 j1 + ( 1) i1 pbj pb1 + ( 1) pbi pb1 j1 + ( 1)2 pbi pbj pb21
+i2 j2 + ( 1) i2 pbj pb2 + ( 1) pbi pb2 j2 + ( 1)2 pbi pbj pb22
+i3 j3 + ( 1) i3 pbj pb3 + ( 1) pbi pb3 j3 + ( 1)2 pbi pbj pb23

ij (p) = i k k j + ( 1) k i pbj pbk + ( 1) pbi pbk k j + ( 1)2 pbi pbj pbk pbk
= i j + ( 1) pbj pbi + ( 1) pbi pbj + ( 1)2 pbi pbj pbk pbk
h i pi pj
= gij + ( 1) pbi pbj 2 + ( 1) pbk pbk = gij + ( 1) 2 [2 + ( 1)]
p

where we have used the fact that pbk is normalized so that pbk pbk = 1.
pi pj pi pj 
ij (p) = gij + ( 1) 2 ( + 1) = gij + 2 2 1
p p
 2 2

pi pj p + m pi pj p2
= gij + 2 1 = g ij
+
p m2 p2 m2
pi pj
ij (p) = gij + 2 (6.72)
m
6.3. SPIN ONE VECTOR FIELDS 173

From Eqs. (6.28, ,6.27, 6.71) we also have

i0 (p) = Li 1 (p) L0 1 (p) + Li 2 (p) L0 2 (p) + Li 3 (p) L0 3 (p)


p1 p2 p3
= [i1 + ( 1) pbi pb1 ] + [i2 + ( 1) pbi pb2 ] + [i3 + ( 1) pbi pb3 ]
m m m
 p p2 p3  ( 1)
1
= i1 + i2 + i3 + pbi [b
p1 p1 + pb2 p2 + pb3 p3 ]
m m m m

pk ( 1) pi ( 1) pi ( 1) pi
i0 (p) = i k + pbi pbk pk = + pbi pbk pbk |p| = + |p|
m m m m m m |p|
pi ( 1) i p0 i
= + p = pi = p
m m m m2
p i p0
i0 (p) = 0i (p) = 2 (6.73)
m

finally, from Eqs. (6.28, 6.71) we have

00 (p) = L0 1 (p) L0 1 (p) + L0 2 (p) L0 2 (p) + L0 3 (p) L0 3 (p)


p1 p1 p2 p2 p3 p3 p2 p2 + m2 p0 p0
= + + = 2 = 1 = 1 +
mm mm mm m m2 m2
0
p p 0
00 (p) = g00 + 2 (6.74)
m

picking up Eqs. (6.72, 6.73, 6.74) we obtain

p p
(p) = g + (6.75)
m2
In addition Eqs. (6.65, 6.75) shows that (p) is the projection matrix on the space orthogonal to the four-
vector p

p p p m2 p
(p) p = g p + = p

m2 m2
(p) p = 0

in a similar way we can show that the four-vectors e (p, ) are orthogonal to p . To see it, we use Eqs. (6.65,
6.66). Using these equations for = 0 we obtain

e (p, ) p L (p) e (0, ) p


e (p, 0) p L (p) e (0, 0) p = L 3 (p) e3 (0, 0) p = L 3 (p) p = L0 3 (p) p0 + Li 3 (p) pi
p3   p3
= p0 + i 3 + ( 1) pbi pb3 pi = p0 + p3 + ( 1) pbi pi pb3
m  0  m  0 
p3 0 p i p3 0 p
= p + p3 + 1 pb pbi |p| pb3 = p + p3 + 1 p3
m m m m
e (p, 0) p = 0

and we can proceed similarly for = 1. Then we obtain

e (p, ) p = 0 (6.76)
174 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

On the other hand, the commutator or anticommutator (6.68) can be written in terms of the function +
defined in Eq. (5.9), as
Z Z  
 +  d3 p d3 p p p
(x) , (y) = eip(xy)
(p) = eip(xy)
g
+
(2)3 2p0 (2)3 2p0 m2
Z Z
d3 p 1 d3 p
= g 3 eip(xy)
+ 2 3 p p eip(xy)
(2) 2p 0 m (2) 2p 0
Z 3 Z
d p 1 d3 p
= g 3 eip(xy) 2 3 eip(xy)
(2) 2p 0 m (2) 2p 0

 
 + 
(x) , (y) = g + (x y) (6.77)
m2
by now, what really matters is that this expression does not vanish for x y space-like and is even in x y. Hence,
we can repeat the reasoning of section 5 to construct a causal field: we start by forming a linear combination of
annihilation and creation fields
v (x) + (x) + (x) (6.78)
and for space-like separations of x and y, we obtain
 

[v (x) , v (y)] = [1 1] g + (x y)
m2
h i  

2 2
v (x) , v (y) || ||
g + (x y)
m2

once again, for both commutators or anticommutators to vanish at space-like separations x y, it is necessary
and sufficient that the spin one particles be bosons and that || = ||. By a suitable choice of the phase of the
one-particle states we can settle = , and drop the common factor by redefining the overall normalization of
the field. With those facts, the causal vector field (6.78) for a massive particle of spin one yields

v (x) = + (x) + + (x) (6.79)

note that v (x) is real


v (x) = v (x) (6.80)

6.4 Spin one vector fields with internal symmetries


The previous framework is suitable for totally neutral spin one particles. If the particles described carry a non-zero
value of some conserved quantum number Q, we cannot construct an interaction that conserves Q out of such a
field. Hence, we must suppose that there is another boson of the same mass and spin, which carries an opposite
value of Q. We then construct the causal field in the form

v (x) = + (x) + c+ (x) (6.81)

which can be expanded as

1 XZ d3 p h i

v (x) = q p e (p, ) a (p, ) eipx + e (p, ) ac (p, ) eipx (6.82)
(2)3 2p0

where the superscript c indicates operators that create the antiparticle that is charge-conjugate to the particle
annihilated by + (x). This is a causal field but not real (hermitian) anymore. Once again, we can use Eq. (6.82)
6.4. SPIN ONE VECTOR FIELDS WITH INTERNAL SYMMETRIES 175

for the case of a purely neutral spin one particle that it is its own antiparticle. In that case we should take simply
ac (p) = a (p) . In either case, the commutator of a vector field with its adjoint gives
h i 


v (x) , v (y) = g (x y) (6.83)
m2
where (x y) is the function defined by Eq. (5.32).

6.4.1 Field equations for spin one particles


The real and complex fields we have constructed obey some interesting field equations. For example, since p in
the exponential of Eq. (6.82) represents a physical four-momentum, we have p2 = m2 . Thus, v (x) are the
components of a field with definite mass. Therefore, according with the discussion given at section 4.6, the field
v (x) must satisfy the Klein-Gordon equation (4.72) page 150

 m2 v (x) = 0 (6.84)
as it also happens for scalar fields.
On the other hand, we have already seen that [see Eq. (6.76)]
e (p, ) p = 0 (6.85)
therefore by applying on both sides of (6.82) and using (6.85), we see that the spin one field v (x) obeys
another field equation
v (x) = 0 (6.86)
Note that when v (x) corresponds to the four-vector potential A (x) of electrodynamics, and taking m 0, the
Klein Gordon equation (6.84) becomes the wave equation, while equation (6.86) becomes the Lorentz gauge. In
other words, in the limit of small mass, Eqs. (6.84, 6.86) become the equations for the potential four-vector of
electrodynamics in the Lorentz gauge.
Notwithstading we do not obtain the electrodynamics by simply taking the limit of m going to zero. We can
see that by taking the rate of production of a spin one particle through an interaction density H = J v where J
is an arbitrary four-momentum current6 . Squaring the matrix elements and summing over the three-component
of spins we find a rate proportional to
3
X 2
|hJ i e (p, ) | = hJ i hJ i (p)
=1

where p is the three-momentum of the emitted spin-one particle, and hJ i is the matrix element of the current
(say at x = 0) between the initial and final states of all other particles. From Eq. (6.75) we see that (p)
contains a term of the form p p /m2 that clearly blows up when m 0. The only way to avoid the problem is
by assuming that hJ i p vanishes, which in the coordinate space is equivalent to
J = 0
which is the continuity equation that leads to the conservation of the current J or more precisely, the conservation
of the generalized charge Z Z
Q dV = J 0 dV

We can see the need for the conservation of the current by counting degrees of freedom: A massive spin one
particle has three spin states of helicities = +1, 0, 1. By contrast, any massless particle like the photon have
only two helicity states 1. Thus, the current conservation ensures that the zero helicity states of the spin-one
particle (of very small mass) are not emitted in the limit of zero mass.
6
This will be the form of any interaction density that depends linearly on the field v (x). Of course, J must be a four-vector for
H to be Lorentz invariant. Further for v (x) hermitian, J (x) must be hermitian as well.
176 CHAPTER 6. CAUSAL VECTOR FIELDS FOR MASSIVE PARTICLES

6.5 Inversion symmetries for spin-one fields


The procedure is quite similar to the case of scalar fields. To evaluate the effect of space inversion, we need a
formula that connects e (p, ) with e (p, ). To do it, we shall use Eq. (1.265) page (58)

L (p) = P L (p) P (6.87)

as well as Eq. (6.65) and the fact that e0 (0, ) = 0, we find

e (p, ) L (p) e (0, ) = L m (p) em (0, ) = L m (p) P m m em (0, )


= L (p) P e (0, ) = L (p) P e (0, ) = (P L (p) P ) P e (0, )
= P L (p) (P P ) e (0, ) = P L (p) e (0, ) = P L (p) e (0, )
e (p, ) = P e (p, )

Then we obtain the desired formula


e (p, ) = P e (p, ) (6.88)

on the other hand, to to evaluate the time-reversal effects, we need to relate e (p, ) with e (p, ). To do
it, we use the identity
(1)1+ e (0, ) = e (0, ) (6.89)

which can be checked explicitly from Eqs. (6.66). For example for = 1, equations (6.66) yield

1
1 i
(1)1+(1) e (0, +1) = e (0, +1) =
= e (0, 1)

2 0
0

and similarly for = 0, +1. Combining Eqs. (6.88, 6.89) and using the definition (6.65) we find

(1)1+ e (p, ) = (1)1+ [L (p) e (0, )] = (1)1+ L (p) e (0, )


= L (p) e (0, ) = e (p, ) = P e (p, )

once again we obtain the relation desired

(1)1+ e (p, ) = P e (p, ) (6.90)

from these results, along with the transformation properties of creation and annihilation operators under space-
inversion [Eq. (3.28) page 122] and time-reversal [see Eq. (3.29)] we obtain the space-time inversion transformation
properties of the annihilation and creation fields (6.67).
For example, for space inversion the annihilation field (6.67) transforms as

1 XZ d3 p  
1
P +
(x) P = q p e (p, ) P a (p, ) P 1 eipx
2p 0
(2)3

1 XZ d3 p
e (p, ) [ a (p, )] ei(pxp x )
0 0
= q p
2p 0
(2)3
6.5. INVERSION SYMMETRIES FOR SPIN-ONE FIELDS 177

changing p p, and using (6.88) we obtain

XZ d3 p
e (p, ) a (p, ) ei(pxp x )
+ 1 0 0
P (x) P = q p
(2)3 2p0

XZ
d3 p
[P e (p, )] a (p, ) ei[(p,p )(x,x )]
0 0
= q p
2p 0
(2)3

1 X Z d3 p
= P q p e (p, ) a (p, ) ei[pPx]
2p0
(2)3

then we obtain finally


P + (x) P 1 = P + (Px) (6.91)
As before, for the causal fields to be transformed into other fields with which they commute at space-like
separations, it is necessary that the intrinsic space inversion, time-reversal and charge-conjugation phases for
spin-one particles and their antiparticles be related by

c = (6.92)
c
= (6.93)

= (6.94)

and all phases must be real if the spin-one particle is its own antiparticle. Under the phase conditions (6.92, 6.93,
6.94) the causal vector field (6.82) has the following properties

P v (x) P 1 = P v (Px) (6.95)


1
Cv (x) C = v (x) (6.96)
1
T v (x) T = P v (Px) (6.97)

In particular, Eq. (6.95) says that a vector field that transforms as a polar vector, with no extra phases or signs
multiplying the matrix P , describes a spin-one particle with intrinsic parity = 1. Effectively, if the three
spatial components of v (x) define a polar three-vector field we have

P v i (x) P 1 = v i (Px) (6.98)

while equation (6.95) yields

P v i (x) P 1 = P i v (Px) = P i i v i (Px) = v i (Px) (6.99)

equating Eqs. (6.98, 6.99) we obtain = 1. We recall once again that the relations developed in this section are
valid only if the inversion symmetries involved, are good symmetries of the physical system.
Chapter 7

Causal Dirac fields for massive particles

We have taken the point of view that the structure and properties of any quantum field theory are given by the
irreducible representation of the Homogeneous Lorentz group under which it transforms. Following that principle,
we should mention that from the mathematical point of view, there are two broad classes of representations of
the rotation and Lorentz groups (more precisely, of their covering groups) that are generically called tensor and
spinor representations.
We shall describe briefly the difference between the spinor and tensor representations of SO (3) [or more
precisely, of its covering group SU (2)]. We can use (for instance) the Euler angles to define an element of SO (3).
On one hand, for the representations with j integer (tensor representations), we can use the same Euler angles
and obtain a one-to-one representation with the required properties of periodicity. On the other hand, for spinor
representations (with j being a half-odd integer) we have for one of the Euler angles that U ( + 2) = U ()
which has no the periodicity required on geometrical grounds. Owing to it these representations are discarded
when we deal with space variables, this is the reason to rule out half-odd integer values of the orbital angular
momentum. We can recover the one-to-one character of the representation if we define an extended manifold
in which 0 4, in that way we arrive to the group SU (2) which is the universal covering group of SO (3).
A similar reasoning shows that the spinor representation of the Lorentz group cannot be used to describe
purely orbital systems. However, it was already known from non-relativistic quantum mechanics that the spinor
representation of SO (3) was adequate to describe the electronic spin. Therefore, it is reasonable to use the spinor
representation of the Lorentz group to describe relativistic electrons. Historically, the spinor representation was
first introduced in Physics by Dirac in his theory of relativistic electrons.

7.1 Spinor representations of the Lorentz group


In this section, we shall treat the spinor representation (that we call the Dirac formalism) from the mathematical
point of view. We do it because according with our approach, we start with the representation to construct the
fields and then predict what kind of particles can be described with this kind of fields.
Then we shall start by forming a representation U ()
 
U () U = U

in a similar way in which we characterized the unitary representations U (). That is by starting with infinitesimal
transformations1 as in Eqs. (1.88, 1.89), page 25

= + (7.1)
= (7.2)
1
We omit the generators of translations because we shall deal only with the homogeneous Lorentz group.

178
7.1. SPINOR REPRESENTATIONS OF THE LORENTZ GROUP 179

so that the infinitesimal transformations can be written in terms of antisymmetric generators J as in Eqs. (1.99,
1.101), page 27
i
D (1 + ) = 1 + J ; J = J (7.3)
2
the set of matrices J satisfy the commutation relations (1.123), page 30

i [J , J ] = g J g J g J + g J (7.4)

in order to find the set of matrices J , we first construct matrices that satisfies the anticommutation
relations
{ , } = 2g (7.5)

and we define by now


i
J = [ , ] (7.6)
4
which has the antisymmetry required for J .
It is convenient to define the commutator between two matrices. By using Eq. (7.5) we can show that

+ = 2g
+ 2 = 2g
[ , ] = 2 (g ) (7.7)

then from Eqs. (7.6, 7.7) we have


 
i i i i
[J , ] = [ , ] , = [ , ] = [ , ] [ , ]
4 4 4 4
i i i i
= [ , ] + [ , ] [ , ] [ , ]
4 4 4 4
i i i i
= (g ) + (g ) (g ) (g )

2 2 2 2
i
[J , ] = [ g + g g g ]
2
i
+ [ + + ]
2

and using (7.5) we have

i
[J , ] = [ ( + ) + ( + ) ]
2
i
= [ { , } + { , } ]
2
= i [g + g ]

obtaining finally,
[J , ] = i g + i g (7.8)

from Eq. (7.8) we can verify in turn that Eq. (7.6) satisfies the required commutation relations (7.4). We do it
as follows
180 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

 
i 1 1
i [J ,J ] = i J , [ , ] = [J , [ , ]] = [J , ]

4 4 4
1 1
= [J , ] [J , ]
4 4
1 1 1 1
= [J , ] + [J , ] [J , ] [J , ]
4 4 4 4
1 1
i [J , J ] = (i g + i g ) + (i g + i g )

4 4
1 1
(i g + i g ) (i g + i g )
4 4
i i i i
i [J , J ] = g + g + g g
4 4 4 4
i i i i
g + g + g g
4 4 4 4
i i
i [J , J ] = ( + ) g + ( ) g
4 4
i i
+ ( ) g + ( ) g

4 4
i i i i
= [ , ] g + [ , ] g + [ , ] g + [ , ] g
4 4 4 4
= J g J g J g J g
i [J , J ] = g J g J + g J g J
so that J are valid representations of the generators of the homogeneous Lorentz group.
It can be proved that the matrices are irreducible, in the sense that there is not any proper subspace
that is left invariant under all these matrices2 . If these matrices were reducible, we would be able to choose some
smaller sets of fields components, which would transform as in Eqs. (7.3) and (7.6) with an irreducible set of s.
A set of matrices that satisfy the relations (7.5) or the Euclidean analog with the Kronecker delta instead of

g , is called a Clifford algebra. From the mathematical point of view, it can be shown that the most general
irreducible representation of the Lorentz group (more precisely of its covering group) is either a tensor or a spinor
representation transforming as in Eqs. (7.3) and (7.6) or a direct product of a spinor and a tensor.
From Eq. (7.3), and using (7.2, 7.8) we can see how transforms under a homogeneous Lorentz transformation
      
i i i i
D () D 1 () = 1 + J 1 J = + J 1 J
2 2 2 2
i i 
= + J J + O 2
2 2
i i 
= + J J + O 2
2 2
i  i 
= + [J , ] + O 2 = + (i g + i g ) + O 2

2 2
1 1 1 1
= + g g = + + g
2 2 2 2
1 1
= + + = +
2 2
1
D () D () = ( + )
2
We should emphasize however, that it does not mean that the group representation is irreducible, since the matrices are not
the generators of the group [recall that the generators are the operators J defined in Eq. (7.6)]. Indeed, we shall see later that the
representation we are constructing is reducible.
7.1. SPINOR REPRESENTATIONS OF THE LORENTZ GROUP 181

which according with Eq. (7.1) gives


D () D 1 () = (7.9)
and comparing with Eq. (1.96) page 27, we can see that
transform as a four-vector operator3 . Similarly,
the unit matrix is obviously a scalar
D () 1D 1 () = 1 (7.10)
from the definition (7.6) of the generators and the rule of transformation (7.9) we have
i i
J D () J D 1 () = D () [ , ] D 1 () = D () [ ] D 1 ()
4 4
i 1 1

= D () D () D () D ()
4
i      
= D () D 1 () D () D 1 () D () D 1 () D () D 1 ()
4
i i
= {[ ] [ ] [ ] [ ]} = { }
4 4
i
= [ , ] = J
4
hence, under a homogeneous proper orthochronus Lorentz transformation, the generators J transform as

D () J D 1 () = J (7.11)

then Eqs. (7.3, 7.11) say that J is an antisymmetric tensor. Other totally antisymmetric tensors can be
constructed from as follows:

A [ ] (7.12)
[ ]
P (7.13)

where the brackets indicate a sum over all permutations of the indices within the brackets, with a minus sign for
odd permutations. For example, Eq. (7.12) can be written explicitly as [homework write Eq. (7.13) explicitly]

A + + (7.14)

by using repeatedly Eq. (7.5) we can express any product of matrices as a sum of antisymmetrized products of
s times a product of metric tensors. Therefore, the totally antisymmetric tensors form a complete basis for the
set of all matrices that can be constructed from the Dirac matrices.
By defining the matrix
i 0 (7.15)
a similarity transformation of the Dirac matrices through the matrix yields
 1 1  0 i  1
i 1 = i 0 i i 0 = 0i 0 = , i0 0
  1  1
= g0i i 0 0 = i0 0 = i
 1 1
0 1 = i 0 0 i 0 = 00 0 = 0

then we obtain
i 1 = i ; 0 1 = + 0 (7.16)
that we can rewrite as  
1 = , 0 P , 0 (7.17)
3 0
 
We recall that it means that the array i , has the same property of transformation as the four-vector operator Pi , H , where
each represents a single operator at the Minkowski space (see section 1.8.1 page 26).
182 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

thus can be considered as a parity transformation for the four-vector operator . Note that here we are
ordering the label as = 0, 1, 2, . . .(this order is more convenient because the dimensionality is arbitrary so far).
We can say alternatively that anticommutes with i and commutes with 0 . Consequently, the same similarity
transformation applied to any product of matrices gives a minus or plus sign according whether the product of
s contain even or odd number of s with space-like indices, respectively. For instance,

0 1 0 2 1 = 0 1 0 2 ; 0 1 3 2 1 = 0 1 3 2

in particular we have
J ij 1 = J ij ; J i0 1 = J i0 ; J 00 = 0 (7.18)
all the properties we shall develop in this section are valid in any number of space-time dimensions and for any
metric g . However, in four spacetime dimensions we have the particular feature that no totally antisymmetric
tensor can have more than four indices [homework!!(14a)], so that the complete sequence of tensors is given by

1, , J , A , P (7.19)

Each of these tensors transform differently under Lorentz and/or parity transformations so that they are all linearly
independent. Another way to check such a linear independence is the following: we define the scalar product of
two matrices by the trace of the product that is
X
(B, C) = T r [B C] = (B C) = B C (7.20)

it can be shown that the rule (7.20) satisfies the axioms of a scalar product, and that the matrices (7.19) form an
orthogonal set under this scalar product[homework!!(15)]4 . Now an orthogonal set is linearly independent unless
some of its vectors are null. However, none of the matrices (7.19) vanish, since each component of each of these
tensors is proportional to a product of different matrices, and in turn such a product has a square equal to
plus or minus the product of the associated squares, and then equal to 1.
Now we count the number of linearly independent components on each of the matrices contained in the set
(7.19). The identity has only one independent component, has four, J has six, A has four, while P
has only one independent component [homework!!(14b) explain this counting in detail]. So we have a total of 16
independent components. Thus an arbitrary 44 matrix can be written as a complex linear combination of the
16 linearly independent matrices given by (7.19).
According with the Clifford algebra (7.5) we have
2 2
i = 144 ; 0 = 144 (7.21)

and recalling the definition of i 0 we obtain that

2 = 144 (7.22)

Then we shall not distinguish between and 1 from now on. Hence, all similarity transformations through the
parity operator , will be written as (. . .) .

7.2 Some additional properties of the Dirac matrices


It is convenient to write the totally antisymmetric tensors (7.12, 7.13) in a more suitable way. It is well-known
that in a ndimensional space there is only one linearly independent totally antisymmetric tensor of n indices.
4
It is easier to show that by using a specific representation (for instance, the one that we shall develop later). The results obtained
in a given representation are valid for any other representation, because the trace is invariant under a similarity transformation.
7.2. SOME ADDITIONAL PROPERTIES OF THE DIRAC MATRICES 183

Consequently, the totally antisymmetric matrix (7.13) must be proportional to the pseudotensor , defined as
a totally antisymmetric quantity with 0123 = +1. Setting , , , equal to 0, 1, 2, 3 respectively, we find

P 0123 = 0 1 2 3 permut (7.23)

for an arbitrary , we have that P is different from zero only if all symbols are different. Thus, for
a non-null value of P , the set of symbols is a permutation P of 0, 1, 2, 3. When writing P it is
clear that we obtain exactly the same 4! terms as in Eq. (7.23)

P = permut (7.24)

however all 4! terms in (7.24) will have opposite sign with respect to (7.23) if the permutation P is odd (P
is the permutation from the sequence 0, 1, 2, 3 to the sequence , , , ). It owes to do with the fact that for P
we assign by definition the positive sign to the sequence , while in Eq. (7.24) this sequence will be
assigned the negative sign if the permutation from 0, 1, 2, 3 to , , , is odd, then we obtain

P = P 0123 = 0 1 2 3 permut (7.25)

we now take into account that anticommutes with as long as 6= . In that case an odd number of jumps
simply gives a minus sign (as long as we jump over a different matrix). For instance we have

P 0123 = 0 1 2 3 0 1 3 2 0 2 1 3 + 0 3 1 2 other permut


= 0 1 2 3 + 0 1 2 3 + 0 1 2 3 0 1 3 2 other permut
= 0 1 2 3 + 0 1 2 3 + 0 1 2 3 + 0 1 2 3 other permut

we see that all terms are identical, therefore



P 0123 = 4! 0 1 2 3 (7.26)

substituting (7.26) in (7.25) we find


 
P = 4! 0 1 2 3 = 4!i i 0 1 2 3
then we obtain finally

P = 4!i 5 (7.27)
0 1 2 3
5 i (7.28)

in the same way A must be proportional to contracted with some matrix M . To show it, let us evaluate
the antisymmetric tensor A of Eq. (7.14), , , = 0, 1, 2

A012 0 1 2 0 2 1 1 0 2 + 2 0 1 + 1 2 0 2 1 0
= 0 1 2 + 0 1 2 + 0 1 2 + 0 1 2 + 0 1 2 + 0 1 2
A012 = 3! 0 1 2 = 3!0123 0 1 2 (7.29)

now defining
g (7.30)
and taking into account the Clifford algebra (7.5) we have
2
i = gii = 1 ; i i = i i gii = 1 No sum over i (7.31)
184 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

using (7.31) in Eq. (7.29) we have


 
A012 = 3!0123 0 1 2 3 3 = 3!i0123 i 0 1 2 3 3
A012 = 3!i0123 5 3

and we can do a similar process for all possible sets of indices , , in Eq. (7.14). It can be checked that by
setting , , equal to 0,1,2 or 0,1,3 or 0,2,3 or 1,2,3 we find

A = 3!i 5 (7.32)

it is easy to show that the square of 5 is the unit matrix and that 5 anticommutes with all

52 = 1 , {5 , } = 0 (7.33)

it can be shown as follows


     
5 , 0 = i 0 1 2 3 0 + 0 0 1 2 3 = i 0 0 1 2 3 + 0 0 1 2 3 = 0
     
5 , 1 = i 0 1 2 3 1 + 1 0 1 2 3 = i 0 1 1 2 3 0 1 1 2 3 = 0

Proceeding similarly with the other matrices we obtain {5 , } = 0. Moreover


 
52 = 0 1 2 3 0 1 2 3 = 0 1 2 3 0 1 2 3 = 0 0 1 2 3 1 2 3
2 1 2 2 2 3 2
= 0 0 1 1 2 3 2 3 = 0 0 1 1 2 2 3 3 = 0
 0 0  1 1  2 2  3 3
, , , ,
= = g00 g11 g22 g33 = 1
2 2 2 2
since 5 anticommutes with , and using the definition of in Eq. (7.15), we see that
  2
5 1 = 5 = i 0 5 i 0 = 0 5 0 = 0 5 = 5

and the commutator between and 5 becomes

[ , 5 ] = 5 5 = 5 + 5
[ , 5 ] = 2 5 (7.34)

we can obtain the commutator between the generators J and 5 by taking into account Eqs. (7.33, 7.34) and
the definition (7.6)

4i [J , 5 ] = [[ , ] , 5 ] = [ , 5 ] [ , 5 ] = [ , 5 ] + [ , 5 ] [ , 5 ] [ , 5 ]
= 2 5 + 2 5 2 5 2 5 = 2 5 2 5 2 5 + 2 5
4i [J , 5 ] = 0

Therefore, the matrix 5 is a pseudoscalar in the sense that

[J , 5 ] = 0 (7.35)
1
5 = 5 = 5 (7.36)

Note that 5 is a Casimir of the homogeneous proper orthochronus Lorentz group since it commutes with all
the generators. However it is not in general proportional to the identity at the Minkowski space because this
representation is not irreducible in such a space. We should remember that Schurs lemmas are only valid within
minimal vector spaces associated with irreducible representations.
7.3. THE CHIRAL REPRESENTATION FOR THE DIRAC MATRICES 185

From Eqs. (7.27, 7.32) we can see that the 16 independent 4 4 matrices defined in Eq. (7.19) can be
reorganized in matrices with more suitable properties as follows

1 : scalar
5 : pseudoscalar
: vector

J : antisymmetric tensor
5 : axial vector (7.37)

the notation 5 has to do with the fact that the anticommutation relations (7.33) along with (7.5) show that the
set
0 , 1 , 2 , 3 , 5

provide a Clifford algebra in five space-time dimensions.

7.3 The chiral representation for the Dirac matrices


We shall choose an explicit set of 4 4 matrices as follows
   
0 022 122 022
i ; i ; (1 , 2 , 3 ) (7.38)
122 022 022

where i are the usual Pauli matrices


     
0 1 0 i 1 0
1 = , 2 = , 3 = (7.39)
1 0 i 0 0 1

let us recall some basic properties of the Pauli matrices

[i , j ] = 2iijk k ; {i , j } = 2ij
i = i ; T ri = 0
i j = ik with i, j, k a cyclic permutation of 1, 2, 3 (7.40)

it can be shown that any other irreducible set of matrices that satisfy the Clifford algebra (7.5), are related with
this through a similarity transformation. Thus a given set of matrices define a unique irreducible inequivalent
representation of the Clifford algebra. However, several representations (though equivalent) of the matrices are
used.
By returning to our specific representation, we first observe from (7.38) that the 0 matrix is anti-hermitian
while the i matrices are hermitian
     
0 022 122 0 022 022
i = ; i =i = (7.41)
122 022 022 022
we can write it in a shorten notation as

= g no sum over (7.42)

we should keep in mind that the hermitian or anti-hermitian property of any matrix is preserved under a similarity
transformation carried out by a unitary matrix.
186 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

We shall now calculate the product of two gamma matrices. For instance from Eq. (7.38) we can obtain the
product between two Dirac matrices with space-like components
    
022 i 022 j i j 022
ij = =
i 022 j 022 022 i j
 
i j 022
ij =
022 i j

and we also have


    
0 i 022 122 022 i i 022
= =
122 022 i 022 022 i
    
i 0 022 i 022 122 i 022
= =
i 022 122 022 022 i
    
0 0 022 122 022 122 122 022
= =
122 022 122 022 022 122

picking up all these results we have


   
i j i j 022 0 i i 0 i 022
= ; = = ; 0 0 = 144 (7.43)
022 i j 022 i

from the properties of the Pauli matrices (7.40) we can verify that the representation (7.38) reproduces the Clifford
algebra (7.5) as it must be
   
 i j {i , j } 022 ij 022 ij
, = =2 = 244 = 2gij
022 {i , j } 022 ij
 i 0
, = 0 i + i 0 = 0 i 0 i = 0 = 2gi0
 0 0
, = 2 0 0 = 2 144 = 2g00 144

on the other hand, we can calculate the Lorentz group generators (7.6). From Eq. (7.38), and using Eqs. (7.40,
7.43), they are given by
 
ij i  i j i [i , j ] 022
J = , =
4 4 022 [i , j ]
   
ij i 2iijk k 022 1 k 022
J = = ijk
4 022 2iijk k 2 022 k
 
i  i 0 i  i 0 i  i i i 022
J i0 = i , 0 = i0 = + 0 i = 0 i =
4 4 4 2 2 022 i
i 
J 00 = 0 , 0 = 044
4
then we obtain for the generators in this representation
   
1 k 022 i i 022
J ij = ijk ; J i0 = ; J 00 = 044 (7.44)
2 022 k 2 022 i

note that the generators (7.44) are block-diagonal. Therefore, the Dirac matrices give a reducible representation
of the proper orthochronus Lorentz group. The four-dimensional representation is thus decomposed as the direct
sum of two irreducible two-dimensional representations with J ij = iijk J k0 .
7.3. THE CHIRAL REPRESENTATION FOR THE DIRAC MATRICES 187

For the specific representation given by Eq. (7.38) the 5 matrix becomes

5 i 0 1 2 3
     
5 022 122 022 1 022 2 022 3
= (i)
122 022 1 022 2 022 3 022
    
1 022 2 3 022 1 2 3 022
= i = i
022 1 022 2 3 022 1 2 3

and the product of Pauli matrices can be obtained by explicit calculation from (7.39) or by using the properties
(7.40)5
i
1 2 3 = (1 2 ) 3 = i3 3 = {3 , 3 } = i (7.45)
2
such that 5 in this representation becomes
 
122 022
5 = (7.46)
022 122

this representation has the advantage of reducing J and 5 to block-diagonal form (this is the so-called chiral
representation). Note that both 2 2 submatrices are proportional to the identity. It is because in the basis
chosen the representation is decomposed in two bidimensional irreducible representations. Since J and 5 have
the same block-diagonal form, equation (7.35) shows that the 2 2 submatrices that represent 5 are Casimirs in
each bidimensional subspace. Since these Casimirs are associated with irreducible representations they must be
proportional to the identity within each minimal invariant bidimensional subspace.
From the physical point of view, the chiral representation is suitable to describe particles in the ultra-relativistic
limit v c. However, in the non-relativistic limit v << c, it is more convenient to choose a representation in
which 0 is diagonal instead of 5 , which is the case in the Dirac representation6 .
The representation of the homogeneous Lorentz group that we have constructed is not unitary, since the
generators J are not all represented by hermitian matrices. In the specific representation given by Eqs. (7.38)
J ij are hermitian while J i0 are anti-hermitian. It can be seen from the fact that 0 is anti-hermitian while the
k s are hermitian [see Eq. (7.42)]
i  i h i   i i h i i
J i = + i , = i = i
4 4 4
i  i i
 i
 i 
= g g = g ,
4 4
J i = g J i (7.47)

It is also convenient to define the matrix i 0 of Eq. (7.15), such that the reality conditions are manifestly
Lorentz -invariant. In the representation (7.38) has the form
 
0 1
= (7.48)
1 0

recalling that = 1 , all similarity transformations through the parity operator can be written as (. . .) .
Using again the fact that 0 is anti-hermitian while the k s are hermitian, we see that
  
0 = 0 = i 0 0 i 0 = 0 0 0 = 0
 
k = k = i 0 k i 0 = 0 k 0 = 0 0 k = k
5
The advantage of using the properties (7.40) instead of the explicit forms (7.39), is that we can be sure that the result (7.45) is
independent of the representation chosen.
6
There is still another widely used representation: the so-called Majorana representation.
188 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

that can be written as


= (7.49)
and equation (7.49) in turn leads to
 
i i  
J = ( ) =
4 4
i h      i i
= = [( ) ( ) ( ) ( )]
4 4
i i
= [ ] = [ , ] = J
4 4
so we also obtain the identity
J = J (7.50)
it is also important to characterize the transformation of the matrices D () under the parity transformation taking
into account that such matrices are not unitary. To do it, we use the expression (7.3) page 179 for infinitesimal
Lorentz transformations, as well as Eq. (7.50)
   
i i i
D () = 1 + J = 1 J = 1 J
2 2 2
i
D () = 1 J = D 1 ()
2
and since any finite transformation can be written as a product of infinitesimal transformations, we can extend
this result for any finite proper orthochronus homogeneous Lorentz transformation. In conclusion, though the
matrices D () are not unitary, they satisfy the pseudounitarity condition

D () = D ()1 (7.51)

moreover, it is easy to check that 5 is hermitian and anticommutes with

5 = 5 , {5 , } = 0 (7.52)

from which we obtain


5 = 5 (7.53)
combining equations (7.33, 7.49, 7.53) we also obtain
  
(5 ) = 5 = 5 = ( ) (5 ) = 5 = 5

so we have
(5 ) = 5 (7.54)
The Dirac and related matrices obey some symmetry properties. From Eqs. (7.38) and (7.39) and the fact that
2 is antisymmetric while 1 and 3 are symmetric, we see that is symmetric for = 0, 2 and antisymmetric
for = 1, 3.

     
0 022 122 0 2 022 e 2 022 2

e i = ; e i = i = 2
122 022
e2 022 2 022
   
k 022 e
k 022 k

e = i = i = k ; k = 1, 3

ek 022 k 022
0 =
e0 ; 2 =
e2 ; 1 = e
1 ; 2 = e
2 (7.55)
7.3. THE CHIRAL REPRESENTATION FOR THE DIRAC MATRICES 189

Note that the property that is independent of the basis is the hermiticity of the Pauli matrices. In the particular
representation described by equation (7.39), 2 is antisymmetric because its non-null elements are purely imagi-
nary, while 1 and 3 are symmetric because they are real. Consequently, equation (7.55) can be guaranteed only
for the representation given by Eqs. (7.38) and (7.39). This can be summarized as
 
1 2 2 0
e = C C
; C = i (7.56)
0 2

to show that, it is convenient to express Eqs. (7.38) in a more condensed notation as


   
0
;
k
, 1 = (k , 1) ; no sum over (7.57)
g 0

from the properties (7.40) of the Pauli matrices and the definition (7.57) of , we obtain
 2
k = k2 = 1 ; ( )2 = 1 (7.58)

from which we can easily check that


C1 = C (7.59)
using the definition of C Eq. (7.56) as well as Eqs. (7.57, 7.59), we have
        
1 2 0 0 2 0
P C C = C C = i i i
0 2 g 0 0 2
  
2 0 0 2
= i
0 2 g 2 0

  
0 2 2 0 Z
P = i i (7.60)
g 2 2 0 g Z 0

now we can evaluate the products of Pauli matrices by using properties (7.40) and (7.58). If = 0, 2; we find

Z 0 = 2 0 2 = 2 2 = 1 ; Z 2 = 2 2 2 = 2 2 2 = 2

and if = k = 1, 3; we obtain

2 k 2 = 2 k 2 = k 2 2 = k ; k = 1, 3

so that we have
Z 0 = 1 , Z 2 = 2 , Z 1 = 1 , Z 3 = 3 (7.61)
substituting Eqs. (7.61) in (7.60) for = 0, 1, 2, 3; and taking into account Eqs. (7.55), we obtain P = e
showing the validity of Eq. (7.56).
Moreover, from Eq. (7.56) we can obtain the transposes of the matrices in the basis (7.37). For instance
i i    
Je = [e e e
e
] = C C1 C C1 C C1 C C1
4 4
i  1  i
= C C C C1 = C [ ] C1 = CJ C1 = CJ C1
4 4
therefore, from Eq. (7.56) we derive the following results

Je = CJ C1 (7.62)
1
5 = +C5 C
e (7.63)
^
( 1
5 ) = +C5 C (7.64)
190 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

the signs in Eqs. (7.62, 7.63, 7.64) will be important when we consider the charge-conjugation properties of
currents formed with these matrices. The results obtained for adjoints and transposes can be combined to obtain
the properties of complex conjugate of the Dirac and related matrices. For example, by combining Eqs. (7.49,
7.56) and taking into account that is real in the chiral representation [see Eq. (7.48)] we obtain
h i
C C1 = e ) = =
= (e

and with an analogous procedure we obtain the conjugate of the matrices in the basis (7.37)

= C C1 (7.65)
1
J = CJ C (7.66)
5 = C5 C 1
(7.67)
1
(5 ) = C5 C (7.68)

for future purposes we also observe that


 
{C, } = C + C 2 + 2 = 2 2 2 = 2 2 = 0
   
[C, 5 ] = C5 5 C i 2 0 5 5 i 2 0 = i 2 0 5 i 2 0 5 = 0

where we have used the fact that 5 anticommutes with . We then obtain

{C, } = [C, 5 ] = 0 (7.69)

7.4 Causal Dirac fields


We shall assume since the beginning that the particle does not coincide with the associated antiparticle, and
assume that we run over a single species of particle (so we have two species n and n corresponding to the particle
and its antiparticle respectively). As customary, we shall construct particle annihilation and antiparticle creation
fields, that in this case transform under the Lorentz group according with the Dirac (or spinor) representation.
By using Eqs. (4.39, 4.40) we write
XZ
k+ (x) = (2)3/2 d3 p uk (p, ) eipx a (p, ) (7.70)

3/2
XZ
kc (x) = (2) d3 p vk (p, ) eipx ac (p, ) (7.71)

where the particle species has been omitted7 . We calculate the coefficient functions u (p, ) and v (p, ) as usual:
we start with Eqs. (4.49, 4.50) to find uk and vk for zero momentum, then we apply Eqs. (4.45, 4.46) to find them
for arbitrary momenta, we should use Dkk () in this case as the 4 4 Dirac representation of the homogeneous
Lorentz group constructed from the generators J of Eq. (7.3).
The zero momentum conditions (4.49) become
X (j)
X
uk (0, ) J = Jkk uk (0, )
k
X (j)

u (0, ) J = J u (0, ) = J 0 u0 (0, ) + J i ui (0, )

7
It worths pointing out that the value of the coefficients uk (p, ) and vk (p, ) only depends on the irreducible representation of
the Lorentz group to which the fields are associated. They do not depend on the specific species of particles. Therefore, since particles
and antiparticles transform under the same irreducible representation, their coefficients are the same. Thus the only difference between
kc (x) and k (x) is the creation operators a and ac as can be seen in Eq. (7.71).
7.4. CAUSAL DIRAC FIELDS 191

and using the explicit form of the cartesian generators of SO (3), Eqs. (6.9, 6.10), page 163 we have
X (j)
u (0, ) (Jk ) = (Jk ) i ui (0, )

for = 0, n respectively we have


X (j)
X (j)
u0 (0, ) (Jk ) = (Jk )0 i ui (0, ) ; un (0, ) (Jk ) = (Jk )n i ui (0, )

X (j)
X (j)
0
u (0, ) (Jk ) = 0 ; u (0, ) (Jk ) = inik ui (0, )
n


???
-
-

Something similar can be done for the zero momentum condition (4.50). On the other hand, the four dimen-
sional representation we are dealing with, is reducible. Thus, it is convenient to replace the four component index
by a couple of indices: one 2-valued index m that labels the rows and columns of the submatrices in Eqs. (7.44)
and a second index that takes the values , and that labels the rows and columns of the supermatrix in Eqs.
(7.44). With this notation, we finally obtain [homework!!(16)].
X (j)
X1
um (0, ) J = mm um (0, ) (7.72)
m
2
X (j)
X1
vm (0, ) J = mm um (0, ) (7.73)
m
2

we can rewrite these equations by regarding um (0, ) and vm (0, ) as the m, elements of matrices U and
V , that is
(U+ )m um+ (0, ) , (U )m um (0, )
(V+ )m vm+ (0, ) , (V )m vm (0, ) (7.74)
By now, such matrices (of dimension m) are rectangular but not neccesarily square, because m takes two values
but we do not know yet how many values takes the quantum number [the number of values of is 2j + 1 but
we have not determined j so far]. In the matrix notation described by (7.74), equations (7.72, 7.73) are written as
X (j)
X1
(U )m J = mm (U )m (0, )
m
2
X (j)
X1
(V )m J = mm (U )m (0, )
m
2

which can be finally written as


1
U J(j) = U (7.75)
2
1
V J(j) = V (7.76)
2
now, the (2j + 1) dimensional matrices J(j) and J(j) provide irreducible representations of the Lie algebra of
rotation8 , the same occurs for the 2 2 matrices /2. Consequently, according with Schurs Lemma 1 [see Lemma
8
If we have a given matrix representation D (G) of a group, it is straightforward to see that the conjugate matrices D (G) also give
a representation on the same vector space. The natural question is whether the conjugate representation is equivalent to the original
representation or not. In the case of SO (3), all conjugate representations are equivalent to the original ones.
192 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

1, page 14] the matrices U and V either vanish (which is not a case in which we are interested) or is square
and non-singular. Therefore, m takes the same number of values as (since U and V are matrices of dimension
m ). Consequently, since m takes two values we have that 2j + 1 = 2, and the Dirac field can only describe
particles of spin j = 1/2. Each matrix U and V in Eqs. (7.75, 7.76) is a 22 square matrix. Further the matrices
J(1/2) , J(1/2) and /2 must describe the same irreducible representation because there is only one irreducible
inequivalent representation of SO (3) in a given vector space. In other words, the matrices J(1/2) and J(1/2) must
be the same as /2 up to a similarity transformation. Indeed, in the standard (or canonical) representation Eqs.
(1.165-1.168) of the rotation generators, we have J(1/2) = /2. From which we have

1 1 1 1 (1/2)
2 k 2 = 2 2 k = k = k = Jk ; k = 1, 3
2 2 2 2
1 1 1 (1/2)
2 2 2 = 2 = 2 = J2 ; k=2
2 2 2

Thus in the canonical representation described by Eqs. (1.165-1.168), we have

1
J(1/2) = and J(1/2) = 2 2 . (7.77)
2 2

combining Eqs. (7.75, 7.76) with Eq. (7.77), it follows that

1 1 1
U J(1/2) = U U = U [U , ] = 0
2 2 2
1 1 1 1 1
V J(1/2) = V V 2 2 = V V 2 22 = V 2
2 2 2 2 2
1 1
V 2 = V 2 [V 2 , ] = 0
2 2
therefore
[U , ] = [V 2 , ] = 0 (7.78)

so that according with Schurs lemma 2 [see 2, page 14] U and V 2 must be proportional to the unit matrix9 ,
hence

U = c I22 ; V 2 = id I22 V = id 2
(U ) = c m ; (V )m = id (2 )

and recalling definitions (7.74) we find

um, (0, ) = c m ; vm, (0, ) = id (2 )m (7.79)

as a matrix index it is convenient to redefine 1/2, 1/2 1, 2. From which the first of Eqs. (7.79) can be
written explicitly as
 
u1,+ 0, 12  c+ 1,1 c+ u1,+ 0, 12  c+ 1,2 0
u2,+ 0, 12  c+ 2,1 0 u2,+ 0, 12  c+ 2,2 c+
= ; =
u1, 0, 12  c 1,1 = c u1, 0, 12  c 1,2 = 0
u2, 0, 12 c 2,1 0 u2, 0, 12 c 2,2 c
9
Of course, the proportionality with the unit matrix is for each two-component matrix U+ and U ,. because the four component
matrix is reducible and the Schurs lemmas are only valid for irreducible representations. Owing to it, we have two different coefficients
for the + and parts in Eq. (7.79). Similar argument follows for V .
7.5. DIRAC COEFFICIENTS AND PARITY CONSERVATION 193

By using the explicit form of 2 in Eq. (7.39), the second of Eqs. (7.79) yields

v1,+ 0, + 12  id+ (2 )1,1 0 0
v2,+ 0, + 1 id+ (2 )2,1 id+ i d+
2 =
v1, 0, + 1 = id (2 ) = 0 0
2 1,1
1
v2, 0, + 2 id (2 )2,1 id i d

v1,+ 0, 12  id+ (2 )1,2 id+ (i) d+
v2,+ 0, 1 id+ (2 )2,2 0 0
2
v1, 0, 1 = id (2 ) = id (i) = d
2 1,2
v2, 0, 12 id (2 )2,2 0 0

thus the coefficients at zero momentum yield



  c+   0
1 0 1 c+
u 0, =
c ; u 0, = 0
(7.80)
2 2
0 c

  0   d+
1 d+ 1 0
v 0, =
0
; v 0, = d

(7.81)
2 2
d 0
from Eqs. (4.45, 4.46), the spinors at arbitrary momentum read
r
m
u (p, ) = D (L (p)) u (0, ) (7.82)
p0
r
m
v (p, ) = D (L (p)) v (0, ) (7.83)
p0

7.5 Dirac coefficients and parity conservation


In general the constants c and d in Eqs. (7.80, 7.81) are rather arbitrary. We could for instance choose c+ and
d+ (or d and c ) to be zero, such that the Dirac field would have only two non-vanishing components. The only
way to say something else about the relative values of c or the d on physical grounds, is by considering the
parity conservation scenario. Equations (5.33, 5.34) page 158, tell us how the particle annihilation and antiparticle
creation operators transform under space inversion
P a (p, ) P 1 = a (p, ) (7.84)
c 1 c c
P a (p, ) P = a (p, ) (7.85)
substituting (7.84, 7.85) in the field expansions (7.70) we have
XZ
+ 1
P k (x) P = q d3 p uk (p, ) eip(Px) a (p, ) (7.86)
3
(2)
c XZ
P kc (x) P 1 = q d3 p vk (p, ) eip(Px) ac (p, ) (7.87)
3
(2)

then we have to characterize the coefficients uk (p, ) and vk (p, ). We can do it by observing from Eqs. (6.27)
page 165 that
Li k (p) = Li k (p) ; Li 0 (p) = L0 i (p) = L0 i (p) ; L0 0 (p) = L0 0 (p)
194 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

from this and Eq. (7.18), we find


D (L (p)) = D (L (p)) (7.88)
combining Eqs. (7.82, 7.88) we have
r r
m m
u (p, ) = 0
D (L (p)) u (0, ) = D (L (p)) u (0, )
p p0

a similar procedure can be done for v (p, ) from Eq. (7.83). We then obtain
r
m
u (p, ) = D (L (p)) u (0, ) (7.89)
p0
r
m
v (p, ) = D (L (p)) v (0, ) (7.90)
p0

substituting (7.89, 7.90) in (7.86, 7.87) we find


r Z
m X
P k+ (x) P 1 = q d3 p [D (L (p)) u (0, )] eip(Px) a (p, ) (7.91)
3 p0
(2)
r Z
1 c m X
c
P k (x) P = q d3 p [D (L (p)) v (0, )] eip(Px) ac (p, ) (7.92)
3 p0
(2)

once again, in order to preserve causality, the parity operator should transform the annihilation and creation
fields at the point x into something proportional to these fields evaluated at Px. To satisfy that condition, it is
necessary that u (0, ) and v (0, ) be proportional to u (0, ) and v (0, ) respectively, then

u (0, ) = bu u (0, ) ; v (0, ) = bv v (0, ) (7.93)

where Eq. (7.93) says that u (0, ) is eigenvector of with eigenvalue bu , and v (0, ) is eigenvector of with
eigenvalue b . Since 2 = 1, its eigenvalues are 1, hence bu and bv are sign factors b2u = b2v = 1. In this case, by
substituting (7.93) in (7.91, 7.92) we have
r Z
+ 1 1 m X
P k (x) P = bu q d3 p [D (L (p)) u (0, )] eip(Px) a (p, )
3 p0
(2)
r Z
1 1 m X
c
P k (x) P c
= bv q d3 p [D (L (p)) v (0, )] eip(Px) ac (p, )
3 p0
(2)

the first of this equation can be written as


XZ r 
1 m
k+P (x) P k+ (x) P 1 = bu q d3 p D (L (p)) u (0, ) eip(Px) a (p, )
3 p0
(2)

1 XZ
= bu q d3 p u (p, ) eip(Px) a (p, ) = bu k+ (Px)
3
(2)

where we have used Eqs. (7.70, 7.82). A similar procedure can be done for P kc (x) P 1 . Consequently, the fields
have the following properties under space inversion

P + (x) P 1 = bu + (Px) (7.94)


c 1 c c
P (x) P = bv (Px) (7.95)
7.5. DIRAC COEFFICIENTS AND PARITY CONSERVATION 195

now we shall obtain information about the coefficients c and d in Eqs. (7.80, 7.81). Let us do the exercise
explicitly for u (0, 1/2). Equation (7.93) says that u (0, 1/2) must be an eigenvector of i 0 with eigenvalue bu
   
0 1 1
i u 0, = bu u 0,
2 2
By using Eqs. (7.38, 7.80), we obtain the explicit form of this equation

0 0 1 0 c+ c+
0 0 0 1 0 0
= bu (7.96)
1 0 0 0 c c
0 1 0 0 0 0

as in any
equation for eigenvectors, we can define one of the components of the eigenvector arbitrarily. Let us set
c+ 1/ 2. With this assignment, Eq. (7.96) becomes

0 0 1 0 1/ 2 1/ 2
0 0 0 1 0 0

1 0 0 0 c = bu c
0 1 0 0 0 0
bu
c
2
0 0
1 =
b c
2 u
0 0

which leads to two equations


bu 1
c = ; bu c = (7.97)
2 2
substituting the first of theseequations into the second we obtain the already known
 condition b2u = 1. Hence
using the convention c+ = 1/ 2 and the first of equations (7.97) the vector u 0, 12 yields

  1/ 2
1 0
u 0, =


2 bu / 2
0

the same procedure can be carried out for the other zero momentum vectors u (0, 1/2) and v (0, 1/2).
In summary, we can adjust the overall scale of the fields for the coefficient functions at zero momentum to
have the specific form

  1   0
1 1 0 1 1 1
u 0, =

, u 0, = (7.98)
2 2 bu 2 2 0
0 bu

  0   1
1 1 1 1 1 0

v 0, = , v 0, = (7.99)
2 2 0 2 2 bv
bv 0

in order to build up a causal field we make a linear combination of the annihilation and creation fields

(x) = + (x) + c (x) (7.100)


196 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

that commutes or anticommutes with itself and with its adjoint at space-like separations.
h To fix the
i values of

and , we shall impose causality to the commutation or anticommutation relation k (x) , k (y) . First from

expansions (7.70, 7.71) the fields (x) and (y) become

(x) = + (x) + c (x)


3/2
XZ h i
k (x) = (2) d3 p uk (p, ) eipx a (p, ) + vk (p, ) eipx ac (p, )

XZ h    i
k (y) = (2) 3/2
d3 p uk p , eip y a p , + vk p , eip y ac p ,

such that its commutation or anticommutation relations give.

h i Z XZ
1 X 
I k (x) , k (y) = 3 d 3
p d3 p uk (p, ) eipx a (p, )
(2)
   i
+vk (p, ) eipx ac (p, ) , uk p , eip y a p , + vk p , eip y ac p ,

X Z X Z  h i
1  
I = 3 d3 p d3 p ||2 uk (p, ) uk p , eipx eip y a (p, ) , a p ,
(2)

h i 
 
+ ||2 vk (p, ) vk p , eipx eip y ac (p, ) , ac p ,

X Z X Z n
1  
I = 3 d3 p d3 p ||2 uk (p, ) uk p , eipx eip y p p
(2)

h  io
||2 vk (p, ) vk p , eipx eip y ac p , , ac (p, )

Z n o
1 X 3 2 ipx ipy 2 ipx ipy
I = d p || uk (p, ) u (p, ) e e || v k (p, ) v (p, ) e e
(2)3 k k

Z ( " # " # )
1 2
X 2
X
3 ip(xy) ip(xy)
I = d p || uk (p, ) uk (p, ) e || vk (p, ) vk (p, ) e
(2)3

h i
In summary, the commutation or anticommutation relation k (x) , k (y) gives

h i Z h i
1
k (x) , k (y) = d 3
p || 2
N k k (p) eip(xy)
||2
Mk k (p) eip(xy)
(7.101)
(2)3
X
Nkk (p) uk (p, ) uk (p, ) (7.102)

X
Mkk (p) vk (p, ) vk (p, ) (7.103)

In order to find N (p) we obtain first N (0) and use an apropriate boost transformation to find N (p), and similarly
7.5. DIRAC COEFFICIENTS AND PARITY CONSERVATION 197

for M (p). For example, the matrix Nkk (0) for p = 0, can be evaluated explicitly from Eqs. (7.98)
X
Nkk (0) uk (0, ) uk (0, )

       
1 1 1 1 1 1 1
N11 = u1 0, u1 0, + u1 0, u1 0, = +00=
2 2 2 2 2 2 2
       
1 1 1 1 1 1
N12 = u1 0, u2 0, + u1 0, u2 0, = 0+0 =0
2 2 2 2 2 2
       
1 1 1 1 1 b b
N13 = u1 0,
u3 0, + u1 0,
u3 0, = u + 0 0 = u
2 2 2 2 2 2 2
       
1 1 1 1 1 bu
N10 = u1 0, u0 0, + u1 0, u0 0, = 0+0 =0
2 2 2 2 2 2

proceeding the same for the other components we obtain explicitly


1 1
2 0 2 bu 0
0 1
0 1
N (0) = 2 2 bu
bu 0
1 1
0
2 2
1 1
0 bu 0
2
2

1 0
0 0 0 0 1 0
1 0 1 0 0
+ bu 0 0 0 1

= 1 [1 + bu ]
N (0) = 0 0
2 1 0 1 0 0 0
2

0 0 0 1 0 1 0 0

we can do the same with the vectors vk (0, ) to obtain M (0). In conclusion, by using either the eigenvalue
conditions (7.93) or the expressions (7.98, 7.99) we find for the coefficients Nkk and Mkk at zero momentum that

1 + bu 1 + bv
N (0) = ; M (0) = (7.104)
2 2
Now, to obtain N (p) at arbitrary momentum, we apply (7.82), so we have
" # "r #
X X rm X m X
Nkk (p) uk (p, ) uk (p, ) = Dkm (L (p)) um (0, ) D (L (p)) un (0, )

p0 m p0 n kn
mX X X
= 0
Dkm (L (p)) Dkn (L (p)) um (0, ) un (0, )
p m n
mX X m XX e (L (p))
= D km (L (p)) Dkn (L (p)) N mn (0) = Dkm (L (p)) Nmn (0) Dnk
p0 m n
p 0
m n
m XX e (L (p))
= Dkm (L (p)) [1 + bu ]mn D nk
2p0 m n
m XX
Nkk (p) = 0
Dkm (L (p)) [1 + bu ]mn Dn k (L (p))
2p m n

and similarly for M (p). Therefore, from Eqs. (7.82, 7.83) we obtain those coefficients at arbitrary momentum
m
N (p) = D (L (p)) [1 + bu ] D (L (p)) (7.105)
2p0
m
M (p) = D (L (p)) [1 + bv ] D (L (p)) (7.106)
2p0
198 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

The pseudounitarity condition (7.51) applied to = L (p) gives

D (L (p)) = D (L (p))1 (7.107)

that can be written as

D (L (p)) D (L (p)) = 1
D (L (p)) D (L (p)) 2 =

since 2 = 1 we get
D (L (p)) D (L (p)) =
alternatively Eq. (7.107) also yields

D (L (p)) = D (L (p))1 2 D (L (p)) 2 = D (L (p))1


D (L (p)) = D (L (p))1 D (L (p)) D (L (p)) = D (L (p)) D (L (p))1

therefore the pseudounitarity condition can be written in either of these ways

D (L (p)) D (L (p)) = (7.108)


1
D (L (p)) D (L (p)) = D (L (p)) D (L (p)) (7.109)

and using the fact that = i 0 and the four-vector character of Eq. (7.9) we find
 0
D (L (p)) D 1 (L (p)) = iD (L (p)) 0 D 1 (L (p)) = iL 0 (p) = i L1 (p)
= i [L (p)]0 = i [L (p)]0 0 0 + i [L (p)]0 i i
s s
0 0 2 0 2 2
p (p ) p0 (p ) m i
= i 0 + i b pi 2
1 i = i 0 i pbi
m m m m2
 
p0 0 pbi
= i i |p| i
m m

where we have used Eqs. (1.49, 6.27), we finally obtain

p
D (L (p)) D 1 (L (p)) = iL 0 (p) = i (7.110)
m
substituting (7.108), (7.109) and (7.110) in Eq. (7.105), the factor N (p) becomes
m m m
N (p) = 0
D (L (p)) [1 + bu ] D (L (p)) = 0 D (L (p)) D (L (p)) + 0 bu D (L (p)) D (L (p))
2p 2p 2p
p
m m m
= D (L (p)) D 1 (L (p)) + 0 bu = i 0 + 0 bu
2p0 2p 2p 2p

something similar can be done for M (p) in Eq. (7.106) we then obtain

1
N (p) = [ip + bu m] (7.111)
2p0
1
M (p) = [ip + bv m] (7.112)
2p0
7.5. DIRAC COEFFICIENTS AND PARITY CONSERVATION 199

substituting (7.111, 7.112) in Eq. (7.101) the commutator or anticommutator of the fields become
h i Z 
1 1
I k (x) , k (y) = 3 d 3
p ||2 0 [i p + bu m]km mk eip(xy)
(2) 2p

2 1 ip(xy)
|| [i p + bv m]km mk e
2p0
Z  
1 3 2 1 ip(xy) 2 1 ip(xy)
= d p || [ + bu m] e || [ + bv m] e
(2)3 2p0 2p0 k k
 Z 3   Z 3 
2 1 d p ip(xy) 2 1 d p ip(yx)
= || [ + bu m] e || [ + bv m] e
(2)3 2p0 k k (2)3 2p0

so we obtain
h i n o
k (x) , k (y) = ||2 [ + bu m] + (x y) ||2 [ + bv m] + (y x) (7.113)
k k

where + is the function defined in Eq. (5.9), page 152


Z
1 d3 p ipx
+ (x) e (7.114)
(2)3 2p0

we saw in section 5 that + (x y) is an even function of x y for space-like separations between x and y. It
implies that its first derivatives are odd functions of x y. Then the commutation or anticommutation relations
give
h i n o
k (x) , k (y) = ||2 [ + bu m] + (x y) ||2 [ + bv m] + (x y)
h i n o h i k k
2 2 2 2
(x) , (y) = || || + (x y) + || bu || bv m + (x y)

Therefore, in order that both the derivative and non-derivative terms in the commutator or anticommutator vanish
at space-like separations, it is necessary and sufficient that

||2 = ||2 (7.115)

and
||2 bu = ||2 bv (7.116)
it is clear that (7.115) discards the possibility of a minus sign which corresponds to the case of commutators.
Consequently, the particles described by Dirac fields must be fermions10 . Combining Eqs. (7.115, 7.116)
we see that it is also necessary that ||2 = ||2 and bu = bv . As in the case of scalars we can redefine the phases
such that / be real positive, in that case we have = , and absorbing the overall phase of the field we
obtain finally
==1
In addition, it is possible to replace by 5 , which changes the sign of both bu and bv , then we can choose

bu = bv = +1 (7.117)

from Eqs. (7.70, 7.71) along with (7.100), the Dirac field becomes
1 XZ h i
k (x) = q d3 p uk (p, ) eipx a (p, ) + vk (p, ) eipx ac (p, ) (7.118)
(2)3
10
Further, we already saw that j = 1/2. So we have predicted once more that fermions are half-odd integer spin particles.
200 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

where the coefficients at zero momentum can be obtained by combining (7.98, 7.99, 7.117) and they are given by

  1   0
1 1 0 1 1 1
u 0, = , u 0, = (7.119)
2 2 1 2 2 0
0 1

  0   1
1 1 1 1 1 0
v 0, = , v 0, = (7.120)
2 2 0 2 2 1
1 0
the spin sums are
1
N (p) = [ip + m] (7.121)
2p0
1
M (p) = [ip m] (7.122)
2p0
so that the anticommutator in Eq. (7.101), yields
h i
k (x) , k (y) = {[ + m] }kk (x y) (7.123)
+

where (x y) is defined in Eq. (5.32) page 158. We now come back to the requirement that for a parity
conserving theory, under space inversion the field (x) must transform into something proportional to (Px).
For this to be possible the phases in Eqs. (7.94, 7.95) must be equal. Therefore, the intrinsic parities of particles
and their antiparticles are related by
c =
then we obtain that the intrinsic parity c = ||2 of a state consisting of a spin 1/2 particle and its antiparticle
is odd. Now, equations (7.94, 7.95) provide the transformation of the field (x) under space inversion

P (x) P 1 = (Px)

Applying u (p, ) on both sides of Eq. (7.110) and using Eq. (7.82) we find
p
D (L (p)) D 1 (L (p)) u (p, ) = i u (p, )
r m
m p
D (L (p)) u (0, ) = i u (p, )
p0 m
and using Eqs. (7.93, 7.117) and (7.82) we obtain
r
m p
bu D (L (p)) u (0, ) = i u (p, )
p0 m
p
u (p, ) = i u (p, )
m
a similar procedure can be done for v (p, ), and we get
p
v (p, ) = i v (p, )
m
therefore, we see that u (p, ) and v (p, ) are eigenvectors of ip /m with eigenvalues +1 and 1 respectively.
We can rewrite these equations as

(ip + m) u (p, ) = 0 , (ip + m) v (p, ) = 0 (7.124)


7.6. CHARGE-CONJUGATION PROPERTIES OF DIRAC FIELDS 201

applying the operator ( + m) on the field (x) of Eq. (7.118), and using Eqs. (7.124), we obtain

3/2
XZ n    o

( + m) (x) = (2) d3 p ( + m) eipx u (p, ) a (p, ) + ( + m) eipx v (p, ) ac (p, )

3/2
XZ n
= (2) d3 p eipx [(i p + m) u (p, )] a (p, ) + eipx [(i p + m) v (p, )] ac (p, )

= 0

Therefore, from Eqs. (7.124) we find that the field (7.118) satisfies the following differential equation

( + m) (x) = 0 (7.125)

which is the so-called Dirac equation for a free particle of spin 1/2. In this approach the free-particle Dirac
equation is the Lorentz invariant way in which we have put together the two irreducible representations of the
proper orthochronus Lorentz group in order to form a field with a simple transformation rule under space inversion.

7.6 Charge-conjugation properties of Dirac fields


In order to characterize the charge conjugation and time-reversal properties of the Dirac field, we require expres-
sions for the complex conjugates of u and v. According with Eqs. (7.119, 7.120) they are real for zero momentum.
To obtain them at arbitrary momentum we should multiply these coefficients with the complex matrix D (L (p)).
Hence we require first an expression for D (L (p)). To obtain it, we start with Eq. (7.66) for a general real value
of , to get
   
1 1 1
Z = iJ = i J
= i KJ K1 = KZK1 ; K C
2 2 2
1
 1
 2 1
Z 2 = KZK KZK = KZ K Z n = KZ n K1

hence the same transformation (7.66) can be applied to a power series of iJ /2 as long as it is convergent.
In particular, we have
     
1 1
exp iJ = C exp iJ C1 (7.126)
2 2
the LHS of this equation gives an arbitrary element D () of the spinor representation of the proper orthochronus
homogeneous Lorentz group. Therefore, we have in particular

D (L (p)) = CD (L (p)) C1 = (C) D (L (p)) (C)1 (7.127)

then for an arbitrary p, the coefficient u (p, ) yields


r  r
m m
u (p, ) = 0
D (L (p)) u (0, ) = D (L (p)) u (0, )
p p0
r
m
u (p, ) = CD (L (p)) C1 u (0, ) (7.128)
p0

On the other hand, by using Eqs. (7.93, 7.117) as well as Eqs. (7.56, 7.59), we have
 
1 1 2 0
C u (0, ) = bu C u (0, ) = Cu (0, ) = i u (0, )
0 2
202 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

evaluating at = 1/2, we obtain



  0 i 0 0 1 0  
1 i i 0 0
0 0 1 1 1
C1 u 0, = = = v 0,
2 2 0 0 0 i 1 2 0 2
0 0 i 0 0 1
and similarly for u (0, 1/2) and also for v (0, 1/2). Then we have the properties
C1 u (0, ) = v (0, ) ; C1 v (0, ) = u (0, ) (7.129)
substituting (7.129) in (7.128) we obtain
r
m
u (p, ) = CD (L (p)) v (0, ) = Cv (p, )
p0
and we can do the same for v (p, m). We finally obtain
u (p, ) = Cv (p, ) ; v (p, ) = Cu (p, ) (7.130)
therefore, the adjoint of the field (7.118), gives
1 XZ h i

(x) = q d3 p uT (p, ) eipx a (p, ) + v T (p, ) eipx ac (p, )
(2)3
1 XZ h i
e (x) = q d3 p [Cv (p, )]T eipx a (p, ) [Cu (p, )]T eipx ac (p, )
(2)3
1 XZ h i

(x) = q d3 p Cv (p, ) eipx a (p, ) Cu (p, ) eipx ac (p, )
(2)3
1 XZ h i
(x) = C q d3 p v (p, ) eipx a (p, ) + u (p, ) eipx ac (p, ) (7.131)
(2)3
As always, for the field to transform under charge-conjugation into another field C (x) with which it commutes
at space-like separations, we require that the charge-conjugation parities of the particle and antiparticle be related
by
c = (7.132)
Now,using Eqs. (3.27, 7.132), the charge-conjugation of the field (7.118) becomes
1 XZ h i
1
C (x) C (x) C = q d3 p u (p, ) eipx Ca (p, ) C 1 + v (p, ) eipx Cac (p, ) C 1
(2)3
1 XZ h i
C (x) = q d3 p u (p, ) eipx ac (p, ) + v (p, ) eipx c a (p, )
(2)3
XZ h i
C (x) = q d3 p u (p, ) eipx ac (p, ) + v (p, ) eipx a (p, ) (7.133)
(2)3
and comparing Eqs. (7.131, 7.133) we have
Z h i
X
(x) = C q d3 p v (p, ) eipx a (p, ) + u (p, ) eipx ac (p, )
(2)3
(x) = CC (x) (x) = 2 CC (x) (x) = CC (x) = C1 C (x)
C (x) = C (x)
7.6. CHARGE-CONJUGATION PROPERTIES OF DIRAC FIELDS 203

and from Eq. (7.69) page 190, we obtain finally

C (x) C 1 = C (x) = C (x) (7.134)


where we write (x) instead of (x) on the RHS of this equation to emphasize that it is a column vector, not
a row.
For a system consisting of a particle and its antiparticle, there is an important difference between fermions
and bosons concerning the intrinsic charge-conjugation phase. Such a state can be written by applying a creation
operator for the particle and a creation operator for the antiparticle on the vacuum state |0i as follows
XZ Z
 
|i = d p d3 p p, , p , a (p, ) ac p , |0i
3
(7.135)
,

Now, we shall assume that the vacuum is invariant under charge conjugation

C |0i = |0i

Therefore, under charge conjugation this state transforms as


XZ Z
 
C |i = d p d3 p p, , p , C a (p, ) ac p , |0i
3

,
XZ Z
 h ih  i
= d3 p d3 p p, , p , C a (p, ) C 1 C ac p , C 1 C |0i
,
XZ Z
 
C |i = c 3
d p d3 p p, , p , ac (p, ) a p , |0i
,

interchanging the variables of integration and summation, and using Eq. (7.132) we have
XZ Z
 
3
C |i = d p d3 p p, , p , ac (p, ) a p , |0i
,
XZ Z
 
= d3 p d3 p p, , p , ac (p, ) a p , |0i
,

and utilizing the anticommutation relations for the creation operators we obtain
XZ Z
 
3
C |i = d p d3 p p, , p , a p , ac (p, ) |0i
,

and interchanging the dummy indices and p p


XZ Z
 
C |i = d p d3 p p , , p, a (p, ) ac p , |0i
3
(7.136)
,

Now, if the wave function of the state is even or odd under the interchange of the momenta and spins of the
particle and antiparticle, we have  
p , , p, = p, , p , (7.137)
in that case substituting (7.137) in (7.136) and comparing with (7.135) we obtain

C |i = |i
204 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

from which we conclude that the charge-conjugation parity of a state consisting of a particle described by a
Dirac field and its antiparticle is odd, in the sense that if the wave function of the state is even or odd under
the interchange of the momenta and spins of the particle and antiparticle, then the charge-conjugation operator
applied on such a state gives a sign (1) or (+1) respectively.
A classical example is the positronium, which is the bound state of an electron and a positron. The two
lowest states of positronium are a pair of nearly degenerate states with total spin S = 0 and S = 1, called
para-positronium and ortho-positronium respectively. The wave function of these two states is even under the
interchange of momenta and odd or even respecively under the interchange of spin three-components. Hence the
para- and ortho-positronium have C = +1 and (1) respectively. it leads to very different decay pattern of them.
The para-positronium decays rapidly into a pair of photons each of which has C = 1. The ortho-positronium
can only decay much more slowly into three or more photons.
Another example is given by the 0 and 0 mesons. They are produced as resonances coming from high-energy
electron-positron annihilation, with one photon as an intermediate state, showing that they must have C = 1,
which is consistent with the interpretation of these mesons as a pair of quark anti quark bound states with L = 0
and S = 1.

7.7 Time-reversal properties of Dirac fields


We start again from the transformation properties of the particle annihilation and antiparticle creation operators
give by Eqs. (3.29) page 122, but with j = 1/2
1
T a (p, ) T 1 = (1) 2 a (p, )
1
T ac (p, ) T 1 = c (1) 2 ac (p, ) (7.138)

and recalling the antilinearity of T , the field (7.118) transforms under time-reversal as

1 XZ h i
1
T k (x) T = q d3 p T uk (p, ) eipx a (p, ) + vk (p, ) eipx ac (p, ) T 1
(2)3

1 XZ h i
= q d3 p uk (p, ) eipx T a (p, ) T 1 + vk (p, ) eipx T ac (p, ) T 1
(2)3

and applying (7.138) we get

3/2
XZ 1
h i
T k (x) T 1
= (2) d3 p (1) 2 uk (p, ) eipx a (p, ) + c vk (p, ) eipx ac (p, )

by redefining the variables of summation and integration as p and we find


XZ 1
h
d3 p (1) 2 + uk (p, ) ei(p,p )(x,x ) a (p, ) + c vk (p, ) ei(p,p )(x,x ) ac
1 3/2 0 0 0 0
T k (x) T = (2)

XZ 1
h
d3 p (1) 2 + uk (p, ) ei(p,p )(x,x ) a (p, ) + c vk (p, ) ei(p,p )(x,x ) ac
3/2 0 0 0 0
= (2)

3/2
XZ 1
h
T k (x) T 1 = (2) d3 p (1) 2 + uk (p, ) eip(Px) a (p, )

i
+ c vk (p, ) eip(Px) ac (p, )
7.7. TIME-REVERSAL PROPERTIES OF DIRAC FIELDS 205

so that we need formulas for uk (p, ) and vk (p, ) in terms of uk (p, ) and vk (p, ). To do it, we use
Eqs. (7.82, 7.83)
r r
m m
u (p, ) = D (L (p)) u (0, ) ; v (p, ) = D (L (p)) v (0, ) (7.140)
p0 p0
r r
m m
u (p, ) = 0
D (L (p)) u (0, ) ; v (p, ) = D (L (p)) v (0, ) (7.141)
p p0

where we have recalled that u (0, ) and v (0, ) are real. Then we need expressions for D (L (p)) and also for
u (0, ) , v (0, ) in terms of u (0, ) , v (0, ).
We start by obtaining D (L (p)) in terms of D (L (p)). For this, we take into account that L (p) is a pure
boost, then

L1 (p) = L (p) D (L (p)) = D L1 (p) = D 1 (L (p))
so that D (L (p)) is the representation of a pure boost. Hence according with Eqs. (1.129, 7.3), in an infinitesimal
transformation of the type D (L (p)) only the generators Ki = J 0i appear in the expansion of D (L (p))
 
1 0i i0
  
D (L (p)) = 1 i 0i J + i0 J = 1 ii0 J i0 (7.142)
2

we shall also use equations (7.59, 7.69, 7.18, 7.35, 7.127)



C1 = C , {C, } = [C, 5 ] = , J i0 = [5 , J ] = 0
D (L (p)) = CD (L (p)) C1

from the previous properties we obtain


 
D (L (p)) = CD (L (p)) C1 = CD 1 (L (p)) C1 = C 1 ii0 J i0 C1
   
= C 1 ii0 J i0 C = C 1 ii0 J i0 C
   
= C 1 + ii0 J i0 C = C 1 + ii0 J i0 C
D (L (p)) = CD (L (p)) C1

alternatively, it can be written as


     
D (L (p)) = C 1 + ii0 J i0 C = C (5 )2 1 + ii0 J i0 C = C 5 1 + ii0 J i0 5 C
D (L (p)) = 5 CD (L (p)) C1 5

or equivalently as
 
D (L (p)) = 5 CD (L (p)) C1 5 = 5 CD (L (p)) C1 5 = 5 D (L (p)) 5

from which the matrix representation D (L (p)) can be written in several equivalent forms

D (L (p)) = CD (L (p)) C1 = 5 CD (L (p)) C1 5 = 5 D (L (p)) 5 (7.143)


The next step is to characterize u (0, ) , v (0, ) in terms of u (0, ) , v (0, ). For this we can combine
Eqs. (7.56, 7.59, 7.93, 7.117) to obtain

5 C1 u (0, ) = 5 Cu (0, ) = 5 2 u (0, ) = bu 5 2 u (0, )


5 C1 u (0, ) = 5 2 u (0, )
206 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

and using the explicit expressions (7.46, 7.38, 7.39, 7.119) we see that

   0 
1 1 5 2 1
5 C1 u 0, = 5 2 u 0, =
2 2 2 0
1

1 0 0 0 0 0 0 i 0 1
i 0 1 0 0 0 0 i 0 1
= 1 0
=
2 0 0 1 0 0 i 0 0 0 2 1
0 0 0 1 i 0 0 0 1 0

  1 0 0 0 0 0 0 i 1 0
1 i 0 1 0
0 0 0
i 0 0 1 1
5 C1 u 0, + = =
2 2 0 0 1 0 0 i 0 0 1 2 0
0 0 0 1 i 0 0 0 0 1

then we find    
1 1 1
5 C u 0, = u 0,
2 2
which can also be written as
1
5 C1 u (0, ) = (1) 2 u (0, )

proceeding similarly for v (0, ) we find


1
5 C1 u (0, ) = (1) 2 u (0, ) (7.144)
1
1
5 C v (0, ) = (1) 2 v (0, ) (7.145)

Now, substituting Eqs. (7.143, 7.144) in Eq. (7.141) we can obtain the coefficients u (p, ) , v (p, ) in
terms of u (p, ) , v (p, ) as required
r r
1
+ 1
+ m 1
+ m 
(1) 2 u (p, ) = (1) 2
0
D (L (p)) u (0, ) = (1) 2 0
5 CD (L (p)) C1 5 u (0, )
p p
r
1 m
= (1) 2 + 5 CD (L (p)) 5 C1 u (0, )
p0
r n o r
1 m 1
m
= (1) 2 + 5 CD (L (p)) (1) 2 u (0, ) = 5 C D (L (p)) u (0, )
p0 p0
= 5 C u (p, )

and similarly for v (p, ). From which we obtain the desired relations ???
1
(1) 2 + u (p, ) = 5 Cu (p, ) (7.146)
1
+
(1) 2 v (p, ) = 5 Cv (p, ) (7.147)

again, in order for time-reversal to take the Dirac field into a field proportional to itself evaluated at the time
reversed point (with which it would anticommute at space-like separations) it is necessary that the time-reversal
phases be related by
c = (7.148)
7.8. MAJORANA FERMIONS AND FIELDS 207

in that case, by substituting (7.146, 7.147) and (7.148) in (7.139) we find

3/2
XZ 1
h i
T k (x) T 1
= (2) d3 p (1) 2 + uk (p, ) eip(Px) a (p, ) + vk (p, ) eip(Px) ac (p, )

3/2
XZ h i
T k (x) T 1 = (2) d3 p 5 Cu (p, ) eip(Px) a (p, ) 5 Cv (p, ) eip(Px) ac (p, )

3/2
XZ h i
1
T k (x) T = (2) 5 C d3 p u (p, ) eip(Px) a (p, ) + v (p, ) eip(Px) ac (p, )

and comparing with (7.118) we have finally

T (x) T 1 = 5 C (Px) (7.149)

7.8 Majorana fermions and fields


We have been distinguishing particles and antiparticles. However, it not not ruled out the scenario in which they
are identical. Spin 1/2 particles that coincide with their antiparticles are called Majorana fermions. Using the
same reasoning that led to (7.134), the Dirac field of a Majorana particle must satisfy the reality condition

(x) = C (x) for Majorana fermions

In addition, the intrinsic space-inversion parity of a Majorana particle must be imaginary, = i, while charge
conjugation parity must be real = 1.

7.9 Scalar interaction densities from Dirac fields


The Lorentz transformation of the field (7.118), can be obtained from the Lorentz transformation of the creation
field Eq. (3.26) page 122, applied on a homogeneous Lorentz transformation
s
(p)0 X (j)

U0 () a (pn) U01 () = D (W (, p)) a (p n) (7.150)
p0

The field transforms as


U [] (x) U 1 [] = D 1 () (x) (7.151)
and its adjoint transforms as
  n o
U [] (x) U 1 [] = U 1 [] (x) U [] = U [] (x) U []
n o 
= U [] (x) U 1 [] = U [] (x) U 1 []
n o
= {U []} (x) {U []}1 ????

For future purposes, it is important to find out how to construct interaction densities out of Dirac fields and
their adjoints. Since the Dirac representation is not unitary, the bilinear form (x) (x) is not a scalar. To solve
that problem it is convenient to define another kind of adjoint

(x) (7.152)
208 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

from the pseudounitarity condition (7.51) we could show that the bilinear forms (x) M (x) have the following
transformation properties
 
U0 () (x) M (x) U01 () = (x) D () M D 1 () (x) (7.153)

and under space inversion


 
P (x) M (x) P 1 = (Px) M (Px) (7.154)

by comparing Eqs. (7.153, 7.154) we see that both transformations have a similar structure by making the
assignments U0 () P , P and D () , emphasizing once more the role of as a parity operation for
Dirac fields. Taking M as the basis matrices I, , J , 5 , 5 we obtain bilinears that transform as

(x) I (x) scalar



(x) (x) four-vector

(x) 5 (x) axial or pseudo four-vector
(x) 5 (x) pseudoscalar
(x) J (x) second rank tensor

we recall that the terms axial or pseudo mean that they have opposite properties of transformation under space-
inversion with respect to ordinary vectors and scalars. A pseudoscalar has negative parity. Further, the space and
time components of an axial vector have positive and negative parity respectively. These results apply also when
the two fermion fields in the bilinear refer to different particle species (and not particle and antiparticle), except
that in the latter case a space inversion also gives a ratio of the intrinsic parities. By the law of transformation of
these matrices we can find any bilinear form (x) M (x) because such matrices form a basis.
For future purposes, it is important to characterize the charge-conjugation properties of these bilinears. From
Eqs. (7.134, 7.59, 7.56) as well as Eqs. (7.62-7.64) we have
 
^
C (x) M (x) C 1 = (C)M ^ ) M
(C ) = (C f C = (x) C1 M
fC = (x) M (x)

where the sign in the last expression is + for the matrices I, 5 , 5 and () for the matrices and J . Note
that the first minus in this sequence comes from the Fermi statistics, and we ignore a cnumber anticommutator.
Consequently, if charge-conjugation is preserved, a boson field that interacts with the current must have C = +1
for scalars, pseudoscalars or axial vectors, and C = 1 for polar vectors or antisymmetric tensors. For example,
we can see in this way that the neutral pion 0 which couples with pseudoscalar or axial vector nucleon currents,
must have C = +1, while the photon that couples with polar vectors has C = 1.
The currents of the form (x) M (x) are very important in modelling the interactions. For example, the
original Fermi theory of beta decay contain a couple of polar vector currents (or an interaction density) of the
form
p n e

which was parity conserving. It was discovered later that the most general non-derivative Lorentz-invariant and
parity-conserving beta decay interactions has the form of a linear combination of products of two currents like
this but with replaced by any one of the five covariant types of 4 4 matrices I, 5 , J , 5 or 5 . We are
assuming that the space-inversion operator is defined so that the proton, electron and neutron all have intrinsic
parity +1. If we consider the neutrino as massless (which is only an approximation) its parity can also be defined
as +1, if necessary the neutrino field can be replaced by 5 . When it was realized that weak interactions are
non-parity conserved, the list of non-derivative interactions added ten terms proportional to

p M n e M and p M n e M 5 ; M I, , J , 5 , 5
7.10. THE CPT THEOREM 209

7.10 The CPT theorem


We have seen that the combination of quantum mechanics with special relativity leads to the existence of antipar-
ticles. It is necessary that each particle has an antiparticle, though of course it is possible in some cases that a
given particle be its own antiparticle. The CPT theorem relates properties of particles and antiparticles in the
following way

Theorem 7.1 (CPT theorem) for an appropiate choice of inversion phases under C, P and T , the product CP T
of all inversions is conserved.

To prove it, we begin by characterizing the action of CP T on scalar, vector and Dirac fields. They can be
obtained by composition of the inversions studied individually for each field [see sections 5.3, 6.5, and 7.5-7.7].
The results are

[CP T ] (x) [CP T ]1 = (x) (7.155)


1
[CP T ] (x) [CP T ] = (x) (7.156)
[CP T ] (x) [CP T ]1 = 5 (x) (7.157)

the phases , and depend on the species of particle described by each field. We can choose the phases so that
fo all particles we have
= 1 (7.158)
taking it into account, a tensor 1 2 n must transform as a superposition of fields of the form i as follows
n on o
[CP T ] {1 (x) 2 (x) . . . n (x)} [CP T ]1 = [CP T ] 1 (x) [CP T ]1 [CP T ] 2 (x) [CP T ]1 . . .
n o n on o n o
. . . [CP T ] n (x) [CP T ]1 = 1 (x) 2 (x) . . . n (x)

[CP T ] {1 (x) 2 (x) . . . n (x)} [CP T ]1 = (1)n 1 (x) 2 (x) . . . n (x)


therefore, any tensor 1 2 n formed from any set of scalar and vector fields and their derivatives transforms
into
[CP T ] 1 2 n (x) [CP T ]1 = (1)n 1 2 n (x) (7.159)
it is important to take into account that any complex number appearing in these tensor is transformed into its
complex conjugate since CP T is an antiunitary operator. It can be checked that the same rule is applied to
bilinear forms of Dirac fields. Using Eq. (7.157) such bilinears have the following rule of transformation
   
[CP T ] 1 (x) M 2 (x) [CP T ]1 = e1 (x) 5 M 5 2 (x) = 1 (x) 5 M 5 2 (x) (7.160)

note that the minus sign coming from the anticommutation of and 5 is cancelled by the anticommutation of
fermionic operators. If the bilinear is a tensor of rank n, we have that M is a product of n modulo 2 Dirac
matrices. Consequently
5 M 5 = (1)n M
so that the bilinear satisfies relation (7.159).
On the other hand, equation (4.16) page 139, says that a Hermitian scalar interaction density H (x) must
have the same number of creation fields and annihilation fields. Therefore, a Hermitian scalar interaction density
H (x) must be formed from tensors with an even total number of space-time indices (i.e. tensors of even rank), so
that
[CP T ] [H (x)] [CP T ]1 = H (x) (7.161)
210 CHAPTER 7. CAUSAL DIRAC FIELDS FOR MASSIVE PARTICLES

AB (x) belonging to one or


it is easy to see that the same is true for Hermitian scalar constructed from the fields ab
more of the general irreducible representations of the homogeneous Lorentz group.From the effects of inversions
on these fields we find  AB 
[CP T ] ab (x) [CP T ]1 = (1)2B abAB
(x)
For the Dirac field, the factor (1)2B is supplied by the matrix 5 in Eq. (7.157). Since the scalar interaction
density H (x) is constructed out of products of the form aA11bB1 1 (x) aA22bB2 2 (x) . . .. In order to couple the fields in
this way it is necessary that both A1 + A2 + . . .and B1 + B2 + . . .be integers. Therefore

(1)2B1 +2B2 +... = 1

so that a hermitian scalar automatically satisfies Eq. (7.161). From Eq. (7.161) it is straightforward that CP T
commutes with the interaction V give by Z
V d3 x H (x, 0)

hence
[CP T ] V [CP T ]1 = V
and in any theory CP T commutes with the free-hamiltonian H0 . Consequently, CP T commutes with the total
Hamiltonian H so that it is a constant of motion. Therefore, the operator CP T that has been defined here by its
action on free-particle operators, acts on in and out states in the way described in sections 2.3.2-2.3.6. The
physical consequences of CPT theorem has been discussed in sections 2.3.6 and 2.7.1.
Chapter 8

Massless particle fields

We have constructed fields associated with massive particles so far. For scalar and Dirac fields, it is not difficult
to construct zero mass states through the zero mass limit. However, we have already discussed some difficulties
in taking the zero mass limit for vector fields of spin one. It owes to do with the fact that at least one of the
polarization vectors blows up in this limit. More generally, it can be shown that the creation and annihilation
operators for physical masssless particles of spin j 1 cannot be used to construct all of the irreducible (A, B)
fields that can be constructued for finite mass. This particular limitation of field types lead us naturally to the
introduction of gauge invariance.
We shall construct a general free field for a massless particle by means of a linear combination of the annihilation
operators a (p, ) for particles of momentum p and helicity , and the associated creation operators ac (p, ) for
the antiparticles. We shall deal with only one species of particle so that we drop the label n. In addition, we shall
introduce at once the linear combination of creation and annihilation fields with coefficients , that we shall
adjust to preserve causality. Under all those considerations we find
Z Xh i
3/2
k (x) = (2) d3 p a (p, ) uk (p, ) eipx + ac (p, ) vk (p, ) eipx (8.1)

for massless particles we have p0= |p|. The rule of transformation of creation operators under homogeneous
Lorentz transformations is given generically for equations 3.26, page 122
s
(p)0 X (j)
U0 () a (p, ) U01 () = D (W (, p)) a (p ) (8.2)
p0
(j)
where D (W (, p)) is an irreducible representation of the little group characterized by the elements W (, p). We
recall that the little group associated with the massless particles is ISO (2) and its representations are determined
by Eqs. (1.238) page 51
D (W ) = exp (i) (8.3)
Combining equations (8.2, 8.3) we obtain the transformation rule of creation operators for massless particles under
homogeneous Lorentz transformations
s
(p)0
U () a (p, ) U 1 () = exp [i (p, )] a (p , ) (8.4)
p0
s
(p)0
U () ac (p, ) U 1 () = exp [i (p, )] ac (p , ) (8.5)
p0
and for the annihilation operator we have
s
1 (p)0
U () a (p, ) U () = exp [i (p, )] a (p , ) (8.6)
p0

211
212 CHAPTER 8. MASSLESS PARTICLE FIELDS

where p p, and the angle (p, ) is defined by Eq. (1.240) page 52. If we want the field to transform
according with a given representation D () of the homogeneous Lorentz group, we have [see Eq. (7.151)]
X 
U () k (x) U 1 () = Dkk 1 k (x) (8.7)
k

By using four-translations, we obtained equations (4.43, 4.44) that are in terms of irreducible representations
of the little group elements W (, p). Therefore, such equations are valid for either the massive or massless case,
by replacing the apropriate little group in each case. Hence, substituting the little group representation (8.3)
associated with the massless case in equations (4.43, 4.44), we see that the coefficients u and v have to satisfy the
conditions
s
p0 X
uk (p , ) exp [i (p, )] = Dkk () uk (p, ) (8.8)
(p)0 k
s
p0 X
vk (p , ) exp [i (p, )] = Dkk () vk (p, ) (8.9)
(p)0 k

By a procedure similar to the one follows in section 4.2.2, we can start from the standard four-momentum
(0, 0, k, k), and apply a standard boost L (p) that takes such a four momentum to an arbitrary physical massless
four-momentum (p, |p|). The result is given by
s
|k| X
uk (p, ) = Dkk (L (p)) uk (k, ) (8.10)
p0
k
s
|k| X
vk (p, ) = Dkk (L (p)) vk (k, ) (8.11)
p0
k

with k (0, 0, k) being the standard three-momentum. These equations are the massless analogous of Eqs. (4.45,
4.46) for massive particles. Note that here we have started from a standard non-zero three-momentum k, instead
of a zero three-momentum as we did in section 4.2.2. It owes to the fact that we cannot have physical massless
particles at zero three-momentum, since they travel at the speed of light in the vacuum. We can also see it by
observing that four-vectors of massless particles are of the form (p, |p|) which for p = 0 would indicate a zero
four-momentum i.e. no particle at all.
Now in section 4.2.3 we obtained a relation between coefficients at zero three-momentum for massive particles,
by using a transformation of the little group (i.e. rotations for the massive case). The analogous procedure for
massless particles permits to obtain a relation between coefficients at the standard three-momentum k, by using
transformations associated with the little group ISO (2). Therefore, the role of Eqs. (4.47, 4.48) [or equivalently
equations (4.49, 4.50)] for the coefficients at zero momentum for massive particles, is taken by the following
relations for the coefficient functions evaluated at the standard momentum k
X
uk (k, ) exp [i (k, W )] = Dkk (W ) uk (k, ) (8.12)
k
X
vk (k, ) exp [i (k, W )] = Dkk (W ) vk (k, ) (8.13)
k

where W is an arbitrary element of the little group associated with the four-momentum k = (k, |k|), i.e. a
transformation that leaves such a standard four-momentum invariant.
We can extract the content of Eqs. (8.12, 8.13) by considering separately the two transformations that provides
the little group as shown in Eqs. (1.206) page 45

W (, , ) = S (, ) R () (8.14)
213

First, we consider the rotation R () around the three-axis given by Eq. (1.205) page 45

cos sin 0 0
sin cos 0 0
R () =


(8.15)
0 0 1 0
0 0 0 1
from this rotation Eqs. (8.12, 8.13) yield
X
uk (k, ) ei = Dkk (R ()) uk (k, ) (8.16)
k
X
vk (k, ) ei = Dkk (R ()) vk (k, ) (8.17)
k

in addition, by using the other transformation S (, ) (which is a combination of a rotation and a boost) in the
x y plane, given by Eq. (1.204) page 45

1 0
2 2
S (, ) =
0 1 ; + (8.18)
1 2
1 +
equations (8.12, 8.13) yield
X
uk (k, ) = Dkk (S (, )) uk (k, ) (8.19)
k
X
vk (k, ) = Dkk (S (, )) vk (k, ) (8.20)
k

in summary, Eqs. (8.16, 8.17) and (8.19, 8.20) are the ones that determine the coefficient functions at the standard
momentum k. Then, Eqs. (8.10, 8.11) give us the corresponding coefficients at arbitrary momenta. Note that the
equations for v are just the complex conjugates of the equations for u. With a suitable choice of phases of and
it is possible to settle the coefficient functions such that
vk (p, ) = uk (p, ) (8.21)
the problem is that we cannot obtain a coefficient uk that satisfies Eq. (8.19) for general representations of the
homogeneous Lorentz group, even for those representations for which we were able to construct fields for particles
of a given helicity in the case of massive particles.
We shall see the inconsistency by trying to construct the field for a massless particle of helicity 1 from the
four-vector representation. Such a representation is characterized by
D () = (8.22)
since we shall be in the four-vector representation, it is convenient to change the index notation from k, k, . . . . to
the usual Minkowski notation , , . . .. As in the case of massive particles Eqs. (6.64, 6.65) page 170, it is
customary to write the coefficient function u in terms of a polarization vector e (p, )
1/2
u (p, ) 2p0 e (p, ) = (2 |p|)1/2 e (p, ) (8.23)
using (8.22, 8.23), equation (8.10) yields
s s
|k| |k| e (k, )
u (p, ) = D (L (p)) u (k, ) = L (p) p
p0 |p| 2 |k|
s
e (p, ) 1
p = L (p) e (k, )
2 |p| 2 |p|
214 CHAPTER 8. MASSLESS PARTICLE FIELDS

so that
e (p, ) = L (p) e (k, ) (8.24)
on the other hand, from Eq. (8.16) and using Eqs. (8.22, 8.23) we have
e (k, )
u (k, ) ei = D (R ()) u (k, ) = R () p
2 |k|

e (k, ) i
e (k, )
p e = R () p
2 |k| 2 |k|

and we can proceed the same with Eq. (8.19). Hence, in the four-vector representation Eqs. (8.16, 8.19) become

e (k, ) ei = R () e (k, ) (8.25)



e (k, ) = S (, ) e (k, ) (8.26)

setting = +1, and using = 1, 2, 3, 0 in Eq. (8.25) and using (8.15) we get

e (k, +1) ei = R () e (k, +1)


e1 (k, +1) ei = R ()1 1 e1 (k, +1) + R ()1 2 e2 (k, +1) + R ()1 3 e3 (k, +1) + R ()1 0 e0 (k, +1)
e1 (k, +1) ei = cos e1 (k, +1) + sin e2 (k, +1)
e2 (k, +1) ei = R ()2 1 e1 (k, +1) + R ()2 2 e2 (k, +1) = sin e1 (k, +1) + cos e2 (k, +1)
e3 (k, +1) ei = R ()3 3 e3 (k, +1) = e3 (k, +1) ; e0 (k, +1) ei = R ()0 0 e0 (k, +1) = e0 (k, +1)

in summary we obtain

e1 (k, +1) ei = cos e1 (k, +1) + sin e2 (k, +1)


e2 (k, +1) ei = sin e1 (k, +1) + cos e2 (k, +1)
e3 (k, +1) ei = e3 (k, +1) ; e0 (k, +1) ei = e0 (k, +1)

the only way that the last two equations can be satisfied for all is by setting

e3 (k, +1) = e0 (k, +1) = 0

and the first two equations for e1 (k, +1) and e2 (k, +1) can be rewritten as
 
cos ei e1 (k, +1) + e2 (k, +1) sin = 0 (8.27)
 
e1 (k, +1) sin + e2 (k, +1) cos ei = 0 (8.28)

then we should find a non-trivial solution for this homogeneous system of two equations with two variables. Setting
e1 (k, +1) = 1, we obtain from equation (8.28)
   
sin + cos ei e2 (k, +1) = 0 cos ei e2 (k, +1) = sin

using the identity1 cos ei = i sin , we find

ie2 (k, +1) sin = sin ie2 (k, +1) = 1 e2 (k, +1) = i
1
The identity follows from
ei + ei 2ei ei + ei ei ei
cos ei = = = i = i sin
2 2 2i
215

and replacing e1 (k, +1) = 1 and e2 (k, +1) = i in Eq. (8.27) we see that it is a consistent solution. Then we have
e1 (k, +1) = 1, e2 (k, +1) = i, e3 (k, +1) = e0 (k, +1) = 0. From Eq. (8.25) it is clear that we can multiply this
vector by a constant and it remains being a valid solution. Thus we normalize it to obtain finally
1
e (k, +1) = (1, +i, 0, 0)
2
we can do the same for e (k, 1). We conclude that equation (8.25) requires that (up to a constant that can be
absorbed in the coefficients and ) the polarization vector be
1
e (k, 1) = (1, i, 0, 0) (8.29)
2
on the other hand, substituting (8.18) and (8.29) in Eq. (8.26) we obtain the condition

e (k, 1) = S (, ) e (k, 1)

1 1 0 1
i 0 1 i

0 = 1 0
0 1 + 0

1 1
i
= i
0 i
0 i

In summary, Eq. (8.25) leads to a solution of the form (8.29) for the polarization vectors, but then Eq. (8.26) also
requires that i = 0 which cannot be true for arbitrary real values of and (recall that the parameters , ,
of our little group must be all real). Consequently, we cannot satisfy the requirement (8.19) or (8.12). Instead,
we can obtain the transformation of e (k, 1) through a complete transformation W (, , ) of the Little group

D (W (, , )) e (k, 1) = S (, ) R () e (k, 1)


1 0 cos sin 0 0 1
1 0 1 sin cos 0 0
i

=


2 1 0 0 1 0 0
1 + 0 0 0 1 0

1 0 cos i sin cos i sin
1 0 1
i cos sin
1 i cos sin
=

= 2 (cos i sin ) (sin i cos )

2 1 0
1 + 0 (cos i sin ) (sin i cos )


ei ei 1
1 iei ei i
=
i (cos i sin ) = 1 =
2 e i (cos i sin ) 2 e 2 i
i i ie i

ei i (cos i sin ) ei iei i



1 0 1 0
ei i i
+ e ( i) 0 = ei 1 i + ei ( i) 0

=
2 0 2 1 2 0 2 |k| |k|
0 1 0 |k|
216 CHAPTER 8. MASSLESS PARTICLE FIELDS

then we obtain finally


 
( i)
D (W (, , )) e (k, 1) = S (, ) R () e (k, 1) = exp (i) e (k, 1) + k (8.30)
2 |k|
hence we conclude that we cannot construct a four-vector field from the annihilation and creation operators for a
particle of null mass and helicity 1.
Let us ignore this difficulty for now. From Eqs. (8.24, 8.29) we are able to define a polarization vector at
arbitrary momentum. Moreover taking into account Eqs. (8.21, 8.23) the field (8.1) can be written in terms of
the polarization vectors as follows
Z X h i
a (x) = (2)3/2 d3 p a (p, ) u (p, ) eipx + ac (p, ) u (p, ) eipx
=1
Z
1/2 X h i
a (x) = (2)3/2 d3 p 2p0 e (p, ) eipx a (p, ) + e (p, ) eipx ac (p, ) (8.31)
=1

we shall see later the utility of this field in physical theories. Since the field (8.31) is defined on a single species of
particle plus its antiparticle, it is clear that such a field satisfies the Klein-Gordon equation [see discussion below
Eq. (4.72) page 150]
a (x) = 0 (8.32)
which for null mass becomes the wave equation. Other properties of the field follow from the properties of the
polarization vector.
On the other hand, we saw in section 1.11.4, that the Lorentz transformation L (p) that takes a massless
particle momentum from k to p can be decomposed as a boost B (|p| / |k|) along the X3 axis that takes the
particle from energy |k| to energy |p|; followed by a rotation R (b
p) that takes the X3 direction to the p direction
[see Eqs. (1.241, 1.242) page 52, and Eq. (1.177), page 40].

1 0 0 0
  0 1 0 0
|p|
L (p) = R (b p) B ; B (u) = ( u 2 +1
) (u 2 1
) (8.33)
|k| 0 0 2u 2u
(u 1) (u +1)
2 2
0 0 2u 2u

pb1 pb3 2 pb2 2 pb1 0
cos cos sin cos sin 0 1b p3 1bp3
cos sin cos sin sin 0 pb2 pb3 pb1
pb2 0
R (b
p) = = 2 2 (8.34)
sin 0 cos 0 p1bp3 1b p3
1 pb 2 0 pb3 0
0 0 0 1 3
0 0 0 1
On the other hand, from Eq. (8.29) it is clear that e (k, 1) is a purely spatial vector with only x1 and x2
components. Therefore, e (k, 1) is left invariant under the transformation of a boost along the X3 axis. It can
be checked explicitly from Eqs. (8.33, 8.29)

1 0 0 0 1 1
0 1 0 0
i
= i = e (k, 1)
B (u) e (k, 1) = (u2 +1) (u2 1)
0 0 2u 2u 0 0
(u2 1) (u2 +1) 0 0
0 0 2u 2u

Consequently, equation (8.24) gives

e (p, 1) = L (p) e (k, 1) = R (b


p) B (|p| / |k|) e (k, 1) = R (b
p) e (k, 1)

p) e (k, 1)
e (p, 1) = R (b (8.35)
217

we also see from Eq. (8.29) that

e0 (k, 1) = 0 k e (k, 1) = ki ei (k, 1) = |k| e3 (k, 1) = 0

hence
e0 (k, 1) = 0 and k e (k, 1) = k e (k, 1) = 0
we can see that the same is true for an arbitrary p. To show it, we substitute Eqs. (8.29, 8.34) in Eq. (8.35) to
have

p) e (k, 1)
e (p, 1) = R (b

pb1 pb3 2 pb2 2 pb1 0 pb1 pb3 2 i pb2 2
1b p 3 1b p3 1 1b p3 1b
p3
pb2 pb3 pb1
pb2 0 i i pb1 + pb2 pb3
=
p1bp3
2 1b p23
0 = 1b p23
p p23
1b

1 pb2 0 pb3 0 1 pb3 2
3 0
0 0 0 1 0

pb1 pb3 ib p2
1 ib p 1 + p
b 2pb3 pi pi
e (p, 1) = p ; pbi = (8.36)
1 pb3 2 1 pb3 2 |p| E
0

therefore e0 (p, 1) = 0 and then



i i p1 pb3 ib
pb1 (b p2 ) + pb2 (ibp1 + pb2 pb3 ) pb3 1 pb23
p e (p, 1) = pi e (p, 1) = |p| pbi e (p, 1) = |p| p
1 pb23
 

pb21 + pb22 + pb23 1 pb3 p|2 1 pb3
|b
= |p| p = |p| p =0
1 pb23 1 pb23

now since e (k, 1) = e (k, 1), and R (b


p) is real, we also obtain

e (p, 1) = e (p, 1) (8.37)

then we also obtain p e (p, 1) = 0. Summarizing we have

e0 (p, 1) = 0 ; p e (p, 1) = p e (p, 1) = p e (p, 1) = 0 (8.38)

from which we conclude that the polarization vector is purely spatial and orthogonal to the direction of propagation.
Setting = 0 in (8.31) and using e0 (p, 1) = 0, we see that a0 (x) = 0. Applying the operator on both sides
of Eq. (8.31) we have
Z X n  o
3/2 3

0 1/2

ipx
 
ipx c
a (x) = (2) d p 2p e e (p, ) a (p, ) + e e (p, ) a (p, )
=1
Z X n o

0 1/2
a (x) = (2)3/2 d3 p 2p ip e (p, ) eipx a (p, ) ieipx p e (p, ) ac (p, )
=1

and using (8.38) we obtain a (x) = 0, but a0 (x) = 0, so that i ai (x) = 0. We summarize it as follows

a0 (x) = 0 ; a (x) = 0 ; a (x) a (x) , a0 (x) (8.39)

in quantum electrodynamics, these are the conditions that are satisfied by the vacuum potential four-vector in
the Coulomb or radiation gauge.
218 CHAPTER 8. MASSLESS PARTICLE FIELDS

Since a0 (x) is null in all reference frames, it is clear that a (x) is not a Lorentz four-vector. Instead, Eq.
(8.30) show that for an arbitrary momentum p and an arbitrary homogeneous Lorentz transformation , in place
of Eq. (8.8), the polarization vector must hold the relation

e (p , 1) exp [i (p, )] = D () e (p, 1) + p (p, )

then, for a general homogenous Lorentz transformation we have

U () a (x) U 1 () = a (x) + (x, ) (8.40)

where (x, ) is constructed from a linear combination of annihilation and creation operators. However, its
explicit form is not relevant for our present discussion. A field like a (x) can be part in a Lorentz invariant
Physical theory if the couplings of a (x) besides being formally Lorentz invariant2 , is also invariant under the
gauge transformation a a + . It is carried out by demanding the couplings of a to have the form a j
where j is a four-vector current that satisfies the continuity equation j = 0.
Despite there is no ordinary four-vector field associated with massless particles of helicities 1, we can construct
an antisymmetric tensor field associated with such particles. Now we recall that

D () D () = (8.41)

provides the tensor representation of the homogenous Lorentz group coming from the tensor product of two vector
representations. We shall apply the representation (8.41) on an antisymmetric tensor Bk defined by

Bk = k e (k, 1) k e (k, 1)

from Eq. (8.30) and using the invariance of the standard four-momentum k under the little group, we obtain

D (W (, , )) D (W (, , )) [k e (k, 1) k e (k, 1)]


= D (W (, , )) D (W (, , )) k e (k, 1) D (W (, , )) D (W (, , )) k e (k, 1)
= [D (W (, , )) k ] [D (W (, , )) e (k, 1)] [D (W (, , )) k ] [D (W (, , )) e (k, 1)]

= k [D (W (, , )) e (k, 1)] k [D (W (, , )) e (k, 1)]


   
i ( i) i ( i)
= e k e (k, 1) + k e k e (k, 1) + k
2 |k| 2 |k|
then we obtain finally

D (W (, , )) D (W (, , )) [k e (k, 1) k e (k, 1)] = ei [k e (k, 1) k e (k, 1)](8.42)


D (W (, , )) D (W (, , )) Bk = ei Bk (8.43)

Hence, Eqs. (8.42, 8.43) show that under an appropriate choice of normalization, the coefficient function that
satisfies Eq. (8.8) for the antisymmetric tensor representation of the homogeneous Lorentz group is
3/2
u (p, 1) = i (2)3/2 2p0 Bp
3/2
u (p, 1) = i (2)3/2 2p0 [p e (p, 1) p e (p, 1)]

with e (p, 1) given by Eq. (8.35). From this along with Eq. (8.31) we obtain the general antisymmetric tensor
field for massless particles of helicity 1, as follows

f (x) = a a (8.44)
2
It means that a must be invariant under formal Lorentz transformations under which a a .
219

it worths pointing out that this is a tensor despite a is not a four-vector. It owes to the fact that the extra term
in Eq. (8.40) that prevents a to be a four-vector is cancelled in Eq. (8.44). For future purpose, by using the
total antisymmetry of we evaluate the quantities
a = a = a = a
a = a = a = a
where in the last step of each line we have used the fact that all indices are dummy. Hence such quantities are
null
a = a = 0 (8.45)
In addition, from Eqs. (8.44, 8.32) and (8.39) we see that
f (x) = ( a a ) = ( a a ) = a ( a )
f (x) = 0
and using (8.45) we obtain
f = ( a a ) = a a = 0
hence from the field equations (8.44, 8.32, 8.39) for the vector field a (x), and the antisymmetric nature of ,
we derive the following field equations for the tensor field f
f = 0 ; f = 0 (8.46)
which are identical in form to the vacuum Maxwell equations.
Now, it is important to calculate the commutation relations for the tensor fields. To do it, we require to sum
over helicities of the bilinears e e .
X X
Tkij ei (k, ) ej (k, ) = ei (k, ) ej (k, )
=1 =1

where we have used ei (k, 1) = ei (k, 1). With the assignment it becomes
X X
Tkij ei (k, ) ej (k, ) = ej (k, ) ei (k, )
=1 =1
X
Tkij = Tkji i j
e (k, ) e (k, )
=1

. Using Eq. (8.29) and taking into account that k1 = k2 = 0 and k3 = |k|, we obtain
X 1 1 1 1 k1 k1
e1 (k, ) e1 (k, ) = + = 1 = 11
=1
2 2 2 2 |k|2
X 1 1 k1 k2
e1 (k, ) e2 (k, ) = i i = 0 = 12
=1
2 2 |k|2
X k1 k3
e1 (k, ) e3 (k, ) = 0 = 13
=1 |k|2

X 1 1 k2 k2
e2 (k, ) e2 (k, ) = i (i) + (i) i = 1 = 22
=1
2 2 |k|2
X k2 k3
e2 (k, ) e3 (k, ) = 0 = 13
=1 |k|2
X |k| |k| k3 k3
e3 (k, ) e3 (k, ) = 0 = 33 = 33
=1
|k|2 |k|2
220 CHAPTER 8. MASSLESS PARTICLE FIELDS

so that we obtain
X ki kj
Tkij ei (k, ) ej (k, ) = ij (8.47)
=1
|k|2

p)i 0 = e0 (k, 1) = 0, we find


further, using (8.35) and R (b
X X h ih i
Tpij ei (p, ) ej (p, ) = R (bp)i m em (k, ) R (b
p)j n en (k, )
=1 =1
X  
i j m n i j km kn
= R (b
p) m R (b
p) n e (k, ) e (k, ) = R (b
p) m R (b
p) n mn
=1 |k|2
h ih i
i j
R (bp) m R (b
p) km n kn
= R (bp)i n R (b
p)j n 2
|k|
h ih i
i m j n
R (bp) m k p) n k
R (b
Tpij = ij
|k|2
p) k by
where we have used the orthogonality condition for rotations in the last step. Let us do explicitly R (b
using (8.34)
pb1 pb3 2 pb2 2 pb1 0
1b p 3 1b
p 3 0 |k| pb1
pb2 pb3 pb1 2 pb2 0
R (b p) =
p) k = R (b 2 0 = |k| pb2
p1bp3 1bp3 |k| |k| pb3
1 pb2 0 pb3 0
3 |k| |k|
0 0 0 1
then we have   
X |k| pbi |k| pbj
Tpij i j
e (p, ) e (p, ) = ij 2 = ij pbi pbj
=1 |k|
thus, we obtain finally
X pi pj
Tpij ei (p, ) ej (p, ) = ij (8.48)
=1 |p|2
now we are able to calculate the commutation relations between the tensor fields. First, from the expression (8.31)
for a the explicit form of the tensor f gives

f (x) = a a
Z
3/2 1/2 X 
= (2) d3 p 2p0 a (p, ) [e (p, ) e (p, ) ] eipx
=1
  o
+ a (p, )c
e (p, ) e (p, ) eipx

and defining


;
x y
the commutator between two of those tensors yields
h i Z Z
3 
0 1/2
1/2 X X 
f (x) , f (y) = (2) 3
d p 2p d3 p 2p0 a (p, ) {e (p, ) e (p, ) } eipx
=1 =1
   
+a (p, ) e (p, ) e (p, ) eipx , a p , e p , e p , eip
c

   i
+ ac p , e p , e p , eip y
221

h i Z Z
1/2 1/2 X X

f (x) , f (y) = (2)3 d3 p 2p0 d3 p 2p0
=1 =1
n   
h i
||2 {e (p, ) e (p, ) } e p , e p , eipx eip y a (p, ) , a p ,
   
h 
+ ||2 e (p, ) e (p, ) e p , e p , eipx eip y ac (p, ) , ac p ,

h i Z
3 1 X
f (x) , f (y) = (2) d3 p 2p0
=1
n 
||2 {e (p, ) e (p, ) } e (p, ) e (p, ) eip(xy)
  o
||2 e (p, ) e (p, ) e (p, ) e (p, ) eip(xy)

but eip(xy) = eip(xy) , therefore


h i Z
3 1 X
f (x) , f (y) = (2) d3 p 2p0
=1
n 
2
|| {e (p, ) e (p, ) } e (p, ) e (p, ) eip(xy)
 o
+ ||2 e (p, ) e (p, ) {e (p, ) e (p, ) } eip(xy)

h i Z
3 1 X 
f (x) , f (y) = (2) d3 p 2p0 {e (p, ) e (p, ) } e (p, ) e (p, )
=1
n o
||2 eip(xy) + ||2 eip(xy)

h i Z
3 1 X 
f (x) , f (y) = (2) d3 p 2p0 e (p, ) e (p, ) e (p, ) e (p, )
=1

e (p, ) e
(p, ) + e (p, ) e (p, )
n o
||2 eip(xy) + ||2 eip(xy)

Z (
h i 1 X X

f (x) , f (y) = (2)3 d3 p 2p0 e (p, ) e (p, ) + e (p, ) e (p, )
=1 =1
)
X X
+ e (p, ) e (p, ) e (p, ) e (p, )
=1 =1
n o
||2 eip(xy) ||2 eip(xy)

obtaining finally ???


h i

f (x) , f (y) = (2)3 [g + g + g g ]
Z
1 h i
d3 p 0 ||2 eip(xy) ||2 eip(xy)
2p
222 CHAPTER 8. MASSLESS PARTICLE FIELDS

which clearly vanishes for x0 = y 0 if and only if

||2 = ||2 (8.49)

in that case since f is a tensor, the commutator also vanishes for all space-like separations. On the other hand,
Eq. (8.49) also implies that the commutators of the a vanish at equal times, and it can be shown that it is
enough to provide the Lorentz invariance of the Smatrix.
Once again, the relative phase between the creation and annihilation operators can be defined so that = .
Moreover, if the particles are their own antiparticles the fields will be hermitian. This is the case of the photon.
A priori we could think in using only fields of the type f instead of fields of the type a (x). After all f is a
Lorentz tensor while a is not. However, and interaction density constructed only from f and its derivatives will
have matrix elements that vanish very quickly (faster than in theories that uses the a field) for small energy and
momentum of massless particles, those interactions at large distances would fall off faster than the usual inverse
square law. It owes to the presence of derivatives in the definition (8.44) of f . Though it is in principle possible,
gauge invariant theories that use vector fields for massless spin one particles represent a more general class of
theories, that in particular include the interactions realized in nature.
Similar features appear when apply the theory to the (hypothetical) gravitons, which are massless particles
of helicity 2. From the creation and annihilation operators of those particles, we can construct a tensor R
with the algebraic properties of the Riemann-Christoffel curvature, that is they are antisymmetric within the
pairs , and , and symmetric between the pairs. However, to obtain an interaction with the inverse square
law characteristic of gravitational interactions, we have to introduce a field h that transforms as a symmetric
tensor upon to gauge transformations [analogous to Eq. (8.40)] associated with the general relativity with general
coordinate transformations.
As in the case of electromagnetic gauge invariance, we achieve long-range interactions by requiring general
covariance that is satisfied by coupling the field to a conserved tensor current satisfying = 0. Roughly
speaking, the tensor structure of the formalism is duplicated with respect to electrodynamics.
Chapter 9

The Feynman rules

The requirement that the Smatrix satisfies Lorentz invariance and the Cluster Decomposition Principle (CDP)
has led us naturally to the construction of the Hamiltonian density based on covariant free fields. This construction
has the advantage of satisfying automatically the Lorentz invariance and the CDP properties of the Smatrix
in each order of the interaction density, whatever the form of the perturbation theory chosen. However, in the
approach called old-fashioned perturbation theory described in Sec. 2.6, the satisfaction of such conditions is
not manifest at each stage of the calculation. In the present chapter we shall describe the perturbative approach
developed by Feynman, Tomonaga and Schwinger in the formed described by Dyson (1949). Such an approach
has the advantage that Lorentz invariance and clustering conditions are apparent at each step of the calculations.

9.1 General framework


The starting point is obtained by combining the Dyson series (2.192) for the Smatrix, with Eq. (3.8) for
free-particle states

Sp1 1 n1 ;p2 2 n2 ; ,p1 1 n1 ;p2 2 n2 ;


X Z
(i)N  
= d4 x1 d4 xN h0| a p2 2 n2 a p1 1 n1 T {H (x1 ) H (xN )} a (p1 1 n1 ) a (p2 2 n2 ) |0i
N!
N =0
(9.1)

we recall that p, , n denote particle momenta, spin and species, |0i denotes the free-particle vacuum state, the
primes denote labels for particles in the final state, while non-prime denotes particles in the initial state. Further,
a and a are annihilation and creation operators, T means time-ordering such that the interaction densities H (x)
are put in an order in which the arguments x0 decrease from left to right. On the other hand, according with Eq.
(4.16), page 139, the Hamiltonian density can be written as a polynomial in the fields and their adjoints1
X
H (x) = gi Hi (x) (9.2)
i

where Hi (x) is a product of definite numbers of fields and field adjoints of each type. Moreover, Eqs. (4.39, 4.40)
page 143 along with Eq. (4.56) page 147, describe the field (and the field adjoint) associated with a particle species
n and its associated antiparticle nc , that transforms under a given representation of the homogeneous Lorentz
1
In Eq. (4.16) Pag. 139, the Hamiltonian density is written in terms of the creation and annihilation fields k+ (x) and k (x).
However, to show manifest Lorentz covariance and the conservation of internal symmetries, such creation and annihilation fields should
be written in terms of fields and field adjoints of the type given by Eq. (5.23) Pag. 156.

223
224 CHAPTER 9. THE FEYNMAN RULES

group (with or without space inversions)


X Z h i
3/2
k (x) = (2) d3 p uk (p,, n) a (p,, n) eipx + vk (p,, n) a (p,, nc ) eipx (9.3)

X Z h i
k (x) = (2)3/2 d3 p uk (p,, n) a (p,, n) eipx + vk (p,, n) a (p,, nc ) eipx (9.4)

the exponential exp (ip x) is calculated by taking


p
p0 = p2 + m2n

in chapters 5, 6, 7, we saw that the coefficients uk and vk depend on the Lorentz transformation properties of the
field and the spin of the particle that it describes. For example Eq. (5.4) page 151 show that for the scalar field
with energy E the coefficient uk is given by (2E)1/2 . We also recall that the index k, on the field should denote
the particle type and the representation of the Lorentz group under which the field transforms, and also includes
a running index over the components in this representation. We shall not consider separately the interactions
involving derivatives of fields, they will just be considered as other fields of the type (9.3) with different coefficients
uk and vk .
We shall call by convention certain species as particles such as electrons, protons or neutrons, and their
corresponding conjugates will be called antiparticles such as positrons, antiprotons and antineutrons. The field
operators that destroy particles and create antiparticles are called by convention simply as fields, while their
corresponding adjoints which destroy antiparticles and create particles are called field adjoints. As we have
seen, there are some particle species that coincide with their antiparticles. In that case, the field adjoints are
proportional to the fields.
Now, with a procedure similar to the one described in section 3.6.1 page 129, we proceed to move all annihilation
operators to the right in Eq. (9.1). It can be done by repeated use of the commutation or anticommutation relations
(3.17, 3.18, 3.19) page 120
  
a (p, , n) a p n = a p n a (p, , n) + 3 p p n n (9.5)



a (p, , n) a p n = a p n a (p, , n) (9.6)
 
a (p, , n) a p n = a p n a (p, , n) (9.7)

+1 if n and/or n are bosons
= (9.8)
1 if n, n are both fermions

also, as discussed in section 3.6.1, when an annihilation operator appears on the extreme right (or a creation
operator appears on the extreme left) its contribution associated with Eq. (9.1) vanishes since according with Eq.
(3.13) these operators annihilate the vacuum state

a (p, , n) |0i = 0 h0| a (p, , n) = 0 (9.9)

and the remaining contributions to Eq. (9.1) are the ones associated with Dirac functions terms on the RHS of
Eq. (9.5), with every creation and annihilation operator in the initial or final states or in the interaction density
paired in this way with some other annihilation or creation operator.
Therefore, each pairing of the form a (p, , n) a (p n ) is finally replaced by the corresponding delta factors
given by Eq. (9.5)  
a (p, , n) a p n 3 p p n n
or equivalently according with Eq. (9.5)
   
a (p, , n) a p n 3 p p n n = a (p, , n) a p n a p n a (p, , n)
 h i
a (p, , n) a p n a (p, , n) , a p n

9.1. GENERAL FRAMEWORK 225

so that the pairing a (p, , n) a (p n ) should be associated with its corresponding commutator or anticommu-
tator.
Note that in the Dyson series (9.1) several kind of pairings of operators can occur, the sequence of operators
in such a series has the order
 
O a p2 2 n2 a p1 1 n1 T {H (x1 ) H (xN )} a (p1 1 n1 ) a (p2 2 n2 ) (9.10)

for instance, an annihilation operator a (p n ) of a final particle could be paired with a field k (x) (contained
in a given interaction Hi (x)), as discussed above, such a pairing should be associated with its commutator or
anticommutator
X Z nh i

  3/2 
a p n , k (x) = (2) d3 p a p n , uk (p,, n) a (p,, n) eipx + vk (p,, n) a (p,, nc ) eipx

X Z h i
3/2 
= (2) d3 p a p n , a (p,, nc ) vk (p,, n) eipx

X Z
   
a p n , k (x) = (2)3/2

n nc d3 p vk (p,, n) p p eipx


 
a p n , k (x) = 0

this commutator or anticommutator vanishes since n is a particle so that n nc = 0. For the pairing of a (p n )
with a field adjoint we have
h i X Z nh i

 3/2 
a p n , k (x) = (2) d3 p a p n , uk (p,, n) a (p,, n) eipx + vk (p,, n) a (p,, nc ) eipx


X Z h i
3/2 
= (2) d3 p a p n , a (p,, n) uk (p,, n) eipx

h i X Z
 
a p n , k (x)

= (2) 3/2
n n d3 p uk (p,, n) p p eipx


h  i 

a pn , k (x) = (2) 3/2
uk p , , n eip x

note that the term [a (p n ) , k (x)] would be nonzero if the particle coincides with its own antiparticle.
However, in that case the field musth be proportional to i the field adjoint so that we would be able to keep only

the commutator or anticommutator a (p n ) , k (x) .

In addition, an annihilation operator a (p nc ) of a final antiparticle could be paired with a field k (x) or
field adjoint k (x). With the same reasoning as before we obtain
h  i
c
a pn , k
(x) = 0

   
a p nc , k (x) = (2)3/2 eip x vk p n

It can also occurs the pairing of a creation operator a (pn) of an initial particle with a field k (x) or a field
adjoint k (x) in an interaction Hi (x). From Eq. (9.10) we see that this pairing occurs with the field to the left
of the creation operator
h i
k (x) , a (pn) = 0
h i
k (x) , a (pn) = (2)3/2 eipx uk (pn)

226 CHAPTER 9. THE FEYNMAN RULES

the pairing of a creation operator a (pnc ) of an initial antiparticle with a field k (x) or a field adjoint k (x)
yields
h i
k (x) , a (pnc ) = (2)3/2 eipx vk (pn)

h i
k (x) , a (pnc ) = 0

pairing of a final particle (or antiparticle) with an initial particle (or antiparticle) yields2
h  i 
a p n , a (pn) = 3 p p n n

it is clear that such a paring vanishes when it occurs between a particle and an antiparticle.
Finally the paring between a field (or field adjoint) in an interaction term Hi (x) with a field (or field adjoint)
in another interaction term Hj (x). In this case, it is not convenient to express the pairing in terms of fields or field
adjoints. It is because when expanding a Hamiltonian density the creation and annihilation fields are in a very
specific order (normal order) as can be seen in Eq. (4.16), page 139. By contrast, when replacing the creation and
annihilation fields by the covariant fields and field adjoints, it is not clear the order in which the latter appear.
Thus, it is more convenient to express this pairing in terms of creation and annihilation fields. From Eq. (4.4)
page 137, we then see that h i
 + 
k (x) , m +
(y) = k+ (x) , m +
(y)

for the non-trivial pairings we should preserve causality then we write


h i h i
(x y) k+ (x) , m
+
(y) (y x) m
(y) , k (x) ikm (x, y) (9.11)


+1 if x0 > y 0
(x y) (9.12)
0 if x0 < y 0

the sign will be explained later.

9.2 Rules for the calculation of the Smatrix


From the previous framework, the contribution
 to Eq. (9.1) of a given order in each of the terms Hi (x) in

the polynomial (9.2) H (x) , (x) is described by a sum over all ways of pairing creation and annihilation
operators, of the integrals of products of factors, in this way

1. Pairing of a final particle with quantum numbers p , , n with a field adjoint k (x) in Hi (x) provides a
factor h  i 

a p n , k (x) = (2)3/2 eip x uk p n (9.13)

2. Pairing of a final antiparticle with quantum numbers p , , nc with a field k (x) in Hi (x) provides a factor
   
a p nc , k (x) = (2)3/2 eip x vk p n (9.14)

3. Pairing of an initial particle with quantum numbers p, , n with a field k (x) in Hi (x) provides a factor
h i
k (x) , a (pn) = (2)3/2 eipx uk (pn) (9.15)

2
In Eq. (9.10), we see that such a pairing occurs either (a) when N = 0, i.e. when there are no vertices of interaction or (b) when
the corresponding annihilation operator has jumped over all interaction operators.
9.3. DIAGRAMMATIC RULES FOR THE SMATRIX 227

4. Pairing of an initial antiparticle state with quantum numbers p, , nc with a field adjoint k (x) in Hi (x)
gives a factor h i
k (x) , a (pnc ) = (2)3/2 eipx vk (pn) (9.16)

5. Pairing of a final particle (or antiparticle) with numbers p , , n with an initial particle (or antiparticle)
with numbers p, , n yields a factor
h  i 
a p n , a (pn) = 3 p p n n (9.17)


6. Pairing of a field k (x) in Hi (x) with a field adjoint m (x) in Hj (y) provides a factor
h i h i
(x y) k+ (x) , m
+
(y) (y x) m (y) , k (x) ikm (x, y) (9.18)


+1 if x0 > y 0
(x y) (9.19)
0 if x0 < y 0

where + (x) and are the terms in that destroy particles and create antiparticles respectively. They
are given by Eqs. (4.39, 4.40) page 143
Z X
k+ (x) = (2)3/2 d3 p uk (p, , n) eipx a (p, , n) (9.20)

Z X

m (x) = (2)3/2 d3 p vm (p, , n) eipx a (p, , nc ) (9.21)

the step function in (9.18) arise from the time ordering in (9.1).

It worths saying that if the interaction density H (x) is written in the normal ordered form as in Eq. (4.16),
there will be no pairing of fields and field adjoints in the same interaction Hi (x), so there are no terms of the form
h i
k (x) , m (x) (x x)

Otherwise, some kind of regularization is required to give a meaning to km (0). Moreover, we can encounter
a pairing between an annihilation field + (x) in H (x) with a creation field + (y) in H (y) only if H (x) was
initially to the left of H (y) in Eq. (9.1) or equivalently if x0 > y 0 . Similarly, we have a pairing of an annihilation
field (y) in H (y) with a creation field (x) in H (x) only if H (y) was initially to the left of H (x) in (9.1)
or equivalently, if y 0 > x0 . The appearance of in the second term of Eq. (9.18) will be clarified later. The
quantity (9.18) is called a propagator.
From Eq. (9.1), it is clear that the Smatrix is obtained by multiplying all those factors, along with additional
numerical factors that we shall see later, then integrating over x1 xN , summing over all pairings, and then over
the numbers of interactions of each type.

9.3 Diagrammatic rules for the Smatrix


It is convenient to represent each factor described in section 9.2 to calculate the Smatrix, through suitable
diagrams (see Fig. 9.1). To do it, we construct diagrams consisting of points called vertices, where each vertex
represents one of the Hi (x), and lines that represent the pairing of a creation with an annihilation operator. All
this with the following algorithm
228 CHAPTER 9. THE FEYNMAN RULES

Figure 9.1: For each pairing of operators that arise in the coordinate-space evaluation of the Smatrix, there is a
factor and a diagram associated. The lines represent the diagrams and the expressions on the right represent the
factor associated.

1. The pairing of a final particle with a field adjoint in one of the H (x), will be represented by a line running
from the vertex representing that H (x), upwards out of the diagram, carrying an arrow pointed upwards
[see Fig. 9.1 (1)].

2. The pairing of a final antiparticle with a field in one of the H (x), will be represented by a line running from
the vertex that represents this H (x) upwards out of the diagram, and carrying an arrow pointed downwards
[see Fig. 9.1 (2)]. Arrows are omitted for particles that coincide with their antiparticles (sometimes arrows
pointing in both directions are used for these kind of particles).

3. The pairing of an initial particle with a field in one of the H (x) will be represented by a line running into
the diagram from below, ending in the vertex that represents this H (x), and carrying an arrow pointing
upwards [see Fig. 9.1 (3)].

4. The pairing of an initial antiparticle with a field in one of the H (x) is represented by a line running into the
diagram from below, ending in the vertex that represents this H (x), carrying an arrow pointing downwards
[see Fig. 9.1 (4)].

5. The pairing of a final particle or antiparticle with an initial particle or antiparticle will be represented by a
line running clear through the diagram from bottom to top, not touching any vertex, with an arrow pointing
upwards for particles and downwards for antiparticles [see Fig. 9.1 (5)].

6. The pairing of a field in H (x) with a field adjoint in H (y) will be represented by a line joining the vertices
associated with H (x) and H (y), with an arrow pointing from y to x [see Fig. 9.1 (6)].

Observe that arrows always point in the direction in which a particle is moving and in the opposite direction
of motion of an antiparticle. Thus it is reasonable not to put any arrow (or put arrows in both senses), when
the particle coincides with its own antiparticle. Note that the arrow direction in rule 6 is consistent with this
convention since a field adjoint in Hj (y) can either create a particle destroyed by a field in Hi (x), or destroy an
antiparticle created by a field in Hi (x). Further, since every field or field adjoint in Hi (x) must be paired with
9.4. CALCULATION OF THE SMATRIX FROM THE FACTORS AND DIAGRAMS 229

something, the total number of lines at a vertex of type i, associated with a term Hi (x) in Eq. (9.2), is equal to
the total number of field or field adjoint factors in Hi (x). Of these lines, the number with arrows pointed into the
vertex or out of it, equals the number of fields or field adjoints respectively in the associated interaction term.
For example, each diagram in Fig. 9.4, page 232, contains two vertices, each one associated with an interaction
term of the type given by Eq. (9.24) which contains two fields and one field adjoint. Each vertex in each diagram
contains three lines equal to the total number of fields (two) or field adjoint factors (one) in each Hi (x).

9.4 Calculation of the Smatrix from the factors and diagrams


From the rules previously established we can calculate the contribution to the Smatrix for a given process, of a
given order Ni in each of the interaction terms Hi (x) in Eq. (9.2), as follows

1. We start by drawing all Feynman diagrams that contain Ni vertices of each type i, and containing a line
coming into the diagrams from below for each particle or antiparticle in the initial state, and a line going
upward out of the diagram for each particle or antiparticle in the final state. In addition, the diagram should
contain internal lines connecting one vertex to another, in a number so that we give each vertex the proper
number of lines attached to it. The lines carry arrows pointing upwards or downwards as we described in
section 9.3. Each vertex is labelled with an interaction type i and a space coordinate x . Each line (internal
or external) is labelled at the end where it runs into a vertex with a field (or a field adjoint) type k associated
with a field k (x) or a field adjoint k (x) which creates or destroys the particle or antiparticle at the given
vertex. Finally, each external line that enters or leave the diagram are labelled with the quantum numbers
p, , n or p , , n associated with the initial or final particle (or antiparticle) respectively [See Fig. 9.1, page
228].

2. For each vertex of type i, we must include a factor (i) [coming from the factor (i)N in Eq. (9.1)] and a
factor gi [that defines the coupling constant in the product of fields (9.2) in Hi (x)]. For each line running
upwards out of the diagram, include a factor (9.13) or (9.14) for the arrow pointing up or down respectively.
For each line running from below into the diagram, include a factor (9.15) or (9.16) for the arrow pointing
up or down respectively. For each line running straight through the diagram include a factor (9.17). Finally,
for each internal line connecting two vertices include a factor (9.18).

3. We then integrate the product of these factors over the coordinates x1 , x2 , . . . of each vertex.

4. We then sum the contribution of each Feynman diagram. The complete perturbation series for the Smatrix
is obtained by adding up all the contributions of each order in each interaction type, up to the order we
want to calculate.

Note that we have not included a factor 1/N !, that appears in Eq. (9.1). It is because the time-ordered product
in (9.1) is a sum over the N ! permutations of x1 x2 xN ., where each permutation gives the same contribution
to the final result. In terms of diagrams it means that a given Feynman diagram with N vertices is on a set of
N ! identical diagrams, that differ from each other only by permutations of the labels of the vertices. Hence, by
adding up all these N ! identical diagrams we cancel the 1/N ! in Eq. (9.1). In summary, we do not include more
than one diagram that differ from the others only by the labelling of the vertices.
However, there are some exceptions to the rules described above. In some cases there are additional combina-
toric factors or signs that must be included in the contribution of a given Feynman diagram

Suppose that an interaction Hi (x) contains among other fields and field adjoints, M factors of the same
field. Now suppose that each of these fields is paired with a field adjoint in a different interaction (different
for each one), or in the initial or final state. Thus, the first of these field adjoints can be paired with any of
the M identical fields in Hi (x); the second with any of the remaining M 1 identical fields, the third with the
230 CHAPTER 9. THE FEYNMAN RULES

remaining M 2 identical fields, and so on. It gives an extra factor M !. It is customary to compensate it by
defining coupling constants gi so that an explicit factor 1/M ! appears in any Hi (x) containing M identical
fields or field adjoints. For instance, the interaction of M thorder in a scalar single field (x) would be
written gM /M !, because of the presence of the M identical fields (x). Further, it is customary to display
a 1/M ! factor when the interaction involves a sum of M factors of fields from the same symmetry multiplet,
or when for any reason the coupling coefficient is totally symmetric or antisymmetric under permutations
of M boson or fermion fields.
We have described in the previous item the situation in which each of the M identical fields in Hi (x), is
paired with a field adjoint in a different interaction. When this is not the case, the cancellation of the M !
is not complete. Let us take the opposite scenario with a Feyman diagram in which the M identical fields in
one interaction Hi (x) are paired with M corresponding field adjoints in a single other interaction Hj (y).
In this case, we find only M ! different pairings cancelling only one of the two factors of 1/M ! in the two
different interactions. This is due to the fact that it makes no difference which of the field adjoints is called
the first, second, third, and so on. In this case, we would require to insert an extra factor M ! by hand
into the contribution of that given Feynman diagram. For instance, Fig. 9.2 shows a diagram in which
three identical fields in Hi (x) are paired with three corresponding field adjoints in a single other interaction
Hj (y). Thus an assignment of the form gi gi /3!, would give a contribution of the form gi gj / (3!)2 due to
the presence of two vertices. However, the multiplicity is only 3! (the 3! permutations of the internal lines
in the diagram).

Figure 9.2: This particular diagram requires an extra combinatoric factor in the Smatrix. For an interaction
including say, three factors of some field (plus other fields) it is customary to include a factor 1/3! in the interaction
Hamiltonian, in order to cancel factors coming from sums over ways of pairing these fields with their adjoints in
other interactions. However, in this diagram there are two of those factors of 1/3! and only 3! different pairings.
Hence, we are left with an extra factor of 1/3!.

Another important scenario arises when some of the permutations of the vertices have no effect on the Feynman
diagram. In that case other combinatoric factors appear because the cancellation of the factor 1/N ! in the series
(9.1) is not complete when relabelling the vertices does not give a new diagram. An example is given by Fig. 9.3
which is a Feynman diagram of N th order in H with the shape of a ring with N corners. It is clear that any
cyclic change of the labels of the vertices give the same diagram. Thus, there are only (N 1)! different diagrams
since a permutation of labels that moves each label to the next vertex around the ring provides the same diagram.
Therefore, such a diagram is accompanied by the factor
(N 1)! 1
= (9.22)
N! N
Physically, this situation is usual in the calculation of vacuum-to-vacuum Smatrix elements in a theory with a
quadratic interaction
H = k Mkm m (9.23)
9.4. CALCULATION OF THE SMATRIX FROM THE FACTORS AND DIAGRAMS 231

Figure 9.3: This is an eighth-order graph describing a vacuum-to-vacuum amplitude with particles interacting with
only one external field represented by wiggly lines. For a given set of labels for the vertices a cyclic change of
such labels leads to the same diagram, applying other cyclic relabelling provides the same diagram again and so
on until we perform seven succesive cyclic permutations (the eighth cycle would lead to the same labelling as the
beginning). Therefore, there are 7! of these eight-order diagrams differing only by relabelling the vertices, but we
do not count as different those labellings that simply rotates the ring. Consequently, the 1/8! factor appearing in
the Dyson series (9.1) is not completely cancelled, leaving us with an extra factor 1/8.

where M could depend on external fields.

In theories with fermion fields, the use of Eqs. (9.5), (9.6) and (9.7) to move annihilation operators to the
right and creation operators to the left, introduces minus signs into the contributions of several pairings. In
particular, we get a minus sign when the permutation of the operators in Eq. (9.1) required to put all paired
operators in the appropriate ordered (with annihilation operators just to the left of the paired creation
operators) involves an odd number of interchanges of fermion operators. We see that by taking into account
that to calculate the contribution of a given pairing, we first permute all operators in (9.1) in such a way
that each annihilation operator is just to the left of the creation operator with which it is paired, ignoring
all commmutators and anticommutators of unpaired operators, and replace each product of paired operators
with their commutators or anticommutators. Therefore, there will arise a minus sign in the relative sign of
the two terms in Eq. (9.18) for the fermion propagator. Given the permutation that puts the annihilation
part + (x) of a field in H (x) just to the left of the creation part + (y) of a field adjoint in H (y), the
associated permutation that puts the annihilation part (y) of the field adjoint just to the left of the
creation part (x) of the field, involves an additional interchange of fermion operators, giving a minus sign
in the second term of Eq. (9.18) for fermions.

Some additional minus signs can appear in the contribution of whole Feynman diagrams. To see it, let us
take a theory in which the sole interaction of fermions has the form
X
H (x) = gkmn k (x) m (x) n (x) (9.24)
kmn
232 CHAPTER 9. THE FEYNMAN RULES

Figure 9.4: The connected second-order diagrams for fermion-fermion scattering for a type of interaction described
by (9.24). The fields and field adjoints are associated with the vertices y (left) and x (right). Straight lines represent
fermions and dotted lines represent neutral bosons. There is a relative minus sign when adding the contributions
of both diagrams, which arises from an extra interchange of fermion operators in the pairings associated with the
second diagram.

where gkmn are general coupling constants, k (x) are a set of complex fermion fields, and n (x) are a set of
real bosonic (not necessarily scalar) fields. Let us study the process of fermion-fermion scattering 12 1 2 ,
up to second order in H. The fermion operators in the second order term of Eq. (9.1) appear in the order3
 
a 2 a 1 (x) (x) (y) (y) a (1) a (2) (9.25)
there are two connecting diagrams associated with the two pairings4
h  ih  ih ih i
a 2 (x) a 1 (y) (y) a (1) (x) a (2) (9.26)
h  ih  ih ih i
a 1 (x) a 2 (y) (y) a (1) (x) a (2) (9.27)

these two pairings are displayed in Fig. 9.4. In order to go from (9.25) to (9.26) we require an even
permutation of fermionic operators. For example, by moving (x) past three operators to the right and
then move a (1 ) past one operator to the right. Consequently, there is no any minus sign in going from
(9.25) to (9.26) i.e. in the contribution of the pairing (9.26). On the other hand, the only difference between
the pairings (9.26) and (9.27) is the interchange of two fermionic operators a (1 ) and a (2 ). In turn, this
relative minus sign is what we require to be compatible with Fermi statistics, since it makes the scattering
amplitude antisymmetric under the interchange of particles 1 and 2 (or 1 and 2). On the other hand, we
recall that the overall sign of the Smatrix is irrelevant in calculating transition rates (such a global sign
depends on sign conventions for the initial and final states). Therefore, what really matters is the relative
sign between the pairings (9.26) and (9.27) instead of the sign of each pairing.
Nevertheless, not all sign factors are so simply related to the antisymmetry of the final or initial states,
even in the lowest order of perturbation theory. To see it, we take as an example the fermion-antifermion
scattering 1 2c 1 2c , to second order in the same interaction (9.24). The fermionic operators in the
second-order term in Eq. (9.1) appear in the order
 
a 2c a 1 (x) (x) (y) (y) a (1) a (2c ) (9.28)
and again we have two Feynman diagrams associated with the pairings
  h  ih ih i
a 2c (x) a 1 (x) (y) a (1) (y) a (2c ) (9.29)
  h  ih ih i
a 2c (x) a 1 (y) (y) a (1) (x) a (2c ) (9.30)
3
Note that the order of the fields and field adjoints here are well defined because of the form in which we proposed our Hamiltonian
density Eq. (9.24). Otherwise, the order that is well defined within a Hamiltonian density is the one concerning the creation and
annihilation fields as we can see in the expansion (4.16), page 139.  
4
Note for instance that pairings of the form [a (2 ) (x)] or of the form (x) a (2) gives no contribution as discussed in section
9.1, page 223.
9.4. CALCULATION OF THE SMATRIX FROM THE FACTORS AND DIAGRAMS 233

Figure 9.5: The connected second-order diagrams for fermion-antifermion scattering for a type of interaction
described by (9.24). Straight lines pointing upward (downward) represent fermions (antifermions), while dotted
lines represent neutral bosons. There is a relative minus sign when adding the contributions of both diagrams,
which arises from an extra interchange of fermion operators in the pairings associated with the second diagram.

as displayed by Fig. 9.5, where Fig. 9.5(a) corresponds to the pairings 9.30, and Fig. 9.5(b) corresponds
to the pairings in (9.29). In order to go from (9.28) to Eq. (9.29) we require an even permutation of
fermionic operators. For example, we can move (x) past two operators to the left and move (y) past
two operators to the right. Hence, there is not extra minus sign in the contribution of the pairing (9.29).
But to go from (9.28) to (9.30) requires an odd permutation [since in passing from (9.29) to (9.30) we require
the interchange of (x) and (y)]. Hence, both pairings (9.29) and (9.30) must have opposite signs. This
sign has an indirect relation with the Fermi statistics5 .

Since the same field can destroy a particle and create an antiparticle, there is a crossing symmetry between
processes in which initial particles or antiparticles are exchanged with final antiparticles and particles. For
instance, the amplitudes for the process 1 2c 1 2c are related with the amplitudes of the crossed process
1 2 1 2; the two pairings (9.29) and (9.30) are associated with the two diagrams for this process, that
differ by an interchange of 1 and 2 (or 1 and 2). Hence, the antisymmetry of the scattering amplitude under
the interchange of initial (or final) particles requires a minus sign in the relative contribution of these two
pairings. Nevertheless the crossing symmetry in general requires an analytic continuation in kinematic
variables, making it very difficult to use in practice.

When we consider higher order contributions, additional signs appear. For example, in a theory of the
type described by (9.24), the fermion lines form either chains of lines that pass through the diagram with
arbitrary numbers of interactions with boson fields as displayed by Fig. 9.6, or fermionic loops as shown in
Fig. 9.7. Let us see the effect of adding a fermionic loop with M corners to the Feynman diagram for any
process. The diagram is associated with the following pairing of fermionic operators
    
(x1 ) (x2 ) (x2 ) (x3 ) (xM ) (x1 ) (9.31)

that is a given vertx (say x1 ) starts a sequence x1 x2 . . . xM x1 that ends into itself. But in the
Dyson series Eq. (9.1) these operators appear in the order

(x1 ) (x1 ) (x2 ) (x2 ) (xM ) (xM ) (9.32)


5
Note however that the initial and final states in both diagrams of Fig. 9.5 are in the same position. We cannot see a difference
by checking only the external lines. The difference has to do with the form in which the internal bosonic line connects the initial and
final states.
234 CHAPTER 9. THE FEYNMAN RULES

Figure 9.6: The connected second-order diagrams for boson-fermion scattering for a type of interaction described
by (9.24). Once again, straight lines represent fermions and dashed lines represent neutral bosons.

and going from (9.32) to (9.31) requires an odd number of permutations of fermionic operators (for instance,
we can move (x1 ) to the right past 2M 1 operators). Therefore, the contribution of each such fermionic
loop is accompanied by a minus sign.

The Feynman rules described above provides the full Smatrix, including processes in which several clusters of
particles are in space-time regions widely separated from each other. According with the discussion in chapter 3,
we can exclude the contributions of such disconnected clusters by taking only the connected Feynman diagrams.
It excludes in particular lines passing clean through the diagram without interacting, that are associated with the
factors (9.17).
We shall illustrate the use of the Feynman rules by calculating the low-order contributions to the Smatrix
for particle scattering using two different theories

9.5 A fermion-boson theory


We shall consider first the theory described by Eq. (9.24) involving interactions of fermions and self-charge
conjugate bosons. Each vertex in this theory contains three lines, two fermion lines and one boson line. Based on
the Feynman rules we shall construct the Smatrix associated with fermion-boson and fermion-fermion scattering.

9.5.1 Fermion-boson scattering


The lowest order connected diagrams for fermion boson scattering are the ones shown in Fig. 9.6. Using the rules
described in Fig. 9.1 and Secs. 9.2, 9.3, 9.4, we obtain the associated Smatrix element. We show in Fig. 9.8
all labels associated with external lines and vertices. Since all are particles, the fermion arrows point upwards (it
is not usual to paint arrows for scalars). Now following the rules illustrated in Fig. 9.1 we begin by adding the
factors associated with each line and vertex. We start with diagram 9.8(a).

1. For the initial (fermion) particle p1 1 n1 with a line pointing upwards, with index n for its field component
, and ending into the vertex (x, m) we put a term

eip1 x un (p1 , 1 , n1 )
(2)3/2

2. For the initial (boson) particle p2 2 n2 with a line pointing upwards, with index k for its field component,
9.5. A FERMION-BOSON THEORY 235

Figure 9.7: The lowest order connected diagram for boson-boson scattering with an interaction density of the type
(9.24). These fermion loop diagrams give an extra minus sign coming from permutations of the paired fermion
fields.

and ending into the vertex (x, m) we put a term

eip2 x uk (p2 , 2 , n2 )
(2)3/2

3. Now for the vertex (x, m) we add a factor


(igmnk )

where m is the index of the vertex, n is the index component of the external incoming fermion and k the
index of component of the field associated with the external incoming boson. Note that the order of the
labels is consistent with the one given in Eq. (9.24) in which the last index correspond to the index of the
boson.

4. For the internal line connecting the vertices (x, m) and (y, m ) we put a propagator

im m (y x)

5. Now for the vertex (y, m ) we put a factor


ign m k

6. For the final (fermion) state p1 1 n1 with a line starting at the vertex (y, m ) going upwards, with component
index n we put a factor

eip1 y un (p1 , 1 , n1 )
(2)3/2
236 CHAPTER 9. THE FEYNMAN RULES

Figure 9.8: Labels to apply the Feynman rules for the lowest order diagrams associated with the fermion-boson
scattering in Fig. 9.6.

7. For the final (boson) state p2 2 n2 with a line starting at the vertex (y, m ) going upwards, with component
index k we put a factor

eip2 y uk (p2 , 2 , n2 )
(2)3/2
8. We put all this stuff together
"
#"
#
eip2 y uk (p2 , 2 , n2 ) eip1 y un (p1 , 1 , n1 )
(ign m k ) [im m (y x)]
(2)3/2 (2)3/2
" #" #
eip2 x uk (p2 , 2 , n2 ) eip1 x un (p1 , 1 , n1 )
(igmnk )
(2)3/2 (2)3/2

9. We then integrate the product of these factors over the coordinates x and y of each vertex and sum over the
components of the fields
" #" #
X XZ Z
e ip2 y u (p , , n )
eip1 y u (p , , n )

2 2 2 1 1 1
D1 d4 x d4 y k
3/2
n
3/2
(ign m k ) [im m (y x)]
k n m knm (2) (2)
" #" #
eip2 x uk (p2 , 2 , n2 ) eip1 x un (p1 , 1 , n1 )
(igmnk )
(2)3/2 (2)3/2

X X Z Z
6 2 
D1 (2) (i) gn m k gmnk un p1 , 1 , n1 un (p1 , 1 , n1 ) 4
d x d4 y [im m (y x)]
k n m knm
n  o
ip1 y ip1 x
e e eip2 y uk p2 , 2 , n2 eip2 x uk (p2 , 2 , n2 ) (9.33)

A similar procedure should be done for the second diagram to obtain its associated contribution D2 . For the
diagram 2 [see Fig. 9.8(b)], the initial particles p1 1 n1 (fermion) and p2 2 n2 (boson) with field components n and
k respectively and connected to the vertices (x, m) and (y, m ) respectively, yield a contribution
" #" #
eip2 y uk (p2 , 2 , n2 ) eip1 x un (p1 , 1 , n1 )
(2)3/2 (2)3/2
9.5. A FERMION-BOSON THEORY 237

the vertices (x, m) and (y, m ) and the propagator (internal line) yield

(igmnk ) , (ign m k ) , [im m (y x)]

and the external particles p1 1 n1 (fermion) and p2 2 n2 (boson) with field components n and k respectively and
connected to the vertices (y, m ) and (x, m) respectively, yield a contribution
"
#"
#
eip1 y un (p1 , 1 , n1 ) eip2 x uk (p2 , 2 , n2 )
(2)3/2 (2)3/2

making the product


" #" #
eip2 y uk (p2 , 2 , n2 ) eip1 x un (p1 , 1 , n1 )
(igmnk ) (ign m k )
(2)3/2 (2)3/2
"
#"
#
eip1 y un (p1 , 1 , n1 ) eip2 x uk (p2 , 2 , n2 )
[im m (y x)]
(2)3/2 (2)3/2

integrating over the vertices and summing over field components


" #" #
X XZ Z
e ip2 y u (p , , n )
k 2 2 2 e ip1 x u (p , , n )
n 1 1 1
D2 d4 x d4 y 3/2 3/2
(igmnk ) (ign m k )
k n m knm (2) (2)
"
#"
#
eip1 y un (p1 , 1 , n1 ) eip2 x uk (p2 , 2 , n2 )
[im m (y x)]
(2)3/2 (2)3/2

X X 
D2 (2)6 (i)2 gn m k gmnk un p1 , 1 , n1 un (p1 , 1 , n1 )

Z Zk n m knm n o

d4 x d4 y [im m (y x)] eip1 y eip1 x eip2 x uk p2 , 2 , n2 eip2 y uk (p2 , 2 , n2 )

we can interchange k k within the sum since they are dummy indices
X X Z Z
6 2 
D2 (2) (i) gn m k gmnk un p1 , 1 , n1 un (p1 , 1 , n1 ) 4
d x d4 y [im m (y x)]
k n m knm
n  o
ip1 y
e eip1 x eip2 x uk p2 , 2 , n2 eip2 y uk (p2 , 2 , n2 ) (9.34)

Then, the Smatrix element associated with fermion-boson scattering at lowest order, is the sum D1 + D2 (in
this case there is no a minus sign coming from odd interchanges of fermion operators). Thus adding Eqs. (9.33)
and (9.34) we obtain

X X 
Sp1 1 n1 p2 2 n2 ;p1 1 n1 p2 2 n2 = (2)6 (i)2 gn m k gmnk un p1 , 1 , n1 un (p1 , 1 , n1 )
k n m knm
Z Z h i

d4 x d4 y im m (y x) eip1 y eip1 x
n 
eip2 y uk p2 , 2 , n2 eip2 x uk (p2 , 2 , n2 )
 o
+eip2 x uk p2 , 2 , n2 eip2 y uk (p2 , 2 , n2 ) (9.35)

where according with Fig. 9.8, the labels 1 and 2 are used for fermions and bosons respectively.
238 CHAPTER 9. THE FEYNMAN RULES

9.5.2 Fermion fermion scattering


For fermion-fermion scattering there are also two second-order diagrams displayed in Fig. 9.4. The associated
Smatrix yields (Homework!! B1)
X X
Sp1 1 n1 p2 2 n2 ;p1 1 n1 p2 2 n2 = (2)6 (i)2 gm mk gn nk
k n m knm
 
um p2 , 2 , n2 un p1 , 1 , n1 um (p2 , 2 , n2 ) un (p1 , 1 , n1 )
Z Z

4
d x d4 y eip2 x eip1 y eip2 x eip1 y [ik k (x y) ]
 
1 2 (9.36)

where the last term indicates the interchange of the two fermions 1 and 2 (or the fermions 1 and 2) in the previous
terms accompanied by a minus sign.

9.5.3 Boson-boson scattering


There are no second-order graphs for boson-boson scattering in this theory. It is because second order graphs for
a boson-boson scattering requires vertices of three lines with two bosons and one fermion. But Eq. (9.24) shows
that only vertices with two fermions and one boson appears in this theory. The lowest order terms for this process
are of fourth order, such as the ones displayed in Fig. 9.7. Note that each vertex in Fig. 9.7, contains three lines
with two fermions and one boson as demanded by the theory in Eq. (9.24).

9.6 A boson-boson theory

Figure 9.9: The connected second-order diagrams for boson-boson scattering with an interaction of the type (9.37).

In the theory described by the interaction (9.24), the three fields are all different6 . Thus, it is useful to consider
an example with a trilinear interaction involving three identical fields or at least entering into the interaction in
a symmetrical way. In that case, more combinatoric factors are introduced as discussed in Sec. 9.4.
We shall consider an interaction density that is a sum of terms that are thrilinear in a set of real bosonic fields
k (x)
1 X
H (x) = gkmn k (x) m (x) n (x) (9.37)
3!
kmn
where gkmn are coupling constants that are real and totally symmetric. Note that we have included a correction
factor 1/3! to the coupling constant according with the discussion in section 9.4. We shall characterize a scattering
6
The fermions fields are a field and an adjoint field. They are different even if they describe the same species of fermions.
9.6. A BOSON-BOSON THEORY 239

process 1 2 1 2 , up to second order in this interaction. Each of the two vertices must have two of the four
external lines attached to it. A priori, the other possibility is that one of the external lines is attached to one of
the vertices while the three remaining external lines are attached to the other vertex. Nevertheless, in the latter
case we would not have remaining lines to connect the two vertices, so that it would be a disconnected diagram
i.e. a disconnected contribution. Coming back to our connected graph, the additional line required at each vertex
should be used to connect the two vertices between them. For the vertex attach to line 1, the same vertex could
be attach to either line 2 or line 1 or line 2 . Therefore, three graphs of that type arise, as can be seen in Fig.
9.9. From the Feynman rules, the contribution of these three diagrams to the Smatrix reads (homework!! B2)
X Z Z
2 6
S Sp1 1 n1 p2 2 n2 ;p1 1 n1 p2 2 n2 = (i) (2)
gnn n gmm m
d x d4 y [in m (x, y)]
4

nn n mm m
h  
ip1 x
un p1 , 1 , n1 e
un p2 , 2 , n2 eip2 x um (p1 , 1 , n1 ) eip1 y um (p2 , 2 , n2 ) eip2 y

 
+un p1 , 1 , n1 eip1 x un (p1 , 1 , n1 ) eip1 x um p2 , 2 , n2 eip2 y um (p2 , 2 , n2 ) eip2 y
 
i
+un p2 , 2 , n2 eip2 x un (p1 , 1 , n1 ) eip1 x um p1 , 1 , n1 eip1 y um (p2 , 2 , n2 ) eip2 y
(9.38)
we can simplify this model by assuming that the boson particles in this theory are spinless (scalars) of a single
species. In that case all indices of field components dissapear, and the interaction (9.37) becomes
3 (x)
H (x) = g (9.39)
3!
substituting (9.39) in (9.38) the Smatrix becomes
Z Z
3 6 2
S Sp1 p2 ;p1 p2 = (i) (2) g d x d4 y [F (x, y)] 4

h  
u p1 eip1 x u p2 eip2 x u (p1 ) eip1 y u (p2 ) eip2 y
 
+u p1 eip1 x u (p1 ) eip1 x u p2 eip2 y u (p2 ) eip2 y
 
i
+u p2 eip2 x u (p1 ) eip1 x u p1 eip1 y u (p2 ) eip2 y (9.40)

where we have taken into account that only can takes one value and there is only one species of particle. Hence
the Smatrix element only depend on the momenta. Now, for a scalar neutral field, we have [see Eq. (5.4) Page
151]
1
u (p) = v (p) = (9.41)
2E
substituting (9.41) in (9.40) we find

Z Z "
3 6 2 4 eip1 x eip2 x eip1 y eip2 y
4
S Sp1 p2 ;p1 p2 = (i) (2) g d x d y [F (x, y)] p p
2E1 2E2 2E1 2E2

#
eip1 x eip1 x eip2 y eip2 y eip2 x eip1 x eip1 y eip2 y
+p p +p p
2E1 2E1 2E2 2E2 2E2 2E1 2E1 2E2
and we obtain finally
Z Z
ig2 4
Sp1 p2 ;p1 p2 = 6
p
d x d4 y F (x y)
(2) 16E1 E2 E1 E2
h i
i(p1 +p2 )x i(p1 +p2 )y
e e + ei(p1 p1 )x ei(p2 p2 )y + ei(p1 p2 )x ei(p2 p1 )y
240 CHAPTER 9. THE FEYNMAN RULES

Figure 9.10: Connected third-order diagrams for boson-boson scattering would have at least one vertex attached to
four lines, but such kind of vertices are not included in the theory described by (9.37).

where F (x y) is the scalar field propagator that we shall calculate soon. There are no terms of third order
or of any odd order in H (x). To see it, we observe that third order connected diagrams would have at least one
vertex connected to four lines (see Fig. 9.10), but the theory described by (9.37) only contains vertices attached
to three lines. Diagrams of fourth order are similar to the one shown in Fig. 9.7, page 235, except that all lines
involved are boson type lines.

9.7 Calculation of the propagator



As we explained in Sec. 9.2, in the pairing of a field k (x) with a field adjoint m (y), a factor of the form (9.18)
that is called the propagator arises. Using Eqs. (9.20) and (9.21) and the commutation and anticommutation
relations for annihilation and creation operators we have

"Z Z #
h i 1 X X  
3 ip y
k+ (x) , +
m (y) = 3
d p uk (p, , n) e ipx
a (p, , n) , d p um
p , ,n
e a
p , ,n
(2)3
Z Z h
1 XX  ip y
i
= d3 p d3 p uk (p, , n) um p , , n
e eipx a (p, , n) , a p , , n
(2)3
Z Z XX
1  
= d3 p d3 p uk (p, , n) um p , , n eip y eipx p p nn
(2)3
h i Z X
1
k+ (x) , m
+
(y) = d3 p uk (p, , n) um (p, , n) eip(xy) (9.42)
(2)3

similarly Z
h i 1 X

m (y) , k (x) = d3 p
vm (p, , n) vk (p, , n) eip(yx) (9.43)
(2)3

substituting (9.42) and (9.43) in Eq. (9.18), we obtain


Z X
3
ikm (x, y) = (x y) (2) d3 p uk (pn) um (pn) eip(xy)

Z X
(y x) (2)3 d3 p
vm (pn) vk (pn) eip(yx) (9.44)

when we derived the commutation and anticommutation relations for fields and field adjoints, we also found
expressions for the sum over spins of bilinear forms of the coefficients. For vector fields they were given by Eqs.
9.7. CALCULATION OF THE PROPAGATOR 241

(6.64, 6.69) and (6.75) page 173, while for Dirac fields they were given by Eqs. (7.102, 7.103) and Eqs. (7.121,
7.122) page 200. Such bilinear forms posses the following general structure7
 p 
X Pkm p, p2 + m2n
uk (pn) um (pn) = p (9.45)
2 p2 + m2
n
 p 
X Pkm p, p2 + m2n

vk (pn) vm (pn) = p (9.46)
2 p2 + m2n
where Pkm (p, ) is a polinomial in p and . In Eqs. (9.44, 9.45, 9.46) the signs refer to bosonic and fermionic
fields respectively. If k (x) (x) and m (y) (y) are scalar fields for a particle of spin zero, Eqs. (5.4)
page 151, shows that
X 1 1
uk (pn) um (pn) = u (p) u (p) = 0 = p (9.47)

2p 2 p2 + m 2
n
comparing Eqs. (9.45, 9.47) the polynomial P (p, ) for scalar fields becomes

P p, p0 = P (p) = 1 (9.48)
if k (x) (x) and m (y) (y) are Dirac fields for a particle of spin 1/2, Eqs. (7.102, 7.103) and Eqs.
(7.121, 7.122) page 200, can be combined to obtain
X 1
Nkk (p) uk (p, ) uk (p, ) = 0 [(ip + m) ]kk (9.49)

2p
X 1
Mkk (p) vk (p, ) vk (p, ) = 0 [(ip m) ]kk (9.50)

2p

comparing Eqs. (9.45, 9.46) with Eqs. (9.49, 9.50) we find that for Dirac fields of spin 1/2, the polynomial reads
P (p) = [(i p + m) ] (9.51)

note that the matrix appears here because we are considering the pairing of (x) with (y) instead of a
pairing of (x) with (y) (y) .
Finally, if k (x) and m (y) are vector fields V (x) and V (y) for a particle of spin one, the combination of
Eqs. (6.64, 6.69) and (6.75) lead to
X Xh 1/2 i   p p
(p) e (p, ) e (p, ) = 2p0 u (p, ) 2p0 u (p, ) = g +

m2
X  
1 p p
u (p, ) u (p, ) = 0 g + (9.52)

2p m2

once again comparison of Eqs. (9.45) and (9.52) shows that the polynomial for spin one vector fields is given by
p p
P (p) = g + (9.53)
m2
Now, substituting Eqs. (9.45) and (9.46) in Eq. (9.44) we have
 p 
Z Pkm p, p2 + m2n
ikm (x, y) = (x y) (2)3 d3 p eip(xy)
2p0
 p 
Z Pkm p, p2 + m2n
+ (y x) (2)3 d3 p eip(yx) (9.54)
2p0
7
For scalar fields we can obtain similar expressions from Eqs. (5.4) page 151.
242 CHAPTER 9. THE FEYNMAN RULES

now taking into account that  n



i eipx = pn eipx
x
we can replace a polynomial in p by
 
0
 ip(xy)
Pkm p, p e = Pkm i eip(xy) (9.55)
x

substituting (9.55) in (9.54) the propagator becomes


Z
 ip(xy) Z
 ip(yx)
(x y) Pkm i x
3 e (y x) 3 Pkm i x e
ikm (x, y) = d p + d p
(2)3 2p 0
(2) 3 2p 0
 Z 3   Z
d p eip(xy) d3 p eip(yx)
ikm (x, y) = (x y) Pkm i 3 + (y x) Pkm i
x 2p0 (2) x 2p0 (2)3

then the propagator takes the form

 

ikm (x, y) = (x y) Pkm i + (x y)
x
 

+ (y x) Pkm i + (y x) (9.56)
x

where + (x) is the function defined in Eq. (5.9) page 152


Z p
1 d3 p ipx
+ (x) e ; p0 + p2 + m2 (9.57)
(2)3 2p0

Now we have to extend the definition of the polynomial P (p). The definition of p P (p) given by Eqs. (9.45) and
0
(9.46) are only valid for four-momenta on the mass shell. That is, when p = p2 + m2 . We require such an
extension because as we shall see later,
2n some 0internal
2n+1 lines will carry momenta that
n are off the mass
n shell.
Since any power of the form p0 or p can be written as p2 + m2 or p0 p2 + m2 ; we see that
any polynomial function of such four-momentum can be taken as linear in p0 . Consequently, we can define a
generalized polynomial P (L) (p) as follows
p
P (L) (p) = P (p) for p0 = p2 + m2 (9.58)
(L) (0) 0 (1)
P (q) = P (q) + q P (q) for general q (9.59)

with P (0,1) being polynomials that depend only on q. Now, using the properties

  
0
x0 y 0 = 0 y 0 x0 = x0 y 0
x x
we are able to move the derivative operators to the left of the functions in Eq. (9.56) to obtain

open
????
 
(L)
Pkm i = P (0) (i) iP (1) (i)
x x0
9.7. CALCULATION OF THE PROPAGATOR 243

 
(L)  (1)
E = Pkm i F (x y) + x0 y 0 Pkm (i) [+ (x y) + (y x)]
x
 
 (1)
= P (i) iP (i) 0 {i (x y) + (x y) + i (y x) + (y x)} + x0 y 0 Pkm (i) [+ (x
(0) (1)
x
   
(0) (1) (0) (1)
= P (i) iP (i) 0 {i (x y) + (x y)} + P (i) iP (i) 0 {i (y x) + (y x
x x
0 0
 (1)
+ x y Pkm (i) [+ (x y) + (y x)]
h i h i
= P (0) (i) {i (x y) + (x y)} iP (1) (i) 0 {i (x y) + (x y)} + P (0) (i) {i (y x) + (
x
 (1)
iP (i) 0 {i (y x) + (y x)} + x y 0 Pkm (i) [+ (x y) + (y x)]
(1) 0
x

h i  
(0) (1) (x y) + (x y)
E = P (i) {i (x y) + (x y)} iP (i) i + (x y) + i (x y)
x0 x0
h i  
(y x) + (y x)
+ P (0) (i) {i (y x) + (y x)} iP (1) (i) i + (y x) + i (y x)
x0 x0
 (1)
+ x0 y 0 Pkm (i) [+ (x y) + (y x)]

h i  
 + (x y)
E = P (0) (i) {i (x y) + (x y)} iP (1) (i) i x0 y 0 + (x y) + i (x y)
x0
h i  
(0) (1) 0 0
 + (y x)
+ P (i) {i (y x) + (y x)} iP (i) i x y + (y x) + i (y x)
x0
 (1)
+ x0 y 0 Pkm (i) [+ (x y) + (y x)]

h i 
(0) (1)
 0 0
 (1) + (x
E = P (i) {i (x y) + (x y)} iP (i) i x y + (x y) iP (i) i (x y)
x0
h i 
  +
+ P (0) (i) {i (y x) + (y x)} iP (1) (i) i x0 y 0 + (y x) iP (1) (i) i (y x)
 (1)  (1)
+ x0 y 0 Pkm (i) + (x y) x0 y 0 Pkm (i) + (y x)

h i 
(0) 0 0
 (1) (1) + (x y)
E = P (i) {i (x y) + (x y)} + x y P (i) + (x y) iP (i) i (x y)
x0
h i 
 + (y
+ P (0) (i) {i (y x) + (y x)} x0 y 0 P (1) (i) + (y x) iP (1) (i) i (y x)
x0
 (1)  (1)
+ x0 y 0 Pkm (i) + (x y) x0 y 0 Pkm (i) + (y x)

h i  
(0) (1) + (x y)
E = P (i) {i (x y) + (x y)} iP (i) i (x y)
x0
h i  
(0) (1) + (y x)
+ P (i) {i (y x) + (y x)} iP (i) i (y x)
x0
 (1) 
+ x0 y 0 Pkm (i) + (x y) + x0 y 0 P (1) (i) + (x y)
 (1) 
x0 y 0 Pkm (i) + (y x) x0 y 0 P (1) (i) + (y x)
244 CHAPTER 9. THE FEYNMAN RULES

????

close
 
(L)  (1)
km (x, y) = Pkm i F (x y) + x0 y 0 Pkm (i) [+ (x y) + (y x)] (9.60)
x

where F is defined by

iF (x) (x) + (x) + (x) + (x) (9.61)

and it is called the Feynman propagator. Now, we observe that for x0 = 0, the function + (x) is even with
respect to x. We see it by noting that the change x x in Eq. (9.57) can be compensated by a change p p
in the integration variable. Thus, we can omit the second term in Eq. (9.60) and write

 
(L)
km (x, y) = Pkm i F (x y) (9.62)
x

It is convenient to express the Feynman propagator as a Fourier integral. To do it, we write the step function
in its Fourier representation
Z
1 exp (ist)
(t) = ds (9.63)
2i s + i

the validity of Eq. (9.63) can be shown as follows: if t > 0 the numerator yields

exp (ist) = ei(Re s+iIm s)t = eit Re s t Im s


e

we can close the contour of integration with a large clockwise semi-circle in the lower half-plane (such that Im s < 0
and the numerator does not diverge), from which the integral has a contribution of 2i from the pole at s = i.
Now, if t < 0 we can close the contour by a large counter-clockwise semi-circle in the upper half-plane, within
which the integrand is analytic, so that the integral vanishes.
Now to express the Feynman propagator as a Fourier integral, we combine Eq. (9.63) with Eq. (9.61) as well
as the Fourier integral (9.57) for + (x). Doing it we obtain

iF (x) (x) + (x) + (x) + (x)


"Z  # Z 3  "Z  # Z 3 
exp isx0 exp isx0
1 1 d p ipx 1 1 d p ipx
= ds e + ds e
2i s + i (2)3 2p0 2i s + i (2)3 2p0

By redefining the variables q p and q 0 = p0 + s, in the first term of Eq. (9.61) we have8

8
Note that by assuming that the spatial components of q coincides with the spatial components of p, while the temporal components
do not coincide, means that q is off the mass shell.
9.7. CALCULATION OF THE PROPAGATOR 245

"Z  # Z 3 
1 exp isx0

1 d p
(x) + (x) = ds exp [ip x]
2i s + i (2)3 2p0
"Z    # Z 3 
exp i q 0 p0 x0  
1 0 1 d p 0 0
= dq exp ip x ip x
2i (q 0 p0 ) + i (2)3 2p0
h  p  i " #
Z exp i q 0 p2 + m2 x0 Z 3p h p i
1 1 d
=  p  dq 0 3
p exp ip x i p2 + m2 x0
2i 0 2
q p + m + i 2 (2) 2 p2 + m2

h  p  i " #
Z exp i q 0 q2 + m2 x0 Z 3q h p i
1 0 1 d 0
=  p  dq p exp iq x i q2 + m2 x
2i q 0 q2 + m2 + i (2)3 2 q2 + m2
h p i h p i
Z 3 Z exp iq 0 x0 + i q2 + m2 x0 exp iq x i q2 + m2 x0
1 1 d q
= 3
p dq 0  p 
2i 2 (2) q2 + m2 q 0 q2 + m2 + i
 
Z Z 0 x0
1 exp iq x iq 1
(x) + (x) = d3 q dq 0 p  p 
2i 2 (2)3 q2 + m2 q 0 q2 + m2 + i

we can calculate (x) + (x) with a similar procedure, so that the Feynman propagator becomes
-

Z Z  
1 3
iq x iq 0 x0
0 exp
iF (x) = d q dq p
2i 2 (2)3 q2 + m2

1 1
 p + p  (9.64)
q0 2 2
q + m + i 0 2 2
q q + m + i

the denominators in parenthesis give


1 1
D  p + p 
q0 q2 + m2 + i q q2 + m2 + i
0

1 1
= h p 
i h p  i
q 0 q2 + m2 + i q 0 + q2 + m2 i
h p  i h p  i
q 0 + q2 + m2 i q 0 q2 + m2 + i
=  p 2     
2
p p
0
(q ) 2
q +m 2 0 2 2 0
i q q + m + i q + q + m + O ( ) 2 2 2

p p
2 q2 + m2 2i 2 q2 + m2 2i
= h p i = h p i
(q 0 )2 (q2 + m2 ) + 2i q2 + m2 (q 0 )2 q2 m2 + 2i q2 + m2
p
2 q2 + m2
D = h p i (9.65)
q2 (q 0 )2 + m2 2i q2 + m2

replacing (9.65) in (9.64) we have


246 CHAPTER 9. THE FEYNMAN RULES

Z Z
 
0 x0
p

1 exp iq x iq 2 q2 + m2
iF (x) = d3 q dq 0 p h  p i (9.66)
2i 2 (2)3 q2 + m2 q2 (q 0 )2 + m2 2i q2 + m2

Z Z  
0 x0
1 3 0 exp iq x iq 1
iF (x) = d q dq h  p i (9.67)
2i (2)3 q2 (q 0 )2 + m2 2i q2 + m2

Z Z  0 0

i exp i q x q x 1
iF (x) = d3 q dq 0 h i (9.68)
2 (2)3 q2 (q 0 )2 + m2 i

p
Note that in the denominator we have replaced 2 q2 + m2 with , because the only relevant thing about
that quantity is that it is a positive infinitesimal. In a four-dimensional notation F (x) can be written as
Z
1 exp (iq x) 2
F (x) = 4 d4 q 2 2
; q 2 q2 q 0 (9.69)
(2) q + m i
Moreover, it shows that F is a Greens function for the Klein-Gordon differential operator, to see it we apply
the Klein gordon operator on F (x)
Z  Z 
2
 1 4  m2 exp (iq x) 1 4 q 2 m2 exp (iq x)
 m F (x) = d q = d q
(2)4 q 2 + m2 i (2)4 q 2 + m2 i
Z
 1
 m2 F (x) = 4 d4 q exp (iq x)
(2)

 m2 F (x) = 4 (x) (9.70)
with boundary conditions determined by the infinitesimal quantity i inh the p
denominator,i since Eq. (9.61)
shows that F (x) for x + involves only positive frequency terms exp ix p + m , while for x0
0 0 2 2
h p i
involves only negative terms exp +ix0 p2 + m2 .
By substituting Eq. (9.69) in Eq. (9.62) the propagator becomes
    Z
(L) (L) 1 exp [iq (x y)]
km (x, y) = Pkm i F (x y) = Pkm i d4 q 2
x x (2)4 q + m2 i
Z (L)

1 4 Pkm i x exp [iq (x y)]
= d q
(2)4 q 2 + m2 i
Z (L)
1 4 P (q) eiq(xy)
km (x, y) = d q km2 (9.71)
(2)4 q + m2 i
The polynomial P (p) is Lorentz-covariant when p is on the mass shell. That is, when p2 = m2 . Nevertheless,
in Eq. (9.71) we are integrating over all q , such that p is off the mass shell. On the other hand, Eq. (9.59)
(L)
shows that the polynomial Pkm (q) for a general q is linear in q 0 , such a condition clearly does not respect Lorentz
covariance unless the polynomial is also linear in each spatial component q k . Hence we could alternatively define
the extension of P (p) to general momentum q , such that P (q) is Lorentz-covariant for general q as follows

Pkm (q) = Dkk () Dmm () Pk m (q)

where is a general Lorentz transformation on the Minkowski space, while D () is the associated representation
of the Lorentz group.
9.7. CALCULATION OF THE PROPAGATOR 247

In the case of scalar, Dirac and vector fields the covariant extensions of the polynomial are obtained by
substituting p by a general four-momentum q in Eqs. (9.48), (9.51) and (9.53). In the case of scalar and Dirac
fields they are already linear in q 0 such that P (L) (q) and P (q) are identical
(L)
Pkm (q) = Pkm (q) for scalar and Dirac fields

but for a vector field of a particle of spin one, we see that the 00 components of the covariant polynomial
q q
P (q) = g +
m2

are quadratic in q 0 . Thus, both polynomials are different


0 0 q 2 q2 m2
 0 0 q 2 + q2 + m2

(L) q q 0 q q 0
P (q) = g + 2
= g + 2 +
m  m m2
2 2
q + m 0 0
(L)
P (q) = P (q) + for vector bosons of spin one (9.72)
m2
the additional term is fixed by imposing two conditions (a) the cancellation of the quadratic term in P00 (q) and
(b) it must vanish when q is on the mass shell. Now, substituting (9.72) in Eq. (9.71) yields the propagator of a
vector field of spin one
Z
1 P (q) eiq(xy) 4 (x y) 0 0
(x, y) = d4 q + (9.73)
(2)4 q 2 + m2 i m2

where the first term is manifestly covariant. As for the second term, it is not covariant, but it is local, so that it
can be cancelled by a local non-covariant term added to the Hamiltonian density. If a vector field V (x) interacts
with other fields by means of a coupling of the form V (x) J (x) in H (x), the second term in Eq. (9.73) produces
an effective interaction described by
" #
0 0
1
iHef f (x) = [iJ (x)] [iJ (x)] i 2 (9.74)
2 m

where the factors i are those factors that appear for each vertex and propagator. The factor 1/2 appears
because there are two ways to pair other fields with Hef f (x), that differ in the interchange of J (x) and J (x).
Consequently, the effect of the non-covariant second term in Eq. (9.73) can be cancelled by adding to H (x) the
term
1  0 2
HN C (x) = Hef f (x) = 2
J (x) (9.75)
2m
further, the singularity that appears at equal-time commutators of vector fields at zero separation requires to use
a wider class of interactions than those of the scalar density. It is in that way that we can obtain a totally Lorentz
invariant Smatrix.
It worths pointing out that the previous phenomenon not only occurs for spins j 1. For example, in the
case of vector fields associated with j = 0 (see section 6.2, page 165) which consists of the derivative of a scalar
field (x), its pairing with a scalar (y) generates the polynomial P (p) on the mass shell given by

P (p) = ip (9.76)

while the pairing of (x) with (y) generates the polynomial

P, (p) = p p (9.77)
248 CHAPTER 9. THE FEYNMAN RULES

Once again, the polynomial associated to the off shell four-momentum q is obtained by replacing q for p
in Eqs. (9.76) and (9.77). Further, Eq. (9.76) shows that P (q) is already linear in q0 , from which P (q) and
(L)
P (q) are identical. By contrast, they differ in Eq. (9.77) such that
(L) 
P, (q) = q q q02 q2 m2 0 0

= P, (q) + q 2 + m2 0 0
from which the propagator becomes
Z
1 q q eiqx
, (x, y) = d4 q + 0 0 4 (x y)
(2)4 q 2 + m2 i
and we can also cancel the effect of the non-covariant second term by adding to the interaction another non-
covariant term given by
1 0 2
HN C (x) = J (x)
2

with J (x) being the current that couples to (x) in the covariant part of H (x).
It is in general always possible to cancel the effects of non-covariant parts of the propagator of a massive
(L)
particle by adding non-covariant local terms to the Hamiltonian density. It is because the numerator Pkm (q) in

the propagator must equal the covariant polynomial Pkm (q) when q is on the mass shell, so that the difference
(L)
between Pkm (q) and Pkm (q) must contain a factor q 2 + m2 . Such a factor cancels the denominator q 2 + m2 i
in the contribution of this difference to Eq. (9.71), hence Eq. (9.71) is always equal to a covariant term plus
a term proportional to 4 (x y) or its derivatives. The effect of the latter term can be cancelled by adding to
the interaction density a term which is quadratic in the currents to which the paired fields couple (or in their
derivatives). From now on, we shall assume that such a term has been properly included, and therefore we shall use
the covariant polynomial Pkm (q) in the propagator (9.71). Hence we shall omit the label (L) in that polynomial.
At this step, we have simply added the non-covariant term to recover the covariance of the theory, by this
addition has not been justified satisfactorily. We shall see that in the canonical formalism the non-covariant
term in the Hamiltonian density required to cancel non-covariant terms in the propagator, appears automatically.
Indeed, it is one of the main motivations for the introduction of the canonical quantization.

9.7.1 Other definitions of the propagator


There are other definitions of the propagator in the literature, equivalent to Eq. (9.44). By taking the vacuum
expectation value (VEV) of Eq. (9.18) page 227, we have
h i  h i 
+ +
ikm (x, y) = (x y) k (x) , m (y) (y x) m (y) , k (x) (9.78)
0 0
where
hABC . . .i0 h0| ABC . . . |0i

is the vacuum expectation value (VEV) of the product of operators ABC . . .. The fields k+ (x) and m (y)
annihilate the vacuum. Consequently, only one term in each commutator or anticommutator contributes to the
propagator in Eq. (9.78)
D E D E
ikm (x, y) = (x y) k+ (x) , m
+
(y) (y x) m
(y) , k (x) (9.79)
0 0

in addition, and + would annihilate the vacuum state on the right, while and +
would annihilate
the vacuum state on the left. Consequently, + and could be substituted everywhere in Eq. (9.79) with the
complete field + + , we then obtain
D E D E

ikm (x, y) = (x y) k (x) , m (y) (y x) m (y) , k (x)
0 0
9.8. FEYNMAN RULES AS INTEGRATIONS OVER MOMENTA 249

which can be written in the form


D n oE

ikm (x, y) = T k (x) m (y)
0

where T means a time-ordered product extended to all fields, with a minus sign for any odd permutation of
fermionic operators.
Notice that this definition of a generalized time-ordered product is not consistent with our definition in Eq.
(2.188) page 105 of the time-ordered product of Hamiltonian densities, because the Hamiltonian density only
contains even numbers of fermionic field factors.

9.8 Feynman rules as integrations over momenta

Figure 9.11: Feynman diagram in which we only indicate the flux of momenta and their associated exponential.
There is another exponential coming from the internal line, owing to the Fourier representation of the propagator.
The four-momentum q of the internal line flows from the vertex y to the vertex x.

The Feynman rules for a diagram of N th order have been described by means of integrals over N space-time
coordinates where the integrands are in turn products of space-time dependent factors. Nevertheless, for many
reasons integrations over momenta are more advantageous. On one hand, experiments in particle physics usually
measure momenta but not positions or times. On the other hand, momenta can be related through kinematics
coming from the principle of conservation of four-momentum, and the relation p2 = m2 is very useful when we
are dealing with on-shell particles.
When establishing those Feynman rules we found that for a final particle or antiparticle line with momentum
p leaving a vertex with space-time coordinate x , we obtain a factor proportional to exp (ip x). We also
found that for an initial particle line with momentum p entering a vertex with a space-time coordinate x the
associated factor becomes exp (+ip x). Finally, we saw that the factor associated with an internal line running
from y to x can be written as a Fourier integral over off-shell four momenta q with an integrand proportional to
exp [iq (x y)]. We could consider that the momentum q is flowing along the internal line in the direction of
250 CHAPTER 9. THE FEYNMAN RULES

the arrow i.e. from y to x. For the diagram in Fig. 9.11, we show the exponentials associated with each external
and internal lines. Since the four-momentum must be conserved at each vertex we have

p1 + p2 = q = p1 + p2

and the product P of all exponentials coming from external and internal lines yield

P eip1 x eip2 x eiq(xy) eip1 y eip2 y = ei(p1 +p2 )x eiqx eiqy ei(p1 +p2 )y

P = eiqx eiqx eiqy eiqy = 1

therefore, all exponential cancel each other and should not be included when we express the propagator explicitly
as a Fourier expansion. Besides, in order to account for conservation of four-momentum, the integral over each
vertexs space-time position provides a factor
X X X X 
(2)4 4 p+ q p q (9.80)
P P
where p and
P p denote
P the total four-momentum of all the final or initial particles leaving or entering the
vertex, while q and q denote the total four-momentum of all the internal lines with arrows leaving or entering
the vertex respectively. Then instead of integrals over spacetime coordinates x , we have to do integrals over the
Fourier variables q (in general off shell) one for each internal line (momenta associated with external lines are on
the mass shell, so they are fixed).
From the discussion above, it is convenient to contextualize the Feynman rules to adapt them for calculations
of contributions to the Smatrix by means of integrals over momenta (see Fig. 9.12).

1. The Feynman diagrams of a given order are the ones described in Sec. 9.3. However, we shall not label
each vertex with spacetime coordinates. Instead, each internal line is labelled with an off-mass-shell four-
momentum, that by convention flows in the direction of the arrow. For neutral particle lines without arrows
the momentum flows in either direction.

2. For a given vertex of type i, we put a factor


X X X X 
i (2)4 gi 4 p+ q p q (9.81)

with the conventions defined below equation (9.80). Note that the Dirac delta guarantees the conservation
of momentum at each vertex in the diagram.

(a) For each external line running upwards out of the diagram we include a factor

1 
3/2
uk p , , n for arrows pointing upwards (particle)
(2)
1 
v
3/2 k
p , , n for arrows pointing downwards (antiparticle)
(2)

(b) For each external line running from below into the diagram the associated factor is

1
uk (p, , n) for arrows pointing upwards (particle)
(2)3/2
1
vk (p, , n) for arrows pointing downwards (antiparticle)
(2)3/2
9.8. FEYNMAN RULES AS INTEGRATIONS OVER MOMENTA 251

Figure 9.12: Feynman representations of pairings of operators in the momentum-space evaluation of the Smatrix.
On the right we have the factors that must be included in the momentum space integrand of the Smatrix for each
line of the Feynman diagram.

(c) For each internal line running with edges labelled as k and m, with the arrow flowing from m to k,
with a momentum label q , we include as a factor the coefficient associated with eiqx in the integral
for ikm (x) [see Eq. (9.71) page 246]9 :

i (2)4 Pkm (q)



q 2 + m2km i

where m2km denotes the mass of the particle in the internal line defined between the vertices m and k.
We recall that the factors u (p,, n) and v (p,, n) and the polynomial P (q) are given by
1
u (q) = v (q) = p ; P (q) = 1 for scalars and pseudoscalars
2q 0

For Dirac spinors of mass M and four-momentum q, the u (q,, n) and v (q,, n) factors are the ones
described in Sec. 7.4, and the polynomial is the matrix P (q) = (i q + M ) .

3. Once all these factors are obtained, we integrate the product of all of them over the four-momenta carried
by internal lines, then sum over all field indices k, m etc.

4. Add up the result obtained from each diagram.


9
keep in mind that the exponential has been cancelled with exponentials of other lines.
252 CHAPTER 9. THE FEYNMAN RULES

5. As discussed in section 9.4, additional combinatoric factors and fermionic signs are required.

According to these rules we have a four-momentum integration variable for each internal line. However,
the Dirac factors associated with the vertices described in Eq. (9.81) eliminates many integrals. Energy and
momentum are separately conserved for each connected part of a Feynman diagram. Thus, there are C remaining
delta functions in a graph with C connected parts (C = 1 for each connected diagram). Therefore in a diagram
with I internal lines and V vertices, we shall have I (V C) independent four-momenta (i.e. not fixed by the
delta functions), and as we discussed in section 3.6.2 [see Eq. (3.59) page 133], this is precisely the number of
independent loops L
L = I V +C (9.82)
which is defined as the maximum number of internal lines that can be cut without disconnecting the diagram, since
any such (and only such) internal lines are associated with an independent four-momentum. From this discussion
the independent momenta characterizes the momenta that circulate in each loop. We call a tree graph as a
diagram without loops, for such diagrams after integrating the delta functions there are no momentum-space
integrals.

9.9 Examples of application for the Feynman rules with integration over
four-momenta variables
We take as a first example, a theory with an interaction of the type (9.24)
X
H (x) = gkmn k (x) m (x) n (x) (9.83)
kmn

As before, we shall calculate the Smatrix elements associated with fermion-boson, fermion-fermion, and boson-
boson scattering. But we shall express such matrix elements as integrals over momenta.

9.9.1 Fermion-boson scattering


The diagrams for fermion boson scattering in the theory (9.83) are the ones in Fig. 9.8, Page 236. Let us start
with the diagram in Fig. 9.8(a)

1. For the two initial particles (p1 , 1 , n1 ) and (p2 , 2 , n2 ) (fermion and boson respectively) we associate factors

un (p1 , 1 , n1 ) uk (p2 , 2 , n2 )
3/2
and
(2) (2)3/2

2. For the vertex (x, m) we add a factor

i (2)4 gmnk 4 (p1 + p2 q)

3. For the internal line with the arrow flowing from m to m , we add the coefficient of the propagator

i (2)4 Pm m (q)

q 2 + m2m m i

where mm m refers to the mass of the particle in the internal line between vertices m and m .

4. For the vertex (y, m ) we add a factor

i (2)4 gn m k 4 (q p1 p2 )
9.9. EXAMPLES OF APPLICATION FOR THE FEYNMAN RULES WITH INTEGRATION OVER FOUR-MOMENT

5. For the final fermion and boson states (p1 , 1 , n1 ) and (p2 , 2 , n2 ) we associate factors
un (p1 , 1 , n1 ) uk (p2 , 2 , n2 )
and
(2)3/2 (2)3/2

6. Let us put all this stuff together


" #" # " #
u (p , , n ) u (p , , n ) h i i (2) 4
P (q)
n 1 1 1 k 2 2 2 m m
P(a) i (2)4 gmnk 4 (p1 + p2 q) 2
(2)3/2 (2)3/2 q + m2m m i
" #" #
h i u (p , , n ) u (p , , n )

i (2)4 gn m k 4 (q p1 p2 ) n 1 1 1 k 2 2 2
(2)3/2 (2)3/2

7. This product is integrated over the momenta of internal lines, and sum over all field indices (knm) and
(k n m ). In this case we have only one internal line carrying a momentum q flowing upwards in the diagram
in Fig. 9.8(a). We then obtain
Z X X
(a) 
SF B d4 q P(a) = (i)2 (2)8 gn m k gmnk un p1 , 1 , n1 un (p1 , 1 , n1 )
k n m knm k n m knm
Z " #
i (2)4 Pm m (q)   
d4 q 2 2 (2)6 uk p2 , 2 , n2 uk (p2 , 2 , n2 ) 4 (p1 + p2 q) 4 (q p1 p2 )
q + mm m i

8. Now we add the contribution of diagram 9.8(b) (there are no extra combinatoric factors for these diagrams).
From which we finally obtain the Smatrix (9.35) page 237 for fermion-boson scattering with the momentum-
space rules
(a) (b)
SF B SF B + SF B = Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2
X Z  
 Pm m (q)
SF B = (i) (2) gn m k gmnk un p1 1 n1 un (p1 1 n1 ) d q i (2)4
2 8 4
q + m2m m i
2
k n m knm
 
(2)6 uk p2 2 n2 uk (p2 2 n2 ) 4 (p1 + p2 q) 4 (q p1 p2 )
 
+uk p2 2 n2 uk (p2 2 n2 ) 4 (p2 p1 + q) 4 (p1 p2 q)

where the labels 1 and 2 are denoting fermions and bosons respectively. After integrating over the off-mass-
shell momentum q, the Dirac delta functions demand that

q = p1 + p2 = p1 + p2 = p1 p2 = p2 p1

and we are left with a single Dirac function that provides the global conservation of four-momentum (because
this is a connected diagram). In addition, we have no integrals over off-shell momenta (owing to the absence
of loops, since this is a tree graph without independent momenta).

SF B Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
X 
gn m k gmnk un p1 1 n1 un (p1 1 n1 )
k n m knm
"
Pm m (p1 + p2 )

2 u k p2 2 n2 uk (p2 2 n2 )
(p1 + p2 ) + m2m m i
#
Pm m (p2 p1 ) 
+ uk p2 2 n2 uk (p2 2 n2 ) (9.84)
(p2 p1 )2 + m2m m i
254 CHAPTER 9. THE FEYNMAN RULES

It is convenient to define a more compact notation as follows, we can define the fermion-boson coupling matrix
as
[k ]nm gnmk (9.85)

in this matrix notation, the matrix element (9.84) for fermion-boson scattering could be rewritten in the form

-


SF B Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
(
X  Pm m (p1 + p2 ) 
[k ]n m [k ]mn un p1 1 n1 un (p1 1 n1 ) 2 2
u
k

p2 2 n2 uk (p2 2 n2 )

k n m knm (p 1 + p 2 ) + m mm i
)
X  Pm m (p2 p1 ) 
+ [k ]n m [k ]mn un p1 1 n1 un (p1 1 n1 ) 2 2
uk p2 2 n2 uk (p2 2 n2 )
k n m knm
(p 2 p 1 ) + m mm i


SF B Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
(
X  Pm m (p1 + p2 ) 
un p1 1 n1 [k ]n m 2 2
[k ]mn un (p1 1 n1 ) uk p2 2 n2 uk (p2 2 n2 )
k n m knm
(p1 + p2 ) + mm m i
)
X  Pmm (p 2 p 1 ) 
+ un p1 1 n1 [k ]n m 2 2
[k ]mn un (p1 1 n1 ) uk p2 2 n2 uk (p2 2 n2 )
k n m knm
(p 2 p 1 ) + m m m i

we have the following matrix multiplications


X  Pm m (p1 + p2 )
M1k k un p1 1 n1 [k ]n m [k ]mn un (p1 1 n1 )
n m nm
(p1 + p2 )2 + m2m m i
 P (p1 + p2 )
= u p1 1 n1 k k u (p1 1 n1 )
(p1 + p2 )2 + M 2 i
X  Pm m (p2 p1 )
M2k k un p1 1 n1 [k ]n m 2 2
[k ]mn un (p1 1 n1 )
(p 2 p1 ) + m i
n m nm mm
 P (p 2 p 1 )
= u p1 1 n1 k k u (p1 1 n1 )
(p2 p1 )2 + M 2 i

-
-

SF B Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
X   P (p1 + p2 ) 


u p1 1 n1 k k u (p1 1 n1 ) uk p2 2 n2 uk (p2 2 n2 )
k k
(p1 + p2 )2 + M 2 i
" #)

 P (p2 p1 )

+ u p1 1 n1 k k u (p1 1 n1 ) uk p2 2 n2 uk (p2 2 n2 ) (9.86)
(p1 p2 )2 + M 2 i

where M 2 is the diagonal mass matrix of the fermions in the internal lines associated with each propagator in
Eqs. (9.86).
9.9. EXAMPLES OF APPLICATION FOR THE FEYNMAN RULES WITH INTEGRATION OVER FOUR-MOMENT

9.9.2 Fermion-fermion scattering


Similarly, in the same theory the Smatrix element for fermion-fermion scattering given by Eq. (9.36) page 238,
and Fig. 9.4, page 232, yields (Homework!! B3)

SF F Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
X Pk k (p1 p1 )
gm mk gn nk
k n m knm
(p1 p1 )2 + m2k k i
   
um p2 2 n2 un p1 1 n1 um (p2 2 n2 ) un (p1 1 n1 ) 1 2 (9.87)

in the matrix notation defined by Eq. (9.85), the matrix element (9.87) for fermion-fermion scattering could be
rewritten in the form (Homework!! B4)

SF F Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
X Pk k (p1 p1 ) h  ih  i

u p2 2 n 2 k u (p2 2 n2 ) u p1 1 n 1 k u (p1 1 n 1 )
k k
(p1 p1 )2 + m2k k i
 
1 2 (9.88)

where m2 is the diagonal mass matrix of the bosons in the internal lines in Eq. (9.88) page 255, and Fig. 9.4,
page 232.

9.9.3 Boson-boson scattering

Figure 9.13: Feynman diagram (box diagram) with four vertices and four internal lines. Since this diagram has a
loop. we have one independent off-shell momentum q over which we shall integrate.

The rule to write the Smatrix contributions in matrix notation, is that we write coefficient functions, coupling
matrices and propagators in an ordered ruled by following fermion lines backwards from the ordered determined
256 CHAPTER 9. THE FEYNMAN RULES

by the arrows10 . For example, in the matrix notation, the Smatrix for boson-boson scattering in the same
theory is given by the sum of one loop diagrams, as shown in Fig. 9.7 page 235. We show in Fig. 9.13, the flow
of momenta in that loop diagram. Note that with the labelling of momenta in each internal line of Fig. 9.13, the
conservation of four-momentum is guaranteed in each vertex. We see it by using the labelling of vertices in Fig.
9.7, and using positive sign for momenta flowing into the vertex, and negative sign for momenta flowing out of
the vertex, we see that
 
Vk1 q + p1 p1 q + p1 + p1 = 0

Vk1 q + p1 p1 q = 0

Vk2 q q p2 p2 = 0
  
Vk2 q p2 + p2 q + p1 p1 = (p1 + p2 ) p1 + p2 = q q = 0

Thus, we shall not introduce the Dirac delta for each vertex but only the global Dirac delta (p1 + p2 p1 p2 ) associated
with the global conservation of four-momentum. In that case the Smatrix element can be written as follows
-
-

1. Initial particles (p1 1 n1 ) and (p2 2 n2 ) (both are bosons) provide factors
uk1 (p1 1 n1 ) uk2 (p2 2 n2 )
3/2
;
(2) (2)3/2

2. Final particles (p1 1 n1 ) and (p2 2 n2 ) (both are bosons) provide factors
uk (p1 1 n1 ) uk (p2 2 n2 )
1 2
;
(2)3/2 (2)3/2

3. The vertex k1 gives (not Dirac delta is introduced)

i (2)4 gn1 n4 k1

4. The propagator from k1 to k1 yields


i (2)4 Pk1 k1 (q + p1 )

(q + p1 )2 + m2k k1 i
1

5. Vertex k1 yields
i (2)4 gn2 n1 k1

6. Propagator from k1 to k2 gives


i (2)4 Pk2 k1 (q)

q 2 + m2k k i
2 1

7. Vertex k2 yields
i (2)4 gn3 n2 k2

8. Propagator from k2 to k2 gives


i (2)4 Pk2 k2 (q p2 )

(q p2 )2 + m2k2 k i
2

10
The reader can check such a rule by contrasting Eq. (9.86) with Fig. 9.8, Page 236, as well as Eq. (9.88) with Fig. 9.4, page 232.
9.9. EXAMPLES OF APPLICATION FOR THE FEYNMAN RULES WITH INTEGRATION OVER FOUR-MOMENT

9. Vertex k2 gives
i (2)4 gn4 n3 k2

10. Propagator from k2 to k1 yields


i (2)4 Pk1 k2 (q + p1 p1 )

(q + p1 p1 )2 + m2k1 k2 i

11. Let us multiply all these factors, put a minus sign associated with the fermionic loops, and the global Dirac
delta function
" #" #
 uk1 (p
1 1 1n ) uk2 (p n
2 2 2 ) uk (p1 1 n1 ) uk (p2 2 n2 )
P p1 + p2 p1 p2 1 2

(2)3/2 (2)3/2 (2)3/2 (2)3/2


" #
h i i (2) 4
Pk k (q + p )
h i i (2) 4
P k k (q)
1
i (2)4 gn1 n4 k1 1 1 i (2)4 gn n k 2 1

(q + p1 )2 + m2k k1 i q 2 + m2k k i
2 1 1

1 2 1
" #
h i 4 h i 4 p )
i (2) Pk k (q p ) i (2) P (q + p
2 k k 1
i (2)4 gn3 n2 k2 i (2)4 gn4 n3 k2 1
2 2 1 2

(q p2 )2 + m2k2 k i (q + p1 p1 )2 + m2k1 k2 i
2

(p1 + p2 p1 p2 ) h

i
P uk p 1 1 n 1 u k p 2 2 n 2 [uk1 (p1 1 n1 ) uk2 (p2 2 n2 )]
(2)6 1 2

" #
h i Pk2 k1 (q) h i Pk1 k1 (q + p1 )
gn3 n2 k2 gn2 n1 k1
q 2 + m2k k i (q + p ) 2 + m2
1 k k i
2 1 1 1
" #

Pk2 k2 (q p2 )
Pk1 k2 (q + p1 p1 )
[gn1 n4 k1 ] 2 [gn4 n3 k2 ]
2
(q + p1 p1 ) + mk1 k2 i (q p2 )2 + m2k2 k i
2

we can replace the couple of indices for vertices by the corresponding index of the internal line thus

k2 k1 n2 ; k1 k1 n1 ; k1 k2 n4 ; k2 k2 n3

using this, and applying the matrix notation we have


(p1 + p2 p1 p2 ) h

i
P uk p1 1 n1 uk p2 2 n2 [uk1 (p1 1 n1 ) uk2 (p2 2 n2 )]
(2)6 1 2

  " #
h i Pn2 (q) h i Pn1 (q + p1 )
k2 k1
n3 n2 q 2 + m2 n2 i n2 n1 (q + p )2 + m2 i
1 n1
" # " #
Pn4 (q + p1 p1 ) Pn3 (q p2 )
[k1 ]n1 n4 [k2 ]n4 n3
(q + p1 p1 )2 + m2n4 i (q p2 )2 + m2n3 i

(p1 + p2 p1 p2 ) h

i
P u
k1 p n
1 1 1 uk2 p n
2 2 2 [uk1 (p1 1 n1 ) uk2 (p2 2 n2 )]
(2)6
  " #
P (q) P (q + p1 )
k2 2
q + M 2 i n3 n2 k1 (q + p1 )2 + M 2 i
n2 n1
" # " #

P (q + p1 p1 ) P (q p2 )
k 1 k 2
(q + p1 p1 )2 + M 2 i n n (q p2 )2 + M 2 i n n
1 4 4 3
258 CHAPTER 9. THE FEYNMAN RULES

note that the chain of matrix multiplications (in the two last lines) starts and ends in the same index (n3 in
this case), thus when summing over the indices n1 , n2 , n3 , n4 what we obtain is the trace of such product of
matrices

12. Now we integrate over the off-shell momentum q (after integrating over all four momenta associated with
each internal line, we are left with only one integral and one global delta because of the Dirac delta functions
over each vertex). We also sum over field indices (n1 , n2 , n3 , n4 ) and (k1 , k2 , k1 , k2 )
Z X X
S(1) = d4 q P
k1 k2 k1 k2 n1 n2 n3 n4

the first set of sums gives the trace of the product of matrices as mentioned above. Thus we finally obtain


SBB Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = (2)6 4 p1 + p2 p1 p2
X  
uk p1 , 1 , n1 uk p2 , 2 , n2 uk1 (p1 , 1 , n1 ) uk2 (p2 , 2 , n2 )
1 2
k1 k2 k1 k2
Z (
P (q) P (q + p1 )
d4 q T r k2
q 2 + M 2 i k1 (q + p1 )2 + M 2 i
)
P (q + p1 p1 ) P (q p2 )
k 1 k2
(q + p1 p1 )2 + M 2 i (q p2 )2 + M 2 i
+... (9.89)

the ellipsis in the last line indicates terms obtained by permuting bosons 1 , 2 , and 2. The minus sign at the
beginning of the RHS is the extra minus sign associated with the fermionic loops. After eliminating all delta
functions we are left with one momentum-space integral, as it must be for a one loop diagram.
By comparing (9.89) with Fig. 9.13 the reader can check once again that the rule to write the Smatrix
contributions in matrix notation, is that we write coefficient functions, coupling matrices and propagators in an
ordered ruled by following lines backwards from the ordered determined by the arrows.

9.10 Examples of Feynman rules as integrations over momenta


Consider a theory involving a Dirac spinor field (x) of mass M and a pseudoscalar field (x) of mass m, with
an interaction given by
ig5 (9.90)

where the factor i is included for the interaction to be hermitian with a real coupling constant g. Recalling that
the polynomial P (q) is the unity for scalars and [i q + M ] for the Dirac spinors, and that the coefficient
functions are u = (2E)1/2 for a scalar of energy E while for the normalized Dirac spinors u is the one shown in
Sec. 7.4. With this input, equations (9.86), (9.88) and (9.89) give the lowest-order connected Smatrix elements
for fermion-boson scattering, fermion-fermion scattering and boson-boson scattering
9.10. EXAMPLES OF FEYNMAN RULES AS INTEGRATIONS OVER MOMENTA 259

9.10.1 Fermion-boson scattering


Let us start with Eq. (9.86)

SF B Sp1 1 n1 , p2 2 n2 ; p1 1 n1 , p2 2 n2 = i (2)2 4 p1 + p2 p1 p2
X   P (p1 + p2 ) 


u p1 1 n1 k k u (p1 1 n1 ) uk p2 2 n2 uk (p2 2 n2 )
k k
(p1 + p2 )2 + M 2 i
" #)

 P (p2 p1 )

+ u p1 1 n1 k k u (p1 1 n1 ) uk p2 2 n2 uk (p2 2 n2 ) (9.91)
(p1 p2 )2 + M 2 i

by taking into account that in theory (9.90) there is only one type of fermion and one type of boson and that
there is only one value of for scalar bosons, we write

SF B Sp1 1 , p2 ; p1 1 , p2 = i (2)2 4 p1 + p2 p1 p2
 

 P (p1 + p2 )

u p1 1 u (p1 1 ) u p2 u (p2 )
(p1 + p2 )2 + M 2 i
" #)
 P (p 2 p1 ) 
+ u p1 1 u (p1 1 ) u p2 u (p2 ) (9.92)
(p1 p2 )2 + M 2 i
we start by inserting the polynomials for fermions, the coefficient functions for scalars and the fermion-fermion-
boson coupling according with Eq. (9.90)
1
P (q) = [i q + M ] ; u (p) = ; = ig5 (9.93)
2E
inserting (9.93) in (9.92) we have

SF B Sp1 1 , p2 ; p1 1 , p2 = i (2)2 4 p1 + p2 p1 p2
(" #

 [i (p1 + p2 ) + M ] 1 1
u p1 1 [ig5 ] [ig5 ] u (p1 1 ) p
(p1 + p2 )2 + M 2 i 2E2 2E2
" #)
 [i (p p ) + M ]
1 1 1
+ u p1 1 [ig5 ] 2
[ig5 ] u (p1 1 ) p (9.94)
(p1 p2 )2 + M 2 i 2E2 2E2

 1 1
SF B Sp1 1 , = i (i)2 (2)2 g2 4 p1 + p2 p1 p2 p
p2 ; p1 1 , p2
2E2 2E2
 
 [i (p1 + p2 ) + M ]
u p1 1 5 5 u (p1 1 )
(p1 + p2 )2 + M 2 i
" #)
 [i (p p ) + M ]
1
+ u p1 1 5 2
5 u (p1 1 ) (9.95)
(p1 p2 )2 + M 2 i
???
1 
SF B Sp1 1 , = i (2)2 g2 p 4 p1 + p2 p1 p2
p2 ; p1 1 , p2
4E2 E2
 

 [i (p1 + p2 ) + M ]
u p1 1 5 5 u (p1 1 )
(p1 + p2 )2 + M 2 i
" #)
 [i (p p ) + M ]
1
+ u p1 1 5 2
5 u (p1 1 ) (9.96)
(p1 p2 )2 + M 2 i
260 CHAPTER 9. THE FEYNMAN RULES

1/2 4 
SF B Sp1 1 p2 ;p1 1 p2 = i (2)2 g2 4E2 E2 p1 + p2 p1 p2

 i (p1 + p2 ) + M
u p1 1 5 5 u (p1 1 )
(p1 + p2 )2 + M 2 i
)
 i (p p ) + M
1
+ u p1 1 5 2
5 u (p1 1 )
(p1 p2 )2 + M 2 i

9.10.2 Fermion-fermion and Boson-boson scattering


For fermion-fermion and boson-boson scattering we can use Eq. (9.88) page 255 and Eq. (9.89) page 258, to
obtain the Smatrix elements for these processes in the framework of the theory (9.90). We obtain


SF F Sp1 1 p2 2 ;p1 1 p2 2 = i (2)2 g2 4 p1 + p2 p1 p2
    
u p2 2 5 u (p2 2 ) u p1 1 5 u (p1 1 )
1  
2 1 2
(p1 p1 ) + m2 i

1/2 4 
SBB Sp1 p2 ;p1 p2 = (2)6 g4 16E1 E2 E1 E2 p1 + p2 p1 p2
Z (
i q + M i (q + p1 ) + M
d4 q T r 5 2 5
q + M 2 i (q + p1 )2 + M 2 i
)
i (q + p1 p1 ) + M i (q p2 ) + M
5 5 + ...
(q + p1 p1 )2 + M 2 i (q p2 )2 + M 2 i

once again the ellipsis mean a sum over permutations of particles 2, 1 , 2 . The factors u have been replaced by
u by using the definition (7.152) page 207.

9.11 Topological structure of the lines


The topological structure of the diagrams suggests a kind of conservation law of lines. It is useful if we consider
the internal and external lines as being created at vertices and destroyed in pairs at the centers of internal lines
or when external lines leave the diagram. Note that it is not related with the direction of the arrows carried by
the lines. Let I and E be the number of internal and external lines respectively, where Vi denotes the numbers
of vertices of various types labelled by i, and ni are the number of lines attached to each vertex. Equating the
number of lines that are created and destroyed we obtain
X
2I + E = ni Vi (9.97)
i

this also holds separately for fields of each type.


Let us take an example, assume that we have a theory with four types of interactions described as follows

T ype 1 f ermion f ermion boson vertex(trilinear) ; n1 = 3


T ype 2 boson boson boson vertex (trilinear); n2 = 3
T ype 3 boson boson f ermion vertex(trilinear); n3 = 3
T ype 4 f our bosons vertex ; n4 = 4
9.11. TOPOLOGICAL STRUCTURE OF THE LINES 261

Figure 9.14: Two connected diagrams for a theory with four types of interactions. (a) Scattering from fermion-
fermion to fermion-boson. (b) Scattering from boson-boson to fermion-boson.

In Fig. 9.14 (a), we have a connected diagram of the scattering from fermion-fermion to fermion-boson. This
diagram as a whole cannot describe a physical process11 , but it could be a piece of a greater diagram. For this
diagram we have

E = 4 ; I = 7 ; V1 = 4 (vertices a, d, e, f ) ;
V2 = 1 (vertex b) ; V3 = 1(vertex c) ; V4 = 0

thus Eq. (9.97) gives in this case


27+4 = 34+31+31+40
In Fig. 9.14 (b), we have a connected diagram of the scattering from boson-boson to fermion-boson. For the
same reasons as before, this diagram cannot account on a physical process, but can be part of a greater diagram.
For this diagram we have

E = 4 ; I = 6 ; V1 = 3 (vertices c, d, e) ;
V2 = 0 ; V3 = 1 (vertex b) ; V4 = 1 (vertex a)

thus Eq. (9.97) gives in this case


26+4 = 33+30+31+41
11
The initial state (fermion-fermion) has an integer spin, while the final process (fermion- boson) has a half-odd integer spin.
Therefore, the difference between final and initial spins is a half-odd integer, and we cannot balance such a difference with an orbital
angular momentum (since orbital angular momenta are always integer). Thus, the process is forbidden by conservation of total angular
momentum.
262 CHAPTER 9. THE FEYNMAN RULES

In the special case in which all interactions involve the same number of fields, we have ni = n so that Eq.
(9.97) becomes
2I + E = nV (9.98)
where V is the total number of all vertices. In this case we can combine Eqs. (9.82) and (9.98) to eliminate I,
and for the case of a connected diagram (C = 1) we have

L = I V +C =I V +1 I =L+V 1
nV = 2I + E = 2 (L + V 1) + E
(n 2) V = 2L + E 2
2L + E 2
V = (9.99)
n2
For example, if we have a trilinear interaction, the diagrams for a scattering process of two particles into two
particles (E = 4) with L = 0, 1, 2, . . . we have V = 2, 4, 6, . . . . In general, the expansion in powers of the coupling
constants is an expansion in increasing numbers of loops.

9.12 Off-shell and on-shell four-momenta

Figure 9.15: (a) A one-loop diagram. (b) The previous diagram could be constructed by connecting two pairs of
external lines of tree graph diagrams.

In the Feynman diagrams for the Smatrix, the external lines are on the mass shell (or simply on-shell),
i.e. the four-momentum associated with each external line obeys the restriction p2 = m2 for a particle of mass
m. It is however important to consider also external lines in which the associated energies are not related with
the three-momenta (like the case of internal lines) i.e. diagrams off the mass shell (or simply off shell). This off
shell diagrams are usually part of a larger Feynman diagram. For example, a loop appearing as an insertion in
some internal line of a diagram could be seen as a Feynman diagram with two external lines, in which both are off
the mass shell. Figure 9.15, shows a one loop diagram that can be formed by connecting two diagrams in which
the external lines that form the loop should be off-shell.
In the path integral approach, it is usual to derive first the Feynman rules with all external lines off-shell and
then construct the Smatrix elements by applying the on-shell limit.
Once the contribution of a diagram off the mass shell is calculated, the associated contribution to the S
matrix can be calculated by takingp the on shell restriction and considering the four-momentump flowing along the
0 2 2 0
line into the diagram with p = p + m for particles in the initial state and with p = p + m2 for particles
2

in the final state inserting the apropriate external line factors:


uk vk
q ; q for initial particles or antiparticles respectively
(2)3 (2)3
uk vk
q ; q for final particles or antiparticles respectively
(2)3 (2)3
9.12. OFF-SHELL AND ON-SHELL FOUR-MOMENTA 263

Feynman diagrams off the mass shell are a particular case of some types of diagrams that take into account the
effects of various possible external fields. Let us consider some additional terms in the Hamiltonian involving some
external fields a (x), such that the interaction V (t) in the Dyson series (2.188) for the Smatrix is replaced by
XZ
V (t) = V (t) + d3 x a (x) Oa (x, t) (9.100)
a

where the currents Oa (t) have the time-dependence typical of operators in the interaction picture
Oa (t) = exp (iH0 t) Oa (0) exp (iH0 t) (9.101)
but otherwise such operators are arbitrary. The Smatrix for a given transition becomes a functional
S [] in terms of the cnumber function a (t). We require an extension of the usual Feynman rules to obtain
such a functional. Besides the usual vertices obtained from V (t) we must include some extra vertices:

1. If the current Oa (x) is a product of na field factors, each Oa vertex with position label x, must be attached
to na lines of the corresponding types. So its contribution to the position-space Feynman rules, is equal
to ia (x) times numerical factors that appear in Oa (x). (Note that a (x) is acting as though it were a
constant coupling).
2. The r th variational derivative of S [] with respect to a1 (x1 ) , a2 (x2 ) , . . . , ar (xr ) at = 0 is given
by position space diagrams with r additional vertices, to which internal lines na1 , na2 , . . . are attached
respectively (i.e. the number nak of fields associated with each current Oak (x, t)), and no external lines
(since we are evaluating at = 0 i.e. eliminating the external fields a (x)). These vertices have position
labels x, y, . . . over which we do not integrate. Each of these vertices provides a contribution of i times
numerical factors that appear in the corresponding current Oa . To see it, we observe that the non-vanishing
terms of S [] when we take the r th variational derivatives are precisely the terms that contain the
rexternal fields a1 (x1 ) , a2 (x2 ) , . . . , ar (xr ) (and nothing else). On one hand, the variational derivative
with respect to ak (xk ) vanishes if such a field is not contained in the term12 . On the other hand, if there
are some extra fields in the term the derivative vanishes when evaluating at = 0.

-
An important particular case appears when these currents are all single field factors, it can be written as
XZ
V (t) = V (t) + d3 x k (x, t) k (x, t)
k
the r th variational derivative of S [] with respect to k1 (x1 ) , k2 (x2 ) , . . . , kr (xr ) at = 0 is represented by
space diagrams with r additional vertices carrying space-time labels x1 , x2 , . . . , xr and to each of then is attached
a single internal particle line of type k1 , k2 , . . . , kr We can figure out them as off shell external lines, but with the
difference that their contribution to the matrix element is not a coefficient of the form (2)3/2 uk (p, ) eipx or
of the form (2)3/2 uk (p, ) eipx but a propagator13 but also a factor (i) from the vertex at the end of the
line. The result is then a momentum space Feynman diagram with particles in states and on the mass shell
with the addition of r external lines of type k1 , k2 , . . . , kr which carry momenta p1 , p2 , . . . , pr from the variational
derivative  
r S []
k1 (x1 ) k2 (x2 ) . . . kr (xr ) =0
by removing the propagators on each off-shell line and taking the adequate Fourier transform , and multiplying
with the proper coefficient functions uk , uk etc, and a factor (i)r .
12
Thus in the Dyson series (2.188), page 105, only survives terms with rvertices V1 (t1 ) V2 (t2 ) Vr (tr ) which contains precisely
the set of external fields a1 (x1 ) a2 (x2 ) . . . ar (xr ), through the vertex correction (9.100).
13
It has to do with the fact that the external lines a (x) of these diagrams will be part of another diagrams in which such external
lines become internal lines, as illustrated in Fig. 9.15, page 262.
264 CHAPTER 9. THE FEYNMAN RULES

9.12.1 The rth derivative theorem


There is a simple relation between the sum of contributions coming from all diagrams associated with perturbation
theory for any off shell amplitude and a matrix element, between eigenstates of the full Hamiltonian, of a time-
ordered product of corresponding operators in the Heisenberg picture. The theorem that provides such a relations,
states that to all orders of perturbation theory
 
r S []

r S = (i)r T {Oa1 (x1 ) Oar (xr )} + (9.102)
a1 (x1 ) a2 (x2 ) . . . ar (xr ) . . . =0

where Oa (x), are the counterparts of Oa (x) in the Heisenberg picture

Oa (x, t) = exp (iHt) Oa (x, 0) exp (iHt) = (t) Oa (x, t) 1 (t) (9.103)
iHt iH0 t
(t) e e (9.104)

we recall that | + i and | i are in and out eigenstates of the full hamiltonian H.
We prove it as follows. From Eqs. (2.36, 2.188) we have
" Z #
D E D X (i)N E
(0) (0) (0)
S [] = S [] = d1 d2 dN T {V1 (1 ) V2 (2 ) VN (N )} (0)
N!
N =0

X (i)N Z D E

S [] = d1 d2 dN (0) T {V1 (1 ) V2 (2 ) VN (N )} (0)
N!
N =0

as discusssed above, when we apply the r th derivative only terms with roperators of the form Oak (xk , tk )
survive in the Dyson series

N
Y XZ XZ
3
P VI (I ) [V1 (1 ) V2 (2 ) VN (N )] = V (1 ) + d y b1 (y1 ) Ob1 (y1 ) V (2 ) + d3 y b2 (y2 )
I=1 b1 b2

XZ
V (N ) + d3 y bN (yN ) ObN (yN )
bN

r P
= (i)r V (1 ) V (N ) Oa1 (x1 ) Oa2 (x2 ) Oar (xr )
a1 (x1 ) . . . ar (xr )
Therefore, from the Dyson series Eq. (2.188) page 105 we observe that the left-hand side of Eq. (9.102) reads
  X Z
r S [] (i)N +r
r S = d1 d2 dN
a1 (x1 ) . . . ar (xr ) =0 N!
N =0
D E

(0) T {V (1 ) V (N ) Oa1 (x1 ) Oar (xr )} (0) (9.105)

let us assume that x01 x02 . . . x0r such that the currents Oak (xk ) are not reordered by the time ordering
operator. However, the original interaction V (1 ) V (N ) are not necessarily ordered neither they are in order
with respect to the currents. Hence, a given subset of the N vertices are on left of Oa1 (x1 ), another subset is on
the left of Oa2 (x2 ) and so on. Of course there could also be a subset of vertices on the right of Oar (xr ).
We shall denote as 01 0N0 all s that are greater than x01 ; as 11 1N1 all s between x01 and x02 and so
on. We finally denote as r1 rNr all s that are less than x0r . With this notation, the time ordering yields

T T {V (1 ) V (N ) Oa1 (x1 ) Oar (xr )} = T {V (01 ) V (0N0 )} Oa1 (x1 ) T {V (11 ) V (1N1 )} Oa2 (x2 )
 
T V (r1,1 ) V r1,Nr1 Oar (xr ) T {V (r1 ) V (rNr )}
9.12. OFF-SHELL AND ON-SHELL FOUR-MOMENTA 265

note that we have r + 1 cases to fill between the rcurrents Oa1 (x1 ) Oar (xr ) in the time ordering process.
That is
{ 1 } Oa1 (x1 ) { 2 } Oa2 (x2 ) . . . { r } Oar (xr ) { r + 1 }
with

{ 1 } = T {V (01 ) V (0N0 )} ; { 2 } = T {V (11 ) V (1N1 )}


 
{ r } = T V (r1,1 ) V r1,Nr1 ; { r + 1 } = T {V (r1 ) V (rNr )}

it is clear that all N vertices V (i ) must be sort out into the r+1 cases, and that the order in each case is not relevant
(because of the presence of the time ordering operator for each subset). Therefore this is a typical combinatory
and the possible ways of sorting the N vertices into the r + 1 subsets each one containing N0 , N1 , . . . , Nr vertices
is given by
N!
with N0 + N1 + . . . + Nr = N
N0 !N1 ! Nr !
for each subset we can reformulate the limits of integration in time according with its position with respect to the
currents. For instance, for the first subset, we had by definition that 01 0N0 are greater than x01 then we can
replace Z Z

d01 d0N0 d01 d0N0
x01

we also had that 11 1N1 are between x01 and x02 thus
Z Z x01
d11 d1N1 d11 d1N1
x02

similarly
Z Z x0r1
dr1,1 dr1,Nr1 dr1,1 dr1,Nr1
x0r

finally, r1 rNr are less than x0r so


Z Z x0r
dr1 drNr d11 d1N1

with all these considerations, Eq. (9.105) becomes


  X
r S [] (i)N X N ! N,N0 +N1 ++Nr
r S = (i)r
a1 (x1 ) . . . ar (xr ) =0 N! N0 !N1 ! Nr !
N =0 N0 N1 Nr
Z Z x0 Z x0 Z x0r
1 r1
d01 d0N0 d11 d1N1 dr1,1 dr1,Nr1 dr1 drNr
x01 x02 x0r
h| T {V (01 ) V (0N0 )} Oa1 (x1 ) T {V (11 ) V (1N1 )} Oa2 (x2 )
 
T V (r1,1 ) V r1,Nr1 Oar (xr ) T {V (r1 ) V (rNr )} |i

where the factor


N ! N,N0 +N1 ++Nr
N0 !N1 ! Nr !
is the combinatoric factor which provides the number of ways of sorting N s into r + 1 subsets, where each
subset contains N0 , N1 , . . . , Nr of these s, with the constraint N0 + N1 + . . . + Nr = N . Then we sum over N .
Now, since in the process of summing over N such a number takes any non-negative integer, we observe that
summing over N and then summing over N0 , N1 , . . . , Nr with the constraint N0 + . . . + Nr = N , is equivalent to
266 CHAPTER 9. THE FEYNMAN RULES

sum over N0 , N1 , . . . , Nr independently, that is with each Nk taking all non-negative integer values (i.e. omitting
the constraint). Further we can decompose (i)N as (i)N0 (i)Nr . From this process we obtain
  XX X (i)N0 (i)N1 (i)Nr
r S []
r S = (i)r h|
a1 (x1 ) . . . ar (xr ) =0 N0 !N1 ! Nr !
N0 N1 Nr
Z Z x0 Z x0 Z x0r
1 r1
d01 d0N0 d11 d1N1 dr1,1 dr1,Nr1 dr1 drNr
x01 x02 x0r
T {V (01 ) V (0N0 )} Oa1 (x1 ) T {V (11 ) V (1N1 )} Oa2 (x2 )
 
T V (r1,1 ) V r1,Nr1 Oar (xr ) T {V (r1 ) V (rNr )} |i

separating each sum and respecting the time ordering we find



 r  X N0 Z
S [] (i)
r S = (i)r h| d01 d0N0 T {V (01 ) V (0N0 )} Oa1 (x1 )
a1 (x1 ) . . . ar (xr ) =0 N0 ! x 0
N0 1

X (i)N1 Z x01
d11 d1N1 T {V (11 ) V (1N1 )} Oa2 (x2 )
N1 ! x 0
N1 2

X (i)Nr1 Z x0r1  
dr1,1 dr1,Nr1 T V (r1,1 ) V r1,Nr1
Nr1 ! x0r
Nr1
" #
X (i)Nr Z x0r
Oar (xr ) dr1 drNr T {V (r1 ) V (rNr )} |i
Nr !
Nr

defining

X Z

(i)N t
U t ,t d1 dN T {V (1 ) V (N )} (9.106)
N! t
N =0
we can write the r th derivative as
 
r S []  
r S = (i)r h| U , x01 Oa1 (x1 ) U x01 , x02
a1 (x1 ) . . . ar (xr ) =0
  
Oa2 (x2 ) U x02 , x03 U x0r1 , x0r Oar (xr ) U x0r , |i (9.107)

Note that the operator (9.106) has a structure similar to the Dyson series Eq. (2.188) page 105 (except that
U (t , t) has finite limits of integration). The operator U (t , t) obeys the differential equation
d   

U t , t = iV t U t , t ; U (t, t) = 1
dt
whose solution reads
    
U t , t = exp iH0 t exp iH t t exp (iH0 t) = 1 t (t) (9.108)

where is defined by Eq. (9.104). Substituting Eq. (9.108) in Eq. (9.107) we obtain

 
r S []     
r S = (i)r h| 1 () x01 Oa1 (x1 ) 1 x01 x02
a1 (x1 ) . . . ar (xr ) =0
        
Oa2 (x2 ) 1 x02 x03 1 x0r1 x0r Oar (xr ) 1 x0r () |i (9.109)
9.12. OFF-SHELL AND ON-SHELL FOUR-MOMENTA 267

taking into account that 1 () = () and using Eq. (9.103) we have


 
r S []   
r S = (i)r h () | x01 Oa1 (x1 ) 1 x01
a1 (x1 ) . . . ar (xr ) =0
       
x02 Oa2 (x2 ) 1 x02 x03 1 x0r1 x0r Oar (xr ) 1 x0r | () i (9.110)
 
r S []
r S = (i)r h () | Oa1 (x1 ) Oa2 (x2 ) Oar (xr ) | () i (9.111)
a1 (x1 ) . . . ar (xr ) =0
On the other hand we saw in 2.1 that [in the sense of Eq. (2.16) page 68] the in and out states are related
with the free states in the form given by Eq. (2.19) page 68
E
= () (0) (9.112)

hence Eq. (9.111) becomes


 
r S []

r S = (i)r Oa1 (x1 ) Oa2 (x2 ) Oar (xr ) + (9.113)
a1 (x1 ) . . . ar (xr ) =0

Recalling that we have assumed the condition x01 x02 . . . x0r , we could replace the product of operators
on the right-hand side of Eq. (9.113) with a time-ordered product of operators
 
r S []

r S = (i)r T {Oa1 (x1 ) Oar (xr )} + (9.114)
a1 (x1 ) . . . ar (xr ) =0

Both sides of Eq. (9.114) are completely symmetric (or completely antisymmetric for fermions) in the a s and
x s. Consequently, this expression is satisfied regardless the order of the times x01 , x02 , . . . , x0r . Equation (9.114)
coincides with the result (9.102) we were looking for.
Chapter 10

Canonical quantization

The canonical quantization of postulated Lagrangians has been the first historical approach for quantum field
theories, and it is also the initial approach in most books of quantum field theory. This starting point has
the advantage that most of the known quantum fields theories can be easily formulated in a Lagrangian form.
Moreover, a classical theory with a Lorentz invariant Lagrangian density, leads when canonically quantized to a
Lorentz invariant quantum theory. The canonical formalism will lead to quantum mechanical operators that obey
the commutation relations of the Poincare algebra, leading in turn to a Lorentz-invariant Smatrix.
The preservation of the Lorentz invariance after canonical quantization should not be taken for granted and
it is and outstanding property. On one hand, some symmetries could be broken after a process of quantization
(phenomenon known as anomaly), and on the other hand we saw in section 9.7 that in theories with derivative
couplings or spins j 1, that it does not suffice to construct the interaction Hamiltonian as the integral over
space of a scalar interaction density. It is necessary to add a non-scalar Hamiltonian density to compensate
non-invariant terms in the propagators. The canonical formalism with a scalar Lagrangian density provides the
additional required terms for such a cancellation. Further, in the case of non-abelian gauge theories is particularly
difficult to guess the form of the extra terms without starting with a Lorentz-invariant and gauge-invariant
Lagrangian density.

10.1 Canonical variables


Our present developments will lead to commutation rules and equations of motions proper for a Hamiltonian
version of the canonical formalism. This is the Hamiltonian formalism we require to calculate the Smatrix
regardless we do it with a canonical or path integral formalism. However, it is not in general easy to find
Hamiltonians that provides a Lorentz-invariant Smatrix. Thus, the starting point will be a Lagrangian version
of the canonical formalism in order to derive satisfactory Hamiltonians. To do it, we should identify properly
the canonical fields and their associated conjugates in various field theories. In this way we shall learn how to
separate the free-field terms in the Lagrangian, to finally check that physically realistic theories are possible in
the canonical formalism.
We shall start by proving that the free fields for scalar, vector and Dirac fields provide a system of quantum op-
erators q n (x, t) and canonical conjugates pn (x, t) that satisfy the appropriate (equal-time) canonical commutation
and anticommutation relations

[q n (x, t) , pn (y, t)] = i3 (x y) nn (10.1)


 n 
q (x, t) , q n (y, t) = 0 ; [pn (x, t) , pn (y, t)] = 0 (10.2)

the subscripts indicates commutators if either of the particles that are created and destroyed by the two
operators are bosons, and anticommutators if both particles are fermions.

268
10.1. CANONICAL VARIABLES 269

10.1.1 Canonical variables for scalar fields


Let us start with a real scalar field (x) that describes a selfcharge-conjugate particle of zero spin. By combining
Eqs. (5.8, 5.9) page 152 with Eq. (5.20) page 154, we obtain the commutation relation for such a field
     
[ (x) , (y)] = + (x) + (x) , + (y) + (y) = + (x) , + (y) + + (x) , (y)
   
+ (x) , + (y) + (x) , (y)
   
= + (x) , (y) + (y) , (x) = + (x y) + (y x)
Z 3 h i
1 d p ip(xy) ip(xy)
[ (x) , (y)] = e e
(2)3 2p0
let us recall the reader that + (x) is even in x only for space-like separations of x and y. Hence for time-like
separations this commutator does not vanish in general. Hence the self-charge-conjugate scalar field obeys the
commutation relation

[ (x) , (y)] = (x y) (10.3)


Z 3 h i p
1 d k ikx ikx
(x) 3 0
e e ; k0 k2 + m2 (10.4)
(2) 2k
the function (x) and its time derivative can be written as
Z 3 h i
1 d k i(kxk0 t) i(kxk 0 t)
(x, t) e e
(2)3 2k0
Z 3 h i Z 3 h i
1 d k 0 i(kxk 0 t) 0 i(kxk 0 t) i d k i(kxk0 t) i(kxk 0 t)
(x, t) = ik e ik e = e + e
(2)3 2k0 (2)3 2
Z
(x, t) = i
d3 kei(kxk t)
0
3
(2)
where we have denoted with a dot, the derivative with respect to the time x0 = t. Such functions evaluated at
t = 0 yield
Z 3 h i Z
1 d k ikx ikx 2i d3 k
(x, 0) e e = sin k x
(2)3 2k0 (2)3 2 k2 + m2
Z
i
(x, 0) = 3 d3 k eikx
(2)
(x, 0) is the
the integral of (x, 0) vanishes since the integrand is odd in k. On the other hand, the integral of
Fourier representation of Diracs delta. Hence, we have seen that
(x, 0) = i3 (x)
(x, 0) = 0, (10.5)
if we think that (x, t) is an appropriate generalized coordinate, a good prospect for the canonical conjugate
momenta is (x, t). Let us examine our hypothesis by calculating the commutator of (x, t) and (x, t). First,
by recalling the explicit form of the creation and annihilation fields Eqs. (5.5, 5.6) page 152 we can obtain + (x)
and (y) at equal times i.e. x0 = y 0 = t, so that the time derivative has the same meaning regardless the point
x or y in which we evaluate it
Z Z
d3 p 1 i(pxp0 t) d3 p 1  0
+
(x, t) = q p a (p) e
; (y, t) = q p a p ei(p yp t)
2p 0 2p 0
(2)3 (2)3
Z Z
+ d3 p ip0 i(pxp0 t) d3 p ip0  i(p yp0 t)
(x, t) = q p a (p) e ; (y, t) = q p a p e (10.6)
(2)3 2p0 (2)3 2p0
270 CHAPTER 10. CANONICAL QUANTIZATION

"Z Z #
h i 1 1 0  i(p yp0 t)
+ 3 ipx 3 ip
(x) , (y) = d p p a (p) e , d p p a p e
(2)3 2p0 2p0
Z Z ipx eip y h i
1 e 
= 3 d3 p d3 p p p ip0 a (p) , a p
(2) 0
2p 2p 0
Z Z
d3 p d3 p ipx ip y 3
 1 d3 p
= ip0 3
p e e p p
= 3 ip0 0 eip(xy)
(2) (2p0 2p0 ) (2) 2p
h i Z 3
i d p ip(xy)
+ (x) , (y) = 3 e
(2) 2

since we are evaluating at equal times then x y = x y. Thus


h i Z 3
i d p ip(xy) i
+ (x, t) , (y, t) = 3 e = 3 (x y)
(2) 2 2

in a similar way we can obtain


h i i h i h i
+ (x, t) , (y, t) = 3 (x y) ; + (x) , + (y) = (x) , (y) = 0
2

now we are ready to obtain the commutation relation between (x) and (y)
h i h i h i h i
+ + + +
(x, t) , (y, t) = (x, t) + (x, t) , (y, t) + (y, t) = (x, t) , (y, t) + (x, t) , (y, t)

h i h i i i
= + (x, t) , (y, t) + (y, t) , (x, t) = 3 (x y) + 3 (y x)
2 2
and we finally obtain h i
(x y, 0) = i3 (x y)
(x, t) , (y, t) =

and combining Eqs. (10.5) and (10.3) for equal time events we find

[ (x, t) , (y, t)] = x y, x0 y 0 = (x y, 0) = 0

Then, we can see that the field and its time derivative satisfy the equal-time commutation relations:
h i
(x, t) , (y, t) = i3 (x y)

h i
[ (x, t) , (y, t)] = (x, t) , (y, t) = 0

Consequently, we can define them as the canonical variables

q (x, t) (x, t) , p (x, t) (x, t) (10.7)

that satisfy the canonical commutation relations (10.1, 10.2).


In the case of a complex scalar field of a particle of spin zero (in which the particle is different from the
antiparticle) we have the commutation relations given by Eqs. (5.32) page 158
h i
(x) , (y) = (x y) ; [ (x) , (y)] = 0 (10.8)

10.1. CANONICAL VARIABLES 271

comparing Eq. (10.8) with Eq. (10.3) it is quite natural to define (x) as the generalized coordinate and use
(y) as a trial field for the conjugate canonical momentum. With a procedure similar to the one carried out for
the slef-charge conjugate scalars, we can show that such a definition is consistent. Therefore, we can define the
free-particle canonical variables as the complex operators

q (x, t) (x, t) ; p (x, t) (x, t) (10.9)

Equivalently by defining
1
(1 + i2 ) ; with 1 = 1 and 2 = 2
2
we can define canonical variables as follows

q k (x, t) = k (x, t) ; pk (x, t) = k (x, t) (10.10)

and they satisfy the commutation relations (10.1, 10.2).

10.1.2 Canonical variables for vector fields


Once again let us start with real vector fields of a particle of spin one that coincides with its antiparticle. The
commutation relations can be obtained by combining Eqs. (6.77, 6.79) page 174
 

[v (x) , v (y)] = g (x y) (10.11)
m2
where we shall use the notation v instead of V because the latter will be used for the fields in the Heisenberg
picture. In this case we can take the canonical variables as follows

v 0 (x, t)
q i (x, t) = v i (x, t) ; pi (x, t) = v i (x, t) + ; i = 1, 2, 3 (10.12)
xi
It can be checked that Eqs. (10.12) satisfy the commutation relations (10.1, 10.2). Let us see it


From Eqs. (10.11, 10.12) and taking into account Eq. (10.5) we have
h i h i  
k n k n kn k n
q (x, t) , q (y, t) = v (x, t) , v (y, t) = g (x y, 0) = 0
m2

now for the commutation relations between q i (x, t) and pi (y, t) we have

1 XZ d3 p h 0
+ +
v (x, t) = (x, t) + (x, t) = q p e (p, ) a (p, ) eipxip t
(2)3 2p0
0
i
+e (p, ) a (p, ) eipx+ip t
(10.13)
i XZ d3 p h 0 k 0
v k (x, t) = +k (x, t) + +k (x, t) = q p p e (p, ) a (p, ) eipxip t
2p 0
(2)3
0
i
+p0 ek (p, ) a (p, ) eipx+ip t (10.14)
Z
v 0 (x, t) i X d3 p h k 0 0 0
i
= q p p e (p, ) a (p, ) eipxip t pk e0 (p, ) a (p, ) eipx+ip t (10.15)
xk 2p0
(2)3
272 CHAPTER 10. CANONICAL QUANTIZATION

i XZ d3 q nh k 0 i 0
pk (y, t) = q p q e (q, ) q 0 ek (q, ) a (q, ) eiqyiq t
2q 0
(2)3
h i 0
o
+ q 0 ek (q, ) q k e0 (q, ) a (q, ) eiqy+iq t
1 X Z d3 p h 0
q k (x, t) = v k (x, t) = +k (x, t) + +k (x, t) = q p ek (p, ) a (p, ) eipxip t
3 2p 0
(2)
0
i
+ek (p, ) a (p, ) eipx+ip t

Z Z h i
n i d3 p d3 q X X n
[q (x, t) , pk (y, t)] = p p e (p, ) q 0 ek (q, ) q k e0 (q, )
(2)3 2p0 2q 0
0 0
h i
eipxip t eiqy+iq t a (p, ) , a (q, )
Z Z h i
i d3 p d3 q X X n k 0 0 k
+ p p e (p, ) q e (q, ) q e (q, )
(2)3 2p0 2q 0
0 0
h i
eipx+ip t eiqyiq t a (p, ) , a (q, )

Z Z h i
n i d3 p d3 q X X n
[q (x, t) , pk (y, t)] = p p e (p, ) q 0 ek (q, ) q k e0 (q, )
(2)3 2p0 2q 0
0 0
eipxip t eiqy+iq t (p q)
Z Z h i
i d3 p d3 q X X n k 0 0 k
p p e (p, ) q e (q, ) q e (q, )
(2)3 2p0 2q 0
0 0
eipx+ip t eiqyiq t (p q)

Z h i
i d3 p X n 0 0
n
[q (x, t) , pk (y, t)] = e (p, ) p 0 k
e (p, ) p k 0
e (p, ) eipxip t eipy+ip t
(2)3 2p0
Z 3 X h i
i d p 0 0
3 0
en
(p, ) p k 0
e (p, ) p 0 k
e (p, ) eipx+ip t eipyip t
(2) 2p

Z h i
i d3 p X n
[q n (x, t) , pk (y, t)] = e (p, ) p 0 k
e (p, ) p k 0
e (p, ) eip(xy)
(2)3 2p0
Z 3 X h i
i d p
en
(p, ) p k 0
e (p, ) p 0 k
e (p, ) eip(xy)
(2)3 2p0

changing p p in the first integral, we have


Z 3 X h i
i d p
I1 en
(p, ) p 0 k
e (p, ) + p k 0
e (p, ) eip(xy)
(2)3 2p0
using the relations (6.88) page 176 we have that

e (p, ) = P e (p, )
en (p, ) = P n e (p, ) = P n n en (p, ) = en (p, )
e0 (p, ) = P 0 e (p, ) = P 0 0 e0 (p, ) = e0 (p, )
10.1. CANONICAL VARIABLES 273

so that I1 becomes
Z h i
i d3 p X n
I1 = e (p, ) p 0 k
e (p, ) p k 0
e (p, ) eip(xy)
(2)3 2p0

then we obtain
Z h i
i d3 p X n
n
[q (x, t) , pk (y, t)] = e (p, ) p 0 k
e (p, ) p k 0
e (p, ) eip(xy)
(2)3 2p0
Z 3 X h i
i d p
en
(p, ) p k 0
e (p, ) p 0 k
e (p, ) eip(xy)
(2)3 2p0

Z (
i d3 p Xh i
n
[q (x, t) , pk (y, t)] = p0 en (p, ) ek (p, ) + en (p, ) ek (p, )
(2)3 2p0
)
X 
pk en (p, ) e0 (p, ) + en (p, ) e0 (p, ) eip(xy)

Z 3
i d p
n
[q (x, t) , pk (y, t)] = 3 0
B eip(xy) (10.16)
(2) 2p
Xh i X 
B p0 en (p, ) ek (p, ) + en (p, ) ek (p, ) pk en (p, ) e0 (p, ) + en (p, ) e0 (10.17)
(p, )

let us evaluate the term B. By using the definition (6.69) page 6.69, and Eq. (6.75) page 173 the term in brackets
yield
Xh i X 
B p0 en (p, ) ek (p, ) + en (p, ) ek (p, ) pk en (p, ) e0 (p, ) + en (p, ) e0 (p, )

h i  
0 nk kn k n0 0n
= p (p) + (p) + p (p) + (p) = 2p0 nk (p) + 2pk n0 (p)
   
0 nk pn pk k n0 pn p0 0 nk
n k
0p p
n 0
kp p
= 2p + 2p g + = 2p + 2p 2p
m2 m2 m2 m2
B = 2p0 nk (10.18)

substituting (10.18) in (10.16) we find


Z
n d3 p  0 nk  ip(xy)
i
[q (x, t) , pk (y, t)] = 2p e
(2)3
2p0
Z
nk 1
= i 3 d3 peip(xy)
(2)
n n
[q (x, t) , pk (y, t)] = ik (x y)

which is the expected commutation relation for two canonical variables. It can also be checked that

[pn (x, t) , pk (y, t)] = 0

from which Equations (10.12) defines consistent canonical variables.




274 CHAPTER 10. CANONICAL QUANTIZATION

The Klein Gordon equation (6.84), the field equation (6.86) page 175, along with Eq. (10.12) permits to write
v0 in terms of the other variables as

p
v0 = (10.19)
m2

therefore, v 0 is not independent, and it not seen as one of the canonical variables. The extension to complex
vector fields (in which particles are distinct to antiparticles), is similar to the case of complex scalar fields
-
Let us obtain (10.19). Separating Eqs. (6.84, 6.86) in its time and coordinate derivatives we obtain

i v i (x) + 0 v 0 (x) = 0 (10.20)


i u 0 2
i v + 0 v = m v (10.21)

with = 0 in Eq. (10.21) and using Eq. (10.20) we have

   
m2 v 0 = i i v 0 + 0 0 v 0 = i i v 0 0 i v i = i i v 0 0 v i
 0   0 
i
 0 i
 i v (x, t) v i (x, t) i v (x, t) i
= i v + 0 v = + = + v (x, t)
xi x0 xi
m2 v 0 = i pi (x, t)

which reproduces Eq. (10.19).

10.1.3 Canonical variables for Dirac fields

For the Dirac field of a non-Majorana particle of spin 1/2 (that is with particle different from antiparticle), we
saw in Eq. (7.123) page 200 that the anticommutator yields

h i
n (x) , n (y) = [( + m) ]n,n (x y)
+

and
[n (x) , n (y)]+ = 0


Note that the anticommutator of n and m does not vanish at equal times, so that we cannot take them as
independent canonical variables (that is we cannot take all of them as q s). A consistent possibility is given by

q n (x) n (x) ; pn (x) in (x) (10.22)

we can see that Eqs. (10.22) satisfy the anticommutation relations (10.1, 10.2). For instance, we have

h i 
n
[q (x, t) , pk (y, t)]+ = n (x, t) , ik
= [( + m) ]nk x y, x0 y 0 = 0
(y, t)
+

[q n (x, t) , pk (y, t)]+ = [ ]nk (x y, t = 0) + mnk x y, x0 y 0 = 0

10.2. FUNCTIONAL DERIVATIVES FOR CANONICAL VARIABLES 275

but according to Eq. (10.5) we have (x y, 0) = 0 so the second term on the right-hand-side vanishes. Let us
define z x y. Hence
Z 3 h i
n 1 d k ikz ikz
[q (x, t) , pk (y, t)]+ = [ ]nk e e =
(2)3 2k0 z 0 =0
 Z 3 h i
1 d k ikz ikz
= [ ]nk ik e + ik e
(2)3 2k0 0
 Z 3 h i  z =0
1 d k
= i [ ]nk 3 k eikz + eikz
(2) 2k0 0
 Z 3 h i z =0
  1 d k
[q n (x, t) , pk (y, t)]+ = i 0 nk 3 k0 eikz + eikz
(2) 2k0
 Z 3 i
 i  1 d k h ikz ikz
+i nk ki e +e
(2)3 2k0

the second integral is odd in k so it vanishes. Further

i 0 = 2 = 1

thus
 Z 3 h i
n 1 d k ik(xy) ik(xy)
[q (x, t) , pk (y, t)]+ = nk e +e
(2)3 2
 
1 1
= nk (x y) + (y x)
2 2
[q n (x, t) , pk (y, t)]+ = kn (x y)

10.2 Functional derivatives for canonical variables


Let us f (x, y) be a function of two sets of variables {x} and {y}. We shall denote as F [f (y)] a functional that
depends on the values of f (x, y) for all x at fixed value of y (thus x appears usually integrated). We define a
bosonic functional as a functional for which each term contains only even number of fermionic fields (i.e. so
that the total spin of the system associated is integer). Let us assume a system of canonical variables i.e. that
satisfy the commutation or anticommutation relations (10.1), (10.2). For a set of canonical variables we can define
a quantum mechanical functional derivative as follows: for an arbitrary bosonic functional F [q (t) , p (t)] at a given
time t, we define

F [q (t) , p (t)]
i [pn (x, t) , F [q (t) , p (t)]] (10.23)
q n (x,t)
F [q (t) , p (t)]
i [F [q (t) , p (t)] , q n (x, t)] (10.24)
pn (x,t)

to understand the sense of such a definition we observe that if F [q (t) , p (t)] were written in such a way that all
q s are to the left of all p s, equations (10.23) and (10.24) become the left and right -derivatives respectively with
respect to q k and pk . As a matter of example, let us assume a functional of the form
Z
 1   
F q (t) , q (t) , p1 (t) , p2 (t) d3 y aq 1 (y, t) p1 (y, t) + bq 1 (y, t) q 2 (y, t) p1 (y, t) p2 (y, t)
2
276 CHAPTER 10. CANONICAL QUANTIZATION

this functional is written such that all q i are on the left of all pi the left-derivative with respect to q i gives
  Z Z
F q 1 (t) , q 2 (t) , p1 (t) , p2 (t) 3 q 1 (y, t) q 1 (y, t) 2
1
= a d y 1
p 1 (y, t) + b d3 y 1 q (y, t) p1 (y, t) p2 (y, t)
q (x, t) q (x, t) q (x, t)
Z Z
= a d y (y x) p1 (y, t) + b d3 y (y x) q 2 (y, t) p1 (y, t) p2 (y, t)
3

 
F q 1 (t) , q 2 (t) , p1 (t) , p2 (t)
= ap1 (x, t) + bq 2 (x, t) p1 (x, t) p2 (x, t) (10.25)
q 1 (x, t)
on the other hand we have
 Z 
3
 1 1 2

i [p1 (x, t) , F [q (t) , p (t)]] = i p1 (x, t) , d y aq (y, t) p1 (y, t) + bq (y, t) q (y, t) p1 (y, t) p2 (y, t)
Z
 
= i d3 y p1 (x, t) , aq 1 (y, t) p1 (y, t)
Z
 
+i d3 y p1 (x, t) , bq 1 (y, t) q 2 (y, t) p1 (y, t) p2 (y, t)
Z
 
= ia d3 y p1 (x, t) , q 1 (y, t) p1 (y, t)
Z
 
+ib d3 y p1 (x, t) , q 1 (y, t) q 2 (y, t) p1 (y, t) p2 (y, t)
Z
i [p1 (x, t) , F [q (t) , p (t)]] = ia d3 y (i) (x y) p1 (y, t)
Z
+ib d3 y (i) (x y) q 2 (y, t) p1 (y, t) p2 (y, t)

i [p1 (x, t) , F [q (t) , p (t)]] = ap1 (x, t) + bq 2 (x, t) p1 (x, t) p2 (x, t) (10.26)
comparing Eqs. (10.25, 10.26) we have
 
F q 1 (t) , q 2 (t) , p1 (t) , p2 (t)
= i [p1 (x, t) , F [q (t) , p (t)]]
q 1 (x, t)
and we can obtain similar relations for the remaining variables. Thus definitions (10.23) and (10.24) are now
clearly motivated.
For an arbitrary cnumber variation of q and p yields
Z X F [q (t) , p (t)] F [q (t) , p (t)]

3 n
F [q (t) , p (t)] = d x q (x, t) + pn (x, t)
n
q n (x,t) pn (x,t)

where q k and pk are bosonic or fermionic and q k , pk are assumed to commute or anticommute with all fermionic
operators respectively, and to commute with all bosonic operators. More general functionals require some extra
signs and equal time commutators in their definitions (10.23) and (10.24).
In particular, H0 is the generator of time-translations on free-particle states in the following sense
q n (x, t) = exp (iH0 t) q n (x, 0) exp (iH0 t) (10.27)
pn (x, t) = exp (iH0 t) pn (x, 0) exp (iH0 t) (10.28)
taking the time derivative of Eq. (10.27) we have
q n (x, t) = iH0 {exp (iH0 t) q n (x, 0) exp (iH0 t)} i {exp (iH0 t) q n (x, 0) exp (iH0 t)} H0
= iH0 q n (x, t) iq n (x, t) H0
q n (x, t) = i [H0 , q n (x, t)]
10.3. FREE HAMILTONIANS 277

and appealing to the definition (10.23) and seen H0 as a functional of the canonical variables we find
H0
q n (x, t) = i [H0 , q n (x, t)] =
pn (x, t)
and a similar relation can be obtained for pn . Therefore, the free particle operators has the time-dependence given
by
H0
q n (x, t) = i [H0 , q n (x, t)] = (10.29)
pn (x, t)
H0
pn (x, t) = i [pn (x, t) , H0 ] = n (10.30)
q (x, t)
which has the form of Hamiltons equations similar to the ones in classical mechanics, in terms of either the Poisson
brackets (replaced by commutators) or Hamiltonian derivatives (replaced by hamiltonian functinal derivatives).

10.3 Free Hamiltonians


An energy term has the form p
a (k, , n) a (k, , n) k2 + m2n (10.31)
p
the operator a (k, , n) a (k, , n) extracts the number of particles in the state given by k, , n; and k2 + m2n
is the energy of each of these (free) states. Therefore, when applied to a multiparticle state, a term of the form
(10.31) provides the energy of all particles in the state k, , n. By summing over all possible states (of course,
integrating on the continuum), provides the total energy of the multiparticle state. Thus, the free-Hamiltonian
can be written as XZ p
H0 = d3 k a (k, , n) a (k, , n) k2 + m2n (10.32)
n,

Such a Hamiltonian can be written in terms of the canonical variables (q, p) at time t.

10.3.1 Free Hamiltonian for scalar fields


For example, the free Hamiltonian (10.32) for a real scalar field (x) can be written (up to a constant term) as
the functional Z  
3 1 2 1 2 1 2 2
H0 = d x p + (q) + m q (10.33)
2 2 2
we can see it by applying Eq. (10.7) and the Fourier expansion of the real scalar field (x) Eq. (5.21), then
equation (10.33) gives
Z  
3 1 2 1 i
 1 2 2
H0 = d x (x, t) + (i (x, t)) (x, t) + m (x, t) (10.34)
2 2 2
from the Fourier expansion (5.21) we have
Z h i
1 1
a (p) ei(pxp t) + a (p) ei(pxp t)
0 0
(x, t) q d3 p p
(2)3 2p0
Z h i
i p0
(x, t) = q d3 p p a (p) eipx + a (p) eipx
(2)3 2p0
Z h i
i pi
i (x) = q d3 p p a (p) eipx a (p) eipx
(2)3 2p0
278 CHAPTER 10. CANONICAL QUANTIZATION

so that H0 in Eq. (10.34) gives

Z Z i
i2 1 3 3 k0 h ikx ikx
H0 = d x d k a (k) e + a (k) e
(2)3 2 2k0
Z
q0 h i
d3 q p a (q) eiqx + a (q) eiqx
2q 0
Z Z i
i2 1 3 3 ki h ikx ikx
+ d x d k a (k) e a (k) e
(2)3 2 2k0
(Z )
3 qi h iqx iqx
i
d q p a (q) e a (q) e
2q 0
Z Z i
1 m2 3 3 1 h ikx ikx
+q d x d k a (k) e + a (k) e
2 2k0
(2)3
Z
3 1 h iqx iqx
i
d q p a (q) e + a (q) e
2q 0

Z Z Z h
1 1 k0 q0
H0 = d3 x d3 k d3 q
p a (k) a (q) ei(k+q)x a (k) a (q) ei(kq)x
(2)3 2 0
2k 2q 0
i
a (k) a (q) ei(kq)x + a (k) a (q) ei(k+q)x
Z Z Z h
1 1 k i qi
d3
x d3
k d3
q p a (k) a (q) ei(k+q)x a (k) a (q) ei(kq)x
(2)3 2 2k0 2q 0
i
a (k) a (q) ei(kq)x + a (k) a (q) ei(k+q)x
Z Z Z h
1 m2 1
+ d 3
x d 3
k d 3
q p a (k) a (q) ei(k+q)x + a (k) a (q) ei(kq)x
(2)3 2 2k0 2q 0
i
+a (k) a (q) ei(kq)x + a (k) a (q) ei(k+q)x

integrating over x we obtain Dirac deltas


Z Z Z
1 1 i(k+q)x i(k 0 +q 0 )t i(k 0 +q 0 )t 1
d3 x ei(k+q)x = ei(k +q )t (k + q)
0 0
3 i(k+q)x 3
3 d xe = 3 d xe e =e 3
(2) (2) (2)
Z
1
d3 x ei(k+q)x = ei(k +q )t (k + q)
0 0
3
(2)
Z Z
1 i(k 0 q 0 )t 1
d3 x ei(kq)x = ei(k q )t (k q)
3 i(kq)x 0 0
3 d xe = e (k q) ; 3
(2) (2)

now since all our four-vectors are on-shell we see that if kkk = kqk then k0 = q 0 , therefore
Z
1 0
3 d3 x ei(k+q)x = e2iq t (k + q)
(2)
Z
1
3 d3 x ei(kq)x = (k q)
(2)
10.3. FREE HAMILTONIANS 279

then the Hamiltonian becomes


Z Z h
1 k0 q0 0
H0 = 3
d k d3 q p a (k) a (q) e2iq t (k + q) a (k) a (q) (k q)
2 2k0 2q 0
0
i
a (k) a (q) (k q) + a (k) a (q) e2iq t (k + q)
Z Z h
1 k i qi 0
d3 k d3 q p a (k) a (q) e2iq t (k + q) a (k) a (q) (k q)
2 2k0 2q 0
0
i
a (k) a (q) (k q) + a (k) a (q) e2iq t (k + q)
Z Z h
m2 1 0
+ d3 k d3 q p a (k) a (q) e2iq t (k + q) + a (k) a (q) (k q)
2 2k 2q0 0
0
i
+a (k) a (q) (k q) + a (k) a (q) e2iq t (k + q)

Z
1 q0q0 h 0 0
i
H0 = d3 q 0 a (q) a (q) e2iq t a (q) a (q) a (q) a (q) + a (q) a (q) e2iq t
2 2q
Z
1 q i qi h 0 0
i
d3 q 0 a (q) a (q) e2iq t a (q) a (q) a (q) a (q) + a (q) a (q) e2iq t
2 2q
Z
m 2 1 h 0 0
i
+ d3 q 0 a (q) a (q) e2iq t + a (q) a (q) + a (q) a (q) + a (q) a (q) e2iq t
2 2q

Z ( 2 " # " 2 #
1 3 q0
2iq 0 t q2 m2 2iq 0 t q0 q2 m2
H0 = d q a (q) a (q) e 0 + 0 + a (q) a (q) e 0+ 0
2 2q 0 2q 2q 2q 0 2q 2q
" 2 # " 2 #)
q0 q2 m2 q0 q2 m2
+a (q) a (q) + + + a
(q) a (q) + +
2q 0 2q 0 2q 0 2q 0 2q 0 2q 0

Z ( " 2 #
1 h i q0 q2 m2
3 2iq 0 t 2iq 0 t
H0 = d q a (q) a (q) e + a (q) a (q) e 0+ 0
2 2q 0 2q 2q
" 2 #)
h i q0 q2 m2
+ a (q) a (q) + a (q) a (q) + 0+ 0
2q 0 2q 2q
Z h i
1
H0 = d3 q a (q) a (q) + a (q) a (q) q 0
2
therefore from (10.33) we obtain the Hamiltonian
Z h i Z  
1 3 0 3 0 1 3
H0 = d k k a (k) , a (k) = d k k a (k) a (k) + (k k) (10.35)
2 + 2
which differs from (10.32) only by a constant divergent term. Such terms only affect the zero of energy and are
not physically observable in the absence of gravity or in the case in which we change the boundary conditions for
the fields. For instance, quantizing in the space between parallel plates instead of infinite space, leads to divergent
significant terms.
Equation (10.35) could be used to test whether a free-field Lagrangian is valid for the given theory. Any
valid free-field Lagrangian must lead to Eq. (10.35) up to a constant. For instance, returning to the case of the
scalar field (x), we could start by finding a free-field Lagrangian that leads to (10.35) for spinless particles or
280 CHAPTER 10. CANONICAL QUANTIZATION

equivalently leads to the free-Hamiltonian (10.32) (up to an unphysical constant). We could check it by starting
with the free-field Lagrangian and applying the Legendre transformation
XZ
L0 [q (t) , q (t)] = d3 x pn (x, t) q n (x, t) H0 (10.36)
n

where pn must be replaced by its expression in terms of q n and q n (and in some cases some auxiliary fields).
In particular from the Hamiltonian (10.33) and the Legendre transformation (10.36) we can derive the free-field
scalar Lagrangian
Z  
1 1 1
L0 = d3 x pq p2 (q)2 m2 q 2
2 2 2
and from the canonical relations (10.7) this free Lagrangian could be written in terms of the field and its derivatives
Z  
3 1 2 1 i
 1 2 2
L0 = d x (i ) m
2 2 2
Z  
3 1 1 i
 1 2 2
= d x (i ) m
2 2 2
Z  
3 1 1 i
 1 2 2
= d x (0 ) (0 ) (i ) m
2 2 2
Z  
3 1 0
 1 i
 1 2 2
= d x (0 ) (i ) m
2 2 2
In summary, from the Hamiltonian (10.33) and from the canonical relations (10.7) we can derive the free-field
Lagrangian given by
Z  
3 1 2 1 2 1 2 2
L0 = d x pq p (q) m q (10.37)
2 2 2
Z  
1 1
L0 = d3 x m2 2 (10.38)
2 2
in terms of the scalar field (x) and its derivatives. Whatever interaction we add to the theory, this free-field
Lagrangian must be taken as the zeroth-order term in a perturbative approach.

10.4 Interacting Hamiltonians


We have formulated only free-field theories in canonical form so far. To do the same for interacting fields as well,
we introduce canonical variables in the Heisenberg picture defined by

Qn (x, t) exp (iHt) q n (x, 0) exp (iHt) (10.39)


Pn (x, t) = exp (iHt) pn (x, 0) exp (iHt) (10.40)

where H is the full Hamiltonian. Note that we are turning on the interaction at time t = 0, so that Q s and P s
coincide with q s and p s at t = 0. This is a similarity transformation that commutes with H and also preserves
all products of canonical variables, for instance

(QP ) exp (iHt) (QP ) exp (iHt) = [exp (iHt) Q exp (iHt)] [exp (iHt) P exp (iHt)]
(QP ) = Q P

form this preservation of products we can see that


  
A BAB 1 F A = F (A) , A , B = [A, B] (10.41)
10.5. THE LAGRANGIAN FORMALISM 281

From these facts, the full Hamiltonian is the same functional of the Heisenberg picture operators in terms of the
canonical variables (q, p)

H [Q, P ] = eiHt H [q, p] eiHt = H [q, p]


Further, the similarity transformations (10.39, 10.40) preserve the commutators or anticommutators with respect
to (q, p), so that they are also canonical variables

[Qn (x, t) , Pn (y, t)] = i3 (x y) nn (10.42)


 
Qn (x, t) , Qn (y, t) = [Pn (x, t) , Pn (y, t)] = 0 (10.43)

with the same procedure that led from Eqs. (10.27, 10.28) to Eqs. (10.29, 10.30) we can start from Eqs. (10.39,
10.40) to obtain the time-dependence of the (interacting) canonical variables

H
Qn (x, t) = i [H, Qn (x, t)] = (10.44)
Pn (x, t)
H
Pn (x, t) = i [Pn (x, t) , H] = n (10.45)
Q (x, t)

As a matter of example, an interacting Hamiltonian for a real scalar field can be constructed by using the free-
particle term (10.33) plus the integral of a scalar interaction density H. Hence, in terms of the Heisenberg picture
the full Hamiltonian reads Z  
3 1 2 1 2 1 2 2
H= d x P + (Q) + m Q + H (Q) (10.46)
2 2 2
in this example, the canonical conjugate to Q has the same expression as for free fields

P = Q (10.47)

but we shall see later that in the general case, the relation between canonical conjugates Pn (x) and the field
variables and their time-derivatives is different from the case of free particle operators. Such relation can be
inferred from Eqs. (10.44) and (10.45).

10.5 The Lagrangian formalism


Now we are challenged to choose appropriate Hamiltonians for realistic theories. A simple form to ensure Lorentz
invariance and other symmetries is by choosing a suitable Lagrangian and use it to derive the Hamiltonian. In
general, we can derive Lagrangians from Hamiltonians and vice versa. Equation (10.36) is the clue for a derivation
in either direction.
In general it is more feasible to explore physically realistic theories by listing possible Lagrangians instead of
Hamiltonians.
The Lagrangian is a functional of a set of fields k (x, t) and their time derivatives (x, t). The conjugate
fields are defined in a way analogous with classical mechanics as the variational derivatives
h i
L (t) , (t)
k (x, t) (10.48)
k (x, t)
where we are using upper case greek letters to indicate that these are interacting instead of free fields. As in
classical mechanics, the field equations can be generated in terms of a Hamiltons variational principle. Thus we
start with the action Z h i
I [] dt L (t) , (t) (10.49)

282 CHAPTER 10. CANONICAL QUANTIZATION

Under an arbitrary variation of (x) the change in I [] yields


Z Z  
3 L k L k
I [] = dt d x (x) + (x)
k (x) k (x)
assuming that k (x) vanishes for t , we can integrate by parts and obtain
Z   Z  
4 L d L k 4 L d
I [] = d x (x) = d x k (x, t) k (x)
k (x) dt k (x) k (x) dt
where we have used the definition (10.48) of conjugate canonical momenta. Now, imposing the stationarity of
the action with respect to all variations k (x) that vanish at t , we see that the necessary and sufficient
condition is that the conjugate canonical momenta satisfy the following equations of motion
h i
L (t) , (t)
k (x, t) = (10.50)
k (x, t)
Now, since the action is the generator of the equations of motion, a natural choice to obtain a hLorentz invariant
i
theory is to make I [] a functional Lorentz scalar. Further, since I [] is a time-integral of L (t) , (t) , we
could expect that L should be a space-integral of an ordinary scalar function of (x) and (x) /x known as
the Lagrangian density L.
h i Z  
L (t) , (t) = d3 x L (x, t) , (x, t) , (x, t) (10.51)

such that the action can be written in a manifestly invariant way in terms of the scalar density L
Z  
4 (x)
I [] = d x L (x) , (10.52)
x
All field theories of elementary particles have Lagrangians of this form.
As in any (classic or quantum) field theory, it is convenient to express the equations of motion in terms of
local quantities (generalized densities) instead of global ones, (e.g. Lagrangian densities instead of Lagrangians).
Thus, we intend to write equation of motion (10.50) in terms of the Lagrangian density L. To do it, we start by
making the variation of m (x) by an amount m (x) and integrating by parts, we find a variation in L given by
Z  
3 L m L m L m
L = d x + +
m (m ) m
Z   
L L L
= d3 x m
+ m
m (m ) m
(homework!! B5 make the integration by parts) so that
Z "  #
L L L m (y, t) L m (y, t)
= d3 y +
k (x, t) m (y, t) (m (y, t)) k (x, t) m (y, t) k (x, t)
Z  
L 3 L L L L
= d y mk (x y) =
k (x, t) m (y, t) (m (y, t)) k (x, t) (k (x, t))
L
a similar procedure can be done for k
and we obtain

L L L
= (10.53)
k (x, t) k (x, t) (k (x, t))
L L
= (10.54)
k
(x, t) k
(x, t)
10.6. FROM LAGRANGIAN TO HAMILTONIAN FORMALISM 283

using definition (10.48) and Eq. (10.54) we have


h i
L (t) , (t) L
k (x, t) =
k
(x, t) k
(x, t)
   
L L
k (x, t) = = (10.55)
x0 k (x, t) x0 k (x, t) /x0

on the other hand, by using the field equations (10.50) along with Eq. (10.53), we find

h i
L (t) , (t) L L
k (x, t) = =
k (x, t) k
(x, t) (k (x, t))
L L
k (x, t) = (10.56)
k (x, t) xi (k (x, t) /xi )

equating Eqs. (10.55, 10.56) we obtain


 
L L L
=
x0 k (x, t) /x0 k (x, t) xi (k (x, t) /xi )

Hence, the field equations (10.50), become

L L
k
= (10.57)
x ( /x ) k

which are the well known Euler-Lagrange equations. If L is a scalar these equations are Lorenz-invariant.
Another important condition for the action (besides being a Lorentz scalar) is that we require it to be real.
This owes to the fact that we want as many fields equations as the number of fields. By splitting the complex fields
into their real and imaginary parts we can figure out I as being a functional of N real fields. If I were complex
with independent real and imaginary parts, we could settle the stationary condition for the real and stationary
parts leading to a set of 2N Euler-Lagrange equations of motion for N fields, which overdetermines the problem
in general. We shall see later that the reality condition for the action also guarantees that we obtain Hermitian
generators associated with several symmetry transformations.

10.6 From Lagrangian to Hamiltonian formalism


We have already said that the Lagrangian formalism is easier to construct realistic Lorentz invariant (and other
symmetries invariant) theories. However, to calculate the Smatrix we should calculate the interaction Hamil-
tonian. Such interaction Hamiltonian (as in the case of the free Hamiltonian) is connected with the interacting
Lagrangian through a Legendre transformation

XZ h i
H= d3 x k (x, t) k (x, t) L (t) , (t) (10.58)
k

Equation (10.48) does not in general allow to express k (x) uniquely in terms of k (x) and k . However, it
can be seen that Eq. (10.58) has null variational derivatives with respect to (x) for any (x) that satisfies Eq.
284 CHAPTER 10. CANONICAL QUANTIZATION

(10.48)

H XZ m (x, t) X
Z
m (y, t) m
= d3 y m (y, t) + d3 y (x, t)
k (x) m k (x) m k (x)
h i
L (t) , (t) Z X
L m (y, t)
d3 y
k (x, t) m
m (y, t) k (x, t)
h i
XZ L (t) , (t)
= d3 y m (y, t) mk (x y)
m k (x, t)
H
= k (x, t) k (x, t) = 0
k (x)

Hence, the Hamiltonian (10.58) is only a functional of k (x) and k . Then, its variational derivatives with respect
to these two sets of variables read
Z
H X m (y, t) L
3
= d y (y, t)
k (x, t) k (x, t)
m k (x, t)
m


Z
X L m (y, t)
d3 y
m (y, t) k (x, t)
m

Z
H X m (y, t)

= k (x, t) + d3 y m (y, t) k (x, t)
k (x, t) m


Z
X L m
d3 y (y, t)
m k (x, t)
m (y, t)

the subscripts denote the variables that are kept fixed in the variational derivatives. From the definition (10.48)
of k , such derivatives become
H L
=
k
(x, t) (x, t)
k

and
H
= k (x, t) (10.59)
k (x, t)
and the equations of motion (10.50) are equivalent to

H
= k (x, t) (10.60)
k (x, t)

so these are the equations of motion (10.50) but in terms of the Hamiltonian. We could a priori identify the
field variables k (x) and their conjugates k with the canonical variables Qk and Pk , and impose the same
canonical commutation relations (10.42, 10.43) on them, such that Eqs. (10.59) and (10.60) are the same as the
Hamiltonian equations of motion (10.44) and (10.45). This is not the case in the most general context as we shall
see. Nevertheless, this association is correct for the simple case of the scalar field with non-derivative coupling.
For this, we consider the Lagrangian density
1 1
L = m2 2 H () (10.61)
2 2
10.6. FROM LAGRANGIAN TO HAMILTONIAN FORMALISM 285

It worths remarking that we are not including a free constant factor in the term (1/2) because such a
constant (if positive) can be absorbed in the normalization of , and a negative constant would lead to an spectrum
not bounded from below. Note that the Lagrangian density (10.61) is obtained by adding a real function H ()
of to the free-field Lagrangian density for the scalar free-field (10.38) (of course replacing the free-field (x) by
the interacting field (x)).
From the Lagrangian density (10.61) we obtain
  h i
L 1 1 2 2 1
= ( ) ( ) m H () = g ( ) ( )
x (k /x ) ( ) 2 2 2 ( )
 
1 ( ) ( ) 1 h i
= g ( ) + g ( ) = g + g
2 ( ) ( ) 2
L 1 h i
k
= 2g = (10.62)
x ( /x ) 2
 
L 1 1 2 2 H ()
= m H () = m2 (10.63)
k 2 2

taking into account the Euler-Lagrange equation (10.57) we equate Eqs. (10.62, 10.63), so that the field equation
becomes
 H ()
 m2 = (10.64)

The canonical conjugate variable associated with can be calculated from the Lagrangian density (10.61) as
 
L 1 0 1 i 1 2 2
= = 0 i m H ()
2 2 2
 
1 1 2
= 0 0 =
2 2

L
= = (10.65)

which is the same association given by Eq. (10.47) if we identify and with the canonical (interacting) variables
Q and P . From the Legendre transformation Eq. (10.58) the full Hamiltonian yields
Z   Z  
3 3 1 2 1 2 1 2 2
H = d x L = d x + () + m + H () (10.66)
2 2 2

which is the Hamiltonian (10.46). In order to interpret the Hamiltonian as an energy is must be bounded from
below. The positivity of the first two terms in Eq. (10.66) shows that the sign postulated in the first term of Eq.
(10.61) was correct. We also require the condition that (1/2) m2 2 + H () must be bounded from below as a
function of . As discusssed below, this example basically validates the Lagrangian (10.61) as a possible theory
for scalar fields (as well as the association of , with the canonical variables Q, P ).
There are however field variables (such as the time component of a vector field or the hermitian conjugate of
a Dirac field) that are not canonical field variables Qn neither have canonical conjugates. Nevertheless, Lorentz
invariance requires that they must be present in the Lagrangians for the vector and Dirac fields respectively.
This type of variables have the feature that they appear in the Lagrangian but not their time-derivatives. The
field variables k whose time-derivatives do not appear in the Lagrangian will be denoted as C r . The remaining
independent field variables are the canonical variables Qn . The canonical conjugates of Qn are given by
h i
L Q (t) , Q (t) , C (t)
Pn (x, t) = (10.67)
Qn (x, t)
286 CHAPTER 10. CANONICAL QUANTIZATION

The pairs Qn , Pk satisfy the canonical commutation relations (10.42)-(10.43). But as we already said, there are
no canonical conjugates associated with C r . Since L/C r = 0, the Hamiltonian (10.58) reads
XZ h i
H= d3 x Pn Qn L Q (t) , Q (t) , C (t)
n

but it is not useful until we express the variables C r and Qk in terms entirely of Q s and P s. For the variables
C r the left-hand side of the equations of motion (10.50) vanishes
h i
L Q (t) , Q (t) , C (t)
0= (10.68)
C r (x, t)

hence the equations of motion associated with C r involve only fields and their first time-derivatives. We shall
treat here simple cases in which Eqs. (10.67) and (10.68) can be solved directly to express the C r and Qk in terms
of Q s and P s. In gauge theories we can solve it by either choosing a particular gauge or by using more modern
covariant methods.

10.6.1 Setting the Hamiltonian for the use of perturbation theory


Once we construct a Hamiltonian as a functional in the Heisenberg picture in terms of canonical variables Qn , Pn ,
we should pass to the interaction picture in order to use perturbation theory. Since the Hamiltonian is time-
independent1 we can express it in terms of Qn and Pn at t = 0, which coincides with the asssociated operators
pn and q n in the interaction picture at t = 0. Then we can express the Hamiltonian in terms of the q s and p s
in the interaction picture, and split it into two parts, one corresponding to a free-Hamiltonian H0 and another
associated with an interaction V . Finally, we can use the time-dependent equations (10.29) and (10.30) as well as
the commutation or anticommutation relations (10.1, 10.2) in order to express the q s and p s in V (t) as linear
combinations of annihilation and creation operators.
By now we shall provide a simple example of this procedure: the scalar field with Hamiltonian (10.66). We
start by splitting it into a free-particle Hamiltonian plus an interaction

H = H0 + V (10.69)
Z  
3 1 2 1 2 1 2 2
H0 = d x + () + m (10.70)
2 2 2
Z
V = d3 x H () (10.71)

where and are evaluated at the same time t. H is independent of t despite H0 and V usually depend on t.
As described above, the next step is to pass to the interaction picture. Thus, by taking t = 0 in Eqs. (10.70)
and (10.71) we can replace , with the interaction picture variables , because according with Eqs. (10.39)
and (10.40) they coincide at t = 0. Now, to calculate the interaction V (t) in the interaction picture we apply the
similarity transformation given by Eq. (2.180) page 103
Z
V (t) = exp (iH0 t) V (t = 0) exp (iH0 t) = d3 x exp (iH0 t) H ( (x, 0)) exp (iH0 t)
Z Z
= d x H (exp (iH0 t) (x, 0) exp (iH0 t)) = d3 x H ( (x, t))
3

1
Recall that in the Heisenberg picture we absorb all time dependence through a similarity transformation with the (complete)
time-evolution operator.
10.6. FROM LAGRANGIAN TO HAMILTONIAN FORMALISM 287

where in the last step we use the fact that the similarity transformations preserve the product [see Eq. (10.41)].
Thus we obtain Z
V (t) = exp (iH0 t) V exp (iH0 t) = d3 x H ( (x, t)) (10.72)

the same transformation must be applied to H0 , but it is clearly kept unchanged


Z  
3 1 2 1 2 1 2 2
H0 = exp (iH0 t) H0 exp (iH0 t) = d x (x, t) + ( (x, t)) + m (10.73)
2 2 2

We can relate the variables and by substituting (10.73) in Eq. (10.29)

H0
(x, t) = = (x, t) (10.74)
(x, t)

Note that this is the same relation as (10.65) but this is not a general feature. The equation of motion for is
given by combining Eqs. (10.73) and (10.30)

H0
(x, t) = = 2 (x, t) m2 (x, t) (10.75)
(x, t)

from Eq. (10.74) we have

(x, t) = 0 (x, t)
(x, t) = 0 0 (x, t) = 0 0 (x, t) (10.76)

and comparing Eqs. (10.75, 10.76) we obtain

0 0 (x, t) = i i (x, t) m2 (x, t)

therefore, by combining Eqs. (10.74) and (10.75) we find the field equation

 m2 (x) = 0

whose general real solution is given by [see procedure in section 4.6, page 149]
Z
3/2 d3 p h ipx i
(x) = (2) p e a (p) + eipx a (p) (10.77)
2p0
p
with p0 = p2 + m2 , and a (p) are some unknown operator function of p to be determined. The canonical
conjugate is given by Eq. (10.74)
Z r
3/2 p0 h ipx i
(x) = (x) = i (2) d3 p e a (p) eipx a (p) (10.78)
2

then we adjust the unknown operators a (p) and a (q) in order to obtain the desired commutation relations

[ (x, t) , (y, t)] = i3 (x y) (10.79)


[ (x, t) , (y, t)] = [ (x, t) , (y, t)] = 0 (10.80)



288 CHAPTER 10. CANONICAL QUANTIZATION

by imposing the first of conditions (10.80) on the expansions (10.77) and (10.78), we obtain
" Z Z
d3 p n ipx o d3 q n iqy o
0 = [ (x, t) , (y, t)] = (2)3/2 p e a (p) + eipx a (p) , (2)3/2 p e a (q) + eiqy a (q)
2p0 2q 0
Z Z
d3 p d3 q hn ipx o n oi
0 = (2)3 p p e a (p) + eipx a (p) , eiqy a (q) + eiqy a (q)
2p0 2q 0
Z Z
d3 p d3 q n ipx iqy h i
0 = (2)3 p p e e [a (p) , a (q)] + eipx eiqy a (p) , a (q)
2p0 2q 0
h i h io
+eipx eiqy a (p) , a (q) + eipx eiqy a (p) , a (q)

I
1 + I2 + I3 + I4

interchanging the dummy variables q p in I


2 we have
Z Z
3 d3 p d3 q ipx iqy h i
I2 (2) p p e e a (p) , a (q)
2p0 2q 0
Z Z
3
d q d p iqx ipy h
3 i
= (2)3 p p e e a (q) , a (p)
2q 0 2p0
Z Z
d3 p d3 q iqx ipy h i
= (2)3 p p e e a (p) , a (q)
2p0 2q 0
Z Z
3 d3 p d3 q  ipx iqy h i
I2 + I3 = (2) p p e e eiqx eipy a (p) , a (q)
2p0 2q 0
applying the second of conditions (10.80) on the expansions (10.77) and (10.78), we have
" Z r Z r
p 0 n o q 0 n iqy
0 = [ (x, t) , (y, t)] = i (2)3/2 d3 p eipx a (p) eipx a (p) , i (2)3/2 d3 q e a (q) e
2 2
Z r Z r
3 p0 q 0 hn ipx o n oi
= (2) 3
d p 3
d q e a (p) eipx a (p) , eiqy a (q) eiqy a (q)
2 2
Z r Z r
p 0 q n ipx iqy h
0 i h i
0 = (2)3 d3 p d3 q e e a (p) , a (q) eipx eiqy a (p) , a (q)
2 2
h io
e e [a (p) , a (q)] + eipx eiqy a (p) , a (q)
ipx iqy

by imposing condition (10.79) on the expansions (10.77) and (10.78), we obtain

" Z Z r
3 3/2 d3 p n ipx ipx
o
3/2 3 q 0 n iqy
i (x y) = [ (x, t) , (y, t)] = (2) p e a (p) + e a (p) , i (2) d q e a (q)
2p0 2
Z Z r
3 d3 p q 0 hn ipx o n oi
= i (2) p d3 q e a (p) + eipx a (p) , eiqy a (q) eiqy a (q)
2p0 2
ZZ r
d3 p
3 q 0 n ipx iqy h i
3
i (x y) = i (2) p d3 q e e [a (p) , a (q)] eipx eiqy a (p) , a (q)
2p0 2
h i h io
+eipx eiqy a (p) , a (q) eipx eiqy a (p) , a (q)

????
10.7. GAUGES OF THE LAGRANGIAN FORMALISM 289

Therefore, to satisfy Eqs. (10.79, 10.80) the unknown operators must fulfill the commutation relations (Home-
work!! B6) h i
a (p) , a (q) = 3 (p q) ; [a (p) , a (q)] = 0 (10.81)

we have already seen that expansion (10.73) leads to the usual free Hamiltonian Eq. (3.23) page 121, up to an
unphysical constant. It is important to emphasize that this should not be considered as an alternative derivation
of Eqs. (10.77), and (10.81) but rather as a proof of validation of the first two terms in Eq. (10.61) as the correct
free-particle Lagrangian for a real scalar field. The way is paved to use perturbation theory to calculate the
Smatrix, by taking Eq. (10.72) as V (t), with the scalar field (x) given by Eq. (10.77) where a (p) and a (p)
obeys the commutation relations (10.81).

10.7 Gauges of the Lagrangian formalism


Since the Hamiltons stationary principle usually requires an integration by parts and we usually assume that the
fields vanish at infinity, Lagrangian densities that differ only by total derivatives F do not contribute to the
action leading to the same field equations. It is also clear that a space derivative term F in the Lagrangian
density does not contribute to the Lagrangian so that it does not affect the quantum field theory defined by such
a Lagrangian. It is important however to emphasize that it is not true anymore when the fields do not vanish in
the boundaries (it happens in some theories in which we define boundaries not at infinity). Furthermore, a time
derivative 0 F 0 in the Lagrangian density does not affect the quantum structure of the theory, though such a fact
is less apparent. We shall see it by considering the effect of adding a more general term to the Lagrangian in the
form
Z
L (t) = d3 x Dn,x [Q (t)] Q (x, t) (10.82)

where D is an arbitrary functional of the values of Q at a given time that in general depend on n and x. According
with Eq. (10.48) the formula for the canonical conjugate momenta P (t) change as a functional of Q (t) and Q (t)
as
L (t)
Pn (x, t) = = Dn,x [Q (t)] (10.83)
Qn (x, t)
But taking into account Eqs. (10.82, 10.83) we see that there is no change in the Hamiltonian expressed as a
functional of Q (t) and Q (t) [so that the change in variables only appears through the canonical momenta Pn ]
h i Z
H Q (t) , Q (t) = d3 x Pn (x, t) Qn (x, t) L (t)
Z Z
= d x Dn,x [Q (t)] Q (x, t) d3 x Dn,x [Q (t)] Q (x, t) = 0
3 n

h i Z
H Q (t) , Q (t) = d3 x Pn (x, t) Qn (x, t) L (t) = 0 (10.84)

Therefore, the Hamiltonian written as a functional of the old canonical variables Qn and Pn do not change either.
We should notice however that the Hamiltonian is not the same functional of the new canonical variables Qn and
Pn + Pn , and in a theory described by the new Lagrangian density L + L it is the new canonical set Qn and
Pn + Pn instead of the old one Qn and Pn , that satisfies the canonical commutation relations. The commutators
of Qn with Qk and of Qn with Pk obey the usual commutation relations, but now the commutators of Pn with Pk
are given by
290 CHAPTER 10. CANONICAL QUANTIZATION

[Pn (x, t) , Pm (y, t)] = [Pn (x, t) + Pn (x, t) , Pm (y, t) + Pm (y, t)] [Pn (x, t) , Pm (y, t) + Pm (y, t)]
[Pn (x, t) + Pn (x, t) , Pm (y, t)] + [Pn (x, t) , Pm (y, t)]
Dn,x [Q (t)] Dm,y [Q (t)]
[Pn (x, t) , Pm (y, t)] = i m
+i (10.85)
Q (y, t) Qn (x, t)
which is in general non-null. Nevertheless, if the additional term in the Lagrangian is a total time-derivative
Z
dG G [Q (t)] n
L = = d3 x Q (x, t) (10.86)
dt Qn (x, t)

we see that D in Eq. (10.82) acquires a particular form

G [Q (t)]
Dn,x [Q] = (10.87)
Qn (x, t)

and in that case the commutator (10.85) vanishes, from which the variables Qn and Pk satisfies the usual
commutation relations. Thus we have shown that a change of the form (10.82) in the Lagrangian does not change
the form of the Hamiltonian as a functional of Qn and Pk , and taking into account that the commutation relations
of those variables do not change, we conclude that the addition to the Lagrangian of the term (10.82) does lead
to the same quantum field theory. Thus, different Lagrangian densities obtained from each other by partial
integration are equivalent in both quantum and classical field theories.

10.8 Global symmetries


Like in classical mechanics, Lagrangian formalism is very suitable in quantum mechanics to implement symmetry
principles. It has to do with the fact that the dynamical equations of motion has the form of a variational principle
in the Lagrangian formalism i.e. the Hamiltons variational principle.
We shall start with an infinitesimal transformation of the fields

k (x) k (x) + iF k (x) (10.88)

that keeps the action (10.49) invariant


Z Z
I [] I [] k
0 = I = d4 x k k (x) = i d4 x k F (x) (10.89)
(x) (x)

when is constant we call them as global symmetries. The factor F k (x) depends in general on the fields and their
derivatives at x. Equation (10.89) is automatically satisfied for all infinitesimal variations of the fields if those
fields obey the dynamical equations. Thus we define an infinitesimal symmetry transformation as one that leaves
the action invariant even when the dynamical equations are not satisfied.
Let us now consider a local transformation, that is, one in which is a function of x

k (x) k (x) + i (x) F k (x) (10.90)

in this case the variation of the action does not vanish, but it acquires the form
Z
(x)
I = d4 x J (x) (10.91)
x
for it to vanish when becomes constant. Here J is a functional of the fields and its first-derivatives. If we
assume that the fields in I [] satisfy the field equations we obtain that I [] is stationary with respect to arbitrary
10.8. GLOBAL SYMMETRIES 291

field variations that vanish at large space-time distances, including local variations of the type (10.90), hence in
that case the variation (10.91) should vanish. Integrating by parts we find that J must satisfy a conservation
law (Homework!! B7)
J (x)
0= (10.92)
x
it is straightforward that
Z
dF
0= , with F d3 x J 0 (x) (10.93)
dt
so that there is one conserved current J and one constant of the motion F for each independent infinitesimal
symmetry transformation. This leads to the fact that continuous symmetries imply conservation laws2 , statement
usually called Noethers theorem.
For some symmetry transformations the Lagrangian is invariant which is a condition stronger than the invari-
ance of the action. Good examples are the translations and rotations in space and also isospin transformations
as well as other internal symmetry transformations (but it is not the case for general Lorentz transformations).
For the stronger condition of invariance of the Lagrangian we correspondingly obtain a stronger form of the
conservation law, i.e. we can obtain an explicit formula for the conserved quantities F .
To see it, let us consider a field variation of the type (10.90) with (x) depending on t but not on x. Under
this restricted transformation, the variation of the action becomes

h i h i
Z h i Z Z L (t) , (t) L (t) , (t)
I = dt L (t) , (t) = dt d3 x k
(x, t) + k
(x, t)
k (x, t) k (x, t)
h i h i
Z Z L (t) , (t) L (t) , (t) d h i
I = i dt d3 x (t) F k (x, t) + (t) F k (x, t)
k (x, t) k
(x, t) dt
h i
Z Z L (t) , (t)
I = i dt d3 x (t) F k (x, t)
k (x, t)
h i h i
L (t) , (t) L (t) , (t) d k
+ (t) F k (x, t) + (t) F (x, t) (10.94)
k (x, t) k (x, t) dt

and the condition of invariance of the Lagrangian under this transformation when becomes constant gives
h i h i
Z L (t) , (t) L (t) , (t) d
0= d3 x F k (x, t) + F k (x, t) (10.95)
k (x, t) k (x, t) dt

assuming that the transformations F k (x, t) are independent we obtain


h i h i
L (t) , (t) L (t) , (t) d
F k (x, t) = F k (x, t) (10.96)
k (x, t) k (x, t) dt

2
The condition of continuity is essential because we are using infinitesimal transformations. However, in practice we often find
conservation laws associated with discrete symmetries. Nevertheless, the latter case is not guaranteed by Noethers theorem.
292 CHAPTER 10. CANONICAL QUANTIZATION

substituting (10.96) in Eq. (10.94) we obtain


h i
Z Z
L (t) , (t) d k
I = i dt d3 x (t) F (x, t)
k (x, t) dt
h i h i
L (t) , (t) L (t) , (t) d
+ (t) F k (x, t) + (t) F k (x, t)
k (x, t) k (x, t) dt

so regardless the field equations are satisfied or not, the variation of the action gives
h i
Z Z L (t) , (t)
I = i dt d3 x (t) F k (x, t) (10.97)
k (x, t)

Now let us rewrite Eq. (10.91) recalling that depends on t but not on x

Z Z   Z Z
3 0 (t) i (t)
I = dt d x J (x) + J (x) = dt (t) d3 x J 0 (x)
x0 xi
Z
I = dt (t) F (10.98)

and comparing Eqs. (10.97) and (10.98) we obtain


h i
Z L (t) , (t)
F = i d3 x F k (x, t) (10.99)
k (x, t)

so we have obtained an explicit expression for the conserved generalized charge F . As a matter of consistency, by
taking the symmetry condition (10.95) it could be checked that F is time-independent for any fields that satisfy
the dynamical equations (10.50).
An even stronger condition that posseses some symmetry transformations (e.g. isospin symmetry) leave the
action, Lagrangian, and Lagrangian density invariant. In that case we can additionally obtain an explicit formula
for the current J (x). The action (10.52) (written in terms of the Lagrangian density), has a variation under the
transformation (10.90) with the local infinitesimal parameter (x) of the form

Z  
L ( (x) , (x)) k
4 L ( (x) , (x))  k
I [] = i d x (x) + (x)
k (x) ( k (x))
Z  i
4 L ( (x) , (x)) k L ( (x) , (x)) h k
I [] = i d x F (x) (x) + F (x) (x)
k (x) ( k (x))

the invariance of the Lagrangian density when becomes constant, provides the condition

L ( (x) , (x)) k L ( (x) , (x))


0= k
F (x) + F k (x) (10.100)
(x) ( k (x))

so for arbitrary fields, the variation of the action gives


Z
L ( (x) , (x)) k
I [] = i d4 x F (x) (x)
( k (x))
10.9. CONSERVED QUANTITIES IN QUANTUM FIELD THEORIES 293

and comparing with (10.91) we see that


L
J = i Fk (10.101)
(k /x )
As a matter of consistency we can check that the symmetry condition (10.100) leads to the continuity equation
J = 0 when the fields obey the Euler-Lagrange equation (10.57) [homework!! B8]. It can also be checked that
the time component of the current (10.101) integrated over spatial coordinates has the value calculated above Eq.
(10.99) in agreement with Eq. (10.93) [homework!! B9].

10.9 Conserved quantities in quantum field theories


The developments done so far are equally valid for classical and quantum field theories. The quantum properties of
the conserved quantities F are more apparent when they come from symmetries of the Lagrangian (not necessarily
of the Lagrangian density) that transform the canonical fields Qn (x, t) (i.e. the ones of the k whose time-
derivatives appear in the Lagrangian) into xdependent functionals of themselves at the same time. Those types
of transformations yield

F n (x, t) = F n [Q (t) ; x] (10.102)


we shall see later that infinitesimal translations, rotations as well as infinitesimal internal symmetry transforma-
tions have the form of Eqs. (10.88) and (10.102), with F n being a linear functional of the Q s, though we shall not
require the linearity here. For those symmetries in quantum mechanics besides being conserved, such operators
are generators of the symmetry.
To see it, we first observe that when k is a canonical variable Qk it has an associated canonical momentum
Pk given by the functional derivative L/k , while when k is an auxiliary field C r , such a functional derivative
is null. From these facts Eq. (10.99) can be written as
Z Z
F = i d x Pn (x, t) F (x, t) = i d3 x Pn (x, t) F n [Q (t) ; x]
3 n
(10.103)

To calculate the commutator (not the anticommutator) of F with the canonical field Qk (x, t) at an arbitrary
time t, we can use Eq. (10.93) to evaluate F as a functional of the Q s and P s (at the time t), and then use
the equal-time canonical commutation relations (10.42) and (10.43) and obtain [homework!! Eq. (10.104) (with
commutator) should be true also for fermionic variables Q, P ]
 Z  Z
[F, Q (x, t)] = i d y Pm (y, t) F [Q (t) ; y] , Q (x, t) = i d3 y [Pm (y, t) F m [Q (t) ; y] , Qn (x, t)]
n 3 m n

Z Z
= i d3 y Pm (y, t) [F m [Q (t) ; y] , Qn (x, t)] i d3 y [Pm (y, t) , Qn (x, t)] F m [Q (t) ; y]
Z

= i d3 y i3 (x y) nm F m [Q (t) ; y]

[F, Qn (x, t)] = F n (x, t) (10.104)


where we have assumed that for Qn being bosonic or fermionic the variation F n is bosonic or fermionic respectively,
from which F is bosonic. Nevertheless, in the case of some symmetries known as supersymmetries, for which F
is fermionic and (10.104) is an anticommutator if Qn is fermionic as well. It is in the sense of Eq. (10.104) that
F can be considered the generator of the transformation with (10.102) and (10.103). The canonical commutation
rules also lead to [homework!! B10 do it for fermionic operators Q, P ]
 Z  Z
[F, Pn (x, t)] = i d3 y Pm (y, t) F m [Q (t) ; y] , Pn (x, t) = i d3 y Pm (y, t) [F m [Q (t) ; y] , Pn (x, t)]

Z
= i d3 y [Pm (y, t) F m [Q (t) ; y] , Pn (x, t)]
294 CHAPTER 10. CANONICAL QUANTIZATION

and using the definition (10.23) page 275 we finally obtain


Z
F m (Q (t) ; y)
[F, Pn (x, t)] = d3 y Pm (y, t) (10.105)
Qn (x, t)

when F k is a linear functional, equation (10.105) shows that Pn transforms contragradiently with respect to Qn ,
it justifies the notation of indices followed so far for Qn and Pn .

10.9.1 Space-time translation symmetry


We shall first examine the space-time translation symmetry

k (x) k (x + ) = k (x) + k (x) (10.106)


equation (10.106) has the structure of Eq. (10.88) with four independent parameters with four associated
transformation functions

Fk = i k (10.107)
so that we have four independent conserved currents which are conveniently grouped into the energy-momentum
tensor T
T = 0
where we can derive four generalized charges which are global conserved quantities written as spatial integrals
of the time components of each of the four translation currents denoted by P .
Z
d
P = d3 x T 0 ; P = 0 (10.108)
dt

In order to avoid confusion with the canonical variables Pn (x, t) we have used greek letters for the subscript of
the four generalized charges.

Spatial translations
The Lagrangian
h L (t)i is invariant under spatial translations (there is an integration over all spatial components to
define L (t) , (t) ), so that the spatial components Pk of Eq. (10.108) acquires an explicit form according with
Eq. (10.99) or equivalently Eq. (10.103). Hence, combining Eqs. (10.103) and (10.107) the spatial components of
P acquires the form (we use for a moment the notation Pk to avoid confusion with the canonical variable Pk )
Z Z
Pk = i d3 x Pn (x, t) Fkn [Q (t) ; x] = i d3 x Pn (x, t) [ik n ]

Z
P d3 x Pn (x, t) Qn (x, t) (10.109)

from the canonical equal-time comutation relations (10.42) and (10.43) we can find the commutator of P with the
canonical variables
 Z  Z
[P, Qn (x, t)] = d3 y Pm (y, t) Qm (y, t) , Qn (x, t) = d3 y [ Pm (y, t) Qm (y, t) , Qn (x, t)]

Z Z
= d y [ Pm (y, t) , Q (x, t)] Q (y, t) = i d3 y 3 (x y) mn Qm (y, t) = iQn (x, t)
3 n m
10.9. CONSERVED QUANTITIES IN QUANTUM FIELD THEORIES 295

 Z  Z
[P, Pn (x, t)] = d3 y Pm (y, t) Qm (y, t) , Pn (x, t) = d3 yPm (y, t) [Pn (x, t) , Qm (y, t)]

Z
= i d3 yPm (y, t) 3 (x y) nm Pn (x, t)

[P, Qn (x, t)] = iQn (x, t) (10.110)


[P, Pn (x, t)] = iPn (x, t) (10.111)

from Eqs. (10.110, 10.111), it can be shown that for a function of the canonical variables Qk , Pk that does not
also depend on x, we have
[P, (x)] = i (x) (10.112)
from these results we conclude that P is the generator of spatial translations.

Time translations
In contrast to the space translations, time translations do not leave the Lagrangian L (t) invariant. But we
already know that its generator is the Hamiltonian P 0 H. And we also know that the Hamiltonian satisfies the
commutation relation

[H, (x, t)] = i (x, t)


for any function (x, t) of Heisenberg picture operators.

10.9.2 Conserved currents and Lagrangian densities for space-time symmetries


By assuming that the Lagrangian is a space integral of a Lagrangian density, we can also obtain an explicit formula
for the energy-momentum tensor T . Nevertheless, the Lagrangian density L is not invariant under space-time
translations, so we cannot apply Eq. (10.101). Hence, we shall go back to examine the change in the action under
a space-time dependent translation of the form

k (x) k (x + (x)) = k (x) + (x) k (x) (10.113)

the variation of the action under transformation (10.113) reads


Z  h i
4 L k L k
I [] = d x + (10.114)
k ( k )

from the Euler-Lagrange equations (10.57) we see that the terms proportional to add up to L, so that the
variation (10.114) becomes
Z    
4 L k L k L k
I [] = d x + + ( )
k ( k ) ( k )
Z       
4 L k L k L k
= d x + + ( )
k ( k ) ( k )
Z         L    
4 L k L k k L k
= d x + + ( )
k ( k ) ( k ) ( k )
Z            
4 L L k L k L k
= d x + + ( )
k ( k ) ( k ) ( k )
Z      
4 L k L k
= d x + ( )
( k ) ( k )
296 CHAPTER 10. CANONICAL QUANTIZATION

Z  
L L
I [] = d4 x + k
x ( k )
and integrating by parts we are led to the form of Eq. (10.91) (Homework!! B11)
Z
I [] = d4 x T

with the associated currents


L
T = L k (10.115)
( k )
by consistency we can see that the spatial components of Eq. (10.108) coincide with the formula (10.109) for P,
while for = 0 Eq. (10.108) gives the well known expression for the Hamiltonian
Z " #
X
0 3 n
H P = P0 = d x Pn Q L (10.116)
n

10.9.3 Additional symmetry principles


In some theories there are other symmetry principles that provides the invariance of the action under a set of
coordinate independent transformations of the canonical fields

Qn (x) Qn (x) + ia (ta )n m Qm (x) (10.117)

along with a set of transformations of the auxiliary fields C r

C r (x) C r (x) + ia (a )r s C s (x) (10.118)

where ta and a denote Hermitian matrices associated with representations of the Lie algebra of the symmetry
group, with sum over repeated group indices a, b, . . .. For example, we shall see that for electrodynamics we have
a symmetry of this kind with a single diagonal matrix tn m , with the charges associated with each field in the
diagonal.
For these kind of symmetries we can find a set of conserved currents Ja

Ja = 0 (10.119)

where the time components are the densities of a set of time-independent generalized charge operators (quantized
charges) Z
Ta = d3 x Ja0 (10.120)

As before, when the Lagrangian as well as the action is invariant under the transformation (10.117) then Eq.
(10.99) provides an explicit expression for the generalized charged operator
Z
Ta = i d3 x Pn (x, t) (ta )n m Qm (x, t) (10.121)

in this case the equal-time commutation relations give

[Ta , Qn (x)] = (ta )n m Qm (x) (10.122)


m
[Ta , Pn (x)] = + (ta ) n Pm (x) (10.123)
10.9. CONSERVED QUANTITIES IN QUANTUM FIELD THEORIES 297

where ta is diagonal. These relations show that Qn and Pn lower and raise the value of Ta by an amount equal to
(ta )n n . From the previous results, we can calculate the commutator between the generalized charge operators
 Z Z 
3 n m 3 k p
[Ta , Tb ] = i d x Pn (x, t) (ta ) m Q (x, t) , i d y Pk (y, t) (tb ) p Q (y, t)

Z Z
[Ta , Tb ] = d3 x d3 y (ta )n m (tb )k p [Pn (x, t) Qm (x, t) , Pk (y, t) Qp (y, t)] (10.124)

let us evaluate the commutators

C [Pn (x, t) Qm (x, t) , Pk (y, t) Qp (y, t)]


= Pn (x, t) [Qm (x, t) , Pk (y, t) Qp (y, t)] + [Pn (x, t) , Pk (y, t) Qp (y, t)] Qm (x, t)
= Pn (x, t) Pk (y, t) [Qm (x, t) , Qp (y, t)] + Pn (x, t) [Qm (x, t) , Pk (y, t)] Qp (y, t)
+Pk (y, t) [Pn (x, t) , Qp (y, t)] Qm (x, t) + [Pn (x, t) , Pk (y, t)] Qp (y, t) Qm (x, t)
C = i3 (x y) mk Pn (x, t) Qp (y, t) i3 (x y) np Pk (y, t) Qm (x, t) (10.125)

substituting (10.125) in (10.124) we obtain


Z Z
[Ta , Tb ] = i d x 3
d3 y (ta )n m (tb )k p 3 (x y) [mk Pn (x, t) Qp (y, t) np Pk (y, t) Qm (x, t)]
Z n o
= i d3 x (ta )n m (tb )m p Pn (x, t) Qp (x, t) + (ta )n m (tb )k n Pk (x, t) Qm (x, t)
Z h i
[Ta , Tb ] = i d3 x Pm (ta )m n (tb )n k Qk + Pn (tb )n k (ta )k m Qm (10.126)

if we define P, Q as matrix arrangements (column vectors) for Pm , Qm , and ta the matrix associated with the
transformation (10.117) we can write
Z h i
[Ta , Tb ] = i d3 x P e ta tb Q + P e tb ta Q
Z n o
[Ta , Tb ] = i d3 x P e [ta , tb ] Q

and hence, if the matrices ta form a Lie algebra with structure constants fab c

[ta , tb ] = ifab c tc (10.127)


Z n o Z
[Ta , Tb ] = i 3
d x e
P [ta , tb ] Q = fab c e tc Q = fab c Tc
d3 x P

where we have used (10.121) in the last step. Hence, the quantum operators Ta form the same Lie algebra as the
matrices ta
[Ta , Tb ] = ifab c Tc (10.128)
which shows that the operators (10.121) are properly normalized to be considered as generators of the symmetry
group.
Once again when there is a Lagrangian density that is invariant under the transformations (10.117) and
(10.118) we can obtain an expression for the current J associated with the global symmetry. From Eq. (10.101)
and using Eqs. (10.117) and (10.118) we have

L L
Ja i n
(ta )n m Qm i (a )r s C s (10.129)
(Q /x ) (C r /x )
298 CHAPTER 10. CANONICAL QUANTIZATION

note that the construction of the current requires the auxiliary fields C r as well as their transformation properties
(10.118), despite Ta in Eq. (10.121) only depends on the fields Q and their canonical conjugates P .
The explicit expression (10.129) for the current lead to other useful commutation relations. In particular,
when the Lagrangian density does not contain time-derivatives of the auxiliary fields, we obtain from (10.129)

L L
Ja0 i n 0
(ta )n m Qm = i (ta )n m Qm
(Q /x ) Qn
Ja0 = iPn (ta )n m Qm

then we can derive the equal-time commutators of general fields not only with the symmetry operators Ta , but
also with the densities Ja0
 
Ja0 (x, t) , Qn (y, t) = 3 (x y) (ta )n m Qm (x, t) (10.130)
  n
Ja0 (x, t) , Pm (y, t) 3
= (x y) (ta ) m Pn (x, t) (10.131)

Moreover, if the auxiliary fields are constructed such that they are local functions of Q s and P s, such that
they transform according with a representation of the Lie symmetry algebra with generators a , we also have
commutation relations involving auxiliary fields of the form
 
Ja0 (x, t) , C r (y, t) = 3 (x y) (a )r s C s (x, t) (10.132)

It is usual to shorten the set of commutation relations (10.130) and (10.132) as follows
h i
Ja0 (x, t) , k (y, t) = 3 (x y) (ta )k m m (x, t)

Commutation relations of the type given by (10.130)-(10.132) are used to derive some relations for matrix elements
related with the current J known as Ward identities.

10.9.4 Conserved current for a two scalars Lagrangian


As a matter of example, let us take the case of two real fields with the same mass, with a Lagrangian density
given by

1 1 1 1 
L = 1 1 m2 21 2 2 m2 22 H 21 + 22 (10.133)
2 2 2 2
this Lagrangian density is invariant under a transformation of the type (10.117) given by3

1 1 1 = 2 ; 2 2 2 = 1 (10.134)
3
This transformation correspond to an infinitesimal rotation in two-dimensions as we can see from
    
cos sin 1 1
=
sin cos 2 2

at first order cos 1, sin , so that


      
1 1 1 1 2
= =
2 1 2 2 + 1
10.9. CONSERVED QUANTITIES IN QUANTUM FIELD THEORIES 299

we see it as follows
1 1 1 1 2 2 
L = 1 1 2 2 m2 2
1 m 2 H 1 + 2
2 2
2 2 2 2
1 1 1
= (1 2 ) (1 2 ) (2 + 1 ) (2 + 1 ) m2 (1 2 )2

2 2 2
1 2  
m (2 + 1 )2 H (1 2 )2 + (2 + 1 )2
2
1 1 1 1
L = 1 1 + 1 2 + 2 1 2 2 2
2 2 2 2
1 1 1 1
2 2 2 1 1 2 2 1 1
2 2 2 2
1 2 2 21 2 2 1 2 2 1
m 1 + m 1 2 m 2 m 2 m 2 1 2 m2 21
2 2
2 2 2  2
H 2 22 21 2 + 21 + 2 21 + 21 2 + 22

1 1 1 1
L = 1 1 2 2 2 2 2 2 1 1
2 2 2 2
1 2 2 1 2 2 21 2 2 21 2 2

m 1 m 2 m 2 m 1 H 21 + 22 + 2 21 + 22
2  2 2 2
2
L = L+O

comparing Eq. (10.134) with Eq. (10.117) we have

1 1 2 ; 2 = 2 + 1
Q1 (x) Q1 (x) + ia (ta )1 m Qm (x) = Q1 (x) Q2
Q1 (x) + it1 m Qm (x) = Q1 (x) Q2
Q1 (x) + it1 1 Q1 (x) + it1 2 Q2 (x) = Q1 (x) Q2
it1 1 = 0 ; it1 2 = 1

similarly

Q2 (x) + it2 m Qm (x) = Q2 (x) + Q1


it2 1 = 1 ; it2 2 = 0

note that despite we are transforming two fields, we only have one independent infinitesimal parameter and so
only one matrix of the form ta . The matrix ta yields

t1 1 = t2 2 = 0 ; t1 2 = i ; t2 1 = i (10.135)

in this case there is only one current that we obtain by replacing (10.135) in (10.129)

L L L
Ja i n
(ta )n m Qm i r
(a )r s C s = i tn m Qm
(Q /x ) (C /x ) (Qn /x )
L L
= i 1
t1 1 Q1 i t1 2 Q2
(Q /x ) (Q1 /x )
L L
i t2 1 Q1 i t2 2 Q2
(Q2 /x ) (Q2 /x )
L L L L
= i t1 2 2 i t2 1 1 = 2 1
( 1 ) ( 2 ) ( 1 ) ( 2 )
300 CHAPTER 10. CANONICAL QUANTIZATION

and using the Lagrangian density (10.133) we finally obtain


   
12 1 1 12 2 2
J = 2 1 = 2 1 + 1 2
( 1 ) ( 2 )

in summary Lagrangian (10.133) posseses a symmetry transformation (10.134) that lead to the following conserved
current
J = 2 1 1 2

10.10 Lorentz invariance


We shall see that the Lorentz invariance of the Lagrangian density implies the Lorentz invariance of the Smatrix.
We proceed in three steps (a) We generate the currents and charges generated by Lorentz symmetry. (b) We
construct the generators of the Lorentz group by using the previous currents and (c) We prove that such generators
commute with the Soperator.

10.10.1 Currents and time-independent operators


We start as usual with infinitesimal Lorentz transformations

= + ; =

From our previous analysis the invariance of the action under these transformations lead to a set of conserved
currents that we denote as M
M = 0 ; M = M
i.e. one current for each independent component of (six independent components owing to the antisymmetry
of ). We have generalized charges in the form of time-independent tensors of the form
Z
d
J d3 x M0 ;

J =0
dt

The tensors J will become the generators of the homogeneous Lorentz group.
It is not immediate to guess the explicit expression for the tensor M , because Lorentz transformations act on
the coordinates so that they do not leave the Lagrangian density invariant. Notwithstanding, translation invariance
permits to formulate Lorentz invariance as a symmetry of the Lagrangian density under certain transformations
on the fields and field derivatives alone (not including coordinate transformations). The transformations on the
fields of the form (10.117) acquires the matrix form given by [note that the indices are making the role of the
index a in Eq. (10.117)]
i
k = (J )k m m (10.136)
2
where the matrices J satisfy the Lie algebra of the Lorentz homogeneous group

[J , J ] = iJ g iJ g iJ g + iJ g

for instance, in the case of a scalar field we have = 0 then J = 0, while for a covariant vector field we have

V = V (10.137)

so the tensors are given by


(J ) = ig + ig (10.138)
10.10. LORENTZ INVARIANCE 301

we can check the validity of the tensor (10.138) by calculating Vk using the covariant form of Eq. (10.136) and
using the tensor (10.138)
i i
V = (J ) m Vm = [ig m + ig m ] Vm
2 2
1 1
= [g V g V ] = [g V g V ]
2 2
1 1
= [g V + g V ] = [ V + V ]
2 2
V = V
hence the tensor (10.138) generates the correct covariant variation (10.137) of the vector field V .
Now, the derivative of a field that transforms as (10.136) has the same structure of transformation but adding
an extra vector index
1
( k ) = i (J )k m m + k (10.139)
2
Since we assume that the Lagrangian density is invariant under the combined transformations (10.136) and (10.139)
we have  
L k L k
0= + (10.140)
k ( k )
substituting (10.136) and (10.139) in (10.140) we have
L i L i L
0= (J )k m m + (J )k m m + k (10.141)
k 2 ( k ) 2 ( k )
the first two terms are written in terms of but the latter is written in terms of . It is convenient to write
the third term in terms of in order to factorize it. To do it, we rewrite the third term on the right-hand side
of Eq. (10.141) as follows
1 1 1 
k = k + k = g k + g k
2 2 2
k 1 k
= (g g ) (10.142)
2
Replacing (10.142) in (10.141) we can factorize as we desired
 
i L k i L k 1 L
0 = (J ) m m
+ (J )
m m
+ (g
g
) k
2 k 2 ( k ) 2 ( k )
is infinitesimal but otherwise arbitrary, setting the coefficients of to zero we have
i L i L 1 L
0= k
(J )k m m + k
(J )k m m + (g g ) k (10.143)
2 2 ( ) 2 ( k )
By applying the Euler-Lagrange equations (10.57) and the expression (10.115) for the energy-momentum tensor
T equation (10.143) can be rewritten as
   
i L k m i L k m 1 L k k
0 = (J ) m + (J ) m + g g
2 ( k ) 2 ( k ) 2 ( k )
   
i L k m 1 L k L k
= (J ) m + g g
2 ( k ) 2 ( k ) ( k )
 
i L 1 
= k
(J )k m m + g ( L T ) g L T
2 ( ) 2
 
i L 1
0 = k
(J )k m m + [(g L T ) (g L T )]
2 ( ) 2
302 CHAPTER 10. CANONICAL QUANTIZATION

 
i L k m 1
0 = (J ) m (T T ) (10.144)
2 ( k ) 2
which suggest to redefine a new energy-momentum tensor, called the Belinfante tensor
i
T Z (10.145)
2
L L L
Z (J )k m m (J )k m m (J )k m m (10.146)
( k ) ( k ) ( k )

because of the antisymmetry of J , the term Z is clearly antisymmetric in the indices and .
L L L
Z = k
(J )k m m k
(J )k m m (J )k m m
( ) ( ) ( k )
L L L
= k
(J )k m m + k
(J )k m m + (J )k m m
( ) ( ) ( k )
Z = Z

therefore we have
Z = 0
because is symmetric in , while Z is antisymmetric in those indices. Hence, obeys the same
conservation law as T
i
T Z = T = 0
2
= 0 (10.147)
Owing to the antisymmetry in , we have Z 00 = 0. Hence when we set = 0 in Eq. (10.145), the other index
runs over space components only
i i
0 T 0 Z 0 = T 0 j Z j0
2 2
therefore by using the divergence theorem and taking into account that all fields vanish at infinity, the derivative
term in the Belinfante tensor disappear when we integrate over all space
Z Z Z Z Z
0 3 0 3 i j0
 3 0 3 i
d x T d x j Z d x= T d x Z j0 dSj
2 2
Z Z
0 d3 x = T 0 d3 x = P (10.148)

from Eqs. (10.115) and (10.148) the time component of P yields

Z Z Z Z
P0 00 d x = T00 d x = g0 T 0 d x = g00 T 0 0 d3 x =
3 3 3

Z Z   Z 
0 3 0 L k 3 3 L k
= T 0 d x= 0 L 0 d x = L d x
(0 k ) k
Z 
P0 = L d3 x k = H

we conclude that
P0 H
from which we can regard as the energy-momentum tensor just as well as T .
10.10. LORENTZ INVARIANCE 303

Now we shall show that the Belinfante tensor is symmetric (in contrast to the tensor T ). To prove it,
we use the definition (10.145) to evaluate

i L L
I = T T (J )k m m (J )k m m
2 ( k ) ( k )

L k m L k m L k m L k m
(J ) m + (J ) m (J ) m + (J ) m
( k ) ( k ) ( k ) ( k )

the last four terms cancel directly. Now from the antisymmetry of J and using Eq. (10.144) we obtain
 
i L k m
I = T T 2 (J ) m =0
2 ( k )

Consequently, we can see from Eq. (10.144) that the Belinfante tensor is symmetric

= (10.149)

in contrast to the tensor T . Indeed in gravitational theories is and not T the appropriate energy-
momentum tensor. Because of the symmetry of the Belinfante tensor we can obtain another conserved tensor
density as
M x x (10.150)
it is immediate to see that such a tensor is conserved i.e. leads to a continuity equation by using (10.149, 10.147
 
M = x x = + x x
= = = 0
M = 0 (10.151)

So Lorentz invariance leads to another set of conserve charges (time-independent operator)


Z Z
0 3
 
J = M d x = d3 x x 0 x 0 (10.152)

10.10.2 Generators and Lie algebra between the homogeneous and inhomogeneous genera-
tors
Recalling the assignments of generators for the Lorentz group in section 1.8.5, Eqs. (1.126)-(1.129), page 30

P P 1, P 2, P 3 (10.153)

J J 23 , J 31 , J 12 {J1 , J2 , J3 } J km = kmn Jn (10.154)
H = P0 (10.155)

K J 01 , J 02 , J 03 {K1 , K2 , K3 } J 0i = Ki (10.156)
we shall show that by assigning the time-independent operators P , J of Eqs. (10.148, 10.152) as generators
we recover the Lie algebra of the inhomogeneous Lorentz group through definitions (10.153)-(10.156).
Since the rotation operator it not time-dependent neither has explicit time-dependence it is a constant of
motion so that it commutes with the Hamiltonian
1
[H, J] = 0 ; Jk = ijk J ij
2
304 CHAPTER 10. CANONICAL QUANTIZATION

applying the commutation identity (10.112) to the tensor 0 we have


h i 1  Z  
1 mk 3 m 0k k 0m
[Pj , Ji ] = imk Pj , J = imk Pj , d x x x
2 2
Z h  i Z  
1 i
= imk d3 x Pj , xm 0k xk 0m = imk d3 x j xm 0k xk 0m
2 2
Z  
i 3 m 0k m 0k k 0m k 0m
= imk d x j + x + j x
2 xj xj
Z Z
i i
= imk d3 x jm 0k + imk d3 x jk 0m
2 2
Z  
i
+ imk d3 x xm j 0k xk j 0m
2 x x
by applying the antisymmetry of ijk we see that the first term cancel each other
Z Z Z Z
i i i i
A imk d x j + imk d x j = ijk d x + imj d3 x 0m
3 m 0k 3 k 0m 3 0k
2 2 2 2
Z Z
i i
= ijk d3 x 0k ijm d3 x 0m = 0
2 2
then we have Z   Z
i 3 m 0k
[Pj , Ji ] = imk d x x j
xk j 0m = iijk d3 x 0k
2 x x
where we have integrated by parts in the last step. We finally obtain the correct Lie algebra Eq. (1.133) page 31
[Pj , Ji ] = iijk Pk
By contrast, the boost generator Kk J k0 is time-independent but has an explicit time-coordinate
Z h i
Kk J = d3 x xk 00 x0 0k
k0

that can be rewritten as Z Z


3 k 00 0
Kk d x x x d3 x 0k
Z
K = tP + d3 x x 00 (x, t) (10.157)

K is a constant so that
0 = K = P + i [H, K]
[H, K] = iP (10.158)
so that we obtain the proper Lie algebra Eq. (1.136). Using again the identity (10.112) we have
 Z  Z h i
[Pj , Kk ] = Pj , tPk + d x x (x, t) = d3 x Pj , xk 00 (x, t)
3 k 00

Z h i Z Z
= i d x j x (x, t) = i d x j (x, t) + i d3 x xk j 00 (x, t)
3 k 00 3 k 00

Z Z

= i d3 x xk j 00 = ijk d3 x 00
x
where we integrate by parts in the last step. So that
[Pj , Kk ] = ijk H
10.10. LORENTZ INVARIANCE 305

10.10.3 Invariance of the Smatrix


For most of realistic Lagrangian densities, the operator (10.157) is smooth in the sense explained in Sec. 2.3.1 [see
discussion below Eq. (2.94) page 80], that is the interaction terms
Z 
eiH0 t
d x x (x, 0) eiH0 t
3 00
(10.159)

vanish at t . i.e. the matrix elements between states that are smooth superpositions of energy eigenstates
vanish in the limit t . Indeed, the interaction term in Eq. (10.159) must vanish if we want to define
proper in and out states as well as the Smatrix. With the same arguments of section 2.3.1, we can derive
the Smatrix Lorentz invariance by using such a smoothness assumption along with the commutation relation
(10.158).

10.10.4 Lie algebra within the homogeneous generators


In addition the same arguments were used in section 2.3.1, to show that the remaining commutation relations
of J between themselves, acquire the apropriate form. To prove it, we should write the operators J ij in terms
of fields. We start by using equations (10.145, 10.146) and (10.115) to express the Belinfante tensor 0j in terms
of fields
i i i i
0j = T 0j Z 0j = T 0 gj Z 0j = T 0 j gjj Z 0j = T 0 j Z 0j
 2  2  2 2
L i L k
= j0 L j k J 0j m m
(0 k ) 2 ( k )

L 
j k m L 
0 k m
J m J m
(0 k ) (j k )


L i L k
0j
=   j k J 0j m m
2 ( )k
k

L 
j k m L 
0 k m
J m J m
(0 k ) (j k )
then such a tensor is replaced in Eq. (10.152) to obtain J ij in terms of fields
Z
 
J = d3 x xi 0j xj 0i =???
ij

the rotation generators in terms of fields take the form


Z
L h i k i
J = d3 x
ij
x j k + xj i k i J ij m m (10.160)
k
Now, by definition there are no present time-derivatives of the auxiliary fields in the Lagrangian density so that
L/ C r = 0, and the rotation generators do not mix the canonical and auxiliary fields. From those facts we see
that in Eq. (10.160) auxiliary fields are absent. Thus, we can rewrite (10.160) in terms of canonical variables only
Z h i
n
J = d3 x Pn xi j Qn + xj i Qn i J ij m Qm
ij
(10.161)

From Eq. (10.161) we can also verify the correctness of the commutation relations for rotation generators. For
instance, from the canonical commutation relations and Eq. (10.161) we obtain
 ij n  n
J , Q (x) = i (xi j + xj i ) Qn (x) J ij m Qm (x) (10.162)
 ij   m
J , Pn (x) = i (xi j + xj i ) Pn (x) + J ij n Pm (x) (10.163)
306 CHAPTER 10. CANONICAL QUANTIZATION

from the results (10.162) and (10.163) we can obtain the commutation relations between J ij by themselves and
with other generators (since all generators are written as functions of Q s and P s). For example, J ij commutes
with H and with Pn Qn so that it commutes with the Lagrangian L. Thus, the commutator of J ij with the
auxiliary fields (when they exist) must be consistent with the rotational invariance of L. In the absence of
auxiliary fields we can apply the same reasoning to the boosts generators to prove that the generators P and
J obey the commutation relations of the Poincare group. Notwithstanding, when there are auxiliary fields the
boost matrices mix auxiliary fields with canonical ones (e.g. the components V i and V 0 of a vector field), since
boosts mix naturally time-components with spatial components. Hence, in the presence of auxiliary fields the
commutation relations of the J i0 between themselves and with other generators should be checked for the specific
case. However, it is not necessary for the proof of Lorentz-invariance of the Smatrix obtained in section 2.3.1.

10.11 The transition to the interaction picture

We have already seen in Sec. 10.6.1, that we have to go to the interaction picture to have a theory ready for the
use of perturbation theory, and we saw how to construct the structure of the Lagrangian including the free and
interaction parts. We shall solve some additional more complicated examples here.

10.11.1 Scalar field with derivative coupling

We shall start again with a neutral scalar field, but with derivative coupling. Let us assume the Lagrangian density

1 1
L = m2 2 J H ()
2 2

with J being either a cnumber external current (not to be confused with the conserved currents worked pre-
viously) or a functional of several fields different from (in the latter case we should add terms involving these
other fields to the Lagrangian density). The canonical conjugate of yields

 
L 1 0 1 i 1 2 2 0 i
= = 0 i m J 0 J i H ()
2 2 2
 
1 2
= J 0
2
= J 0 (10.164)

The Hamiltonian is obtained from the Legendre transformation and using (10.164) to obtain the Hamiltonian in
terms of and

Z h i Z  
3 3 1 0

1 2 2
H = d x L = d x + J + + m + J + H ()
2 2
Z  
3 0
 1 2 1 2 1 2 2 0
= d x + J + () + m + J + J + H ()
2 2 2
Z  
3 0
 1 
0 2 1 2 1 2 2 0 0

H = d x +J +J + () + m + J + J + J + H () (10.165)
2 2 2
10.11. THE TRANSITION TO THE INTERACTION PICTURE 307

now, and J 0 commute, because J 0 contains other fields different from . From this fact the Hamiltonian can
be reorganized it as
Z 
1 1 0 2 1
H = d x 2 + J 0 2
3
J J 0 + ()2
2 2 2

1 2
+ m2 2 + J 0 + J 0 + J + H ()
2
Z 
3 1 2 1 0 2
H = d x + J 0 + J
2 2

1 2 1 2 2
+ () + m + J + H ()
2 2

This Hamiltonian can be separated into the free and interacting parts as follows

H = H0 + V
Z  
3 1 2 1 2 1 2 2
H0 = d x + () + m
2 2 2
Z  
3 0 1 0 2
V = d x J + J + J + H ()
2

as we saw in Sec. 10.6.1, we can go to the interaction picture by simply replacing the set , with the set ,
(which corresponds to evaluate the Hamiltonian at t = 0). Of course, we should do the same replacement in the
fields contained in J though we do not indicate it explicitly.
Z  
1 2 1 1
H0 = d3 x (x, t) + [ (x, t)]2 + m2 2 (x, t) (10.166)
2 2 2
Z  
3 0 1 0 2
V (t) = d x (x, t) J (x, t) + J (x, t) (x, t) + J (x, t) + H ( (x, t)) (10.167)
2

The free Hamiltonian H0 coincides with the Hamiltonian (10.73) of Sec. 10.6.1, thus it leads to Eqs. (10.74)-
(10.81). Indeed, regardless the form of the total Hamiltonian we should take (10.166) as the free part of it,
because it is this Hamiltonian the one that leads to the appropriate expansion (10.77) of the scalar field in terms
of creation an annihilation operators that obey the commutation relations (10.81). Next, we should substitute in
the interaction Hamiltonian with its value in the interaction picture (not with its value J 0 in the Heisenberg
picture), so Eq. (10.167) becomes
Z  
1 0 2
V (t) = d3 x 0 (x, t) J 0 (x, t) + J i (x, t) i (x, t) + J (x, t) + H ( (x, t))
2
Z  
3 1 0 2
V (t) = d x J (x, t) (x, t) + J (x, t) + H ( (x, t)) (10.168)
2
note that the non-invariant term in Eq. (10.168) is precisely the one required to recover Lorentz invariance in the
propagator of in section 9.7.

10.11.2 Spin one massive vector field


We shall start the canonical quantization of the vector field V for a particle of spin one, by writing a quite general
Lagrangian density
1 1 1
L = V V V V m2 V V + J V (10.169)
2 2 2
308 CHAPTER 10. CANONICAL QUANTIZATION

where , , m2 are arbitrary constants so far. As before, J is either a cnumber external current, or an operator
that depends on other fields. Once again, in the latter case we should add to the Lagrangian density, terms
involving the other fields.
By taking the Euler-Lagrange equation for the vector field
L L
=
( V ) V
and applying it to the Lagrangian density (10.169) the left-hand side of the Euler-Lagrange equation yields
 
L 1 1
A1 = V V V V
( V ) ( V ) 2 2
 
1 1
= ( V ) g g ( V ) ( V ) g g ( V )
( V ) 2 2
 
1 1
= g g ( V ) ( V ) ( V ) ( V )
( V ) 2 2

1 1
= g g ( V ) ( V )
2 2

1 1
( V ) ( V )
2 2

1 1
A1 = g g ( V ) ( V ) g g
2 2

1 1
g g ( V ) ( V ) g g
2 2
 
1 1 1 1
= ( V ) ( V ) ( V ) ( V )
2 2 2 2
A1 = V ( V )
L
A1 = V ( V ) (10.170)
( V )
the right-hand side of the Euler-Lagrange equation gives
   
L 1 2 1 2
A2 = m V V + J V = m g V V + J V
V V 2 V 2
   
1 2 1 2
= m g ( V + V ) + J = m (g V + g V ) + J
2 2
L
A2 = m2 V + J (10.171)
V
equating the left and right sides of the Euler-Lagrange equation (10.170) and (10.171) we find
V ( V ) = m2 V + J
In summary, the Euler-Lagrange equations (10.57) applied to the vector field V with the Lagrangian density
(10.169) yields
V ( V ) + m2 V = J (10.172)
and taking the divergence we have
 V ( V ) + m2 V = J
 V  ( V ) + m2 V = J
10.11. THE TRANSITION TO THE INTERACTION PICTURE 309

( + )  V + m2 V = J (10.173)
 
m2 J
 V = (10.174)
( + ) ( + )

which provides the equation for an ordinary scalar field ( V is clearly scalar) with mass m2 / ( + ) and source
J / ( + ). We shall describe a particle of spin one (not of spin zero). Thus, we can avoid the appearance of
an independently propagating scalar field term V by taking = . In that case we cannot divide by ( + )
and according with Eq. (10.173) we have
m2 V = J

therefore, when = , the term V can be written in terms of an external current or other fields, as J /m2 .
The Lagrangian density (10.169) becomes

1 1 1
L = V V + V V m2 V V + J V (10.175)
2 2 2

so that the scalar can be absorbed in the normalization of the vector field V (recall that m2 by now is just a
coefficient and J can also be normalized), then we can define

= = 1 (10.176)

so that the Lagrangian density becomes

1 1 1
L = ( V ) V + ( V ) V m2 V V + J V (10.177)
2 2 2

and taking into account that indices and are dummy

1 1
L = [2 ( V ) V 2 ( V ) ( V )] m2 V V + J V
4 2
1 1 2
= [( V ) V ( V ) ( V ) ( V ) V + ( V ) V ]

m V V + J V
4 2
1 1 2
= [( V ) V ( V ) ( V ) ( V ) V + ( V ) V ] m V V + J V
4 2
1 1
L = ( V V ) ( V V ) m2 V V + J V
4 2

so the Lagrangian density yields

1 1
L = F F m2 V V + J V (10.178)
4 2
F V V (10.179)

it is clear that F is an antisymmetric tensor. Under the assumption (10.176) the Euler-Lagrange equations
become

V + ( V ) + m2 V = J
[( V ) V ] + m2 V = J
F + m2 V = J (10.180)
310 CHAPTER 10. CANONICAL QUANTIZATION

The derivative of the tensors F and F with respect to V yields

F
= ( V V ) = 0 0
V (0 V )
F  
= V V = g g ( V V )
V (0 V ) (0 V )
F
= g g (0 0 )
V
F
= g0 g g g0
V

thus, the derivative of the Lagrangian density (10.178) with respect to V gives
!
L 1 
 1 F 1 F
= F F = F F
V 4 V 4 V 4 V
1   1
= F g0 g g g0 (0 0 ) F
4 4
L 1  1 
= F 0 F 0 F 0 F 0
V 4 4

and taking into account the antisymmetry of F we finally obtain

L
= F 0 (10.181)
V

which is non-zero for spatial components i. Thus, V i are canonical fields, and their conjugates are given by

L
i = = F 0i = F i0 = i V 0 0 V i = i V 0 + 0 V i
V i

i = F i0 = V i + i V 0 (10.182)
However, because of the antisymmetry, F 00 = 0. Consequently, V 0 does not appear in the Lagrangian density, so
that V 0 is an auxiliary field. The fact that L/ V 0 = 0, implies that the field equation for V 0 does not contain
second time-derivatives, from which it can be used as a constraint that eliminates the field variable. For = 0,
the Euler-Lagrange equation (10.180) gives

F 0 + m2 V 0 = J 0
0 F 00 i F i0 = m2 V 0 + J 0

so we have finally
i F i0 = m2 V 0 J 0 (10.183)
combining Eqs. (10.183) and Eq. (10.182) we find

1  1 
V0 = i F i0
+ J 0
= i i
+ J 0
m2 m2
1  ~ + J0

V0 = (10.184)
m2
10.11. THE TRANSITION TO THE INTERACTION PICTURE 311

showing again the fact that V 0 is not an independent variable. Now we calculate the Hamiltonian. For this, we
~ and J 0
first observe that Eqs. (10.182) and (10.184) permits to express V in terms of
 
V = ~ V 0 = ~ 1 ~ + J0 (10.185)
m2
Now, we intend to evaluate the Hamiltonian. We start by reexpressing the Lagrangian density (10.178) in terms of
i and V i (canonical variables) and eliminating V 0 . We do this by using Eqs. (10.184, 10.182) as followsevaluating
the quantity

1 1
L = F F m2 V V + J V
4 2
1 1 1 1
= F0 F 0 Fi F i m2 V0 V 0 m2 Vi V i + J0 V 0 + Ji V i
4 4 2 2
1 0i 1 i0 1 ij 1 2 0 2 1 2 2
= F0i F Fi0 F Fij F + m V m V + J0 V 0 + J V
4 4 4 2 2
1 1 1  2 1  
= i i Fij F ij + ~ + J 0 m2 V2 + J0 ~ + J0 + J V
2 4 2 2 m2
1~2 1 1   0  
~ + J 0 1 m2 V 2 J
2
L = Fij F ij + ~ + J0 + J V

2 4 2 2 m2
we can show that
1 1
Fij F ij = ( V)2
4 2
and the Lagrangian density yields

1~ 2 1 1 2
~ + J 0 1 m2 V 2 J
0  
L= ( V)2 + ~ + J0 + J V
(10.186)
2 2 2 2 m2

using Eqs. (10.185) and (10.186), the Hamiltonian becomes
Z   Z     
3 ~ 3 ~ ~ 1 ~ 0
H = d x V L = d x 2 + J L
m
Z    
= 3
d x ~2 + 1 ~ ~ + J0 1 ~ 2 + 1 ( V)2
m2 2 2
1 2 2 1   2 1 0  
+ m V ~
+J 0
J V+ 2 J + J~ 0
2 2m2 m
then we split it into the free and interacting part

H = H0 + V

then we pass to the interaction picture by replacing the quantities V, ~ (in the Heisenberg picture) with the
corresponding interaction picture variables v and ~ (the same replacements should be done for the fields contained
in J ).

Z  
1 1 1 1
H0 = d x ~ 2 +
3
2
( ~ )2 + ( v)2 + m2 v2 (10.187)
2 2m 2 2
Z  
1 1 2
V = d3 x J v + 2 J 0 ~ + 2
J0 (10.188)
m 2m
312 CHAPTER 10. CANONICAL QUANTIZATION

and the relation between ~ and v is given by

H0 (v, ~ ) 1
v = = ~ 2 ( ~ ) (10.189)
~ m
while the field equation yields
H0 (v, ~ )
~ = = 2 v ( v) m2 v (10.190)
v
as for V 0 , since it is not an independent field variable, it is not related with an interaction picture field v 0 through
a similarity transformation. It is convenient to define the following quantity

~
v0 = (10.191)
m2
then combining Eqs. (10.189) and (10.191) we can express ~ in the form
 
~
~ = v +
m2

~ = v + v 0 (10.192)

Thus, we see that by defining v 0 as in Eq. (10.191) we obtain a relation (10.192) between ~ and v similar to the
associated relation in the Heisenberg picture Eq. (10.182).
Taking divergence in Eq. (10.192) and using Eq. (10.191) we have

~ = v + 2 v 0 (10.193)
2 0 2 0
m v = v + v (10.194)

on the other hand, deriving (10.192) with respect to time



~ = v + v 0

and inserting it in Eq. (10.190) we obtain



v + v 0 = 2 v ( v) m2 v

therefore, the field equations become

2 v 0 + v m2 v 0 = 0 (10.195)

2 v ( v) v v 0 m2 v = 0 (10.196)

that can be combined to be written in the covariant form. To do it, we rewrite Eq. (10.195) as

2 v 0 + v m2 v 0 = 0
i i v 0 + i 0 v i m2 v 0 = 0

i i v 0 + 0 0 v 0 0 0 v 0 0 i v i m2 v 0 = 0
v 0 0 v m2 v 0 = 0
v 0 0 v m2 v 0 = 0 (10.197)
10.11. THE TRANSITION TO THE INTERACTION PICTURE 313

and the k th component of Eq. (10.196) is rewritten as


.. k
2 v k v k v 0 k ( v) m2 v k = 0
i i v k 0 0 v k k 0 v 0 k i v i m2 v k = 0
i i v k + 0 0 v k 0 k v 0 i k v i m2 v k = 0
v k k v m2 v k = 0
v k k v m2 v k = 0 (10.198)

thus Eqs. (10.197) and (10.198) can be written in a single covariant equation as

v v m2 v = 0 (10.199)

taking the divergence we have

v v m2 v = 0
 v  v m2 v = 0
 v  v m2 v = 0

which leads to
v = 0 (10.200)
from which the second term in Eq. (10.199) vanishes. Thus such an equation becomes

 m2 v = 0 (10.201)

a real vector field that obeys Eqs. (10.200) and (10.201) can be expanded as a Fourier transform in the form

3/2
XZ d3 p n o
ipx ipx
v (x) = (2) p e (p, ) a (p, ) e + e (p, ) a (p, ) e (10.202)
2p0
p
with p0 = p2 + m2 ; the terms e (p, ) with = 1, 0, 1 are three independent vectors with the constraint [see
Eq. (6.76) page 173]
p e (p, ) = 0 (10.203)
and normalized in the form [see Eqs. (6.69, 6.75) page 171]
X p p
e (p, ) e (p, ) = g + (10.204)

m2

and a (p, ) , a (p, ) are operator coefficients to be determined. From Eqs. (10.192), (10.202) and (10.204) we
can check that v and ~ follow the correct commutation relations
 i 
v (x, t) , j (y, t) = iij 3 (x y)
 i   
v (x, t) , v j (y, t) = i (x, t) , j (y, t) = 0

as long as the operators a (p, ) , a (p, ) satisfy the commutation relations


h i
a (p, ) , a q, = 3 (p q)
  h i
a (p, ) , a q, = a (p, ) , a q, = 0.
314 CHAPTER 10. CANONICAL QUANTIZATION

since we already knew that the vector field for a neutral spin one particle must have the form (10.202), the
derivation of our present results permits to verify that Eq. (10.187) provides a correct form for the free Hamiltonian
for a massive particle of spin one. It could also be checked that the Hamiltonian (10.187) can be written up to a
constant term in the standard form for a free-particle energy, i.e. as (homework!! B12)
XZ
H0 = d3 p p0 a (p, ) a (p, )

Further, applying Eq. (10.191) in Eq. (10.167) gives us the interaction Hamiltonian in the interaction picture
Z  
3 1 
0 2
V (t) = d x J v + J (10.205)
2m2

the non-invariant term in Eq. (10.205) is the one we needed to cancel the non-invariant term coming from the
propagator of the vector field [see Eqs. (9.73), (9.74), and (9.75) on page 247, and discussion around them].

10.11.3 Dirac Fields of spin 1/2


For a Dirac field of spin 1/2, we shall start with the trial Lagrangian density

L = ( + m) H ,

where H , is a real function of and . The Lagrangian is not real but the action is because

A = ( ) ( )
 
= ( ) [ ] = + ()
 
= + 2
 
= + =

therefore the field equations obtained from the stationary condition for the action with respect to are the adjoints
of the equations obtained applying the stationary condition with respect to . Thus both sets of equations are
not independent as we require to avoid too many field equations that overdetermine the problem. The canonical
conjugate of yields
L
= = 0 (10.206)

we should not consider as a field like but as proportional to the canonnical conjugate momentum of . The
Hamiltonian is given by
Z h i Z 
H = d x L = d3 x 0 [ + m] + H
3

as customary we should split it in a free and interacting Hamiltonian

H = H0 + V (10.207)
Z Z
H0 = d x [ + m] ; V = d3 x H
3 0
(10.208)

as we know, the next step is to pass to the interaction picture. We observe that Eq. (10.206) does not involve the
time, thus the similarity transformation (10.39), (10.40) gives

= 0
10.11. THE TRANSITION TO THE INTERACTION PICTURE 315

in the same way, H0 and V (t) can be calculated by substituting , by , in Eqs. (10.208). From this we
obtain the equation of motion
H0
= = 0 ( + m)

or equivalently
( + m) = 0 (10.209)
as explained above, the other equation of motion

H0
=

provides just the adjoint of Eq. (10.209). A solution of Eq. (10.209) can be expanded as a Fourier transform
Z Xn o
1 3 ipx ipx
(x) = d p u (p, ) e a (p, ) + v (p, ) e b (p, ) (10.210)
(2)3/2
p
p0 p2 + m2

where a (p, ) and b (p, ) are operator coefficients, u p, 12 are two independent solutions of

(i p + m) u (p, ) = 0 (10.211)

similarly v p, 12 are two independent solutions of

(i p + m) v (p, ) = 0 (10.212)
P P
The matrix i p has eigenvalues m. Consequently, the bilinears uu and vv must be proportional to
the projection matrices
X (i p + m) X (i p + m)
u (p, ) u (p, ) ; v (p, ) v (p, )

2m
2m

thus the proportionality factor can be fit up to a sign by absorbing it into the definition of u (p, ) and v (p, ).
Then we can fit the sign by a positivity condition
X X
Tr u (p, ) u (p, ) = u (p, ) u (p, ) > 0

and similarly for v (p, ). We then will normalize the spinors as


X (i p + m)
u (p, ) u (p, ) =

2p0
X (i p + m)
v (p, ) v (p, ) =

2p0

Once again we can determine the properties of the undetermined operators a (p, ) and b (p, ) by imposing the
canonical conditions to the fields and canonical conjugates . So in order to obtain the canonical anticommu-
tation relations
   
(x, t) , (y, t) + = [ (x, t) , (y, t)]+ 0 = i 0 3 (x y)
[ (x, t) , (y, t)]+ = 0
316 CHAPTER 10. CANONICAL QUANTIZATION

we must fit the anticommutation relations for a (p, ) , a (p, ) and b (p, ) , b (p, ) as follows
h i h i
a (p, ) , a q, = b (p, ) , b q, = 3 (q p)
 +   +
a (p, ) , a q, + = b (p, ) , b q, + = 0
  h i
a (p, ) , b q, + = a (p, ) , b q, =0
+

and their adjoints. These developments are in agreement with the results obtained in chapter 7. Hence, it confirms
the validity of H0 in Eq. (10.208) as a free Hamiltonian for Dirac fields of spin 1/2. In terms of the operators
a (p, ) and b (p, ) the Hamiltonian reads (homework!! B13)
XZ h i
H0 = d3 p p0 a (p, ) a (p, ) b (p, ) b (p, ) (10.213)

once again we can rewrite it as a more standard (normal ordered) Hamiltonian plus an infinite cnumber
XZ h i
H0 = d3 p p0 a (p, ) a (p, ) + b (p, ) b (p, ) (10.214)

the cnumber divergent term in Eq. (10.214) only matters in graviational theories. For our case like in the case
of the scalar theory, this is just a shift of the zero of energy. When we drop out the cnumber, H0 is a positive
operator as it was the case for bosons.

10.12 Constraints and Dirac Brackets


Since the starting point is usually the Lagrangian, we should confront the task of passing to the Hamiltonian,
which implies to pass to the system of canonical variables Q, P . This task is particularly difficult when we have
constraints. We shall use the analysis of Dirac to manage this problem. As a matter of example we shall use the
vector field theory described by Lagrangian (10.178), and we shall use the same analysis in more complex (and
realistic) problems later.
Primary constraints appear in two cases (a) when we impose a constraint to the system (this is the case when
we choose a particular gauge in any gauge field theory), or (b) when the constraint comes from the structure of
the Lagragian itself. For this second case we shall take as an example the Lagrangian (10.178) that describes a
real vector field V of spin one, coupled to a current J
1 1
L = F F m2 V V + J V (10.215)
4 2
F V V (10.216)

Let us try to work with the four components on V on the same foot. We then define an extension of the conjugates
(10.182)
L
= F 0
(0 V )
because of the antisymmetry of the F tensor we arrive to the primary constraint

0 = 0 (10.217)

In general we get a primary constraint when the equations


L
k =
(0 k )
10.12. CONSTRAINTS AND DIRAC BRACKETS 317

cannot be solved to obtain all the 0 k (at least locally) in terms of k and k . The necessary and sufficient
condition for it, is that the matrix
2 L
(0 k ) (0 m )
be singular i.e. its determinant must be zero. Lagrangians of this kind are called irregular.
On the other hand, we have secondary constraints which come from the condition that the primary constraints
be consistent with the equations of motion. In the case of the massive vector field, it is the Euler-Lagrange
equation (10.183) for V 0
i i = m2 V 0 J 0 (10.218)
these are all constraints we shall encounter in the present example. However, in other theories further constraints
could appear by requiring consistency of the secondary constraints will field equations, and so on. Nevertheless,
the distinction between primary, secondary
h i and other constraints will not be relevant in our present approach.
Let us assume a Lagrangian L , that depends on a set of variables a (t) and time-derivatives a (t).
The Lagrangians of a Field theory are a special case in which the index a runs over all pairs of k and x. For all
these variables we can define canonical conjugate momenta as

L
a
a
The set of variables and are not in general independent but can be connected by several constraints equations,
both primary or secondary (Hamiltons approach requires that the canonical variables be independent so that we
should get rid of the constrained variables). We define the Poisson bracket as

A B B A
[A, B]P
a a a a

where we ignore the constraints in calculating the derivatives with respect to a and a (the calculation of
a partial derivative implies to ignore constraints, since we are moving each single variable without moving the
others). It is clear that
[a , b ]P = ba
once again we emphasize that in a field theory this delta has kronecker deltas for discrete indices and a Dirac
delta for positions. In evaluating the Poisson bracket, all fields are evaluated at the same time, thus we omit the
time argument. The Poisson brackets possess the same algebraic properties of commutator including the Jacobi
identity (Homework!! B14)

[A, B]P = [B, A]P (10.219)


[A, BC]P = B [A, C]P + [A, B]P C (10.220)
[A, [B, C]P ]P + [B, [C, A]P ]P + [C, [A, B]P ]P = 0 (10.221)

for example property (10.220) can be verified as follows


   
A (BC) (BC) A A C B C B A
[A, BC]P = a
a
= a
B + C B a
+ a
C
a a a a a
   
A C A B C A B A
= B a
+ a
C B + C
a a a a a
a
   
A C C A A B B A
= B + C
a a a a a a a a
[A, BC]P = B [A, C]P + [A, B]P C
318 CHAPTER 10. CANONICAL QUANTIZATION

as usual we define a function of operators as the corresponding series expansion of the ordinary function

X
F (O) fnn On
n=0

and its derivative with respect to O as



dF (O) X
nfn On1
dO
n=0

we can define functions and derivatives of two operators f (A, B) accordingly. If we have two operators O1 and
O2 that commutes with their commutators that is

[O1 , [O1 , O2 ]] = [O2 , [O1 , O2 ]] = 0

we find the property

dF (O2 ) dG (O1 )
[O1 , F (O2 )] = [O1 , O2 ] ; [G (O1 ) , O2 ] = [O1 , O2 ]
dO2 dO1
a very important particular case is the following

[O1 , O2 ] = I

which is clearly the case when we have canonical variables. In that case we have

dF (O2 ) dG (O1 )
[O1 , F (O2 )] = ; [G (O1 ) , O2 ] =
dO2 dO1
?????
-
From the previous results ????
if we were able to adopt the usual commutation relations
h i
[a , b ] = ba ; a , b = [a , b ] = 0

the commutator of any two functions of s and s would satisfy

[A, B] = i [A, B]P

but the constraints do not permit it in general.


We can generically express the constraints in the form

N (, ) = 0 (10.222)

where N is a set of functions of the canonical variables. Since we are including both primary and secondary
constraints, the set (10.222) of all constraints is necessarily consistent with the equations of motion

A = [A, H]P

so that
N = [N , H]P
from which
[N , H]P = 0 (10.223)
10.12. CONSTRAINTS AND DIRAC BRACKETS 319

when we have the constraint N = 0.


A constraint is of first class if its poisson bracket with all the other constraints vanishes when (after calculating
the Poisson bracket) we impose the constraint. We shall see an example in quantum electrodynamics where the
first class constraint comes from gauge invariance which is a symmetry of the action. In general the set of first
class constraints N = 0 is always associated with a group of symmetries, from which any quantity A undergoes
the infinitesimal transformation X
N A N [N , A]P
N
recalling that in field theory the index N has a space-time coordinate, these theories contain local transformations.
From Eq. (10.223) it can be seen that this transformation leaves the Hamiltonian invariant, and for first class
constraints it respects all other constraints as well. First class constraints can be dropped out by an appropriate
gauge choice.
When all first class constraints are dropped out by choicing the adequate gauge, the remaining constraints
N = 0 satisfy the condition that any non-trivial linear combination of the Poisson brackets between each other
X
uN [N , M ]P
N

does not vanish. Therefore, the matrix associated with the Poisson brackets of the remaining constraints must
have non-null determinant

det C 6= 0 ; CN M [N , M ]P
such a matrix is then non-singular. These types of constraints are second class constraints. It is clear that CN M
is antisymmetric. Besides, an antisymmetric matrix of odd dimensionality has null determinant4 . Consequently,
the number of second class constraints must be even.
We have seen in Eqs. (10.217) and (10.218) that for the case of the massive vector field the constraints are
given by
1x = 2x = 0 (10.224)
with
1x = 0 (x) ; 2x = i i (x) m2 V 0 (x) + J 0 (x)
the Poisson brackets of these constraints read
 
C1x,2y = [1x , 2y ]P = 0 (x) , i i (y) m2 V 0 (y) + J 0 (y) P
 
0 (x) i i (y) m2 V 0 (y) + J 0 (y) i i (y) m2 V 0 (y) + J 0 (y) 0 (x)
=
V V (y)
 2 0 0

i i (y) m V (y) + J (y) 0 (x) 0
V (y) 0 (x)
=
= m2
V (y) V 0 (y) 0 (y)

C1x,2y = C2y,1x = [1x , 2y ]P = m2 3 (x y) (10.225)


C1x,1y = C2x,2y = 0 (10.226)
 
0 m2 3 (x y)
C= (10.227)
m2 3 (x y) 0
4
For an antisymmetric matrix we have
 
det A = e = (1)n det A
det A e

det A = (1)n det A

which for n even becomes a triviality and for n odd gives det A = det A, so that the determinat becomes zero.
320 CHAPTER 10. CANONICAL QUANTIZATION

It is clear that this matrix is non-singular5 . Then, the constraints (10.224) are of second class.
Dirac suggested that when all constraints are of second class, the commutation relations satisfy the property

[A, B] = i [A, B]D (10.228)

where [A, B]D is a generalization of the Poisson bracket, known as the Dirac bracket6
N M
[A, B]D [A, B]P [A, N ]P C 1 [M , B]P (10.229)

We recall again that the indices N and M include the position in space, taking values of the form 1, x and 2, x
in the example of the massive vector field. The Dirac brackets satisfy the same basic algebraic properties of
commutators and Poisson brackets, that is (homework!! B15)

[A, B]D = [B, A]D


[A, BC]D = [A, B]D C + B [A, C]D
[A, [B, C]P ]P + [B, [C, A]P ]P + [C, [A, B]P ]P = 0

now setting A = N in the definition (10.229) we have

QM
[N , B]D [N , B]P [N , Q ]P C 1 [M , B]P
 QM
= [N , B]P C N Q C 1 [M , B]P
= [N , B]P N M [M , B]P = 0

so we obtain the additional relations


[N , B]D = 0 (10.230)
property (10.230) makes the commutation relations (10.228) consistent with the constraints N = 0. It can also
be shown that the Dirac brackets are left invariant when we substitute N with any functions N for which the
equations N = 0 and N = 0 generate the same submanifold of phase space. But none of the previous properties
prove Diracs conjecture (10.228).
The conjecture is illuminated by a theorem proved by Maskawa and Nakajima. They proved that for any set
of canonical variables a and a governed by second class constraints, it can always be constructed by a canonical
transformation7 two sets of variables Qn , Qr and their corresponding canonical conjugates Pn , Pr , such that the
constraints give
Qr = Pr = 0
while Qn and Pn are unconstrained canonical variables8 . Now, using these coordinates to calculate the Poisson
brackets, and redefining the constraint functions as

1r = Qr ; 2r = Pr

we obtain
5
Strictly speaking, this matrix has continuous entries. We see it by recalling that CNM corresponds to indices N and M in which
N contains a discrete index and a position index. In Eq. (10.227), we see that the non-diagonal submatrices are proportional to the
identity in the continuum.
6
Note that the definition (10.229) requires that the constraints be of second class for C 1 to exist.
7
A canonical transformation keeps the poisson bracket invariant as well as the structure of Hamiltons equations.
8
In other words exists a canonical transformation for which a given subset of the canonical variables absorbs all the constraints
when all cosntraints are of second class.
10.12. CONSTRAINTS AND DIRAC BRACKETS 321

C1r,2s = [Qr , Ps ]P = sr ; C2s,1r = C1r,2s


C1r,1s = [Qr , Qs ]P = 0 ; C2r,2s = [Pr , Ps ]P = 0
   
0 (x y) 1 0 (x y)
C = ; C = = C
(x y) 0 (x y) 0

where we have taken into account that sr implies a kronecker delta and a Dirac delta in the position. For arbitrary
functions A and B of the complete set of canonical variables Qn , Qr and Pn , Pr we obtain
   
r A Qr Qr A A Qr Qr A
[A, 1r ]P = [A, Q ]P = +
Qn Pn Qn Pn Qs Ps Qs Ps
Qr A A A
= = sr =
Qs Ps Ps Pr

and similarly for [A, 2r ]P then for arbitrary functions A and B of the canonical variables we have9

A A
[A, 1r ]P = ; [A, 2r ]P = (10.231)
Pr Qr

This Cmatrix is non-singular since C 1 = C, as it must be for second class constraints. Hence, the Dirac
brackets (10.229) give here
N M
[A, B]D [A, B]P [A, N ]P C 1 [M , B]P

1 1r,1s
1r,2s
= [A, B]P [A, 1r ]P C [1s , B]P [A, 1r ]P C 1 [2s , B]P

1 2s,1r

1 2r,2s
[A, 2s ]P C [1r , B]P [A, 2r ]P C [2s , B]P
[A, B]D = [A, B]P + [A, 1r ]P sr [2s , B]P [A, 2s ]P sr [1r , B]P

[A, B]D = [A, B]P + [A, 1r ]P [2r , B]P [A, 2r ]P [1r , B]P (10.232)
and applying Eqs. (10.231) to Eq. (10.232) we have

A B A B
[A, B]D = [A, B]P + r

Pr Q Qr Pr
A B B A
[A, B]D = [A, B]P r
+ (10.233)
Q Pr Qr Pr
taking into account that the Poisson bracket must be taken with respect to the whole system of canonical variables
Qn , Qr and Pn , Pr ; we find
   
A B B A A B B A A B B A
[A, B]D = n
n
+ r
r
r
+
Q Pn Q Pn Q Pr Q Pr Q Pr Qr Pr
A B B A
[A, B]D = (10.234)
Q Pn Qn Pn
n

In words we can say that the Dirac bracket coincides with the Poisson bracket calculated in terms of the reduced
set of unconstrained canonical variables Qn , Pn . It shows the utility of the Dirac brackets, since the Dirac bracket
of two arbitrary functions A, B can be evaluated as a Poisson bracket and only with the independent degrees of
freedom Qn , Pn .
9
In this point we emphasize again that by taking partial derivatives we are ignoring the constraints even for the set Q, P.
322 CHAPTER 10. CANONICAL QUANTIZATION

Now, if we assume that the unconstrained canonical variables satisfy the canonical commutation relations, we
see that the commutators of general operators A, B are given by Eq. (10.228) in terms of the Dirac brackets.
It is important however, to say that it is not clear yet whether we should adopt canonical commutation relations
for the unconstrained variables Qn and Pn constructed from theMaskawa-Nakajima canonical transformation.
Indeed the test of such canonical commutation relations is their consistency with the free-field commutation
relations derived for scalar, vector and Dirac fields (see chapters 5, 6, 7, 8). Nevertheless, in order to apply this
test it is necessary toknow what the canonical variables Qn and Pn are. We shall restrict to show it for some
classes of theories. We shall also see that for these classes of theories the Hamiltonian expressed in terms of the
unconstrained s and P s can also be written in terms of the constrained variables.
Let us go back to our special case of massive vector field, and quantize it in terms of Dirac brackets. For this
example it is easy to write the constrained variables V 0 and 0 in terms of the unconstrained variables V i and
i , since they are given by Eqs. (10.217) and (10.218).
1 
0 = 0 ; V 0 =
2
J 0 + i i (10.235)
m
From Eqs. (10.225) and (10.226) we see that the matrix CN M has an inverse given by
1x,2y 2y,1x 1
C 1 = C 1 = 2 3 (x y)
m
1x,1y 
1 2x,2y
C 1 = C =0

From which the Dirac prescription gives the following equal-time commutators
Z
i   
[A, B] = i [A, B]P + 2 d3 z [A, 0 (z)]P i i (z) m2 V 0 (z) J 0 (z) , B A B
m
By definition we have

[V (x) , (y)]P = 3 (x y) ; [V (x) , V (y)]P = [ (x) , (y)]P = 0

so that
      i
V i (x) , V j (y)
V 0 (x) , V 0 (y) = 0 ; V i (x) , V 0 (y) = 2 i 3 (x y)
=
 i    m
V (x) , j (y) = iji 3 (x y) ; V 0 (x) , j (y) = [V (x) , 0 (y)] = 0
[ (x) , (y)] = 0

which are the commutation relations that we obtain by assuming that the unconstrained variables obey the
canonical commutation relations
 i 
V (x) , j (y) = iji 3 (x y)
 i 
V (x) , V j (y) = [i (x) , j (y)] = 0

and using the constraints (10.235) to evaluate the commutators involving the constrained variables 0 and V 0 .
Chapter 11

Quantum electrodynamics

The usual starting point to construct quantum electrodynamics is to start with Maxwells equations and quantize
them. In the present approach we shall instead start showing the necessity of a gauge invariance principle coming
from the difficulties that arise when trying to quantize massless particles with spin. From the gauge invariance
principle we shall infer many of the properties of the quantum theory of electromagnetism, in particular to deduce
the existence of a vector potential to describe massless particles of spin one.

11.1 Gauge invariance


In trying to construct a four-vector covariant field as a linear combination of creation and annihilation fields for
helicity 1, we saw in chapter 8, that it was not possible to build up a truly covariant four-vector of this kind,
but a four-vector field a (x) that transforms as a four-vector up to a gauge transformation [see Eq. (8.40) page
218]
U0 () a (x) U01 () = a (x) + (x, ) (11.1)
it is because in trying to adjust the coefficients of the expansion, we had to use the generators of the little group
ISO (2), whose elements can be written as

W (, , ) = S (, ) R ()

then we adjust the coefficients by using only the subgroup R (), and after trying to do the same with the subgroup
S (, ) we obtain a contradiction.
However, it was possible to construct a truly covariant antisymmetric second-rank tensor free-field of the form

f (x) = a (x) a (x)

The impossibility of building a true four-vector permits to understand the presence of singularities at m = 0 in
the Feynman propagator of a massive vector field of spin one
Z q q

4 g + m 2
(x, y) = (2) d4 q eiq(xy) 2 2
q + m i

which forbids us to obtain the propagator of massless vector particles of helicity 1 through the limit m 0.
From now on we shall use capitol letters because we shall be dealing with interacting fields. We can avoid the
problems above by using only
F (x) = A (x) A (x) (11.2)
because it is a truly covariant second-rank tensor. Nevertheless, this is not the most general possibility neither
is realized in nature. Thus, instead of getting rid of A (x) in the action, we shall demand for the part of the

323
324 CHAPTER 11. QUANTUM ELECTRODYNAMICS

action IM that describes the matter and the radiation-matter interaction to be invariant under the general gauge
transformation
A (x) A (x) + (x) (11.3)
at least when the matter fields satisfy the field equations. In that way the extra term in Eq, (11.1) has no physical
consequences. The change in the matter action under the transformation (11.3) reads
Z
IM
IM = d4 x (x) (11.4)
A (x)
-open
-
????
We first observe that under the condition of null fields at infinity, the four dimensional integrals of any four-
divergence F of any function F involving those fields must vanish. In particular
Z   Z    Z
4 IM 4 IM IM
0 = d x (x) = d x (x) + d4 x (x)
A (x) A (x) A (x)
Z   
4 IM
0 = d x (x) + IM (11.5)
A (x)
where we have used Eq. (11.4). Demanding Lorentz invariance we have IM = 0 and appealing to the arbitrariness
of (x) the integrand in Eq. (11.5) must vanish
????
-
close
Thus, Lorentz invariance of IM demands
IM
=0 (11.6)
A (x)
Let us take first the case in which IM depends only on F (x) and its derivatives, together with matter fields.
In that case it can be shown that

IM IM
= 2 (11.7)
A (x) F (x)
by applying a divergence on both sides we obtain
IM IM
= 2 =0
A (x) F (x)
the latter step comes from the symmetric nature of the tensor derivative and the antisymmetric nature of
F . We conclude that the Lorentz invariance condition (11.6) is obviously satisfied if IM depends only on F (x)
and its derivatives, together with matter fields. Nevertheless, if IM depends on A (x) itself, equation (11.6)
becomes a non-trivial constraint on the theory.
open

????
Now we shall prove Eq. (11.7). If IM depends only on F (x) and its derivatives, together with matter
fields, we take into account that matter fields do not contribute to the variation since they are independent of A ,
therefore
IM F I ( A A ) I
= =
A A F A F
IM ( A ) I ( A ) I
= (11.8)
A A F A F
11.1. GAUGE INVARIANCE 325

now the variation under a Lorentz transformation of the derivative of a field is given by Eqs. (10.136) and (10.139)
pages 300 and 301, that can be combined as

( A ) = A + A

( A ) = A + g A (11.9)
Applying Eq. (11.9) in Eq. (11.8) gives
    
IM A A I A A I
= + g + g
A A A F A A F
   
IM A A I A A I
= + g g (11.10)
A A A F A A F

since A and their derivatives are consider independent (as in the Euler-Lagrange equations), only the first term
in parenthesis contributes in Eq. (11.10)

IM A I A I
=
A A F A F
I I I I
= =
F F F F
IM I I I
= + = 2
A F F F

which proves Eq. (11.7).



-close

11.1.1 Currents and their coupling with A


We wonder now about a theory that provides a conserved current to be coupled to the field A (x). In section
10.8, we saw that infinitesimal internal symmetries of the action lead to the existence of conserved currents. In
particular assuming that the transformation

k (x) = i (x) qk k (x) no sum over k (11.11)

leaves the matter action invariant for constant , the change in the matter action for arbitrary infinitesimal values
(x) must have the form [see Eqs. (10.90) and (10.91) page 290]
Z
IM = d4 x J (x) (x) (11.12)

if the matter fields satisfy the field equations, the matter action is stationary with respect to arbitrary variations
of the k , thus in that case IM in Eq. (11.12) must vanish, therefore

J = 0 (11.13)

further, if IM is the integral of a function LM of k and k , the conserved current becomes [see section 10.8]
X LM
J = i qk k (11.14)
( k )
k
326 CHAPTER 11. QUANTUM ELECTRODYNAMICS

Where k runs over all independent fields different from A . The use of capitol indicates that they are
fields in the Heisenberg picture, whose time-dependence includes the effects of interactions. This generates the
transformations (11.11) i.e. [see Eqs. (4.65), page 148]
h i
Q, k (x) = qk k (x)

where Q is the generalized conserved charge operator


Z
Q= d3 x J 0

Thus, we can construct a Lorentz-invariant theory by coupling A with the conserved current J , in the sense
that IM /A (x) is taken as proportional to J (x) as can be seen by comparing Eqs. (11.6) and (11.13). Any
constant of proportionality can be absorbed in the overall scale of the charges qk , hence we can settle the constant
of proportionality equal to unity
IM
= J (x) (11.15)
A (x)
the conservation of the electric charge only permits to adjust the values of all charges in terms of the value of a
given one. The reference charge is usually taken to be the electron charge e. Equation (11.15) is the one that
provides a definite meaning to the value of e. It is important to take into account that Eq. (11.15) ony fixes the
definition of e after we have carried out the normalization of the field A (x).
We can express the condition (11.15) in the form of a principle of invariance: The matter action is invariant
under the joint transformations

A (x) = (x) (11.16)


k (x) = i (x) qk k (x) (11.17)

this kind of symmetry is called a local symmetry because of the local nature of (x). It is also called a gauge
invariance of the second kind. When is constant we call them global symmetries or gauge invariance of the
first kind. We have many examples of exact local symmetries while global symmetries appear only as accidental
symmetries coming from other principles.

11.1.2 Action for the photons (radiation)


As for the action of the photons (radiation) themselves, we can take it as the one for massive vector fields with
m=0 Z
1
I = d4 x F F (11.18)
4
which coincides with the action of classical electrodynamics. However, we justify it by observing that (up to a
constant) it is the unique gauge invariant functional quadratic in F , without higher derivatives. We shall see
later that it leads to a consistent quantum field theory. Any terms in the action with higher derivatives and/or of
higher order in F can be translated into the so called matter action.
It can be shown from Eq. (11.18) that
I
= F (11.19)
A
From Eqs. (11.15) and (11.19) the field equation for electrodynamics gives


0= [I + IM ] = F + J (11.20)
A
11.1. GAUGE INVARIANCE 327

which are the well-known inhomogeneous Maxwell equations with source J . From the definition (11.2) of F ,
other homogeneous Maxwell equations arise

0 = ( A A ) + ( A A ) + ( A A )
0 = ( A A ) + ( A A ) + ( A A )
0 = ( A A ) + ( A A ) + ( A A )

0 = F + F + F (11.21)
which is obtained from cyclic permutation of the symbols , , .

11.1.3 General overview of gauge invariance


We have started with the existence of massless spin one particles and we arrived to the invariance of the matter
action under the combined local gauge transformations described by Eqs. (11.16) and (11.17). However, in the
standard literature it is usual to start with a global internal symmetry

k (x) = i q k k (x) (11.22)

and wonder about the conditions to extend it to a local symmetry

k (x) = i (x) q k k (x) (11.23)

For the case of a Lagrangian density L that depends only on the fields k (x) but not on their derivatives,
the local or global nature of is irrelevant. Notwithstanding, most of realistic Lagrangian densities depend on
the fields and their derivatives. In the latter case we should work out the problem that the derivatives of fields
transform differently under global or local gauges (i.e. transform differently from fields themselves)

k (x) = i (x) qk k (x) + iqk k (x) (x) (11.24)

we can cancel the second term that spoils gauge invariance by introducing a vector field A (x) with transfor-
mation rule
A (x) = (x) (11.25)
and demand that the Lagrangian density depend on k and A only through the combination

D k k iqk A k (11.26)

which transforms like the fields k themselves as we wanted


     
D k (x) = k iqk A k = k iqk A k
 
= k iqk (A ) k iqk A k

and applying Eqs. (11.23), (11.24) and (11.25) we have

D k (x) = i (x) qk k (x) + iqk k (x) (x) iqk [ (x)] k + qk A (x) q k k (x)
= i (x) qk k (x) + qk A (x) q k k (x)
h i
D k (x) = i (x) qk k (x) iA q k k (x)

D k (x) i (x) qk D k (x) (11.27)


328 CHAPTER 11. QUANTUM ELECTRODYNAMICS

A matter Lagrangian LM (, D) that only depends on k and D k will be invariant under the transformations
(11.23) and (11.25), with (x) an arbitrary function, if it is invariant when is constant. With a Lagrangian
density of this form we obtain
IM X LM   X LM
k
= iq k = i qk k
A D k ( k )
k k

and from Eq. (11.14) we have


IM
= J (x)
A
which coincides with Eq. (11.15). Note that we could include F and its derivatives in LM . With this approach,
the masslessness of the particles described by A is a consequence of the gauge invariance (instead of being an
assumption), since a term in the Lagragian density of the form
1
Lm m2 A A
2
would violate gauge invariance.
From an effective point of view, the global symmetry (11.22) becomes a local symmetry (11.23), by replacing
the ordinary derivative by the covariant derivative D

D iqk A (11.28)

and imposing the rule of transformation (11.25) on the gauge field A .

11.2 Constraints and gauge conditions


We shall adopt a notation in which Qn and Pn denotes canonical matter fields and their canonical conjugates,
while Ai , i are the canonical electromagnetic fields and their canonical conjugates. Thus, we could start by
defining the canonical conjugates of the electromagnetic vector potential field as
L

(0 A )
imposing canonical quantization conditions we would have

[A (x, t) , (y, t)] = i 3 (x y)

but we cannot do it, because A and posses several constraints. We see it by observing that for
F F
L =
4
with a procedure similar to the one that led to Eq. (10.181) page 310, we obtain
L
= F 0
(0 A )
which vanishes for = 0, owing to the antisymmetry of F . As for the matter Lagrangian LM , if it involves only
k and D k , the prescription (11.26) says that LM does not depend on any derivatives of any A . Furthermore,
even if LM depends also on F we have that LM / ( A ) is also antisymmetric in and , so that it vanishes
for = = 0. Thus we have
(L + LM )
0 =0
(0 A0 )
11.2. CONSTRAINTS AND GAUGE CONDITIONS 329

From the discussion above, we find a first constraint coming from the fact that the Lagrangian density
L L + LM
does not depend on the time-derivative of A0 , consequently
0 (x) = 0 (11.29)
this is a primary constraint, since it comes directly from the structure of the Lagrangian. We also have a
secondary constraint in this theory, that comes from the field equation (Euler.Lagrange equation) for the
quantity fixed by the primary constraint. With a procedure similar to the one that led to Eq. (10.218) page
10.218, we obtain
L L
i i = i = = J 0 (11.30)
Fi0 A0
once again the time-derivative term does not appear becuase F00 = 0. Despite the matter Lagrangian density
could depend on A0 , the charge density J 0 only depends on the canonical matter fields and momenta Qn , Pn . We
can see it by using Eq. (11.14) page 325 with = 0
X L X
J 0 = i q n n
= i Pn qn Qn (11.31)
n
(0 n ) n

therefore, Eq. (11.30) is a functional relation between canonical variables. Both Eqs. (11.29) and (11.30) are not
consistent with the usual quantization relations
[A (x, t) , (y, t)] = i 3 (x y) ; [Qn (x, t) , (y, t)] = [Pn (x, t) , (y, t)] = 0 (11.32)
for instance by using = = 0 in the first set of relations (11.32) we obtain
 
A0 (x, t) , 0 (y, t) = i3 (x y)
but constraint (11.29) says that 0 = 0 so that this commutation relation must vanish. On the other hand, by
applying in the second of relations (11.32) and using the constraint (11.30) we have
   
[Qn (x, t) , (y, t)] = 0 Qn (x, t) , i i (y, t) = 0 Qn (x, t) , J 0 (y, t) = 0
and using (11.31) we obtain
X
i qm [Qn (x, t) , Pm (y, t) Qm (y, t)] = 0
m
X
i qm [Qn (x, t) , Pm (y, t)] Qm (y, t) = 0
m

the left-hand side of this equation is not zero in general.


In the case of massive vector fields we can solve that problem in two equivalent forms (a) By using the Dirac
brackets or (b) Treating only Ai and i as canonical variables solving the analog of Eq. (11.30) in order to
obtain A0 in terms of such variables. However, in our present context we cannot use Dirac brackets. We see it by
observing that our constraint functions are
1 = 0 ; 2 = i i + J 0 (11.33)
for massive vector fields the second constraint is i i + J 0 m2 A0 [see Eq. (10.218) page 10.218]. But constraints
(11.33) have vanishing Poisson brackets
 
 0 0
 0 i i + J 0 i i + J 0 0
[1x , 2y ]P = , i i + J P = +
Aa a Aa a
0
[1x , 2y ]P = 0 (since = 0)
330 CHAPTER 11. QUANTUM ELECTRODYNAMICS

consequently, constraints (11.29) and (11.30) are of first class. On the other hand, we cannot either eliminate
A0 as a dynamical variable by writing it in terms of the other variables. Note that Eq. (11.30) is just an initial
condition (not an equation for A0 at all times). Nevertheless, if (11.30) is valid at a given time, it is valid for all
times, because the field equations for the other fields Ai yield
 
L 0 L L
0 i J = 0 i 0 J 0 = i 0 0 J 0
Fi0 F0i F0i
L
= i j i Ji 0 J 0
Fji
 
L 0 L
0 i J = i j J (11.34)
Fi0 Fji
the first term on the right-hand side vanishes because of the symmetry of i j and the antisymmetry of Fji , while
the second term vanishes because of the current coservation condition. Then Eq. (11.34) becomes
 
L 0
0 i J =0
Fi0
note that we have three field equations for four components of A . It is because we have a local gauge symmetry
that prevents us to predict the values of the fields at all times from their values and rates of change at a given time.
In other words, we do not have a unique time evolution for the field A (x, t) even knowing the initial conditions
for A and its derivatives. Instead, for a given solution A (x, t) of the three field equations, by choosing (x, t)
so that its first and second derivatives vanish at t = 0, we can obtain another solution of the form

A (x, t) + (x, t) (11.35)

each one with the same value and time-derivative at t = 0. Indeed there are infinite solutions of the type (11.35)
with the same initial conditions. Of course each solution of the form (11.35) differs from each other [and from
A (x, t)] for later times.
Owing to the partial arbitrariness of A (x, t) we cannot impose the quantization canonical conditions to that
field directly (or for finite mass to Ai ). The most common strategies to work it out are (a) The Lorentz-invariant
method of BRST-quantization (very useful in the quantization of Yang-Mills theories) and (b) By making profit
of the gauge invariance of the theory to choose an apropriate gauge.
We shall follow the second procedure here, that is we shall quantize in a particular gauge. Then we should do
a particular transformation of the form

A (x) A (x) + (x) ; k (x) exp [iqk (x)] k (x)

in order to impose a condition on A (x) that permits to apply the methods of canonical quantization.
In principle we have an infinite set of choices. However, the most common gauges are the following

Lorentz (or Landau) gauge : A = 0


Coulomb Gauge : A= 0
Temporal Gauge : A0 = 0
Axial Gauge : A3 = 0
Unitarity Gauge : real

In the latter gauge, is a complex scalar field with q 6= 0. The unitarity gauge is used when the gauge symmetry
is spontaneously broken by a non-zero vacuum expectation value of .
For the canonical quantization the most convenient gauges are the axial and the Coulomb gauges. We shall
choose the Coulomb gauge because the rotational invariance is more manifest in such a gauge.
11.2. CONSTRAINTS AND GAUGE CONDITIONS 331

We first should be sure that this gauge is possible. To see it, we observe that if A (x, t) does not already
satisfy the condition i Ai = 0, we can make a gauge transformation

A (x, t) A (x, t) + (x, t)

in particular for the spatial coordiante it yields

Ai (x, t) Ai (x, t) + i (x, t)

by applying divergence on both sides and demanding the condition A = 0 we have

0 = i Ai (x, t) i Ai (x, t) + i i (x, t)


0 = A (x, t) + 2 (x, t)

hence, to satisfy the Coulomb gauge the function (x, t) must be chosen such that it satisfies the equation

2 (x, t) = A (11.36)

that is we basically have to solve a Poisson equation with source A. It is the existence of a solution for this
differential equation that guarantees the possibility of choosing such a gauge.
Indeed, we only have to mind about the existence of the solution, but we do not need to solve equation (11.36)
explicitly. We just impose the condition1
A =0 (11.37)
and perform a quantization that respect that constraint.
For simplicity we shall limit ourselves to theories in which the matter Lagrangian density LM could depend
on matter fields and their time-derivatives and on A but not on derivatives of A . For instance, this is the
case when LM depends only on and D with D defined by Eq. (11.28). Quantum Electrodynamics has a
Lagrangian density that fulfills this condition. From this assumption, the only term that could depend on F is
the kinematic term
1
L F F
4
and the constraint equation (11.30) gives

i i = i F 0i = J 0
i F i0 = J 0 (11.38)

and applying the Coulomb condition (11.37) on the right-hand side of this equation we have
 
i F i0 = i i A0 0 Ai = 2 A0 0 i Ai
i F i0 = 2 A0 (11.39)

equating Eqs. (11.38) and (11.39) we have


2 A0 = J 0 (11.40)
Equation (11.40) is the well-known Poisson equation with source J 0 , which can be solved as
Z
J 0 (y, t)
A (x, t) = d3 y
0
(11.41)
4 |x y|
where the remaining three degrees of freedom Ai are subject to the gauge condition (11.37). Note that Eq. (11.40)
can conversely show that J 0 cannot depend exclusively on matter fields.
As we already said, the charge density J 0 only depends on the matter variables Qn and Pn as we see in Eq.
(11.31). Hence, Eq. (11.41) provides an explicit solution for the auxiliary field A0 .
1
We forget the prime notation A , since we just impose the Coulomb condition since the beginning.
332 CHAPTER 11. QUANTUM ELECTRODYNAMICS

11.3 Quantization in Coulomb Gauge


11.3.1 Canonical quantization of the constrained variables
We shall omit the time argument, since all functions and operators are evaluated at the same time. Despite we
have been able to eliminate 0 and A0 through Eqs. (11.29) and (11.41), we still have two more constraints on
the set of variables Ai and i that do not permit to apply the canonical quantization yet. They are the Coulomb
gauge condition
1x i Ai (x) = 0 (11.42)
and the secondary constraint (11.30) that provides the constraint

2x i i (x) + J 0 (x) = 0 (11.43)

these constraints are not compatible with the quantization conditions

[Ai (x) , j (y)] = iij 3 (x y) (11.44)

since by operating on the right-hand side with either /xi or /y j does not give zero.
We shall see that the constraints (11.42) and (11.43) are of second class, from which we can follow the
prescription described in section 10.12 for the commutation relations2 . We shall see it by observing that the
constraint functions have Poisson brackets that forms a non-singular matrix CN M . In this context, the definition
of the Poisson bracket for two arbitrary functionals U and V is3
Z  
3 U V V U
[U, V ]P d x (11.45)
Ai (x) i (x) Ai (x) i (x)

open
????
By applying the definition (11.45) of the Poisson brackets and the constraints (11.42) and (11.43) we have

Z  
3 1x 2y 2y 1x
C1x,2y [1x , 2y ]P = d z
Ai (z) i (z) Ai (z) i (z)
Z "        #
3 i Ai (x) i i (y) + J 0 (y) i i (y) + J 0 (y) i Ai (x)
C1x,2y = d z
Ai (z) i (z) Ai (z) i (z)

Since according with Eq. (11.31), J 0 only depends on the matter canonical variables Qn and Pn , then J 0 does
not contribute to the Poisson brackets, thus
Z "     #
[ A p (x)] k (y) k (y)
[ Ap (x)]
p k k p
C1x,2y = d3 z
Ai (z) i (z) Ai (z) i (z)
Z  k

p
3 [p A (x)] k (y)
C1x,2y = d z
Ai (z) i (z)
by using Eq. (11.9) and assuming that the derivatives A are independent of the fields A we have
   
i Ai (x) Ai (x)
= i
Aj (z) Aj (z)
2
Recall that the constraints (11.33) were of first class, so that we were not able to use the formalism of Dirac brackets there.
3
What really matters is that this Poisson brackets obey the algebraic rules (10.219, 10.220, 10.221) page 317. So that we can apply
the results obtained in section 10.12.
11.3. QUANTIZATION IN COULOMB GAUGE 333

and similarly for the canonical field momenta


   
i i (y) i (y)
= i
Aj (z) Aj (z)

therefore we have [Homework!! C1 prove equation 11.46)]


Z    Z
3 Ap (x) k (y)   
C1x,2y = d z p i k = d3 z [pi p (x z)] ki k 3 (y z)
A (z) i (z)
Z
   
C1x,2y = d3 z i 3 (x z) i 3 (y z) = i i 3 (x y) (11.46)

C1x,2y = 2 3 (x y) (11.47)

close
We can do a similar procedure for the other Poisson brackets to obtain

C1x,2y = C2y,1x [1x , 2y ]P = 2 3 (x y)


C1x,1y [1x , 1y ]P = 0 ; C2x,2y [2x , 2y ]P = 0
 
0 2 3 (x y)
C= (11.48)
2 3 (x y) 0
The matrix CN M is non-singular, so the constraints (11.42) and (11.43) are of second class as we anticipated.
On the other hand, the field variables Ai can be written in terms of independent canonical variables. For instance,
we can take
Q1x A1 (x) , Q2x A2 (x)
as independent variables while A3 (x) is given by the solution of Eq. (11.42) that makes A compatible with the
Coulomb gauge.
Z x3
3
  
A (x) = ds 1 A1 x1 , x2 , s + 2 A2 x1 , x2 , s

in the same way, we can write i and Ai in terms of the canonical (independent) conjugates P1x , P2x and Q1x , Q2x
by using Eq. (11.43). As we said in section 10.12, it can be shown that if the independent variables Q1x , Q2x
and P1x , P2x satisfy the usual commutation relations, the commutators of the constrained variables and their
canonical conjugates are given by the Dirac brackets according with Eqs. (10.228) and (10.229). This prescription
is convenient since we do not have to write explicit expressions of the dependent variables in terms of independent
ones.
In order to calculate the Dirac brackets we start by writing the inverse of the Cmatrix. For this we first
write the quantity 2 3 (x y) in its Fourier representation
Z Z
2 3 d3 p ip(xy)
2 d3 p 2 ip(xy)
(x y) = e = e
(2)3 (2)3
Z
2 3 d3 p 2 ip(xy)
(x y) = p e
(2)3

from this it can be shown that the inverse of such a quantity is given by
Z
 1 d3 p eip(xy)
2 3 (x y) = (11.49)
(2)3 p2
334 CHAPTER 11. QUANTUM ELECTRODYNAMICS

now the matrix (11.48) has the texture


   
0 a 0 2 3 (x y)
C= =
a 0 2 3 (x y) 0
whose inverse has the form
 1
  1 !
0 a 0 2 3 (x y)
C= =  2 3 1 (11.50)
a1 0 (x y) 0

from Eqs. (11.49) and (11.50) the inverse of CN M reads


Z
1
 1
 d3 k eik(xy) 1
C 1x,2y
= C 2y,1x
= 3 2
=
(2) k 4 |x y|
1
 1

C 1x,1y
= C 2x,2y
=0 (11.51)

And the non-zero Poisson brackets of the Ai and i variables with the constraint functions yield
  3
Ai (x) , 2y P
= (x y) (11.52)
xi
3
[i (x) , 1y ]P = (x y) (11.53)
xi
now, the commutator can be evaluated by taking equations (10.228) and (10.229) for the Dirac brackets. Let us
evaluate the commutator of Ai (x) with j (y). From Eqs. (10.228) and (10.229) we have
 i        N M
A (x) , j (y) = i Ai (x) , j (y) D = i Ai (x) , j (y) P i Ai (x) , N P C 1 [M , j (y)]P
 i   i  1z,1w
= i A (x) , j (y) P i A (x) , 1z P C 1 [1w , j (y)]P
 i   1z,2w   2w,1z
i A (x) , 1z P C 1 [2w , j (y)]P i Ai (x) , 2w P C 1 [1z , j (y)]P
 i   2z,2w
i A (x) , 2z P C 1 [2w , j (y)]P

but Eqs. (11.52) and (11.53) give us the only non-zero Poisson brackets, then we have
 i      2w,1z
A (x) , j (y) = i Ai (x) , j (y) P i Ai (x) , 2w P C 1 [1z , j (y)]P
 i   i   i   2w,1z
A (x) , j (y) = i A (x) , j (y) P + i A (x) , 2w P C 1 [j (y) , 1z ]P
Z    
 i    1 3
A (x) , j (y) = i Ai (x) , j (y) P + i d3 w d3 z i 3 (x w) (y z)
x 4 |x y| y j
in the last step we take into account that we sum over repeated indices, that in the case of the indices z and w
implies integration. Further, taking into account that
 i 
A (x) , j (y) P = ji 3 (x y)

we finally obtain [Homework!! C2]


 
 i
 2 1
A (x) , j (y) = iji 3 (x y) + i j i
x x 4 |x y|
Therefore, from Eqs. (10.228) and (10.229), the equal-time commutators read
 
 i  i 3 2 1
A (x) , j (y) = ij (x y) + i j i (11.54)
x x 4 |x y|
 i j

A (x) , A (y) = [i (x) , j (y)] = 0 (11.55)
11.3. QUANTIZATION IN COULOMB GAUGE 335

which are consistent with the Coulomb gauge conditions (11.42) and (11.43), as it must be from the general
properties of the Dirac Bracket. Comparing (11.54) with (11.44), we see that the constraints prevent us for using
the usual quantization rules for the constrained variables. Indeed, we modify the quantization conditions for the
constrained variables, precisely to mantain the usual quantization rules on the independent variables.

11.3.2 ~
Quantization with the solenoidal part of
~ in electrodynamics. For the class of theories in which only the kinematic
We then wonder about the meaning of
term Z
1
d3 x F F
4
in the Lagrangian depend on A, by varying the Lagrangian with respect to A without considering the constraint
given by the Coulomb gauge (11.42), yields
Z Z
L 1 3 1 F (y)
j = = d y [F (y) F (y)] = d3 y F (y)
j
A (x) 4 j
A (x) 2 Aj (x)
Z Z  
1 3 [ A (y) A (y)] 1 3 [ A (y)] [ A (y)]
= d x F (y) = d x F (y)
2 Aj (x) 2 [0 Aj (x)] [0 Aj (x)]
Z
1   1
= d3 x 0 j 3 (y x) 0 j 3 (y x) F (y) = [0 j 0 j ] F (x)
2 2
1 1  
= [0 j F (x) + 0 j F (x)] = F 0j (x) + F 0j (x) = F j0 (x)
 2j 0 0 j
 0
2
j
= A (x) A (x) = j A (x) + 0 A (x)

L
j = = Aj (x) + j A0 (x) (11.56)
Aj (x) x
However, when the Coulomb condition is taken into account, variational derivatives with respect to A are not
well-defined. If the variation of L under A in A is given by
Z
L = d3 x P A (11.57)

let us assume an arbitrary scalar function F (x) with the only condition that it vanishes at infinity. Therefore, we
can construct a null integral via divergence theorem (in three-dimensions)
Z h i Z n o

0 = d x i F (x) Ai = d3 x
3 i
i F i (x) Ai + F i (x) i Ai
Z n o
0 = d3 x F (x) A + F (x) A
Z n o
0 = d3 x F (x) A

but taking into account that A = 0, we then have


Z n o
0 = d3 x F (x) A (11.58)

the we can add a zero on both sides of Eq. (11.57) through Eq. (11.58) to obtain
Z
L = d3 x [P + F (x)] A
336 CHAPTER 11. QUANTUM ELECTRODYNAMICS

for an arbitrary scalar function F (x) with the only condition that it vanishes at infinity. Thus, after applying the
Coulomb gauge condition, by examining the Lagrangian we only can say that
~ = A (x) + A0 (x) + F (x)
(11.59)
where the scalar F (x) is arbitrary. By applying condition (11.43) we can remove the ambiguity. Such a condition
along with Eq. (11.40) (valid only in the Coulomb gauge) requires
~ = J 0 = 2 A0 (11.60)
now, applying the divergence on Eq. (11.59) and taking into account Eq. (11.60) and that A = 0, we have
~ = 2 A0 (x) + 2 F (x) = 2 A0

it can be seen that Eq. (11.56) provides a correct (though perhaps not unique) expression for i .
~ does
The commutation relations (11.54, 11.55) are relatively simple but we should confront the fact that
n
not commute with matter fields and their canonical conjugates Q , Pn . Thus, if F is any functional of matter
~ yields
variables, its Dirac bracket with A is null (Homework!! C3), but its Dirac bracket with
h i h i  h i
~ (z) = F,
F, ~ (z) [F, N ] C 1 N M M , ~ (z) (11.61)
P
D P P
from the definition of Poisson brackets in our present context Eq. (11.45) page 332 it is clear that
h i
~ (z) = 0
F,
P
~ and A (it only depends on matter fields). With these considerations Eq. (11.61)
since F does not depend on
becomes
h i 1x,1y h i  h i
~ (z)
F, = [F, 1x ]P C 1 ~ (z) [F, 1x ] C 1 1x,2y 2y ,
1y , ~ (z)
P
D
 h iP 2x,2y h iP
1 2y,1x ~ (z) [F, 2x ] C 1 ~ (z)
[F, 2y ]P C 1x , P 2y ,
P P

but recalling that the only non-zero Poisson brackets between ~ and N are the ones in Eqs. (11.53) we have
h i  h i  h i
~ (z) = [F, 1x ] C 1 1x,1y 1y ,
F, ~ (z) [F, 2y ] C 1 2y,1x 1x ,
~ (z)
P P
D P P
but from Eq. (11.51) we see that the diagonal terms of C 1
are null. Hence
h i  h i
F, ~ (z) = [F, 2y ] C 1 2y,1x 1x ,
~ (z)
D P P
Besides, we should take into account that there is sum over repeated indices 2y and 1x. But we should recall that
these indices contain a continuous part and a discrete part, thus we shall also have integrals in x and y. From
these facts and using expression (11.51) of C 1 , we have
h i Z h i
~ (z) = d3 x d3 y [F, 2y ] 1 ~
F, P 4 |x y| 1x , (z)
D P

and interchanging the dummy indices x and y we obtain


h i Z h i
~ 1 ~ (z)
F, (z) = d3 x d3 y [F, 2x ]P 1y ,
D 4 |x y| P
Z    
3 3 i 0 1 k ~
= d x d y F, i (x) + J (x) k
A (y) , (z)
Z
x P 4 |x y| y P
  1
= d3 x d3 y F, J 0 (x) P 3 (y z)
4 |x y|
Z
 
= d3 y F, A0 (y) P 3 (y z)
   
= F, A0 (z) P = F, A0 (z) D
11.3. QUANTIZATION IN COULOMB GAUGE 337

in order to obtain a redefined canonical momentum whose Dirac bracket with F vanishes, it is natural to define
~ as
the solenoidal part of
~
~ A0 (11.62)
combining (11.56) con (11.62) we see that
 
~ j (x)
~ j (x) A0 (x) = Aj (x) + j A0 (x) j A0 (x) = Aj (x)
j
~ (x)
~ (x) A0 (x) = A (x) (11.63)

note one additional advantage of the use of ~ (x): Its relation with its canonical coordinate is simpler [compare
Eq. (11.63) with Eq. (11.56)]. We shall write the Hamiltonian in terms of A and ~ (instead of A and ) ~
because it is more convenient for the pass to the interaction picture. From this definition it is clear that
h i
~ (z)
F, = 0 (11.64)
D
h i h i
~ (x) , A0 (y) =
~ (x) ,
~ (y)
~ (y) = 0 (11.65)
 
i A0 (x) , j A0 (y) = 0 (11.66)

~ (x) obeys the same commutation relations as


from Eqs. (11.65) and (11.66) it can be verified that ~ (x).
~
Further by applying Eq. (11.60) we find that obeys a constraint
 
~ = ~ A0 = ~ 2 A0 = 2 A0 2 A0

~ = 0
(11.67)
~ is divergenceless, it justifies the name solenoidal part of .
since ~
-

11.3.3 Constructing the Hamiltonian


In order to construct the Hamiltonian we can use the customary relation between the Hamiltonian and the
Lagrangian using the constrained variables A and ~ , without first having to express explicitly the Hamiltonian
n
in terms of the unconstrained variables Q and Pn . For electrodynamics we obtain
Z h i Z h  i
H= d x 3 ~ i A + Pn Q L = d3 x
i n ~ i + A0 Ai + Pn Qn L

with a procedure similar to the one that lead to Eq. (11.58) by using A = 0, and taking F (x) A0 we can
show that the term with A0 vanishes. Hence
Z h i
H = d3 x ~ i Ai + Pn Qn L (11.68)

as we said before Qn and Pn will denote the matter canonical fields and their canonical conjugates respectively.
As a matter of example let us consider a Lagrangian density of the form
1
L = F F + J A + Lmatter (11.69)
4
where the current J does not depend on A , and Lmatter is the Lagrangian density that involves any other fields
that appear in J aside from their electromagnetic interactions [given explicitly by the term J A in Eq. (11.69)].
338 CHAPTER 11. QUANTUM ELECTRODYNAMICS

The electrodynamics of particles of spin 1/2 has the form of the Lagragian density (11.69). By substituting A
~ 2 [according with Eq. (11.63)], the Hamiltonian (11.68) becomes
everywhere with
Z  
3 ~ ~ i n 1
H = d x i + Pn Q + F F J A Lmatter
4
Z   Z h i
~ ~ 1
H = d3 x i + F F J A + d3 x Pn Qn Lmatter
i
4
Z   Z h i
~ 1 
H = d3 x 2
+ F0i F + Fi0 F + Fij F J0 A Ji A + d3 x Pn Qn Lmatter
0i i0 ij 0 i
4
Z   Z h i
~ 1 
H = d3 x 2
+ 2F0i F + Fij F + J A J A + d3 x Pn Qn Lmatter
0i ij 0 0
4
Z   Z h i
~ 1 1 2
H = d3 x + F0i F + ( A) + J A J A + d3 x Pn Qn Lmatter
2 0i 0 0
(11.70)
2 2

let us develop the term

1 1  1
F0i F 0i = (0 Ai i A0 ) 0 Ai i A0 = (0 Ai i A0 ) (0 Ai i A0 )
2 2 2
1 1
= (0 Ai ) (0 Ai ) + (0 Ai ) (i A0 ) (i A0 ) (i A0 )
2 2
1 ~  ~  ~ 1 1~ 2 ~
 1 2
= i i + i (i A0 ) (A0 )2 = A
0
A0
2 2 2 2
1 1 ~ 2
F0i F 0i = + A 0
(11.71)
2 2

substituting (11.71) in (11.70) the Hamiltonian becomes


Z  
3 ~ 2 1 2 1 ~ 0
2
0 0
H= d x + ( A) + A J A + J A + HM
2 2

with HM being the Hamiltonian for the matter fields, without their electromagnetic interactions
Z  
HM d3 x Pn Qn Lmatter

Applying the expression for A0 in Eq. (11.41) we have


Z  
1~2 1 1
H= d3 x + ( A)2 J A + J 0 A0 + HM (11.72)
2 2 2

from Eq. (11.41), it can be seen that the term (1/2) J 0 A0 is the Coulomb energy
Z Z Z
1 1 J 0 (x) J 0 (y)
VCoul = d3 x J 0 A0 = d3 x d3 y (11.73)
2 2 4 |x y|

~
By applying the commutation relations (11.54, 11.55) it can be checked that any operator function F of A and
obeys the field equation (Homework!! C4)
iF = [F, H]

as expected.
11.4. FORMULATION OF QED IN THE INTERACTION PICTURE 339

11.4 Formulation of QED in the interaction picture


As customary, we split the Hamiltonian (11.72) into a free term H0 and an interacting term V

H = H0 + V (11.74)
Z  
3 1~ 2 1 2
H0 = d x + ( A) + Hmatter,0 (11.75)
2 2
Z
V = d3 x J A + VCoul + Vmatter (11.76)

where Hmatter,0 is the free-particle term for matter and Vmatter its corresponding interacting term. Finally, VCoul is
the Coulomb interaction term (11.73). Recalling that the total Hamiltonian H in Eq. (11.74) is time-independent,
we can evaluate H0 and V in Eqs. (11.75) and (11.76) at any time. In particular we can evaluate those terms at
t = 0. As usual, the transition to the interaction picture is carried out through a similarity transformation

h i
~ , Q, P
V (t) = exp (iH0 t) V A, exp (iH0 t)
t=0
= V [a (t) , (t) , q (t) , p (t)]

Where we are omitting the subscript on (x). Here P means the canonical conjugates to the matter fields
Q. If O (x) is an operator in the interaction picture we denote as O (x), its counterpart in the Heisenberg picture.
In particular, O (x, t) in the interaction picture can be expressed by its value O (x, 0) at t = 0 in the Heisenberg
picture as
O (x, t) = exp [iH0 t] O (x, 0) exp [iH0 t]

deriving with respect to time on both sides


 
O (x, t) = iH0 exp [iH0 t] O (x, 0) exp [iH0 t] + exp [iH0 t] O (x, 0) exp [iH0 t] (iH0 )
O (x, t) = iH0 {O (x, t)} iO (x, t) H0

then its equation of motion yields


iO (x, t) = [O (x, t) , H0 ] (11.77)

of course, a similarity transformation does not change the equal-time commutation relations, so they are the same
as in the Heisenberg picture
 
  2 1
ai (x, t) , j (y, t) = i ij 3 (x y) + i j (11.78)
x x 4 |x y|
 i j
  i j

a (x, t) , a (y, t) = (x, t) , (y, t) = 0 (11.79)

and similarly for the matter fields and their canonical conjugates. For the same reason, the constraints (11.37)
and (11.67) preserve their form (recall that ~ means ~ in this context)

a = 0 ; ~ = 0 (11.80)

now, we can find the relation between ~ and a, by evaluating a through Eq. (11.77) and using the commutation
340 CHAPTER 11. QUANTUM ELECTRODYNAMICS

relations (11.78) and (11.79) we have


 Z   
1 2 1
iai (x, t) = [ai (x, t) , H0 ] = ai (x, t) , d3 y ~ (y) + ( a (y))2 + Hmatter,0
2 2
 Z   Z
3 1 2 1 2 1  
= ai (x, t) , d y ~ (y) + ( a (y)) = d3 y ai (x, t) , ~ 2 (y)
2 2 2
Z
1
= d3 y {[ai (x, t) , j (y)] j (y) + j (y) [ai (x, t) , j (y)]}
2
Z    
3 3 2 1
= d y i ij (x y) + i j j (y)
x x 4 |x y|
and we obtain
Z  
2 1
iai (x, t) = [ai (x, t) , H0 ] = i d3 y ij 3 (x y) + j (y, t)
xi xj 4 |x y|

we can replace /xj by /y j , and integrate by parts then use the second of Eqs. (11.80) to obtain [Homework!!
C5]
a = ~ (11.81)
like in the Heisenberg picture [see Eq. (11.63)]. Similarly, the field equation is given by
 Z   
3 1 2 1 2
ii (x, t) = [i (x, t) , H0 ] = i (x, t) , d y ~ (y) + ( a (y)) + Hmatter,0
2 2
Z h i
1
= d3 y i (x, t) , ( a (y))2
2
Z  
3 3 2 1
ii (x, t) = [i (x, t) , H0 ] = i d y ij (x y) + i j ( a (y, t))j (11.82)
x x 4 |x y|
from Eq. (11.82) and using the first of Eqs. (11.80) and Eq. (11.81) we arrive at the usual wave equation

a=0 (11.83)

recalling that in the Heisenberg picture A0 is not an independent variable but a functional (11.41) of the matter
fields and their canonical conjugates that vanishes in the limit of zero charges, we do not introduce a corresponding
degree of freedom in the interaction picture. Instead, for convenience we define

a0 = 0 (11.84)

The most general real solution of the first of Eqs. (11.80), along with Eqs. (11.83) and (11.84) gives
Z
1 d3 p X h ipx i

a (x) = 3 p e e (p, ) a (p, ) + eipx e (p, ) a (p, ) (11.85)
2 2p0

with p0 = kpk. The coefficients e (p, ) are two independent degrees of freedom that satisfy the relations

p e (p, ) = 0 ; e0 (p, ) = 0 (11.86)

we call these coefficients polarization vectors, name that we shall justify later. Again, a (p, ) are a pair of
operator coefficients, where is a two-valued index. By an appropriate normalization of a (p, ) we can normalize
the polarization vectors e (p, ) so that the completeness relation gives
X pi pj
ei (p, ) ej (p, ) = ij (11.87)
kpk2
11.4. FORMULATION OF QED IN THE INTERACTION PICTURE 341

in particular we can choose the polarization vectors that we obtained in chapter 8



1/ 2
i/ 2
p)
e (p, 1) = R (b

(11.88)
0
0

where R (b b . From Eqs. (11.87)


p) is the standard rotation that takes the unitary vector u3 into the direction p
and (11.81) we obtain that the commutation relations (11.78) and (11.79) are fulfilled if and only if the operator
coefficients in Eq. (11.85) hold the conditions [Homework!! C6]
h i
a (p, ) , a q, = 3 (p q) (11.89)
h i
a (p, ) , a q,
= 0 (11.90)

as we have said several times, this is not an alternative derivation of Eqs. (11.89) and (11.90) but a validation of
Eq. (11.75) as the correct free Hamiltonian for massless particles of helicity 1. In a similar way we can apply
Eqs. (11.81), and (11.85) in Eq. (11.75) to calculate the free-photon Hamiltonian (Homework!! C7)

Z X1 h i
H0 = d3 p p0 a (p, ) , a (p, )

2 +
Z X  1

H0 = d3 p p0 a (p, ) a (p, ) + 3 (p p) (11.91)

2

which has the expected form of the free-Hamiltonian in terms of creation and annihilation operators [see Eq.
(10.32) page 277], plus an irrelevant divergent term. Thus, Eq. (11.91) completes the validation of Eq. (11.75) as
a correct expression for the Free Hamiltonian.
The interaction term in the Heisenberg picture is given by Eq. (11.76), the corresponding interaction term in
the interaction picture reads
Z
V (t) = d3 x j (x, t) a (x, t) + VCoul (t) + Vmatter (t) (11.92)

in Eq. (11.92) we write a j instead of a j because a0 = 0. With respect to the current in the Heisenberg picture
J , the associated current j in the interaction picture reads

j (x, t) exp [iH0 t] J (x, t) exp [iH0 t]

and VCoul (t) is the associated Coulomb term

VCoul (t) = exp [iH0 t] VCoul exp [iH0 t]


Z
1 j 0 (x, t) j 0 (x, t)
VCoul (t) = d3 x d3 y (11.93)
2 4 kx yk

finally, (in the interaction picture) Vmatter (t) is the non-electromagnetic part of the matter field interaction

Vmatter (t) = exp [iH0 t] Vmatter exp [H0 t]


342 CHAPTER 11. QUANTUM ELECTRODYNAMICS

11.5 The propagator of the photon


According with the Feynman rules discusssed in chapter 9, an internal photon line in a Feynman diagram provides
a factor in the Smatrix associated with the process, with the propagator

i (x y) h0| T {a (x) , a (y)} |0i (11.94)

with T denoting the time-ordered product. Substituting the expansion (11.85) in (11.94) the propagator becomes
Z h i
d3 p ip(xy) ip(yx)
i (x y) = P (p) e (x y) + e (y x) (11.95)
(2)3 (2 kpk)
with
X
P (p) e (p, ) e (p, ) ; p0 = kpk
=1

From Eqs. (11.87) and (11.86) we have

pi pj
Pij (p) = ij ; P0i (p) = Pi0 (p) = P00 (p) = 0 (11.96)
kpk2
We showed in Chapter 9, that the theta function in Eq. (11.95) can be written in terms of integrals over an
independent time component q 0 belonging to an off-shell four-momentum q . From this, equation (11.95) can be
expressed by an integral over a four-momentum but with the off-mass-shell condition i.e. q 0 is independent of q
Z
4 P (q) iq(xy)
(x y) = (2) d4 q 2 e
q i
we are considering an internal photon line carrying a four-momentum q running between the vertices in which the
photon is created and destroyed by fields a and a . According with the Feynman rules in momentum space, the
contribution of such an internal line gives
i P (q)
(2)4 q 2 i
It is convenient to reexpress Eq. (11.96) as follows

q 0 q n + q 0 q n q q + q 2 n n
P (q) = g + (11.97)
kqk2
2
n (0, 0, 0, 1) ; q 2 = q2 q 0 with q 0 arbitrary (11.98)

thus n is a constant time-like vector. Since q 0 is arbitrary we shall determine it in Eq. (11.97) from four-
momentum conservation. That is, it will be taken as the difference of the matter p0 s flowing in and out of the
vertex in which the photon is created. Thus, the terms proportional to q and/or q do not contribute to the
Smatrix. It owes to the fact that the factors q or q act like derivatives and , while the photon fields a
and a are coupled to currents j and j that satisfy the conservation condition j = 0. On the other hand, the
term proportional to n n has a factor q 2 that is cancelled by the factor q 2 in the denominator of the propagator4 ,
giving a term that coincides with the one that would be generated by the term in the action
Z Z Z
1 4 4
 0
 0
 i d4 q iq(xy)
i d x d y i j (x) i j (y) e
2 (2)4 kqk2
4 q2
A term of the form q 2 i
= 1
1i 2
= 1
1i
shows that we can cancel q 2 completely, after all what really matters for it that it
q
tends to zero from the positive side.
11.6. FEYNMAN RULES IN SPINOR ELECTRODYNAMICS 343

the integration over q 0 provides a delta function in time. Thus, this is equivalent to a correction for V (t) given by
Z Z
1 j 0 (x, t) j 0 (y, t)
d3 x d3 y
2 4 kx yk

which is the precise term to cancel the Coulomb interaction Eq. (11.93). In summary, the covariant quantity
Z
ef f 4 g iq(xy)
(x y) = (2) d4 q 2 e
q i

can be used as an effective photon propagator. That is, the Coulomb term does not appear henceforth. We
can see that the apparent violation of Lorentz invariance in the (instantaneous) Coulomb interaction is cancelled
by another apparent violation of the Lorentz invariance coming from the fact that the fields a are not truly
four-vectors, leading to a non-covariant propagator. After this cancellation is carried out, we are left with a term
of the form
i g
4 q 2 i
(2)
as the contribution of an internal photon line in the momentum space Feynman rules. The Coulomb interaction
term is then absent in the rule.

11.6 Feynman rules in spinor electrodynamics


We shall study the electrodynamics of a single species of particles of spin 1/2, charge q = e and mass m. We
shall make the treatment for electrons but it works equally well for muons, taos and other similar particles.
We shall start with the simplest gauge and Lorentz invariant Lagrangian density for this theory. Such a theory
contains a kinetic term for the photons and a (locally gauge invariant) Diracs Lagrangian density. Since the
Diracs Lagrangian density is locally gauge invariant, it will contain a covariant derivative D . Therefore, the
Diracs Lagrangian density contains the matter radiation coupling plus the matter part of the Lagrangian density

1
L = F F ( D + m)
4
1
L = F F ( [ + ieA ] + m) (11.99)
4
The four-vector that describes the electric current is given by

L
J = = ie (11.100)
A

and the interaction term (11.92) in the interaction picture yields


Z
 
V (t) = ie d3 x (x, t) (x, t) a (x, t) + VCoul (t) (11.101)

and we have no a Vmatter term5 . As already discussed, the Coulomb term cancels with the non-covariant part of
the photon propagator (where both terms are local in time i.e. instantaneous).
We shall then define the Feynman rules in the momentum space to calculate the connected part of the Smatrix
in this spinor electrodynamics.
5
That is this Lagrangian does not contain terms of interaction with only matter fields (the purely matter terms contains the matter
propagator and the mass term for fermions). All interaction terms are of the matter-radiation type.
344 CHAPTER 11. QUANTUM ELECTRODYNAMICS

11.6.1 Drawing the Feynman diagrams


We start by drawing all connected Feynman diagrams for a given number of vertices.

1. There are two types of lines: electron lines carrying arrows and drawed as continuous lines, and photon lines
that do not carry arrows and represented by waving lines6 . Lines are joined at vertices

2. Each vertex contains three lines: one incoming electron line, one outcoming electron line and one photon
line.

3. For each initial particle we have an external line coming from below into the diagram.

4. For each final particle we have an external line going upwards out of the diagram.

5. Electrons in the diagram correspond to lines with arrows pointing upwards (regardless whether they are
internal lines or external lines into or out of the diagram).

6. Positrons correspond to lines with arrows pointing downwards (regardless whether they are internal lines or
external lines into or out of the diagram).

7. There are as many internal lines as the ones to attach each vertex to the required number of lines (three
lines for each vertex).

8. Each internal line will be labeled by an off-mass-shell four-momentum flowing in a given sense on the line
(conventionally in the same direction of the arrow for electron lines)7

9. Each external line is labelled with the momentum and third component of spin or helicity for an electron or
a photon respectively, in the initial and final states.

11.6.2 Factors associated with vertices


For a given vertex we shall have three labels (one for each attached line): (a) one four-component Dirac index
at the electron line with arrow coming into the vertex (b) A four-component Dirac index at the electron line
with arrow going out of the vertex, and (c) a space-time vertex associated with the photon line. Then from
these three labels , , ; we construct an associated vertex factor given by

(2)4 e ( ) 4 k k + q (11.102)

with k and k denoting the electron four-momenta entering and leaving the vertex respectively, and q denotes the
photon four-momentum entering the vertex (if the photon four-momentum is leaving the vertex, we put a minus).

11.6.3 Factors associated with external lines


1. We label a given external line with the three-momentum p and third component of spin or helicity (for
electrons and photons respectively) of the particle in the initial or final state.

2. For an electron line in the final state going out of a vertex with a Dirac label on this line, we associate a
factor
u (p, )
(11.103)
(2)3/2
6
In general, it is conventional to draw scalar bosons as dashed lines, fermions as continuous lines, and vector bosons as waving lines.
7
The arrow mentioned before (upward for electrons and downward for positrons), defines what we call the fermionic flux. The
arrows of the fermionic flux do not necessarily coincide with the arrows of the momentum flux. It is convenient (but not mandatory)
to make them coincide.
11.6. FEYNMAN RULES IN SPINOR ELECTRODYNAMICS 345

observe that we have extracted a matrix from the interaction (11.102). It is because of this, that u and v
will appear in the Feynman rules instead of u and v . The four-component spinors u (p, ) and v (p, ) are
the ones dicussed in section 7.4.
3. For a positron line in the final state coming into a vertex with a Dirac label on this line, we associate a
factor
v (p, )
(2)3/2
4. For an electron line in the initial state coming into a vertex with a Dirac label on this line, the associated
factor will be
u (p, )
(2)3/2
5. For a positron line in the initial state going out of the vertex with a Dirac label on this line, the factor
associated yields
v (p, )
(2)3/2
6. For a photon line in the final state attached to a vertex with space-time index on this line, the factor is
e (p, )
p
(2)3/2 2p0
where e (p, ) are the polarization vectors described in section 11.4.
7. For a photon line in the initial state attached to a vertex with space-time label on this line, the factor
reads
e (p, )
p
(2)3/2 2p0
If we call u (p, ) and v (p, ) as spinors, and u (p, ) and v (p, ) as adjoint spinors, we could say that we
use spinors when the arrow of the electron line comes into a vertex and we use adjoint spinors when the electron
arrow goes out of the vertex. Further we use u (p, ) or u (p, ) for electrons (particles i.e. upward lines) and
v (p, ) or v (p, ) for positrons (antiparticles i.e. downward lines).
Note also that the fermionic flux never changes its sense so that it forms a fermionic current. For instance,
the two fermion arrows in a given vertex never goes both out (or both come into) the vertex, always one of them
is entering and the other one is leaving the vertex.

11.6.4 Factors associated with internal lines


From now on we shall use a Dirac slash notation for a given four vector k defined by
6 k k (11.104)
1. For each internal electron line labelled by a four momentum k and running from a vertex with a Dirac label
to another vertex with Dirac label , the factor is given by
i [i 6 k + m]
(2)4 k2 + m2 i

2. For each internal photon line labelled by a four-momentum q running between two vertices with spacetime
labels and the factor gives
i g
4 q 2 i
(2)
346 CHAPTER 11. QUANTUM ELECTRODYNAMICS

11.6.5 Construction of the Smatrix process


To construct the Smatrix assocaited with the process we then proceed as follows

1. Integrate the product of all the previous factors over the four-momenta associated with internal lines (since
they are off the mass shell), and sum over all Dirac and spacetime indices (we usually used the convention
of sum over repeated indices).

2. Add up the results above for each Feynman diagram.

3. Some additional combinatoric factors and fermionic signs might appear as discussed in section 9.4.

11.7 General features of the Feynman rules for spinor QED


As the number of internal lines and vertices increases, there is a corresponding increase in the difficulty of evalauting
the diagram. Therefore, it is useful to have some idea of the numerical order of magnitude of the contribution of
the diagram. We shall estimate these numerical factors by including factors e coming from the electronic charge
associated with the vertices and also factors 2 and that comes from vertices, propagators, and momentum space
integrals.
We have already seen some diagrammatical identities concerning the number of vertices V , internal lines I,
external lines E and loops L. They are given by [see Eqs. (9.82) and (9.98)) pages 252, 262]

L = I V + 1 ; 2I + E = 3V (11.105)

solving for I on both sides we have

3V E
L+V 1 =
2
V = 2L + E 2 (11.106)

we have a factor e (2)4 from each vertex, a factor (2)4 from each internal line, and a fourdimensional momen-
tum space integral for each loop. The element of volume in the four-dimensional euclidean space with a radius
parameter k is 2 k2 dk2 . Consequently, each loop gives a factor 2 . From these facts and using Eqs. (11.105,
11.106), the diagram contains a numerical factor F given by
" #
h iV 
1
I
 2 L h iV  1 I (2)4 L
4 V 4
F = e (2) = [e] (2)
(2)4 (2)4 16 2
h iV I+L  1 L   
1 L

2 L E2
= [e]2L+E2 (2)4 = e e (2) 4
16 2 16 2
 L
4 E2 e2
F = (2) e (11.107)
16 2
the number E of external lines is the same for each diagram associated with the same physical process. Then
from Eq. (11.107), we find that the expansion parameter that gives us an idea of the suppression of the Feynman
diagrams for each additional loop reads

e2
2
= = 5.81 104
16 4
this number is small enough to ensure a good perturbative behavior of the Feynman diagrams.
11.7. GENERAL FEATURES OF THE FEYNMAN RULES FOR SPINOR QED 347

11.7.1 Photon polarization


We should take into account the spin states in which photons and electrons are typically found in experiments. In
most of experiments the electrons and photons have no well-defined values of the zcomponent of spin neither the
helicity. In the case of photons, they are usually found in a state of transverse or elliptical polarization (instead
of helicity states). For photons states with well-defined helicity the polarization vectors are given by Eq. (11.88)

1/ 2
i/ 2
p)
e (p, 1) = R (b

(11.108)
0
0

the most general photon state is a linear superposition of helicity states8

+ p,+1 + p,1 (11.109)

if the helicity states are normalized, the normalization of this superposition yields

|+ |2 + | |2 = 1 (11.110)

If we want to calculate the Smatrix element for absorbing or emitting a photon in the state (11.109) we should
replace the vector polarization e (p, 1) (that describes a helicity eigenstate) with

e (p) = + e (p, +1) + e (p, 1) (11.111)

in the Feynman rules. The polarization vectors with well defined helicity satisfy the normalization condition

e p, e (p, ) = (11.112)

so that if the normalization condition (11.110) is fulfilled, the polarization vector (11.111) also satisfies a normal-
ization condition of the type
e (p) e (p) = 1
we have two extreme cases of polarizations of the type(11.111) (a) when either + = 0 or = 0 corresponding
to circular polarization, (b) the case |+ | = | | = 1/ 2, corresponding to linear polarization.
For linear polarization, by properly choosing the overall phase of the state (11.109), we can settle the coefficients
+ and to be complex conjugates
1
= exp (i) (11.113)
2
and substituting (11.113, 11.108) in (11.111) the polarization vector yields
i

e /2 ei /2

+iei /2 iei /2

e (p) = + e (p, +1) + e (p, 1) = R (b p)
+




0 0

0 0

cos
sin
p)
e (p) = R (b 0
(11.114)
0
8
Of course we are still assuming it as a three-momentum eigenstate.
348 CHAPTER 11. QUANTUM ELECTRODYNAMICS

then we can use the polarization vector (11.114) in the Feynman rules. So that is the azimuthal angle of the
photon polarization in the plane perpendicular to p [recall that e (p, ) is perpendicular to p, according with Eq.
(11.86) page 340]. The photon polarization vector (11.114) is real, and it is only possible for linear polarization.
The most general state correspond to elliptical polarization in the case in which both + and are non-zero
and they are non-equal.
In a more general experimental context, the initial photons could be prepared in a statistical mixture of helicity
(r)
states. In the most general case an initial photon could have any number of possible polarization vectors e (p)
each one with probability Pr . The rate of abosortion of such a photon in a given process will have the form9
X 2 X    

= Pr e(r)
(p) M = Pr e(r)
(p) M
e(r)
(p) M

r r
X

= M M Pr e(r)
(p) e(r)
(p)
r

X 2

= Pr e(r)
(p) M = M M (11.115)
r
X
Pr e(r) (r)
(p) e (p) (11.116)
r

where is thePdensity matrix. Let us recall that is a Hermitian positive matrix of unit trace (that account on
the fact that r Pr = 1). Further, from Eqs. (11.86) the density matrix satisfies

0 = 0 = 0 and p = p = 0 (11.117)

the matrix density in its canonical form, can be expressed as


X
= s e (p; s) e (p; s)
s=1,2

where e (p; s) are the two orthonormal eigenvectors of , and s denote the associated eigenvalues, with

e0 (p; s) = e (p; s) p = 0

owing to the positivity of the matrix and the conservation of probability (unit trace), the associated eigenvalues
s must satisfy
X
s 0 ; s = 1
s=1,2

therefore, the rate for the photon absorption process can be reexpressed by
X
= s |e (p; s) M |2
s=1,2

so that the existence of a statistical mixture of initial photon states is equivalent to the superposition of two
orthonormal polarizations e (p; s) with probabilities s .
In many experiments the photons and electrons are not prepared in well defined polarization states, neither is
measured the final polarization states of such particles.
9
Note that the expression (11.115) for the rate of absorption is a sum over probabilities and not a sum over amplitudes, indicating
that we are considering a statistical mixture of helicity states instead of a linear superposition of quantum helicity states.
11.7. GENERAL FEATURES OF THE FEYNMAN RULES FOR SPINOR QED 349

If we have no information about the initial photon polarization, we consider that both helicity states have the
same probability 1 = 2 = 12 . Using that assumption and Eq. (11.87), the density matrix (and so the absorption
rate) which is an average over both possible initial polarizations, becomes
1 X 1
ij = ei (p; s) ej (p; s) = (ij pbi pbj )
2 2
s=1,2

where we have used the fact that the unitary transformation that takes the density matrix to its canonical
form, leaves the identity invariant as well as the factor pbi pbj since this rotation is in the helicity space but not
in the momentum space. Consequently, this result is independent of the particular pair of polarization vectors
ei (p; s) chosen to make the average. Thus, for unpolarized photons the absorption rate can be averaged over any
pair of orthonormal polarization vectors. Further if we also ignore the final polarization state of the photon, we
can calculate the rate by summing over any pair of orthonormal final photon polarization vectors.

11.7.2 Electron and positron polarization


A similar result follows for electrons and positrons. If they are not prepared in well defined states of spin, the rate
is calculated by averaging over any two orthonormal initial spin states10 , such as those with z component of spin
equal to = 1/2. If we have no information about the final spin states of the electrons and positrons, we sum
the rate over any two orthonormal spin states, such as the ones with zcomponent of spin given by = 1/2.
Those sums are obtained by using Eqs. (7.102, 7.103) as well as Eqs. (7.121, 7.122)
X  
i 6 p + m
u (p, ) u (p, ) = (11.118)

2p0
X  
i 6 p m
v (p, ) v (p, ) = (11.119)

2p0
p
where p0 = p2 + m2 . As a matter of example, if we have an electron in the state |p, i and a positron in the
state |p , i in the initial state, the Smatrix of the process will have the form
v (p, ) M u (p, )
if the electron and positron spin states are not determined, the rate is proportional to
1 1 X  2
v p , M u (p, )
2 2
,

where each 1/2 factor comes from the average of electron and positron initial spin states.
1 X  2 1 X      
v p , M u (p, ) = v p , M u (p, ) v p , M u (p, )
4 4
, ,
1 Xh  i  
= u (p, ) M v p , v p , M u (p, )
4
,
 h 
1X
 i   
= u (p, ) M v p , v p , M u (p, )
4
,
1 X n o   
= u (p, ) M v p , v p , M u (p, )
4
,
1 Xn o   
= u (p, ) M v p , v p , M u (p, )
4
,
10
This is another way to say that if we have no information about spin states, any direction can be equally well chosen to be the
direction of the axis of quantization.
350 CHAPTER 11. QUANTUM ELECTRODYNAMICS

where we have inserted a factor 2 = 1, in the last step

1 Xn o   
u (p, ) M v p , v p , M u (p, )
4
,
1 X n o  
= u (p, ) M v p , v p , M u (p, )
4
,
( )( )
1 X X   n o
= u (p, ) u (p, )
v p , v p ,
M M
4

and using Eqs. (11.118) and (11.119), we have


    n o
1 i 6 p + m i 6 p m
M M
4 2p0 2p0
(     )
1 i 6 p m i 6 p + m
= M M
4 2p0 2p0

    
1 X
 2 1
i 6 p m i 6 p + m
v p , M u (p, ) = T r M M (11.120)
4 4 2p0 2p0
,

11.8 Example of application: Feynman diagrams for electron-photon (Comp-


ton) scattering

Figure 11.1: At the tree-level, there are two diagrams that contribute to the electron-photon (Compton) scattering.

At the tree level, there are two Feynman diagrams for electron-photon scattering shown in Fig. 11.1. Let us
write the Feynman rules for the diagram in Fig. 11.1(a).
11.8. EXAMPLE OF APPLICATION: FEYNMAN DIAGRAMS FOR ELECTRON-PHOTON (COMPTON) SCATTER

According with the rules explained above, for the initial electron line coming into a vertex we have a Dirac
label on this line, with momentum p and third component of spin : Thus, this external line gives a factor
u (p, )
(2)3/2
the initial photon state has label on this line, momentum k and helicity e . This external line provides a factor
e (k)

(2)3/2 2k0
now for the vertex 1, we have the Dirac indices , for the electron lines coming into and going out of the vertex
respectively, the photon index is . The momenta p and k are entering, while q is leaving the vertex. Hence, the
vertex factor yields
(2)4 e ( ) 4 (p + k q)
the internal electron line is labelled by a momentum q running from a vertex with Dirac label to a vertex with
Dirac label . The factor of the internal line becomes
i [i 6 q + m]
(2)4 q 2 + m2 i
now for the vertex 2, we have Dirac indices , for the electron lines coming into and going out of the vertex
respectively, the photon index is . The momenta p and k are leaving, while q is entering the vertex. Hence,
the vertex factor yields 
(2)4 e ( ) 4 p k + q
the final state of electron goes out of the vertex with Dirac label on this line. Further, its momentum is p and
its third component of spin is , so the factor gives
u (p , )
(2)3/2
the final photon line state has space-time index on this line with momentum k and helicity e thus the factor
gives
e
(k )

(2)3/2 2k0
putting all these factors together we find
! !  
, ) h i
e
(k) u (p 4 4 i [i 6 q + m]
P (2) e ( ) p k + q
(2)3/2 2k0 (2)3/2 (2)4 q 2 + m2 i
! !
h i e (k) u (p, )
4 4
(2) e ( ) (p + k q)
(2)3/2 2k0 (2)3/2

the amplitude associated with this diagram is obtained by summing over Dirac and photon indices , , , , ,
and integrating over momenta of internal lines which is this case is only q.

! !
X X XZ e u (p , ) h i
4 (k ) 4 4
Ma = d q (2) e ( )
p k + q
, , , (2)3/2 2k0 (2)3/2
  ! !
i [i 6 q + m] h i e (k) u (p, )
(2)4 e ( ) 4 (p + k q)
(2)4 q 2 + m2 i (2)3/2 2k0 (2)3/2
352 CHAPTER 11. QUANTUM ELECTRODYNAMICS

using convention of sum over repeated indices and reordering terms we have

Z   
u (p , )
e
(k )
u (p, ) e (k) 4 i i 6 q + m
Ma = d q
(2)3/2 (2)3/2 2k0 (2)3/2 (2)3/2 2k0 (2)4 q 2 + m2 i
h i h i
e (2)4 ( ) 4 q p k e (2)4 ( ) 4 (q p k) (11.121)

adding the contribution of the second diagram we obtain the tree-level contribution of the Smatrix element for
Compton scattering.
Z   

 u (p , ) e
(k ) u (p, ) e (k) 4 i i 6 q + m
S (p,; k, e) p , ; k , e = d q
(2)3/2 (2)3/2 2k0 (2)3/2 (2)3/2 2k0 (2)4 q 2 + m2 i
nh  i h i
4 4 4 4
e (2) ( ) q p k e (2) ( ) (q p k)
h  ih io
+ e (2)4 ( ) 4 q + k p e (2)4 ( ) 4 q + k p (11.122)

11.9 Calculation of the cross-section for Compton scattering


We have shown the Feynman diagrams for electron-photon (Compton) scattering to lowest order (tree-level) in
e. Integrating over the internal line momentum in Eq. (11.122) we replace q p + k for the first diagram and
q p k for the second

h i2
ie2 (2)4 1  
S = h i4 4
u p , e
k u (p, ) e (k)
(2)3/2 (2) 2k0 2k0
(  h
i (6 p+ 6 k) + m 4
i
( ) p + k p k ( )
(p + k)2 + m2 i
  h )
i (6 p 6 k ) + m 4
i
+ ( ) p k + k p ( )
(p k )2 + m2 i
factorizing the delta and simplifying factors
ie2 4 (p + k p k ) 
S = 2
u p ,
(2) 2k0 2k0
(  

 i (6 p+ 6 k) + m
e k e (k) ( ) ( )
(p + k)2 + m2 i
  )

 i (6 p 6 k ) + m
+ e k e (k) ( ) ( ) u (p, )
(p k )2 + m2 i
reordering terms
ie2 4 (p + k p k ) 
S = 2
u p ,
(2) 2k0 2k0
(  

 i (6 p+ 6 k) + m
e (e )
(p + k)2 + m2 i
  )
i (6 p 6 k ) + m 
+ (e ) e
u (p, )
(p k )2 + m2 i
11.9. CALCULATION OF THE CROSS-SECTION FOR COMPTON SCATTERING 353

using the notation (11.104) and defining


6 e e
where we emphasize that 6 e is not (6 e) . With this notation we find

ie2 4 (p + k p k ) 
S = 2
u p ,
(2) 2k 0 2k 0
(  

 i (6 p+ 6 k) + m
6 e (6 e)
(p + k)2 + m2 i
  )
i (6 p 6 k ) + m 
+ (6 e) 6 e u (p, )
(p k )2 + m2 i

and rewriting the result in matrix notation gives


  
ie2 4 (p + k p k)
 i (6 p+ 6 k) + m
S = h i u p , 6 e 6e
(2)2 2 k0 k0 (p + k)2 + m2
  
i (6 p 6 k ) + m
+6e 6 e u (p, ) (11.123)
(p k )2 + m2
where we have dropped the factor i since the denominators here do not acquire a singularity. Since external lines
refer to on-shell particles and phtons are massless, we have

p2 = m2 ; k2 = k2 = 0

with m being the electron mass. Thus, the denominators become

(p + k)2 + m2 = p2 + k2 + 2p k + m2
= m2 + 0 + 2p k + m2 = 2p k

and similarly for the other denominator, hence

(p + k)2 + m2 = 2p k (11.124)
2
p k + m2 = 2p k (11.125)

The Compton Smatrix element becomes


  
4
 e2
 i (6 p+ 6 k) + m
S = 2i p + k p k h i u p , 6 e 6e
(2)3 2 k0 k0 2p k
  
i (6 p 6 k ) + m
6e 6 e u (p, ) (11.126)
2p k

11.9.1 Feynman amplitude for Compton scattering


The Feynman amplitude M defined in Eq. (2.61) page 76 is given by

S = 2i4 p + k p k M (11.127)

comparing (11.127) with (11.126) the Feynman amplitude becomes


     
e2
 i (6 p+ 6 k) + m i (6 p 6 k ) + m
M= u p , 6 e 6 e 6 e 6 e u (p, ) (11.128)
4 (2)3 k0 k0 pk p k
354 CHAPTER 11. QUANTUM ELECTRODYNAMICS

To calculate the square of this amplitude we additionally sum over and because we shall later use an
average over initial states and sum over final states. Then we shall utilize the normalization
X (i 6 p + m)
u (p, ) u (p, ) =

2p0

Further, for an arbitrary 4 4 matrix M, with a procedure similar to the one used to obtain Eq. (11.120) we have
X  X  h i
u p , M u (p, ) 2 = u p , M u (p, ) u (p, ) M u p ,
, ,
X    
= M u (p, ) u (p, ) M u p , u p ,

,
X     

u p , M u (p, ) 2 i 6 p + m i 6 p + m
= Tr M M (11.129)
2p0 2p0
,

Combining Eqs. (11.128) and (11.129) the square of the amplitude (sum over spin states) yields
      2
X e4 X 
2
|M | = u p , 6 e i (6 p+ 6 k) + m 6 e 6 e i (6 p 6 k ) + m 6 e u (p, )
,
16 (2)6 k0 k0 , pk p k
      
e4 i (6 p+ 6 k) + m i (6 p 6 k ) + m i 6 p + m
= Tr 6 e 6 e 6 e 6e
16 (2)6 k0 k0 pk p k 2p0
        )
i (6 p+ 6 k) + m i (6 p 6 k ) + m i 6 p + m
6e 6 e 6 e 6e (11.130)
pk p k 2p0

and recalling the property (7.49) page 188 = , we have


 
6 z = (z ) = z = z = z
6 z = 6 z (11.131)

more generally for a product of terms with slash

(6 z1 6 z2 6 z3 . . . 6 zn ) = 6 zn . . . 6 z3 6 z2 6 z1 = 6 zn . . . 6 z3 6 z2 6 z1
     
= 6 zn . . . 6 z3 6 z2 6 z1
= ( 6 zn ) . . . ( 6 z3 ) ( 6 z2 ) ( 6 z1 )
(6 z1 6 z2 6 z3 . . . 6 zn ) = (1)n 6 zn . . . 6 z3 6 z2 6 z1 (11.132)

from the property (11.132) and taking into account that p, p , k, k are all real, the first three products of the last
line of Eq. (11.130) becomes
     
i (6 p+ 6 k) + m i (6 p 6 k ) + m
F 6e 6 e 6 e 6e
pk p k
   
i 6 e (6 p+ 6 k) 6 e + m 6 e 6 e i 6 e (6 p 6 k ) 6 e + m 6 e 6 e
=
pk p k
" ! !#
i [6 e (6 p+ 6 k) 6 e] + m [6 e 6 e] i [6 e (6 p 6 k ) 6 e ] + m [6 e 6 e ]
=
pk p k
11.9. CALCULATION OF THE CROSS-SECTION FOR COMPTON SCATTERING 355

! !
i [6 e (6 p+ 6 k) 6 e] + m [6 e 6 e] i [6 e (6 p 6 k ) 6 e ] + m [6 e 6 e ]
F =
pk p k
! !
i (1)3 [6 e (6 p+ 6 k) 6 e ] + m (1)2 [6 e 6 e ] i (1)3 [6 e (6 p 6 k ) 6 e ] + m (1)2 [6 e 6 e ]
=
pk p k
   
i (6 p+ 6 k) + m i (6 p 6 k ) + m
= 6 e 6 e 6 e 6 e
pk p k

therefore we have
     
i (6 p+ 6 k) + m
i (6 p 6 k ) + m
F 6e 6 e 6 e 6e
pk p k
  

i (6 p+ 6 k) + m i (6 p 6 k ) + m
= 6e 6e6e 6 e (11.133)
pk p k

substituting (11.133) in Eq. (11.130) we have

X e4
|M |2 =
,
64 (2)6 p0 p0
 
[i (6 p+ 6 k) + m] [i (6 p 6 k ) + m]
T r 6 e 6 e 6 e 6e (i 6 p + m)
pk p k
  
[i (6 p+ 6 k) + m] [i (6 p 6 k ) + m]

6e 6e6e 6e i 6 p + m (11.134)
pk p k

we shall use the following gauge


e p = e p = e p = e p = 0 (11.135)

such as for instance the Coulomb gauge in the laboratory frame, in which e0 = e0 = 0 and p = 0. Using these
facts as well as Eqs. (11.200) and (11.201) we obtain

[i 6 p + m] 6 e [i 6 p + m] = 6 e [i 6 p + m] [i 6 p + m]
   
= 6 e 6 p2 + m2 =6 e p2 + m2 = 0
[i 6 p + m] 6 e [i 6 p + m] = 0

and similarly for 6 e , 6 e and 6 e because all of them are orthogonal to p

[i 6 p + m] 6 e [i 6 p + m] = [i 6 p + m] 6 e [i 6 p + m] = 0
[i 6 p + m] 6 e [i 6 p + m] = [i 6 p + m] 6 e [i 6 p + m] = 0 (11.136)

From Eq. (11.136), many products in Eq. (11.134) vanish and we can express Eq. (11.134) in a simpler form

X     
e4 6 e 6 k 6 e 6 e 6 k 6 e 6 e 6 k 6 e 6 e 6 k 6 e 
|M |2 = 6 Tr + (i 6 p + m) + i 6 p + m
64 (2) p0 p0 pk p k pk p k

(11.137)
The trace of an odd number of Dirac matrices vanishes. As a consequence, terms linear in m vanish since they
contain seven Dirac matrices. We also see that terms independent of m contains eight Dirac matrices while terms
356 CHAPTER 11. QUANTUM ELECTRODYNAMICS

proportional to m2 contain six Dirac matrices. Hence, Eq. (11.137) splits into terms of zeroth and second order
in m. [Homework!! C8, goes from Eq. (11.137) to Eq. (11.138)]
X 
2 e4 T1 T2 T3
|M | = 6 2 + (p k) (p k ) + (p k) (p k )

64 (2) p p 0 0 (p k)

2 
T4 m t1 m2 t2 m2 t3 m2 t4
+ (11.138)
(p k )2 (p k)2 (p k) (p k ) (p k) (p k ) (p k )2

with the definitions



T1 = T r 6 e 6 k 6 e 6 p 6 e 6 k 6 e 6 p (11.139)

T1 = T r 6 e 6 k 6 e 6 p 6 e 6 k 6 e 6 p (11.140)

T3 = T r 6 e 6 k 6 e 6 p 6 e 6 k 6 e 6 p (11.141)

T4 = T r 6 e 6 k 6 e 6 p 6 e 6 k 6 e 6 p (11.142)

t1 = T r 6 e 6 k 6 e 6 e 6 k 6 e (11.143)

t2 = T r 6 e 6 k 6 e 6 e 6 k 6 e (11.144)

t3 = T r 6 e 6 k 6 e 6 e 6 k 6 e (11.145)

t4 = T r 6 e 6 k 6 e 6 e 6 k 6 e (11.146)

we shall see later how to calculate the traces of the type T r (6 a1 6 a2 6 a3 . . .) as a sum of products of scalar products
of the four-vectors a1 , a2 , a3 . . .. Traces of products of 6 or 8 gamma matrices (as is the case of ti and Tk ) contain
15 and 105 terms in the sum respectively. However, in this case many scalar products vanish reducing the number
of terms considerably. We see it by taking into account Eq. (11.135) as well as the relations (coming form the
masslessness of photons)
k k = k k = 0 (11.147)
moreover, the normalization condition
e e = e e = 1 (11.148)
also simplifies the calculation.

11.9.2 Feynman amplitude for the case of linear polarization


We make an additional simplification by assuming linear polarization. Therefore, e and e are real. Hence, we
can omit the asterisk in Eqs. (11.139)-(11.146). For instance

T1 = T r 6 e 6 k 6 e 6 p 6 e 6 k 6 e 6 p (11.149)

we shall see that by virtue of some orthonormal relations the trace (11.149) will be reduced to the trace of only
four slash four-vectors. We start by using e p = 0 with e e = 1, along with properties (11.200, 11.201) we
obtain
6 e 6 p 6 e = 6 p 6 e 6 e = 6 pe2 = 6 p
and T1 becomes 
T1 = T r 6 e 6 k 6 p 6 k 6 e 6 p
since the photons are massless we have k k = 0, and using (11.199) then

6 k 6 p 6 k = 6 k (6 p 6 k) =6 k [2 (p k) 6 k 6 p] = 6 k 6 k 6 p + 2 6 k (p k) = k2 6 p + 2 6 k (p k)
6 k 6 p 6 k = 2 6 k (p k)
11.9. CALCULATION OF THE CROSS-SECTION FOR COMPTON SCATTERING 357

from which

T1 = 2 (p k) T r 6 e 6 k 6 e 6 p

Now using the identity (11.202) T1 yields

T r {6 a 6 b 6 c 6 d} = 4 [(a b) (c d) (a c) (b d) + (a d) (b c)]

 
    
T1 = 8 (p k) e k
e p e e k p + e p k e
   
T1 = 8 (p k) 2 e k e p k p (11.150)

where we have used (11.148) with e real. We shall make the following substitutions coming from the conservation
of four-momentum and the orthogonal conditions (11.135) page 355
 
e p = e p + k k
e p = e k (11.151)

now the conservation of four-momentum yields

p + k = p + k p k = p k
2 2
p k = p k
p2 2p k + k2 = p2 2k p + k2
m2 2p k = m2 2k p

obtaining finally
p k = k p (11.152)

substituting (11.151) and (11.152) in (11.150) we find

   
T1 = 8 (p k) 2 e k e k p k
2 
T1 = 16 (p k) e k + 8 (p k) p k (11.153)

with a similar procedure we have [Homework!! C9, obtain expressions (11.154)-(11.157)]


2 2  2 
T2 = T3 = 8 e k (p k) + 16 e e p k (p k) + 8 e e k k m2
   2 
8 e e m2 k e k e + 8 e k p k
   
4 (k p)2 + 4 k k p p 4 k p p k (11.154)
 2 
T4 = 16 p k e k + 8 (p k) p k (11.155)
t1 = t4 = 0 (11.156)
    2 
t2 = t3 = 8 e e k e k e + 8 k k e e 4 k k (11.157)

substituting all these terms in Eq. (11.138) we obtain [Homework!! C10, obtain Eq. (11.158)]
" #
X 2 e4 8 (k k )2 2
|M | = 6
+ 32 e e (11.158)
64 (2) p0 p0 (k p) (k p)

358 CHAPTER 11. QUANTUM ELECTRODYNAMICS

11.9.3 Differential cross-section for Compton scattering


The differential cross-section for processes with two particles in the initial state, is given by Eq. (2.143) page 95

d = (2)4 u1 |M |2 4 p + k p k d3 p d3 k (11.159)

let us recall the expression (2.145) page 96 for the relative velocity u
q
(p1 p2 )2 m21 m22
u=
E1 E2
where particles 1 and 2 refer to initial states. For our case, since the photons are massless, the initial relative
velocity reads
|p k|
u= 0 0 (11.160)
p k
For high-energy photon-electron dispersion (i.e. X ray or gamma ray regime), it is reasonable to assume that the
electron can be taken at rest with respect to the laboratory. Thus we shall take a reference frame in which the
electron is initially at rest

p = 0, p0 = m (11.161)
0 0 0
p k = p k0 = p k (11.162)

and the velocity u in Eq. (11.160) yields


u=1 (11.163)
we denote the energy of the initial photon as , by using (11.162) such an energy yields

p0 k 0 pk
= k0 = |k| = =
p0 m
and similarly for the final photon, thus
pk
= k0 = |k| = (11.164)
m
p k
= k0 = k = (11.165)
m
taking into account that we are in the frame in which p = 0, the Diracs delta function in (11.159) becomes
  
4 p + k p k = 3 p + k p k p0 + k0 p0 k0
 
= 3 p + k k p0 + k0 p0 k0

after integrating d3 p we settle p = k k , and we are left only with the energy part of the delta function
 p 
p0 + k0 p0 k0 = p2 + m2 + m
q 
0 0 0 0
 2
p +k p k = 2
(k k ) + m + m (11.166)

from which (the final photon energy) must satisfy


q
(k k )2 + m2 = m +
p
k2 + k2 2 |k| |k | cos + m2 = + m +
q
(k0 )2 + (k0 )2 2k0 k0 cos + m2 = + m +
11.9. CALCULATION OF THE CROSS-SECTION FOR COMPTON SCATTERING 359

so that p
2 2 cos + 2 + m2 = + m
with being the angle between k and k , i.e. the angle that measures the deflection of the photon. Squaring both
sides and cancelling terms yields

2 2 cos + 2 + m2 = 2 + m2 + 2 + 2m 2 2m
2 cos = 2m 2 2m
(m + cos ) = m
m
= c () (11.167)
m + (1 cos )
by using Eq. (11.167) and the property (2.161) page 99, the energy delta function (11.166) can be rewritten as
q 
0 0 0 0
 2
p +k p k = 2
(k k ) + m + m
p 
= 2 2 cos + 2 + m2 + m
 ( c ())
p0 + k0 p0 k0 =

h i

2 2 cos + 2 + m 2 +
( c ())
=
( cos )
1 + p 0
 p0 
p0 + k0 p0 k0 = c () (11.168)
m
from the previous facts, after integrating Eq. (11.159) over d3 p we are left in that equation with the energy delta
function (11.168) and taking into account the constraint p = k k (coming from the three-momentum part of
the delta) and using Eq. (11.163), the equation (11.159) for the differential cross section reads

d = (2)4 u1 |M |2 4 p + k p k d3 p d3 k

= (2)4 |M |2 p0 + k0 p0 k0 d3 k
p0 
d = (2)4 |M |2 c () d3 k with p = k k (11.169)
m
and the differential d3 k can be written is spherical coordinates as

d3 k = 2 d d (11.170)

where d is the solid angle within which the final photon is scattered. The delta function in (11.169) permits to
eliminate the differential d in Eq. (11.170) obtaining a differential cross-section

p0 
d = (2)4 |M |2 c () 2 d d
m
p0 3
d = (2)4 |M |2 d (11.171)
m
with
p = k k ; p0 = m + and = c () (11.172)
and c () is given by Eq. (11.167).
360 CHAPTER 11. QUANTUM ELECTRODYNAMICS

As we already discussed, it is not usual to measure the spin zcomponent of the initial or final electron. Then
we sum over and average over . Hence,
  1 X  
d (p,; k, e) p , ; k , e d (p,; k, e) p , ; k , e (11.173)
2
,

It worths noticing that expression (11.167) can be written in the form

1 1 1 cos
= (11.174)
m
so that there is an increase in the wavelength. Equation (11.174) is the usual form of the Compton formula in te
scattering of Xrays by electrons.

11.9.4 Differential cross-section in the laboratory frame


All previous expressions are covariant, i.e. valid in any inertial Lorentz reference frame. Since results are usually
measured and analized in the laboratory, it is convenient to especialize these results to the Laboratory reference
frame. In such a frame we have

k k = (k,) k , = k k = |k| k cos
 
(cos 1) 1 1
k k = cos = m = m
m

where we have used Eq. (11.174). From this fact and picking up equations (11.164) and (11.165) we obtain
 
1 1
k k = (cos 1) = m


p k = m ; p k = m (11.175)

substituting equations (11.158) and (11.175) in Eq. (11.171) page 359, we obtain the laboratory frame cross-section
given by

1X   (2)4 X p0 3
d d (p,; k, e) p , ; k , e = |M |2 d
2 2 m
, ,
( " #)
(2)4 e4 8 (k k )2
 2 p0 3
= + 32 e e d
2 64 (2)6 p0 p0 (k p) (k p) m
( "  2 #)
(2)4 e4 8 m 1 1
 2 p0 3
d = + 32 e e d
2 64 (2)6 p0 p0 (m) (m ) m

And we finally obtain [Homework!! C11]


 
1X 
 e4 2 d 
2
d (p,; k, e) p , ; k , e = + 2+4 ee (11.176)
2 64 2 m2 2

This formula was derived by O. Klein and Y. Nishina in 1929 by using old-fashioned perturbation theory.
Assuming that the initial photon is not prepared in a given polarization eigenstate, we must average over the
two orthonormal vectors e. It yields
1X 1 bi k
bj

ei ej = ij k
2 e 2
11.10. TRACES OF DIRAC GAMMA MATRICES 361

so that the differential cross-section becomes [Homework!! C12]


  2 
1 XX   e4 2 d b e
d (p,; k, e) p , ; k , e = + 2 k (11.177)
4 e
64 2 m2 2
,

it occurs that the scattered photon is prererentially polarized in a direction perpendicular to the initial and also
the final photon direction. So it is polarized mainly perpendicular to the plane in which the scattering of the
photon occurs.
Further, if the final photon polarization is not measured in the experiment, we must sum (11.177) over both
polarization states, such that X
b k
ei ej = ij k b
i j
e

then we obtain [Homework!! C13]


 
1 XX 
 e4 2 d 2
d (p,; k, e) p , ; k , e = + 1 + cos (11.178)
4 32 2 m2 2
e,e ,

b and k
with being the angle between k b (scattered angle of the photon). In the non-relativistic limit << m,
Eq. (11.178) becomes

1 XX   e4 d  
d (p,; k, e) p , ; k , e = 1 + cos 2
(11.179)
4 32 2 m2
e,e ,

by integrating over the solid angle we have


Z Z Z 2
 2
  2
 16
1 + cos d = 1 + cos sin d d =
0 0 3

so that the total cross-section for << m yields

e4 8 2 e2
T = = r ; r0 2.818 1015 m (11.180)
6m2 3 0 4m
where r0 is known as the classical electron radius. Equation (11.180) is called the Thompson cross-section.
Equations (11.179) and (11.180) were originally derived using classical mechanics and electrodynamics, calculating
the reemission of light by a non-relativistic point charge in a plane-wave electromagnetic field.

11.10 Traces of Dirac gamma matrices


In calculating many observables related with Dirac fields, we encounter traces of products of Dirac gamma matrices.
We start by proving that for an even number of gamma matrices the trace is given by
X Y
T r {1 2 2N } = 4 P gpaired s (11.181)
pairings

where the sum is over all ways of pairing the indices 1 , 2 , . . . , 2N . We can see a pairing as a permutation of
the integers 1, 2, . . . , 2N into some order P1 , P2 , . . . , P2N , in which we pair as folows
 
(P1 , P2 ) , (P3 , P4 ) , , P2N1 , P2N
362 CHAPTER 11. QUANTUM ELECTRODYNAMICS

It is clear that permuting complete pairs or permuting the two s into a given pair gives the same pairing.
Therefore, the number of pairings is

(2N )!
NN = = (2N 1) (2N 3) 1 (2N 1)!! (11.182)
N ! 2N
In order to avoid summing over equivalent pairings we can demand

P1 < P2 , P3 < P4 , P2N 1 < P2N (11.183)

and
P1 < P3 < P5 < (11.184)
the requirement (11.183) order the elements within a pair, avoiding duplication of pairs. The requirement (11.184)
avoids replications due to permutations of complete sets of pairs. With this convention the factor P = 1
according whether the pairing involves an even or odd permutation of indices. Thus, the product in Eq. (11.181)
is over all N pairs, and the nth pair contributes with a factor gP2n1 P2N .
As an example, for N = 1 there is only one possible pairing

{(1 , 2 )} (11.185)

which coincides with the result of formula (11.182)

N1 = (2 1 1)!! = 1!! = 1

for N = 2 in Eq. (11.182), the number of pairings is

N2 = (2 2 1)!! = 3!! = 3

that could be listed as

{(1 , 2 ) , (3 , 4 )} ; {(1 , 3 ) , (2 , 4 )} {(1 , 4 ) , (2 , 3 )} (11.186)

note that each pairing in (11.186) has been chosen to satisfy conditions (11.183) and (11.184).
For N = 3 the number of pairings yields

N3 = (2 3 1)!! = 5!! = 5 3 1 = 15

and the list of all 15 pairings that follows the rules (11.183) and (11.184) is the following

{(1 , 2 ) , (3 , 4 ) , (5 , 6 )} ; {(1 , 2 ) , (3 , 5 ) , (4 , 6 )} ; {(1 , 2 ) , (3 , 6 ) , (4 , 5 )}


{(1 , 3 ) , (2 , 4 ) , (5 , 6 )} ; {(1 , 3 ) , (2 , 5 ) , (4 , 6 )} ; {(1 , 3 ) , (2 , 6 ) , (4 , 5 )}
{(1 , 4 ) , (2 , 3 ) , (5 , 6 )} ; {(1 , 4 ) , (2 , 5 ) , (3 , 6 )} ; {(1 , 4 ) , (2 , 6 ) , (3 , 5 )}
{(1 , 5 ) , (2 , 3 ) , (4 , 6 )} ; {(1 , 5 ) , (2 , 4 ) , (3 , 6 )} ; {(1 , 5 ) , (2 , 6 ) , (3 , 4 )}
{(1 , 6 ) , (2 , 3 ) , (4 , 5 )} ; {(1 , 6 ) , (2 , 4 ) , (3 , 5 )} ; {(1 , 6 ) , (2 , 5 ) , (3 , 4 )}

the parity of each term is dictated by the permutation from the standard order to the order induced by the pairing
for instance for the pairing
P9 {(1 , 4 ) , (2 , 6 ) , (3 , 5 )}
the associated permutation is  
1 2 3 4 5 6
1 4 2 6 3 5
11.10. TRACES OF DIRAC GAMMA MATRICES 363

which is a even permutation. And writing

, , , , , 1 , 2 , 3 , 4 , 5 , 6

we find

T r { } = 4g (11.187)
T r { } = 4 [g g g g + g g ] (11.188)
T r { } = 4 [g g g g g g + g g g g g g + g g g
g g g + g g g g g g + g g g g g g
+g g g g g g + g g g g g g + g g g ] (11.189)

On the other hand, the trace of an odd product of Dirac gamma matrices is zero

T r 1 2 2N+1 = 0

We can prove Eq. (11.181) by using mathematical induction. First of all, by using the Clifford algebra of
gamma matrices and the cyclic invariance of traces we have11

T r { } = T r {2g I44 } = 2g T r {I44 } T r { }


T r { } = 8g T r { } 2T r { } = 8g
T r { } = 4g (11.190)

as we already discussed for N = 1 (which is the case here) we have only one pairing of the form (11.185) so that
Eq. (11.190) coincides with (11.181) for N = 1.
Now that we have proved that formula (11.181) works for N = 1, we assume that it is true for N M 1.
Under such a hypothesis we find

T r {1 2 2M } = 2g1 g2 T r {3 2M } T r {2 1 3 2M }
T r {1 2 2M } = 2g1 g2 T r {3 2M } 2g1 3 T r {2 4 2N }
+T r {2 3 1 4 2M }
T r {1 2 2M } = 2g1 g2 T r {3 2M } 2g1 3 T r {2 4 2M }
+2g1 g4 T r {2 3 5 2M }

+2g1 g2M T r 2 2M 1 T r {2 2M 1 }

Now, recalling that the trace is cyclic invariant, the last term substracted in the last expression coincides with the
left hand side. Consequently

2T r {1 2 2M } = 2g1 g2 T r {3 2M } 2g1 3 T r {2 4 2M }

+2g1 g4 T r {2 3 5 2M } + 2g1 g2M T r 2 2M 1

so that

T r {1 2 2M } = g1 g2 T r {3 2M } g1 3 T r {2 4 2M }

+g1 g4 T r {2 3 5 2M } + g1 g2M T r 2 2M 1 (11.191)

now if we assume that Eq. (11.181) provides the correct trace for any product of 2N 2 gamma matrices, we
conclude that Eq. (11.191) shows that Eq. (11.181) provides the correct trace for any product of 2N gamma
matrices.
11
We should take into account that g is NOT a matrix in this context, since and are fixed indices. So g is a number and as
a matrix is g times the identity 44.
364 CHAPTER 11. QUANTUM ELECTRODYNAMICS

We can see that the trace of an odd number of gamma matrices vanishes by recalling the property

= 5 (5 )1

So for an od number 2N + 1 of gamma matrices we have


 
5 1 2 2N 2N+1 (5 )1 = 5 1 (5 )1 5 2 (5 )1 5 (5 )1 5 2N (5 )1 5 2N+1 (5 )1
h ih i h ih i
= 5 1 (5 )1 5 2 (5 )1 5 2N (5 )1 5 2N+1 (5 )1
= (1)2N +1 1 2 2N 2N+1
 
5 1 2 2N 2N+1 (5 )1 = 1 2 2N 2N+1 (11.192)

taking traces on both sides of Eq (11.192) and using the fact that traces are invariant under a similarity transfor-
mation, we obtain
n   o 
T r 5 1 2 2N 2N+1 (5 )1 = T r 1 2 2N 2N+1
 
T r 1 2 2N 2N+1 = T r 1 2 2N 2N+1

T r 1 2 2N 2N+1 = 0 (11.193)
We also encounter traces of the form
T r {5 1 2 n } (11.194)
taking into account that
5 i0 1 2 3 (11.195)
if n is odd in Eq. (11.194), then substituting Eq. (11.195) into Eq. (11.194) we obtain the trace of n + 4 gamma
matrices which is also odd. Thus the trace (11.194) vanish for odd n

T r 5 1 2 2N+1 = 0
It also vanishes for n = 0 and n = 2
T r {5 } = T r {5 } = 0
we can show that by observing that from expression (11.195) we cannot pair the indices of

T r {0 1 2 3 } or T r {0 1 2 3 } (11.196)

in such a way that the spacetime indices in each pair are equal. Thus all possible pairings of

0, 1, 2, 3 or 0, 1, 2, 3, ,

contain at least one non-diagonal gi j = 0 in the product of g s. We can see it explicitly by applying Eqs.
(11.188) and (11.189) to obtain the traces (11.196).
By contrast, for n = 4 we can pair the indices in

T r {5 } = iT r {0 1 2 3 } (11.197)

in such a way that the spacetime indices in each pair are all equal so that all g s in the product are diagonal.
However, it is possible only if the set , , , is some permutation of the set 0, 1, 2, 3. Now, since gamma matrices
with different indices anticommute, we see that the trace (11.197) must be odd under permutations of , , , .
Thus by defining
T T r {5 }
11.11. SOME PROPERTIES OF SLASH MOMENTA 365

This is a totally antisymmetric structure of four indices in which each index can take four values. Therefore, the
trace (11.197) must be proportional to the totally antisymmetric tensor . We can determine the constant of
proportionality by setting , , , 0, 1, 2, 3 recalling that
0123 = 0123 = 1
then we obtain
T r {5 } = 4i (11.198)
using the same procedure to obtain Eq. (11.181) we can obtain the trace of products of 5 with six, eight or more
gamma matrices.

11.11 Some properties of slash momenta


From the Cliffrod algebra of the Dirac matrices we have

6 a 6 b = (a ) (b ) = a b = a b (2g )
6 a 6 b = (2a g b b a )
6 a 6 b = 2a b 6 b 6 a (11.199)
in particular if the four-momenta a and b are orthogonal we obtain
6a6b=6b6a if ab=0 (11.200)
another interesting particular case appears from Eq. (11.199) by setting a = b
6 a2 = 2a2 6 a2 6 a2 = a2 (11.201)
of course, in the scalars that appear in these identities we should multiply by a 44 unit matrix.
We shall also encounter traces of products of slash momenta. We can calculate them by using the traces of
products of Dirac gamma matrices. For instance, using Eq. (11.193) we have
 
T r {6 a1 6 a2 6 a2N +1 } = T r (a1 1 ) (a2 2 ) a2N+1 2N+1

= a1 a2 a2N+1 T r 1 2 2N+1
T r {6 a1 6 a2 6 a2N +1 } = 0
Using Eq. (11.187) we have
T r {6 a 6 b} = T r {a a } = a a T r { } = 4a a g
T r {6 a 6 b} = 4a b
this result can also be obtained from Eq. (11.199) and the cyclic invariance of the trace
T r {6 a 6 b} = T r {2 (a b) 144 6 b 6 a} = 2 (a b) T r {144 } T r {6 b 6 a}
T r {6 a 6 b} = 8 (a b) T r {6 a 6 b} 2T r {6 a 6 b} = 8 (a b)
T r {6 a 6 b} = 4 (a b)
now, from Eq. (11.188) we find
T r {6 a 6 b 6 c 6 d} = a b c d T r { } = 4a b c d [g g g g + g g ]
= 4 [(a g b ) (c g d ) (a g c ) (b g d ) + (a g d ) (b g c )]
T r {6 a 6 b 6 c 6 d} = 4 [(a b) (c d) (a c) (b d) + (a d) (b c)] (11.202)
Chapter 12

Path integral approach for bosons in


quantum field theory

By applying the canonical quantization methods we were able to derive the Feynman rules for many theories.
However, in the case of scalar fields with derivative couplings or the vector fields, we saw that the interaction
Hamiltonian contains a non-covariant term that is cancelled by a non-covariant term in the propagator1 . In
quantum electrodynamics, the non-covariant term in the interacting Hamiltonian (the Coulomb energy) is not
spatially local but it is local in time. Forgetting the cancellation process we can manage to use only the covariant
terms to obtain the Feynman rules. However, this process of cancellation of non-covariant terms could be very
difficult to manage in non-abelian theories and in general relativity. Consequently, it is preferable a method that
uses the Lagrangian directly to derive the Feynman rules in a manifestly covariant form.
The path integral approach is a good alternative to work directly with a Lagrangian rather than a Hamiltonian.
Indeed their are useful in non-abelian theories even when they exhibit spontaneous symmetry breaking, as is the
case of the standard model of weak and electromagnetic interactions. Moreover, path integral methods permit to
account on the contributions to the Smatrix with an essential singularity at zero coupling constant that cannot
be discovered at any givne order in perturbation theory.
The reason to start with a canonical formalism lies in the fact that in such a formalism the Smatrix is
clearly unitary. On the other hand, the path-integral formalism provides manifestly Lorentz invariance in the
diagrammatic rules but the only way to show that the path-integral approach gives a unitary Smatrix is by
reconstructing the canonical formalism in which unitarity is apparent. Hence, in the canonical formalism unitarity
is apparent while Lorentz invariance is obscure and the opposite occurs with the path-integral formalism. The
advantage of deriving the path-integral formalism from the canonical approach (as we shall do) is that we are
certain of having the same Smatrix in both formalisms. Consequently, we can guarantee for the Smatrix both
the Lorentz-invariance and unitarity.
A second reason to start with the canonical formalism has to do with the fact that for some important
theories the simplest version of the path-integral approach (without starting with a canonical formalism) in which
propagators and interaction vertices are taken directly from the Lagragian are wrong. This is the case with the
non-linear model
1  
L = gkm k ( m )
2
in which the Feynman ruels derived directly from the Lagrangian density would lead to a non-unitary and wrong
Smatrix, that even depends on the way in which we define the scalar field. By deriving the path-integral
approach from a canonical formalism we are able to see the additional sorts of vertices required to correct the
simplest version of the Feynman path-integral formulation.
1
The covariant term in the interaction Hamiltonian equals the negative of the interaction term in the Lagrangian.

366
12.1. THE GENERAL PATH-INTEGRAL FORMULA FOR BOSONIC OPERATORS 367

12.1 The general path-integral formula for bosonic operators


Let us assume a general quantum mechanical system with Hermitian bosonic operator coordiantes Qa and
conjugate momenta Pb that satisfy canonical (bosonic) commutation (and NOT anticommutation) relations

[Qa , Pb ] = iab (12.1)


[Qa , Qb ] = [Pa , Pb ] = 0 (12.2)

we are assuming that any first class constraints are eliminated by choosing an apropriate gauge, and that the
remaining second class constraints are solved by expressing the constrained variables in terms of the unconstrained
variables Qa and Pa . Thus, we are dealing with independent canonical variables only.
The index a, is condensing continuous (position) indices as well as discrete ones (discrete Lorentz and species
m indices)

Qa Qx,m Qm (x) (12.3)


Pa Px,m Pm (x) (12.4)

in the same way the kronecker delta in Eq. (12.1) means

ab x,m ; y,n 3 (x y) mn (12.5)

the operators defined so far are in the Schrodinger picture taken at a fixed time (e.g. t = 0).
Since the set of Qa s commute, we can find a basis of simultaneous eigenstates |qi with eigenvalues qa

Qa |qi = qa |qi (12.6)

note that the lower case notation qa denotes eigenvalues instead of operators in the interaction picture. Such a
notation is not misleading since we shall not use the interaction picture by now.
The eigenvectors of such a basis can be taken as orthonormal (in the extended sense)

Y  
q |qi = qa qa q q
a

with the corresponding completeness relation


Z Y
1= dqa |qi hq|
a

in the same way we can find an orthonormal basis |pi of eigenvectors common to all Pa s:

Pa |pi = pa |pi (12.7)



Y  
p |pi = pa pa p p (12.8)
Za Y
1 = dpa |pi hp| (12.9)
a

Now from Eq. (12.1) it can be shown that Pb acts as a derivative operator i/qb on wave functions on the
qbasis. Then the wave function |pi in the basis {|qi} is given by
Y 1
hq |pi = exp (iqa pa ) (12.10)
a 2
368 CHAPTER 12. PATH INTEGRAL APPROACH FOR BOSONS IN QUANTUM FIELD THEORY

Q
where the factor 1 is obtained from the normalization condition (12.8). On the other hand, Eq. (12.10) can
a 2
also be seen as the scalar product between the two complete orthonormal sets {|qi} and {|pi}.
In the Heisenberg picture, the canonical operators acquires time-dependence

Qa (t) exp (iHt) Qa exp (iHt) (12.11)


Pa (t) exp (iHt) Pa exp (iHt) (12.12)

where H is the total Hamiltonian. The canonical operators in the Heisenberg picture have eigenstates |q; ti and
|p; ti

Qa (t) |q; ti = qa |q; ti


Pa (t) |p; ti = pa |p; ti

it is easy to obtain the basis |q; ti and |p; ti in terms of |qi and |pi

Qa (t) |q; ti = qa |q; ti exp (iHt) Qa (t) exp (iHt) exp (iHt) |q; ti = qa exp (iHt) |q; ti
Qa [exp (iHt) |q; ti] = qa [exp (iHt) |q; ti]

and comparing with (12.6) we obtain


|qi = exp (iHt) |q; ti

and similarly for momenta eigenstates then we have

|q; ti = exp (iHt) |qi ; |p; ti = exp (iHt) |pi (12.13)

it worhts emphasizing that |q; ti is the eigenstate of Qa (t) with eigenvalue qa and it is NOT the time-evolution
of the state |qi for a time t. Owing to it, its time dependence is given by a factor exp (iHt) (i.e. the inverse
of the time-evolution operator) instead that exp (iHt). Once again, these states satisfy the completeness and
orthonormality conditions

Qa (t) |q; ti = qa |q; ti ; Pa (t) |p; ti = pa |p; ti



Y  

q ; t |q; ti = qa qa q q ; p ; t |p; ti p p
Z Y Za Y
dqa |q; ti hq; t| = dpa |p; ti hp; t| = 1
a a

as well as the product


Y 1
hq; t |p; ti = exp (iqa pa ) (12.14)
a 2

If after measuring at time t, our system becomes prepared at the state |q; ti, the probability amplitude of finding
it at the state |q ; t i when we measure at time t , is given by the scalar product

A q ; t |q; ti (12.15)

we shall focus on calculating this probability amplitude.


12.1. THE GENERAL PATH-INTEGRAL FORMULA FOR BOSONIC OPERATORS 369

12.1.1 Probability amplitude for infinitesimal time-intervals


The amplitude (12.15) is easier to calculate when the time interval involved is infinitesimal. From Eq. (12.13) we
have



q ; + d |q; i = q ; exp (iH d ) |q; i (12.16)
Originally the Hamiltonian is expressed as a function H (Q, P ), but taking into account that the Hamiltonian
commutes with itself (and so with eiHt ), and that Eqs. (12.11) and (12.12) are similarity transformations, the
Hamiltonian can equally be written as the same function H (Q (t) , P (t))

H H (Q, P ) = eiHt H (Q, P ) eiHt = H (Q (t) , P (t))

The Hamiltonian can be expressed in several forms through the commutation relations (12.1) and (12.2) by moving
the operators Q and P to obtain a given order. We shall settle it in a standard form or normal order with all
Q s to the left of all P s. As a matter of example, if a term of the form Pa Qb Pc appears in the Hamiltonian, by
using Eq. (12.1), we have
Pa Qb Pc = Qb Pa Pc iab Pc
by setting this normal order the Qa (t) s in the Hamiltonian of Eq. (12.16) could be replaced by the eigenvalue qa of
Qa (t) associated with the bra. Notice that such a replacement is possible only because the operator exp [iH d ]
is linear in H for infinitesimal d . Now to deal with the P (t) operators (on the right) it is convenient to insert
an identity involving their eigenvectors |p; i
Z Y



q ; + d |q; i = dpa q ; exp [iH (Q ( ) , P ( )) d ] |p; i hp; | q; i
a
Z Y


= dpa q ; {1 iH (Q ( ) , P ( )) d } |p; i hp; | q; i
a
Z Y Z Y



= dpa q ; p; i hp; | q; i i

dpa q ; {H (Q ( ) , P ( )) d } |p; i hp; | q; i
a a

taking into account the order chosen for the Q, P operators in the Hamiltonian, we can replace H (Q ( ) , P ( ))
by H (q , p). In addition, by using the relation (12.14) we obtain
Z Y " #

dpa  X 
q ; + d |q; i = exp iH q , p d + i qa qa pa (12.17)
a
2 a

with pa integrated over the interval (, ).

12.1.2 Probability amplitude for finite time intervals


We shall calculate hq ; t |q; ti, for t < t , by dividing the time interval from t to t into N + 1 intervals of equal
time length
(t t)
k+1 k = ; 0 = t, N +1 = t (12.18)
N +1
and insert a set of N identities associated with each time in the interior of the interval (so that we do not insert
identities associated with the edges 0 = t and N +1 = t )

Z



q ; t |q; ti = dq1 dqN q ; t qN ; N i hqN ; N | qN 1 ; N 1 i hqN 1 ; N 1 | |q1 ; 1 i hq1 ; 1 |q; ti
(12.19)
370 CHAPTER 12. PATH INTEGRAL APPROACH FOR BOSONS IN QUANTUM FIELD THEORY

by defining
q0 q ; qN +1 = q
along with definitions (12.18) for N +1 and 0 . We can rewrite (12.19) as
Z


q ; t |q; ti = dq1 dqN hqN +1 ; N +1 | qN ; N i hqN ; N | qN 1 ; N 1 i hqN 1 ; N 1 | |q1 ; 1 i hq1 ; 1 |q0 ; 0 i
(12.20)

We have inner products of the type

hqk+1 ; k+1 | qk ; k i ; k = 0, 1, 2, . . . , N

if N is large enough, the time intervals become small enough to apply the identity (12.17) valid for infinitesimal
time intervals
Z Y " #
dpa X
hqk+1 ; k+1 |qk ; k i exp iH (qk+1 , pk ) + i (qk+1,a qk,a) pk,a (12.21)
a
2 a
substituting (12.21) in (12.20) we find

Z "Y
N Y
#" N Y
Y
#

dpk,a
q ; t |q; ti dqk,a
2
k=1 a k=0 a
( N
" #)
X X
exp i (qk+1,a qk,a) pk,a H (qk+1 , pk ) (12.22)
k=0 a

where we have defined


qa (k ) qk,a ; pa (k ) pk,a (12.23)
now we take the limit in which N so that d 0. In that case the aproximation (12.22) becomes
exact and we obtain
N
( )
X X
N (qk+1,a qk,a) pk,a H (qk+1 , pk ) d
k=0 a
( )
N
X X  
= qa (k ) pa (k ) H (q (k ) , p (k )) d + O (d )2
k=0 a
Z ( )
t X
N qa ( ) pa ( ) H (q ( ) , p ( )) d (12.24)
t a

additionally we can define integrals over the functions q ( ) and p ( ) as follows


Z Y Y dpb ( ) Z Y Y dpk,b
dqa ( ) lim dqk,a (12.25)
,a
2 d 0 2
,b k,a k,b

therefore, Eq. (12.22) becomes a constrained path integral


Z Y Y dpb ( )


q ; t |q; ti = qa(t)=qa dqa ( )
2
qa (t )=qa ,a ,b
( Z " #)
t X
exp i d qa ( ) pa ( ) H (q ( ) , p ( )) (12.26)
t a
12.1. THE GENERAL PATH-INTEGRAL FORMULA FOR BOSONIC OPERATORS 371

we call it a path integral because we are integrating over all paths that take q ( ) from q at = t to q at = t ,
and also over all p ( ). We shall see that by writing the matrix elements as path integrals, we can calculate them
easily by expanding in powers of the coupling constants in H.

12.1.3 Calculation of matrix elements of operators through the path-integral formalism


In addition to amplitudes of probabilities we can obtain matrix elements between states hq ; t | and |q; ti of time-
ordered products of general operators of the form O (P ( ) , Q ( )). In this case it is more convenient to define
these operators with all P s moved to the left of all Q s. Thus, by inserting such an operator in Eq. (12.17) we
obtain


M q ; + d O (P ( ) , Q ( )) |q; i
Z Y


= dpa q ; exp [iH (Q ( ) , P ( )) d ] |p; i hp; | O (P ( ) , Q ( )) |q; i
a

Owing to the order chosen for the operators Q, P in the Hamiltonian and in the operator O, we can make the
replacements

H (Q ( ) , P ( )) H q , p and O (P ( ) , Q ( )) O (p, q)

within the inner products. Then we can use the relation (12.14). The procedure is similar to the one that led to
Eq. (12.26)
Z Y " #

dpa  X 
q ; + d O (P ( ) , Q ( )) |q; i =

exp iH q , p d + i qa qa pa O (p, q) (12.27)
a
2 a

Now, to calculate the matrix element of a product of time ordered operators

OA1 (P (A1 ) , Q (A1 )) OA2 (P (A2 ) , Q (A2 )) ; tA1 > tA2 >

we should insert these operators between the apropriate states in Eq. (12.19) [or Eq. (12.20)], and then utilize
Eq. (12.27). For example, if tA1 lies between k and k+1 , we should insert OA1 (P (A1 ) , Q (A1 )) between
hqk+1 ; k+1 | and |qk ; k i. We observe that in Eq. (12.19) each sucessive sum over states is at a later time, and it is
possible because of the assumption that tA1 > tA2 > . With the same procedure as before, we find the general
path-integral formula given by


MO q ; t OA1 (P (A1 ) , Q (A1 )) OA2 (P (A2 ) , Q (A2 )) |q; ti
Z Y Y dpb ( )
= qa(t)=qa dqa ( ) OA1 (p (A1 ) , q (A1 )) OA2 (p (A2 ) , q (A2 ))

2
qa (t )=qa ,a ,b
( Z " #)
t X
exp i d qa ( ) pa ( ) H (q ( ) , p ( )) (12.28)
t a

in which we are assuming the time ordering given by

t > tA1 > tA2 > > t

nevertheless, there is nothing on the right-hand side of Eq. (12.28) that refers to the order of time-arguments.
Therefore, if we have a path integral of the form of the right-hand side of Eq. (12.28) but with tA1 , tA2 in
arbitrary order (except that they are all between t and t with t < t ), such a path integral is equal to a matrix
372 CHAPTER 12. PATH INTEGRAL APPROACH FOR BOSONS IN QUANTUM FIELD THEORY

element of the type of the left-hand side of Eq. (12.28), but with the operators ordered from left to right in
decreasing time. Consequently for tA1 , tA2 , in arbitrary order, the matrix element become


MT O q ; t T {OA1 (P (A1 ) , Q (A1 )) OA2 (P (A2 ) , Q (A2 )) } |q; ti
Z Y Y dpb ( )
= qa(t)=qa dqa ( ) OA1 (p (A1 ) , q (A1 )) OA2 (p (A2 ) , q (A2 ))
2
qa (t )=qa ,a ,b
( Z " #)
t X
exp i d qa ( ) pa ( ) H (q ( ) , p ( )) (12.29)
t a

where as customary, T represents time-ordering of the operators.


Note that the cnumber functions qa ( ) and pa ( ) in Eq. (12.29) are unconstrained variables of integration
(each one is swept independently) and in particular they are not constrained to obey the classical equations of
motion
H (q ( ) , p ( )) H (q ( ) , p ( ))
qa ( ) = 0 ; pa ( ) + =0 (12.30)
pa ( ) qa ( )
owing to it, the Hamiltonian H (q ( ) , p ( )) in Eq. (12.29) is not constant in . Notwithstanding, in some sense
path integrals respect those equations of motion. Let us assume that one of the functions in Eq. (12.29) e.g.
OA1 (P (A1 ) , Q (A1 )), is the left hand side of either of Eqs. (12.30). It can be seen that for t < tA1 < t we have
 
H (q (tA1 ) , p (tA1 ))
qa (tA1 ) exp {i I [q, p]} = i exp {i I [q, p]}
pa (tA1 ) pa (tA1 )
 
H (q (tA1 ) , p (tA1 ))
pa (tA1 ) + exp {i I [q, p]} = i exp {i I [q, p]}
qa (tA1 ) qa (tA1 )

with iI [q, p] being the argument of the exponential in Eq. (12.29)


Z t (X )
I [q, p] d qa ( ) pa ( ) H (q ( ) , p ( ))
t a

whenever tA1 does not approach t nor t , the integration over the variables qa (tA1 ) and pa (tA2 ) are not con-
strained, and if they have well behavior in convergence, the integral of these variational derivatives must vanish.
Consequently, the path integral given by (12.29) vanishes if OA1 (p, q) is taken to be the left.hand side of either of
the equations of motion (12.30).
The rule above is valid only if the integration variables qa (tAi ) , pa (tAi ) are independent of any other pair of
variables qa (tAk ) , pa (tAk ) that appears in any other function OAk different from OAi in Eq. (12.29). Then, the
rule follows only if we forbid tAi to approach tAk with k 6= i, and also different from t and t . If tAi approaches
a given tAk , the path integral will contain a non-zero term proportional to (tAi tAk ) or its derivatives. These
are the same delta function that we encounter in the operator formalism coming from time derivatives of step
functions that are implicit in the definition of time-ordered products.
In order to evaluate the path integrals (12.26) or (12.29) it suffices to know the classical Hamiltonian as a
cnumber function H (q, p). In formulating a path integral approach a natural question is how to choose the
Hamiltonian after the quantization H (Q, P ). Certainly, there are several Hamiltonians that differs only in the
order in which the Q s and P s are put after quantization. According to our developments the apropriate choice
seems to be the one with all Q s on the left and all P s on the right. However, such a choice depends on the
interpretation given to the measure Y Y
dqa ( ) dpb ( )

that appears in the path integrals (12.26) or (12.29). The prescription of all Q s on the left and all P s on the
right is correct only if the measured is interpreted according with Eqs. (12.23), (12.24) and (12.25). With other
12.2. PATH FORMALISM FOR THE SMATRIX 373

measures we would have other prescription for the ordering of the operators. Nevertheless, different prescriptions
for ordering the operators in the Hamiltonian lead to different choices of the constants that appear as coefficients in
the various terms of the Hamiltonian. Since such coefficients are considered arbitrary parameters (to be adjusted
by phenomenology) any prescription is physically equivalent to any other.
The path integral form (12.29) is not adequate for numerical calculations or to prove theorems. In that case it
is more advantageous to use a path integral method in which we calculate amplitudes in the Euclidean space with
t substituted by ix4 , so that the exponential in Eq. (12.29) becomes a negative real quantity. In the Minkowski
space, we have oscillating paths generating rapid oscillations of the integrand from one path to another. By using
the Euclidean space, the oscillating paths are exponentially supressed. Indeed we could start directly from a
path integral formulation in the Euclidean space or pass from one formulation to the other. By now, we use the
Minkowski space since our main goal is to calculate Feynman amplitudes with a perturbation approach.

12.2 Path formalism for the Smatrix


The results so far are valid in the framework of ordinary quantum mechanics. A more suitbale notation for
quantum field theory is obtained by setting the index a to run over points x in space and over a spin and species
index m so that we replace
Qa (t) Qm (x, t) ; Pa (t) Pm (x, t)
further, we also rewrite

H (q (t) , p (t)) H [q (t) , p (t)] ; O (p (t) , q (t)) H [p (t) , q (t)]

to emphasize that the Hamiltonian and operators are functionals of qm (x, t) and pm (x, t) at a given time t. From
these facts, we rewrite Eq. (;;;) as


MT O q ; t T {OA1 [P (tA1 ) , Q (tA1 )] OA2 [P (tA2 ) , Q (tA2 )] } |q; ti
Z Y Y dpm (x, )
= qm(x,t)=qm (x) dqm (x, ) OA1 [p (tA1 ) , q (tA1 )] OA2 [p (tA2 ) , q (tA2 )]
2
qm (x,t )=qm
(x) ,x,m ,x,m
( Z "Z #)
t X
3
exp i d d x qm (x, ) pm (x, t) H [q ( ) , p ( )] (12.31)
t m
Bibliography

[1] Steven Weinberg, The Quantum Theory of Fields. Vol. I: Foundations. Cambridge University Press
(1995).

[2] Steven Weinberg, The Quantum Theory of Fields. Vol. II: Modern Applications. Cambridge University
Press (1996).

[3] Steven Weinberg, The Quantum Theory of Fields. Vol. III: Supersymmetry. Cambridge University
Press (2000).

[4] Lewis H. Ryder, Quantum Field Theory (Second Ed.). Cambridge University Press (1996).

[5] Michio Kaku, Quantum Field Theory, a modern introduction. Oxford University Press (1993).

[6] John Collins, Renormalization. Cambridge Monographs on Mathematical Physics. Cambridge University
Press (1984).

[7] Claude Itzykson, Jean-Bernard Zuber, Quantum Field Theory. McGraw-Hill International Editions, Physics
Series (1980).

374

Vous aimerez peut-être aussi