Vous êtes sur la page 1sur 1464

PARTICLES AND QUANTUM FIELDS

Particles and Quantum Fields


Hagen Kleinert
Professor of Physics
Freie Universitat Berlin

Preface
This book arose from lectures I gave at the Free University Berlin over the last
40 years years. The lectures were intended to prepare graduate students for their
research work in either condensed matter many-body physics or in particle physics.
As a general broadly-based introduction, the book will eventually require a more
detailed and up-to-date account of the exciting developments in nonabelian gauge
theories that have taken place over the last two decades. These have pretxyy much
established quantum chromodynamics (QCD) as a probably true theory of strong
interactions.
Another extension into the strong-coupling limit (or critical limit) of quantum
field theory has already been published as a monograph.1 with interesting applications into the theory od very rare events 2
I thank Axel Pelster for carefully reading parts of these lecture notes and for
their useful comments and corrections. A number of printing errors were corrected
by my wife Dr. Annemarie Kleinert whom I thank for a her patience and sacrifice
of many sunny weekends. Without her persistent ecouragements the book would
never have been finished.

H. Kleinert
Berlin, October 2013

H. Kleinert and V. Schulte-Frohlinde, Critical Properties of 4 -Theories, World Scientific,


Singapore 2001, pp. 1489 (http://klnrt.de/b8).
2
H. Kleinert, Quantum Field Theory of Black-Swan Events, EPL 100, 10001 (2013)
(http://klnrt.de/399/399-TAI-PEH.pdf); a Effective Action and Field Equation for BEC from
Weak to Strong Couplings, (http://klnrt.de/403).

vii

Contents

1 Fundamentals
1.1 Classical Mechanics . . . . . . . . . . . . . . . . . . . . . .
1.2 Relativistic Mechanics in Curved Spacetime . . . . . . . . .
1.3 Quantum Mechanics . . . . . . . . . . . . . . . . . . . . . .
1.3.1
Bragg Reflections and Interference . . . . . . . . .
1.3.2
Matter Waves . . . . . . . . . . . . . . . . . . . . .
1.3.3
Schrodinger Equation . . . . . . . . . . . . . . . .
1.3.4
Particle Current Conservation . . . . . . . . . . . .
1.4 Diracs Bra-Ket Formalism . . . . . . . . . . . . . . . . . .
1.4.1
Basis Transformations . . . . . . . . . . . . . . . .
1.4.2
Bracket Notation . . . . . . . . . . . . . . . . . . .
1.4.3
Continuum Limit . . . . . . . . . . . . . . . . . . .
1.4.4
Generalized Functions . . . . . . . . . . . . . . . .
1.4.5
Schrodinger Equation in Dirac Notation . . . . . .
1.4.6
Momentum States . . . . . . . . . . . . . . . . . .
1.4.7
Incompleteness and Poissons Summation Formula
1.5 Observables . . . . . . . . . . . . . . . . . . . . . . . . . .
1.5.1
Uncertainty Relation . . . . . . . . . . . . . . . . .
1.5.2
Density Matrix and Wigner Function . . . . . . . .
1.5.3
Generalization to Many Particles . . . . . . . . . .
1.6 Time Evolution Operator . . . . . . . . . . . . . . . . . . .
1.7 Properties of the Time Evolution Operator . . . . . . . . .
1.8 Heisenberg Picture of Quantum Mechanics . . . . . . . . .
1.9 Interaction Picture and Perturbation Expansion . . . . . .
1.10 Time Evolution Amplitude . . . . . . . . . . . . . . . . . .
1.11 Fixed-Energy Amplitude . . . . . . . . . . . . . . . . . . .
1.12 Free-Particle Amplitudes . . . . . . . . . . . . . . . . . . .
1.13 Quantum Mechanics of General Lagrangian Systems . . . .
1.14 Particle on the Surface of a Sphere . . . . . . . . . . . . . .
1.15 Spinning Top . . . . . . . . . . . . . . . . . . . . . . . . . .
1.16 Classical and Quantum Statistics . . . . . . . . . . . . . . .
1.16.1 Canonical Ensemble . . . . . . . . . . . . . . . . .
1.16.2 Grand-Canonical Ensemble . . . . . . . . . . . . .
1.17 Density of States and Tracelog . . . . . . . . . . . . . . . .
Appendix 1A Simple Time Evolution Operator . . . . . . . . . .
viii

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

1
1
10
11
11
13
14
18
18
19
21
22
24
25
27
29
31
32
33
34
35
38
40
43
44
47
49
53
58
61
68
69
70
75
77

ix
Appendix 1B Convergence of the Fresnel Integral . . . . . . . . . . . . . . 77
Appendix 1C The Asymmetric Top . . . . . . . . . . . . . . . . . . . . . . 78
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
2 Field Formulation of Many-Body Quantum Physics
82
2.1 Mechanics and Quantum Mechanics for n Nonrelativistic Particles . 82
2.2 Identical Particles Bosons and Fermions . . . . . . . . . . . . . . 85
2.3 Creation and Annihilation Operators for Bosons . . . . . . . . . . . 91
2.4 Schrodinger Equation for Noninteracting Bosons in Terms of Creation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.5 Second Quantization and Symmetrized Product Representation . . . 97
2.6 Bosons with Two-Body Interactions . . . . . . . . . . . . . . . . . . 101
2.7 Quantum Field Formulation of Many-Boson Schrodinger Equations . 102
2.8 Canonical Formalism in Quantum Field Theory . . . . . . . . . . . . 104
2.9 More General Creation and Annihilation Operators . . . . . . . . . 109
2.10 Quantum Field Formulation of Many-Fermion Schrodinger Equations 111
2.11 Free Nonrelativistic Particles and Fields . . . . . . . . . . . . . . . . 113
2.12 Second-Quantized Current Conservation Law . . . . . . . . . . . . . 116
2.13 Free-Particle Propagator . . . . . . . . . . . . . . . . . . . . . . . . 117
2.14 Quantum Statistic of Free Nonrelativistic Fields . . . . . . . . . . . 120
2.14.1 Thermodynamic Quantities . . . . . . . . . . . . . . . . . . 120
2.14.2 The Degenerate Fermi Gas Near T = 0 . . . . . . . . . . . . 126
2.14.3 Degenerate Bose Gas Near T = 0 . . . . . . . . . . . . . . . 131
2.14.4 High Temperatures . . . . . . . . . . . . . . . . . . . . . . . 137
2.15 Noninteracting Bose Gas In Trap . . . . . . . . . . . . . . . . . . . . 138
2.15.1 Bose Gas in Finite Box . . . . . . . . . . . . . . . . . . . . 138
2.15.2 Harmonic and General Power Trap . . . . . . . . . . . . . . 140
2.15.3 Anharmonic Trap in Rotating Bose-Einstein Gas . . . . . . 141
2.16 Temperature Green Functions of Free Particles . . . . . . . . . . . . 143
2.17 Calculating the Matsubara Sum via Poisson Formula . . . . . . . . . 148
2.18 Nonequilibrium Quantum Statistics . . . . . . . . . . . . . . . . . . 149
2.19 Linear Response and Time-Dependent Green Functions for T 6= 0 . . 150
2.20 Spectral Representations of Green Functions for T 6= 0 . . . . . . . 153
2.21 Other Important Green Functions . . . . . . . . . . . . . . . . . . . 155
2.22 Hermitian Adjoint Operators . . . . . . . . . . . . . . . . . . . . . . 158
2.23 Harmonic Oscillator Green Functions for T 6= 0 . . . . . . . . . . . . 159
2.23.1 Creation Annihilation Operators . . . . . . . . . . . . . . . 160
2.23.2 Real Field Operators . . . . . . . . . . . . . . . . . . . . . . 162
Appendix 2A Permutation Group and Representations . . . . . . . . . . . 164
Appendix 2B Treatment of Singularities in Zeta-Function . . . . . . . . . . 169
2B.1
Finite Box . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
2B.2
Harmonic Trap . . . . . . . . . . . . . . . . . . . . . . . . . 172
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

x
3 Interacting Non-Relativistic Particles
3.1 Weakly Interacting Bose Gas . . . . . . . . . . . . .
3.2
Weakly Interacting Fermi Gas . . . . . . . . . . .
3.2.1
Electrons in a Metal . . . . . . . . . . . . .
3.3 Superconducting Electrons . . . . . . . . . . . . . .
3.3.1
Zero Temperature . . . . . . . . . . . . . .
3.4 Renormalized Theory at Strong Interactions . . . .
3.4.1
Finite Temperature . . . . . . . . . . . . . .
3.5 Crossover to Strong Couplings . . . . . . . . . . . .
3.5.1
Bogoliubov Bose Gas at Finite Temperature
3.6 Bose Gas at Strong Interactions . . . . . . . . . . .
3.7 Corrections Due to Omitted Interaction Hamiltonian
Appendix 3A Two-Loop Momentum Integrals . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

176
. 177
. 187
. 187
. 194
. 200
. 204
. 206
. 210
. 211
. 214
. 231
. 234
. 237

4 Relativistic Free Particles and Fields


239
4.1 Relativistic Particles . . . . . . . . . . . . . . . . . . . . . . . . . . 239
4.2 Differential Operators for Lorentz Transformations . . . . . . . . . . 246
4.3 Space Inversion and Time Reversal . . . . . . . . . . . . . . . . . . 255
4.4 Relativistic Free Scalar Fields . . . . . . . . . . . . . . . . . . . . . 257
4.5 Other Symmetries of Scalar Action . . . . . . . . . . . . . . . . . . . 264
4.5.1
Translations . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
4.5.2
Space Inversion . . . . . . . . . . . . . . . . . . . . . . . . . 265
4.5.3
Time Reversal . . . . . . . . . . . . . . . . . . . . . . . . . 267
4.5.4
Charge Conjugation . . . . . . . . . . . . . . . . . . . . . . 270
4.6 Electromagnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . 271
4.6.1
Action and Field Equations . . . . . . . . . . . . . . . . . . 271
4.6.2
Gauge Invariance . . . . . . . . . . . . . . . . . . . . . . . . 273
4.6.3
Lorentz Transformation Properties of Electromagnetic Fields 276
4.7 Other Symmetries of Electromagnetic Action . . . . . . . . . . . . . 278
4.7.1
Translations . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
4.7.2
Space Inversion, Time Reversal, and Charge Conjugation . . 279
4.8 Plane-Wave Solutions of Maxwells Equations . . . . . . . . . . . . 280
4.9 Massive Vector Fields . . . . . . . . . . . . . . . . . . . . . . . . . . 284
4.9.1
Action and Field Equations . . . . . . . . . . . . . . . . . . 284
4.9.2
Plane Wave Solutions for Massive Vector Fields . . . . . . . 285
4.10 Gravitational Field . . . . . . . . . . . . . . . . . . . . . . . . . . . 289
4.10.1 Action and Field Equations . . . . . . . . . . . . . . . . . . 289
4.10.2 Lorentz Transformation Properties of Gravitational Field . . 293
4.10.3 Other Symmetries of Gravitational Action . . . . . . . . . . 293
4.10.4 Translations . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
4.10.5 Space Inversion, Time Reversal, and Charge Conjugation . . 294
4.10.6 Gravitational Plane Waves . . . . . . . . . . . . . . . . . . 294
4.11 Relativistic Free Fermi Fields . . . . . . . . . . . . . . . . . . . . . . 299

xi
4.12 Spin-1/2 Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.13 Other Symmetries of Dirac Action . . . . . . . . . . . . . . . . . . .
4.13.1 Translations and Poincare group . . . . . . . . . . . . . . .
4.13.2 Space Inversion . . . . . . . . . . . . . . . . . . . . . . . . .
4.13.3 Charge Conjugation . . . . . . . . . . . . . . . . . . . . . .
4.13.4 Time Reversal . . . . . . . . . . . . . . . . . . . . . . . . .
4.13.5 Transformation Properties of Currents . . . . . . . . . . . .
4.14 Majorana Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.14.1 Plane-Wave Solutions of Dirac Equation . . . . . . . . . . .
4.15 Lorentz Transformation of Spinors . . . . . . . . . . . . . . . . . .
4.16 Precession . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.16.1 Wigner Precession . . . . . . . . . . . . . . . . . . . . . . .
4.16.2 Thomas Precession . . . . . . . . . . . . . . . . . . . . . . .
4.16.3 Spin Four-Vector and Little Group . . . . . . . . . . . . . .
4.17 Weyl Spinor Calculus . . . . . . . . . . . . . . . . . . . . . . . . . .
4.18 Higher Spin Representations . . . . . . . . . . . . . . . . . . . . . .
4.18.1 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.18.2 Extension to Lorentz Group . . . . . . . . . . . . . . . . . .
4.18.3 Finite Representation Matrices . . . . . . . . . . . . . . . .
4.19 Higher Spin Fields . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.19.1 Plane-Wave Solutions . . . . . . . . . . . . . . . . . . . . .
4.20 Vector Field as Higher Spin Field . . . . . . . . . . . . . . . . . . .
4.21 Rarita-Schwinger field for Spin 3/2 . . . . . . . . . . . . . . . . . . .
Appendix 4A Derivation of Baker-Campbell-Hausdorff Formula . . . . . .
Appendix 4B Wigner Rotations and Lobatschewksi Geometry of Rapidities
4B.1
Wigner Precession . . . . . . . . . . . . . . . . . . . . . . .
4B.2
Thomas Precession . . . . . . . . . . . . . . . . . . . . . . .
4B.3
Calculation in Four-Dimensional Representation . . . . . .
4B.4
Hyperbolic Geometry . . . . . . . . . . . . . . . . . . . . .
Appendix 4C Clebsch-Gordan Coefficients . . . . . . . . . . . . . . . . . .
Appendix 4D Spherical Harmonics . . . . . . . . . . . . . . . . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5 Classical Radiation
5.1 Electromagnetic Waves . . . . . . . . . . . . . . . . .
5.2 Wave Equation . . . . . . . . . . . . . . . . . . . . . .
5.3 Gravitational Waves . . . . . . . . . . . . . . . . . . .
5.3.1
Gravitational Field of Matter Source . . . . .
5.3.2
Quadrupole Moment . . . . . . . . . . . . . .
5.3.3
Average Radiated Energy . . . . . . . . . . .
5.3.4
Relation to Gravitational Interaction Energy
5.4 Simple Models for Sources of Gravitational Radiation
5.4.1
Vibrating Quadrupole . . . . . . . . . . . . .
5.4.2
Two Rotating Masses . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

300
310
310
310
321
324
325
326
329
341
343
343
345
346
349
351
351
354
356
363
365
367
368
369
371
372
374
374
375
377
381
383
385
385
385
390
390
393
396
396
398
398
400

xii
5.4.3
Particle Falling into Star . . . . . . . . . . . . . . . . . . .
5.4.4
Cloud of Colliding Stars . . . . . . . . . . . . . . . . . . .
5.5 Orders of Magnitude of Different Radiation Sources . . . . . . . .
5.6 Detection of Gravitational Waves . . . . . . . . . . . . . . . . . . .
5.7 Attractive Gravity versus Repulsive Electromagnetism between Like
Charges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.8 Nonlinear Gravitational Waves . . . . . . . . . . . . . . . . . . . .
Appendix 5A Nonexistence of Gravitational Waves for D = 3 and D = 2
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.

406
407
410
412

.
.
.
.

416
417
418
423

6 Relativistic Particles and Fields in External Electromagnetic Potential


466
6.1 Charged Point Particles . . . . . . . . . . . . . . . . . . . . . . . . . 466
6.1.1
Coupling to Electromagnetism . . . . . . . . . . . . . . . . 467
6.1.2
Spin Precession in Atom . . . . . . . . . . . . . . . . . . . . 469
6.1.3
Relativistic Equation of Motion for Spin Vector and Thomas
Precession . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472
6.2 Charged Particle in Schrodinger Theory . . . . . . . . . . . . . . . . 475
6.3 Charged Relativistic Fields . . . . . . . . . . . . . . . . . . . . . . . 478
6.3.1
Scalar Field . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
6.3.2
Dirac Field . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
6.4 Pauli Equation from Dirac Theory . . . . . . . . . . . . . . . . . . . 479
6.5 Relativistic Wave Equations in Coulomb Potential . . . . . . . . . . 481
6.5.1
Reminder of Schrodinger Equation with Coulomb Potential 482
6.5.2
Klein-Gordon Field in Coulomb Potential . . . . . . . . . . 484
6.5.3
Dirac Field in Coulomb Potential . . . . . . . . . . . . . . . 485
6.6 Green Function in External Electromagnetic Field . . . . . . . . . . 491
6.6.1
Scalar Field in Constant Electromagnetic Field . . . . . . . 491
6.6.2
Dirac Field in Constant Electromagnetic Field . . . . . . . . 497
6.6.3
Dirac Field in Electromagnetic Plane-Wave Field . . . . . . 499
Appendix 6A Spinor Spherical Harmonics . . . . . . . . . . . . . . . . . . 502
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503
7 Quantization of Relativistic Free Fields
7.1 Scalar Fields . . . . . . . . . . . . . . . . . . . . .
7.1.1
Real Case . . . . . . . . . . . . . . . . . .
7.1.2
Field Quantization . . . . . . . . . . . . .
7.1.3
Propagator of Free Scalar Particles . . . .
7.1.4
Complex Case . . . . . . . . . . . . . . .
7.1.5
Energy of Free Charged Scalar Particles .
7.1.6
Behavior under Discrete Symmetries . . .
7.2 Spacetime Behavior of Propagators . . . . . . . .
7.2.1
Wick Rotation . . . . . . . . . . . . . . .
7.2.2
Feynman Propagator in Minkowski Space

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

504
505
505
505
511
514
516
518
523
523
526

xiii
7.2.3
Retarded and Advanced Propagators . . . . . . . . . .
7.2.4
Comparison of Singular Functions . . . . . . . . . . .
7.3 Free Dirac Fields . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3.1
Field Quantization . . . . . . . . . . . . . . . . . . . .
7.3.2
Energy of Free Dirac Particles . . . . . . . . . . . . . .
7.3.3
Lorentz Transformation Properties of Particle States .
7.3.4
Behavior under Discrete Symmetries . . . . . . . . . .
7.4 Free Photon Field . . . . . . . . . . . . . . . . . . . . . . . . .
7.4.1
Field Quantization . . . . . . . . . . . . . . . . . . . .
7.4.2
Covariant Field Quantization . . . . . . . . . . . . . .
7.4.3
Behavior under Discrete Symmetries . . . . . . . . . .
7.5 Massive Vector Bosons . . . . . . . . . . . . . . . . . . . . . .
7.5.1
Field Quantization . . . . . . . . . . . . . . . . . . . .
7.5.2
Energy of Massive Vector Particle . . . . . . . . . . .
7.5.3
Propagator of Massive Vector Particle . . . . . . . . .
7.5.4
Behavior under Discrete Symmetry Transformations .
7.6 Spin-3/2 Fields . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.7 Gravitons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.8 Spin-Statistics Theorem . . . . . . . . . . . . . . . . . . . . . .
7.9 CPT-Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.10 Physical Consequences of Vacuum Fluctuations. Casimir Effect
7.11 Zeta Function Regularization . . . . . . . . . . . . . . . . . . .
7.12 Dimensional Regularization . . . . . . . . . . . . . . . . . . . .
7.13 Accelerated Frame . . . . . . . . . . . . . . . . . . . . . . . . .
7.14 Free Green Functions of n Fields . . . . . . . . . . . . . . . . .
7.14.1 Wicks Theorem . . . . . . . . . . . . . . . . . . . . .
7.15 Functional Form of Wicks Theorem . . . . . . . . . . . . . . .
7.15.1 Thermodynamic Version of Wicks Theorem . . . . . .
Appendix 7A Lienard-Wiechert Potential . . . . . . . . . . . . . . . .
Appendix 7B Equal-Time Commutator from Time-Ordered Product
Appendix 7C Euler-Maclaurin formula . . . . . . . . . . . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . .
8 Continuous Symmetries and Conservation Laws; Noethers
rem
8.1 Point Mechanics . . . . . . . . . . . . . . . . . . . . . . . . .
8.1.1
Continuous Symmetries and Conservation Law . . .
8.1.2
Alternative Derivation . . . . . . . . . . . . . . . . .
8.2 Displacement and Energy Conservation . . . . . . . . . . . .
8.3 Momentum and Angular Momentum . . . . . . . . . . . . . .
8.3.1
Translational Invariance in Space . . . . . . . . . . .
8.3.2
Rotational Invariance . . . . . . . . . . . . . . . . .
8.3.3
Center-of-Mass Theorem . . . . . . . . . . . . . . . .
8.3.4
Conservation Laws resulting from Lorentz Invariance

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

528
533
536
537
540
542
551
555
556
562
588
588
590
591
593
597
598
600
601
604
605
611
615
618
619
624
627
632
636
637
639
643

Theo645
. . . . 645
. . . . 645
. . . . 647
. . . . 648
. . . . 650
. . . . 650
. . . . 651
. . . . 652
. . . . 654

xiv
8.4
8.5

8.6

8.7
8.8
8.9

8.10
8.11

8.12
8.13
8.14
Notes

Generating the Symmetries . . . . . . . . . . . . . . . . . . . . . .


Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.5.1
Continuous Symmetry and Conserved Currents . . . . . .
8.5.2
Alternative Derivation . . . . . . . . . . . . . . . . . . . .
8.5.3
Local Symmetries . . . . . . . . . . . . . . . . . . . . . .
Canonical Energy Momentum Tensor . . . . . . . . . . . . . . . .
8.6.1
Electromagnetism . . . . . . . . . . . . . . . . . . . . . .
8.6.2
Dirac Field . . . . . . . . . . . . . . . . . . . . . . . . . .
Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . .
Four-Dimensional Angular Momentum . . . . . . . . . . . . . . . .
Spin Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
8.9.1
Electromagnetic Fields . . . . . . . . . . . . . . . . . . . .
8.9.2
Dirac Field . . . . . . . . . . . . . . . . . . . . . . . . . .
Symmetric Energy-Momentum Tensor . . . . . . . . . . . . . . . .
8.10.1 Gravitational Field . . . . . . . . . . . . . . . . . . . . . .
Internal Symmetries . . . . . . . . . . . . . . . . . . . . . . . . . .
8.11.1 U(1)-Symmetry and Charge Conservation . . . . . . . . .
8.11.2 SU(N)-Symmetry . . . . . . . . . . . . . . . . . . . . . .
8.11.3 Broken Internal Symmetries . . . . . . . . . . . . . . . . .
Generating the Symmetry Transformations on Quantum Fields . .
Energy Momentum Tensor of a Relativistic Massive Point Particle
Energy Momentum Tensor of Massive Charged Particles in an Electromagnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . .
and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

9 Scattering and Decay of Particles


9.1 Quantum-Mechanical Description . . . . . .
9.1.1
Schrodinger Picture . . . . . . . . .
9.1.2
Heisenberg Picture . . . . . . . . . .
9.1.3
Interaction Picture . . . . . . . . . .
9.1.4
Neumann-Liouville expansion . . . .
9.1.5
Mller Operators . . . . . . . . . . .
9.1.6
Lippmann-Schwinger equation . . . .
9.1.7
Discrete States . . . . . . . . . . . .
9.1.8
Gell-Mann - Low Formulas . . . . . .
9.2 Scattering in External Potential . . . . . . .
9.2.1
The T -Matrix . . . . . . . . . . . . .
9.2.2
Asymptotic Behavior . . . . . . . . .
9.2.3
Partial Waves . . . . . . . . . . . . .
9.2.4
Off Shell T -Matrix . . . . . . . . . .
9.2.5
Cross Section . . . . . . . . . . . . .
9.2.6
Partial Wave Decomposition of Cross
9.2.7
Dirac -Function Potential . . . . . .
9.2.8
Spherical Square-Well Potential . . .

. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
. . . . .
Section
. . . . .
. . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

656
658
658
659
660
662
663
664
666
667
669
669
672
674
676
677
677
678
678
678
680

. 681
. 684
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

686
686
686
687
688
688
690
693
695
696
701
701
705
707
713
716
720
720
723

xv
9.3

Two-Particle Scattering . . . . . . . . . . . . . . . . . . . . . . . .
9.3.1
Center-of-Mass Scattering Cross Section . . . . . . . . . .
9.3.2
Laboratory Scattering Cross Section . . . . . . . . . . . .
9.4 Decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.5 Optical Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.6 Initial- and Final-State Interactions . . . . . . . . . . . . . . . . .
9.7 Tests of Time-Reversal Violations . . . . . . . . . . . . . . . . . .
9.7.1
Strong and Electromagnetic Interactions . . . . . . . . . .
9.7.2
Selection Rules in Weak Interactions . . . . . . . . . . . .
9.7.3
Phase of Weak Amplitudes from Time-Reversal Invariance
Appendix 9A Green Function in Arbitrary Dimensions . . . . . . . . . .
Appendix 9B Partial Waves in Arbitrary Dimensions . . . . . . . . . . .
Appendix 9C Spherical Square-Well Potential in D Dimensions . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10 Quantum Field Theoretic Perturbation Theory
10.1 The Interacting n-Point Function . . . . . . .
10.2 Perturbation Expansion for Green Functions .
10.3 Feynman Rules for 4 -theory . . . . . . . . . .
10.3.1 The Vacuum Graphs . . . . . . . . . .
10.4 The Two-Point Function . . . . . . . . . . . .
10.5 The Four-Point Function . . . . . . . . . . . .
10.6 Connected Green Function . . . . . . . . . . .
10.6.1 One-Particle Irreducible Graphs . . . .
10.6.2 Momentum Space Version of Diagrams
10.7 Green Functions and Scattering Amplitudes . .
10.8 Wick Rules for Scattering Amplitudes . . . . .
10.9 Thermal Perturbation Theory . . . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.

725
727
728
732
732
733
734
735
736
737
738
740
745
747

748
. 748
. 751
. 752
. 754
. 757
. 759
. 761
. 765
. 767
. 769
. 776
. 777
. 780

11 Extracting Finite Results from Perturbation Series. Regularization,


Renormalization
781
11.1 Vacuum Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . 781
11.2 The Two- and Four-Point Functions . . . . . . . . . . . . . . . . . . 784
11.3 Divergences, Cutoff, and Counterterms . . . . . . . . . . . . . . . . 786
11.4 Bare Theory and Multiplicative Renormalization . . . . . . . . . . . 793
11.5 Dimensional Regularization of Integrals . . . . . . . . . . . . . . . . 797
11.6 Renormalization of Amplitudes . . . . . . . . . . . . . . . . . . . . . 811
11.7 Additive Renormalization of Vacuum Energy . . . . . . . . . . . . . 814
11.8 Generalization to O(N)-Symmetric Model . . . . . . . . . . . . . . . 814
11.9 Finite S-Matrix Elements . . . . . . . . . . . . . . . . . . . . . . . . 820
Appendix 11AAlternative Proof of Veltman Integral Rule . . . . . . . . . . 822
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 824

xvi
12 Quantum Electrodynamics
12.1 Gauge Invariant Interacting Theory . . . . . . . . . . . . . . . .
12.1.1 Reminder of Classical Electrodynamics of Point Particles
12.1.2 Electrodynamics and Quantum Mechanics . . . . . . . .
12.1.3 Principle of Nonholonomic Gauge Invariance . . . . . . .
12.1.4 Electrodynamics and Relativistic Quantum Mechanics .
12.2 Noethers Theorem and Gauge Fields . . . . . . . . . . . . . . .
12.3 Quantization . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
12.4 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . .
12.5 Ward-Takahashi identity . . . . . . . . . . . . . . . . . . . . . .
12.6 Magnetic Moment of Electron . . . . . . . . . . . . . . . . . . .
12.7 Decay of Atomic State . . . . . . . . . . . . . . . . . . . . . . .
12.8 Rutherford Scattering . . . . . . . . . . . . . . . . . . . . . . . .
12.8.1 Classical Cross Section . . . . . . . . . . . . . . . . . . .
12.8.2 Quantum-Mechanical Born Approximation . . . . . . . .
12.8.3 Relativistic Born Approximation . . . . . . . . . . . . .
12.9 Compton Scattering . . . . . . . . . . . . . . . . . . . . . . . . .
12.9.1 Classical Result . . . . . . . . . . . . . . . . . . . . . . .
12.9.2 Klein-Nishina Formula . . . . . . . . . . . . . . . . . . .
12.10 Electron-Positron Annihilation . . . . . . . . . . . . . . . . . . .
12.11 Positronium Decay . . . . . . . . . . . . . . . . . . . . . . . . . .
12.12 Bremsstrahlung . . . . . . . . . . . . . . . . . . . . . . . . . . .
12.12.1 Classical Bremsstrahlung . . . . . . . . . . . . . . . . .
12.12.2 Bremsstrahlung in Mott Scattering . . . . . . . . . . . .
12.13 Electron-Electron Scattering . . . . . . . . . . . . . . . . . . . .
12.14 Electron-Positron Scattering . . . . . . . . . . . . . . . . . . . .
12.15 Anomalous Magnetic Moment of Electron and Muon . . . . . .
12.15.1 Form Factors . . . . . . . . . . . . . . . . . . . . . . . .
12.15.2 Charge Radius . . . . . . . . . . . . . . . . . . . . . . .
12.15.3 Anomalous Magnetic Moment . . . . . . . . . . . . . . .
12.16 Vacuum Polarization . . . . . . . . . . . . . . . . . . . . . . . .
12.17 Dimensional Regularization . . . . . . . . . . . . . . . . . . . . .
12.18 Two-Dimensional QED . . . . . . . . . . . . . . . . . . . . . . .
12.19 Self-Energy of Electron . . . . . . . . . . . . . . . . . . . . . . .
12.20 Ward-Takahashi identity . . . . . . . . . . . . . . . . . . . . . .
12.21 Lamb Shift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
12.21.1 Rough Estimate of Effect of Vacuum Fluctuations . . . .
12.21.2 Relativistic Estimate . . . . . . . . . . . . . . . . . . . .
12.21.3 Effect of Wave Functions . . . . . . . . . . . . . . . . .
12.21.4 Effect of the Anomalous Magnetic Moment . . . . . . .
Appendix 12ACalculation of Dirac Trace for Klein-Nishina formula . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

825
. 825
. 826
. 828
. 830
. 831
. 832
. 834
. 837
. 842
. 843
. 847
. 851
. 851
. 853
. 853
. 857
. 857
. 859
. 863
. 869
. 870
. 870
. 873
. 876
. 878
. 881
. 886
. 887
. 888
. 892
. 896
. 897
. 898
. 901
. 903
. 903
. 906
. 906
. 915
. 917
. 920

xvii
13 Functional Integral Representation of Quantum Field Theory
13.1 Functional Fourier Transform . . . . . . . . . . . . . . . . . . . .
13.2 Gaussian Functional Integral . . . . . . . . . . . . . . . . . . . .
13.3 Functional Formulation for Free Quantum Fields . . . . . . . . .
13.4 Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.5 Euclidean Quantum Field Theory . . . . . . . . . . . . . . . . .
13.6 Functional Integral Representation for Fermions . . . . . . . . .
13.7 Relation Between Z[j] and Partition Function . . . . . . . . . .
13.8 Bosons and Fermions in a Single State . . . . . . . . . . . . . . .
13.9 Free Energy for Free Scalar Fields . . . . . . . . . . . . . . . . .
13.10 Interacting Nonrelativistic Fields . . . . . . . . . . . . . . . . . .
13.10.1 Functional Formulation . . . . . . . . . . . . . . . . . .
13.10.2 Grand-Canonical Ensembles at Zero Temperature . . . .
13.11 Interacting Relativistic Fields . . . . . . . . . . . . . . . . . . . .
13.12 Plasma Oscillations . . . . . . . . . . . . . . . . . . . . . . . . .
13.12.1 General Formalism . . . . . . . . . . . . . . . . . . . . .
13.12.2 Physical Consequences . . . . . . . . . . . . . . . . . . .
13.13 Gauge Fields and Gauge Fixing . . . . . . . . . . . . . . . . . .
13.14 Nontrivial Gauge and Fadeev-Popov Ghosts . . . . . . . . . . . .
13.15 Functional Formulation of Quantum Electrodynamics . . . . . .
13.15.1 Decay Rate of Dirac Vacuum in Constant Electric Field
13.15.2 Constant Electric and Magnetic Background Fields . . .
13.15.3 Decay Rate in Constant Electromagnetic Field . . . . .
13.15.4 Effective Action in Purely Magnetic Field . . . . . . . .
13.15.5 Heisenberg-Euler Lagrangian . . . . . . . . . . . . . . .
13.15.6 Alternative Derivation for Constant Magnetic Field . . .
Appendix 13APropagator of the Bilocal Pair Field . . . . . . . . . . . .
Appendix 13BFluctuations around the Composite Field . . . . . . . . .
Appendix 13CTwo-Loop Heisenberg-Euler Effective Action . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

923
. 923
. 924
. 926
. 929
. 932
. 933
. 937
. 942
. 944
. 945
. 947
. 948
. 954
. 956
. 956
. 960
. 963
. 970
. 973
. 974
. 978
. 981
. 982
. 983
. 986
. 990
. 992
. 994
. 995

14 Formal Properties of Perturbation Theory


1004
14.1 Connectedness Structure of Feynman Diagrams . . . . . . . . . . . . 1004
14.2 Functional Differential Equations . . . . . . . . . . . . . . . . . . . . 1005
14.3 Decomposition of Green Functions into Connected Green Functions 1007
14.4 Functional Differential Equation for W [j[ . . . . . . . . . . . . . . . 1009
14.5 Iterative Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1009
14.6 Vertex Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1011
14.7 The Generating Functional for Vertex Functions . . . . . . . . . . . 1011
14.8 Functional Differential Equation for [] . . . . . . . . . . . . . . . 1017
14.9 Effective Action as a Basis for Variational Calculations . . . . . . . 1020
14.10 Effective Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1020
14.11 Higher Effective Actions . . . . . . . . . . . . . . . . . . . . . . . . . 1021
14.12 High Orders in Simple Model . . . . . . . . . . . . . . . . . . . . . . 1026

xviii
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1029
15 Path Integral Calculation of Effective Action.
15.1 General Formalism . . . . . . . . . . . . . . .
15.2 Quadratic Fluctuations . . . . . . . . . . . .
15.3 Massless Theory . . . . . . . . . . . . . . . .
15.4 Effective Action to Second Order in h
. . . .
15.5 Effective Action to all Orders in h
. . . . . .
Notes and References . . . . . . . . . . . . . . . . .

Loop
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

Expansion
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

16 Systematic Graphical Construction of Feynman Diagrams . . .


16.1 Generalized Scalar 4 -Theory . . . . . . . . . . . . . . . . . . . .
16.2 Basic Graphical Operations . . . . . . . . . . . . . . . . . . . . .
16.2.1 Cutting Lines . . . . . . . . . . . . . . . . . . . . . . . .
16.2.2 Removing Lines . . . . . . . . . . . . . . . . . . . . . . .
16.3 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . .
16.4 Functional Differential Equation for W = ln Z . . . . . . . . . .
16.5 Recursion Relation and Graphical Solution . . . . . . . . . . . .
16.6 Scalar 2 A-Theory . . . . . . . . . . . . . . . . . . . . . . . . . .
16.7 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . .
16.8 Functional Differential Equation for W = ln Z . . . . . . . . . .
16.9 Recursion Relation and Graphical Solution . . . . . . . . . . . .
16.10 Computer Generation of Diagrams . . . . . . . . . . . . . . . . .
16.11 Matrix Representation of Diagrams . . . . . . . . . . . . . . . .
16.12 Practical Generation . . . . . . . . . . . . . . . . . . . . . . . . .
16.12.1 Connected Vacuum Diagrams . . . . . . . . . . . . . . .
16.12.2 Two- and Four-Point Functions from Cutting Lines . . .
16.12.3 Two- and Four-Point Function from Removing Lines . .
17 Spontaneous Symmetry Breakdow
17.1 Scalar O(N)-Symmetric 4 -Theory . . . .
17.2 Nambu-Goldstone Particles . . . . . . . .
17.2.1 The Mechanism . . . . . . . . . .
17.2.2 General Considerations . . . . .
17.2.3 Experimental Consequences . . .
17.3 Domain Walls in O(1)-Symmetric Theory
17.4 Vortex Lines in O(2)-Symmetric Theory .
Notes and References . . . . . . . . . . . . . . .
18 Scalar Quantum Electrodynamics
18.1 Action and Generating Functional
18.2 Meissner-Higgs Effect . . . . . . .
18.3 Spatially Varying Ground States .
18.4 Two Natural Length Scales . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

1030
. 1030
. 1033
. 1038
. 1042
. 1046
. 1048

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

1050
. 1051
. 1053
. 1053
. 1056
. 1057
. 1058
. 1060
. 1062
. 1064
. 1064
. 1065
. 1067
. 1067
. 1069
. 1069
. 1072
. 1073

.
.
.
.
.
.
.
.

1084
. 1084
. 1090
. 1090
. 1092
. 1094
. 1095
. 1100
. 1106

.
.
.
.

1107
. 1107
. 1110
. 1116
. 1117

xix
18.5
18.6
18.7
18.8
18.9
Notes

Planar Domain Wall . . . . . . . . . . . . . . . . . . . .


Surface Energy . . . . . . . . . . . . . . . . . . . . . . .
Single Vortex Line and Critical Field Hc1 . . . . . . . .
Critical Field Hc2 where Superconductivity is Destroyed
Order of the Superconducting Phase Transition . . . . .
and References . . . . . . . . . . . . . . . . . . . . . . .

19 Exactly Solvable O(N)-Symmetric 4 -Theory for


19.1 Introduction of a Collective Field . . . . . . . . .
19.2 The Limit of Large N . . . . . . . . . . . . . . .
19.3 Variational Equations . . . . . . . . . . . . . . .
19.3.1 Non-trivial Ground States . . . . . . . .
19.4 Special Features of Two Dimensions . . . . . . .
19.5 Experimental Consequences . . . . . . . . . . . .
19.6 Correlation Functions for Large N . . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . .

Large
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .
. . . .

.
.
.
.
.
.

.
.
.
.
.
.

N
. .
. .
. .
. .
. .
. .
. .
. .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

1119
1125
1126
1132
1135
1136

.
.
.
.
.
.
.
.

1137
. 1137
. 1140
. 1147
. 1148
. 1152
. 1154
. 1157
. 1160

20 Non-Linear -Model
1161
20.1 Definition of Classical Heisenberg Model . . . . . . . . . . . . . . . . 1161
20.2 Spherical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1164
20.3 Free Energy and Gap Equation in D > 2 Dimensions . . . . . . . . . 1165
20.3.1 High-Temperature Phase . . . . . . . . . . . . . . . . . . . 1167
20.3.2 Low-Temperature Phase . . . . . . . . . . . . . . . . . . . . 1168
20.4 Approaching the Critical Point . . . . . . . . . . . . . . . . . . . . . 1170
20.5 Physical Property of Bare Temperature . . . . . . . . . . . . . . . . 1171
20.6 Spherical Model on Lattice . . . . . . . . . . . . . . . . . . . . . . . 1173
20.7 Background Field Treatment of Cold Phase . . . . . . . . . . . . . . 1177
20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in
D Spacetime Dimensions . . . . . . . . . . . . . . . . . . . . . . . . 1179
20.8.1 Two-Dimensional Model . . . . . . . . . . . . . . . . . . . . 1179
20.8.2 Four-Dimensional Model . . . . . . . . . . . . . . . . . . . . 1184
20.8.3 Any Dimension . . . . . . . . . . . . . . . . . . . . . . . . . 1185
20.9 Criteria for Onset of Fluctuations in Ginzburg-Landau Theories . . . 1190
20.9.1 Ginzburgs Criterion . . . . . . . . . . . . . . . . . . . . . . 1191
20.9.2 Kleinerts Criterion . . . . . . . . . . . . . . . . . . . . . . . 1192
20.9.3 Experimental Consequences . . . . . . . . . . . . . . . . . . 1194
21 The
21.1
21.2
21.3
21.4
21.5
21.6

Renormalization Group
1197
Example for Redundancy in Parametrization of Renormalized Theory1198
Renormalization Scheme . . . . . . . . . . . . . . . . . . . . . . . . 1200
The Renormalization Group Equation . . . . . . . . . . . . . . . . . 1202
Calculation of Coefficient Functions from Counter Terms . . . . . . 1203
Solution of Renormalization Group Equations for Vertex Functions . 1207
Renormalization Group for Effective Action and Potential . . . . . . 1211

xx
21.7 Approach to Scaling . . . . . . . . . . . . . . . . . . . .
21.8 Explicit Solution of the RGE close to D = 4 Dimensions
21.9 Further Critical Relations . . . . . . . . . . . . . . . . .
21.9.1 Scaling Relations Above Tc . . . . . . . . . . .
21.9.2 Scaling Relations Below Tc . . . . . . . . . . .
21.10 Comparison of Scaling Relations with Experiment . . .
21.11 Higher-Order Expansion . . . . . . . . . . . . . . . . .
21.12 Mean-Field Results for Critical Indices . . . . . . . . .
21.13 Effective Potential in the Critical Regime to Order . .
21.14 O(N)-Symmetric Theory . . . . . . . . . . . . . . . . .
21.15 Direct Scaling Form in the Limit of Large N . . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

1213
1215
1219
1219
1221
1223
1225
1226
1229
1234
1236

22 Critical Properties of Non-linear -Model


1238
22.1 Introductory Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . 1238
22.2 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 1240
22.3 Symmetry Properties of Renormalized Effective Action . . . . . . . 1245
22.4 Critical Behavior in 2 + Dimensions . . . . . . . . . . . . . . . . . 1248
22.5 Critical Exponents . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1249
22.6 Two- and Three-Loop Results . . . . . . . . . . . . . . . . . . . . . 1256
22.7 Variational Resummation of -Expansions . . . . . . . . . . . . . . . 1259
22.7.1 Strong-Coupling Theory . . . . . . . . . . . . . . . . . . . . 1261
22.7.2 Interpolation . . . . . . . . . . . . . . . . . . . . . . . . . . 1263
22.8 Relation of -Model to Quantum Mechanics of a Point Particle on
a Sphere . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1268
22.9 Generalization of the Model . . . . . . . . . . . . . . . . . . . . . . . 1274
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1277
23 Exactly Solvable O(N)-Symmetric Four-Fermion Theory in
Dimensions
23.1 Scalar Self-Interaction . . . . . . . . . . . . . . . . . . . . . . .
23.2 Spontaneous Symmetry Breakdown . . . . . . . . . . . . . . .
23.3 Dimensionally Transmuted Coupling Constant . . . . . . . . .
23.4 Scattering Amplitude for Fermions . . . . . . . . . . . . . . . .
23.5 Nonzero Bare Fermion Mass . . . . . . . . . . . . . . . . . . .
23.6 Pairing Model and Dynamically Generated Goldstone Bosons .
23.7 Spontaneously Broken Symmetry . . . . . . . . . . . . . . . . .
23.8 Relation between Pairing and Gross-Neveu Model . . . . . . .
23.9 Comparison with O(N)-Symmetric 4 -Theory . . . . . . . . . .
23.10 Two Phase Transitions in Chiral Gross-Neveu Model . . . . . .
23.11 Finite-Temperature Properties . . . . . . . . . . . . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . .

2+
1280
. . . 1280
. . . 1285
. . . 1286
. . . 1287
. . . 1294
. . . 1295
. . . 1302
. . . 1306
. . . 1307
. . . 1312
. . . 1315
. . . 1326

xxi
24 Internal Symmetries of Strong Interactions
24.1 Classification of Elementary Particles . . . . . . . . . . . . . .
24.2 Isospin in Nuclear Physics . . . . . . . . . . . . . . . . . . . .
24.3 Isospin in Pion Physics . . . . . . . . . . . . . . . . . . . . . .
24.4 SU(3)-Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . .
24.5 New Quarks . . . . . . . . . . . . . . . . . . . . . . . . . . . .
24.6 Tensor Representations and Young Tableaux . . . . . . . . . .
24.7 Effective Interactions among Hadrons . . . . . . . . . . . . . .
24.7.1 The Pion-Nucleon Interaction . . . . . . . . . . . . . .
24.7.2 The Decay (1232) N . . . . . . . . . . . . . . .
24.7.3 Vector Meson Decay (770) . . . . . . . . . . .
24.7.4 Vector Meson Decays (783) and (783)
24.7.5 Vector Meson Decays K (892) K . . . . . . . . . .
24.7.6 Axial Vector Meson Decay a1 (1270) . . . . . . .
24.7.7 Coupling of (770)-Meson to Nucleons . . . . . . . . .
Appendix 24AUseful SU(3) Formulas . . . . . . . . . . . . . . . . . .
Appendix 24BDecay rate for a1 . . . . . . . . . . . . . . . . . .
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

1328
. 1328
. 1331
. 1336
. 1339
. 1358
. 1359
. 1364
. 1364
. 1367
. 1372
. 1373
. 1373
. 1374
. 1375
. 1376
. 1377
. 1379

25 Symmetries Linking Internal and Spacetime Structure


25.1 Approximate SU(4)-Symmetry of Nuclear Forces . . . .
25.2 Approximate SU(6)-Symmetry in Strong Interactions .
25.3 From SU(6) to Current Algebra . . . . . . . . . . . . .
25.4 Supersymmetry . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

1380
. 1380
. 1388
. 1396
. 1401

26 Weak Interactions
26.0.1 Fermi Theory . . . . .
26.0.2 Lepton Conservation .
26.0.3 Cabibbo Angle . . . .
26.0.4 Cabibbo Mass Matrix
26.0.5 Heavy Vector Bosons .
26.1 Standard Model of Electroweak
26.2 Quantum Oscillations . . . . .
26.2.1 Kaon Oscillations . . .
26.2.2 General Flavor Mixing
Notes and References . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

1404
. 1404
. 1408
. 1408
. 1410
. 1410
. 1412
. 1417
. 1417
. 1420
. 1421

.
.
.
.
.
.

1422
. 1422
. 1424
. 1425
. 1432
. 1438
. 1441

. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
Interactions
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

27 Nonabelian Gauge Theory of Strong Interactions


27.0.3 Local Color Symmetry . . . . . . . . . . .
27.0.4 Gluon Action . . . . . . . . . . . . . . . .
27.1 Quantization in the Coulomb gauge . . . . . . . .
27.2 Direct Functional Quantization of Gauge Fields . .
27.3 Equivalence of Landau and Coulomb gauges . . .
27.3.1 Approximate Chiral Symmetry . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

xxii
Notes and References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1442
Index

1443

List of Figures
1.1
1.2
1.3
1.4

Probability distribution of particle behind a double slit . . . . .


P
2in
Relevant function N
in Poissons summation formula
n=N e
Illustration of time-ordering procedure . . . . . . . . . . . . . . .
Triangular closed contour for Cauchy integral . . . . . . . . . . .

.
.
.
.

12
30
37
78

2.1
2.2
2.3
2.4
2.5

Average Bose occupation number . . . . . . . . . . . . . . . . . . .


Average Fermi occupation number . . . . . . . . . . . . . . . . . . .
Temperature behavior of specific heat of a free Fermi gas . . . . . .
Temperature behavior of the chemical potential of a free Bose gas .
Temperature behavior of fraction of degenerate bosons in the zeromomentum state of a free Bose gas . . . . . . . . . . . . . . . . . . .
Temperature behavior of the specific heat of a free boson gas. For
comparison we show the specific heat of the strongly interacting Bose
liquid 4 He, scaled down by a factor 2 to match the Dulong-Petit limit
of the free Bose gas . . . . . . . . . . . . . . . . . . . . . . . . . . .
Rotating trap potential for 2 > 0 and 2 < 0 . . . . . . . . . . . .
Contour C in the complex z-plane . . . . . . . . . . . . . . . . . . .
Finite-size corrections to critical temperature for N = 300 to infinity

122
123
131
132

2.6

2.7
2.8
2.9

.
.
.
.

Plot of the quasiparticle energies as a function of the momenta in


an interacting Bose gas . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Common volume of two spheres at a distance q in momentum space
3.3 Energy density in units Ry of electron gas in uniform background of
positive charge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4 Historical evolution of critical temperatures of onset of superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5 Approximate energy of a free electron near the Fermi surface in a
grand canonical ensemble . . . . . . . . . . . . . . . . . . . . . . . .
3.6 Gap in the energy spectrum caused by attraction of pairs of electrons
with opposite spin and momenta . . . . . . . . . . . . . . . . . . .
3.7 Spectrum of quasiparticles and holes . . . . . . . . . . . . . . . . .
3.8 Gap in energy spectrum . . . . . . . . . . . . . . . . . . . . . . . . .
3.9 Solution of gap equation for weak attraction between electrons . . .
3.10 Plot of the gap function and of the chemical potential against the
inverse s-wave scattering length . . . . . . . . . . . . . . . . . . . .
3.11 Temperature dependence of the normal fraction u / in Bose gas . .

133

134
142
147
172

3.1

xxiii

186
192
194
196
198
199
200
200
210
211
214

xxiv
a as a function of the reduced s-wave scattering
3.12 Reduced gap s /
length a
s = 8as /a = 8as 1/3 . . . . . . . . . . . . . . . . . . . .
3.13 Reduced energy per particle w1e = W1 /Na as a function of the
reduced s-wave scattering length, compared with Bogoliubovs weakcoupling result . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.14 Temperature dependence of the normal particle density . . . . . .
3.15 Diagrams picturing the Wick contractions . . . . . . . . . . . . .
4.1
4.2
4.3
4.4

. 222

. 223
. 230
. 232

Six leptons and quarks . . . . . . . . . . . . . . . . . . . . . . . . . 313


Asymmetry observed in the distribution of electrons from the decay of polarized 60
27 Co . . . . . . . . . . . . . . . . . . . . . . . . . 314
+ and L
upon the states |s, mi 354
of raising and lowering operators L
Triangle formed by rapidities in a hyperbolic space. The sum of
angles is smaller than 1800 . The angular defect yields the angle of
Thomas precession . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376

5.1

Two equal masses M oscillating at the ends of a spring as a source


of gravitational radiation . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Two spherical masses in circular orbits about their center of mass. .
5.3 Gravitational Amplitudes arriving at the earth from possible sources
5.4 Two pulsars orbiting around each other . . . . . . . . . . . . . . . .
5.5 Shift of time of periastron passage of PSR 1913+16 . . . . . . . . .
5.6 Two masses in a Keplerian orbit around the common center-of-mass
5.7 Spectrum of the gravitational radiation emitted by a particle of mass
m falling radially into a black hole of mass M. . . . . . . . . . . . .
5.8 Particle falling radially toward a mass. . . . . . . . . . . . . . . . . .
5.9 Distortions of a circular array of mass points by the passage of a
gravitational quandrupole wave . . . . . . . . . . . . . . . . . . . .
5.10 Field lines of tidal forces of a gravitational wave . . . . . . . . . . .

399
401
403
404
405
405
408
408
414
415

6.1

Hydrogen spectrum according to Diracs theory. The fine-structure


splitting of the 2P -levels is about 10 times as big as the hyperfine
splitting and Lamb shift . . . . . . . . . . . . . . . . . . . . . . . . 488

7.1

Pole positions in the complex p0 -plane in the integral representations


of Feynman propagators . . . . . . . . . . . . . . . . . . . . . . . . .
Wick rotation of the contour of integration in the complex p0 -plane
Integration contours in the complex p0 -plane of the Fourier integral
for the various propagators . . . . . . . . . . . . . . . . . . . . . . .
Different coupling schemes for two-particle states of total angular
momentum j and helicity m. . . . . . . . . . . . . . . . . . . . . .
Geometry of the silver plates for the calculation of the Casimir effect

7.2
7.3
7.4
7.5
9.1

524
524
536
547
605

Behavior of wave function for different positions of a bound state


near the continuum . . . . . . . . . . . . . . . . . . . . . . . . . . . 721

xxv
9.2
9.3

Behavior of binding energy and scattering length in attractive


square-well potential . . . . . . . . . . . . . . . . . . . . . . . . . . 724
Geometry of particle beams in a collider . . . . . . . . . . . . . . . . . 731

11.1 Singularities in complex q0 -plane of Feynman propagator. There are


poles and cuts due to three- and more-particle intermediate states
in the diagram. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 821
12.1 An electron on the mass shell absorbing several photons and leaving
again on the mass shell . . . . . . . . . . . . . . . . . . . . . . . . . 839
12.2 An electron on the mass shell absorbing several photons, plus an
additional photon, and leaving again on the mass shell . . . . . . . . 840
12.3 An internal electron loop absorbing several photons, plus an additional photon, and leaving again on the mass shell . . . . . . . . . . 841
12.4 Transition of an atomic state from a state n with energy En to a
lower state n with energy En , thereby emitting a photon with a
frequency = (En En )/h . . . . . . . . . . . . . . . . . . . . . . 847
12.5 Kinematics of Rutherford cross section . . . . . . . . . . . . . . . . . 852
12.6 Lowest-order Feynman Diagrams contributing to Compton Scattering and giving rise to the Klein-Nishina formula . . . . . . . . . . . 857
12.7 Illustration of the photon polarization sum in Compton scattering . 861
12.8 Ratio between total relativistic Compton cross section and nonrelativistic Thomson cross section . . . . . . . . . . . . . . . . . . . . . 864
12.9 Lowest-Order Feynman Diagrams contributing to Electron-Positron
Annihilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864
12.10 Illustration of the photon polarization sum in electron-positron annihilation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 867
12.11 Electron-positron annihilation cross section . . . . . . . . . . . . . . 867
12.12 Lowest-Order Feynman Diagrams contributing to decay of parapositronium decay . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869
12.13 Lowest-Order Feynman Diagrams contributing to decay of orthopositronium decay . . . . . . . . . . . . . . . . . . . . . . . . . . 870
12.14 Trajectories in the simplest classical Bremsstrahlung process: An
electron changing abruptly its momentum . . . . . . . . . . . . . . 871
12.15 Lowest-Order Feynman diagrams contributing to Bremsstrahlung.
The vertical photon line indicates the nuclear Coulomb potential. . 874
12.16 The angles , , in the Bethe-Heitler cross section formula . . . . 875
12.17 Lowest-Order Feynman Diagrams contributing to Electron-Electron
Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876
12.18 Kinematics of electron-electron scattering in the center of mass frame877
12.19 General form of the diagrams contributing to electron-positron scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
12.20 Lowest-order contributions to electron-positron scattering . . . . . . 879

xxvi
12.21 Experimental data for electron-electron and electron-positron scattering at = 900 as a function of the incident electron energy . . . 880
12.22 Cross section for Bhabha scattering at high energy . . . . . . . . . . 881
12.23 The vertex correction responsible for the anomalous magnetic moment882
12.24 Leading hadronic vacuum polarization corrections to a . . . . . . . . 890
12.25 One-loop electroweak radiative corrections to a . The wiggly lines
are gluons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 891
12.26 Measured values of a and prediction of the standard model (SM) . 892
12.27 Lowest-order Feynman diagram for the vacuum polarization . . . . . 892
12.28 Lowest-order Feynman diagram for the vacuum polarization . . . . . 898
12.29 Diagrammatic content in the calculation of the energy shift with the
help of Schrodinger wave functions. A hydrogen atom is represented
by the fat line on the left which results from an infinite sum of photon
exchanges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 910
13.1 The pure current piece of the collective action . . . . . . . . . . . . 958
13.2 The non-polynomial self-interaction terms of the plasmons . . . . . 959
13.3 Free plasmon propagator . . . . . . . . . . . . . . . . . . . . . . . . 959
14.1 Graphical solution of recursion relation (14.30) for the generating
functional W 133j135 of all connected Green functions . . . . . . .
14.2 Tree decompositon of connected Green functions into one-particle
irreducible perts . . . . . . . . . . . . . . . . . . . . . . . . . . . .
14.3 Graphical solution of functional differential equation (14.64) for the
interacting part int 133135 of the effective action . . . . . . . . .
14.4 Recursion relation for two-paticle-irreducible graphs in the effective
action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
14.5
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
14.6 Approximations to F obtained from the extrema of the higher effective action . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
17.1 Effective potential of the 4 -theory for N = 2 in mean-field approxmation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
17.2 Magnetization 0 in mean-field approximation as a function of the
temperature ratio T /TcMF in mean-field approximation . . . . . . .
17.3 Magnetization 0 as a function of external source j in mean-field
approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
17.4 Plot of the symmetric double-well potential . . . . . . . . . . . .
17.5 Classical kink solution in double-well potential connecting the two
degenerate maxima in the reversed potential . . . . . . . . . . . .
17.6 Reversed double-well potential governing the motion of the position
as a function of the imaginary time x . . . . . . . . . . . . . . .
17.7 The order parameter = ||/|0| around a vortex line of strength
n = 1, 2, 3, . . . as a function of the reduced distance r = r/ where r
is the distance from the axis and the healing length. . . . . . . .

. 1010
. 1015
. 1018
. 1024
. 1027
. 1028
. 1086
. 1088
. 1090
. 1096
. 1097
. 1098

. 1102

xxvii
18.1 Dependence of order parameter and magnetic field H on the reduced distance z for a planar domain wall between the normal and
superconductive phases . . . . . . . . . . . . . . . . . . . . . . . .
18.2 Order parameter , and the magnetic field h for an n = 1 vortex
line in a deep type-II superconductor with K = 10. . . . . . . . . .
18.3 Critical field hc1 at which a vortex line of strength n forms when it
first invades a type-II superconductor, as a function of the parameter
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
18.4 Spatial distribution magnetization of the order parameter (x) in
a typical mixed state in which the vortex lines form a hexagonal
lattice (from W.M. Kleiner et al., See References.) . . . . . . . . .
20.1 Free energy as a function of for D = 2. The gap equation deter of the maximum. . . . . . . . . . . . . . . . .
mines the position
20.2 Free energy as a function of for D > 2. The gap equation deter of the maximum. . . . . . . . . . . . . . . . .
mines the position
20.3 Solution of the gap equation (20.48) for = 1 and large volume LD
20.4 Temperature behavior of the correlation length . . . . . . . . . .

. 1122
. 1131

. 1131

. 1132
. 1166
. 1167
1170
. 1171

21.1 Curves in the (, g)-plane corresponding to the same physical


fermion mass Mf = , . . . , 5. These curves are the renormalization group trajectories of the O(N)-symmetric four-fermion model
in the limit N . . . . . . . . . . . . . . . . . . . . . . . . . . . 1200
21.2 Flow of the coupling constant g() as the scale parameter approaches zero (infrared limit) . . . . . . . . . . . . . . . . . . . . . . 1216
22.1
22.2
22.3
22.4
22.5
22.6
22.7
22.8
22.9
22.10

Two-loop diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . 1257


Three-loop diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . 1257
Integrands of the Pade-Borel transform for the Pade approximants . 1267
The inverse of the critical exponent for the classical Heisenberg
model in the O(3)-universality class is plotted as a function of = 4D1268
The inverse of the critical exponent for the O(3)-universality class
is plotted as a function of = 4 D . . . . . . . . . . . . . . . . . . 1269
The inverse of the critical exponent for the O(5)-universality class
is plotted as a function of = 4 D . . . . . . . . . . . . . . . . . . 1269
The inverse of the critical exponent for the O(1)-universality class
(of the Ising model) is plotted as a function of = 4 D . . . . . . 1270
The three successive approximations for at n = 3 (Heisenberg
model) plotted as a function of x = M 2 . . . . . . . . . . . . . . . 1270
The three successive approximations for at n = 4 plotted as a
function of x = M 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 1271
The three successive approximations for at n = 5 plotted as a
function of x = M 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 1271

xxviii
22.11 The
three
successive
approximations
(1 , 2 , 3 , 4 ) = (0.862357, 0.665451, 2.08686, 0.802416) for n = 1
plotted as a function of x = M 2 . . . . . . . . . . . . . . . . . . . 1271
22.12 The highest approximations (M = 4) for n = 3, 4, 5, and the 1/nexpansions to order 1/n2 . . . . . . . . . . . . . . . . . . . . . . . . 1272
23.1 One-loop Feynman diagram in the inverse propagator of the -field 1288
23.2 The function J(z) + 2 in the denominator pf the -propagator (23.60)1290
23.3 The two transition lines in the N g-plane of the chiral Gross-Neveu
model in 2 + dimensions . . . . . . . . . . . . . . . . . . . . . . . . 1314
23.4 Solution of the temperature dependent gap equation, showing the
decrease of the fermion mass M(T ) = (T ) with increasing temperature T /Tc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1319
24.1
24.2
24.3
24.4
24.5
24.6

24.7
24.8
24.9

24.10
24.11
24.12
24.13
24.14
24.15
24.16
24.17
24.18

The total and elastic + -proton cross section showing . . . . . . . 1329


The total and elastic -proton cross section . . . . . . . . . . . . . 1330
The photon-proton and photon-deuteron total cross sections . . . . . 1332
Mirror nuclei 5 B11 and 6 C11 with their excited states . . . . . . . . . 1333
Singlets and triplets of isospin in the nuclei 6 C14 , 7 N14 , 8 O14 . . . . . 1334
Pseudoscalar meson octet states associated with the pions. The same
picture holds for the vector meson octet states with the replacement
(24.58) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1341
Baryon octet states associated with the nucleons. . . . . . . . . . . 1341
Baryon decuplet states associated with the first resonance of the
nucleons. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1342
Quark content of the pseudoscalar meson octet fields. The particle
and quark symbols denote the annihilation parts of the corresponding fields. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1344
Effect of raising and lowering operators on quark and antiquark states1348
Adding the fundamental weights in product representation space of
3 and 3 vectors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1349
The states of the 3-representation with de Swart phases. . . . . . . 1350
The quark-antiquark content of the meson octet meson states with
de Swart phase convention. . . . . . . . . . . . . . . . . . . . . . . 1350
Combination of indices a in the pseudoscalar octet field Ma . . . . 1351
The quark content in the reduction of the product 3 3 = 6 + 3. . 1353
The octet and singlet states obtained from 3 3 in the product space
of tree-quarks. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1353
The irreducible three-quark states 10 and 8 in the product 3 6 (the
symbol (. . .)s denotes complete symmetrization). . . . . . . . . . . 1355
The four quarks u, d, s, c and their position in the three-dimensional
weight space with the new quantum number charm . . . . . . . . 1359

xxix
25.1 The would-be SU(4) partner of the deuteron, with spin-1 and
isospin-0, hiding in the complex energy plane just below the twoparticle cut . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1385
25.2 The pseudoscalar and vector mesons of the 35 representation of SU(6)1389
25.3 The SU(3) content of the particles in the 56 representation of SU(6)
containing the J P = 1/2+ nucleons and the J P = 3/2+ resonances. 1390
25.4 The nucleon resonance of negative parity in the 70 representation of
SU(6) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390
+
25.5 Octet of spin-parity 12 baryons . . . . . . . . . . . . . . . . . . . . 1391
26.1 Quark diagrams for K + and K 0 decays involving strangeness changing neutral currents . . . . . . . . . . . . . . . . . . . . . . . . . .
26.2 Diagrams for the K 0 + decay with compensating strangenesschanging neutral currents . . . . . . . . . . . . . . . . . . . . . . .
26.3 Oscillation in decay rate into + of K 0 beam . . . . . . . . . .
26.4 Asymmetry of the number of mesons as a function of time . . . . .

. 1416
. 1416
. 1419
. 1419

27.1 Propagators in the Yang-Mills theory . . . . . . . . . . . . . . . . . 1438


27.2 Vertices in the Yang-Mills theory . . . . . . . . . . . . . . . . . . . . 1438

List of Tables
4.1 The transformation properties of various composite fields . . . . . . 326
4.2 The lowest Clebsch-Gordan coefficients. . . . . . . . . . . . . . . . . . 379
5.1 Binary systems as sources of gravitational radiation . . . . . . . . . . 402
5.2 Some observed parameters of PSR 1913+16 . . . . . . . . . . . . . . 403
5.3 Typical Astrophysical Sources of Gravitational Radiation . . . . . . . 409
16.1 Connected vacuum diagrams and their multiplicities of the 4 -theory
up to five loops . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
16.2 Connected diagrams of the two-point function and their multiplicities
of the 4 -theory up to four loops . . . . . . . . . . . . . . . . . . .
16.3 Connected diagrams of the four-point function and their multiplicities
of the 4 -theory up to three loops . . . . . . . . . . . . . . . . . . .
16.4 Connected vacuum diagrams and their multiplicities of the 2 Atheory up to four loops . . . . . . . . . . . . . . . . . . . . . . . . .
16.5 Unique matrix representation of all connected vacuum diagrams of
4 -theory up to the order p = 4. . . . . . . . . . . . . . . . . . . . .
16.6 Unique matrix representation of all connected two-point function of
4 -theory up to the order p = 4 . . . . . . . . . . . . . . . . . . . .
16.7 Unique matrix representation of all connected two-point function of
4 -theory up to the order p = 4 . . . . . . . . . . . . . . . . . . . .

. 1075
. 1076
. 1077
. 1079
. 1080
. 1081
. 1082

18.1 The different critical magnetic fields in units of gauss for various impurities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1131
20.1 Values of the lattice Yukawa potential vlD2 (0) of mass l2 at the origin
for different dimensions and l2 . . . . . . . . . . . . . . . . . . . . . . 1174
22.1 Equations determining the coefficients bn (
g0 ) in the strong-coupling
expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
22.2 Coefficients of the successive extension of the expansion coefficients
for n = 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
22.3 Coefficients of the successive extension of the expansion coefficients
for n = 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
22.4 Coefficients of the successive extension of the expansion coefficients
for n = 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
xxx

. 1266
. 1266
. 1266
. 1267

xxxi
22.5 Coefficients of the successive extension of the expansion coefficients
for n = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1267
24.1
24.2
24.3
24.4

Masses and lifetimes of the octet states associated


Structure Constants of SU(3) . . . . . . . . . . .
The symmetric couplings dabc . . . . . . . . . . .
Isoscalar Factors of SU(3) . . . . . . . . . . . . .

with the
. . . . .
. . . . .
. . . . .

nucleons
. . . . .
. . . . .
. . . . .

.
.
.
.

1342
1346
1347
1378

25.1 Action of the different interchange operators . . . . . . . . . . . . . . 1381


25.2 Action of spin and isospin operators in the expansion (25.7) . . . . . 1382
25.3 Eigenvalues of charge and other operators on quark states . . . . . . 1392

xxxii

Any wide piece of ground is the potential site of a palace,


but theres no palace till its built.
Fernando Pessoa (1888-1935), The Book of Disquiet

1
Fundamentals
Before turning to the actual subject of this text it is useful to recall some basic
theoretical background underlying the quantum field theory to be developed.1

1.1

Classical Mechanics

The orbits of a classical-mechanical system are described by a set of time-dependent


generalized coordinates q1 (t), . . . , qN (t). A Lagrangian
L(qi , qi , t)

(1.1)

depending on q1 , . . . , qN and the associated velocities q1 , . . . , qN governs the dynamics of the system. The dots denote the time derivative d/dt. The Lagrangian is at
most a quadratic function of qi . The time integral
A[qi ] =

tb

dt L(qi (t), qi (t), t)

ta

(1.2)

of the Lagrangian along an arbitrary path qi (t) is called the action of this path. The
path being actually chosen by the system as a function of time is called the classical
path or the classical orbit qicl (t). It has the property of extremizing the action in
comparison with all neighboring paths
qi (t) = qicl (t) + qi (t)

(1.3)

having the same endpoints q(tb ), q(ta ). To express this property formally, one
introduces the variation of the action as the linear term in the Taylor expansion of
A[qi ] in powers of qi (t):
A[qi ] {A[qi + qi ] A[qi ]}lin term in qi .

(1.4)

The extremal principle for the classical path is then



A[qi ]

qi (t)=qicl (t)

=0

This chapter is a small excerpt of the introductory sections to the textbook [1].

(1.5)

1 Fundamentals

for all variations of the path around the classical path qi (t) qi (t) qicl (t), which
vanish at the endpoints, i.e., which satisfy
qi (ta ) = qi (tb ) = 0.

(1.6)

Since the action is a time integral of a Lagrangian, the extremality property can
be phrased in terms of differential equations. Let us calculate the variation of A[qi ]
explicitly:
A[qi ] = {A[qi + qi ] A[qi ]}lin
=

tb

tb

tb

ta

ta

ta

dt {L (qi (t) + qi (t), qi (t) + qi (t), t) L (qi (t), qi (t), t)}lin


)

dt

L
L
qi (t) +
qi (t)
qi
qi

dt

tb
L
L
d L
qi (t) +

qi (t) .
qi dt qi
qi
ta

(1.7)

The last expression arises from a partial integration of the qi term. Here, as in the
entire text, repeated indices are understood to be summed (Einsteins summation
convention). The endpoint terms (surface or boundary terms) with the time t equal
to ta and tb may be dropped, due to (1.6). Thus we find for the classical orbit qicl (t)
the Euler-Lagrange equations:
L
d L
=
.
(1.8)
dt qi
qi
There is an alternative formulation of classical dynamics which is based on a
Legendre-transformed function of the Lagrangian called the Hamiltonian
H

L
qi L(qi , qi , t).
qi

(1.9)

Its value at any time is equal to the energy of the system. According to the general
theory of Legendre transformations [2], the natural variables which H depends on
are no longer qi and qi , but qi and the generalized momenta pi , the latter being
defined by the N equations

L(qi , qi , t).
(1.10)
pi
qi
In order to express the Hamiltonian H (pi , qi , t) in terms of its proper variables pi , qi ,
the equations (1.10) have to be solved for qi by a velocity function
qi = vi (pi , qi , t).

(1.11)

This is possible provided the Hessian metric


hij (qi , qi , t)

2
L(qi , qi , t)
qi qj

(1.12)

1.1 Classical Mechanics

is nonsingular. The result is inserted into (1.9), leading to the Hamiltonian as a


function of pi and qi :
H (pi , qi , t) = pi vi (pi , qi , t) L (qi , vi (pi , qi , t) , t) .

(1.13)

In terms of this Hamiltonian, the action is the following functional of pi (t) and qi (t):
A[pi , qi ] =

tb

ta

dt pi (t)qi (t) H(pi (t), qi (t), t) .

(1.14)

This is the so-called canonical form of the action. The classical orbits are now speccl
ified by pcl
i (t), qi (t). They extremize the action in comparison with all neighboring
orbits in which the coordinates qi (t) are varied at fixed endpoints [see (1.3), (1.6)]
whereas the momenta pi (t) are varied without restriction:
qi (t) = qicl (t) + qi (t),

qi (ta ) = qi (tb ) = 0,

(1.15)

pi (t) = pcl
i (t) + pi (t).
In general, the variation is
A[pi , qi ] =

tb

tb

ta

ta

"

H
H
pi
qi
dt pi (t)qi (t) + pi (t) qi (t)
pi
qi
dt

("

"

H
H
pi p i (t) +
qi
qi (t)
pi
qi

(1.16)

tb

+ pi (t)qi (t) .
ta

cl
Since this variation has to vanish for the classical orbits, we find that pcl
i (t), qi (t)
must be solutions of the Hamilton equations of motion

H
,
qi
H
=
.
pi

p i =
qi

(1.17)

These agree with the Euler-Lagrange equations (1.8) via (1.9) and (1.10), as can
easily be verified. The 2N-dimensional space of all pi and qi is called the phase
space.
An arbitrary function O(pi (t), qi (t), t) changes along an arbitrary path as follows:
O
O
O
d
O (pi (t), qi (t), t) =
p i +
qi +
.
dt
pi
qi
t

(1.18)

If the path coincides with a classical orbit, we may insert (1.17) and find
dO
H O O H O

+
=
dt
pi qi
pi qi
t
O
.
{H, O} +
t

(1.19)

1 Fundamentals

Here we have introduced the symbol {. . . , . . .} called Poisson brackets:


{A, B}

A B B A

,
pi qi
pi qi

(1.20)

again with the Einstein summation convention for the repeated index i. The Poisson
brackets have the obvious properties
{A, B} = {B, A}
{A, {B, C}} + {B, {C, A}} + {C, {A, B}} = 0

antisymmetry,

(1.21)

Jacobi identity.

(1.22)

If two quantities have vanishing Poisson brackets, they are said to commute.
The original Hamilton equations are a special case of (21.179):
d
pi = {H, pi } =
dt
d
qi = {H, qi } =
dt

pi H
H
H pi

=
,
pj qj
pj qj
qi
H qi
qi H
H

=
.
pj qj
pj qj
pi

(1.23)

By definition, the phase space variables pi , qi satisfy the Poisson brackets


{pi , qj } = ij ,

{pi , pj } = 0,

(1.24)

{qi , qj } = 0.

A function O(pi , qi ) which has no explicit dependence on time and which, moreover, commutes with H (i.e., {O, H} = 0), is a constant of motion along the classical
path, due to (21.179). In particular, H itself is often time-independent, i.e., of the
form
H = H(pi , qi ).
(1.25)
Then, since H commutes with itself, the energy is a constant of motion.
The Lagrangian formalism has the virtue of being independent of the particular
choice of the coordinates qi . Let Qi be any other set of coordinates describing the
system which is connected with qi by what is called a local 2 or point transformation
qi = fi (Qj , t).

(1.26)

Certainly, to be of use, this relation must be invertible, at least in some neighborhood


of the classical path,
Qi = f 1 i (qj , t).
(1.27)
2

The word local means here at a specific time. This terminology is of common use in field
theory where local means, more generally, at a specific spacetime point .

1.1 Classical Mechanics

Otherwise Qi and qi could not both parametrize the same system. Therefore, fi
must have a nonvanishing Jacobi determinant:
det

fi
Qj

6= 0.

(1.28)

In terms of Qi , the initial Lagrangian takes the form


L Qj , Q j , t L fi (Qj , t) , fi (Qj , t) , t
and the action reads

A =

tb
ta
tb
ta

(1.29)

dt L Qj (t), Q j (t), t

(1.30)

dt L fi (Qj (t), t) , fi (Qj (t), t) , t .




By performing variations Qj (t), Q j (t) in the first expression while keeping


Qj (ta ) = Qj (tb ) = 0, we find the equations of motion
L
d L

= 0.
dt Q j
Qj

(1.31)

The variation of the lower expression, on the other hand, gives


A =

tb

tb

ta

ta

dt
dt

L
L
fi +
fi
qi
qi

L
L tb
d L
fi +

fi .
ta
qi dt qi
qi

(1.32)

If qi is arbitrary, then so is fi . Moreover, with qi (ta ) = qi (tb ) = 0, also fi


vanishes at the endpoints. Hence the extremum of the action is determined equally
well by the Euler-Lagrange equations for Qj (t) [as it was by those for qi (t)].
Note that the locality property is quite restrictive for the transformation of the
generalized velocities qi (t). They will necessarily be linear in Q j :
fi
fi
Qj +
qi = fi (Qj , t) =
.
Qj
t

(1.33)

In phase space, there exists also the possibility of performing local changes of
the canonical coordinates pi , qi to new ones Pj , Qj . Let them be related by
pi = pi (Pj , Qj , t),
qi = qi (Pj , Qj , t),
with the inverse relations

Pj = Pj (pi , qi , t),
Qj = Qj (pi , qi , t).

(1.34)

(1.35)

1 Fundamentals

However, while the Euler-Lagrange equations maintain their form under any local
change of coordinates, the Hamilton equations do not hold, in general, for any transformed coordinates Pj (t), Qj (t). The local transformations pi (t), qi (t) Pj (t), Qj (t)
for which they hold, are referred to as canonical . They are characterized by the form
invariance of the action, up to an arbitrary surface term,
tb

ta

dt [pi qi H(pi , qi , t)] =

tb

ta

dt Pj Q j H (Pj , Qj , t)
tb

+ F (Pj , Qj , t)

ta

(1.36)

where H (Pj , Qj , t) is some new Hamiltonian. Its relation with H(pi , qi , t) must be
chosen in such a way that the equality of the action holds for any path pi (t), qi (t)
connecting the same endpoints (at least any in some neighborhood of the classical
orbits). If such an invariance exists then a variation of this action yields for Pj (t)
and Qj (t) the Hamilton equations of motion governed by H :

H
P i =
,
Qi
(1.37)

H
.
Q i =
Pi
The invariance (1.36) can be expressed differently by rewriting the integral on the
left-hand side in terms of the new variables Pj (t), Qj (t),
Z

tb

ta

dt

pi

qi
qi
qi
Pj +
Qj +
Pj
Qj
t

H(pi (Pj , Qj , t), qi (Pj , Qj , t), t) ,

(1.38)

and subtracting it from the right-hand side, leading to


Z

tb

ta

qi
P j pi
Qj

dQj pi

qi
dPj
P!j )

qi
H dt
H + pi
t

tb

F (Pj , Qj , t) .

(1.39)

ta

The integral is now a line integral along a curve in the (2N + 1)-dimensional space,
consisting of the 2N-dimensional phase space variables pi , qi and of the time t.
The right-hand side depends only on the endpoints. Thus we conclude that the
integrand on the left-hand side must be a total differential. As such it has to satisfy
the standard Schwarz integrability conditions [3], according to which all second
derivatives have to be independent of the sequence of differentiation. Explicitly,
these conditions are
pi qi
qi pi

= kl ,
Pk Ql Pk Ql
pi qi
qi pi

= 0,
Pk Pl Pk Pl
qi pi
pi qi

= 0,
Qk Ql Qk Ql

(1.40)

1.1 Classical Mechanics

and

qi pi
(H H)
pi qi

=
,
t Pl
t Pl
Pl
qi pi
(H H)
pi qi

=
.
t Ql
t Ql
Ql

(1.41)

The first three equations define the so-called Lagrange brackets in terms of which
they are written as
(Pk , Ql ) = kl ,
(Pk , Pl ) = 0,
(Qk , Ql ) = 0.

(1.42)

Time-dependent coordinate transformations satisfying these equations are called


symplectic. After a little algebra involving the matrix of derivatives

its inverse

J =

J 1 =

Pi /pj

Pi /qj

Qi /pj

Qi /qj

pi /Pj

pi /Qj

qi /Pj

qi /Qj

and the symplectic unit matrix

E=

0
ij

ij
0

(1.43)

(1.44)

,
,

(1.45)

we find that the Lagrange brackets (1.42) are equivalent to the Poisson brackets
{Pk , Ql } = kl ,
{Pk , Pl } = 0,
{Qk , Ql } = 0.

(1.46)

This follows from the fact that the 2N 2N matrix formed from the Lagrange
brackets

(Qi , Pj )
(Qi , Qj )

L
(1.47)
(Pi , Pj )
(Pi , Qj )

can be written as (E 1 J 1 E)T J 1 , while an analogous matrix formed from the


Poisson brackets

{Pi , Qj }
{Pi , Pj }

P
(1.48)
{Qi , Qj }
{Qi , Pj }
is equal to J(E 1 JE)T . Hence L = P 1 , so that (1.42) and (1.46) are equivalent to
each other. Note that the Lagrange brackets (1.42) [and thus the Poisson brackets

1 Fundamentals

(1.46)] ensure pi qi Pj Q j to be a total differential of some function of Pj and Qj in


the 2N-dimensional phase space:
d
pi qi Pj Q j = G(Pj , Qj , t).
dt

(1.49)

The Poisson brackets (1.46) for Pi , Qi have the same form as those in Eqs. (21.182)
for the original phase space variables pi , qi .
The other two equations (1.41) relate the new Hamiltonian to the old one. They
can always be used to construct H (Pj , Qj , t) from H(pi , qi , t). The Lagrange brackets (1.42) or Poisson brackets (1.46) are therefore both necessary and sufficient for
the transformation pi , qi Pj , Qj to be canonical.
A canonical transformation preserves the volume in phase space. This follows
from the fact that the matrix product J(E 1 JE)T is equal to the 2N 2N unit
matrix (1.48). Hence det (J) = 1 and
YZ

[dpi dqi ] =

YZ

[dPj dQj ] .

(1.50)

It is obvious that the process of canonical transformations is reflexive. It may


be viewed just as well from the opposite side, with the roles of pi , qi and Pj , Qj
exchanged [we could just as well have considered the integrand (1.39) as a complete
differential in Pj , Qj , t space].
Once a system is described in terms of new canonical coordinates Pj , Qj , we
introduce the new Poisson brackets
{A, B}

A B
B A

,
Pj Qj
Pj Qj

(1.51)

and the equation of motion for an arbitrary observable quantity O (Pj (t), Qj (t), t)
becomes with (21.183)
O
dO

= {H , O} +
,
(1.52)
dt
t
by complete analogy with (21.179). The new Poisson brackets automatically guarantee the canonical commutation rules
{Pi , Qj }
{Pi , Pj }

= ij ,
= 0,

(1.53)

{Qi , Qj } = 0.

A standard class of canonical transformations can be constructed by introducing


a generating function F satisfying a relation of the type (1.36), but depending
explicitly on half an old and half a new set of canonical coordinates, for instance
F = F (qi , Qj , t).

(1.54)

1.1 Classical Mechanics

One now considers the equation


Z

tb
ta

dt [pi qi H(pi , qi , t)] =

replaces Pj Q j by Pj Qj +

tb

ta

"

d
dt Pj Q j H (Pj , Qj , t) + F (qi , Qj , t) , (1.55)
dt

d
PQ,
dt j j

defines

F (qi , Pj , t) F (qi , Qj , t) + Pj Qj ,
and works out the derivatives. This yields
Z

tb

ta

dt pi qi + P j Qj [H(pi , qi , t) H (Pj , Qj , t)]


Z

tb

ta

F
F
F
(qi , Pj , t)qi +
(qi , Pj , t)P j +
(qi , Pj , t) .
dt
qi
Pj
t

(1.56)

A comparison of the two sides yields the equations for the canonical transformation
pi =
Qj

F (qi , Pj , t),
qi

F (qi , Pj , t).
=
Pj

(1.57)

The second equation shows that the above relation between F (qi , Pj , t) and
F (qi , Qj , t) amounts to a Legendre transformation.
The new Hamiltonian is
H (Pj , Qj , t) = H(pi , qi , t) +

F (qi , Pj , t).
t

(1.58)

Instead of (1.54) we could, of course, also have chosen functions with other mixtures
of arguments such as F (qi , Pj , t), F (pi , Qj , t), F (pi , Pj , t) to generate simple canonical
transformations.
A particularly important canonical transformation arises by choosing a generating function F (qi , Pj ) in such a way that it leads to time-independent momenta
Pj j . Coordinates Qj with this property are called cyclic. To find cyclic coordinates we must search for a generating function F (qj , Pj , t) which makes the
transformed H in (1.58) vanish identically. Then all derivatives with respect to the
coordinates vanish and the new momenta Pj are trivially constant. Thus we seek a
solution of the equation

F (qi , Pj , t) = H(pi , qi , t),


t

(1.59)

where the momentum variables in the Hamiltonian obey the first equation of (1.57).
This leads to the following partial differential equation for F (qi , Pj , t):
t F (qi , Pj , t) = H(qi F (qi , Pj , t), qi , t),

(1.60)

10

1 Fundamentals

called the Hamilton-Jacobi equation. Here and in the sequel we shall often use the
short notations for partial derivatives t /t, qi /qi .
A generating function which achieves this goal is supplied by the action functional
(1.14). When following the classical solutions starting from a fixed initial point and
running to all possible final points qi at a time t, the associated actions of these
solutions form a function A(qi , t). Expression (1.14) show that if a particle moves
along a classical trajectory, and the path is varied without keeping the endpoints
fixed, the action changes as a function of the end positions (1.16) by
A[pi , qi ] = pi (tb )qi (tb ) pi (ta )qi (ta ).

(1.61)

From this we deduce immediately the first of the equations (1.57), now for the
generating function A(qi , t):
pi =

A(qi , t).
qi

(1.62)

Moreover, the function A(qi , t) has the time derivative


d
A(qi (t), t) = pi (t)qi (t) H(pi (t), qi (t), t).
dt

(1.63)

Together with (1.62) this implies


t A(qi , t) = H(pi , qi , t).

(1.64)

If the momenta pi on the right-hand side are replaced according to (1.62), A(qi , t)
is indeed seen to be a solution of the Hamilton-Jacobi differential equation:
t A(qi , t) = H(qi A(qi , t), qi , t).

1.2

(1.65)

Relativistic Mechanics in Curved Spacetime

The classical action of a relativistic spinless point particle in a curved fourdimensional spacetime is usually written as an integral
A = Mc

d L(q, q)
= Mc

d g q ( )q ( ),

(1.66)

where is an arbitrary parameter of the trajectory. It can be chosen in the final


trajectory to make L(q, q)
1, in which case it coincides with the proper time of
the particle. For arbitrary , the Euler-Lagrange equation (1.8) reads
"

1
1
d
g q =
( g) q q .
dt L(q, q)

2L(q, q)

(1.67)

If is the proper time where L(q, q)


1, this simplifies to
1
d
(g q ) = ( g ) q q ,
dt
2

(1.68)

11

1.3 Quantum Mechanics

or

1
g q =
g g q q .
2

(1.69)

For brevity, we have denoted partial derivatives /q by . This partial derivative


is supposed to apply only to the quantity right behind it. At this point one introduces
the Christoffel symbol
1 ( g + g g ),

(1.70)

and the Christoffel symbol of the second kind [6]:


g
.

(1.71)

q q = 0.
q +

(1.72)

Then (1.69) can be written as

Since the solutions of this equation minimize the length of a curve in spacetime,
they are called geodesics.

1.3

Quantum Mechanics

Historically, the extension of classical mechanics to quantum mechanics became


necessary in order to understand the stability of atomic orbits and the discrete
nature of atomic spectra. It soon became clear that these phenomena reflect the
fact that at a sufficiently short length scale, small material particles such as electrons
behave like waves, called material waves. The fact that waves cannot be squeezed
into an arbitrarily small volume without increasing indefinitely their frequency and
thus their energy, prevents the collapse of the electrons into the nucleus, which
would take place in classical mechanics. The discreteness of the atomic states of an
electron are a manifestation of standing material waves in the atomic potential well,
by analogy with the standing waves of electromagnetism in a cavity.

1.3.1

Bragg Reflections and Interference

The most direct manifestation of the wave nature of small particles is seen in diffraction experiments on periodic structures, for example of electrons diffracted by a crystal. If an electron beam of fixed momentum p passes through a crystal, it emerges
along sharply peaked angles. These are the well-known Bragg reflections. They
look very similar to the interference patterns of electromagnetic waves. In fact, it
is possible to use the same mathematical framework to explain these patterns as in
electromagnetism. A free particle moving with momentum
p = (p1 , p2 , . . . , pD ).

(1.73)

12

1 Fundamentals

through a D-dimensional Euclidean space spanned by the Cartesian coordinate vectors


x = (x1 , x2 , . . . , xD )

(1.74)

is associated with a plane wave, whose field strength or wave function has the form
p (x, t) = eikxit ,

(1.75)

where k is the wave vector pointing into the direction of p and is the wave frequency. Each scattering center, say at x , becomes a source of a spherical wave
with the spatial behavior eikR /R (with R |x x | and k |k|) and the wavelength = 2/k. At the detector, all field strengths have to be added to the total
field strength (x, t). The absolute square of the total field strength, |(x, t)|2, is
proportional to the number of electrons arriving at the detector.
The standard experiment where these rules can most simply be applied consists
of an electron beam impinging vertically upon a flat screen with two parallel slits
a with spacing d. At a large distance R behind these, one observes the number of
particles arriving per unit time (see Fig. 1.1)

2 1
1
1
dN
|1 + 2 |2 eik(R+ 2 d sin ) + eik(R 2 d sin ) 2 ,
dt
R

(1.76)

where is the angle of deflection from the normal.

dN
dt


2
1
1


eik(R+ 2 d sin ) + eik(R 2 d sin )

eikx

Figure 1.1 Probability distribution of particle behind double slit, being proportional to
the absolute square of the sum of the two complex field strengths.

Conventionally, the wave function (x, t) is normalized to describe a single particle. Its absolute square gives directly the probability density of the particle at the
place x in space, i.e., d3 x |(x, t)|2 is the probability of finding the particle in the
volume element d3 x around x.

13

1.3 Quantum Mechanics

1.3.2

Matter Waves

From the experimentally observed relation between the momentum and the size of
the angular deflection of the diffracted beam of the particles, one deduces the
relation between momentum and wave vector
p=h
k,

(1.77)

where h
is the universal Planck constant whose dimension is equal to that of an
action,
h
= 1.0545919(80) 1027 erg sec
(1.78)
h

2
(the number in parentheses indicating the experimental uncertainty of the last two
digits before it). A similar relation holds between the energy and the frequency of
the wave (x, t). It may be determined by an absorption process in which a light
wave hits an electron and kicks it out of the surface of a metal, the well-known
photoelectric effect. From the threshold property of this effect one learns that an
electromagnetic wave oscillating in time as eit can transfer to the electron the
energy
E=h
,
(1.79)
where the proportionality constant h
is the same as in (1.77). The reason for this lies
in the properties of electromagnetic waves. On the one hand, their frequency and
the wave vector k satisfy the relation /c = |k|, where c is the light velocity defined
to be c 299 792.458 km/s. The energy and momentum are related by E/c = |p|.
Thus, the quanta of electromagnetic waves, the photons, certainly satisfy (1.77) and
the constant h
must be the same as in Eq. (1.79).
With matter waves and photons sharing the same relations (1.77), it is suggestive
to postulate also the relation (1.79) between energy and frequency to be universal for
the waves of all particles, massive and massless ones. All free particles of momentum
p are described by a plane wave of wavelength = 2/|k| = 2h/|p|, with the
explicit form
p (x, t) = N ei(pxEp t)/h ,
(1.80)
where N is some normalization constant. In a finite volume, the wave function
is normalized to unity. In an infinite volume, this normalization makes the wave
function vanish. To avoid this, the current density of the particle probability
j(x, t) i

h

(x, t) (x, t)
2m

(1.81)

is normalized in some convenient way, where is a short notation for the difference
between forward- and backward-derivatives

(x, t) (x, t) (x, t) (x, t) (x, t) (x, t)


(x, t)(x, t) [ (x, t)] (x, t).

(1.82)

14

1 Fundamentals

The energy Ep depends on the momentum of the particle in the classical way,
i.e., for nonrelativistic
material particles of mass M it is Ep = p2 /2M, for relativistic
2
ones Ep = c p + M 2 c2 , and for massless particles such as photons Ep = c|p|. The
common relation Ep = h
for photons and matter waves is necessary to guarantee
conservation of energy in quantum mechanics.
In general, both momentum and energy of a particle are not sharply defined as
in the plane-wave function (1.80). Usually, a particle wave is some superposition of
plane waves (1.80)
Z
d3 p
f (p)ei(pxEp t)/h .
(1.83)
(x, t) =
(2h)3
By the Fourier inversion theorem, f (p) can be calculated via the integral
f (p) =

d3 x eipx/h (x, 0).

(1.84)

With an appropriate choice of f (p) it is possible to prepare (x, t) in any desired


form at some initial time, say at t = 0. For example, (x, 0) may be a function
sharply centered around a space point x
. Then f (p) is approximately a pure phase
f (p) eipx/h , and the wave contains all momenta with equal probability. Conversely, if the particle amplitude is spread out in space, its momentum distribution
is confined to a small region. The limiting f (p) is concentrated at a specific mo . The particle is found at each point in space with equal probability, with
mentum p
the amplitude oscillating like (x, t) ei(pxEp t)/h .
In general, the width of (x, 0) in space and of f (p) in momentum space are
inversely proportional to each other:
x p h
.

(1.85)

This is the content of Heisenbergs principle of uncertainty. If the wave is localized


in a finite region of space while having at the same time a fairly well-defined average
, it is called a wave packet. The maximum in the associated probability
momentum p
density can be shown from (1.83) to move with a velocity
= Ep / p
.
v

(1.86)

.
This coincides with the velocity of a classical particle of momentum p

1.3.3

Schr
odinger Equation

Suppose now that the particle is nonrelativistic and has a mass M. The classical
Hamiltonian, and thus the energy Ep , are given by
p2
.
2M
We may therefore derive the following identity for the wave field p (x, t):
H(p) = Ep =

d3 p
f (p) [H(p) Ep ] ei(pxEp t)/h = 0.
(2h)3

(1.87)

(1.88)

15

1.3 Quantum Mechanics

The arguments inside the brackets can be removed from the integral by observing
that p and Ep inside the integral are equivalent to the differential operators
= ihx ,
p

E = iht

(1.89)

outside. Then, Eq. (1.88) may be written as the differential equation


[H(ihx ) iht )](x, t) = 0.

(1.90)

This is the Schrodinger equation for the wave function of a free particle. The equation suggests that the motion of a particle with an arbitrary Hamiltonian H(p, x, t)
follows the straightforward generalization of (1.90)


iht (x, t) = 0,
H

(1.91)

is the differential operator


where H
H(ihx , x, t).
H

(1.92)

from the classical Hamiltonian H(p, x, t) by the substitution


The rule of obtaining H
= ihx will be referred to as the correspondence principle.3
pp
This simple correspondence principle holds only in Cartesian coordinates. A
slight generalization is possible to generalized coordinates qi (t) which are of the
quasi-Cartesian type. For these, the so-called dynamical metric, or Hessian), defined
in the Lagrangian formalism by
gij (q)

2
L(q, q),

qi qi

(1.93)

2
H(p, q),
pi pj

(1.94)

and in the Hamiltonian formalism by


g ij (q)

must be qi -independent and proportional to the unit matrix. Then the momentum
operators are, as in (1.89),
pi i
3

.
qi

(1.95)

Our formulation of this principle is slightly stronger than the historical one used in the initial
phase of quantum mechanics, which gave certain translation rules between classical and quantummechanical relations. The substitution rule for the momentum runs also under the name Jordan
rule.

16

1 Fundamentals

For such quasi-Cartesian generalized coordinates, the system may be quantized alternatively `a la Heisenberg by assuming pi (t) and qi (t) to be Heisenberg operators
pi H (t) and qi H (t) satisfying the canonical commutation rules (1.275):
[
piH (t), qjH (t)] = ihij ,

[
piH (t), pjH (t)] = 0,

(1.96)

[
qiH (t), qjH (t)] = 0.
This pecularity of the canonical quantization rules will be discussed further in Sections 1.131.15.
The Schrodinger operators (1.89) of momentum and energy satisfy with x and t
the so-called canonical commutation relations
[
pi , xj ] = ih,

t] = 0 = ih.
[E,

(1.97)

The linear combinations of the solutions of the Schrodinger equation (1.91)


form, at each time t, a Hilbert space. If the Hamiltonian does not depend explicitly on time, the Hilbert space can be spanned by the energy eigenstates
En (x, t) = eiEn t/h En (x), where En (x) are time-independent stationary states,
which solve the time-independent Schrodinger equation
p, x)En (x) = En En (x).
H(

(1.98)

The validity of the Schrodinger theory (1.91) is confirmed by experiment, most


notably for the Coulomb Hamiltonian
p2
e2
,
(1.99)
2M
r
which governs the quantum mechanics of the hydrogen atom in the center-of-mass
coordinate system of electron and proton, where M is the reduced mass of the two
particles.
Since the square of the wave function, |(x, t)|2 , is interpreted as the probability
density of a single particle in a finite volume, the integral over the entire volume
must be normalized to unity:
H(p, x) =

d3 x |(x, t)|2 = 1.

(1.100)

For a stable particle, this normalization must remain the same at all times. If
(x, t) is to follow the Schrodinger equation (1.91), this is assured if and only if the
Hamiltonian operator is Hermitian,4 i.e., if it satisfies for arbitrary wave functions
1 , 2 the equality
Z
4

2 (x, t)] 1 (x, t) =


d x [H
3

1 (x, t).
d3 x 2 (x, t)H

(1.101)

Problems arising from unboundedness or discontinuities of the Hamiltonian and other


quantum-mechanical operators, such as restrictions of the domains of definition, are ignored here
since they are well understood. Correspondingly we do not distinguish between Hermitian and selfadjoint operators (see J. vNeumann, Mathematische Grundlagen der Quantenmechanik , Springer,
Berlin, 1932).

17

1.3 Quantum Mechanics

of the operator H,
which satThe left-hand side defines the Hermitian-adjoint H
isfies the identity
Z

1 (x, t)
x 2 (x, t)H

2 (x, t)] 1 (x, t)


d3 x [H

(1.102)

is Hermitian, if it coincides
for all wave functions 1 (x, t), 2 (x, t). An operator H

:
with its Hermitian-adjoint H
=H
.
H

(1.103)

Let
us calculate the time change of the integral over two arbitrary wave functions,
Z
3
d x 2 (x, t)1 (x, t). With the Schrodinger equation (1.91), this time change van is Hermitian:
ishes indeed as long as H
ih

d
dt
=

d3 x 2 (x, t)1 (x, t)


1 (x, t)
d3 x 2 (x, t)H

(1.104)
2 (x, t)] 1 (x, t) = 0.
d3 x [H

This
also implies the time independence of the normalization integral
R 3
d x |(x, t)|2 = 1.
is not Hermitian, one can always find an eigenstate of H
whose
Conversely, if H

norm changes with time: any eigenstate of (H H )/i has this property.
will automatically
= ihx and x are themselves Hermitian operators, H
Since p
be a Hermitian operator if it is a sum of a kinetic and a potential energy:
H(p, x, t) = T (p, t) + V (x, t).

(1.105)

This is always the case for nonrelativistic particles in Cartesian coordinates x. If p


and x appear in one and the same term of H, for instance as p2 x2 , the correspon Then
dence principle does not lead to a unique quantum-mechanical operator H.
there seem to be, in principle, several Hermitian operators which, in the above exam and two x
operators [for instance
ple, can be constructed from the product of two p
2 + x
2p
2 + p
x
2p
with ++ = 1]. They all correspond to the same classical

p2 x
p2 x2 . At first sight it appears as though only a comparison with experiment could
select the correct operator ordering. This is referred to as the operator-ordering
problem of quantum mechanics which has plagued many researchers in the past. If
the ordering problem is caused by the geometry of the space in which the particle moves, there exists a surprisingly simple geometric principle which specifies the
ordering in the physically correct way. These are explained in Chapter 10 of the
textbook [1]. Here we avoid such ambiguities by assuming H(p, x, t) to have the
standard form (1.105), unless otherwise stated.

18

1.3.4

1 Fundamentals

Particle Current Conservation

The conservation of the total probability (1.100) is a consequence of a more general


local conservation law linking the current density of the particle probability
j(x, t) i

(x, t) (x, t)
2m

(1.106)

with the probability density


(x, t) = (x, t)(x, t)

(1.107)

t (x, t) = j(x, t).

(1.108)

via the relation


By integrating this current conservation law over a volume V enclosed by a surface
S, and using Greens theorem, one finds
Z

d3 x t (x, t) =

d3 x j(x, t) =

dS j(x, t),

(1.109)

where dS are the directed infinitesimal surface elements. This equation states that
the probability in a volume decreases by the same amount by which probability
leaves the surface via the current j(x, t).
By extending the integral (1.109) over the entire space and assuming the currents
to vanish at spatial infinity, we recover the conservation of the total probability
(1.100).
More general dynamical systems with N particles in Euclidean space are
parametrized in terms of 3N Cartesian coordinates x ( = 1, . . . , N). The Hamiltonian has the form
N
X
p2
+ V (x , t),
(1.110)
H(p , x , t) =
=1 2M
where the arguments p , x in H and V stand for all p s, x with = 1, 2, 3, . . . , N.
The wave function (x , t) satisfies the N-particle Schrodinger equation
(

1.4

N
X

=1

"

h
2
x 2 + V (x , t)
2M

#)

(x , t) = iht (x , t).

(1.111)

Diracs Bra-Ket Formalism

Mathematically speaking, the wave function (x, t) may be considered as a vector in


an infinite-dimensional complex vector space called Hilbert space. The configuration
space variable x plays the role of a continuous index of these vectors. An obvious
contact with the usual vector notation may be established, in which a D-dimensional
vector v is given in terms of its components vi with a subscript i = 1, . . . D, by
writing the argument x of (x, t) as a subscript:
(x, t) x (t).

(1.112)

19

1.4 Diracs Bra-Ket Formalism

The usual norm of a complex vector is defined by


|v|2 =

vi vi .

(1.113)

(1.114)

The continuous version of this is


2

|| =

x x (t)x (t)

d3 x (x, t)(x, t).

The normalization condition (1.100) requires that the wave functions have the norm
|| = 1, i.e., that they are unit vectors in the Hilbert space.

1.4.1

Basis Transformations

In a vector space, there are many possible choices of orthonormal basis vectors bi a
labeled by a = 1, . . . , D, in terms of which5
vi =

bi a va ,

(1.115)

with the components va given by the scalar products


va

bi a vi .

(1.116)

The latter equation is a consequence of the orthogonality relation 6


X

bi a bi a = aa ,

(1.117)

which in a finite-dimensional vector space implies the completeness relation


X

bi a bj a = ij .

(1.118)

In the space of wave functions (1.112) there exists a special set of basis functions
called local basis functions of particular importance. It may be constructed in the
following fashion: Imagine the continuum of space points to be coarse-grained into
a cubic lattice of mesh size , at positions
xn = (n1 , n2 , n3 ),

n1,2,3 = 0, 1, 2, . . . .

(1.119)

Let hn (x) be a function that vanishes everywhere in space, except in a cube of size
3 centered around xn , i.e., for each component xi of x,
(
1/ 3
|xi xn i | /2, i = 1, 2, 3.
hn (x) =
(1.120)
0
otherwise.
P
(b)
Mathematicians would expand more precisely vi = a bi a va , but physicists prefer to shorten
the notation by distinguishing the different components via different types of subscripts, using for
the initial components i, j, k, . . . and for the b-transformed components a, b, c, . . . .
6
An orthogonality relation implies usually a unit norm and is thus really an orthonormality
relation but this name is rarely used.
5

20

1 Fundamentals

These functions are certainly orthonormal:


Z

(1.121)

hn (x)n (t)

(1.122)

d3 x hn (x) hn (x) = nn .

Consider now the expansion


(x, t) =

X
n

with the coefficients


n (t) =

d3 x hn (x) (x, t)

3 (xn , t).

(1.123)

It provides an excellent approximation to the true wave function (x, t), as long as
the mesh size is much smaller than the scale over which (x, t) varies. In fact, if
(x, t) is integrable, the integral over the sum (1.122) will always converge to (x, t).
The same convergence of discrete approximations is found in any scalar product,
and thus in any observable probability amplitudes. They can all be calculated with
arbitrary accuracy knowing the discrete components of the type (1.123) in the limit
0. The functions hn (x) may therefore be used as an approximate basis in the
same way as the previous basis functions f a (x), g b(x), with any desired accuracy
depending on the choice of .
In general, there are many possible orthonormal basis functions f a (x) in the
Hilbert space which satisfy the orthonormality relation
Z

d3 x f a (x) f a (x) = aa ,

(1.124)

in terms of which we can expand


X

(x, t) =

f a (x)a (t),

(1.125)

with the coefficients


a (t) =

d3 x f a (x) (x, t).

(1.126)

Suppose we use another orthonormal basis fb (x) with the orthonormality relation
Z

d3 x fb (x) fb (x) = bb ,

X
b

fb (x)fb (x ) = (3) (x x ),

to re-expand
(x, t) =

(1.127)

b (t),
fb (x)

(1.128)

d3 x fb (x) (x, t).

(1.129)

X
b

with the components


b (t) =

Inserting (1.125) shows that the components are related to each other by
b (t) =

X Z
a

d3 x fb (x) f a (x) a (t).

(1.130)

21

1.4 Diracs Bra-Ket Formalism

1.4.2

Bracket Notation

It is useful to write the scalar products between two wave functions occurring in the
above basis transformations in the so-called Dirac bracket notation as
hb|ai

d3 x fb (x) f a (x).

(1.131)

In this notation, the components of the state vector (x, t) in (1.126), (1.129) are
a (t) = ha|(t)i,

(1.132)

b (t) = hb|(t)i.

The transformation formula (1.130) takes the form


hb|(t)i =

X
a

hb|aiha|(t)i.

(1.133)

The right-hand side of this equation may be formally viewed as a result of inserting
the abstract relation
X
|aiha| = 1
(1.134)
a

between hb| and |(t)i on the left-hand side:

hb|(t)i = hb|1|(t)i =

X
a

hb|aiha|(t)i.

(1.135)

Since this expansion is only possible if the functions f b (x) form a complete basis,
the relation (1.134) is alternative, abstract way of stating the completeness of the
basis functions. It may be referred to as completeness relation a` la Dirac.
Since the scalar products are written in the form of brackets ha|a i, Dirac called
the formal objects ha| and |a i, from which the brackets are composed, bra and ket,
respectively. In the Dirac bracket notation, the orthonormality of the basis f a (x)
and g b(x) may be expressed as follows:

ha|a i =
hb|b i =

d3 x f a (x) f a (x) = aa ,

d x fb (x) fb (x) = bb .

(1.136)

In the same spirit we introduce abstract bra and ket vectors associated with the
basis functions hn (x) of Eq. (1.120), denoting them by hxn | and |xn i, respectively,
and writing the orthogonality relation (1.121) in bracket notation as
hxn |xn i

d3 x hn (x) hn (x) = nn .

The components n (t) may be considered as the scalar products

n (t) hxn |(t)i 3 (xn , t).

(1.137)

(1.138)

22

1 Fundamentals

Changes of basis vectors, for instance from |xn i to the states |ai, can be performed
according to the rules developed above by inserting a completeness relation a` la
Dirac of the type (1.134). Thus we may expand
n (t) = hxn |(t)i =

X
a

hxn |aiha|(t)i.

(1.139)

Also the inverse relation is true:


ha|(t)i =

X
n

ha|xn ihxn |(t)i.

(1.140)

This is, of course, just an approximation to the integral


Z

d3 x hn (x) hx|(t)i.

(1.141)

The completeness of the basis hn (x) may therefore be expressed via the abstract
relation
X
|xn ihxn | 1.
(1.142)
n

The approximate sign turns into an equality sign in the limit of zero mesh size,
0.

1.4.3

Continuum Limit

In ordinary calculus, finer and finer sums are eventually replaced by integrals. The
same thing is done here. We define new continuous scalar products
1
hx|(t)i hxn |(t)i,
3

(1.143)

where xn are the lattice points closest to x. With (1.138), the right-hand side is
equal to (xn , t). In the limit 0, x and xn coincide and we have
hx|(t)i (x, t).

(1.144)

The completeness relation can be used to write


ha|(t)i

ha|xn ihxn |(t)i

3 ha|xihx|(t)i

which in the limit 0 becomes


ha|(t)i =

(1.145)
x=xn

d3 x ha|xihx|(t)i.

(1.146)

This may be viewed as the result of inserting the formal completeness relation of
the limiting local bra and ket basis vectors hx| and |xi,
Z

d3 x |xihx| = 1,

(1.147)

23

1.4 Diracs Bra-Ket Formalism

evaluated between the vectors ha| and |(t)i.


With the limiting local basis, the wave functions can be treated as components
of the state vectors |(t)i with respect to the local basis |xi in the same way as any
other set of components in an arbitrary basis |ai. In fact, the expansion
ha|(t)i =

d3 x ha|xihx|(t)i

(1.148)

may be viewed as a re-expansion of a component of |(t)i in one basis, |ai, into


those of another basis, |xi, just as in (1.133).
In order to express all these transformation properties in a most compact notation, it has become customary to deal with an arbitrary physical state vector in a
basis-independent way and denote it by a ket vector |(t)i. This vector may be specified in any convenient basis by multiplying it with the corresponding completeness
relation
X
|aiha| = 1,
(1.149)
a

resulting in the expansion

|(t)i =

X
a

|aiha|(t)i.

(1.150)

This can be multiplied with any bra vector, say hb|, from the left to obtain the
expansion formula (1.135):
hb|(t)i =

X
a

hb|aiha|(t)i.

(1.151)

The continuum version of the completeness relation (1.142) reads


Z

d3 x |xihx| = 1,

(1.152)

and leads to the expansion


|(t)i =

d3 x |xihx|(t)i,

(1.153)

in which the wave function (x, t) = hx|(t)i plays the role of an xth component
of the state vector |(t)i in the local basis |xi. This, in turn, is the limit of the
discrete basis vectors |xn i,
1
(1.154)
|xi |xn i ,
3
with xn being the lattice points closest to x.
A vector can be described equally well in bra or in ket form. To apply the above
formalism consistently, we observe that the scalar products
ha|bi =
hb|ai =

d3 x f a (x) fb (x),
d3 x fb (x) f a (x)

(1.155)

24

1 Fundamentals

satisfy the identity

hb|ai ha|bi .

(1.156)

Therefore, when expanding a ket vector as


|(t)i =

|aiha|(t)i,

(1.157)

h(t)| =

h(t)|aiha|,

(1.158)

or a bra vector as
a

a multiplication of the first equation with the bra hx| and of the second with the ket
|xi produces equations which are complex-conjugate to each other.

1.4.4

Generalized Functions

Diracs bra-ket formalism is elegant and easy to handle. As far as the vectors |xi are
concerned there is, however, one inconsistency with some fundamental postulates of
quantum mechanics: When introducing state vectors, the norm was required to be
unity in order to permit a proper probability interpretation of single-particle states.
The limiting states |xi introduced above do not satisfy this requirement. In fact,
the scalar product between two different states hx| and |x i is
hx|x i

1
1
i =
hx
|x
nn ,
n
n
3
3

(1.159)

where xn and xn are the lattice points closest to x and x . For x 6= x , the states are
orthogonal. For x = x , on the other hand, the limit 0 is infinite, approached
in such a way that
X 1
3
(1.160)
nn = 1.
3
n
Therefore, the limiting state |xi is not a properly normalizable vector in the Hilbert
space. For the sake of elegance, it is useful to weaken the requirement of normalizability (1.100) by admitting the limiting states |xi to the physical Hilbert space.
In fact, one admits all states which can be obtained by a limiting sequence from
properly normalized state vectors.
The scalar product between states hx|x i is not a proper function. It is denoted
by the symbol (3) (x x ) and called Dirac -function:
hx|x i (3) (x x ).

(1.161)

The right-hand side vanishes everywhere, except in the infinitely small box of width
around x x . Thus the -function satisfies
(3) (x x ) = 0

for

x 6= x .

(1.162)

At x = x , it is so large that its volume integral is unity:


Z

d3 x (3) (x x ) = 1.

(1.163)

25

1.4 Diracs Bra-Ket Formalism

Obviously, there exists no proper function that can satisfy both requirements, (1.162)
and (1.163). Only the finite- approximation in (1.159) to the -function are proper
functions. In this respect, the scalar product hx|x i behaves just like the states |xi
themselves: Both are 0 -limits of properly defined mathematical objects.
Note that the integral Eq. (1.163) implies the following property of the function:
1 (3)
(3) (a(x x )) =
(x x ).
(1.164)
|a|
In one dimension, this leads to the more general relation
(f (x)) =

X
i

1
|f (xi )|

(x xi ),

(1.165)

where xi are the simple zeros of f (x).


In mathematics, one calls the -function a generalized function or a distribution.
It defines a linear functional of arbitrary smooth test functions f (x) which yields
its value at any desired place x:
[f ; x]

d3 x (3) (x x )f (x ) = f (x).

(1.166)

Test functions are arbitrarily often differentiable functions with a sufficiently fast
falloff at spatial infinity.
There exist a rich body of mathematical literature on distributions [4]. They
form a linear space. This space is restricted in an essential way in comparison with
ordinary functions: products of -functions or any other distributions remain undefined. However, in Chapter 10 of the textbook [1] it was found that physics forces
us to go beyond these rules. An important requirement of quantum mechanics is
coordinate invariance. If one imposes this condition upon the path integral formulation of quantum mechanics, we can derive an extension of the existing theory of
distributions, which specifies uniquely integrals over products of distributions.
In quantum mechanics, the role of the test functions is played by the wave packets
(x, t). By admitting the generalized states |xi to the Hilbert space, we also admit
the scalar products hx|x i to the space of wave functions, and thus all distributions,
although they are not normalizable.

1.4.5

Schr
odinger Equation in Dirac Notation

In terms of the Dirac bracket notation, the Schrodinger equation can be expressed
in a basis-independent way as an operator equation

, t)|(t)i = iht |(t)i,


H|(t)i
H(
p, x

(1.167)

to be supplemented by the following specifications of the canonical operators:


hx|
p ihx hx|,
hx|
x xhx|.

(1.168)
(1.169)

26

1 Fundamentals

Any matrix element can be obtained from these equations by multiplication from
the right with an arbitrary ket vector; for instance with the local basis vector |x i:
hx|
p|x i = ihx hx|x i = ihx (3) (x x ),

(1.170)

hx|
x|x i = xhx|x i = x (3) (x x ).

(1.171)

The original differential form of the Schrodinger equation (1.91) follows by multiplying the basis-independent Schrodinger equation (1.167) with the bra vector hx|
from the left:
, t)|(t)i = H(ihx , x, t)hx|(t)i
hx|H(
p, x
= iht hx|(t)i.

(1.172)

and x
are Hermitian matrices in any basis,
Obviously, p
ha|
p|a i = ha |
p|ai ,

(1.173)

ha|
x|a i = ha |
x|ai ,

(1.174)

i = ha |H|ai
,
ha|H|a

(1.175)

and so is the Hamiltonian


as long as it has the form (1.105).
The most general basis-independent operator that can be constructed in the
, x
, t,
generalized Hilbert space spanned by the states |xi is some function of p
O(
, t).
O(t)
p, x

(1.176)

In general, such an operator is called Hermitian if all its matrix elements have
this property. In the basis-independent Dirac notation, the definition (1.101) of a
(t) implies the equality of the matrix elements
Hermitian-adjoint operator O

(t)|a i ha |O(t)|ai

ha|O
.

(1.177)

Thus we can rephrase Eqs. (1.173)(1.175) in the basis-independent form


= p
,
p
= x
,
x

(1.178)

= H
.
H
The stationary states in Eq. (1.98) have a Dirac ket representation |En i, and satisfy
the time-independent operator equation
n i = En |En i.
H|E

(1.179)

27

1.4 Diracs Bra-Ket Formalism

1.4.6

Momentum States

. Its eigenstates are given by the eigenvalue


Let us now look at the momentum p
equation
|pi = p|pi.
p
(1.180)

By multiplying this with hx| from the left and using (1.168), we find the differential
equation
hx|
p|pi = ihx hx|pi = phx|pi.
(1.181)

The solution is

hx|pi eipx/h .

(1.182)

Up to a normalization factor, this is just a plane wave introduced before in Eq. (1.75)
to describe free particles of momentum p.
In order for the states |pi to have a finite norm, the system must be confined
to a finite volume, say a cubic box of length L and volume L3 . Assuming periodic
boundary conditions, the momenta are discrete with values
pm =

2h
(m1 , m2 , m3 ),
L

mi = 0, 1, 2, . . . .

(1.183)

Then we adjust the factor in front of exp (ipm x/h) to achieve unit normalization
1
hx|pm i = exp (ipm x/h) ,
L3

(1.184)

and the discrete states |pm i satisfy


Z

d3 x |hx|pm i|2 = 1.

(1.185)

The states |pm i are complete:


X
m

|pm ihpm | = 1.

(1.186)

We may use this relation and the matrix elements hx|pm i to expand any wave
function within the box as
(x, t) = hx|(t)i =

X
m

hx|pm ihpm |(t)i.

(1.187)

If the box is very large, the sum over the discrete momenta pm can be approximated
by an integral over the momentum space [7].
X
m

d3 pL3
.
(2h)3

(1.188)

In this limit, the states |pm i may be used to define a continuum of basis vectors
with an improper normalization

(1.189)
|pi L3 |pm i,

28

1 Fundamentals

in the same way as |xn i was used in (1.154) to define |xi (1/ 3 )|xn i. The
momentum states |pi satisfy the orthogonality relation
hp|p i = (2h)3 (3) (p p ),

(1.190)

with (3) (pp ) being again the Dirac -function. Their completeness relation reads
Z

d3 p
|pihp| = 1,
(2h)3

(1.191)

such that the expansion (1.187) becomes


Z

(x, t) =

d3 p
hx|pihp|(t)i,
(2h)3

(1.192)

with the momentum eigenfunctions


hx|pi = eipx/h .

(1.193)

This coincides precisely with the Fourier decomposition introduced above in the
description of a general particle wave (x, t) in (1.83), (1.84), with the identification
hp|(t)i = f (p)eiEp t/h .

(1.194)

The frequent appearance of factors 2h with -functions and integration measures in momentum space makes it convenient to define the modified -functions
and integration measures
(D)
- (p) (2h)D (D) (p),

d-D p

dD p
,
(2h)D

(1.195)

the latter in analogy with h


h/2. Then we may write orthogonality and completeness relations as
(3)
hp|p i = - (p p ),
(1.196)
and

d-3 p|pihp| = 1,

(1.197)

The bra-ket formalism accommodates naturally the technique of Fourier transforms. The Fourier inversion formula is found by simply inserting into hp|(t)i a
R
completeness relation d3 x|xihx| = 1 which yields
hp|(t)i =
=

d3 x hp|xihx|(t)i
3

ipx/
h

d xe

(1.198)

(x, t).

The amplitudes hp|(t)i are referred to as momentum space wave functions.

29

1.4 Diracs Bra-Ket Formalism

By inserting the completeness relation


Z

d3 x|xihx| = 1

(1.199)

between the momentum states on the left-hand side of the orthogonality relation
(1.190), we obtain the Fourier representation of the -function (1.190):
hp|p i =
=

1.4.7

d3 x hp|xihx|pi

(1.200)

d3 x ei(pp )x/h = (2h)3 (3) (p p ).

Incompleteness and Poissons Summation Formula

For many physical applications it is important to find out what happens to the
completeness relation (1.152) if one restricts the integral to a subset of positions.
Most relevant will be the one-dimensional integral,
Z

dx |xihx| = 1,

(1.201)

restricted to a sum over equally spaced points xn = na:


N
X

n=N

|xn ihxn | = 1.

(1.202)

Taking this sum between momentum eigenstates |pi, we obtain


N
X

n=N

hp|xn ihxn |p i =

N
X

ei(pp )na/h .

(1.203)

n=N

For N we can perform the sum with the help of Poissons summation formula:7

n=

e2in =

m=

( m).

(1.204)

Identifying with (p p )a/2h, we find using Eq. (1.164):

X
2h
2hm
a(pp )
. (1.205)
m =
pp
hp|xn ihxn |p i =

2h
a
m= a
n=
m=

In order to prove the Poisson formula (1.204), we observe that the sum s()
side is periodic in with a unit period and has
m ( m) on the right-hand
P
2in
the Fourier series s() =
. The Fourier coefficients are given by
n= sn e

The proof of this formula is taken from Section 2.15 of the textbook in Ref. [1] on p. 80.

30

1 Fundamentals

2in in Poissons summation formula. In the


Figure 1.2 Relevant function N
n=N e
limit N , is squeezed to the integer values.

R 1/2

sn = 1/2 d s()e2in 1. These are precisely the Fourier coefficients on the


left-hand side.
For a finite N, the sum over n on the left-hand side of (1.204) yields
N
X

e2in = 1 + e2i + e22i + . . . + eN 2i + c.c.

n=N

1 e2i(N +1)
= 1 +
+ c.c.
1 e2i
= 1+

(1.206)

e2i e2i(N +1)


sin (2N + 1)
+ c.c. =
.
2i
1e
sin

This function is well known in wave optics (see Fig. 2.4). It determines the diffraction pattern of light behind a grating with 2N + 1 slits. It has large peaks at
= 0, 1, 2, 3, . . . and N 1 small maxima between each pair of neighboring peaks, at = (1 + 4k)/2(2N + 1) for k = 1, . . . , N 1. There are zeros at
= (1 + 2k)/(2N + 1) for k = 1, . . . , N 1.
Inserting = (p p )a/2h into (1.206), we obtain
N
X

n=N

hp|xn ihxn |pi =

sin (p p )a(2N + 1)/2h


.
sin (p p )a/2h

(1.207)

Let us see how the right-hand side of (1.206) turns into the right-hand side of
(1.204) in the limit N . In this limit, the area under each large peak can
be calculated by an integral over the central large peak plus a number n of small
maxima next to it:
Z n/2N
Z n/2N
sin (2N + 1)
sin 2N cos +cos 2N sin
d
d
=
.
sin
sin
n/2N
n/2N

31

1.5 Observables

(1.208)
Keeping a fixed ratio n/N 1, we may replace in the integrand sin by and
cos by 1. Then the integral becomes, for N at fixed n/N,
sin (2N + 1) N Z n/2N
sin 2N Z n/2N
d

d cos 2N
+
sin

n/2N
n/2N
n/2N
Z n
Z n
N
N 1
sin x
1
dx
dx cos x
1,
(1.209)

+
n
x
2N n

n/2N

where we have used the integral formula


Z

dx

sin x
= .
x

(1.210)

In the limit N , we find indeed (1.204) and thus (1.212).


There exists another useful way of expressing Poissons formula. Consider an
arbitrary smooth function f () which possesses a convergent sum

f (m).

(1.211)

m=

Then Poissons formula (1.204) implies that the sum can be rewritten as an integral
and an auxiliary sum:

m=

f (m) =

e2in f ().

(1.212)

n=

The auxiliary sum over n squeezes to the integer numbers.

1.5

Observables

Changes of basis vectors are an important tool in analyzing the physically observable
content of a wave vector. Let A = A(p, x) be an arbitrary time-independent real
function of the phase space variables p and x. Important examples for such an
A are p and x themselves, the Hamiltonian H(p, x), and the angular momentum
L = x p. Quantum-mechanically, there will be an observable operator associated
with each such quantity. It is obtained by simply replacing the variables p and x in
and x
:
A by the corresponding operators p
).
A A(
p, x

(1.213)

This replacement rule is the extension of the correspondence principle for the Hamiltonian operator (1.92) to more general functions in phase space, converting them
into observable operators. It must be assumed that the replacement leads to a
unique Hermitian operator, i.e., that there is no ordering problem of the type discussed in context with the Hamiltonian (1.105).8 If there are ambiguities, the naive
8

Note that this is true for the angular momentum

= x p.

32

1 Fundamentals

correspondence principle is insufficient to determine the observable operator. Then


the correct ordering must be decided by comparison with experiment, unless it can
be specified by means of simple geometric principles. The probelem is solved in the
textbook [1].
Once an observable operator A is Hermitian, it has the useful property that the
set of all eigenvectors |ai obtained by solving the equation
= a|ai
A|ai

(1.214)

can be used as a basis to span the Hilbert space. Among the eigenvectors, there is
always a choice of orthonormal vectors |ai fulfilling the completeness relation
X
a

|aiha| = 1.

(1.215)

The vectors |ai can be used to extract physical information concerning the observable A from arbitrary state vector |(t)i. For this we expand this vector in the
basis |ai:
X
|(t)i =
|aiha|(t)i.
(1.216)
a

The components
ha|(t)i

(1.217)

yield the probability amplitude for measuring the eigenvalue a for the observable
quantity A.
The wave function (x, t) itself is an example of this interpretation. If we write
it as
(x, t) = hx|(t)i,
(1.218)
it gives the probability amplitude for measuring the eigenvalues x of the position
, i.e., |(x, t)|2 is the probability density in x-space.
operator x
The expectation value of the observable operator (1.213) in the state |(t)i is
defined as the matrix element

h(t)|A|(t)i

1.5.1

d3 xh(t)|xiA(ih, x)hx|(t)i.

(1.219)

Uncertainty Relation

We have seen before [see the discussion after (1.83), (1.84)] that the amplitudes in
real space and those in momentum space have widths inversely proportional to each
other, due to the properties of Fourier analysis. If a wave packet is localized in real
space with a width x, its momentum space wave function has a width p given
by
x p h
.
(1.220)

33

1.5 Observables

From the Hilbert space point of view this uncertainty relation can be shown to be
and p
do not commute with each
a consequence of the fact that the operators x
other, but the components satisfy the canonical commutation rules
[
pi , xj ] = ihij ,
[
xi , xj ] = 0,
[
pi , pj ] = 0.

(1.221)

In general, if an observable operator A is measured to have a sharp value a in one


state, this state must be an eigenstate of A with an eigenvalue a:
= a|ai.
A|ai

(1.222)

This follows from the expansion


|(t)i =

X
a

|aiha|(t)i,

(1.223)

in which |ha|(t)i|2 is the probability to measure an arbitrary eigenvalue a. If this


probability is sharply focused at a specific value of a, the state necessarily coincides
with |ai.
we may ask under what circumstances
Given the set of all eigenstates |ai of A,

another observable, say B, can be measured sharply in each of these states. The

requirement implies that the states |ai are also eigenstates of B,

B|ai
= ba |ai,

(1.224)

with some a-dependent eigenvalue ba . If this is true for all |ai,


A|ai
= ba a|ai = aba |ai = AB|ai,

(1.225)

necessarily commute:
the operators A and B
B]
= 0.
[A,

(1.226)

Conversely, it can be shown that a vanishing commutator is also sufficient for


two observable operators to be simultaneously diagonalizable and thus to allow for
simultaneous sharp measurements.

1.5.2

Density Matrix and Wigner Function

An important object for calculating observable properties of a quantum-mechanical


system is the quantum mechanical density operator associated with a pure state
(t) |(t)ih(t)|,

(1.227)

34

1 Fundamentals

and the associated density matrix associated with a pure state


(x1 , x2 ; t) = hx1 |(t)ih(t)|x2i.

(1.228)

) can be calculated from the trace


The expectation value of any function f (x, p
)|(t)i = tr[f (x, p
)
h(t)|f (x, p
(t)] =

d3 xh(t)|xif (x, ih)hx|(t)i. (1.229)

If we decompose the states |(t)i into stationary eigenstates |En i of the Hamiltonian
[recall (1.179)], |(t)i = Pn |En ihEn |(t)i, then the density matrix has
operator H
the expansion
(t)

n,m

|En inm (t)hEm | =

n,m

|En ihEn |(t)ih(t)|Em ihEm |.

(1.230)

Wigner showed that the Fourier transform of the density matrix, the Wigner function
W (X, p; t)

d3 x ipx/h
e
(X + x/2, X x/2; t)
(2h)3

(1.231)

satisfies, for a single particle of mass M in a potential V (x), the Wigner-Liouville


equation


p
,
(1.232)
t + v X W (X, p; t) = Wt (X, p; t), v
M
where
Wt (X, p; t)

2
h

d3 q
W (X, p q; t)
(2h)3

d3 x V (X x/2)eiqx/h .

(1.233)

In the limit h
0, we may expand W (X, p q; t) in powers of q, and V (X x/2)
in powers of x, which we rewrite in front of the exponential eiqx/h as powers of
ihq . Then we perform the integral over x to obtain (2h)3 (3) (q), and perform
the integral over q to obtain the classical Liouville equation for the probability
density of the particle in phase space


t + v X W (X, p; t) = F (X)p W (X, p; t),

p
,
M

(1.234)

where F (X) X V (X) is the force associated with the potential V (X).

1.5.3

Generalization to Many Particles

All this development can be extended to systems of N distinguishable mass points


with Cartesian coordinates x1 , . . . , xN . If H(p , x , t) is the Hamiltonian, the
Schrodinger equation becomes
, t)|(t)i = iht |(t)i.
H(
p , x

(1.235)

35

1.6 Time Evolution Operator

We may introduce a complete local basis |x1 , . . . , xN i with the properties


hx1 , . . . , xN |x1 , . . . , xN i = (3) (x1 x1 ) (3) (xN xN ),
Z

and define

d3 x1 d3 xN |x1 , . . . , xN ihx1 , . . . , xN | = 1,
hx1 , . . . , xN |
p = ihx hx1 , . . . , xN |,
hx1 , . . . , xN |
x = x hx1 , . . . , xN |.

(1.236)

(1.237)

The Schrodinger equation for N particles (1.111) follows from (1.235) by multiplying
it from the left with the bra vectors hx1 , . . . , xN |. In the same way, all other formulas
given above can be generalized to N-body state vectors.

1.6

Time Evolution Operator

If the Hamiltonian operator possesses no explicit time dependence, the basisindependent Schrodinger equation (1.167) can be integrated to find the wave function
|(t)i at any time tb from the state at any other time ta

|(tb )i = ei(tb ta )H/h |(ta )i.

(1.238)

U (tb , ta ) = ei(tb ta )H/h

(1.239)

The operator
is called the time evolution operator . It satisfies the differential equation
b , ta ) = H
U(t
b , ta ).
ihtb U(t

(1.240)

Its inverse is obtained by interchanging the order of tb and ta :

U 1 (tb , ta ) ei(tb ta )H/h = U (ta , tb ).

(1.241)

As an exponential of i times a Hermitian operator, U is a unitary operator satisfying


U = U 1 .

(1.242)

Indeed,
(tb , ta ) = ei(tb ta )H /h
U

= ei(tb ta )H/h = U 1 (tb , ta ).

(1.243)

, t) depends explicitly on time, the integration of the Schrodinger equation


If H(
p, x
(1.167) is somewhat more involved. The solution may be found iteratively: For
tb > ta , the time interval is sliced into a large number N + 1 of small pieces of
thickness with (tb ta )/(N + 1), slicing once at each time tn = ta + n for

36

1 Fundamentals

n = 0, . . . , N + 1. We then use the Schrodinger equation (1.167) to relate the wave


function in each slice approximately to the previous one:


|(ta + )i


E
i Z ta +

dt H(t) (ta ) ,
h
ta

i Z ta +2

1
dt H(t)
|(ta + )i,
h
ta +

|(ta + 2)i
..
.

i
1
h

|(ta + (N + 1))i

ta +(N +1)

ta +N

(1.244)

dt H(t)
|(ta + N)i.

From the combination of these equations we extract the evolution operator as a


product
b , ta ) 1 i
U(t
h

tb

tN

dtN +1

) 1 i
H(t
N +1
h

t1

ta

dt1

) .
H(t
1

(1.245)

By multiplying out the product and going to the limit N we find the series
i
i Z tb

dt1 H(t1 ) +
U(tb , ta ) = 1
h
ta
h

i
+
h

3 Z

tb

ta

dt3

t3
ta

dt2

2 Z

t2

ta

tb

ta

dt2

t2

ta

)H(t
)
dt1 H(t
2
1
(1.246)

)H(t
)H(t
) + ... ,
dt1 H(t
3
2
1

known as the Neumann-Liouville expansion or Dyson series.


Note that each integral has the time arguments in the Hamilton operators ordered
causally: Operators with later times stand to left of those with earlier times. It is
useful to introduce a time-ordering operator which, when applied to an arbitrary
product of operators,
n (tn ) O
1 (t1 ),
O
(1.247)
reorders the times successively. More explicitly we define

i1 (ti1 ),
n (tn ) O
1 (t1 )) O
in (tin ) O
T (O

(1.248)

where tin , . . . , ti1 are the times tn , . . . , t1 relabeled in the causal order, so that
tin > tin1 > . . . > ti1 .

(1.249)

Any c-number factors in (1.248) can be pulled out in front of the T operator. With
this formal operator, the Neumann-Liouville expansion can be rewritten in a more
compact way. Take, for instance, the third term in (1.246)
Z

tb

ta

dt2

t2

ta

2 )H(t
1 ).
dt1 H(t

(1.250)

The integration covers the triangle above the diagonal in the square t1 , t2 [ta , tb ]
in the (t1 , t2 ) plane (see Fig. 1.2). By comparing this with the missing integral over

37

1.6 Time Evolution Operator

tb
t2

ta

ta

t1

tb

Figure 1.3 Illustration of time-ordering procedure in Eq. (1.250).

the lower triangle


Z

tb

dt2

ta

tb

2 )H(t
1)
dt1 H(t

t2

(1.251)

we see that the two expressions coincide except for the order of the operators. This
can be corrected with the use of a time-ordering operator T . The expression
Z

tb

ta

dt2

tb

t2

2 )H(t
1)
dt1 H(t

(1.252)

is equal to (1.250) since it may be rewritten as


Z

tb

ta

dt2

tb

1 )H(t
2)
dt1 H(t

t2

(1.253)

or, after interchanging the order of integration, as


Z

tb

ta

dt1

t1
ta

1 )H(t
2 ).
dt2 H(t

(1.254)

Apart from the dummy integration variables t2 t1 , this double integral coincides
with (1.250). Since the time arguments are properly ordered, (1.250) can trivially
be multiplied with the time-ordering operator. The conclusion of this discussion is
that (1.250) can alternatively be written as
tb
1 tb
2 )H(t
1 ).
T
dt2
dt1 H(t
(1.255)
2 ta
ta
On the right-hand side, the integrations now run over the full square in the t1 , t2 plane so that the two integrals can be factorized into

1
T
2

Z

tb

ta

dt H(t)

2

(1.256)

Similarly, we may rewrite the nth-order term of (1.246) as


Z tb
Z tb
1 Z tb
n )H(t
n1 ) H(t
1)
T
dtn
dtn1
dt1 H(t
n! ta
ta
ta

1
= T
n!

"Z

tb

ta

#n

dt H(t)
.

(1.257)

38

1 Fundamentals

b , ta ) has therefore the series expansion


The time evolution operator U(t
i
U (tb , ta ) = 1 T
h

tb

ta

+ 1 i
dt H(t)
2! h

1 i
+...+
n! h

n

Z

tb

ta

2

tb

Z

dt H(t)

ta

n

dt H(t)

2

(1.258)

+ ... .

The right-hand side of T contains simply the power series expansion of the exponential so that we can write
i
U (tb , ta ) = T exp
h

tb

ta

dt H(t)
.

(1.259)

does not depend on the time, the time-ordering operation is superfluous, the
If H
integral can be done trivially, and we recover the previous result (1.239).

b , ta ) by
Note that a small variation H(t)
of H(t)
changes U(t
(

i tb
i tb

) T exp i
U (tb , ta ) =
dt H(t)
H(t
dt T exp
h
ta
h
t
h

Z tb
i
b , t ) H(t
) U (t , ta ).
=
dt U(t
h
ta


ta

dt H(t)

(1.260)

A simple application for this relation is given in Appendix 1A.

1.7

Properties of the Time Evolution Operator

By construction, U (tb , ta ) has some important properties:


a) Fundamental composition law
If two time translations are performed successively, the corresponding operators U
are related by
b , t )U (t , ta ),
U (tb , ta ) = U(t
t (ta , tb ).
(1.261)

This composition law makes the operators U a representation of the abelian group
of time translations. For time-independent Hamiltonians with U (tb , ta ) given by
(1.239), the proof of (1.261) is trivial. In the general case (1.259), it follows from
the simple manipulation valid for tb > ta :
i
T exp
h

"

tb
t

dt T exp i
H(t)
h

ta

dt
H(t)

!#

i Z t
i Z tb

H(t) dt
H(t) dt exp
= T exp
h
t
h
ta


i
= T exp
h

tb

ta

dt .
H(t)

(1.262)

39

1.7 Properties of the Time Evolution Operator

b) Unitarity
The expression (1.259) for the time evolution operator U (tb , ta ) was derived only for
the causal (or retarded ) time arguments, i.e., for tb later than ta . We may, however,
define U (tb , ta ) also for the anticausal (or advanced ) case where tb lies before ta . To
be consistent with the above composition law (1.261), we must have
1
U (tb , ta ) U (ta , tb ) .

(1.263)

Indeed, when considering two states at successive times


|(ta )i = U (ta , tb )|(tb )i,

(1.264)

1 (ta , tb ):
the order of succession is inverted by multiplying both sides by U
|(tb )i = U (ta , tb )1 |(ta )i,

tb < ta .

(1.265)

The operator on the right-hand side is defined to be the time evolution operator
U (tb , ta ) from the later time ta to the earlier time tb .
If the Hamiltonian is independent of time, with the time evolution operator being

U (ta , tb ) = ei(ta tb )H/h ,

ta > tb ,

(1.266)

tb < ta .

(1.267)

the unitarity of the operator U (tb , ta ) is obvious:


b , ta )1 ,
U (tb , ta ) = U(t

Let us verify this property for a general time-dependent Hamiltonian. There, a


direct solution of the Schrodinger equation (1.167) for the state vector shows that
the operator U (tb , ta ) for tb < ta has a representation just like (1.259), except for a
reversed time order of its arguments. One writes this in the form [compare (1.259)]
U (tb , ta ) = T exp

i
h

tb

ta

dt ,
H(t)

(1.268)

where T denotes the time-antiordering operator, with an obvious definition analogous to (1.248), (1.249). This operator satisfies the relation
h

1 (t1 )O
2 (t2 )
T O

i

2 (t2 )O
1 (t1 ) ,
= T O


(1.269)

with an obvious generalization to the product of n operators. We can therefore


conclude right away that
(tb , ta ) = U (ta , tb ),
U

tb > ta .

(1.270)

With U (ta , tb ) U (tb , ta )1 , this proves the unitarity relation (1.267), in general.

40

1 Fundamentals

b , ta )
c) Schrodinger equation for U(t

Since the operator U (tb , ta ) rules the relation between arbitrary wave functions at
different times,
|(tb )i = U (tb , ta )|(ta )i,
(1.271)
b , ta ) satisfies the
the Schrodinger equation (1.235) implies that the operator U(t
corresponding equations
(t, ta ) = H
U(t,
ta ),
iht U
1
1
ta )

iht U(t,
= U (t, ta ) H,
with the initial condition

1.8

U (ta , ta ) = 1.

(1.272)
(1.273)
(1.274)

Heisenberg Picture of Quantum Mechanics

The unitary time evolution operator U (t, ta ) may be used to give a different formulation of quantum mechanics bearing the closest resemblance to classical mechanics.
This formulation, called the Heisenberg picture of quantum mechanics, is in a way
more closely related to classical mechanics than the Schrodinger formulation. Many
classical equations remain valid by simply replacing the canonical variables pi (t)
and qi (t) in phase space by Heisenberg operators, to be denoted by pHi (t), qHi (t).
Originally, Heisenberg postulated that they are matrices, but later it became clear
that these matrices had to be functional matrix elements of operators, whose indices
can be partly continuous. The classical equations hold for the Heisenberg operators
and as long as the canonical commutation rules (1.97) are respected at any given
time. In addition, qi (t) must be Cartesian coordinates. In this case we shall always
use the notation xi for the position variable, as in Section 1.4, rather than qi . The
corresponding Heisenberg operators are xHi (t). Suppressing the subscripts i, the
canonical equal-time commutation rules are
[
pH (t), x
H (t)] = ih,
[
pH (t), pH (t)] = 0,

(1.275)

[
xH (t), x
H (t)] = 0.
According to Heisenberg, classical equations involving Poisson brackets remain
valid if the Poisson brackets are replaced by i/h times the matrix commutators at
equal times. The canonical commutation relations (1.275) are a special case of this
rule, recalling the fundamental Poisson brackets (21.182). The Hamilton equations
of motion (1.23) turn into the Heisenberg equations
i
d
i h
pH (t) =
HH , pH (t) ,
dt
h

i
d
i h
xH (t) =
HH , xH (t) ,
dt
h

(1.276)

1.8 Heisenberg Picture of Quantum Mechanics

41

where
H H(
H
pH (t), xH (t), t)

(1.277)

is the Hamiltonian in the Heisenberg picture. Similarly, the equation of motion for
an arbitrary observable function O(pi (t), xi (t), t) derived in (21.179) goes over into
the matrix commutator equation for the Heisenberg operator
H (t) O(
O
pH (t), xH (t), t),

(1.278)

namely,
d
i

OH = [H
OH .
(1.279)
H , OH ] +
dt
h

t
These rules are referred to as Heisenbergs correspondence principle.
The relation between Schrodingers and Heisenbergs picture is supplied by the
be an arbitrary observable in the Schrodinger detime evolution operator. Let O
scription
O(
O(t)
p, x, t).
(1.280)
If the states |a (t)i are an arbitrary complete set of solutions of the Schrodinger

equation, where a runs through discrete and continuous indices, the operator O(t)
can be specified in terms of its functional matrix elements

Oab (t) ha (t)|O(t)|


b (t)i.

(1.281)

We can now use the unitary operator U (t, 0) to go to a new time-independent basis
|H a i, defined by
|a (t)i U (t, 0)|H a i.
(1.282)
Simultaneously, we transform the Schrodinger operators of the canonical coordinates
p and x into the time-dependent canonical Heisenberg operators pH (t) and xH (t) via
0),
pH (t) U (t, 0)1 p U(t,
xH (t) U (t, 0)1 x U (t, 0).

(1.283)
(1.284)

At the time t = 0, the Heisenberg operators pH (t) and xH (t) coincide with the timeindependent Schrodinger operators p and x, respectively. An arbitrary observable

O(t)
is transformed into the associated Heisenberg operator as
H (t) U (t, ta )1 O(
O
p, x, t)U (t, ta )
O (
pH (t), xH (t), t) .

(1.285)

The Heisenberg matrices OH (t)ab are then obtained from the Heisenberg operators
H (t) by sandwiching O
H (t) between the time-independent basis vectors |H a i:
O
H (t)|H b i.
OH (t)ab hH a |O

(1.286)

42

1 Fundamentals

Note that the time dependence of these matrix elements is now completely due to
the time dependence of the operators,
d
d
OH (t)ab hH a | O
H (t)|H b i.
dt
dt

(1.287)

This is in contrast to the Schrodinger representation (1.281), where the right-hand


side would have contained two more terms from the time dependence of the wave
functions. Due to the absence of such terms in (1.287) it is possible to study the
equation of motion of the Heisenberg matrices independently of the basis by considering directly the Heisenberg operators. It is straightforward to verify that they do
indeed satisfy the rules of Heisenbergs correspondence principle. Consider the time
H (t),
derivative of an arbitrary observable O
!

d 1
U (t, ta )
U (t, ta ) O(t)
dt
!
!
d

1
1

ta ) + U (t, ta )O(t)

O(t)
U(t,
U (t, ta ) ,
+ U (t, ta )
t
dt

d
OH (t) =
dt

which can be rearranged as


"

d 1
ta ) U 1 (t, ta )O(t)
U (t, ta )
U (t, ta ) U(t,
(1.288)
dt
!
h
i
d

1
1
1
(t, ta )O(t)
U(t,
ta ) U
(t, ta ) U (t, ta ) + U
(t, ta )

+ U
O(t)
U (t, ta ).
dt
t

Using (1.272), we obtain


!

i h 1 i 1
d
OH (t) =
O(t) U .
U H U , OH + U
dt
h

(1.289)

After inserting (1.285), we find the equation of motion for the Heisenberg operator:
i
d
i h

OH (t) =
O
HH , OH (t) +
dt
h

(t).

(1.290)

By sandwiching this equation between the complete time-independent basis states


|a i in the Hilbert space, it holds for the matrices and turns into the Heisenberg
equation of motion. For the phase space variables pH (t), xH (t) themselves, these
equations reduce, of course, to the Hamilton equations of motion (1.276).
Thus we have shown that Heisenbergs matrix quantum mechanics is completely
equivalent to Schrodingers quantum mechanics, and that the Heisenberg matrices
obey the same Hamilton equations as the classical observables.

1.9 Interaction Picture and Perturbation Expansion

1.9

43

Interaction Picture and Perturbation Expansion

For some physical systems, the Hamiltonian operator can be split into two contributions
=H
0 + V ,
H
(1.291)
0 is a so-called free Hamiltonian operator for which the Schrodinger equation
where H

H0 |(t)i = iht |(t)i can be solved, and V is an interaction potential which perturbs
these solutions slightly. In this case it is useful to describe the system in Diracs
interaction picture. We remove the time evolution of the unperturbed Schrodinger
solutions and define the states

|I (t)i eiH0 t/h |(t)i.

(1.292)

Their time evolution comes entirely from the interaction potential V . It is governed
by the time evolution operator
UI (tb , ta ) eiH0 tb /h eiHtb /h eiHta /h eiH0 ta /h ,
and reads

(1.293)

|I (tb )i = UI (tb , ta )|I (ta )i.

(1.294)

I (tb , ta ) = VI (tb )UI (tb , ta ),


ihtb U

(1.295)

If V = 0, the states |I (tb )i are time-independent and coincide with the Heisenberg
0.
states (1.282) of the operator H
I (tb , ta ) satisfies the equation of motion
The operator U

where

VI (t) eiH0 t/h V eiH0 t/h

(1.296)

is the potential in the interaction picture. This equation of motion can be turned
into an integral equation
i Z tb
UI (tb , ta ) = 1
dtVI (t)UI (t, ta ).
h
ta

(1.297)

Inserting Eq. (1.296), this reads


i Z tb

I (t, ta ).

dt eiH0 t/h V eiH0 t/h U


UI (tb , ta ) = 1
h
ta

(1.298)

This equation can be iterated to find a perturbation expansion for the operator
UI (tb , ta ) in powers of the interaction potential:
i Z tb

dt eiH0 t/h V eiH0 t/h


UI (tb , ta ) = 1
h
ta

 Z
Z t
i 2 tb


+
dt dt eiH0 t/h V eiH0 (tt )/h V eiH0 t /h + . . . .
h

ta
ta

(1.299)

44

1 Fundamentals

Inserting on the left-hand side the operator (1.293), this can also be rewritten as
eiH(tb ta )/h = eiH0 (tb ta )/h


i
h

2 Z

tb

ta

dt

ta

i
h

tb

ta

dt eiH0 (tb t)/h V eiH0 (tta )/h

dt eiH0 (tb t)/h V eiH0 (tt )/h V eiH0 (t ta )/h + . . . .

(1.300)

This expansion is seen to be the recursive solution of the integral equation


iH(tb ta )/
h

iH0 (tb ta )/
h

=e

tb

ta

dt eiH0 (tb t)/h V eiH(tta )/h .

(1.301)

Note that the lowest-order correction agrees with the previous formula (1.260)
A compact way of writing the expansion (1.300) is
iH(tb ta )/
h

iH0 (tb ta )/
h

=e

i
T exp
h

tb

0 (tb t)/
iH
h

dt e

ta

0 (tta )/
iH
h

Ve

. (1.302)

The right-hand exponential can be expanded with the help of Lies expansion formula
eiA BeiA = 1 i[A, B] +

i2
[A, [A, B]] + . . . ,
2!

(1.303)

and forms the basis of the Campbell-Baker-Hausdorff expansion to be derived in


Appendix 4A.
Equation (1.302) can be used as a basis for a perturbative formula for the energy of a interacting system. Let |E0 i nbe an eigenstate of the free Schrodinger
0 |E0 i = E0 |E0 i. If it is subjected for an infinite amount of time to the
equation H
time-independent interaction V , it will turn into an eigenstate |E i of the total
This has an energy E = E0 + E with9
Hamiltonian H.
iE(tb ta )/
h

1.10

i
= hE0 |T exp
h

tb

ta

0 (tb t)/
iH
h

dt e

0 (tta )/
iH
h

Ve

|E0 i. (1.304)

Time Evolution Amplitude

In the subsequent development, an important role will be played by the matrix


elements of the time evolution operator in the localized basis states,
(xb tb |xa ta ) hxb |U (tb , ta )|xa i.

(1.305)

They are referred to as time evolution amplitudes. The functional matrix (xb tb |xa ta )
is also called the propagator of the system. For a system with a time-independent
Hamiltonian operator where U (tb , ta ) is given by (1.266), the propagator is simply
b ta )/h]|xa i.
(xb tb |xa ta ) = hxb | exp[iH(t
9

See Eqs. (3.506) and (3.515) of the textbook [4]. .

(1.306)

45

1.10 Time Evolution Amplitude

Due to the operator equations (1.272), the propagator satisfies the Schrodinger
equation
[H(ihxb , xb , tb ) ihtb ] (xb tb |xa ta ) = 0.
(1.307)

In the quantum mechanics of nonrelativistic particles, only the propagators from


earlier to later times will be relevant. It is therefore customary to introduce the
so-called causal time evolution operator or retarded time evolution operator :
R

U (tb , ta )

U (tb , ta ),
0,

tb ta ,
tb < ta ,

(1.308)

and the associated causal time evolution amplitude or retarded time evolution amplitude
R (tb , ta )|xa i.
(xb tb |xa ta )R hxb |U
(1.309)
Since this differs from (1.305) only for tb < ta , and since all formulas in the subsequent text will be used only for tb > ta , we shall often omit the superscript R.
To abbreviate the case distinction in (1.308), it is convenient to use the Heaviside
function defined by

1 for t > 0,
(t)
(1.310)
0 for t 0,
and write
U R (tb , ta ) (tb ta )U (tb , ta ),

(xb tb |xa ta )R (tb ta )(xb tb |xa ta ). (1.311)

There exists also another Heaviside function which differs from (1.310) only by the
value at tb = ta :

1 for t 0,
R
(t)
(1.312)
0 for t < 0.
Both Heaviside functions have the property that their derivative yields Diracs
-function
t (t) = (t).
(1.313)
If it is not important which -function is used we shall ignore the superscript.
The retarded propagator satisfies the Schrodinger equation
h

H(ihxb , xb , tb )R ihtb (xb tb |xa ta )R = ih(tb ta ) (3) (xb xa ).

(1.314)

The nonzero right-hand side arises from the extra term


ih [tb (tb ta )] hxb tb |xa ta i = ih(tb ta )hxb tb |xa ta i = ih(tb ta )hxb ta |xa ta i
(1.315)
and the initial condition hxb ta |xa ta i = hxb |xa i, due to (1.274).
If the Hamiltonian does not depend on time, the propagator depends only on the
time difference t = tb ta . The retarded propagator vanishes for t < 0. Functions
f (t) with this property have a characteristic Fourier transform. The integral
f(E)

dt f (t)eiEt/h

(1.316)

46

1 Fundamentals

is an analytic function in the upper half of the complex energy plane. This analyticity
property is necessary and sufficient to produce a factor (t) when inverting the
Fourier transform via the energy integral
f (t)

dE
f (E)eiEt/h .
2h

(1.317)

For t < 0, the contour of integration may be closed by an infinite semicircle in the
upper half-plane at no extra cost. Since the contour encloses no singularities, it can
be contracted to a point, yielding f (t) = 0.
The Heaviside function (t) itself is the simplest retarded function, with a
Fourier representation containing just a single pole just below the origin of the
complex energy plane:
Z
dE
i
(t) =
eiEt ,
(1.318)
2
E
+
i

where is an infinitesimally small positive number. The integral representation is


undefined for t = 0 and there are, in fact, infinitely many possible definitions for the
Heaviside function depending on the value assigned to the function at the origin. A
special role is played by the average of the Heaviside functions (1.312) and (1.310),
which is equal to 1/2 at the origin:

(t)

for t > 0,
for t = 0,

0 for t < 0.
1
2

(1.319)

Usually, the difference in the value at the origin does not matter since the Heaviside
function appears only in integrals accompanied by some smooth function f (t). This
makes the Heaviside function a distribution with respect to smooth test functions

f (t) as defined in Eq. (1.166). All three distributions r (t), l (t), and (t)
define
the same linear functional of the test functions by the integral
[f ] =

dt (t t )f (t ),

(1.320)

and this is an element in the linear space of all distributions.


As indicated after Eq. (1.166), a consitent theory of path integrals specifies, in
addition, integrals over products of distribution and thus give rise to an important
t )
extension of the theory of distributions. In this, the Heaviside function (t
plays the main role.
While discussing the concept of distributions let us introduce, for later use, the
closely related distribution
t ) (t
t),
(t t ) (t t ) (t t) = (t

(1.321)

which is a step function jumping at the origin from 1 to 1 as follows:

1
0
(t t ) =

for
for
for

t > t ,
t = t ,
t < t .

(1.322)

47

1.11 Fixed-Energy Amplitude

1.11

Fixed-Energy Amplitude

The Fourier-transform of the retarded time evolution amplitude (1.309)


(xb |xa )E =

dtb eiE(tb ta )/h (xb tb |xb ta )R =

ta

dtb eiE(tb ta )/h (xb tb |xb ta ) (1.323)

is called the fixed-energy amplitude.


If the Hamiltonian does not depend on time, we insert here Eq. (1.306) and find
that the fixed-energy amplitudes are matrix elements

(xb |xa )E = hxb |R(E)|x


ai

(1.324)

of the so-called resolvent operator

R(E)
=

ih
,
+ i
E H

(1.325)

which is the Fourier transform of the retarded time evolution operator (1.308):

R(E)
=

dtb eiE(tb ta )/h U R (tb , ta ) =

ta

dtb eiE(tb ta )/h U (tb , ta ).

(1.326)

Let us suppose that the time-independent Schrodinger equation is completely


solved, i.e., that one knows all solutions |n i of the equation
n i = En |n i.
H|

(1.327)

These satisfy the completeness relation


X
n

|n ihn | = 1,

(1.328)

which can be inserted on the right-hand side of (1.306) between the Dirac brackets
leading to the spectral representation
(xb tb |xa ta ) =

X
n

n (xb )n (xa ) exp [iEn (tb ta )/h] ,

(1.329)

with
n (x) = hx|n i

(1.330)

being the wave functions associated with the eigenstates |n i. Applying the Fourier
transform (1.323), we obtain
(xb |xa )E =

X
n

n (xb )n (xa )Rn (E) =

X
n

n (xb )n (xa )

ih
.
E En + i

(1.331)

The matrix elements of the resolvent operator


(xb |xa )E = hxb |

1
|x i
+ i a
EH

(1.332)

48

1 Fundamentals

are the Green functions of the Schrodinger equation, since they satisfy
E)(x|x )E = ih (3) (x x ).
(H

(1.333)

as G(E).

For this reason we may also denote the resolvent operator ih/(E H)
The fixed-energy amplitude (1.323) contains as much information on the system
as the time evolution amplitude, which is recovered from it by the inverse Fourier
transformation
Z
dE iE(tb ta )/h
e
(xb |xa )E .
(1.334)
(xb ta |xa ta ) =
h
2
The small i-shift in the energy E in (1.331) may be thought of as being attached
to each of the energies En , which are thus placed by an infinitesimal piece below the
real energy axis. Then the exponential behavior of the wave functions is slightly
damped, going to zero at infinite time:
ei(En i)t/h 0.

(1.335)

This so-called i-prescription ensures the causality of the Fourier representation


(11.6). When doing the Fourier integral (11.6), the exponential eiE(tb ta )/h makes it
always possible to close the integration contour along the energy axis by an infinite
semicircle in the complex energy plane, which lies in the upper half-plane for tb < ta
and in the lower half-plane for tb > ta . The i-prescription guarantees that for
tb < ta , there is no pole inside the closed contour making the propagator vanish. For
tb > ta , on the other hand, the poles in the lower half-plane give, via Cauchys residue
theorem, the spectral representation (1.329) of the propagator. An i-prescription
will appear in another context in Section 7.1.3.
If the eigenstates are nondegenerate, the residues at the poles of (1.331) render
directly the products of eigenfunctions (barring degeneracies which must be discussed separately). For a system with a continuum of energy eigenvalues, there is
a cut in the complex energy plane which may be thought of as a closely spaced sequence of poles. In general, the wave functions are recovered from the discontinuity
of the amplitudes (xb |xa )E across the cut, using the formula
disc

ih
E En

ih
ih

= 2h(E En ).
E En + i E En i

(1.336)

Here we have employed the relation10 , valid inside integrals over E:


P
1
=
i(E En ),
E En i
E En

(1.337)

where P indicates that the principal value of the integral has to be taken.
10

This is often referred to as Sochockis formula. It is the beginning of an expansion in powers


of > 0: 1/(x i) = P/x i(x) + [ (x) idx P/x] + O( 2 ).

49

1.12 Free-Particle Amplitudes

The energy integral over the discontinuity of the fixed-energy amplitude (1.331)
(xb |xa )E reproduces the completeness relation (1.328) taken between the local states
hxb | and |xa i,
Z

X
dE
disc (xb |xa )E =
n (xb )n (xa ) = hxb |xa i = (D) (xb xa ).
2h
n

(1.338)

The completeness relation reflects the following property of the resolvent operator:
Z

dE

disc R(E)
= 1.
2h

(1.339)

In general, the system possesses also a continuous spectrum, in which case the
completeness relation contains a spectral integral and (1.328) has the form
X
n

|n ihn | +

d | ih | = 1.

(1.340)

The continuum causes a branch cut along in the complex energy plane, and (1.338)
includes an integral over the discontinuity along the cut. The cut will mostly be
omitted, for brevity.

1.12

Free-Particle Amplitudes

=p
2 /2M, the spectrum is conFor a free particle with a Hamiltonian operator H
tinuous. The eigenfunctions are (1.193) with energies E(p) = p2 /2M. Inserting
the completeness relation (1.191) into Eq. (1.306), we obtain for the time evolution
amplitude of a free particle the Fourier representation
(xb tb |xa ta ) =

p2
i
dD p
p(x

x
)

exp
(tb ta )
b
a
(2h)D
h

2M
(

"

#)

(1.341)

The momentum integrals can easily be done. First we perform a quadratic completion in the exponent and rewrite it as
2

M (xb xa )2
(tb ta )
.
2 tb ta
(1.342)

Then we replace the integration variables by the shifted momenta p = p


(xb xa )/(tb ta )M , and the amplitude (1.341) becomes
1 2
1 xb xa
1
p(xb xa )
p
p (tb ta ) =
2M
2M
M tb ta


i M (xb xa )2
,
(xb tb |xa ta ) = F (tb ta ) exp
h
2 tb ta
#

(1.343)

d D p
i p 2
exp

(tb ta ) .
(2h)D
h
2M

(1.344)

"

where F (tb ta ) is the integral over the shifted momenta


F (tb ta )

50

1 Fundamentals

This can be performed using the Fresnel integral formula


Z

dp
1
a
exp i p2 = q
2
2
|a|


i,
1/ i,

a > 0,
a < 0.

(1.345)

Here the square root i denotes the phase factor ei/4 : This follows from the Gauss
formula


Z

1
dp
exp p2 = ,
Re > 0,
(1.346)
2

2
by continuing analytically from positive values into the right complex half-plane.
As long as Re > 0, this is straightforward. On the boundaries, i.e., on the positive
and negative imaginary axes, one has to be careful. At = ia + with a >
0 and
<
infinitesimal > 0, the integral is certainly convergent yielding (1.345). But the
integral also converges for = 0, as can easily be seen by substituting x2 = z. See
Appendix 1B.
Note that differentiation of Eq. (1.346) with respect to yields the more general
Gaussian integral formula
Z

dp
1 (2n 1)!!

p2n exp p2 =
2

n
2


Re > 0,

(1.347)

where (2n 1)!! is defined as the product (2n 1) (2n 3) 1. For odd powers
p2n+1 , the integral vanishes. In the Fresnel formula (1.345), an extra integrand p2n
produces a factor (i/a)n .
Since the Fresnel formula is a special analytically continued case of the Gauss formula, we shall in the sequel always speak of Gaussian integrations and use Fresnels
name only if the imaginary nature of the quadratic exponent is to be emphasized.
Applying this formula to (1.344), we obtain
1
F (tb ta ) = q
D,
2ih(tb ta )/M

(1.348)

so that the full time evolution amplitude of a free massive point particle is
1

(xb tb |xa ta ) = q
D
2ih(tb ta )/M

i M (xb xa )2
exp
.
h
2 tb ta
"

(1.349)

In the limit tb ta , the left-hand side becomes the scalar product hxb |xa i =
(D) (xb xa ), implying the following limiting formula for the -function

(D)

(xb xa ) =

lim

tb ta 0

1
q

2ih(tb ta )/M

i M (xb xa )2
.
exp
h
2 tb ta
"

(1.350)

51

1.12 Free-Particle Amplitudes

Inserting Eq. (1.343) into (1.323), we have for the fixed-energy amplitude the
integral representation
p2
i
dD p
p(x

x
)
+
(t

t
)
E

.
exp
(xb |xa )E =
d(tb ta )
b
a
b
a
(2h)D
h

2M
0
(1.351)
Performing the time integration yields
Z

(xb |xa )E =

"

!#)

dD p
ih
exp [ip(xb xa )]
,
D
2
(2h)
E p /2M + i

(1.352)

where we have inserted a damping factor e(tb ta ) into the integral to ensure convergence at large tb ta . For a more explicit result it is more convenient to calculate
the Fourier transform (1.349):
(xb |xa )E =

d(tb ta ) q
D
2ih(tb ta )/M

i
M (xb xa )2
exp
E(tb ta ) +
h

2 tb ta
(

"

#)

(1.353)

For E < 0, we set

2ME/h2 ,

(1.354)

and perform the integral with the help of the formula11


Z

1 it+i/t

dt t

=2

!/2

ei/2 K (2 ),

(1.355)

where K (z) is the modified Bessel function which satifies K (z) = K (z)12 The
result is
2M D2 KD/21 (R)
(xb |xa )E = i
,
(1.356)
h
(2)D/2 (R)D/21
where R |xb xa |. The simplest modified Bessel function is13
K1/2 (z) = K1/2 (z) =

z
e ,
2z

(1.357)

so that we find for D = 1, 2, 3, the amplitudes


i
11

M 1 R
e
,
h

M1
K0 (R),
h

M 1 R
e
.
h
2R

(1.358)

I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series, and Products, Academic Press,
New York, 1980, Formulas 3.471.10, 3.471.11, and 8.432.6
12
ibid., Formula 8.486.16
13
M. Abramowitz and I. Stegun, Handbook of Mathematical Functions, Dover, New York, 1965,
Formula 10.2.17.

52

1 Fundamentals

At R = 0, the amplitude (11.19) is finite for all D 2, where we can use the
small-argument behavior of the associated Bessel function14
K (z) = K (z)

 

z
1
()
2
2

for Re > 0,

(1.359)

to obtain
2M D2
(1 D/2).
(x|x)E = i
h
(4)D/2

(1.360)

This result can be continued analytically to D > 2, which is often of interest.


For E > 0 we set
q
(1.361)
k 2ME/h2

and use the formula15


Z

1 it+i/t

dtt

= i

!/2

(1)

ei/2 H (2 ),

(1.362)

where H(1) (z) is the Hankel function, to find


(xb |xa )E =

M k D2 HD/21 (kR)
.
h
(2)D/2 (kR)D/21

(1.363)

The relation16
K (iz) =

i/2 (1)
ie
H (z)
2

(1.364)

connects the two formulas with each other when continuing the energy from negative
to positive values, which replaces by ei/2 k = ik.
For large distances, the asymptotic behavior17
K (z)

z
e ,
2z

H(1) (z)

2 i(z/2/4)
e
z

(1.365)

shows that the fixed-energy amplitude behaves for E < 0 like


(xb |xa )E i

M D2
1
1

eR/h ,
(D1)/2
(D1)/2
h

(2)
(R)

(1.366)

and for E > 0 like


(xb |xa )E

M D2
1
1
k
eikR/h .
(D1)/2
(D1)/2
h

(2i)
(kR)

For D = 1 and 3, these asymptotic expressions hold for all R.


14

ibid.,
ibid.,
16
ibid.,
17
ibid.,

15

Formula 9.6.9.
Formulas 3.471.11 and 8.421.7.
Formula 8.407.1.
Formulas 8.451.6 and 8.451.3.

(1.367)

1.13 Quantum Mechanics of General Lagrangian Systems

1.13

53

Quantum Mechanics of General Lagrangian Systems

An extension of the quantum-mechanical formalism to systems described by a set


of completely general Lagrange coordinates q1 , . . . , qN is not straightforward. Only
in the special case of qi (i = 1, . . . , N) being merely a curvilinear reparametrization
of a D-dimensional Euclidean space are the above correspondence rules sufficient
to quantize the system. Then N = D and a variable change from xi to qj in the
Schrodinger equation leads to the correct quantum mechanics. It will be useful to
label the curvilinear coordinates by Greek superscripts and write q instead of qj .
This will help when we write all ensuing equations in a form that is manifestly
covariant under coordinate transformations. In the original definition of generalized
coordinates in Eq. (1.1), this was unnecessary since transformation properties were
ignored. For the Cartesian coordinates we shall use Latin indices alternatively as
sub- or superscripts. The coordinate transformation xi = xi (q ) implies the relation
between the derivatives /q and i /xi :
= ei (q)i ,

(1.368)

ei (q) xi (q)

(1.369)

with the transformation matrix

called basis D-ad (in 3 dimensions triad, in 4 dimensions tetrad, etc.). Let
ei (q) = q /xi be the inverse matrix (assuming it exists) called the reciprocal
D-ad , satisfying with ei the orthogonality and completeness relations
ei ei = ,

ei ej = i j .

(1.370)

Then, (1.368) is inverted to


i = ei (q)

(1.371)

and yields the curvilinear transform of the Cartesian quantum-mechanical momentum operators
pi = ihi = ihei (q) .
(1.372)
The free-particle Hamiltonian operator
h
2 2
1 2

=
x
p
H0 = T =
2M
2M

(1.373)

goes over into

2
0 = h
H
,
2M
where is the Laplacian expressed in curvilinear coordinates:
= i2 = ei ei
= ei ei + (ei ei ) .

(1.374)

(1.375)

54

1 Fundamentals

At this point one introduces the metric tensor


g (q) ei (q)ei (q),

(1.376)

g (q) = ei (q)ei (q),

(1.377)

its inverse
defined by g g = , and the so-called affine connection
(q) = ei (q) ei (q) = ei (q) ei (q).

(1.378)

Then the Laplacian takes the form


= g (q) (q) ,

(1.379)

with being defined as the contraction


g .

(1.380)

The reason why (1.376) is called a metric tensor is obvious: An infinitesimal square
distance between two points in the original Cartesian coordinates
ds2 dx2

(1.381)

becomes in curvilinear coordinates


ds2 =

x x
dq dq = g (q)dq dq .
q q

(1.382)

The infinitesimal volume element dD x is given by


dD x =

g dD q,

(1.383)

where
g(q) det (g (q))

(1.384)

is the determinant of the metric tensor. Using this determinant, we form the quantity
1
g 1/2 ( g 1/2 ) = g ( g )
2

(1.385)

and see that it is equal to the once-contracted connection


= .

(1.386)

With the inverse metric (1.377) we have furthermore


= g .

(1.387)

55

1.13 Quantum Mechanics of General Lagrangian Systems

We now take advantage of the fact that the derivatives , applied to the coordinate transformation xi (q) commute causing to be symmetric in , i.e.,
= and hence = . Together with (1.385) we find the rotation
1

= ( g g),
g

(1.388)

which allows the Laplace operator to be rewritten in the more compact form
1

= g g .
g

(1.389)

This expression is called the Laplace-Beltrami operator .


Thus we have shown that for a Hamiltonian in a Euclidean space
H(
p, x) =

1 2
+ V (x),
p
2M

(1.390)

the Schrodinger equation in curvilinear coordinates becomes


h
2

H(q,
t)
+ V (q) (q, t) = iht (q, t),
2M
"

(1.391)

where
V (q) is short for V (x(q)). The scalar product of two wave functions
R D
d x2 (x, t)1 (x, t), which determines the transition amplitudes of the system,
transforms into
Z

dD q g 2 (q, t)1 (q, t).


(1.392)
It is important to realize that this Schrodinger equation would not be obtained
by a straightforward application of the canonical formalism to the coordinatetransformed version of the Cartesian Lagrangian
=
L(x, x)

M 2
x V (x).
2

(1.393)

With the velocities transforming as


x i = ei (q)q ,

(1.394)

the Lagrangian becomes


L(q, q)
=

M
g (q)q q V (q).
2

(1.395)

Up to a factor M, the metric is equal to the Hessian metric of the system, which
depends here only on q [recall (1.12)]:
H (q) = Mg (q).

(1.396)

56

1 Fundamentals

The canonical momenta are


p

L
= Mg q .

(1.397)

The associated quantum-mechanical momentum operators p have to be Hermitian


in the scalar product (1.392) and must satisfy the canonical commutation rules
(1.275):
[
p , q ] = ih ,
[
q , q ] = 0,
[
p , p ] = 0.

(1.398)

An obvious solution is
p = ihg 1/4 g 1/4 ,

q = q .

(1.399)

The commutation rules are true for ihg z g z with any power z, but only z = 1/4
produces a Hermitian momentum operator:
Z

dq g

2 (q, t)[ihg 1/4 g 1/4 1 (q, t)]

d3 q g 1/4 2 (q, t)[ih g 1/4 1 (q, t)]

d3 q g [ihg 1/4 g 1/4 2 (q, t)] 1 (q, t),

(1.400)

as is easily verified by partial integration.


In terms of the quantity (1.385), this can also be rewritten as
p = ih( + 21 ).

(1.401)

Consider now the classical Hamiltonian associated with the Lagrangian (1.395),
which by (1.397) is simply
H = p q L =

1
g (q)p p + V (q).
2M

(1.402)

When trying to turn this expression into a Hamiltonian operator, we encounter the
operator-ordering problem discussed in connection with Eq. (1.105). The correspondence principle requires replacing the momenta p by the momentum operators p ,
but it does not specify the position of these operators with respect to the coordinates q contained in the inverse metric g (q). An important constraint is provided
by the required Hermiticity of the Hamiltonian operator, but this is not sufficient
for a unique specification. We may, for instance, define the canonical Hamiltonian
operator as
can 1 p g (q)
H
p + V (q),
(1.403)
2M
in which the momentum operators have been arranged symmetrically around the
inverse metric to achieve Hermiticity. This operator, however, is not equal to the

57

1.13 Quantum Mechanics of General Lagrangian Systems

correct Schrodinger operator in (1.391). The kinetic term contains what we may
call the canonical Laplacian
can = ( + 21 ) g (q) ( + 21 ).

(1.404)

It differs from the Laplace-Beltrami operator (1.389) in (1.391) by


can = 12 (g ) 41 g .

(1.405)

The correct Hamiltonian operator could be obtained by suitably distributing pairs of


dummy factors of g 1/4 and g 1/4 symmetrically between the canonical operators [8]:
= 1 g 1/4 p g 1/4 g (q)g 1/4 p g 1/4 + V (q).
H
2M

(1.406)

This operator has the same classical limit (1.402) as (1.403). Unfortunately, the
correspondence principle does not specify how the classical factors have to be ordered
before being replaced by operators.
The simplest system exhibiting the breakdown of the canonical quantization rules
is a free particle in a plane described by radial coordinates q 1 = r, q 2 = :
x1 = r cos , x2 = r sin .

(1.407)

Since the infinitesimal square distance is ds2 = dr 2 + r 2 d2 , the metric reads


1 0
0 r2

g =

(1.408)

It has a determinant
g = r2
and an inverse
g

1 0
0 r 2

(1.409)
!

(1.410)

1
1
= r rr + 2 2 .
r
r

(1.411)

The Laplace-Beltrami operator becomes

The canonical Laplacian, on the other hand, reads


1 2

r2
1
1
1
= r 2 + r 2 + 2 2 .
r
4r
r

can = (r + 1/2r)2 +

(1.412)

The discrepancy (1.405) is therefore


can =

1
.
4r 2

(1.413)

58

1 Fundamentals

Note that this discrepancy arises even though there is no apparent ordering problem
in the naively quantized canonical expression p g (q) p in (1.412). Only the need to
introduce dummy g 1/4 - and g 1/4 -factors creates such problems, and a specification
of the order is required to obtain the correct result.
If the Lagrangian coordinates qi do not merely reparametrize a Euclidean space
but specify the points of a general geometry, we cannot proceed as above and derive the Laplace-Beltrami operator by a coordinate transformation of a Cartesian
Laplacian. With the canonical quantization rules being unreliable in curvilinear
coordinates there are, at first sight, severe difficulties in quantizing such a system.
This is why the literature contains many proposals for handling this problem [9].
Fortunately, a large class of non-Cartesian systems allows for a unique quantummechanical description on completely different grounds. These systems have the
common property that their Hamiltonian can be expressed in terms of the generators of a group of motion in the general coordinate frame. For symmetry reasons,
the correspondence principle must then be imposed not on the Poisson brackets of
the canonical variables p and q, but on those of the group generators and the coordinates. The brackets containing two group generators specify the structure of the
group, those containing a generator and a coordinate specify the defining representation of the group in configuration space. The replacement of these brackets by
commutation rules constitutes the proper generalization of the canonical quantization from Cartesian to non-Cartesian coordinates. It is called group quantization.
The replacement rule will be referred to as the group correspondence principle. The
canonical commutation rules in Euclidean space may be viewed as a special case
of the commutation rules between group generators, i.e., of the Lie algebra of the
group. In a Cartesian coordinate frame, the group of motion is the Euclidean group
containing translations and rotations. The generators of translations and rotations
are the momenta and the angular momenta, respectively. According to the group
correspondence principle, the Poisson brackets between the generators and the coordinates are to be replaced by commutation rules. Thus, in a Euclidean space, the
commutation rules between group generators and coordinates lead to the canonical quantization rules, and this appears to be the deeper reason why the canonical
rules are correct. In systems whose energy depends on generators of the group of
motion other than those of translations, for instance on the angular momenta, the
commutators between the generators have to be used for quantization rather than
the canonical commutators between positions and momenta.
The prime examples for such systems are a particle on the surface of a sphere or
a spinning top whose quantization will now be discussed.

1.14

Particle on the Surface of a Sphere

For a particle moving on the surface of a sphere of radius r with coordinates


x1 = r sin cos , x2 = r sin sin , x3 = r cos ,

(1.414)

59

1.14 Particle on the Surface of a Sphere

the Lagrangian reads


L=

Mr 2 2
( + sin2 2 ).
2

(1.415)

The canonical momenta are

p = Mr 2 ,

p = Mr 2 sin2 ,

(1.416)

and the classical Hamiltonian is given by


1
1 2
H=
p2 +
p .
2
2Mr
sin2


(1.417)

According to the canonical quantization rules, the momenta should become operators
1
p = ih 1/2 sin1/2 , p = ih .
(1.418)
sin
But as explained in the previous section, these momentum operators are not expected to give the correct Hamiltonian operator when inserted into the Hamiltonian
(1.417). Moreover, there exists no proper coordinate transformation from the surface of the sphere to Cartesian coordinates18 such that a particle on a sphere cannot
be treated via the safe Cartesian quantization rules (1.275):
[
pi , xj ] = ihi j ,
[
xi , xj ] = 0,
[
pi , pj ] = 0.

(1.419)

The only help comes from the group properties of the motion on the surface of the
sphere. The angular momentum
L=xp

(1.420)

can be quantized uniquely in Cartesian coordinates and becomes an operator


=x
p

(1.421)

whose components satisfy the commutation rules of the Lie algebra of the rotation
group
i, L
j ] = ihL
k
[L

(i, j, k cyclic).

(1.422)

Note that there is no factor-ordering problem since the xi s and the pi s appear
with different indices in each Lk . An important property of the angular momentum
18

There exist, however, certain infinitesimal nonholonomic coordinate transformations which are
multivalued and can be used to transform infinitesimal distances in a curved space into those in a
flat one. They are introduced and applied in the textbook on Multivalued Fields cited in Ref. [6].
leading once more to the same quantum mechanics as the one described here.

60

1 Fundamentals

operator is its homogeneity in x. It has the consequence that when going from
Cartesian to spherical coordinates
x1 = r sin cos , x2 = r sin sin , x3 = r cos ,

(1.423)

the radial coordinate cancels making the angular momentum a differential operator
involving only the angles , :
1 =
L
ih (sin + cot cos ) ,
2 = ih (cos cot sin ) ,
L
3 = ih .
L

(1.424)

There is then a natural way of quantizing the system which makes use of these
i . We re-express the classical Hamiltonian (1.417) in terms of the classical
operators L
angular momenta
L1 = Mr 2 sin sin cos cos ,


L2 = Mr 2 cos sin cos sin ,




L3 = Mr 2 sin2

(1.425)

as

1
L2 ,
(1.426)
2
2Mr
and replace the angular momenta by the operators (1.424). The result is the Hamiltonian operator:
H=

=
H

h
2
1
1 2
1
2
L =
(sin ) +
2 .
2
2
2Mr
2Mr sin
sin


(1.427)

2 are well known.


The eigenfunctions diagonalizing the rotation-invariant operator L
i , for instance
They can be chosen to diagonalize simultaneously one component of L

the third one, L3 , in which case they are equal to the spherical harmonics
Ylm (, ) = (1)m

"

2l + 1 (l m)!
4 (l + m)!

#1/2

Plm (cos )eim ,

(1.428)

with Plm (z) being the associated Legendre polynomials


Plm (z)

l+m
1
2 m/2 d
(z 2 1)l .
= l (1 z )
l+m
2 l!
dx

(1.429)

The spherical harmonics are orthonormal with respect to the rotation-invariant


scalar product
Z

d sin

d Ylm
(, )Ylm (, ) = ll mm .

(1.430)

61

1.15 Spinning Top

Two important lessons can be learned from this group quantization. First, the
correct Hamiltonian operator (1.427) does not agree with the canonically quantized
one which would be obtained by inserting Eqs. (1.418) into (1.417). The correct
result would, however, arise by distributing dummy factors
g 1/4 = r 1 sin1/2 ,

g 1/4 = r sin1/2

(1.431)

between the canonical momentum operators as observed earlier in Eq. (1.406). Second, just as in the case of polar coordinates, the correct Hamiltonian operator is
equal to
h
2

H=
,
(1.432)
2M
where is the Laplace-Beltrami operator associated with the metric
g = r
i.e.,
=

1.15

1
0
0 sin2

(1.433)

1
1
1

(sin

)
+
2 .

r 2 sin
sin2


(1.434)

Spinning Top

For a spinning top, the optimal starting point is again not the classical Lagrangian
but the Hamiltonian expressed in terms of the classical angular momenta. In the
symmetric case in which two moments of inertia coincide, it is written as
H=

1
1
(L 2 + L 2 ) +
L 2 ,
2I
2I

(1.435)

where L , L , L are the components of the orbital angular momentum in the directions of the principal body axes with I , I I , I being the corresponding moments
of inertia. The classical angular momentum of an aggregate of mass points is given
by
X
L=
x p ,
(1.436)

where the sum over runs over all mass points. The angular momentum possesses
a unique operator
X
=
p
,
L
x
(1.437)

i . Since rotations
with the commutation rules (1.422) between the components L
do not change the distances between the mass points, they commute with the constraints of the rigid body. If the center of mass of the rigid body is placed at the
origin, the only dynamical degrees of freedom are the orientations in space. They
can uniquely be specified by the rotation matrix which brings the body from some
standard orientation to the actual one. We may choose the standard orientation

62

1 Fundamentals

to have the principal body axes aligned with the x, y, z-directions, respectively. An
arbitrary orientation is obtained by applying all finite rotations to each point of the
body. They are specified by the 3 3 orthonormal matrices Rij . The space of these
matrices has three degrees of freedom. It can be decomposed, omitting the matrix
indices as
R(, , ) = R3 ()R2 ()R3 (),
(1.438)
where R3 (), R3 () are rotations around the z-axis by angles , , respectively,
and R2 () is a rotation around the y-axis by . These rotation matrices can be
expressed as exponentials
Ri () eiLi /h ,
(1.439)
where is the rotation angle and Li are the 3 3 matrix generators of the rotations
with the elements
(Li )jk = ihijk .
(1.440)
It is easy to check that these generators satisfy the commutation rules (1.422) of
angular momentum operators. The angles , , are referred to as Euler angles.
The 3 3 rotation matrices make it possible to express the infinitesimal rotations around the three coordinate axes as differential operators of the three Euler
angles. Let (R) be the wave function of the spinning top describing the probability
amplitude of the different orientations which arise from a standard orientation by
the rotation matrix R = R(, , ). Under a further rotation by R( , , ), the
wave function goes over into (R) = (R1 ( , , )R). The transformation may
be described by a unitary differential operator

U ( , , ) ei L3 ei L2 ei L3 ,

(1.441)

i is the representation of the generators in terms of differential operators.


where L
To calculate these we note that the 3 3 -matrix R1 (, , ) has the following
derivatives
ih R1 = R1 L3 ,
ih R1 = R1 (cos L2 sin L1 ),
ih R1 = R1 [cos L3 + sin (cos L1 + sin L2 )] .

(1.442)

The first relation is trivial, the second follows from the rotation of the generator
eiL3 /h L2 eiL3 /h = cos L2 sin L1 ,

(1.443)

which is a consequence of Lies expansion formula (4.103), together with the commutation rules (1.440) of the 3 3 matrices Li . The third requires, in addition, the
rotation
eiL2 /h L3 eiL2 /h = cos L3 + sin L1 .
(1.444)

63

1.15 Spinning Top

Inverting the relations (1.442), we find the differential operators generating the
rotations [10]:
1
L
2
L

cos
,
= ih cos cot + sin
sin
!
sin
= ih sin cot cos
,
sin

(1.445)

3 = ih .
L
After exponentiating these differential operators we derive
U (, , )R1 U 1 ( , , )(, , ) = R1 (, , )R(, , ),
1 ( , , ) = R1 ( , , )R(, , ), (1.446)
U ( , , )R(, , )U
so that U ( , , )(R) = (R), as desired.
along the body axes.
In the Hamiltonian (1.435), we need the components of L
They are obtained by rotating the 3 3 matrices Li by R(, , ) into
L = RL1 R1 = cos cos (cos L1 + sin L2 )
+ sin (cos L2 sin L1 ) cos sin L3 ,

L = RL2 R1 = sin cos (cos L1 + sin L2 )


+ cos (cos L2 sin L1 ) + sin sin L3 ,

(1.447)

L = RL3 R1 = cos L3 + sin (cos L1 + sin L2 ),

i in the final expressions. Inserting (1.445), we find the operand replacing Li L


ators
!

= ih cos cot sin + cos ,


L
sin
)
sin

= ih sin cot cos


L
,
sin
= ih .
L

(1.448)

Note that these commutation rules have an opposite sign with respect to those in
i :19
Eqs. (1.422) of the operators L
, L
] = ihL
,
[L

, , = cyclic.

(1.449)

The sign is most simply understood by writing


= ai L

L
i,
19

= ai L

L
i,

= ai L

L
i,

(1.450)

When applied to functions not depending on , then, after replacing and , the
1.
operators agree with those in (1.424), up to the sign of L

64

1 Fundamentals

where ai , ai , ai , are the components of the body axes. Under rotations these behave
i , aj ] = ihijk ak , i.e., they are vector operators. It is easy to check that this
like [L

property produces the sign reversal in (1.449) with respect to (1.422).


The correspondence principle is now applied to the Hamiltonian in Eq. (1.435)
by placing operator hats on the La s. The energy spectrum and the wave functions
, L
, L
. The
can then be obtained by using only the group commutators between L
spectrum is
"
! #
1
1
1
2
EL = h

(1.451)
2 ,
L(L + 1) +

2I
2I
2I
2 , and = L, . . . , L
where L(L + 1) with L = 0, 1, 2, . . . are the eigenvalues of L

are the eigenvalues of L . The wave functions are the representation functions of
the rotation group. If the Euler angles , , are used to specify the orientation of
the body axes, the wave functions are
L
Lm (, , ) = Dm
(, , ).

(1.452)

3 , the magnetic quantum numbers, and D L (, , )


Here m are the eigenvalues of L
m
are the representation matrices of angular momentum L. In accordance with (1.441),
one may decompose

i(m+m ) L
L
dmm (),
Dmm
(, , ) = e

(1.453)

with the matrices


dLmm ()

(L + m )!(L m )!
=
(L + m)!(L m)!
"

cos
2

!m+m

#1/2

sin
2

!mm

(m m,m +m)

PLm

(cos ).

(1.454)

For j = 1/2, these form the spinor representation of the rotations around the y-axis
1/2
dm m ()

cos /2 sin /2
sin /2
cos /2

(1.455)

The indices have the order +1/2, 1/2. The full spinor representation function
D 1/2 (, , ) in (1.453) is most easily obtained by inserting into the general expres i with the
sion (1.441) the representation matrices of spin 1/2 for the generators L
commutation rules (1.422), the famous Pauli spin matrices:
1

0 1
1 0

, =

0 i
i 0

, =

1
0
0 1

(1.456)

Thus we can write


D 1/2 (, , ) = ei3 /2 ei2 /2 ei3 /2 .

(1.457)

65

1.15 Spinning Top

The first and the third factor yield the pure phase factors in (1.453). The function
2
1/2
dm m () is obtained by a simple power series expansion of ei /2 , using the fact
that ( 2 )2n = 1 and ( 2 )2n+1 = 2 :
ei

2 /2

= cos /2 i sin /2 2,

(1.458)

which is equal to (4.853).


For j = 1, the representation functions (1.454) form the vector representation
1
(1 + cos )
2
1 sin
2
1
(1 cos )
2

d1m m () =

12 sin 12 (1 cos )

.
cos
12 sin

1
1
sin
(1
+
cos
)
2
2

(1.459)

where the indices have the order +1/2, 1/2. The vector representation goes over
into the ordinary rotation matrices Rij () by mapping the states |1mi onto the
x i
y)/2 using the matrix elements
spherical unit vectors (0) = z, (1) = (
P1
i
1
hi|1mi = (m). Hence R()(m) = m =1 (m )dm m ().
The representation functions D 1 (, , ) can also be obtained by inserting into
the general exponential (1.441) the representation matrices of spin 1 for the genera i with the commutation rules (1.422). In Cartesian coordinates, these are
tors L
i )jk = iijk , where ijk is the completely antisymmetric tensor with
simply (L
i )ij hj|m i =
i )mm = hm|ii(L
123 = 1. In the spherical basis, these become (L
i )ij j (m ). The exponential (ei L 2 )mm is equal to (1.459).
i (m)(L
(,)
The functions Pl
(z) are the Jacobi polynomials [11], which can be expressed
in terms of hypergeometric functions as
(,)

Pl

(1)l (l + + 1)
F (l, l + 1 + + ; 1 + ; (1 + z)/2),
l!
( + 1)

where
F (a, b; c; z) 1 +

a(a + 1) b(b + 1) z 2
ab
z+
+ ... .
c
c(c + 1)
2!

(1.460)

(1.461)

The rotation functions dLmm () satisfy the differential equation


d2
d
m2 + m 2 2mm cos L
2 cot
dmm () = L(L + 1)dLmm (). (1.462)
+
d
d
sin2
!

The scalar products of two wave functions have to be calculated with a measure of
integration that is invariant under rotations:
h2 |1 i

dd sin d 2 (, , )1 (, , ).

(1.463)

The above eigenstates (1.453) satisfy the orthogonality relation


Z

L1
L2
dd sin d Dm
m (, , )Dm m (, , )
1
2
1

= m1 m2 m1 m2 L1 L2

8 2
.
2L1 + 1

(1.464)

66

1 Fundamentals

Let us also contrast in this example the correct quantization via the commutation
rules between group generators with the canonical approach which would start out
with the classical Lagrangian. In terms of Euler angles, the Lagrangian reads
1
L = [I ( 2 + 2 ) + I 2 ],
2

(1.465)

where , , are the angular velocities measured along the principal axes of the
top. To find these we note that the components in the rest system 1 , 2 , 3 are
obtained from the relation
1
k Lk = iRR
(1.466)
as
1 = sin + sin cos ,
2 = cos + sin sin ,
3 = cos + .

(1.467)

After the rotation (1.447) into the body-fixed system, these become
= sin sin cos ,
= cos + sin sin ,
= cos + .

(1.468)

Explicitly, the Lagrangian is


1
2 ].
L = [I ( 2 + 2 sin2 ) + I ( cos + )
2

(1.469)

Considering , , as Lagrange coordinates q with = 1, 2, 3, this can be written


in the form (1.395) with the Hessian metric [recall (1.12) and (1.396)]:
g
whose determinant is

I sin2 + I cos2 0 I cos

0
I
0
=
,
I cos
0
I

(1.470)

g = I2 I sin2 .

(1.471)

p = L/ = I sin2 + I cos ( cos + ),

p = L/ = I ,
p = L/ = I ( cos + ).

(1.472)

Hence the measure d3 q g in the scalar product (1.392) agrees with the rotationinvariant measure (1.463) up to a trivial constant factor. Incidentally, this is also
true for the asymmetric top with I 6= I 6= I , where g = I2 I sin2 , although the
metric g is then much more complicated (see Appendix 1C).
The canonical momenta associated with the Lagrangian (1.465) are, according
to (1.395),
R

67

1.15 Spinning Top

After inverting the metric to


g

I sin2

1
0
cos

0
sin2
0

2
2
cos
0
cos + I sin /I

(1.473)

we find the classical Hamiltonian


cos2
1
1 1 2
p +
+
H=
2
2 I
I sin I
"

1
2 cos
2
+
p p .
2 p
I sin
I sin2

(1.474)

This Hamiltonian has no apparent ordering problem. One is therefore tempted to


replace the momenta simply by the corresponding Hermitian operators which are,
according to (1.399),
p = ih ,

p = ih(sin )1/2 (sin )1/2 = ih( +

1
cot ),
2

p = ih .

(1.475)

Inserting these into (1.474) gives the canonical Hamiltonian operator


can = H
+H
discr ,
H

(1.476)

with
2
I
h
H
2 + cot +
+ cot2 2
2I
I
#
2 cos
1
2


+
sin2
sin2
"

and

1
3
discr 1 ( cot ) + 1 cot2 =
H
.
2
2
4
4 sin 4

(1.477)

(1.478)

agrees with the correct quantum-mechanical operator derived


The first term H
above. Indeed, inserting the differential operators for the body-fixed angular mo The term H
discr is the
menta (1.448) into the Hamiltonian (1.435), we find H.
discrepancy between the canonical and the correct Hamiltonian operator. It exists
even though there is no apparent ordering problem, just as in the radial coordinate
expression (1.412). The correct Hamiltonian could be obtained by replacing the
classical p 2 term in H by the operator g 1/4 p g 1/2 p g 1/4 , by analogy with the
of Eq. (1.406).
treatment of the radial coordinates in H
As another similarity with the two-dimensional system in radial coordinates and
the particle on the surface of the sphere, we observe that while the canonical quantization fails, the Hamiltonian operator of the symmetric spinning top is correctly
given by the Laplace-Beltrami operator (1.389) after inserting the metric (1.470)

68

1 Fundamentals

and the inverse (1.473). It is straightforward although tedious to verify that this is
also true for the completely asymmetric top [which has quite a complicated metric
given in Appendix 1C, see Eqs. (1C.2), and (1C.4)]. This is an important nontrivial
result, since for a spinning top, the Lagrangian cannot be obtained by reparametrizing a particle in a Euclidean space with curvilinear coordinates. The result suggests
that a replacement
g (q)p p h2

(1.479)

produces the correct Hamiltonian operator in any non-Euclidean space.


What is the characteristic non-Euclidean property of the , , space? It is the
curvature scalar R. For the asymmetric spinning top we find (see Appendix 1C)
R=

(I + I + I )2 2(I2 + I2 + I2 )
.
2I I I

(1.480)

Thus, just like a particle on the surface of a sphere, the spinning top corresponds to a
particle moving in a space with constant curvature. In this space, the correct correspondence principle can also be deduced from symmetry arguments. The geometry
is most easily understood by observing that the , , space may be considered as
the surface of a sphere in four dimensions, as was shown in detal in Chapter 8 of
Ref. [1].
An important non-Euclidean space of physical interest is encountered in the
context of general relativity. Originally, gravitating matter was assumed to move
in a spacetime with an arbitrary local curvature. In newer developments of the
theory one also allows for the presence of a nonvanishing torsion. In such a general
situation, where the group quantization rule is inapplicable, the correspondence
principle has always been a matter of controversy [see the references after (1.413)]. It
was solved only in the textbook [1], where a natural and unique passage from classical
to quantum mechanics in any coordinate frame was given.20 The configuration
space may carry curvature and a certain class of torsions (gradient torsion). Several
arguments suggest that our principle is correct. For the above systems with a
Hamiltonian which can be expressed entirely in terms of generators of a group of
motion in the underlying space, the new quantum equivalence principle will give the
same results as the group quantization rule.

1.16

Classical and Quantum Statistics

Consider a physical system with a constant number of particles N whose Hamiltonian has no explicit time dependence. If it is brought into contact with a thermal
reservoir at a temperature T and has reached equilibrium, its thermodynamic prop20

H. Kleinert, Mod. Phys. Lett. A 4 , 2329 (1989) (http://www.physik.fu-berlin.de/


~kleinert/199); Phys. Lett. B 236 , 315 (1990) (ibid.http/202).

69

1.16 Classical and Quantum Statistics

erties can be obtained through the following rules: At the level of classical mechanics,
each volume element in phase space
dp dq
dp dq
=
h
2h

(1.481)

is occupied with a probability proportional to the Boltzmann factor


eH(p,q)/kB T ,

(1.482)

where kB is the Boltzmann constant,


kB = 1.3806221(59) 1016 erg/Kelvin.

(1.483)

The number in parentheses indicates the experimental uncertainty of the two digits
in front of it. The quantity 1/kB T has the dimension of an inverse energy and is
commonly denoted by . It will be called the inverse temperature, forgetting about
the factor kB . In fact, we shall sometimes take T to be measured in energy units kB
times Kelvin rather than in Kelvin. Then we may drop kB in all formulas.
The integral over the Boltzmann factors of all phase space elements,21
Zcl (T )

dp dq H(p,q)/kB T
e
,
2h

(1.484)

is called the classical partition function. It contains all classical thermodynamic


information of the system. Of course, for a generalZHamiltonian system with many
Y
dpn dqn /2h. The normalized
degrees of freedom, the phase space integral is
n

Boltzmann factor

w(p, q) = Zcl1 (T )eH(p,q)/kB T

(1.485)

is called the classical Gibbs distribution function. The reader may wonder why an
expression containing Plancks quantum h
is called classical . The reason is that
h
can really be omitted in calculating any thermodynamic average. In classical
statistics it merely supplies us with an irrelevant normalization factor which makes
Z dimensionless.

1.16.1

Canonical Ensemble

and the integral


In quantum statistics, the Hamiltonian is replaced by the operator H
over phase space by the trace in the Hilbert space. This leads to the quantumstatistical partition function


Z(T ) Tr eH/kB T Tr eH(p,x)/kB T ,


21

(1.486)

In the sequel we shall always work at a fixed volume V and therefore suppress the argument
V everywhere.

70

1 Fundamentals

denotes the trace of the operator O.


If |ni are eigenstates of the Hamilwhere Tr O
tonian with energy En , the partition function becomes a sum
Z(T ) =

eEn /kB T

(1.487)

The normalized Boltzmann factor


wn = Zcl1 (T )eEn /kB T

(1.488)

is called the quantum-statistical Gibbs distribution function.


is an N-particle Schrodinger Hamiltonian, the quantum-statistical system
If H
is referred to as a canonical ensemble.
The right-hand side of (1.486) contains the position operator x in Cartesian
coordinates rather than q to ensure that the system can be quantized canonically.
In cases such as the spinning top, the trace formula is also valid but the Hilbert
space is spanned by the representation states of the angular momentum operators. In
more general Lagrangian systems, the quantization has to be performed differently
in the way to be described in Chapters 8 and 10 of the textbook [1].
At this point we make an important observation: The quantum partition function is related in a very simple way to the quantum-mechanical time evolution operator. To emphasize this relation we shall define the trace of this operator for
time-independent Hamiltonians as the quantum-mechanical partition function:
h
b , ta ) = Tr ei(tb ta )H/
ZQM (tb ta ) Tr U(t
.

(1.489)

This may be considered as the partition function associated with a quantummechanical Gibbs distribution
1
wn = ZQM
(tb ta )eiEn (tb ta )/h .

(1.490)

Obviously the quantum-statistical partition function Z(T ) may be obtained from


the quantum-mechanical one by continuing the time interval tb ta to the negative
imaginary value
ih
ih.
(1.491)
tb ta =
kB T
This simple formal relation shows that the trace of the time evolution operator
contains all information on the thermodynamic equilibrium properties of a quantum
system.

1.16.2

Grand-Canonical Ensemble

For systems containing many bodies it is often convenient to study their equilibrium
properties in contact with a particle reservoir characterized by a chemical potential
. For this one defines what is called the grand-canonical quantum-statistical partition function



(1.492)
ZG (T, ) = Tr e(HN )/kB T .

71

1.16 Classical and Quantum Statistics

is the operator counting the number of particles in each state of the ensemble.
Here N
The combination of operators in the exponent,
G = H
N,

(1.493)

is called the grand-canonical Hamiltonian.


Given a partition function Z(T ) at a fixed particle number N, the free energy is
defined by
F (T ) = kB T log Z(T ).
(1.494)
Its grand-canonical version at a fixed chemical potential is22
FG (T, ) = kB T log ZG (T, ).

(1.495)

The average energy or internal energy is defined by


BT
H/k
E = Tr He

.

Tr eH/kB T .

(1.496)

It may be obtained from the partition function Z(T ) by forming the temperature
derivative

E = Z 1 kB T 2
Z(T ) = kB T 2
log Z(T ).
(1.497)
T
T
In terms of the free energy (1.494), this becomes

E=T
(F (T )/T ) = 1 T
T
T
2

F (T ).

(1.498)

For a grand-canonical ensemble we may introduce an average particle number


defined by


.


)/kB T
N
(H
(1.499)
N = Tr Ne
Tr e(HN )/kB T .
This can be derived from the grand-canonical partition function as
N = ZG 1 (T, )kB T

ZG (T, ) = kB T
log ZG (T, ),

(1.500)

or, using the grand-canonical free energy, as


N =

FG (T, ).

(1.501)

The average energy in a grand-canonical system,

)/kB T
N
(H
E = Tr He
22

.

Tr e(HN )/kB T ,

(1.502)

The grand-canonical free energy FG (T, ) is als known as thermodynamic Gibbs potential and
denoted by (T, ). In Eq. (1.526) we shall see that it is also equal to pV , where p is the pressure.

72

1 Fundamentals

can be obtained by forming, by analogy with (1.497) and (1.498), the derivative
E N = ZG 1 (T, )kB T 2

1T
T

ZG (T, )
T

(1.503)

FG (T, ).

For a large number of particles, the density is a rapidly growing function of


energy. For a system of N free particles, for example, the number of states up to
energy E is given by
N(E) =

X
pi

(E

N
X

p2i /2M),

(1.504)

i=1

where each of the particle momenta pi is summed over all discrete momenta pm in
(1.183) available to a single particle in a finite box of volume V = L3 . For a large
V , the sum can be converted into an integral23
N(E) = V

"
N Z
Y

N
X
d 3 pi
p2i /2M),
(E

(2h)3
i=1

i=1
3
which is
simply [V /(2h) ]
radius 2ME:

(1.505)

times the volume 3N of a 3N-dimensional sphere of

N(E) =

"

V
(2h)3

#N

"

V
(2h)3

#N

3N
(2ME)3N/2

3
N
2

+1

(1.506)
.

Recall the well-known formula for the volume of a unit sphere in D dimensions:
D = D/2 /(D/2 + 1).

(1.507)

The surface is [see Subsection 8.5.2 in [1] for a derivation]


SD = 2 D/2 /(D/2).

(1.508)

This follows directly from the integral24

23

SD =

dD p (p 1) =

i


D/2

dD p 2(p2 1) =

ei

2 D/2
=
(D/2)

dD p

d i(p2 1)
e
(1.509)

(1.510)

Remember, however, the exception noted in the footnote to Eq. (1.188) for systems possessing
a condensate.
24
I. S. Gradshteyn and I. M. Ryzhik, op. cit., Formula 3.382.7.

73

1.16 Classical and Quantum Statistics

Therefore, the density per energy = N /E is given by


"

V
(E) =
(2h)3

#N

2M

(2ME)3N/21
.
( 23 N)

(1.511)

It grows with the very large power E 3N/2 in the energy. Nevertheless, the integral for
the partition function (1.532) is convergent, due to the overwhelming exponential
falloff of the Boltzmann factor, eE/kB T . As the two functions (e) and ee/kB T
are multiplied with each other, the result is a function which peaks very sharply
at the average energy E of the system. The position of the peak depends on the
temperature T . For the free N particle system, for example,
(E)eE/kB T e(3N /21) log EE/kB T .

(1.512)

This function has a sharp peak at


E(T ) = kB T

3N
3N
1 kB T
.
2
2


(1.513)

The width of the peak is found by expanding (1.512) in E = E E(T ):


)

E(T )
1
3N
3N
log E(T )

(E)2 + . . . .
(1.514)
exp
2
2
kB T
2E (T ) 2

Thus, as soon as E gets to be of the order of E(T )/ N, the exponential is reduced


by a factor
of two with respect to E(T ) kB T 3N/2. The deviation is of a relative
order 1/ N , i.e., thepeak is very sharp. With N being very large, the peak at
E(T ) of width E(T )/ N can be idealized by a -function, and we may write
(E)eE/kB T (E E(T ))N(T )eE(T )/kB T .

(1.515)

The quantity N(T ) measures the total number of states over which the system is
distributed at the temperature T .
The entropy S(T ) is now defined in terms of N(T ) by
N(T ) = eS(T )/kB .

(1.516)

Inserting this with (1.515) into (1.532), we see that in the limit of a large number
of particles N:
Z(T ) = e[E(T )T S(T )]/kB T .
(1.517)
Using (1.494), the free energy can thus be expressed in the form
F (T ) = E(T ) T S(T ).

(1.518)

By comparison with (1.498) we see that the entropy may be obtained from the free
energy directly as

S(T ) = F (T ).
(1.519)
T

74

1 Fundamentals

For grand-canonical ensembles we may similarly consider


ZG (T, ) =
where

dE dn (E, n)e(En)/kB T ,

(E, n)e(En)/kB T

(1.520)
(1.521)

is now strongly peaked at E = E(T, ), n = N(T, ) and can be written approximately as


(E, n)e(En)/kB T (E E(T, )) (n N(T, ))
eS(T,)/kB e[E(T,)N (T,)]/kB T .

(1.522)

Inserting this back into (1.520) we find for large N


ZG (T, ) = e[E(T,)N (T,)T S(T,)]/kB T .

(1.523)

For the grand-canonical free energy (1.495), this implies the relation
FG (T, ) = E(T, ) N(T, ) T S(T, ).

(1.524)

By comparison with (1.503) we see that the entropy can be calculated directly from
the derivative of the grand-canonical free energy

S(T, ) = FG (T, ).
(1.525)
T
The particle number is, of course, found from the derivative (1.501) with respect to
the chemical potential, as follows directly from the definition (1.520).
The canonical free energy and the entropy appearing in the above equations
depend on the particle number N and the volume V of the system, i.e., they are
more explicitly written as F (T, N, V ) and S(T, N, V ), respectively.
In the arguments of the grand-canonical quantities, the particle number N is
replaced by the chemical potential .
Among the arguments of the grand-canonical free energy FG (T, , V ), the volume
V is the only one which grows with the system. Thus FG (T, , V ) must be directly
proportional to V . The proportionality constant defines the pressure p of the system:
FG (T, , V ) p(T, , V )V.

(1.526)

Under infinitesimal changes of the three variables, FG (T, , V ) changes as follows:


dFG (T, , V ) = SdT + dN pdV.

(1.527)

The first two terms on the right-hand side follow from varying Eq. (1.524) at a fixed
volume. When varying the volume, the definition (1.526) renders the last term.
Inserting (1.526) into (1.524), we find Eulers relation:
E = T S + N pV.

(1.528)

The energy has S, N, V as natural variables. Equivalently, we may write


F = N pV,
where T, N, V are the natural variables.

(1.529)

75

1.17 Density of States and Tracelog

1.17

Density of States and Tracelog

In many thermodynamic calculations, a quantity of fundamental interest is the density of states. To define it, we express the canonical partition function

Z(T ) = Tr eH/kB T

(1.530)

as a sum over the Boltzmann factors of all eigenstates |ni of the Hamiltonian:, i.e.
eEn /kB T .

(1.531)

dE (E)eE/kB T .

(1.532)

Z(T ) =

X
n

This can be rewritten as an integral:


Z(T ) =

The quantity
(E) =

X
n

(E En )

(1.533)

specifies the density of states of the system in the energy interval (E, E + dE). It
may also be written formally as a trace of the density of states operator (E):

(E) = Tr (E) Tr(E H).

(1.534)

The density of states is obviously the Fourier transform of the canonical partition
function (1.530):
(E) =

d E  H 
=
e Tr e
2i

The integral
N(E) =

d E
e Z(1/kB ).
2i

dE (E )

(1.535)

(1.536)

is the number of states up to energy E. The integration may start anywhere below
the ground state energy. The function N(E) is a sum of Heaviside step functions
(1.319):
X
N(E) =
(E En ).
(1.537)
n

This equation is correct only with the Heaviside function, which is equal to 1/2 at
the origin, not with the one-sided version (1.312), as we shall see later. Indeed, if
integrated to the energy of a certain level En , the result is
N(En ) = (n + 1/2).

(1.538)

This formula may be used to determine the energies of bound states from approximations to (E), the classical approximation leading to the well-known BohrSommerfeld condition. In order to apply this relation one must be sure that all

76

1 Fundamentals

levels have different energies. Otherwise N(E) jumps at En by half the degeneracy
of this level.
An important quantity related to (E) is frequently used in this text: the trace
E.
of the logarithm, short tracelog, of the operator H
E) =
Trlog(H

X
n

log(En E).

(1.539)

It may be expressed in terms of the density of states (1.534) as


E) = Tr
Trlog(H

log(E E) =
dE (E H)

dE (E ) log(E E).

(1.540)
The tracelog of the Hamiltonian operator itself can be viewed as a limit of an operator

zeta function associated with H:


,
H () = Tr H

(1.541)

whose trace is the generalized zeta-function


) =
H () Tr H () = Tr(H
h

En .

(1.542)

For a linearly spaced spectrum En = n with n = 1, 2, 3 . . . , this reduces to Riemanns


zeta function (2.276).
From the generalized zeta function we can obtain the tracelog by forming the
derivative
= ()| .
Trlog H
H
=0

(1.543)

By differentiating the tracelog (1.539) with respect to E, we find the trace of the
resolvent (11.8):
E) = Tr
E Trlog(H

X
1
1
1 X
1

=
=
Rn (E) = Tr R(E).

ih n
ih
EH
n E En

(1.544)

Recalling Eq. (1.337) we see that the imaginary part of this quantity slightly above
the real E-axis yields the density of states
X
1
E i) =
Im E Trlog(H
(E En ) = (E).

(1.545)

By integrating this over the energy we obtain the number of states function N(E)
of Eq. (1.536):
X
1
=
(E En ) = N(E).
(1.546)
Im Trlog(E H)

77

Appendix 1A

Simple Time Evolution Operator

Appendix 1A

Simple Time Evolution Operator

Consider the simplest nontrivial time evolution operator of a spin-1/2 particle in a magnetic field
0 = B /2, so that the time evolution operator reads,
B. The reduced Hamiltonian operator is H
in natural units with h = 1,

(1A.1)
eiH0 (tb ta ) = ei(tb ta )B/2 .
Expanding this as in (1.300) and using the fact that (B )2n = B 2n and (B )2n+1 = B 2n (B ),
we obtain

 sin B(tb ta )/2 ,


eiH0 (tb ta ) = cos B(tb ta )/2 + iB
(1A.2)
B/|B|. Suppose now that the magnetic field is not constant but has a small timewhere B
dependent variation B(t). Then we obtain from (1.260) [or the lowest expansion term in (1.300)]
e

0 (tb ta )
iH

tb

ta

dt eiH0 (tb t) B(t) eiH0 (tta ) .

(1A.3)

Using (1A.2), the integrand on the right-hand side becomes


i
i
h
h
 sin B(tb t)/2 B(t)  cos B(tta )/2+iB
 sin B(tta )/2 . (1A.4)
cos B(tb t)/2+iB

We simplify this with the help of the formula [recall (23.53)]


i j = ij + iijk k

(1A.5)

so that
 B(t)  = B
 =B
B(t) + i[B
B(t)] , B(t)  B
B(t) i[B
B(t)] , (1A.6)
B
and
h
i
B(t)]  B
B(t) B
 + i[B
 B(t)  B
 = B

B
n
o
B(t)] B
+ [B
B(t)]B
[B
B(t)] B
. (1A.7)
= i[B

2 = 1. Thus
The first term on the right-hand side vanishes, the second term is equal to B, since B
we find for the integrand in (1A.4):
cos B(tb t)/2 cos B(t ta )/2 B(t) 
B(t) + i[B
B(t)] }
+i sin B(tb t)/2 cos B(t ta )/2{B

B(t) i[B
B(t)] }
+i cos B(tb t)/2 sin B(t ta )/2{B
+ sin B(tb t)/2 sin B(t ta )/2 B 

(1A.8)

which can be combined to give


n
o

B(t)]  +i sin B(tb ta )/2 BB(t).(1A.9)


cos B[(tb +ta )/2t] B(t)sin B[(tb +ta )/2t] [B
Integrating this from ta to tb we obtain the variation (1A.3).

Appendix 1B

Convergence of the Fresnel Integral

Here we prove the convergence of the Fresnel integral (1.345) by relating it to the Gauss integral.
According to Cauchys integral theorem, the sum of the integrals along the three pieces of the

78

1 Fundamentals

Figure 1.4 Triangular closed contour for Cauchy integral


2

closed contour shown in Fig. 1.4 vanishes since the integrand ez is analytic in the triangular
domain:
I
Z A
Z B
Z O
2
z 2
z 2
z 2
dze
=
dze
+
dze
+
dzez = 0.
(1B.1)
0

Let R be the radius of the arc. Then we substitute in the three integrals the variable z as follows:
0 A:
B 0:
AB:

and obtain the equation


Z R
Z
2
dp ep + ei/4
0

dz = dp,
z 2 = p2
i/4
dz = dp e
, z 2 = ip2
dz = i Rdp, z 2 = p2 ,

z = p,
z = pei/4 ,
z = R ei ,

dp eip +

/4

d iR eR

(cos 2+i sin 2)+i

= 0.

(1B.2)

The first integral converges rapidly to /2 for R . The last term goes to zero in this limit.
To see this we estimate its absolute value as follows:
Z

Z /4
/4

2
2

R (cos 2+i sin 2)+i
d iR e
d eR cos 2 .
(1B.3)

<R
0

0

The right-hand side goes to zero exponentially fast, except for angles close to /4 where the
cosine in the exponent vanishes. In the dangerous regime (/4 , /4) with small > 0, one
certainly has sin 2 > sin 2, so that
Z /4
Z /4
2
sin 2 R2 cos 2
e
.
(1B.4)
R
d eR cos 2 < R
d
sin
2

The right-hand integral can be performed by parts and yields


h
i=/4
2
2
1
R eR cos 2 +
eR cos 2
,
R sin 2
=
which goes to zero like 1/R for large R. Thus we find from (1B.2) the limiting formula

ei/4 /2, or
Z

2
dp eip = ei/4 ,

(1B.5)
R0

dp eip =
(1B.6)

p
which goes into Fresnels integral formula (1.345) by substituting p p a/2.

Appendix 1C

The Asymmetric Top

The Lagrangian of the asymmetric top with three different moments of inertia reads
L=

1
[I 2 + I 2 + I 2 ].
2

(1C.1)

Appendix 1C

The Asymmetric Top

79

It has the Hessian metric [recall (1.12) and (1.396)]


g11

= I sin2 + I cos2 (I I ) sin2 sin2 ,

g21
g31

= (I I ) sin sin cos ,


= I cos ,

g22
g32

= I + (I I ) sin2 ,
= 0,

g33

= I ,

(1C.2)

rather than (1.470). The determinant is


g = I I I sin2 ,

(1C.3)

and the inverse metric has the components


g 11

g 21

g 31

g 22

g 32

g 33

1
{I + (I I ) sin2 }I ,
g
1
sin sin cos (I I )I ,
g
1
{cos [ sin2 (I I ) I ]}I ,
g
1
{sin2 [I sin2 (I I )]}I ,
g
1
{sin cos sin cos (I I )}I ,
g
1
{sin2 I I + cos2 I I + cos2 sin2 (I I )I }.
g

(1C.4)

From this we find the components of the Riemann connection, the Christoffel symbol defined in
Eq. (1.70):
11 1

21 1

=
=

[cos cos sin (I2 I I I2 + I I )]/I I ,

{cos [sin2 (I2 I2 (I I )I )


+ I (I + I I )]}/2 sin I I ,

31 1

22 1

32 1

33 1

11 2

21 2

31 2

22 2

32 2

33 2

11 3

0,

{cos sin [sin2 (I I (I I ) I (I2 I2 ) + I2 (I I ))

21 3

=
=
=
=
=

{cos sin [I2 I2 + (I I )I ]}/2I I ,


0,

[sin2 (I2 I2 (I I )I ) I (I I + I )]/2 sin I I ,

0,
{cos sin [sin2 (I2 I2 I (I I )) I (I I )]}/I I ,
{cos cos sin [I2 I2 I (I I )]}/2I I ,

{sin [sin2 (I2 I2 I (I I )) I (I I I )]}/2I I ,


0,

[cos sin (I2 I2 I (I I ))]/2I I ,

+ (I2 I2 )I I2 (I I )]}/I I I ,

{sin2 [sin2 (2I I (I I ) + I (I2 I2 ) I2 (I I ))

+I I (I I ) + I I (I I )] sin2 ((I2 I2 )I I2 (I I ))

80

1 Fundamentals

31

22 3

32 3

33 3

sin (I2

I2

[cos cos

cos sin (I I )/I ,

=
=

I I (I + I I )}/2 sin I I I ,
I (I I ))]/2I I ,

{cos [sin2 (I2 I2 + (I I )I ) + I (I I + I )]}/2 sin I I ,


0.

(1C.5)

=
. From this
The other components follow from the symmetry in the first two indices

Christoffel symbol we calculate the Ricci tensor, [see Eq. (10.8) in [1]]
11
R
21
R
31
R

=
=

22
R
32
R

33
R

{sin2 [sin2 (I3 I3 (I I I2 )(I I ))


+ ((I + I )2 I2 )(I I )] + I3 I (I I )2 }/2I I I ,

{sin sin cos [I3 I3 + (I I I2 )(I I )]}/2I I I ,

{cos [(I I )2 I2 ]}/2I I ,

{sin2 [I3 I3 + (I I I2 )(I I )] + I3 (I I )2 I }/2I I I ,


0,

[(I I )2 I2 ]/2I I .

(1C.6)

Contraction with g gives the curvature scalar


= [2(I I + I I + I I ) I 2 I 2 I 2 ]/2I I I .
R

(1C.7)

is equal to the
Since the space under consideration is free of torsion, the Christoffel symbol

and R calculated
full affine connection . The same thing is true for the curvature scalars R

from and , respectively.

Notes and References


For more details see some standard textbooks:
I. Newton, Mathematische Prinzipien der Naturlehre, Wiss. Buchgesellschaft, Darmstadt, 1963;
J.L. Lagrange, Analytische Mechanik , Springer, Berlin, 1887;
G. Hamel, Theoretische Mechanik , Springer, Berlin, 1949;
A. Sommerfeld, Mechanik , Harri Deutsch, Frankfurt, 1977;
W. Weizel, Lehrbuch der Theoretischen Physik , Springer, Berlin, 1963;
H. Goldstein, Classical Mechanics, Addison-Wesley, Reading, 1950;
L.D. Landau and E.M. Lifshitz, Mechanics, Pergamon, London, 1965;
R. Abraham and J.E. Marsden, Foundations of Mechanics, Benjamin, New York, 1967;
C.L. Siegel and J.K. Moser, Lectures on Celestial Mechanics, Springer, Berlin, 1971;
P.A.M. Dirac, The Principles of Quantum Mechanics, Clarendon, Oxford, 1958;
L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon, London, 1965;
A. Messiah, Quantum Mechanics, Vols. I and II, North-Holland , Amsterdam, 1961;
L.I. Schiff, Quantum Mechanics, 3rd ed., McGraw-Hill, New York, 1968;
E. Merzbacher, Quantum Mechanics, 2nd ed, Wiley, New York, 1970;
L.D. Landau and E.M. Lifshitz, Statistical Physics, Pergamon, London, 1958;
E.M. Lifshitz and L.P. Pitaevski, Statistical Physics, Vol. 2, Pergamon, London, 1987.
The particular citations in this chapter refer to:
[1] H. Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets, 5th ed., World Scientific, Singapore, 2009, pp. 1-1579 (klnrt.de/b5).
[2] For an elementary introduction see the book
H.B. Callen, Classical Thermodynamics, John Wiley and Sons, New York, 1960.

Notes and References

81

[3] The integrability conditions are named after the mathematician of complex analysis H.A.
Schwarz, a student of K. Weierstrass, who taught at the Humboldt-University of Berlin from
18921921.
[4] L. Schwartz, Theorie des distributions, Vols.I-II, Hermann & Cie, Paris, 1950-51;
I.M. Gelfand and G.E. Shilov, Generalized functions, Vols.I-II, Academic Press, New YorkLondon, 1964-68.
[5] H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation,
World Scientific, Singapore 2009, pp. 1497
(http://users.physik.fu-berlin.de/~kleinert/b11).
[6] Note that in many textbooks, for instance
S. Weinberg, Gravitation and Cosmology, Wiley, New York, 1972,
the upper index and the third index in (1.70) stand at the first position. Our notation follows
the classic book
J.A. Schouten, Ricci Calculus, Springer, Berlin, 1954.
It will allow for a closer analogy with gauge fields in the construction of the Riemann tensor
as a covariant curl of the Christoffel symbol. See also
H. Kleinert, Gauge Fields in Condensed Matter , Vol. II Stresses and Defects, World Scientific, Singapore 1989, pp. 744-1443 (http://www.physik.fu-berlin.de/~kleinert/b2),
or the texbook [5].
[7] An exception occurs in the theory of Bose-Einstein condensation where the single state
p = 0 requires a separate treatment since it collects a large number of particles in what is
called a Bose-Einstein condensate. See p. 169 in the above-cited textbook by L.D. Landau
and E.M. Lifshitz on Statistical Mechanics. Bose-Einstein condensation will be discussed in
Section 2.14. See also Chpater 7 in the textbook [1].
[8] This was first observed by
B. Podolsky, Phys. Rev. 32, 812 (1928).
[9] B.S. DeWitt, Rev. Mod. Phys. 29, 377 (1957);
K.S. Cheng, J.Math. Phys. 13, 1723 (1972);
H. Kamo and T. Kawai, Prog. Theor. Phys. 50, 680, (1973);
T. Kawai, Found. Phys. 5, 143 (1975);
H. Dekker, Physica A 103, 586 (1980);
G.M. Gavazzi, Nuovo Cimento 101A, 241 (1981).
See also the alternative approach by
N.M.J. Woodhouse, Geometric Quantization, Oxford University Press, Oxford, 1992.
[10] C. van Winter, Physica 20, 274 (1954).
[11] For detailed properties of the representation matrices of the rotation group, see
A.R. Edmonds, Angular Momentum in Quantum Mechanics, Princeton University Press,
1960.
[12] L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon, London, 1965.

A common mistake that people make when trying to design something completely foolproof
is to underestimate the ingenuity of complete fools
Douglas Adams (1952-2001)

2
Field Formulation of Many-Body
Quantum Physics
A piece of matter composed of a large number of microscopic particles is called a
many-body system. The microscopic particles may either be all identical or of different species. Examples are crystal lattices, liquids, and gases, all of these being
aggregates of molecules and atoms. Molecules are composed of atoms which, in turn,
consist of an atomic nucleus and electrons, held together by electromagnetic forces,
or more precisely their quanta, the photons. The mass of an atom is mostly due to
the nucleus, only a small fraction being due to the electrons and an even smaller
fraction due to the electromagnetic binding energy. Atomic nuclei are themselves
bound states of nucleons, held together by mesonic forces, or more precisely their
quanta, the mesons. The nucleons and mesons, finally, consist of the presently most
fundamental objects of nuclear material, called quarks, held together by gluonic
forces. Quarks and gluons are apparently as fundamental as electrons and photons.
It is a wonderful miracle of nature that this deep hierarchy of increasingly fundamental particles allows for a common description with the help of a single theoretical
structure called quantum field theory.
As a first step towards developing this powerful theory we shall start from the
well-founded Schrodinger theory of nonrelativistic spinless particles. We show that
there exists a completely equivalent formulation of this theory in terms of quantum
fields. This formulation will serve as a basis for setting up various quantum field
theoretical models which can eventually be composed to explain the physics of the
entire particle hierarchy described above.

2.1

Mechanics and Quantum Mechanics


for n Distinguishable Nonrelativistic Particles

For a many-body system with only one type of nonrelativistic spinless particles of
mass M, which may be spherical atoms or molecules, the classical Lagrangian has
the form
L(x , x ; t) =

n
X

M 2
x V (x1 , . . . , xn ; t),
=1 2
82

(2.1)

2.1 Mechanics and Quantum Mechanics for n Nonrelativistic Particles

83

where the arguments x , x in L(x , x ; t) stand, pars pro toto, for all positions x
and velocities x , = 1, . . . , n. The general n-body potential V (x1 , . . . , xn ; t) can
usually be assumed to consist of a sum of an external potential V1 (x ; t) and a pair
potential V2 (x x ; t), also called one- and two-body-potentials, respectively:
V (x1 , . . . , xn ; t) =

V1 (x ; t) +

1X
V2 (x x ; t).
2 ,

(2.2)

The second sum is symmetric in and , so that V2 (x x ; t) may be taken as


a symmetric function of the two spatial argumentsany asymmetric part would
not contribute. The symmetry ensures the validity of Newtons third law actio
equal reactio. The two-body potential is initially defined only for 6= , and
the sum is restricted accordingly, but for the development to come it will be useful to include also the = -terms into the second sum (2.2), and compensate
this by an appropriate modification of the one-body potential V1 (x ; t), so that the
total potential remains the same. Such a rearrangement excludes pair potentials
V2 (x x ; t) which are singular at the origin, such as the Coulomb potential between point charges V2 (x x ; t) = e2 /4|x x |. Physically, this is not a serious
obstacle since all charges in nature have really a finite charge radius. Even the light
fundamental particles electrons and muons possess a finite charge radius, as will be
seen in Chapter 12.
At first we shall consider all particles to be distinguishable. This often unphysical
assumption will be removed later.
The Euler-Lagrange equations of motion associated with the above Lagrangian
read [see Eq. (1.8)]
=
Mx

V1 (x ; t) X

V2 (x x ; t).
x
x

(2.3)

The transition to the Hamiltonian formalism proceeds by introducing the canonical


momenta [see (1.8)]
L
p =
= M x
(2.4)
x
and calculating the Legendre transform [see (1.7)]
H(p , x ; t) =

"
X

p x L(x , x ; t)
p2

2M

V1 (x ; t) +

x =p /M

1X
V2 (x x ; t).
2 ,

(2.5)

The Hamilton equations of motions are [see (1.13)]


V1 (x ; t) X

V2 (x x ; t),
x
x
p
H
=
,
x = {H, x } =
p
M

p = {H, p } =

(2.6)

84

2 Field Formulation of Many-Body Quantum Physics

with the Poisson brackets [see (1.16)]


{A, B} =

n
X

=1

A B
B A

p x
p x

(2.7)

An arbitrary observable F (p , x ; t) changes as a function of time according to the


equation of motion [see (1.15)]
dF
F
= {H, F } +
.
dt
t

(2.8)

It is now straightforward to write down the laws of quantum mechanics for the
system. We follow the rules in Eqs. (1.235)(1.237), and take the local basis
|x1 , . . . , xn i

(2.9)

:
as eigenstates of the position operators x
|x1 , . . . , xn i = x |x1 , . . . , xn i,
x

= 1, . . . , n.

(2.10)

They are orthonormal to each other,


hx1 , . . . , xn |x1 , . . . , xn i = (3) (x1 x1 ) (3) (xn xn ),

(2.11)

and form a complete basis in the space of localized n-particle states:


Z

d3 x1 d3 xn |x1 , . . . , xn ihx1 , . . . , xn | = 1.

(2.12)

An arbitrary state is denoted by a ket vector and can be expanded in this basis by
multiplying the state formally with the unit operator (13.242), yielding the expansion
|(t)i 1 |(t)i =

d3 x1 d3 xn |x1 , . . . , xn ihx1 , . . . , xn |(t)i. (2.13)

The scalar products


hx1 , . . . , xn |(t)i (x1 , . . . , xn ; t)

(2.14)

are the wave functions of the n-body system. They are the probability amplitudes
for the particle 1 to be found at x1 , particle 2 at x2 , etc. .
The Schrodinger equation reads, in operator form

; t)|(t)i = iht |(t)i,


H|(t)i
= H(
p , x

(2.15)

are Schrodingers momentum operators, whose action upon the wave funcwhere p
tions is specified by the rule
hx1 , . . . , xn |
p = ihx hx1 , . . . , xn |.

(2.16)

85

2.2 Identical Particles Bosons and Fermions

Multiplication of (13.244) from the left with the bra vectors hx1 , . . . , xn | yields via
(13.245) the Schrodinger differential equation for the wave functions:
H(ihx , x ; t)(x1 , . . . , xn ; t)
"
#
X h
X
2 2
1X
=
x +
V1 (x ; t) +
V2 (x x ; t) (x1 , . . . , xn ; t)
2 ,
2M

= iht (x1 , . . . , xn ; t).

(2.17)

In many physical systems, the potentials V1 (x ; t) and V2 (x x ; t) are independent


of time. Then we can factor out the time dependence of the wave functions with
fixed energy as
(x1 , . . . , xn ; t) = eiEt/h E (x1 , . . . , xn ),
(2.18)
and find for E (x1 , . . . , xn ) the time-independent Schrodinger equation
H(ihx , x ; t)E (x1 , . . . , xn ) = EE (x1 , . . . , xn ).

2.2

(2.19)

Identical Particles Bosons and Fermions

The quantum mechanical rules presented in the last section were written down
under the assumption that all particles are distinguishable. For a realistic n-body
system, however, this is an unphysical assumption. For example, there is no way of
distinguishing the electrons in an atom. Thus not all of the solutions (x1 , . . . , xn ; t)
to the Schrodinger equation (13.246) can be physically permissible. Consider the
case where the system contains only one species of identical particles, for example
electrons. The Hamilton operator in Eq. (13.246) is invariant under all permutations
of the particle labels . In addition, all probability amplitudes h1 |2 i must reflect
this invariance. They must form a representation space of all permutations. Let Tij
be an operator which interchanges the variables xi and xj .
Tij (x1 , . . . , xi , . . . , xj , . . . , xn ; t) (x1 , . . . , xj , . . . , xi , . . . , xn ; t).

(2.20)

It is called a transposition. The invariance may then be expressed as hTij 1 |Tij 2 i =


h1 |2 i, implying that the wave functions |1 i and |2 i can change at most by a
phase:
Tij (x1 , . . . , xn ; t) = eiij (x1 , . . . , xn ; t).
(2.21)
But from the definition (2.20) we see that Tij satisfies
Tij2 = 1.

(2.22)

Thus eiij can only have the values +1 or 1. Moreover, due to the identity of all
particles, the sign must be the same for any pair ij.
The set of all multiparticle wave functions can be decomposed into wave functions
transforming in specific ways under arbitrary permutations P of the coordinates. It
will be shown in Appendix 2A that each permutation P can be decomposed into

86

2 Field Formulation of Many-Body Quantum Physics

products of transpositions. Of special importance are wave functions which are


completely symmetric, i.e., which are obtained by applying the operation
S=

1 X
P
n! P

(2.23)

to an arbitrary n-particle wave function. The sum runs over all n! permutations
of the particle indices. Another important type of wave function is obtained by
applying the antisymmetrizing operator
A=

1 X
P P
n! P

(2.24)

to the arbitrary n-particle wave function. The symbol P is unity for even permutations and 1 for odd permutations. It is called the parity of the permutation.
By applying S or A to an arbitrary product n-particle wave function one obtains
completely symmetric or completely antisymmetric n-particle wave functions. In
nature both signs can occur. Particles with symmetric wave functions are called
bosons, the others fermions. Examples for bosons are photons, mesons, -particles
or any nuclei with an even atomic number, examples for fermions are electrons,
neutrinos, muons, protons, neutrons, or any nuclei with an odd atomic number.
In two-dimensional multi-electron systems in very strong magnetic fields an interesting new situation has been discovered. These systems have wave functions
which look like those of free quasiparticles on which transpositions Tij yield a phase
factor eiij which is not equal to 1: These quasiparticles are neither bosons nor
fermions, and have therefore been called anyons. Their existence in two dimension
is made possible by imagining each particle to introduce a singularity in the plane
which makes the plane multisheeted. A second particle moving by 3600 around
this singularity does not arrive at the initial point but at a point lying in a second
sheet below the initial point. For this reason, the equation (2.22) need no longer be
fulfilled.
The symmetry properties of the wave functions considerably increase the work in
dealing with the Schrodinger equation. Let us illustrate this by looking at the simplest nontrivial case in which the particles have only a common time-independent external potential V1 (x ) but are otherwise noninteracting, i.e., their time-independent
Schrodinger equation (13.249) reads
X

h
2 2

+ V1 (x ) E (x1 , . . . , xn ) = EE (x1 , . . . , xn ).
2M x

"

(2.25)

It can be solved by the factorizing ansatz in terms of single-particle states E of


energy E ,
E (x1 , . . . , xn ) =

n
Y

=1

E (x ),

(2.26)

87

2.2 Identical Particles Bosons and Fermions

with the total energy being the sum of the individual energies:
E=

n
X

E .

(2.27)

=1

The single-particle states E (x ) are the solutions of the one-particle Schrodinger


equation
#
"
h
2 2
(2.28)
+ V1 (x) E (x) = E E (x).

2M x
If the wave functions E (x) form a complete set of states, they satisfy the oneparticle completeness relation
X

E (x)E (x ) = (3) (x x ).

(2.29)

The sum over the labels may, of course, involve an integral over a continuous
part of the spectrum. It is trivial to verify that the set of all products (13.263) is
complete in the space of n-particle wave functions:
X

1 ,...,n

E1 (x1 ) En (xn ) E n (xn ) E 1 (x1 )


= (3) (x1 x1 ) (3) (xn xn ).

(2.30)

For identical particles, this Hilbert space is greatly reduced. In the case of bosons,
only the fully symmetrized products correspond to physical energy eigenstates. We
apply the symmetrizing operation S of Eq. (2.23) to the product of single-particle
Q
wave functions n=1 Ep() (x ) and normalize the result to find
S
{E
(x1 , . . . , xn )
}

S
N{E
S
}

n
Y

=1

Ep() (x ) =

S
N{E
}

n
1 XY
(x ).
E
n! P =1 p()

(2.31)

The sum runs over n! permutations of the indices = 1, . . . , n, the permuted indices
being denoted by p().
S
Note that {E
(x1 , . . . , xn ) no longer depends on the order of the labels on
}
the energies E1 , . . . , En . This is a manifestation of the indistinguishability of the
particles in the corresponding one-particle states. We have indicated this by the
S
curly brackets notation {E
. Also the normalization factor is independent of the
}
S
order and has been denoted by N{E
.
}
For fermions, the wave functions are
A
A
{E
(x1 , . . . , xn ) = N{E
A
}
}

n
Y

=1

A
Ep() (x ) = N{E
}

n
1 X Y
Ep() (x ), (2.32)
P
n! P
=1

where P = 1 for even or odd permutations, respectively.

88

2 Field Formulation of Many-Body Quantum Physics

Instead of the indices on the labels , we may just as well symmetrize or antisymmetrize those on the position variables:
S,A
{E
(x1 , . . . , xn )
}

S,A
N{E
}

1 X
n! P

1
P

n
Y

E (xp() ).

(2.33)

=1

See Appendix 2A for more details.


The completely antisymmetrized products (2.32) can also be written in the form
of a determinant introduced first by Slater:
A
A
{E
(x1 , . . . , xn ) = N{E
}
}

1
n!

E1 (x1 ) E1 (x2 ) . . . E1 (xn )


..
..
..
.
.
.
En (x1 ) En (x2 ) . . . En (xn )

(2.34)

To determine the normalization factors in (2.31)(2.34), we calculate the integral


Z

S
S
d3 x1 d3 xn {E
(x1 , . . . , xn ){E
(x1 , . . . , xn )
}
}
2

S
= N{E
}

n Z
1 XY
d3 x E (x )Eq() (x ).
p()
n!2 P,Q =1

(2.35)

Due to the group property of permutations, the double-sum contains n! identical


terms with P = Q and can be rewritten as
Z

S
S
d3 x1 d3 xn {E
(x1 , . . . , xn ){E
(x1 , . . . , xn )
}
}

S 2
N{E
}

n Z
1 XY
d3 x E (x )Ep() (x ).
n! P =1

(2.36)

If all single-particle states E (x ) are different from each other, then only the
identity permutation P = 1 with p() = survives, and (2.36) results in
2

S
N{E
= n!.
}

(2.37)

Suppose now that two of the single-particle wave functions E , say E1 and E2 ,
coincide, while all others are different from these and each other. Then only two
permutations survive in (2.36): those with p() = , and those in which P is a
transposition Tij , an exchange of two elements (see Appendix 2A). The right-hand
side of (2.37) is then reduced by a factor 2:
2

S
N{E
=
}

n!
.
2

(2.38)

Extending this consideration to n1 identical states E1 , . . . , En1 , all n1 ! permutations among these give equal contributions to the normalization integral (2.36) and
results in
n!
S 2
.
(2.39)
N{E
=
}
n1 !

89

2.2 Identical Particles Bosons and Fermions

Finally, it is easy to see that if there are groups of n1 , n2 , . . . nk identical states, the
normalization factor is
n!
S 2
.
(2.40)
N{E
=
}
n1 !n2 ! nk !
For the antisymmetric states of fermion systems, the situation is much simpler.
Here, no two states can be identical as is obvious from the Slater determinant (2.34).
Similar considerations as in (2.35), (2.36) for the wave functions (2.32) lead to
2

A
N{E
= n!.
}

(2.41)

The projection into the symmetric and antisymmetric subspaces has the following
effect upon the completeness relation (2.30). Multiplying it by the symmetrization
or antisymmetrization operators
P S,A

1 X
=
n! P

1
P

(2.42)

which may be applied upon the particle positions x1 , . . . , xn as in (2.33), we calculate


X h

1 ,...,n

S,A
N{E
}

i2

S,A
S,A
{E
(x1 , . . . , xn ){E
(x1 , . . . , xn ) = (3)S,A (x1 , . . . , xn ; x1 , . . . , xn ),
}
}

(2.43)

where the symmetrized or antisymmetrized -function is defined by

(3)S,A

(x1 , . . . , xn ; x1 , . . . , xn )

1 X

n! P

1
P

(3) (x1 xp(1) ) (3) (xn xp(n) ).


(2.44)

Since the left-hand side of (2.43) is independent of the order of E1 , . . . , En , we


may sum only over a certain order among the labels
X

1 ,...,n

n!

(2.45)

1 >...>n

If there are degeneracy labels in addition to the energy, these have to be ordered in
the same way.
For antisymmetric states, the labels 1 , . . . , n are necessarily different from each
other. Inserting (2.41) and (2.45) into (2.44), this gives directly
X

1 >...>n

A
A
{E
(x1 , . . . , xn ){E
(x1 , . . . , xn ) = (3)A (x1 , . . . , xn ; x1 , . . . , xn ). (2.46)
}
}

For the symmetric case we can order the different groups of identical states and denote their common labels by n1 , n2 , . . . , nk , with the numbers n indicating how
often the corresponding label is present. Obviously, there are n!/n1 !n2 ! nk !
permutations in the completeness sum (2.43) for each set of labels n1 > n2 >

90

2 Field Formulation of Many-Body Quantum Physics

. . . > nk , the divisions by n ! coming from the permutations within each group
of n1 , . . . , nk identical states. This combinatorial factor cancels precisely the normalization factor (2.40), so that the completeness relation for symmetric n-particle
states reads
X

n1 ,...,nk n1 >...>nk

S
S
{E
(x1 , . . . , xn ){E
(x1 , . . . , xn )
}
}

= (3)S (x1 , . . . , xn ; x1 , . . . , xn ).

(2.47)

The first sum runs over the different breakups of the total number n of states into
identical groups so that n = n1 + . . . + nk .
It is useful to describe the symmetrization or antisymmetrization procedure directly in terms of Diracs bra and ket formalism. The n-particle states are direct
products of single-particle states multiplied by the operator P S,A and can be written
as:
|

S,A

S,A

i=P

|E1 i |En i = N

S,A

1 X
n! P

1
P

|Ep(1) i |Ep(n) i.

(2.48)

The wave functions (2.31) and (2.32) consist of scalar products of these states with
the localized boson states |x1 , . . . , xn i which may be written in factorized form as
|x1 , x2 , . . . , xn i = |x1 i
|x2 i
...
|xn i.

(2.49)

In this state, particle one sits at x1 , particle two at x2 , . . . , etc. The symmetrization
process wipes out the distinction between the particles 1 to n.
Let us adapt the symbolic completeness relation to the symmetry of the wave
functions. The general relation
Z

d3 x1 d3 xn |x1 ihx1 |
...
|xn ihxn | = 1

(2.50)

covers all square integrable wave functions in the product space. As far as the
physical Hilbert space is concerned, it can be restricted as follows
Z

d3 x1 d3 xn |x1 , . . . , xn iS,A

S,A

hx1 , . . . , xn | = P S,A ,

(2.51)

where
|x1 , . . . , xn i

S,A

1 X
=
n! P

1
P

|xp(1) , . . . , xp(n) i.

(2.52)

The states are orthonormal in the sense


S,A

hx1 , . . . , xn |x1 , . . . , xn iS,A = (3)S,A (x1 , . . . , xn ; x1 , . . . , xn ).

This basis will play an essential role for the introduction of quantum fields.

(2.53)

2.3 Creation and Annihilation Operators for Bosons

91

While the formalism presented so far is applicable to any number of particles,


practical calculations usually present a tremendous task. The number of particles is
often so large, of the order 1023 , that no existing computer could even list the wave
functions. On the other hand, macroscopic many-body systems containing such a
large number of microscopic particles make up our normal environment and our
experience teaches us that many global phenomena can be predicted quite reliably.
They should therefore also be calculable in simple terms. For example, for most
purposes a crystal follows the laws of a rigid body and nothing in this laws records
the immense number of degrees of freedom inherent in a microscopic description.
If the solid is exited, there are sound waves in which all the many atoms in the
lattice vibrate around their equilibrium positions. Their description requires only a
few bulk parameters such as elastic constants and mass density. Phenomena of this
type are called collective phenomena.
To describe such phenomena, an economic way has to be found which does not
require the solution of the Schrodinger differential equation with 3n 1023 coordinates. We shall see later that field theory provides us with an elegant and economical access to such phenomena. After a suitable choice of field variables, simple
mean-field approximations will often give a rough explanation of many collective
phenomena. In the subsequent sections we shall demonstrate how the Schrodinger
theory of any number of particles can be transformed into the quantum field theory
of a single field.
There is a further motivation at a more fundamental level for introducing fields.
They offer a natural way of accounting for the symmetry properties of the wave
functions, as we shall now see.

2.3

Creation and Annihilation Operators for Bosons

When dealing with n-particle Schrodinger equations, the imposition of symmetry


upon the Schrodinger wave functions (x1 , . . . , xn ; t) seems to be a rather artificial
procedure. There exists an alternative formulation of the quantum mechanics of
n particles in which the Hilbert space automatically carries the correct symmetry.
This formulation may therefore be viewed as a more natural description of such
quantum systems. The basic mathematical structure which will serve this purpose
is first encountered in a particular quantum mechanical description of harmonic
oscillators which we now recall. It is well-known that the Hamilton operator of an
oscillator of unit mass
1 2 2 2

H = p + q
(2.54)
2
2
can be rewritten in the form


1
=h
H
a a
+
,
(2.55)
2
where

q i
p/
q + i
p/

a
=
,
a
=
(2.56)
2h
2h

92

2 Field Formulation of Many-Body Quantum Physics

are the so-called raising and lowering operators. The canonical quantization rules
[
p, p] = [
x, x] = 0,
[
p, x
] = ih

(2.57)

imply that a, a satisfy


[a, a
] = [a , a
] = 0,
[a, a
] = 1.

(2.58)

The energy spectrum of the oscillator follows directly from these commutation rules.
We introduce the number operator

which satisfies the equations

=a
N
a

(2.59)

a
[N,
] = a
,
a
[N,
] = a.

(2.60)

|i = |i,
N

(2.62)

(2.61)

These imply that a


and a
increase and decrease eigenvalues of the number operator
by one unit, respectively. Indeed, if |i is an eigenstate with eigenvalue ,
N
we see that
a |i = (a N
+ a )|i = ( + 1)a |i,
N
a|i = (aN
a
N
)|i = ( 1)a|i.

(2.63)
(2.64)

Moreover, the eigenvalues must all be integer numbers n larger or equal to zero.
To see this we observe that a
a
is a positive operator since for every state |i in the
Hilbert space
h|a a
|i = ||a |i||2 0.
(2.65)
Hence there must be a state, usually denoted by |0i, whose energy cannot be lowered
by one more application of a
, so that it must satisfy
a
|0i = 0.

(2.66)

applied to |0i must be zero. Applying the raising operator


But then the operator N

will cover all integer numbers = n


a
any number of times, the eigenvalues of N
with n = 0, 1, 2, 3, . . . . The corresponding states are denoted by |ni:

N|ni
= n|ni,

n = 0, 1, 2, 3, . . . .

(2.67)

Explicitly, these states are given by


|ni = Nn (a )n |0i,

(2.68)

2.3 Creation and Annihilation Operators for Bosons

93

with some normalization factor Nn which can be calculated using the commutation
rules (2.58), yielding
1
(2.69)
Nn = .
n!
By considering the commutation rules (2.58) between different states |ni and inserting intermediate states, we derive the matrix elements of the operators a
, a
:

(2.70)
hn |a|ni = n n ,n1 ,

(2.71)
hn |a |ni = n + 1 n ,n+1.
In this way, all properties of the harmonic oscillator are recovered by purely algebraic
manipulations using (2.58) and the condition (2.66) defining the ground state.
This mathematical structure can be used to describe the complete set of symmetric localized states (2.31). All we need to do is reinterpret the eigenvalue n of
which in the case of the oscillator is the quantum number of the
the operator N,
single-particle state, as the number of particles contained in an n-particle state. The
operators a
and a which raise and lower n are renamed creation and annihilation
operators which add or take away a single particle in the state |ni. The ground state
|0i contains no particle. It constitutes the vacuum state of the n-body system. In
the states (2.31) there are n particles at places x1 , . . . , xn . We therefore introduce
the spatial degree of freedom by giving a
, a
a spatial label and defining the operators
a
x , a
x ,
which permit the creation and annihilation of a particle localized at the point x.1
The operators at different locations are taken to be independent, i.e., they commute as
(2.72)
x ] = 0,
[ax , a
x ] = [ax , a
[ax , a
x ] = 0,

x 6= x .

(2.73)

[ax , a
x ] = (3) (x x ).

(2.74)

The commutation rule between ax and a


x for coinciding space variables x and x is
specified with the help of a Dirac -function as follows:

We shall refer to these x-dependent commutation rules as the local oscillator algebra.
The -function singularity in (2.74) is dictated by the fact that we want to
preserve the raising and lowering commutation rules (2.60) and (2.61) for the particle
number at each point x, i.e., we want that
, a
[N
x ] = a
x ,
1

(2.75)

The label x in configuration space of the particles bears no relation to the operator q in the
Hamiltonian (2.54) which is an operator in field space, as we shall understand better in Section
2.8.

94

2 Field Formulation of Many-Body Quantum Physics

a
[N,
x ] = ax .

(2.76)

The total particle number operator is then given by the integral


=
N

d3 x a
x a
x .

(2.77)

Due to (2.73), all parts in the integral (2.77) with x different from the x in (2.75)
and (2.76) do not contribute. If the integral is to give the right-hand sides in (2.75)
and (2.76), the commutator has to be equal to the -function.
The use of the -functions is of course completely analogous to that in Subsection 1.4. [recall the limiting process in Eq. (1.159)]. In fact, we could have introduced
local creation and annihilation operators with ordinary unit commutation rules at
each point by discretizing the space into a fine point-lattice of a small lattice spacing
, with discrete lattice points at
n = 0, 1, 2, . . . .

xn = (n1 , n2 , n3 ),

(2.78)

For the creation or annihilation of a particle in the small cubic box around xn we
use the operators a
n or an , which satisfy the discrete commutation rules
n ] = 0,
[an , a
n ] = [an , a

(2.79)

(3)

(2.80)

a
n a
n .

(2.81)

[an , a
n ] = nn ,
and the total particle number operator is
=
N

X
n

Obviously, we may identify an with the discrete subset of the continuous set of
operators a
x as follows:


a
n = 3 a
x
.
(2.82)
xxn

Then the discrete and continuous formulations of the particle number operator are
related by
Z
X
X
=
N
a
n a
n 3
xn
d3 x a
x a
x .
(2.83)
a
xn a
n

xn

In the same limit, the commutator


1 (3)
1
xn ] = 3 nn
[an , a
n ] [axn , a
3

(2.84)

tends to (3) (x x ), as can be seen in the same way as in Eq. (1.159).


We are now ready to define the vacuum state of the many-particle system. It is
given by the unique state |0i of the local oscillator algebra (2.72) and (2.73), which
contains no particle at all places x, thus satisfying
a
x |0i 0,

h0|ax 0.

(2.85)

2.4 Schr
odinger Equation for Noninteracting Bosons in Terms of Creation . . .

95

It will always be normalized to unity:


h0|0i = 1.

(2.86)

We can now convince ourselves that the fully symmetrized Hilbert space of all localized states of n particles may be identified with the states created by repeated
application of the local creation operators a
x :
1
x a
xn |0i.
|x1 , . . . , xn iS a
n! 1

(2.87)

The so generated Hilbert space will be referred to as the second-quantized Hilbert


space for reasons to be seen below. It decomposes into a direct sum of n-particle
sectors. The symmetry of these states in the position variables is obvious due to the
x among each other.
commutativity (2.72) of all a
x , a
Let us check that the generalized orthonormality relation is indeed fulfilled by
these states. Using the local commutation rules (2.72), (2.73), and the definition of
the vacuum state (2.85), we calculate for one particle
S

hx|x iS = h0|ax a
x |0i
= h0| (3) (x x ) + a
x ax |0i
= (3) (x x ).

(2.88)

For two particles we find


S

1
h0|ax2 ax1 ax a
x |0i
1
2
2!
h
i
1 (3)
=
(x1 x1 )h0|ax2 ax |0i + h0|ax2 a
x a
x1 ax |0i
2
1
2
2!
h
i
1 (3)
(x1 x1 ) (3) (x2 x2 ) + (3) (x2 x1 ) (3) (x1 x2 )
=
2!
(2.89)
= (3)S (x1 , x2 ; x1 , x2 ).

hx1 , x2 |x1 , x2 iS =

The generalization to n particles is straightforward but somewhat tedious. It is


left to the reader as an exercise. Later in Section 7.14.1, rules will be derived in a
different context by a procedure due to Wick, which greatly simplify calculations of
this type.

2.4

Schr
odinger Equation for Noninteracting Bosons
in Terms of Creation and Annihilation Operators

Expressing the localized states in terms of the local creation and annihilation operators a
x , a
x does not only lead to an automatic symmetrization of the states. It also
brings about an extremely simple unified form of the Schrodinger equation which
does not involve the initial specification of the particle number n to be considered,

96

2 Field Formulation of Many-Body Quantum Physics

as in Eq. (13.246). This will now be shown for the case of identical particles without
two-body interaction V2 (x x ; t).
In order to exhibit the unified Schrodinger equation for any number of particles
let us first neglect interactions and consider only the motion of the particles in an
external potential with the Schrodinger equation
(

h
2 2
+ V1 (x ; t)

2M x

"

#)

(x1 , . . . , xn ; t) = iht (x1 , . . . , xn ; t).

(2.90)

We shall now demonstrate that the a , a


-form of this equation which is valid for
any n reads

H(t)|(t)i
= iht |(t)i,
(2.91)

where H(t)
is simply the one-particle Hamiltonian sandwiched between creation and
annihilation operators a
x and a
x and integrated over x, i.e.,

H(t)
=

xax

"

h
2 2
x + V1 (x; t) ax .

2M
#

(2.92)

The operator (2.92) is called the second-quantized Hamiltonian, the equation (2.91)
the second-quantized Schrodinger equation, and the state |(t)i is any n-particle
state in the second-quantized Hilbert space, generated by multiple application of a
x
upon the vacuum state |0i, as described in the last section. The operator nature of
a
x , ax accounts automatically for the many-body nature of Eq. (2.91).
This statement is proved by multiplying Eq. (2.91) from the left with
S

1
hx1 , . . . , xn | = h0|axn a
x1 ,
n!

to obtain
1
1

h0|axn a
x1 H(t)|(t)i
= iht h0|axn a
x1 |(t)i.
n!
n!

(2.93)

We now make use of the property (2.85) of the vacuum state to satisfy h0|ax = 0.
As a consequence, we may rewrite the left-hand side of (2.93) with the help of a
commutator as
1

h0|[axn ax1 , H]|(t)i.


n!
This commutator is easily calculated using the operator chain rules
B
C]
= B[
A,
C]
+ [A,
B]
C,

[A,

C]
= A[
B,
C]
+ [A,
C]
B.

[AB,

(2.94)

These rule are easily memorized by noting that their structure is exactly the same as
in Leibnitz rule for derivatives. In the first rule we may imagine A to be a differential
operator applied to the product BC, which is evaluated by first differentiating B
leaving C untouched, then C leaving B untouched. In the second rule we imagine

97

2.5 Second Quantization and Symmetrized Product Representation

C to be a differential operator acting similarly to the left upon the product AB.
Generalizing this rule to products of more than two operators we derive
x1 , a
z ]
y ]az + a
y [axn a
x1 , a
y a
z ] = [axn ax1 , a
[axn a

x2 , a
y ]ax1 a
z + . . .
x2 [ax1 , a
y ]az + [axn ax3 a
= a
xn ax3 a
X

x+1 a
z a
x1 a
x1 .
(3) (x y)axn a

Multiplying both sides by


h

(3) (y z) h2 z 2 /2M + V1 (z; t) ,


and integrating over d3 y d3 z gives, using (2.92),

x1 , H(t)]
=
[axn a

h
2 2
x1 a
x1 , (2.95)

x+1 a
x a
xn a
+ V1 (x ; t) a
2M x

"

so that, indeed,
1
1
h
2 2
x + V1 (x ; t) h0|axn a
x1|(t)i =iht h0|axn a
x1 |(t)i,

2M
n!
n!

(2.96)
which is precisely the n-body Schrodinger equation (2.90) with the wave function
X

"

1
(x1 , . . . , xn ; t) h0|axn a
x1 | (t)i.
n!

2.5

(2.97)

Second Quantization and Symmetrized Product


Representation

It is worth pointing out that the mathematical structure exploited in the process of
second quantization is of a very general nature.
Consider a set of matrices Mi with indices ,
(Mi )
which satisfy some matrix commutation rules, say
[Mi , Mj ] = ifijk Mk .

(2.98)

Let us sandwich these matrices between creation and annihilation operators which
satisfy
[a , a
] = [a , a ] = 0,
[a , a
] = ,

(2.99)

98

2 Field Formulation of Many-Body Quantum Physics

and form the analogs of second-quantized operators


i a
= a
Mi a.
M
(Mi ) a

(2.100)

Repeated indices imply a summation over all . This is commonly referred to as


Einsteins summation convention. In the last expression we have suppressed the
indices , , for brevity. It is then easy to verify, using the operator chain rules
i satisfy the same commutation rules as the matrices
(2.94), that the operators M
Mi :
h

i, M
j
M

a Mi a
, a
Mj a

+ a Mj a
Mi a
, a

a Mi a
, a
Mj a

= a
Mi Mj a a Mj Mi a

k.
= a
[Mi , Mj ] a = ifijk a Mk a
= ifijk M

(2.101)

i generate an operator representation of


Thus the second-quantized operators M
the matrices Mi . They can be sandwiched between states in the second-quantized
Hilbert space generated by applying products of creation operators a
n upon the
vacuum state |0i. Thereby they are mapped into an infinite-dimensional matrix
representation. On each subspace spanned by the products of a fixed number of
creation operators, they generate the symmetrized part of the direct product representation.
i upon the large Hilbert space
The action of the second-quantized operators M
is very simple to calculate. The only commutation rules required are
i, a
M

i
a
, M

= a
(Mi ) ,

= (Mi ) a
.

(2.102)

From this property we calculate directly the action upon single-particle states:
ia
M
|0i =

i, a
M
|0i = a
|0i(Mi ) ,
i

i = (Mi ) h0|a .
i = h0| a , M
h0|a M
h

(2.103)

Thus the states a


|0i span an invariant subspace and are transformed into each
other via the matrix (Mi ) .
Consider now a state with two particles
a1 a
2 |0i.

(2.104)

i to this state yields


Applying M
i, a
M
1 a
2

i, a
i, a
2
2 + a
1 M
M
1 a

= a
a (Mi )1 1 2 2 + 1 1 (Mi )2 2 .
1

(2.105)

2.5 Second Quantization and Symmetrized Product Representation

99

Multiplying Eq. (2.105) by |0i from the right, we find the transformation law for the
two-particle states a
1 a
2 |0i. They are transformed via the representation matrices
(Mi )1 2 ,1 2 = (Mi )1 1 2 2 + 1 1 (Mi )2 2 .

(2.106)

Omitting the indices we may also write the matrices as


(2)

Mi

= Mi 1 + 1 Mi ,

(2.107)

which is the well-known way of forming representations of a matrix algebra in a


direct product space.
Since a1 and a2 commute with each other, the invariant space constructed in
this way contains only symmetric tensors and only the completely symmetrized part
of the matrices (Mi )1 2 ,1 2 can become active.
The alert reader will have realized that the same operator structure can be obtained for the antisymmetrized parts of the matrices (Mi )1 2 ,1 2 by using creation
and annihilation operators a1 and a2 which satisfy the algebraic rules
} = 0,
{a , a
} = {a , a
{a , a
} = ,

(2.108)

where the curly bracket denotes the anticommutator


{A, B} AB + BA.

(2.109)

The first two lines in (2.101) are unchanged since the operator chain rules (2.94) hold
for Bose and Fermi operators. To derive the third line we must use the additional
rules
B
C]
= {A,
B}
C B{
A,
C},

[A,
C]
= A{
B,
C}
{A,
C}
B.

[AB,

(2.110)

This fermionic version of the commutation relation (2.101) will form, in Section 2.10,
the basis for constructing a second-quantized representation for the n-particle wave
functions and Schrodinger operators of fermions.
If Mi are chosen to be representation matrices Li of the generators of the rotation group (to be discussed in detail in Section 4.1), the law (2.107) represents the
quantum mechanical law of addition of two angular momenta:
(2)

Li = Li 1 + 1 Li .
The generalization to any number of angular momenta is obvious.
i
Incidentally, any operator which satisfies the same commutation rules with M
as a
in (2.102),
i, O

i
, M
O

(Mi ) ,
= O

,
= (Mi ) O

(2.111)

100

2 Field Formulation of Many-Body Quantum Physics

will be referred to as a spinor operator. Generalizing this definition, an opera i like a product O
1 O
2 O
n generalizing
1 2 ...n which commutes with M
tor O
Eq. (2.105) is called a multispinor operator of rank n. Another type of operators
which frequently occurs in quantum mechanics and quantum field theory is a vector
j which commutes with M
i in the same way as M
j
operator, which is any operator O
itself in (2.101), i.e.,
i, O
j = ifijk O
k .
M

(2.112)

i as the product of operators


j1 j2 ...jn which commutes with M
Its generalization O
j1 M
j2 . . . M
jn is called a tensor operator of rank n.
M
The many-particle version of the Schrodinger theory is obtained if we view the
one-particle Schrodinger equation
H(ihx , x) (x, t) = iht (x, t)

(2.113)

as a matrix equation in the discretized x-space with the lattice positions x =


(n1 , n2 , n3 ), so that (x, t) correspond to vectors n (t). Then the differential operator i (x) becomes simply i (x) = [ (x + i) (x)] / = [n+i n ] / where
i is the unit vector in the ith direction and the lattice spacing. The Laplacian may
be viewed as the continuum limit of the matrix
i i (x) =

3
1 X
[ (x + i) 2(x) + (x i)]
2 i=1
3
1 X
[n+i 2n + ni ] .
2 i=1

(2.114)

The Schrodinger equation (2.113) is then the 0 -limit of the matrix equation
Hnn n (t) = iht n (t).

(2.115)

The many-particle Schrodinger equation in second quantization form reads

H|(t)i
= iht |(t)i,

(2.116)

=a
H
n Hn n a
n .

(2.117)

with the Hamilton operator

It can be used to find the eigenstates in the symmetrized multispinor representation


space spanned by
a
n1 . . . a
nN |0i.

(2.118)

to it we see that this state is multiplied from the right by a directApplying H


product matrix
H 1 . . . 1 + 1 H . . . 1 + . . . 1 1 . . . H.

(2.119)

Due to this very general relation, the Schrodinger energy of a many-body system
without two- or higher-body interactions is the sum of the one-particle energies.

101

2.6 Bosons with Two-Body Interactions

2.6

Bosons with Two-Body Interactions

We now include two-body interactions. For simplicity, we neglect the one-body


potential V1 (x; t) which can be added at the end and search for the second-quantized
form of the Schrodinger equation
"

h
2 2
1X
x +
V2 (x x ; t) (x1 , . . . , xn ; t) = iht (x1 , . . . , xn ; t). (2.120)
2M
2 ,
#

It is easy to see that such a two-body potential can be introduced into the secondquantized Schrodinger equation (2.91) by adding to the Hamilton operator in (2.92)
the interaction term
Z
1

d3 xd3 x ax a
x V2 (x x ; t)ax ax .
(2.121)
Hint (t) =
2
To prove this we work out the expectation value
1
int (t)|(t)i = 1 h0|[axn . . . ax1 , H
int (t)]|(t)i
h0|axn . . . ax1 H
n!
n!

(2.122)

with the use of the local commutation rules (2.72), (2.74), and the vacuum property
(2.85). First we generalize Eq. (2.95) to
x1 , a
y2 a
y1 az1 a
z2 ]
[axn a

x1 , a
z1 a
z2 ]
y2 a
y1 [axn a
x1 , a
y2 ay1 ]az1 az2 + a
= [axn a
x1 , a
y1 ]az1 a
z2
x1 , a
y2 ]ay1 az1 az2 + a
y2 [axn a
= [axn a
=

z2
x1 a
x1 a
y1 az1 a
x+1 a
(3) (x y2 )axn a

+a
y2

(3) (x y1 )axn a
x+1 a
x1 ax1 az1 az2 .

(2.123)

The second piece does not contribute to Eq. (2.122) since a


y2 annihilates the vacuum
on the left. For the same reason, the first piece can be written as
X

x+1 a
x1 ax1 , a
y1 ]az1 a
z2
(3) (x y2 )[axn a

(2.124)

as long as it stands to the right of the vacuum. Using the commutation rule (2.95),
this leads to:
z1 a
z2 |(t)i =
y1 a
x1 a
y2 a
h0|axn a

X
,

(3) (x y1 ) (3) (x y2 )

(2.125)

h0|axn a
x+1 a
x1 a
x+1 ax1 . . . ax1 az1 az2 |(t)i.
After multiplying this relation by V2 (y2 y1 ; t) (3) (y1 z1 ) (3) (y2 z2 )/2, and
integrating over d3 y1 d3 y2 d3 z1 d3 z2 , we find
X
int |(t)i = 1
h0|axn ax1 H
x1 | (t) i,
V2 (x x ; t)h0|axn a
2 ,

(2.126)

102

2 Field Formulation of Many-Body Quantum Physics

which is precisely the two-body interaction in the Schrodinger equation (2.120).


Adding now the one-body interactions we see that an n-body Schrodinger equation with arbitrary one- and two-body potentials can be written in the form of a
single operator Schrodinger equation

H(t)|(t)i
= iht |(t)i

(2.127)

with the second-quantized Hamilton operator

H(t)
=

xax

"

1
h
2 2
x + V1 (x; t) a
x +

2M
2
#

d3 x d3 x a
x ax V2 (x x ; t)ax a
x .
(2.128)

The second-quantized Hilbert space of the states |(t)i is constructed by repeated


multiplication of the vacuum vector |0i with particle creation operators a
x . The
order of the creation and annihilation operators in this Hamiltonian is such that the
vacuum as a zero-particle state has zero energy:

H(t)|0i
= 0,

h0|H(t)
= 0,

(2.129)

as in the original Schrodinger equation.


A Hamiltonian which is a spatial integral over a Hamiltonian density H(x) as
H=

d3 x H(x),

(2.130)

is called a local Hamiltonian. In (2.128), the free part is local, but the interacting
part is not. It consists of an integral over two spatial variables, thus forming a bilocal
operator.

2.7

Quantum Field Formulation of Many-Boson


Schr
odinger Equations

The annihilation operator a


x can now be used to define a time-dependent quantum
t) as being the Heisenberg picture of the operator ax (which itself is also
field (x,
referred to as the Schrodinger picture of the annihilation operator). According to
Eq. (1.284), the Heisenberg operator associated with ax is

Thus we define

ax H (t) [U (t, ta )]1 a


x U (t, ta ).

(2.131)

t) ax H (t).
(x,

(2.132)

t) coincides with ax at
Choosing the time variable ta = 0, the quantum field (x,
t = 0:
0) a
(x,
x .
(2.133)

103

2.7 Quantum Field Formulation of Many-Boson Schr


odinger Equations

t) is ruled by Heisenbergs equation of motion (1.279):


The time dependence of (x,
t) =
t (x,

i
t)].
[HH (t), (x,
h

(2.134)

For simplicity, we shall at first assume the potentials to have no explicit time dependence, an assumption to be removed later. Then Eq. (2.134) is solved by
h

t) = eiHt/
(x,
(x, 0)eiHt/h = eiHt/h a
x eiHt/h .

(2.135)

The hermitian conjugate of this determines the time dependence of the Heisenberg
picture of the creation operator:

(x, t) = eiHt/h (x, 0)eiHt/h = eiHt/h ax eiHt/h .

(2.136)

t) fulfills the same commutation rules (2.72) and


At each given time t, the field (x,
(2.73) as a
x :
t), (x
, t)] = 0,
[(x,
[ (x, t), (x , t)] = 0,
t), (x , t)] = (3) (x x ).
[(x,

(2.137)

Consider now the Hamilton operator (2.128) in the Heisenberg representation.


Under the assumption of no explicit time dependence in the potentials we may

simply multiply it by eiHt/h and eiHt/h from the left and right, respectively, and see
that
h
2 2
t)

HH (t) =
d x (x, t)
x + V1 (x) (x,
2M
Z
1
, t)(x,
t).
d3 xd3 x (x, t) (x , t)V2 (x x )(x
+
2
Z

"

(2.138)

commutes with itself, the operator H


H (t) is time independent, so that
Since H
H (t) H.

(2.139)

is now that since it


The important point about the expression (2.138) for H
t), it can be viewed as the Hamilton operator
contains the time-dependent fields (x,
of a canonically quantized Heisenberg field theory by analogy with the Hamiltonian
H(
H
pH (t), xH (t), t) in (1.277). Instead of pH (t) and qH (t) we are dealing here
with generalized coordinates and their canonically conjugate momenta of the field
system. They consist of the hermitian and anti-hermitian parts of the field, R (x, t)
and I (x, t), defined by
( + )

R
,
2

( )
I
.
2i

(2.140)

104

2 Field Formulation of Many-Body Quantum Physics

They commute like


[I (x, t), R (x , t)] = i (3) (x x ),
[I (x, t), I (x , t)] = 0,
[R (x, t), R (x , t)] = 0.

(2.141)

These commutation rules are structurally identical to those between the quasiCartesian generalized canonical coordinates qiH (t) and piH (t) in Eq. (1.96).
In fact, the formalism developed there can be generalized to an infinite set of
canonical variables labeled by the space points x rather than i, i.e., to canonical
variables px (t) and qx (t). Then the quantization rules (1.96) take the form
[
px (t), qx (t)] = ih (3) (x x ),
[
px (t), px (t)] = 0,
[
qx (t), qx (t)] = 0,

(2.142)

which is a local version of the algebra (2.57). The replacement i x can of course
be done on a lattice with a subsequent continuum limit as in Eqs. (2.78)(2.84).
When going from the index i to the continuous spatial variable x, the Kronecker ij
turns into Diracs (3) (x x ), and sums become integrals.
By identifying
px (t) h
I (x, t),
qx (t) R (x, t),
(2.143)
we now obtain the commutation relations (2.141). In quantum field theory it is
customary to denote the canonical momentum variable px (t) by the symbol x (t),
and write
px (t) = h
I (x, t)
(x, t).

(2.144)

Thus the many-body nature of the system may be considered as a consequence


of quantizing the fields qx (t) = R (x, t) and px (t) = h
I (x, t) canonically via
Eq. (2.142).

2.8

Canonical Formalism in Quantum Field Theory

t) and (x , t)
So far, the commutation rules have been imposed upon the fields (x,
by the particle nature of the n-body Schrodinger theory. It is, however, possible to
derive these rules from a standard canonical formalism involving the fields R (x, t)
and I (x, t) as generalized Lagrange coordinates. To see this, let us recall once
more the general procedure for finding the quantization rules and the Schrodinger
equation for a general Lagrangian system with an action
A=

dt L(q(t), q(t)),

(2.145)

105

2.8 Canonical Formalism in Quantum Field Theory

where the Lagrangian L is some function of the independent variables q(t) =


(q1 (t), . . . , qN (t)) and their velocities q(t)

= (q1 (t), . . . , qN (t)). The conjugate momenta are defined, as usual, by the derivatives
pi (t) =

L
.
qi (t)

(2.146)

The Hamiltonian is given by the Legendre transformation


H(p(t), q(t)) =

X
i

pi (t)qi (t) L(q(t), q(t)).

(2.147)

If q(t) are Cartesian or quasi-Cartesian coordinates, quantum physics is imposed in


the Heisenberg picture by letting pi (t), qi (t) become operators piH (t), qiH (t) satisfying the canonical equal time commutation rules
[
piH (t), qjH (t)] = iHhij ,
[
piH (t), pjH (t)] = [
qiH (t), qjH (t)] = 0,

(2.148)

and postulating the Heisenberg equation of motion


d
i

OH = [H
OH .
H , OH ] +
dt
h

(2.149)

H (t) O(
O
pH (t), qH (t), t).

(2.150)

for any observable

This formalism holds for any number of Cartesian or quasi-Cartesian variables.


It can therefore be generalized to functions of space variables xn lying on a lattice af
a small width [see (2.78)]. Suppressing the subscripts of xn , the canonical momenta
(2.146) read
L
px (t) =
,
(2.151)
qx (t)
and the Hamiltonian becomes
H=

X
x

px (t)qx (t) L(qx , qx ).

(2.152)

The canonical commutation rules (2.148) become the commutation rules (2.142) of
second quantization.
In quantum field theory, the formalism must be generalized to continuous space
x. The action (2.145) for which the Hamiltonian is (2.152) is
A=

dt L(t) =

dt

d x I (x, t)ht R (x, t)

dt H[I , R ],

(2.153)

where H[I , R ] denotes the classical Hamiltonian associated with the operator
HH (t) in Eq. (2.138). The derivative term can be written in terms of a kinetic
Lagrangian density Lkin (x, t) as
Akin =

dt

d3 x Lkin

dt

d3 x I (x, t)ht R (x, t).

(2.154)

106

2 Field Formulation of Many-Body Quantum Physics

Then the lattice rule (2.151) for finding the canonical momentum has the following
generalization to find the canonical field momentum
px (t) =

L
qx (t)

(x, t)

Lkin
=h
I (x, t),
t R (x, t)

(2.155)

in agreement with the identification (2.143). The canonical quantization rules


[(x, t), R (x , t)] = i (3) (x x ),
[(x, t), (x , t)] = 0,
[R (x, t), R (x , t)] = 0,

(2.156)

coincide with the quantization rules (2.141) of second quantization. Obviously, the
Legendre transformation (2.152) turns L into the correct Hamiltonian H.
More conveniently, one expresses the classical action in terms of complex fields
A=

dt L(t) =

dt

d3 x (x, t) iht (x, t)

dt H[, ],

(2.157)

and defines the canonical field momentum as


(x, t)

Lkin
=h
(x, t),
t (x, t)

(2.158)

Then the canonical quantization rules become


[(x, t), (x , t)] = i (3) (x x ),
[(x, t), (x , t)] = 0,
[ (x, t), (x , t)] = 0.

(2.159)

We have emphasized before that the canonical quantization rules are applicable
only if the field space is quasi-Cartesian.2 For this, the dynamical metric (1.93) has
to be q-independent. This condition is violated by the interaction in the Hamiltonian
(2.138). There are ambiguities in ordering the field operators in this interaction.
These are, however, removed by the requirement that after quantizing the field
system one wants to reproduce the n-body Schrodinger equation, which requires
that the zero-body state has zero energy and thus satisfies Eq. (2.129).
The equivalence of n-body Schrodinger theory and the above derived canonically
quantized field theory requires specification of the ordering of the field operators after
having imposed the canonical commutation rules upon the fields.
By analogy with the definition of a local Hamiltonian we call an action A local ,
if it can be written as a spacetime integral over a Lagrangian density L:
A=
2

See the remark on page 15.

dt

d3 x L(x, t),

(2.160)

2.8 Canonical Formalism in Quantum Field Theory

107

where L(x, t) depends only on the fields (x, t) and their first derivatives. The
kinetic part in (2.157) is obviously local, the interacting part is bilocal [recall (2.138)].
For a local theory, the canonical field momentum (2.161) becomes
(x, t)

L
=h
(x, t),
t (x, t)

(2.161)

The formal application of the rules (2.142) leads again directly to the commutation
rules (2.137) without the prior splitting into kinetic part and remainder.
In the complex-field formulation, only (x, t) has a canonical momentum, not

(x, t). This, however, is an artifact of the use of complex field variables. Later
in Section 7.4.1 we shall encounter a more severe problem where the canonical momentum of a component of the real electromagnetic vector field vanishes as a consequence of gauge invariance, requiring an essential modification of the quantization
procedure.
Let us calculate the classical equations of motion for the continuous field theory.
They are obtained by extremizing the action with respect to (x) and (x). To
do this we need the rules of functional differentiation. These rules are derived
as follows: we take the obvious differentiation rules stating the independence of
generalized Lagrange variables qi (t), which read
qi (t)
= ij ,
qj (t)

(2.162)

and generalize them to lattice variables


qx (t)
= xx
qx (t)

(2.163)

For continuous field variables, these become


(x, t)
= (x x ),
(x , t)

(2.164)

which can further be generalized to


(x, t)
= (3) (x x )(t t ) = (4) (x x ).
(x , t

(2.165)

The functional derivative the action with respect to the field (x, t) is then obtained
by using the chain rule of differentiation and the rules (2.165).
For a local theory with an action of the form (2.160), the extremality conditions
lead to the Euler-Lagrange equations
A
L
L
=
t
= 0,
(x, t)
(x, t)
t (x, t)

(2.166)

L
L
A
=

= 0.
t

(x, t)
(x, t)
t (x, t)

(2.167)

108

2 Field Formulation of Many-Body Quantum Physics

The second equation is simply the complex-conjugate of the first equation.


Note that these equations are insensitive to surface terms. This is why in spite of
the asymmetric appearance of and in the action (2.157), the equations (2.166)
and (2.167) are complex conjugate to each other. Indeed, the latter gives
Z
h
2 2
iht +
x V1 (x) (x, t) = dx (x , t)V2 (x x ; t)(x , t) (x, t),
2M
(2.168)
and it is easy to verify that (2.166) yields the complex conjugate of this.
After field quantization, the above Euler-Lagrange equation becomes an equation
for the field operator (x , t) , and its conjugate (x , t) must be replaced by the
hermitian conjugate field operator (x , t).
Let us also remark that the equation of motion (2.168) can be used directly
to derive the n-body Schrodinger equation once more using time-dependent field
operators. As a function of time, an arbitrary state vector evolves as follows
"

|(t)i = eiHt/h |(0)i.

(2.169)

Multiplying this by the basis bra-vectors


1
h0|axn a
x1 ,
n!

(2.170)

we obtain the time-dependent Schrodinger wave function


(x1 , . . . , xn ; t).

(2.171)

Inserting between each pair of a


x operators the trivial unit factors 1 = eiHt/h eiHt/h ,
each of these operators is transformed into the time-dependent field operators
, t), and one has
(x
1

n , t) (x
1 , t)|(0)i
(x1 , . . . , xn ; t) = h0|eiHt/h (x
n!

(2.172)

Using the zero-energy property (2.129) of the vacuum state, this becomes
(x1 , . . . , xn ; t) = hx1 , . . . , xn ; t|(0)i.

(2.173)

The bra states arising from the application of the time-dependent field operators

(xi , t) to the vacuum state on the left,


1
n , t) (x
1 , t),
h0|(x
n!

(2.174)

define a new time-dependent basis


hx1 , . . . , xn ; t|,

(2.175)

109

2.9 More General Creation and Annihilation Operators

with the property


hx1 , . . . , xn ; t|(0)i hx1 , . . . , xn |(t)i.

(2.176)

If we apply to the states (2.174) the differential operator (2.168) and use the
canonical equal-time commutation rules (2.137), we may derive once more that
(x1 , . . . , xn ; t) obeys the Schrodinger equation (13.246).
The difference between the earlier form (2.97) and the form (2.173) of defining the
wave function is, of course, the second-quantized version of the difference between
the Schrodinger- and the Heisenberg picture for the ordinary quantum mechanical
wave functions. In Eq. (2.97), the states |(t)i are time-dependent but the basis ket
0) = a
vectors hx1 , .., xn | are not, and with them also the field operators (x,
x generating them. In Eq. (2.173), on the contrary, the states h(0)| are time-independent
(and may be called Heisenberg states), but the local basis bra states hx1 , . . . , xn ; t|
t) generating it. Whatever represenare not, and with them the field operators (x,
tation we use, the n-body wave function (x1 , . . . , xn , t) is always in the Schrodinger
representation. The change of picture concerns only the operator structure of the
many-particle description.
Certainly, there is also the possibility of changing the picture for the
(x1 , . . . , xn , t) wave functions but this unitary transformation would take place
in another Hilbert space, namely in the space of square integrable functions of n
arguments where p and x are the differential operators ihx and x.
When going through the latter alternative proof that (2.173) satisfies the nbody Schrodinger equation we realize that at no place we need the assumption of
time-independent potentials. Thus we can conclude that the canonical quantization
scheme for the action (2.157) is valid for an arbitrary explicit time dependence of the
[see (13.234)] and is always equivalent to the Schrodinger description
potentials in H
for an arbitrary number of particles.

2.9

More General Creation and Annihilation Operators

In many applications it is possible to solve the Schrodinger equation with only the
one-body potential V1 (x; t) exactly. Then it is useful to employ, instead of the
creation and annihilation operators of particles at a point, another equivalent set of
such operators which refers right-away to the corresponding eigenstates. We do this
by expanding the field operator into the complete set of solutions of the one-particle
Schrodinger equation
X
t) =
(x,
(2.177)
E (x, t)a .

If the one-body potential is time independent and there is no two-body potential,


the states have the time dependence
E (x, t) = E (x) eiEt/h .

(2.178)

110

2 Field Formulation of Many-Body Quantum Physics

The expansion (2.177) is inverted to give


a =

t),
d3 x E (x, t)(x,

(2.179)

which we shall write shorter in a scalar-product notation as

a
= (E (t), (t)).

(2.180)

As opposed to the Dirac bracket notation to denote basis-independent scalar products, the parentheses indicate more specifically a scalar product between spatial
wave functions.
From the commutation rules (2.137) we find that the new operators a
, a
satisfy
the commutation rules
] = 0,
[a , a
] = [a , a
[a , a
] = , .

(2.181)

Inserting (2.177) into (2.138) with V2 = 0, we may use the orthonormality relation
among the one-particle states E (x) to find the field operator representation for
the Hamilton operator
X
=
H
E a
a .
(2.182)

The eigenstates of the time-independent Schrodinger equation

H|(t)i
= E|(t)i
are now
|n1 , . . . , nn i =

1
(a1 )n1 (ak )nk |0i,
n1 ! nn !

(2.183)

(2.184)

where the prefactor ensures the proper normalization. The energy is


E=

k
X

Ei ni .

(2.185)

i=1

Finally, by forming the scalar products


1
1
h0|axn ax1 (a1 )n1 (ak )nk |0i
,
n1 ! nk !
n!

(2.186)

we recover precisely the symmetrized wave functions (2.31) with the normalization
factors (2.40).
Similar considerations are, of course, possible in the Heisenberg picture of the
operators a
, a
which can be obtained from
a
(t) =

t).
d3 xE (x)(x,

(2.187)

2.10 Quantum Field Formulation of Many-Fermion Schr


odinger Equations

111

In the field operator description of many-body systems, the Schrodinger wave


function (x, t) has become a canonically quantized field object. Observe that the
field (x, t) by itself contains all relevant quantum mechanical information of the
system via the derivative terms of the action (2.157),
2
t) + h
t).
(x, t)iht (x,
(x, t)x2 (x,
2M

(2.188)

This fixes the relation between wavelength and momentum and between frequency
and energy. The field quantization which introduces the additional processes of
particle creation and annihilation is distinguished from this and often referred to as
second quantization.
It should be kept in mind that for a given n-body system, second quantization is
completely equivalent and does not go beyond the usual n-body Schrodinger theory.
It merely introduces a technical advantage of collecting the wave equations for any
particle number n into a single operator representation. This advantage is, nevertheless, of great use in treating systems with many identical particles. In the limit
of large particle densities, it gives rise to approximations which in the Schrodinger
formulation would be very difficult to reproduce. In particular, collective excitations
of many-particle systems find their easiest explanation in terms of a quantum field
formulation.
The full power of quantum fields, however, will unfold itself when trying to
explain the physics of relativistic particles, where the number of particles is no
longer conserved. Since the second-quantized Hilbert space contains any number of
particles, the second-quantized formulation allows naturally for the description of the
emission and absorption of fundamental particles, processes which the Schrodinger
equation is unable to deal with.

2.10

Quantum Field Formulation of Many-Fermion


Schr
odinger Equations

The question arises whether an equally simple formalism can be found which automatically leads to the correct antisymmetric many-particle states
1 X
P |xp(1) i
...
|xp(n) i.
|x1 , . . . , xn iA =
n! P

(2.189)

This is indeed possible. Let us remember that the symmetry of the wave functions
t) for different position
was a consequence of the commutativity of the operators (x,
values x Obviously, we can achieve an antisymmetry in the coordinates by forming
product states
1
|x1 , . . . , xn iA = a
xn |0i
x a
(2.190)
n! 1
and requiring anticommutativity of the particle creation and annihilation operators:
{ax , a
x } = 0,

{ax , a
x } = 0.

(2.191)

112

2 Field Formulation of Many-Body Quantum Physics

The curly bracket denotes the anticommutator defined in Eq. (2.109). To define a
closed algebra, we require in addition as an analog of the third commutation rule
(2.74) for bosons the anticommutation rule
{ax , a
x } = (3) (x x ).

(2.192)

As in the bosonic case we introduce a vacuum state |0i which is normalized as in


(2.86) and contains no particle [cf. (2.85)]:
a
x |0i = 0 , h0|ax = 0.

(2.193)

The anticommutation rules (2.191) have the consequence that each point can at
most be occupied by a single particle. Indeed, applying the creation operator twice
to the vacuum state yields zero:
x |0i = 0.
x }|0i ax a
x |0i = {ax , a
a
x a

(2.194)

This guarantees the validity of the Pauli exclusion principle.


The properties (2.191), (2.192), and (2.193) are sufficient to derive the manybody Schrodinger equations with two-body interactions for an arbitrary number of
fermionic particles. It is easy to verify that the second-quantized Hamiltonian has
the same form as in Eq. (2.128). The proof proceeds along the same line as in
the symmetric case, Eqs. (2.93)(2.128). A crucial tool is the operator chain rule
(2.110) derived for anticommutators. The minus sign by which anticommutators
differ from commutators cancels out in all relevant equations.
As for bosons we define a time-dependent quantum field for fermions in the
Heisenberg picture as
h

t) = eiHt/
(x,
(x, 0)eiHt/h

= eiHt/h a
x eiHt/h ,
and find equal-time anticommutation rules of
relations (2.137):
t), (x
, t)} =
{(x,
{ (x, t), (x , t)} =
t), (x , t)} =
{(x,

(2.195)

the same type as the commutation


0,
0,
(3) (x x ).

(2.196)

The Hamiltonian has again the form Eq. (2.138).


There is only one place where the fermionic case is not completely analogous
to the bosonic one: The second-quantized formulation cannot be derived from a
standard canonical formalism of an infinite number of generalized coordinates. The
standard formalism of quantum mechanics applies only to true physical canonical
coordinates p(t) and q(t), and these can never account for anticommuting properties
of field variables.3 Thus an identification analogous to (2.143),
px (t) I (x, t),
qx (t) R (x, t),
(2.197)
3

For a detailed discussion of classical machanics with supersymmetric Lagrange coordinates see
A. Kapka, Supersymmetrie, Teubner, 1997

113

2.11 Free Nonrelativistic Particles and Fields

is at first impossible.
The canonical formalism may nevertheless be generalized appropriately. We may
start out with exactly the same classical Lagrangian as in the boson case, Eq. (2.157),
but treat the fields formally as anticommuting objects, i.e.,
(x, t)(x , t ) = (x , t )(x, t).

(2.198)

In mathematics, such objects are called Grassmann variables. Using these, we define
again classical canonical momenta
L
= ih (x, t) (x, t).

(x, t)

px (t)

(2.199)

Together with the field variable qx (t) = (x, t), this is postulated to satisfy the
canonical anticommutation rule
{px (t), qx (t)} = ih (3) (x x ).

2.11

(2.200)

Free Nonrelativistic Particles and Fields

An important way to approach interacting theories will be based on perturbative


methods. Usually, these begin with the free theory and prescribe how to calculate
successive corrections due to the interaction energies. A detailed discussion of how
and when this works will be given later. It seems intuitively obvious, however, that
at least for weak interactions, the free theory may be a good starting point for an
approximation scheme. It is therefore worthwhile to study a few properties of the
free theory in detail.
The free-field action is, according to Eqs. (2.157) and (2.138) for V1 (x) = 0 and
V2 (x x ) = 0:
A=

h
2 2
x (x, t).
dtd x (x, t) iht +
2M
3

"

(2.201)

t) satisfies the field operator equation


The quantum field (x,
h
2 2
iht +
x (x, t) = 0,
2M

(2.202)

with the conjugate field satisfying

h
2 2

x
(x, t) ih t +
2M

= 0.

(2.203)

The equal-time commutation rules for bosons and fermions are


t), (x
, t)] = 0,
[(x,

[ (x, t), (x , t)] = 0,


t), (x , t)] = (3) (x x ),
[(x,

(2.204)

114

2 Field Formulation of Many-Body Quantum Physics

where we have denoted commutator and anticommutator collectively by [ . . . , . . . ] ,


respectively.
In a finite volume V , the solutions of the free one-particle Schrodinger equation
are given by the time-dependent version of the plane wave functions (1.184) [compare
(2.178)]:
(
!)
1
pm 2
i
m
pm (x, t) = exp
p x
t
.
(2.205)
h

2M
V
These are orthonormal in the sense
Z

d3 x p m (x, t)pm (x, t) = pm ,pm ,

(2.206)

and complete, as expressed by


X
pm

pm (x, t)p m (x , t) = (3) (x x ),

(2.207)

valid inside the finite volume V . As in Eq. (2.177) we now expand the field operator
in terms of these solution as
t) =
(x,

pm (x, t)apm .

(2.208)

pm

This expansion is inverted with the help of the scalar product (2.180) as
a
pm

=
= (pm (t), (t))

t).
d3 x p m (x, t)(x,

(2.209)

In the sequel we shall usually omit the superscript of the momenta pm if their
discrete nature is evident from the context. The operators a
p and a
p , obey the
canonical commutation rules corresponding to Eq. (2.181):
[ap , a
p ] = 0,
[ap , a
p ] = 0,
[ap , a
p ] = pp ,

(2.210)

where we have used the modified -functions introduced in Eq. (1.195).


In an infinite volume, we use the time-dependent version of the continous wave
functions (1.194)
(
!)
i
p2
p (x, t) = exp
px
t ,
(2.211)
h

2M
which are orthonormal in the sense
Z

d3 x p (x, t)p (x, t) = (2h)3 (3) (p p )

- (3) (p p ),

(2.212)

115

2.11 Free Nonrelativistic Particles and Fields

and complete, as expressed by


Z

Z
d3 p

p (x, t)p (x , t) d-3 p p (x, t)p (x , t) = (3) (x x ).


3
(2h)

(2.213)

In terms of these continuum wave functions, we expand the field operator as


t) =
(x,

d-3 p p (x, t)a(p),

(2.214)

and have the inverse

a
(p) = (p (t), (t))
=

t).
d3 x p (x, t)(x,

(2.215)

The discrete-momentum operators a


p and a
p and the continuous ones a(p) and
a
(p) are related by [recall Eq. (1.189)]

a
(p) = V ap .
(2.216)
a
(p) = V ap ,
For the continuous-momentum operators a
(p) and a
(p) , the canonical commutation
rules Eq. (2.181) take the form
[a(p), a
(p )] = 0,
[a (p), a
(p )] = 0,
(3)
[a(p), a
(p )] = - (p p ).

(2.217)

The time-independent many-particle states are obtained by repeatedly applying as


in (2.184) any number of creation operators ap [or a
(p)] to the vacuum state |0i,
thus creating states
|np1 , np2 , . . . , npk i = N S,A (ap1 )np1 (apk )npk |0i,

(2.218)

where the normalization factor is determined as in Eq. (2.184). For bosons with np1
identical states of momentum p1 , with np2 identical states of momentum p2 , etc.,
the normalization factor is
1
.
NS = q
np1 ! npk !

(2.219)

The same formula can be used for fermions only that then npi can only take the
values 0 or 1, so that N A 1. The time-independent wave functions are obtained
as
N S,A

ES,A (x1 , . . . , xn ) = h0|(x


ap1 )np1 (apk )npk |0i,
n , 0) (x1 , 0)(
n!
(2.220)

116

2 Field Formulation of Many-Body Quantum Physics

and the time-dependent ones as


S,A (x1 , . . . , xn ; t) = ES,A (x1 , . . . , xn )eiEt/h
(2.221)
S,A
N
h np
n , 0) (x
1 , 0)eiHt/
(ap1 ) 1 . . .(apk )npk|0i
= h0|(x
n!
N S,A

= h0|(x
ap1 (t)]np1 [apk (t)]npk |0i,
n , 0) (x1 , 0)[
n!
with the time-dependent creation operators being defined by

p eiHt/h .
ap (t) eiHt/h a

(2.222)

The energy of the states (2.218) is


E=

k
X

ni pi .

(2.223)

i=1

where p p2 /2M are the energies of the single-particle wave functions (2.211).
The many-body states (2.218) form the so-called occupation number basis of the
Hilbert space. For fermions, ni can only be 0 or 1, due to the anticommutativity of
the operators a
p and a
p among themselves. The basis states are properly normalized
hnp1 np2 np3 . . . npk |np1 np2 np3 . . . npk i = np1 np1 np2 np2 np3 np3 . . . npk npk ,

(2.224)

and satisfy the completeness relation


X

p1 p2 p3 ...

np1 np2 np3 ...npk

|np1 np2 np3 . . . npk ihnp1 np2 np3 . . . npk | = 1S,A ,

(2.225)

where the unit operator on the right-hand side covers only the physical Hilbert space
of symmetric or antisymmetric n-body wave functions.

2.12

Second-Quantized Current Conservation Law

In Section 1.3 we have observed as an essential property for the probability interpretation of the Schrodinger wave functions that the probability density (1.107) and
the probability current density (1.106) were related by the local conservation law
Eq. (1.108):
t (x, t) = j(x, t).
(2.226)
This followed directly from the Schrodinger equation (2.168). Since the same equation holds for the field operators, i.e., with (x , t) replaced by (x , t), the field
operators of charge and current density
t),
(x, t) = (x, t)(x,


j(x, t) = i h
t)
(x, t) (x,
2M

(2.227)

117

2.13 Free-Particle Propagator

satisfy the same relation. When integrating (2.226) over all space as in Eq. (1.109),
the resulting global conservation law ensures the time independence of the operator
=
N

d3 x (x, t) =

t).
d3 x (x, t)(x,

(2.228)

Being time-independent we can use (2.133) to rewrite


=
N

d3 x (x, 0) =

0) =
d3 x (x, 0)(x,

d3 x a
x a
x ,

(2.229)

coincides with the particle number operator (2.77). The original form
so that N
(2.228) is the Heisenberg picture of the particle number operator, which coincides
because of (2.229) with the Schrodinger picture since the particle number is conserved.

2.13

Free-Particle Propagator

The perturbation theory of interacting fields to be developed later in Chapter 10)


requires knowledge of an important free-field quantity called the free-particle propagator. It is the vacuum expectation of the time-ordered product of two free field
operators. As we shall see, the calculation of any observable quantities can be reduced to the calculation of some linear combination of products of free propagators
[see Section 7.14.1]. Let us first extend the definition (1.248) of the time-ordered
product of n time-dependent operators to allow for fermion field operators. Suppose
that the times in an operator product An (tn ) A1 (t1 ) have an order
tin > tin1 > . . . > ti1 .

(2.230)

Then the time-ordered product of the operators is defined by


TAn (tn ) A1 (t1 ) P Ain (tin ) Ai1 (ti1 ).

(2.231)

With respect to the definition (1.249), the right hand side carries a sign factor
P = 1 depending on whether an even or an odd permutation P of the fermion
field operators is necessary to reach the time-ordered form. For bosons, P 1.
The definition of the time-ordered products can be given more concisely using
the Heaviside function (t) of Eq. (1.312).
For two operators, we have
1 )B(t
2 ) = (t1 t2 )A(t
1 )B(t
2 )(t2 t1 )B(t
2 )A(t
1 ),
TA(t

(2.232)

with the upper and lower sign applying to bosons and fermions, respectively. The
free-particle propagator can now be constructed from the field operators as the
vacuum expectation value
t) (x , t )|0i.
G(x, t; x , t ) = h0|T (x,

(2.233)

118

2 Field Formulation of Many-Body Quantum Physics

Applying the free-field operator equations (2.202), (2.203) we derive a remarkable property: The free-particle propagator G(x, t; x , t ) coincides with the Green
function of the Schrodinger differential operator. Recall that a Green function of a
homogeneous differential equation is defined by being the solution of the inhomogeneous equation with a -function source (see Section 1.6). This property may easily
be verified for the free-particle propagator, which satisfies the differential equation
h
2 2
iht +
x G(x, t; x , t ) = ih(t t ) (3) (x x ),
2M



h
2 2

G(x, t; x , t ) ih t +
= ih(t t ) (3) (x x ),
x
2M


(2.234)
(2.235)

thus being a Green function of the free-particle Schrodinger equation: The right t) satisfies the Schrodinger
hand side follows directly from the fact that the field (x,
equation and the obvious formula
t (t t ) = (t t ).

(2.236)

With the help of the chain rule of differentiation and Eq. (2.232), we see that


h
2 2
t) (x , t )|0i
x h0|T(x,
2M
h
i
t) (x , t )|0i t (t t)h0| (x , t )(x,
t)|0i
= ih t (t t )h0|(x,


iht +

t), (x , t)] |0i = ih(t t ) (3) (x x ),


= ih(t t )h0|[(x,

(2.237)

where the commuation and anticommutation rules (2.137) and (2.196) have been
used, and the unit notmalization (2.86) of the vacuum state.
In the theory of differential equations, Green functions are introduced to find
solutions for arbitrary inhomogeneous terms, since such a term can be viewed as a
superposition of -function sources. In quantum field theory, the same Green functions serve as propagators in solving inhomogeneous differential equations involving
field operators.
t) anniExplicitly, the free field propagator is calculated as follows: Since (x,
hilates the vacuum, only the first term in the definition Eq. (2.232) contributes so
that we can write
t) (x , t )|0i.
G(x, t; x , t ) = (t t )h0|(x,

(2.238)

Inserting the expansion Eq. (2.214) with the wave functions (2.211) and using
(2.217), the right-hand side becomes

(t t )


2
2
d-3 p d-3 p ei[(pxp x )(p t/2M p t /2M )]/h h0|a(p)a (p )|0i

= (t t )

d-3 p ei[p(xx )p (tt )/2M ]/h .

(2.239)

119

2.13 Free-Particle Propagator

By completing the square and using the Gaussian integral


Z

1
2
d-3 p eap /2h =
3,
2ha

(2.240)

we find
1
iM (xx )2 /2
h(tt )
G(x, t; x , t ) = (t t ) q
3e
2ih(t t )/M
= G(x x , t t ).

(2.241)

The right-hand side is recognized as the usual quantum-mechanical Green function


of the free-particle Schrodinger equation of Eq. (1.349). Indeed, the factor after
(t t ) is simply the one-particle matrix element of the time evolution operator
)/

iH(tt
h
t) (x , t )|0i = h0|(x)e

h0|(x,
(x )|0i
t )|x i.
= hx|U(t,

(2.242)

This is precisely the expression discussed in Eqs. (1.309)(1.311). It describes the


probability amplitude that a single free particle has propagated from x to x in the
time t t > 0. For t t < 0, G vanishes.
There exists a more useful way of writing the Fourier representation of the propagator than that in Eq. (2.239). It is based on the integral representation (1.318)
of the Heaviside function:
(t t ) =

- eiE(tt )/h
dE

ih
.
E + i

(2.243)

As discussed in general in Eqs. (1.316)(1.317), the i in the denominator ensures


the causality. For t > t , the contour of integration can be closed by an infinite
semicircle below picking up the pole at E = i, so that we obtain by the residue
theorem
(t t ) = 1, t > t .
(2.244)

For t < t , on the other hand, the contour may be closed above and since there is
no pole in the upper half-plane:
(t t ) = 0,

t < t .

(2.245)

Relation (2.243) can be generalized to


iE0 (tt )/
h

(t t )e

- eiE(tt )/h
dE

ih
.
E E0 + i

(2.246)

Using this with E0 = p2 /2M we find from (2.239) the integral representation
G(x x , t t ) =

d-3 p

- eip(xx )/hiE(tt )/h


dE

ih
.
E p2 /2M + i

(2.247)

120

2 Field Formulation of Many-Body Quantum Physics

In this form we can trivially verify the equations of motion (2.234) and (2.235). This
expression agrees, of course, with the quantum mechanical time evolution amplitude
(1.343).
The Fourier-transformed propagator
Z

G(p, E) =
=

d3 x

dt ei(pxEt)/h G(x, t)

ih

p2 /2M

(2.248)

+ i

has the property of being singular when the variable E is equal to a physical particle
energy E = p2 /2M . This condition is often called the energy shell condition. It is a
general property of Green functions that their singularities in the energy-momentum
variables display the spectra of the particles of the system.

2.14
2.14.1

Quantum Statistic of Free Nonrelativistic Fields


Thermodynamic Quantities

Consider the grand-canonical partition function introduced in Eq. (1.120):

ZG (T, ) = Tr(e(HN )/kB T ).

(2.249)

The trace has to be taken over the complete set of basis states (2.218):
X

p1 p2 p3 ...

np1 np2 np3 ...npk

hnp1 np2 np3 . . . |e(Enp1np2np3...

npi )/kB T

|np1 np2 np3 . . . i.

Using the additivity of the energies of all single-particle states found in Eq. (2.223),
this can be written as an infinite product
ZG (T, ) =

p2

p2

p2

1 )n
2
3
( 2M
p1 /kB T ( 2M )np2 /kB T ( 2M )np3 /kB T

p1p2p3 ... np1np2np3 ...npk

(2.250)

Each Boltzmann factor leads to the partition function associated with the available single-particle momentum p1 , p2 , . . .. The product is turned into a sum by
considering the grand-canonical free energy
FG (T, ) kB T log ZG (T, ) = kB T

log

p2

e( 2M )n/kB T .

(2.251)

We now distinguish between Bose and Fermi particles. In the first case, the
occupation numbers ni run over all integers 0, 1, 2, . . . up to infinity:

n=0

p2

e( 2M )n/kB T =

p2
( 2M

1e

)/kB T

(2.252)

121

2.14 Quantum Statistic of Free Nonrelativistic Fields

In the second case, ni can be only zero or one, so that


1
X

p2

p2

e( 2M )n/kB T = 1 + e( 2M )/kB T .

(2.253)

n=0

Thus we obtain
FG (T, ) = kB T

X
p

p2

log[1 e( 2M )/kB T ].

(2.254)

Beacause of the frequent appearance of the energy combination p2 /2M , it will


often be useful to use the symbol
p =

p2
= p
2M

(2.255)

to abbreviate calculations in a grand-canonical ensemble.


In a large volume, momentum states lie so close to each other that the sum may
be approximately evaluated as an integral with the help of the limiting formula
X
p

gV
V

d3 p
= gV
(2h)3

d-3 p.

(2.256)

In writing this we have allowed for a degeneracy number g for each momentum state
p to account for extra degrees of freedom of the particles in each momentum state.
In the absence of internal quantum numbers, g counts the different spin polarization
states. If s denotes the spin, then its thrid component can run from s to s so that
g = 2s + 1.

(2.257)

Morover, the limit (2.256) is certainly valid only for sums over sufficiently smooth
functions. We shall see in Section 2.14.3 that the limit fails for a Bose gas near
T = 0, where the limit requires the more careful treatment in Eq. (2.336).
As an alternative to the momentum integral (2.256), we may integrate over the
single-particle energies. With the energy p = p2 /2M, the relation between the
integration measures is
Z

d-3 p =

4
(2)3h
3
1

dp p2

(2.258)

Z
2 Z
= q
d g d.
3
0
2h2 /M

In the last expression we have introduced the quantity


2
1
,
g q
3

2
2h /M

(2.259)

122

2 Field Formulation of Many-Body Quantum Physics

which is the density of states per unit energy interval and volume. With the help
of this quantity we may write (2.254) as an energy integral
FG (T, , V ) = gV

dg F (T, )

(2.260)

where
F (T, ) kB T log[1 e()/kB T ]

(2.261)

is the grand-canonical free energies for an individual single-particle energy value.


According to the thermodynamic rule (1.501), the average particle number is
found from the derivative of FG (T, , V ) with respect to the chemical potential.
Using (2.254) or the integral representation (2.260) we find
N = gV

X
p

gV

1
p2
( 2M

)/kB T

dg f ,

= gV

dg

F (T, )

(2.262)

where

F (T, ) = ()/k T
(2.263)
B

e
1
are the average Bose and Fermi occupation numbers of a level of energy , respectively. They are plotted in Figs. 2.1 and 2.2.
f

Figure 2.1 Average Bose occupation number nB ( ). Note that free bosons have a
negative chemical potential .

The internal energy of the system can be calculated from the integral
E = g

dgf

2 Z
V

= gq
3
0
2h2 /M

(2.264)

d
.
e()/kB T 1

123

2.14 Quantum Statistic of Free Nonrelativistic Fields

nF ()

e()/kB T

Figure 2.2 Average Fermi occupation number nF (). Fermions have a positive chemical
potential .

On the other hand, we find by a partial integration of the integral in (2.260):


Z

d1/2 log[e()/kB T 1] =

d 3/2
2Z
.
3 0 e()/kB T 1

(2.265)

that the grand-canonical partition function can be rewritten as


FG (T, , V ) = g

2
dg F = g
3

dg f .

(2.266)

This implies the general thermodynamic relation for a free Bose or Fermi gas:
2
FG = E.
3

(2.267)

Recalling the definition of the pressure (1.526), we have thus found the equation of
state for a free Bose or Fermi gas:
2
pV = E.
3

(2.268)

To evaluate the energy integral, we introduce the variable z /kB T and write
(2.262) as


V

,
(2.269)
N = N(T, ) g 3
I
(T ) (3/2) 3/2 kB T
where
1
(T ) q
2
2h /MkB T

(2.270)

is the thermal length associated with mass M and temperature T , and In (/kB T )
denotes the function
Z
z n1

.
(2.271)
In ()
dz z
e
1
0

124

2 Field Formulation of Many-Body Quantum Physics

After expanding the denominator in a power series, each term can be integrated
and leads for bosons to a series representation
In () = (n)

ek
.
n
k
k=1

The sum can be expressed in terms of the polylogarithmic function [7]


n (z)

(2.272)
4

zk
n
k=1 k

(2.273)

as
In () = (n)n (e ).

(2.274)

The sum in (2.272) converges only for < 0. In the limit 0, it has the
limit
In (0) = (n)(n),
n > 1.
(2.275)
where (n) is Riemanns zeta function
()

1
.

k=1 k

(2.276)

For n 1, the function In () diverges like ()n1 in the limit 0. In the


opposite limit , it goes to a constant
In () e (n).

(2.277)

This limit is needed to find the high-temperature behavior of the free Bose gas at a
fixed average particle number N, as we see from Eq. (2.269). For large T , the ratio
/kB T diverges to .
In the case of fermions, the expansions (2.272), (2.274) read
In+ () = (n)

(1)k1

k=1

ek
= (n)n (e ),
kn

(2.278)

the only difference with respect to bosons being the alternating signs in the sum.
For = 0, this becomes
In+ (0) =



X
X
1
1
1
1
1
1
1n
1n
(n). (2.279)
n + n n +... =

2
=
1

2
n
n
1 2
3
4
k=0 k
k=0 k

In the opposite limit , the limit is the same constant as in (2.277)


In+ () e (n).
4

A frequently used notation for n (z) is Lin (z).

(2.280)

125

2.14 Quantum Statistic of Free Nonrelativistic Fields

At a fixed particle number N, the chemical potential changes with temperature


in a way determined by the vanishing of the derivative of (2.269), which yields the
equation

I3/2
( kBT )

(2.281)
= 3 .
T T
kB T
I1/2 ( kB T )
Here we have used the property

z In (z) = (n 1)In1
(z),

(2.282)

which follows directly from the series expansion (2.272). For large T , we obtain from
Eqs. (2.275) and (2.277) the limit
T T

.
kB T T 2

(2.283)

Relation (2.268) can also be obtained from the general thermodynamic calculation

ZG
=
FG ,
(2.284)
E N = ZG1

which follows directly from (2.249) by differentiating with respect to = 1/kB T .


The grand-canonical free energy is from (2.266), using (2.259), (2.263), (2.270),
and (2.271):
FG (T, , V ) = pV

V
2 1

= g 3
kB T
I5/2
(T )
3 (3/2)
kB T

2 I5/2 ( kB T )
.
= N(T, ) kB T

3 I3/2
( kBT )


(2.285)

For large T where , the limiting formula (2.277) shows that FG has the
correct Dulong-Petit behavior NkB T implying the ideal-gas law
pV = NkB T.

(2.286)

Due to relation (2.268), the energy is


E = N(T, ) kB T

I5/2
( kBT )

I3/2
( kBT )

(2.287)

For large temperatures, this has the correct Dulong-Petit limit of free particles
3NkB T /2.
We may check the first line in Eq. (2.285) by differentiating it with respect
to and using the relation (2.282) to reobtain the thermodynamic relation N =
FG /.

126

2 Field Formulation of Many-Body Quantum Physics

According to Eq. (1.519), the entropy S is obtained from the negative derivative
of (2.285) with respect to T . This yields
S = kB g

V
3 (T )

IS

kB T


2 IS k B T
= gkB N(T, )   ,
3 I3/2

(2.288)

kB T

where

5
5
3

IS () I5/2
() I5/2
() = I5/2
() I3/2
().
2
2
2
This agrees with the general thermodynamic relation
FG (T, ) = E(T, ) N(T, ) T S(T, ).

(2.289)

(2.290)

Adiabatic processes are defined by the condition S/N = const. which implies by
Eq. (2.289) that also the ratio /kB T is a constant. Inserting this into (2.269) with
(2.270), we find that at constant particle number an adiabatic process has
V T 3/2 |adiab = const.

(2.291)

Combining this with (2.286) leads to


pV 5/3 |adiab = const.

(2.292)

The specific heat at constant volume and particle number is found from the
entropy by forming the derivative CV = T T S of Eq. (2.288). Using (2.281), this
leads to
CV = gkB N

I5/2
()

I3/2
()

1
,
5
9
2 I3/2
()
I1/2 ()

.
kB T

(2.293)

For large T , we use again (2.277) to show that this becomes a constant
3
CV
CVDP = g NkB ,
T
2

(2.294)

which complies with the classical rule of Dulong and Petit (kB /2 per degree of
freedom).

2.14.2

The Degenerate Fermi Gas Near T = 0

Consider the Fermi gas close to zero temperature which is called the degenerate
limit. Then the occupation number (2.263) reduces to
f =

1
0

for

<
>

= ( ).

(2.295)

127

2.14 Quantum Statistic of Free Nonrelativistic Fields

All states with energy lower than are filled, all higher states are empty. The
chemical potential at zero temperature is called Fermi energy F :

T =0

F .

(2.296)

The Fermi energy for a given particle number N is found from (2.259), (2.262), and
(2.295):

3/2
Z
Z F
p3F
2M 3/2 F
=
gV
,
(2.297)
N = gV
dgf = gV
dg = gV
3 2h
3
6 2 h
3
0
0
where
pF =

2MF h
kF

(2.298)

is the Fermi momentum associated with the Fermi energy. Equation (2.297) is solved
for F by
6 2
g

F =

!2/3 

N
V

2/3

h
2
,
2M

(2.299)

and for the Fermi momentum by


pF k F h
=

6 2
g

!1/3 

N
V

1/3

h
.

(2.300)

In two dimensions, we find


F =

2
.
g M

(2.301)

Note that in terms of the particle number N, the density of states per unit energy
interval and volume (2.259) can be written as
3N 1
gg
2 V F

.
F

(2.302)

As the gas is heated slightly, the degeneracy in the particle distribution function
in Eq. (2.295) softens. In order to study this quantitatively it is useful to define a
characteristic temperature associated with the Fermi energy F , the so-called Fermi
Temperature
F
1 pF 2
h
2 kF 2
TF
=
=
.
(2.303)
kB
kB 2M
kB 2M
For electrons in a metal, kF is of the order of 1/
A. Inserting M = me = 9.109558
1028 g, kB = 1.380622 1016 erg/K, and h
= 6.0545919 1027 erg sec, we see
that the order of magnitude of TF is
TF 44 000 K.

(2.304)

128

2 Field Formulation of Many-Body Quantum Physics

Hence the relation


T /TF 1

(2.305)

is quite well fulfilled even far above room temperatures, and T /TF can be used as
an expansion parameter in evaluating the thermodynamic properties of the electron
gas at nonzero temperature.
Let do this to calculate the corrections to the above equations at small T . Eliminating the particle number in (2.269) in favor of the Fermi temperature with the
help of Eqs. (2.297) and (2.303), we find the temperature dependence of the chemical
potential from the equation
1=

T
TF

3/2

3 +

I3/2
2
kB T


(2.306)

For T 0, the chemical potential approaches the Fermi energy F , so that small
T corresponds to large reduced chemical potential
= /kB T . Let us derive an
+
expansion for I3/2 (
) in powers of 1/
in this regime. Setting
z
x

(2.307)

and writing
In+ (
)

(
+ x)n1
dx
=
ex + 1

(
x)n1
dx x
+
e +1

dx

(
+ x)n1
. (2.308)
ex + 1

In the first integral on the right-hand side we substitute


1
1
=1 x
+1
e +1

ex

(2.309)

and obtain
In+ (
) =

dx xn1 +

dx

(
+x)n1 (
x)n1 Z (
x)n1
+
dx
. (2.310)
ex + 1
ex + 1

In the limit
, only the first term survives, whereas the last term
L(
)

dx

(
x)n1
= (1)n
x
e +1

dx

xn1
= (1)n (n)n (e ) (2.311)
x+

e
+1

is exponentially small, so that it can be ignored in an expansion in powers of 1/


.
The second term is expanded as
2



X
n1

k=odd

n1k

dx

X
xk
(n1)! n1k
=
2

(1 2k )(k + 1).
ex + 1
(n1k)!
k=odd
(2.312)

129

2.14 Quantum Statistic of Free Nonrelativistic Fields

At even positive and odd negative integer arguments, the zeta function is related to
the Bernoulli numbers by5
(1 2n) =

B2n
,
2n

(2n) =

(2)2n
|B2n |.
2(2n)!

(2.313)

The two equations go over into each other via the identity
(x) = 2x x1 sin(x/2)(1 x)(1 x),

(2.314)

which can also be written as


x
.
2

(x) = 2x1 x (1 x)/(x) cos

(2.315)

The lowest values of (k + 1) occurring in the expansions (2.312) are6


4
6
2
, (4) = , (6) =
,
6
90
945

(2) =

(2.316)

so that In+ (
) starts out for large
like
In+ (
) =

1 n
1
7

+ 2(n1) (2)
n2 + 2(n1)(n2)(n3) (4)
n4 + . . . . (2.317)
n
2
8

Inserting this with n = 3/2 into Eq. (2.306), we find


T
1=
TF


3/2

"

3 2

2 3 kB T


3/2

2
+
12 kB T


1/2

7 4
+
+
3 320
kB T


5/2

... ,

(2.318)
from which we derive the small-T /TF expansion of /kB T , which yields for =
F [(/kB T )(T /TF)] the expansion
= F

2 T
1
12 TF

"

2

7 4 T
+
720 TF


4

+ ... .

(2.319)

Only exponentially small terms in eTF /T coming from L(


) are ignored. Inserting
this into the grand-canonical free energy FG in the first line of Eq. (2.285) which
+
contains I5/2
(
), this leads to the expansion
5 2
FG (T, , V ) = FG (0, , V ) 1 +
8
"

T
TF

2

7 4 T

384 TF


4

+ ... ,

(2.320)

where
5

These and the subsequent formulas are found in I.S. Gradshteyn and I.M. Ryzhik, op. cit.,
Formulas 9.542 and 9.535.
6
Other often-needed values are (0) = 1/2, (0) = log(2)/2, (2n) = 0, (3)
1.202057, (5) 1.036928, . . . .

130

2 Field Formulation of Many-Body Quantum Physics

FG (0, , V )

k3
2
2
2M 3/2 5/2
gV

gV
.
5
5 6 2
3 2h
3

(2.321)

Here k is the analog of kF, to which it reduces for T = 0 [compare (2.298)]:


k

1q
2M.
h

(2.322)

At T = 0 where F , we see from (2.298) that


2
FG (0, F, V ) = NF .
5

(2.323)

This can also be obtained from (2.285) using the limiting behavior (2.280).
By differentiating FG with respect to the temperature at fixed , we obtain the
low-temperature behavior of the entropy
S = kB

2 T N
+ ... ,
2 TF V

(2.324)

and from this the specific heat


CV

S
2 T
= T
= kB N .
T V,N
2 TF

(2.325)

This differs from the constant high-temperature Dulong-Petit behavior by a factor [recall (2.294)
2 T
.
(2.326)
CV = CVDP
3g TF
This is the typical linear low-temperature behavior of the electronic specific heat
(see Fig. 2.3 for the full temperature behavior). The linear behavior is due to the
progressive softening of the Fermi distribution which makes more and more electrons
thermally excitable. It is detected experimentally in metals at low temperature
where the contribution of lattice vibrations freezes out as (T /TD )3 . Here TD is
the Debye temperature which characterizes the elastic stiffness of the crystal and
ranges from TD 90 K in soft metals like lead over TD 389 K for aluminum
to TD 1890 K for diamond. The experimental size of the slope is usually larger
than the free electron gas value in (2.325). This can be explained mainly by the
interactions with the lattice which result in a large effective mass Meff > M.
Note that the quantity FG (0, , V ) is temperature-dependent via the chemical
potential . Inserting (2.319) into (2.320) we find the complete T -dependence

5 2
FG (T, , V ) = FG (0, F, V ) 1 +
12

T
TF

2

16

with FG (0, F, V ) given by Eq. (2.323) at = F .

kB T
F

!4

+ . . . .

(2.327)

131

2.14 Quantum Statistic of Free Nonrelativistic Fields

Figure 2.3 Temperature behavior of specific heat of a free Fermi gas.

Recalling the relation (2.267), the above equation supplies us also with the lowtemperature behavior of the internal energy:
3
5 2
E = NF 1 +
5
12
"

T
TF

2

4 T

16 TF


4

+ ... .

(2.328)

The first term is the energy of the zero-temperature Fermi sphere. Using the relation
CV = E/T , the second term yields once more the leading T 0 behavior (2.325)
of the specific heat.
This behavior of the specific heat can be observed in metals where the conduction
electrons behave like a free electron gas. Due to Blochs theorem, a single electron
in a perfect lattice behave just like a free particle. For many electrons, this is still
approximately true, if the mass of the electrons is replaced by an effective mass.
Another important macroscopic system where (2.325) can be observed is a liquid
consisting of the fermionic isotope 3 He. There are two electron spins and an odd
number of nucleon spins which make this atom a fermion. The atoms interact
strongly in the liquid, but it turns out that these interactions produce a screening
effect after which the system may be considered approximately as an almost-free gas
of quasiparticles which behave like free fermions whose mass is about 8 times that
of the strongly interacting atoms.

2.14.3

Degenerate Bose Gas Near T = 0

For bosons, the low temperature discussion is quite different. As we can see from
Eq. (2.262), the particle density remains positive for all (0, ) only if the
chemical potential is negative. A positive would also cause a divergence in the
integrals (2.269), (2.271). At large temperatures, the chemical potential has a large
negative value, which moves closer to zero as the temperature decreases (see Fig. 2.4,
compare also Fig. 2.1). The chemical potential vanishes at a critical temperature

132

2 Field Formulation of Many-Body Quantum Physics

Figure 2.4 Temperature behavior of the chemical potential of a free Bose gas.

Tc . From Eqs. (2.262), (2.269), and (2.275), this is determined by the equation
N
V

= g

d-3 p

1
p2

e( 2M )/kB T 1



2
V

,
I
= g 3
(T ) 3/2 kB T

=g

dg f
(2.329)

at = 0, where it yields, via Eq. (2.275), the particle density


N
V

= g

V
3(T )

(3/2).

(2.330)

Thus the critical temperature Tc satisfies the equation


2/3

Tc = [g(3/2)]

N
V

2/3

2h2
,
kB M

(2.331)

with (3/2) = 2.61238 . . . .


It is interesting to rewrite this equation in natural variables. We may introduce
an average distance of the bosons a by the relation N/V 1/a3 . There is an energy
associated with it
a

h
2
,
2Ma2

(2.332)

and a temperature Ta a /kB . In these natural units, the critical temperature


(2.331) of the free Bose gas is simply
Tc = [g(3/2)]2/3 4Ta .

(2.333)

We may rewrite Eq. (2.329) in the form


1=

T
Tc

3/2

I3/2 (
)
,
I3/2 (0)

.
kB T

(2.334)

133

2.14 Quantum Statistic of Free Nonrelativistic Fields

For
between 0 and 1, this equation yields T /Tc > 1, while Eq. (2.293) gives us
the associated specific heat. The result is displayed in Fig. 2.5. To understand the
figure we must realize what happens in the regime of low temperatures T < Tc where
(2.269) has no solution.
A glance at Eq. (2.330) shows that a phase transition takes place when the
average distance a between atoms becomes smaller than the De Broglie wavelength
of thermal motion (2.270). In natural units, this may be expressed as
1
(T ) =
4

Ta
.
T

(2.335)

For helium, this length scale has roughly the value 5.64
A.7

Figure 2.5 Temperature behavior of fraction of degenerate bosons in the zero-momentum


state of a free Bose gas.

For T < Tc , Eq. (2.329) has no solution even though the physical system can be
cooled further. The apparent contradiction has its origin in a failure of the integral
approximation (2.256) to the sum over momenta for T < Tc . The reason is that the
state with p = 0 is missed in the energy integral over all states in Eq. (2.269). To
avoid this we have to write more properly
N = Np=0 + Np6=0 Ns + gV

dg f |=0 .

(2.336)

where Ns Np=0 is the number of Bose particles at zero momentum and energy.
Below Tc , a finite fraction of all particles Ns /N accumulates in this single degenerate
state. It can be calculated from the modified Eq. (2.329):
N Ns = V g

dgf |=0

V
2
T
I3/2
= g 3
(0) = N
(T )
Tc


See p. 256 in the textbook Ref. [8].

3/2

(2.337)

134

2 Field Formulation of Many-Body Quantum Physics

Thus we find that the number of degenerate bosons has the temperature behavior
"

T
Ns = N 1
Tc


3/2 #

(2.338)

which is plotted in Fig. 2.5.

Figure 2.6 Temperature behavior of the specific heat of a free boson gas. For comparison
we show the specific heat of the strongly interacting Bose liquid 4 He, scaled down by a
factor 2 to match the Dulong-Petit limit of the free Bose gas.

The phenomenon of a macroscopic accumulation of particles in a single state


is called Bose condensation and plays a central role in the understanding of the
phenomenon of superfluidity in liquid helium consisting of the bosonic atoms 4 He.
In fact, the temperature Tc calculated from Eq. (2.331) is Tc 3.1 K which is roughly
of the same order as the experimental value8
Tcexp 2.18K.

(2.339)

The discrepancy is due to the strong interactions between the 4 He atoms in the
liquid state which have all been neglected in deriving Eq. (2.339).
There exists a phenomenological two-fluid description of superfluidity in which
the condensate of the p = 0 bosons is identified with the superfluid component of
the liquid. This is the reason why we have used the subscript s in (2.337). The
complementary piece
Nn = N Ns
(2.340)
is usually referred to as the normal component of the superfluid.
8

The mass density is 0.145 g/cm3 . With the mass of the Helium atoms being M4 He
4mp 4 1.6762 1024 g, this implies a volume per atom of V /N 46.2
A3 , and

135

2.14 Quantum Statistic of Free Nonrelativistic Fields

For T < Tc , the energy of the normal liquid is equal to the total energy. Using
Eq. (2.287) with = 0 we find
E = En

 3/2
I5/2
(0)
V
2
T

kB T I5/2 (0) = g
= g 3
NkB T
(T )
I3/2 (0)
Tc

T
g 0.7703 NkB T
Tc


3/2

Ec

T
Tc

3/2

(2.341)

where we have expressed In (0) via Eq. (2.275). From this energy we find the specific
heat below Tc :
CV

E
S
=
= T

T V,N
T V,N
5E
=
T 3/2 .
2T

(2.342)

Integrating this with respect to the temperature gives the entropy


S=

5E
,
3T

(2.343)

and the free energy F = E T S takes the simple form


2
F = E.
3

(2.344)

This is consistent with the general result Eq. (2.267) since for = 0, the grand
canonical free energy FG coincides with the free energy F [recall the Euler relation
(1.528)].
The special role of the Bose condensate lies in the fact that it provides the system
with a particle reservoir, with the relation Eq. (2.269) being replaced by Eq. (2.329)
and Eq. (2.337).
Inserting (2.341) into (2.342), the low-temperature behavior of the specific heat
becomes explicitly
CV =

E
5 (5/2)(5/2) T
= gkB N
T
2 (3/2)(3/2) Tc


3/2

gkB N 1.92567

T
Tc

3/2

(2.345)

Comparing (2.345) with (2.293) and using the fact that I1/2 (0) = , we see hat at
Tc the peak in (2.345) has the same maximal value as the T > Tc -solution (since
I1/2 (0) = ).
As T passes Tc , the chemical potential becomes negative. To calculate the be

havior of in this regime we use Eq. (2.334) replace I3/2


(
) by I3/2
(0) + I3/2
(
)
with

I3/2
(
)

dz z

1/2

1
.
z
z

e
1 e 1
1

(2.346)

136

2 Field Formulation of Many-Body Quantum Physics

This function receives its main contribution from z 0, where it can be approximated by9
Z

=
.
(2.347)
dz 1/2
I3/2 (
)

z (z
)
0
Hence we obtain from (2.334) the relation
T
1=
Tc


3/2

1 +

I3/2

kB T

I3/2
(0)

(2.348)

Inserting here the small-


behavior (2.347) we see that for T slightly above Tc , the
negative chemical potential becomes nonzero with a power
1

2 kB Tc [I3/2
(0)]2

"

T
Tc

3/2

#2

1 ,

(2.349)

where I3/2
(0) is given by Eq. (2.275).
Let us use this result to find the internal energy slightly above the critical temperature Tc . With the help of relation (2.267) we calculate the derivative of the
energy with respect to the chemical potential as

E
3 FG
3
=
= gN.


T,V
2 T,V
2

(2.350)

This allows us to find the internal energy slightly above the critical temperature Tc ,
where is small, as
3
E Ec + gN.
2
" 3/2
#2
3
T

= Ec 2 gNkB Tc [I3/2
(0)]2
1 ,
2
Tc

(2.351)

where Ec is the internal energy at the critical point defined in Eq. (2.341). Forming
the derivative with respect to temperature as in (2.342) we see that at Tc the specific
heat has a kink. Its slope jumps by
CV

27
kB
kB
[I3/2 (0)]2 gN
3.6658 gN ,
2
4
Tc
Tc

(2.352)

the slopes being below Tc from (2.341)


CV
T
9

5 3 I5/2 (0) gNkB


53
gNkB
gNkB

0.7703
2.8885
,
2 2 I3/2 (0) Tc
22
Tc
Tc

T Tc , (2.353)

In general, the small-


expansion of I (
) = (n) (e ) follows from the so-called Robinson
P

1
expansion: (e ) = (1 )

+ () + k=1 (
)k ( k)/k!, derived in Subsec. 2.15.6 of
the textbook Ref. [1].

2.14 Quantum Statistic of Free Nonrelativistic Fields

137

and above Tc from (2.351):


CV
T

27
5 3 I5/2 (0)
2 [I3/2
(0)]2
2 2 I3/2 (0) 4

gNkB
gNkB
0.7715
,
Tc
Tc

T > Tc .

(2.354)
Let us compare the behavior of the specific heat of the free Bose gas with the
experimental results for the Bose liquid 4 He (see Fig. 2.6). The latter also rises like
T 3 for small T , but it has a sharp singularity at T = Tc . Considering the crudeness of
the free-gas approximation the similarity of the curves is quite surprising, indicating
the physical relevance of the above idealized quantum-statistical description.

2.14.4

High Temperatures

At high temperatures, the particles are distributed over a large volume in phase
space, so that the occupation numbers of each energy level are very small. As
a consequence, the difference between bosons and fermions disappears, and the
distribution functions (2.263) become
f

F (T, ) e()/kB T ,

(2.355)

for either statistics. The high-temperature limit of the thermodynamic quantities


can therefore all be calculated from the fermion expressions. The corresponding
limit in the functions In (
) in Eq. (2.271) is
, for which we obtain
In (
)

z n ez = n!e .

(2.356)

Inserting this for n = 1/2 into (2.302), we find that for a fixed particle number N,
the chemical potential has the large-T behavior

3
T
log
kB T
2
TF

!2/3
3
.
4

(2.357)

In the same limit, the grand-canonical free energy (2.285) behaves like
FG (T, , V ) gN(T, ) kB T.

(2.358)

With the definition of the pressure (1.526), this is the equation of state for the ideal
gas. Using the relation (2.267), we obtain from this the internal energy at a fixed
particle number
3
(2.359)
E gN kB T.
2
This equation is a manifestation of the Dulong-Petit law: Each of the 3N degrees of
freedom of the system carries an internal energy kB T /2.

138

2 Field Formulation of Many-Body Quantum Physics

The corresponding specific heat per constant unit volume is


3
CVDP = gNkB .
2

(2.360)

[compare with the fermion formula (2.294)] The low-temperature behavior (2.345)
is related to this by the factor [compare with (2.326)]
CV

2.15

small T

CVDP

(5/2) T
5
(3/2) Tc


3/2

CVDP

T
1.2838
Tc


3/2

(2.361)

Noninteracting Bose Gas In Trap

In 1995, Bose-Einstein condensation was observed in a dilute gas in a way that fits
the above simple theoretical description [2]. When 87 Rb atoms were cooled down
in a magnetic trap to temperatures less than 170 nK, about 50 000 atoms were
observed to form a condensate, a kind of superatom. Recently, such condensates
have been set into rotation and shown to become perforated by vortex lines [3] just
like rotating superfluid helium II.

2.15.1

Bose Gas in Finite Box

Consider first the condensation process in a finite number N of bosons enclosed in


a large cubic box of size L. Then the sum momentum sum in Eq. (2.262) has to
be carried out over the discrete momentum vectors pn = h
(n1 , n2 , . . . , nD )/L with
ni = 1, 2, 3, . . . :
N=

X
VD box
1

(z)

,
2 /2M
D/2
D
p
n
le (h)
1
pn e

(2.362)

where z e is the so-called fugacity, This can be expressed in terms of the onedimensional auxiliary partition function of a particle in a one-dimensional box:
Z1 (b)

ebn

2 /2

n=1

b h2 2 /ML2 = le2 (h)/2L2 ,

(2.363)

in the form of a fugacity expansion


N=

X
VD box

(z)

Z1D (wb)z w ,
D/2
D
le (h)
w

(2.364)

where the function Z1 (b) is related to the elliptic theta function


3 (u, z) 1 + 2

n=1

z n cos 2nu

(2.365)

139

2.15 Noninteracting Bose Gas In Trap

by Z1 (b) = [3 (0, eb/2 ) 1]/2. The small-b behavior of this function is easily
calculated as follows. We rewrite the sum as a sum over integrals
3 (0, eb/2 ) =

ek

2 b/2

k=

m=

dk ek

2 b/2+2ikm

X
2
2 2
e2 m /b .
1+2
b
m=1

Thus, up to exponentially small corrections, we may replace 3 (0, eb/2 ) by

so that for small b (i.e., large L/ ):


Z1 (b) =

1
2
+ O(e2 /b ).
2b 2

(2.366)
q

2/b,

(2.367)

For large b, Z1 (b) falls exponentially fast to zero.


In the sum (2.362), the lowest energy level with p1,...,1 = h
(1, . . . , 1)/L plays a
special role. Its contribution to the total particle number is the number of particles
in the condensate:
1
zD
Ncond (T ) = Db/2
=
, zD eDb/2 .
(2.368)
e
1
1 zD

box
This number diverges for zD 1, where the box function D/2
(z) has a pole
1/(Db/2 ). This pole prevents from becoming exactly equal to Db/2 when
solving the equation (2.362) for the particle number in the box.
For a large but finite system near T = 0, almost all particles will go into the
condensate, so that Db/2 will be very small, of the order 1/N, but not zero. The
thermodynamic limit can be performed smoothly by defining a regularized function
box
D/2
(z) in which the lowest, singular term in the sum (2.362) is omitted. Let us
define the number of normal particles which have not condensed into the state of
zero momentum as Nn (T ) = N Ncond (T ). Then we can rewrite Eq. (2.362) as an
equation for the number of normal particles

Nn (T ) =

VD box
(z()),
D
le (h) D/2

(2.369)

which reads more explicitly


Nn (T ) = SD (zD )

w
[Z1D (wb)ewDb/2 1]zD
.

(2.370)

w=1

A would-be critical point may now be determined by setting here zD = 1 and


equating the resulting Nn with the total particle number N. If N is sufficiently
large, we need only the small-b limit of SD (1) which is calculated in Appendix 2B
[see Eq. (2B.14)], so that the associated temperature Tc(1) is determined from the
equation
s
3

3
N=
(2.371)
(3/2) + (1) log C3 bc + . . . ,
2bc
4bc

140

2 Field Formulation of Many-Body Quantum Physics

where C3 0.0186. In the thermodynamic limit, the critical temperature Tc(0) is


obtained by ignoring the second term, yielding
N=

(0)

2bc

(3/2),

(2.372)

in agreement with Eq. (2.330) for Tc , if we recall b from (2.363). Using this we
rewrite (2.371) as
!3/2
3
Tc(1)
+
log C3 b(0)
(2.373)
1
c .
(0)
2N 2b(0)
Tc
c
Expressing b(0)
c in terms of N from (2.372), this implies
Tc(1)
(0)

Tc

1
2
N 2/3
log
.
2/3 (3/2)N 1/3
C3 2/3 (3/2)

(2.374)

Experimentally, the temperature Tc(1) is not immediately accessible. What is


easy to find is the place where the condensate density has the largest curvature, i.e.,
where d3 Ncond /dT 3 = 0. The associated temperature Tcexp is larger than Tc(1) by a
factor 1 + O(1/N), so that it does not modify the leading finite-size correction to
order in 1/N 1/3
Alternatively we may use the phase space formula

X
dD p
1
Nn = d x
=
(2h)D e[p2 /2M +V (x)] 1 n=1
Z

X
1
dD x enV (x) ,
=
q
D
n=1
2h2 n/M

where the spatial integration produces a factor


side becomes again (T /Tc(0) )D N.

2.15.2

dD x

dD p n[p2 /2M +V (x)]


e
(2h)D
(2.375)

2/M 2 n so that the right-hand

Harmonic and General Power Trap

For a D-dimensional harmonic trap V = M 2 x2 /2, the critical temperature is


reached if Nn is equal to the total particle number N where
kB Tc(0)

"

N
=h

(D)

#1/D

(2.376)

This formula has a solution only for D > 1.


The equation (2.375) for the particle number can be easily calculated for a more
general trap where the potential has the anisotropic power behavior
D
|xi |
M 2 2X

V (x) =
2
ai
i=1

!p i

(2.377)

141

2.15 Noninteracting Bose Gas In Trap


h

where
is some frequency parameter and a
is the geometric average a D
i=1 ai
Inserting (2.377) into (2.375) we encounter a product of integrals
D Z
Y

i=1

dx enM

2a
2 (|x

p
i |/ai ) i /2

D
Y

i=1

ai
(1
2
(M
a
2 /2)1/pi

+ 1/pi ),

i1/D

(2.378)

so that the right-hand side of (2.375) becomes (T /Tc(0) )D N, with the critical temperature
kB Tc(0)

M a2
2
=
2

M a2
2

!D/D
"

is the dimensionless parameter


where D

N D/2
Q
D (1 + 1/pi )
(D)
i=1

#1/D

(2.379)

D
D X
1

D
+
.
2
i=1 pi

(2.380)

which takes over the role of D in the harmonic formula (2.376).


A harmonic
trap with different oscillator frequencies 1 , . . . , D along the D Cartesian axes,
= D, and formula
is a special case of (2.377) with pi 2, i2 =
2a
2 /a2i and D
(2.379) reduces to (2.376) with replaced by the geometric average of the frequencies

(1 D )1/D . The parameter a disappears from the formula. A free Bose gas
Q
D D
in a box of size VD = D
is described by (2.377) in the limit pi
i=1 (2ai ) = 2 a
= D/2. Then Eq. (2.379) reduces to
where D
kB Tc(0)

h2
N
=
2M a2 (D/2)
"

#2/D

N
2h2
=
M VD (D/2)
"

#2/D

(2.381)

in agreement with Eq. (2.330)


Another interesting limiting case is that of a box of length L = 2a1 in the xdirection with p1 = , and two different oscillators of frequency 2 and 3 in the
2
other two directions. To find Tc(0) for such a Bose gas we identify
2a
2 /a22,3 = 2,3
4
2
2 2
2
in the potential (2.377), so that
/a = 2 3 /a1 , and obtain
kB Tc(0)

2.15.3

h
=h

2M

!1/5 "

N
a1 (5/2)

#2/5

21 2
=h

L2

!1/5 "

N
(5/2)

#2/5

. (2.382)

Anharmonic Trap in Rotating Bose-Einstein Gas

Another interesting potential can be prepared in the laboratory by rotating a Bose


condensate [11] with an angular velocity around the z-axis. The vertical trapping
frequencies is z 2 11.0 Hz 0.58 nK, the horizontal one is 6 z .
The centrifugal forces create an additional repulsive harmonic potential, bringing
the rotating potential to the form
4
Mz2 2
r
2
V (x) =
z + 36r
+
2
2 2z

h
z
=
2

2
4
z2
r
r
+
36
+
2z
2z
2 4z

(2.383)

142

2 Field Formulation of Many-Body Quantum Physics

2
2
where r
= x2 + y 2 , 1 2 /
, 0.4, and z 3.245 m 1.42 103 K.
For > , turns negative and the potential takes the form of a Mexican hat as
2
shown in Fig. 17.2, with a circular minimum at rm
= 362z /. For large rotation
speed, the potential may be approximated by a circular harmonic well, so that we
may apply formula (2.382) with a1 = 2rm , to obtain the -independent critical
temperature

2 > 0

2 < 0

V (x)

V (x)

x2

x2
x1

x1

Figure 2.7 Rotating trap potential for 2 > 0 and 2 < 0, pictured for the case of two
components x1 , x2 . The right-hand figure looks like a Mexican hat or the bottom of a
champaign bottle..

kB Tc(0)

h
z

 1/5 "

N
(5/2)

#2/5

(2.384)

For = 0.4 and N = 300 000, this yields Tc 53nK.


q At the critical rotation speed = , the potential is purely quartic r =
(x2 +y 2). To estimate Tc(0) we approximate it for a moment by the slightly different
potential (2.377) with the powers p1 = 2, p2 = 4, p3 = 4, a1 = z , a2 = a3 =
z (/2)1/4 , so that formula (2.379) becomes
kB Tc(0)

=h
z

"

2
164 (5/4)

#1/5 "

N
(5/2)

#2/5

(2.385)

It is easy to change this result so that it holds for the potential r 4 = (x + y)4
rather than x4 + y 4. We use the semiclassical formula for the number of normal
particles in the form
Z
cl (E)
Nn =
dE E/k T
,
(2.386)
e B 1
Emin

143

2.16 Temperature Green Functions of Free Particles

where Emin is the classical ground state energy, and


M
cl (E) =
2h2


D/2

1
(D/2)

dD x [E V (x)]D/21

(2.387)

i the semiclassical density of states. For a harmonic trap, the spatial integral can
be done, after which the energy integral on the right-hand side of (2.386) yields
[kB T /h]D (D) = (T /Tc(0) )D N, thus leading back to (2.376). Hence the critical
temperature for the potential r 4 = (x + y)4 rather than x4 + y 4 is obtained by
multiplying the right-hand side of Eq. (2.386) for N by a factor
4

2 rdrdxdy er
3/2
R
=
.
4
4
dxdy ex y
[5/4]2
R

(2.388)

This factor arrives inversely in front of N in Eq. (2.389), so that we obtain the
critical temperature in the critically rotating Bose gas
kB Tc(0) = h
z

1/5 "

N
(5/2)

#2/5

(2.389)

The critical temperature at = is therefore by a factor 41/5 1.32 larger


than at infinite . Actually, this limit is somewhat academic in a semiclassical
approximation since the quantum nature of the oscillator should be accounted for.
For more details see Chapter 7 in the textbook [1].

2.16

Temperature Green Functions of Free Particles

As argued in Section 1.7, all properties of a system in thermodynamic equilibrium


are calculable by continuing the quantum theory to imaginary times t = i with
real . We shall see later that the calculation of small interaction effects to the freeparticle results presented in the last section can be done perturbatively. It involves
the analytically continued analog of the free-particle propagator (2.233).
In a grand-canonical ensemble, the relevant quantity is the so called finitetemperature Green function or finite-temperature propagator of the free particles:
G(x, ; x ) =

) (x , )
Tr eHG /kB T T (x,

i

Tr(eH G /kB T )

i

) (x , ) ,
= eFG /kB T Tr eHG /kB T T (x,
h

(2.390)

where T is the -ordering operator defined in complete analogy to the time-ordering


), (x, ) are defined in analogy to (2.132)
operator T in (2.231). The fields (x,
and (2.133) via an analytically continued time evolution operator as follows:
) = eH G /h ax eH G /h ,
(x,

(x, ) = eHG /h ax eHG h .

(2.391)
(2.392)

144

2 Field Formulation of Many-Body Quantum Physics

These satisfy canonical commutation relations at equal imaginary time analogous


to (2.204):
), (x
, )] = 0,
[(x,

[ (x, ), (x , )] = 0,
), (x , )] = (3) (x x )
[(x,

(2.393)
(2.394)
(2.395)

The time evolution of these field operators is governed by the grand-canonical


Hamiltonian
G = H
N
.
H
(2.396)

Note that while p is the Hermitian conjugate of p at = 0, this is no longer


true for 6= 0. The advantages of using these non-Hermitian fields will become
apparent later when we discuss perturbation theory.
When it comes to calculating physical observables, only Green functions with an
imaginary time variable in the interval
(0, h
/kB T ) ,

(2.397)

will be needed. In the present discussion, however, we consider the behavior for all
.
Differentiating (2.391) and (2.392) with respect to , we obtain the Heisenberg
equations
) = [H
)],
G , (x,
(x,
)].
G , (x,
(x, ) = [H

(2.398)
(2.399)

) and
Using the canonical field commutation relations (2.394), the fields (x,

(x, ) are seen to satisfy the analytically continued Schrodinger equations


h
2 2
) = 0,
h +
+ (x,
2M x
!

h
2 2
(x, ) h
+
x +
2M

(2.400)

= 0.

(2.401)

If we apply these differential operators to the Green function (2.390) we obtain


a nonzero result:
h
2 2
h +
+ G(x, ; x )
2M x
!

), (x , )
= eFG /kB T Tr eHG /kB T ( ) (x,

Using the commutation rule (2.395), this becomes

io

h
2 2
h +
+ G(x, ; x ) = ( ) (3) (x x ),
2M x
!

(2.402)

(2.403)

145

2.16 Temperature Green Functions of Free Particles

which is the general defining equation of a Green function [recall (1.314)].


The plane-wave solutions of Eqs. (2.400) and (2.401) are the analyticallycontinued versions of the plane-wave solutions (2.205), with particle energies
counted from the chemical potential rather than zero. The canonical field operators solving (2.400) and (2.401) have expansions (2.214). Explicitly:
) =
(x,
(x, ) =

d-3 p eipx/h(p) /h a
p ,

(2.404)

d-3 p eipx/h+(p) /h a
p

Inserting these into (2.390), we now calculate the Green function for >
G(x, ; x , ) = G(x x , )

) (x , )
= eFG /kB T Tr eHG /kB T (x,

= eFG /kB T

(2.405)

d-3 p d-3 p Tr(eHG /kB T a


p a
p )ei[pxp x ]/hi[(p) (p ) ]/h.

Now we observe that




eFG /kT Tr eHG /kB T a


p a
p =

Tr eHG /kB T a
p a
p
Tr(eH G /kB T )

is simply the average particle number of Eq. (2.263) for the energy (p):
f(p) =

1
e(p)/kB T

(2.406)

apart from a Dirac -function in the momenta (3) (p p ). Hence we find for > 0:
G(x, )

d-3 peipx/h(p) /h (1 f(p) ).

(2.407)

For < 0, the operator order is reversed, and we obtain directly


G(x, )

d-3 peipx/h(p) /h f(p) .

(2.408)

From these expressions we can derive an important property of the temperature


Green function. Using the identity
e/kB T f = 1 f ,

(2.409)

we see that G(x, ) satisfies the relation


G(x, ) = G(x, + h
/kB T ),

(h/kB T , h
/kB T ],

(2.410)

in particular G(x, 0) = G(x, h


/kB T ). As a consequence, G(x, ) has a Fourier
transform
kB T X im
e
G(x, m )
(2.411)
G(x, ) =
h
m

146

2 Field Formulation of Many-Body Quantum Physics

with the frequencies


(

2
m m =
h

m
m+

1
2

for bosons,
for fermions,

(2.412)

where m runs through all integer values m = 0, 1, 2, . . .. These are known as


Matsubara frequencies.
The Fourier components are given by the integrals
G(x, m ) =

h/kB T

d eim G(x, ).

(2.413)

The full Fourier representation in space and imaginary time reads


G(x, ) =

Z
kB T X -3 im +ipx/h
d pe
G(p, m )
h
m

(2.414)

with the components


G(p, m ) =

d3 x

h/kB T

d eim ipx/h G(x, )

n
o
h

(1 f(p) ) e[ihm (p)]/kB T 1 .


=
ihm (p)

(2.415)

Inserting (2.406) and the explicit form of the Matsubara frequencies (2.412), we
obtain
h

G(p, m ) =
.
(2.416)
ihm (p)

Due to a marvelous cancellation, this has become very simple. In fact, the result can
be obtained from the Fourier transform (2.248) of the quantum field theoretic Green
function G(p, E) in two steps: First we continue (2.248) analytically in the energy
E to the imaginary off-shell values E = ihm , and second we shift the single-particle
energy from (p) = p2 /2M to (p) = (p) , this being a trivial consequence of
G = H
N
instead of H.

the use of H
As a cross check, let us calculate G(x, ) for a very small negative = from
the Fourier representation (2.414) with the components (2.416) for bosons:
G(x, ) =

kB T X im
h

d-3 p eipx/h
.
e
h
m
im + (p)

(2.417)

The phase factor eim is necessary to ensure convergence of the otherwise logarithmically divergent sums. The sum can be performed by changing it into a contour
integral
kB T

X
m

im

1
kB T Z
1
ez
=
.
dz z/k T
B
im (p)
2i C e
1z

(2.418)

147

2.16 Temperature Green Functions of Free Particles

Figure 2.8 Contour C in the complex z-plane for evaluating the Matsubara sum (13.194).
The semicircles at infinity l and r do not contribute. After shrinking the contours, only
the pole on the right-hand side contributes via Cauchys residue theorem.

The contour of integration C encircles the imaginary z-axis in the positive sense,
thereby enclosing all integer or half-integer valued poles at z = im (see Fig. 2.8).
The upper signs on the right-hand side of (13.193) holds for bosons, the lower for
fermions.
We now close the two branches of the contour C by two semicircles l and r
at infinity. Both contribute nothing since the right-semicircle is suppressed by an
exponential factor ez/kB T , the left-hand by a factor ez . The two closed contours
are now shrunk to zero. There is a pole only on the right-hand side, at z = which
contributes, by Cauchys residue theorem,
kB T

eim

1
1
= (p)/k T
= f(p) ,
B
im (p)
e
1

(2.419)

which are the Bose and Fermi distribution functions, as implied by Eq. (2.408) for
small negative .
In the opposite limit = , the phase factor in the sum would be ein . In this
case, the sum is converted into a the contour integral
kB T

X
n

ein

1
kB T
=
in (p)
2i

dz

ez
ez/kB T

1
,
1z

(2.420)

from which we would find 1 f(p) , corresponding to Eq. (2.407) for small positive
.

148

2 Field Formulation of Many-Body Quantum Physics

While the phase factors eim are needed to make the logarithmically-divergent
sums converge, they become superfluous if the two sums are combined. Indeed,
adding the two sums yields
kB T

X
n

"

"

X
1
1
1
1
ein
= kB T
+ ein
+
e
in
in
in in
n
kB T X 2
=
= 1 2 f
2 + 2
h
m m
in

coth(h/kB T )
tanh(h/kB T )

for

bosons
fermions

(2.421)

The right-hand side is the thermal expecationn value of ap a


p + a
p a
p .

2.17

Calculating the Matsubara Sum via Poisson Formula

There exists another way of calculating the Matsubara sum in the finite-temperature
propagator (2.414). At very low temperature, the Matsubara frequencies m =
2mkB T /h or m = (2m + 1)kB T /h move infinitely close to each other, so that
the sum over m becomes an integral
Z

kB T X
dm

h
n=
2

(2.422)

for Bose and Fermi fields. The propagator (2.414) becomes therefore, with (2.416),
kB T
h

G(x, ) =

i
m + i(p)/h
m=
Z
i
dm
d-3 peim +ipx/h
.
2
m + i(p)/h
d-3 peim +ipx/h

(2.423)

The integral over m can be performed trivially with the help of the residue theorem,
as in Eq. (2.243), yielding
G(p, ) =

dm im
i
e
= ( )e(p) /h .
2
m + i(p)/h

(2.424)

For finite temperatures we make use of Poissons summation formula (1.212) to write

f (m) =

m=

e2in f (),

(2.425)

()n e2in f ().

(2.426)

n=

from which we derive

m=

f (m + 1/2) =

n=

149

2.18 Nonequilibrium Quantum Statistics

A direct application of this formula shows that


kB T
h

im

m=

X
i
=
m + i(p)/h n=

1
(1)n

( + nh)e(p)( +nh)/h .

(2.427)
Thus the finite-temperature Green function is obtained from the zero-temperature
one by making it periodic or antiperiodic by a simple sum over all periods, with
equal or alternating signs. This guarantees the property (2.410).
The sum over n is a geometric series in powers of e(p)nh , which can be performed trivially. For (0, h
), the Heaviside function forces the sum to run only
over positive n, so that we find
G(p, ) =

n=0

1
(1)n

( + nh)e(p) /h en(p) =

e(p) /h
,
1 e(p)

(2.428)

which can also be rewritten in terms of the Bose and Fermi distribution functions
(2.406) as
h

G(p, ) = e(p) /h 1 f(p) .

(2.429)

For a free particles with zero chemical potential where (p) = p2 /2M, the momentum integral can be done at zero temperature as in Eq. (2.239), and we obtain
the imaginary-time version of the Schrodinger propagator (2.241)
1
M (xx )2 /2
h2 ( )
G(x, ; x , ) = ( ) q
3e
2h2 ( )/M
= G(x x , ).

(2.430)

This Gaussian function coincides with the end-to-end distribution of a random walk
of length proportional to h
( ). Thus the quantum-mechanical propagator is a
complex version of a particle performing a random walk. The random walk is caused
by quantum fluctuations. This fluctuation picture of Schrodinger theory is exhibited
best in the path integral formulation of quantum mechanics.10 In the imaginarytime formulation of quantum field theory, we describe ensembles of particles. They
correspond therefore to ensembles of random walks of fixed length. For this reason,
nonrelativistic quantum field theories can be used efficiently to formulate theories
of fluctuating polymers. In this context, they are called disorder field theories.11 .

2.18

Nonequilibrium Quantum Statistics

The physical systems which can be described by the above imaginary-time Green
functions are quite limited. They must be in thermodynamic equilibrium, with a
10

See the textbook in Ref. [1].


For details see the textbooks quoted in in Ref. [1] and Ref. [8]. These will be discussed in
Chapter 20.2
11

150

2 Field Formulation of Many-Body Quantum Physics

constant temperature enforced by a thermal reservoir. Only then can a partition


function and a particle distribution be calculated from an analytic continuation
of quantum-mechanical time evolution amplitudes to an imaginary time tb ta =
ih/kB T . In this section we want to go beyond such equilibrium physics and extend
the path integral formalism to nonequilibrium time-dependent phenomena.

2.19

Linear Response and Time-Dependent Green


Functions for T =
/0

If the deviations of a quantum system from thermal equilibrium are small, the
easiest description of nonequilibrium phenomena proceeds via the theory of linear
response. In operator quantum mechanics, this theory is introduced as follows.

First, the system is assumed to have a time-independent Hamiltonian operator H.


The ground state is determined by the Schrodinger equation, evolving as a function
of time according to the equation

|S (t)i = eiHt |S (0)i

(2.431)

(in natural units with h


= 1, kB = 1). The subscript S denotes the Schrodinger
picture.
a time-dependent external
Next, the system is slightly disturbed by adding to H
interaction,
H
+H
ext (t),
H

(2.432)

iHt
|dist
UH (t)|S (0)i,
S (t)i = e

(2.433)

ext (t)UH (t),


iUH (t) = H
H

(2.434)

ext

ext (t) eiHt


H
H (t)eiHt .
H

(2.435)

ext (t) is assumed to set in at some time t0 , i.e., H


ext (t) vanishes identically
where H
for t < t0 . The disturbed Schrodinger ground state has the time dependence

where UH (t) is the time translation operator in the Heisenberg picture. It satisfies
the equation of motion

with12

H (t) is given by
To lowest-order perturbation theory, the operator U
UH (t) = 1 i
12

t0

ext
H
(t ) + .
dt H

(2.436)

ext
Note that after the replacements H H0 , HH
HIint , Eq. (2.434) coincides with the
equation for the time evolution operator in the interaction picture to appear in Section 9.1.3. In
contrast to that section, however, the present interaction is a nonpermanent artifact to be set equal
to zero at the end, and H is the complicated total Hamiltonian, not a simple free one. This is why
we do not speak of an interaction picture here.

2.19 Linear Response and Time-Dependent Green Functions for T 6= 0

151

In the sequel, we shall assume the onset of the disturbance to lie at t0 = .


whose Heisenberg
Consider an arbitrary time-independent Schrodinger observable O
representation has the time dependence
iHt

H (t) = eiHt
O
Oe
.

(2.437)

Its time-dependent expectation value in the disturbed state |dist


S (t)i is given by
iHt

iHt
dist

hdist
Oe
UH (t)|S (0)i
S (t)|O|S (t)i = hS (0)|UH (t)e

hS (0)| 1 + i


1i

ext (t ) + . . . O
H (t)
dt H
H


ext (t ) + . . . |S (0)i
dt H
H

H (t)|H i ihH |
= hH |O

(2.438)

H (t), H
ext (t ) |H i + . . . .
dt O
H
h

We have identified the time-independent Heisenberg state with the time-dependent


Schrodinger state at zero time in the usual manner, i.e., |H i |S (0)i. Thus the
deviates from equilibrium by
expectation value of O
dist
S (t)i hdist (t)|O(t)|

hS (t)|O|
S
S (t)i hS (t)|O(t)|S (t)i

= i

ext
H (t), H
H
dt hH | O
(t ) |H i.

(2.439)

If the left-hand side is transformed into the Heisenberg picture, it becomes


S (t)i = hH |O
H (t)|H i = hH | O
H (t)|H i,
hS (t)|O|
so that Eq. (2.439) takes the form
H (t)|H i = i
hH | O

H (t), H
ext (t ) |H i.
dt hH | O
H
h

(2.440)

H (t) and H
H (t ) in
It is useful to use the retarded Green function of the operators O
the state |H i
h


GR
OH (t, t ) (t t )hH | OH (t), HH (t ) |H i.

(2.441)

Then the deviation from equilibrium is given by the integral


H (t)|H i = i
hH | O

dt GR
OH (t, t ).

(2.442)

H (t) is capable of undergoing oscillations.


Suppose now that the observable O
H (t) will in general excite these oscillaThen an external disturbance coupled to O
tions. The simplest coupling is a linear one, with an interaction energy
ext (t) = O
H (t)j(t),
H

(2.443)

152

2 Field Formulation of Many-Body Quantum Physics

where j(t) is some external source. Inserting (2.443) into (2.442) yields the linearresponse formula
H (t)|H i = i
hH | O

dt GR
OO (t, t )j(t ),

(2.444)

where GR
OO is the retarded Green function of two operators O:


GR
OO (t, t ) = (t t )hH | OH (t), OH (t ) |H i.

(2.445)

At frequencies where the Fourier transform of GOO (t, t ) is singular, the slightest
disturbance causes a large response. This is the well-known resonance phenomenon
found in any oscillating system. Whenever the external frequency hits an eigenfrequency, the Fourier transform of the Green function diverges. Usually, the eigenfrequencies of a complicated N-body system are determined by calculating (2.445)
and by finding the singularities in .
It is easy to generalize this description to a thermal ensemble at a nonzero
temperature. The principal modification consists in the replacement of the ground
state expectation by the thermal average

Tr(eH/T O)

.
hOiT

Tr(eH/T
)

Using the free energy

F = T log Tr(eH/T ),
this can also be written as

T = eF/T Tr(eH/T

hOi
O).

(2.446)

must be replaced by H
N
and F by its grandIn a grand-canonical ensemble, H
canonical version FG (see Section 1.16). At finite temperatures, the linear-response
formula (2.444) becomes

hO(t)i
T = i

dt GR
OO (t, t )j(t ),

(2.447)

where GR
OO (t, t ) is the retarded Green function at nonzero temperature defined by
[recall (1.312)]

F/T
H (t), O
H (t )
GR
Tr eH/T O
OO (t, t ) GOO (t t ) (t t ) e

io

. (2.448)

i (t) for
In a realistic physical system, there are usually many observables, say O
H
i = 1, 2, . . . , l, which perform coupled oscillations. Then the relevant retarded Green
function is some l l matrix

F/T
i (t), O
H
GR
Tr eH/T O
(t )
ij (t, t ) Gij (t t ) (t t ) e
H

io

(2.449)

2.20 Spectral Representations of Green Functions for T 6= 0

153

After a Fourier transformation and diagonalization, the singularities of this matrix


render the important physical information on the resonance properties of the system.
The retarded Green function at T 6= 0 occupies an intermediate place between
the real-time Green function of field theories at T = 0, and the imaginary-time Green
function used before to describe thermal equilibria at T 6= 0. The Green function
(2.449) depends both on the real time and on the temperature via an imaginary
time.

2.20

Spectral Representations of Green Functions for T =


/0

The retarded Green functions are related to the imaginary-time Green functions of
1 ,
equilibrium physics by an analytic continuation. For two arbitrary operators O
H
2
, the latter is defined by the thermal average
O
H

1
2
H
H
G12 (, 0) G12 ( ) eF/T Tr eH/T T O
( )O
(0) ,

H ( ) is the imaginary-time Heisenberg operator


where O

(2.450)

H
H ( ) eH
Oe
.
O

(2.451)

To see the relation between G12 ( ) and the retarded Green function GR
12 (t), we take
1 2

a complete set of states |ni, insert them between the operators O , O , and expand
G12 ( ) for 0 into the spectral representation
G12 ( ) = eF/T

1|n ihn |O
2|ni.
eEn /T e(En En ) hn|O

n,n

(2.452)

Since G12 ( ) is periodic under + 1/T , its Fourier representation contains only
the discrete Matsubara frequencies m = 2mT :
Z

G12 (m ) =

1/T

0
F/T

= e

d eim G12 ( )
X

n,n

1|n ihn |O
2|ni
eEn /T 1 e(En En )/T hn|O


1
.
im En + En

(2.453)

The retarded Green function satisfies no periodic (or antiperiodic) boundary condition. It possesses Fourier components with all real frequencies :
GR
12 ()

= eF/T

it

dt e
Z

F/T

(t)e

dt eit

Xh

n,n

H/T

Tr e

1
2
H
H
O
(t), O
(0)

i 

1 |n ihn |O
2|ni
eEn /T ei(En En )t hn|O
2 |n ihn |O
1|ni . (2.454)
eEn /T ei(En En )t hn|O
i

154

2 Field Formulation of Many-Body Quantum Physics

In the second sum we exchange n and n and perform the integral, after having
attached to an infinitesimal positive-imaginary part i to ensure convergence The
result is
F/T
GR
12 () = e

n,n

1 |n ihn |O
2|ni
eEn /T 1 e(En En )/T hn|O
i

i
.
En + En + i

(2.455)

By comparing this with (2.453) we see that the thermal Green functions are obtained
from the retarded ones by replacing [12]
i
1

.
En + En + i
im En + En

(2.456)

i (which are not observable).


A similar procedure holds for fermion operators O
There are only two changes with respect to the boson case. First, in the Fourier
expansion of the imaginary-time Green functions, the bosonic Matsubara frequencies m in (2.453) become fermionic. Second, in the definition of the retarded Green
functions (2.449), the commutator is replaced by an anticommutator, i.e., the re i is defined by
tarded Green function of fermion operators O
H

GR
ij (t, t )

GR
ij (t

F/T

t ) (t t )e

H/T

Tr e

j
i
H
H
(t )
O
(t), O

i 
+

. (2.457)

These changes produce an opposite sign in front of the e(En En )/T -term in both of
the formulas (2.453) and (2.455). Apart from that, the relation between the two
Green functions is again given by the replacement rule (2.456).
At this point it is customary to introduce the spectral function
12 ( ) =

1 e /T eF/T

n,n

1 |n ihn |O
2|ni,
eEn /T 2( En + En )hn|O

(2.458)

where the upper and the lower sign hold for bosons and fermions, respectively. Under
an interchange of the two operators it behaves like
12 ( ) = 12 ( ).

(2.459)

Using this spectral function, we may rewrite the Fourier-transformed retarded and
thermal Green functions as the following spectral integrals:
GR
12 () =

G12 (m ) =

i
d
12 ( )
,
2
+ i

d
1
12 ( )
.
2
im

(2.460)
(2.461)

155

2.21 Other Important Green Functions

These equations show how the imaginary-time Green functions arise from the retarded Green functions by a simple analytic continuation in the complex frequency
plane to the discrete Matsubara frequencies, im . The inverse problem of reconstructing the retarded Green functions in the entire upper half-plane of from
the imaginary-time Green functions defined only at the Matsubara frequencies m is
not solvable in general but only if other information is available [13]. For instance,
the sum rules for canonical fields to be derived later in Eq. (2.496) with the ensuing
asymptotic condition (2.497) are sufficient to make the continuation unique [14].
Going back to the time variables t and , the Green functions are
GR
12 (t) = (t)
Z

G12 ( ) =

12 ( )ei t ,
2

(2.462)

X
d
1
eim
12 ( )T
.
2
im
m

(2.463)

The sum over even or odd Matsubara frequencies on the right-hand side of G12 ( )
was evaluated before [see Fig. 2.8] for bosons and fermions as
T

eim

eim

1
1
= Gp,e ( ) = e( 1/2T )
im
2 sin(/2T )

= e (1 + f )

(2.464)

and
T

1
1
= Ga,e ( ) = e( 1/2T )
im
2 cos(/2T )

= e (1 f ),

(2.465)

with the Bose and Fermi distribution functions (13.194)


f =

1
e/T

(2.466)

respectively.

2.21

Other Important Green Functions

In studying the dynamics of systems at finite temperature, several other Green


functions are useful whose spectral functions we shall now derive.
In complete analogy with the retarded Green functions for bosonic and fermionic
operators, we may introduce their counterparts, the so-called advanced Green functions (compare page 39)

GA
12 (t, t )

GA
12 (t

F/T

t ) = (t t)e

H/T

Tr e

1
2
H
H
O
(t), O
(t )

i 

.
(2.467)

156

2 Field Formulation of Many-Body Quantum Physics

Their Fourier transforms have the spectral representation


GA
12 ()

d
i
12 ( )
,
2
i

(2.468)

differing from the retarded case (2.460) only by the sign of the i-term. This makes
the Fourier transforms vanish for t > 0, so that the time-dependent Green function
has the spectral representation [compare (2.462)]
GA
12 (t) = (t)

d
12 ()eit .
2

(2.469)

By subtracting retarded and advanced Green functions, we obtain the thermal


expectation value of commutator or anticommutator:


1
2
H
H
C12 (t, t ) = eF/T Tr eH/T O
(t), O
(t )

i 

= GR
12 (t, t ) G12 (t, t ). (2.470)

Note the simple relations:

GR
12 (t, t ) = (t t )C12 (t, t ),

GA
12 (t, t ) = (t t)C12 (t, t ).

(2.471)
(2.472)

When inserting the spectral representations (2.460) and (2.469) of GR


12 (t) and
into (2.470), and using the identity (1.336),

GA
12 (t)

i
i

= 2
= 2( ),

+ i i
( )2 + 2

(2.473)

we obtain the spectral integral representation for the commutator function:13


C12 (t) =

d
12 ()eit .
2

(2.474)

Thus a knowledge of the commutator function C12 (t) determines directly the spectral
function 12 () by its Fourier components
C12 () = 12 ().

(2.475)

An important role in studying the dynamics of a system in a thermal environment


is played by the time-ordered Green functions. They are defined by

1 (t)O
2 (t ) .
G12 (t, t ) G12 (t t ) = eF/T Tr eH/T T O
H
H

13

(2.476)

Due to the relation (2.471), the same representation is found by dropping the factor (t) in
(2.462).

157

2.21 Other Important Green Functions

Inserting intermediate states as in (2.453) we find the spectral representation


Z

G12 () =

F/T

= e

2 (t)O
1 (0)
dt eit (t)eF/T Tr eH/T O
H
H

F/T

1 (t)O
2 (0)
dt eit (t) eF/T Tr eH/T O
H
H

dt eit

1|n ihn |O
2|ni
eEn /T ei(En En )t hn|O

2 |n ihn |O
1|ni .
eEn /T ei(En En )t hn|O

n,n

dt eit

n,n

(2.477)

Interchanging again n and n , this can be written in terms of the spectral function
(2.458) as
G12 () =

1
1
i
i
d
. (2.478)
12 ( )
+
/T
/T

2
1e
+ i 1 e
i
#

"

Let us also write down the spectral decomposition of a further operator expression complementary to C12 (t) of (2.470), in which boson or fermion fields appear
with the wrong commutator:

F/T

A12 (t t ) e

H/T

Tr e

1 (t), O
2 (t )
O
H
H

i 

(2.479)

1
2
This function characterizes the size of fluctuations of the operators OH
and OH
.
Inserting intermediate states, we find

A12 () =

= eF/T

it F/T

dt e e
Z

H/T

Tr e

Xh
it

dt e

n,n

1 (t), O
2 (0)
O
H
H

i 

1|n ihn |O
2|ni
eEn /T ei(En En )t hn|O

2 |n ihn |O
1|ni . (2.480)
eEn /T ei(En En )t hn|O
i

In the second sum we exchange n and n and perform the integral, which runs now
over the entire time interval and gives therefore a -function:
A12 () = eF/T

n,n

1 |n ihn |O
2|ni
eEn /T 1 e(En En )/T hn|O
h

2( En + En ).

(2.481)

In terms of the spectral function (2.458), this has the simple form
A12 () =

tanh1
12 ( ) 2( ) = tanh1
12 (). (2.482)
2
2T
2T

Thus the expectation value (2.479) of the wrong commutator has the time dependence
Z
i(tt )
d

12 () tanh1
e
.
(2.483)
A12 (t, t ) A12 (t t ) =
2T
2

158

2 Field Formulation of Many-Body Quantum Physics

There exists another way of writing the spectral representation of the various
A
Green functions. For retarded and advanced Green functions GR
12 , G12 , we decompose in the spectral representations (2.460) and (2.468) according to the rule (1.337):
i
P
=
i
i( ) ,
i



(2.484)

where P indicates principal value integration across the singularity, and write
GR,A
12 ()

=i

P
d
12 ( )
i( ) .
2



(2.485)

Inserting (2.484) into (2.478) we find the alternative representation of the timeordered Green function
G12 () = i

P
d
12 ( )
i tanh1
( ) . (2.486)

2T


The term proportional to ( ) in the spectral representation is commonly


referred to as the absorptive or dissipative part of the Green function. The first
term proportional to the principal value is called the dispersive or fluctuation part.
The relevance of the spectral function 12 ( ) in determining both the fluctuation part as well as the dissipative part of the time-ordered Green function is the
content of the important fluctuation-dissipation theorem. In more detail, this may
be restated as follows: The common spectral function 12 ( ) of the commutator
function in (2.474), the retarded Green function in (2.460), and the fluctuation part
of the time-ordered Green function in (2.486) determines, after being multiplied by
a factor tanh1 ( /2T ), the dissipative part of the time-ordered Green function in
Eq. (2.486).
A
The three Green functions iG12 (), iGR
12 (), and iG12 () have the same
real parts. By comparing Eqs. (2.460) and (2.461) we found that retarded and
advanced Green functions are simply related to the imaginary-time Green function
via an analytic continuation. The spectral decomposition (2.486) shows this is not
true for the time-ordered Green function, due to the extra factor tanh1 (/2T ) in
the absorptive term.
Another representation of the time-ordered Green is useful. It is obtained by
expressing tan1 in terms of the Bose and Fermi distribution functions (2.466) as
tan1 = 1 2f . Then we can decompose
G12 () =

2.22

d
i
12 ( )
2f ( ) .
2
+ i
"

(2.487)

Hermitian Adjoint Operators

1
2
H
H
If the two operators O
(t), O
(t) are Hermitian adjoint to each other,

2 (t) = [O
1 (t)] ,
O
H
H

(2.488)

2.23 Harmonic Oscillator Green Functions for T 6= 0

159

the spectral function (2.458) can be rewritten as

12 ( ) = (1 e /T )eF/T
X
1 (t)|n ik2 .
eEn /T 2( En + En )|hn|O

(2.489)

n,n

This shows that


12 ( ) 0
12 ( )

for bosons,
(2.490)

for fermions.

This property makes it possible to derive several useful inequalities between various
diagonal Green functions.
Under the condition (2.488), the expectation values of anticommutators and
commutators satisfy the time-reversal relations

GA
12 (t, t )
A12 (t, t )
C12 (t, t )
G12 (t, t )

=
=
=
=

GR
21 (t , t) ,
A21 (t , t) ,
C21 (t , t) .
G21 (t , t) .

(2.491)
(2.492)
(2.493)
(2.494)

Examples are the corresponding functions for creation and annihilation operators
which will be treated in detail below. More generally, this properties hold for any
1 (t) = p (t), O
2 (t) = (t) of a specific
interacting nonrelativistic particle fields O
H
H
p
momentum p.
Such operators satisfy, in addition, the canonical equal-time commutation rules
at each momentum
p (t), p (t) = 1

(2.495)

Using (2.470), (2.474) we derive from this spectral function sum rule:
Z

d
12 ( ) = 1.
2

(2.496)

For a canonical free field with 12 ( ) = 2( ), this sum rule is of course trivially
fulfilled. In general, the sum rule ensures the large- behavior of imaginary-time,
retarded, and advanced Green functions of canonically conjugate field operators to
be the same as for a free particle, i.e.,
G12 (m )

2.23

i
,
m

GA,R

12 ()

1
.

(2.497)

Harmonic Oscillator Green Functions for T =


/0

As an example, consider a single harmonic oscillator of frequency or, equivalently,


a free particle at a point in the second-quantized field formalism. We shall start
with the second representation.

160

2 Field Formulation of Many-Body Quantum Physics

2.23.1

Creation Annihilation Operators

1 (t) and O
2 (t) are the creation and annihilation operators in the
The operators O
H
H
Heisenberg picture
a
H (t) = a eit ,

a
H (t) = a
eit .

(2.498)

The eigenstates of the Hamiltonian operator







1
1 2

2 2

p + x =
a a
+ aa
= a
(2.499)
H=
2
2
2
are
1
|ni = (a )n |0i,
(2.500)
n!
with the eigenvalues En = (n 1/2) for n = 0, 1, 2, 3, . . . or n = 0, 1, if a
and a
commute or anticommute, respectively. In the second-quantized field interpretation
the energies are En = n and the final Green functions are the same. The spectral
2 = a
function 12 ( ) is trivial to calculate. The Schrodingeroperator O
can connect
1 = a
the state |ni only to hn + 1|, with the matrix element n + 1. The operator O
does the opposite. Hence we have
12 ( ) = 2( )(1 e/T )eF/T

,0
X

e(n1/2)/T (n + 1).

(2.501)

n=0

Now we make use of the explicit partition functions of the oscillator whose paths
satisfy periodic and antiperiodic boundary conditions:
F/T

Z e

,1
X

(n1/2)/T

n=0

[2 sinh(/2T )]1
2 cosh(/2T )

bosons
fermions

for

. (2.502)

These allow us to calculate the sums in (2.501) as follows

n=0
0
X

(n+1/2)/T

1

1 F/T 
(n + 1) = T
e
= 1 e/T
eF/T ,
+
2


e(n1/2)/T (n + 1) = e/2T = 1 + e/T

n=0

1

eF/T .

(2.503)

The spectral function 12 ( ) of the a single oscillator quantum of frequency is


therefore given by
12 ( ) = 2( ).

(2.504)

With it, the retarded and imaginary-time Green functions become

(tt )
GR
,
(t, t ) = (t t )e

G (, ) = T

eim (

m=

= e(

1 n
n

(2.505)
)

1
im
for


,
<

(2.506)

(2.507)

2.23 Harmonic Oscillator Green Functions for T 6= 0

161

with the average particle number f of (2.466). The commutation function, for
instance, is by (2.474) and (2.504):

C12 (t, t ) = ei(tt ) ,

(2.508)

and the correlation function of the wrong commutator is from (2.483) and (2.504):
A (t, t ) = tanh1

i(tt )
e
.
2T

(2.509)

Of course, these harmonic-oscillator expressions could have been obtained directly by starting from the defining operator equations. For example, the commutator function

C (t, t ) = eF/T Tr eH/T [aH (t), a


H (t )]

(2.510)

turn into (2.508) by using the commutation rule at different times


[aH (t), a
H (t )] = ei(tt ) ,

(2.511)

which follows from (2.498). Since the right-hand side is a c-number, the thermodynamic average is trivial:

eF/T Tr(eH/T ) = 1.

(2.512)

After this, the relations (2.471), (2.472) determine the retarded and advanced
Green functions

i(tt )
GR
,
(t t ) = (t t )e

i(tt )
GA
.
(t t ) = (t t)e

(2.513)

For the Green function at imaginary times

G (, ) eF/T Tr eH/T T a
H ( )aH ( ) ,

(2.514)

the expression (2.507) is found using [see (2.515)]

eH = a e ,
a
H ( ) eH a

eH = a
e ,
a
H ( ) eH a

(2.515)

and the summation formula (2.503).


The wrong commutator function (2.509) can, of course, be immediately derived
from the definition


H (t), a
H (t )
A12 (t t ) eF/T Tr eH/T a
and (2.498), by inserting intermediate states.

i 

(2.516)

162

2 Field Formulation of Many-Body Quantum Physics

For the temporal behavior of the time-ordered Green function we find from
(2.478)


G () = 1 e/T

1

/T
GR
() + 1 e

and from this by a Fourier transformation


G (t, t ) =
=

1 e/T

1

1

(t t )ei(tt ) 1 e/T
i

(t t ) (e/T 1)1 ei(tt )

GA
(),

(2.517)

1

(t t)ei(tt )
(2.518)

= [(t t ) f ] ei(tt ) .

The same result is easily obtained by directly evaluating the defining equation using
(2.498) and inserting intermediate states:

G (t, t ) G (t t ) = eF/T Tr eH/T T a


H (t)aH (t )

= (t t )ha a
iei(tt ) (t t)ha a
iei(tt )

= (t t )(1 f )ei(tt ) (t t)f ei(tt ) ,

(2.519)

which is the same as (2.518). For the correlation function with a and a interchanged,

(t, t ) G (t t ) = eF/T Tr eH/T


G
TaH (t)aH (t ) ,

we find in this way

(t, t ) = (t t )ha aiei(tt ) (t t)ha a iei(tt )


G

= (t t )f ei(tt ) (t t)(1 f )ei(tt ) ,

(2.520)

(2.521)

in agreement with (2.494).

2.23.2

Real Field Operators

From the above expressions it is easy to construct the corresponding Green functions
for the position operators of the harmonic oscillator x(t). It will be useful to keep
the discussion more general by admitting oscillators which are not necessarily mass
points in space but can be field variables. Thus we shall use, instead of x(t), the
symbol (t), and call this a field variable. We decompose the field as
x(t) =

i
h
h it
ae
+ a eit .
2M

(2.522)

In this section we use physical units. The commutator function (2.470) is directly
C(t, t ) h[(t),

(t
)] i =

2i sin (t t ),
2M

(2.523)

implying a spectral function [recall (2.474)]


( ) =

1
2 [( ) ( + )].
2M

(2.524)

2.23 Harmonic Oscillator Green Functions for T 6= 0

163

The real operator (t)


behaves like the difference of a particle of frequency and
, with an overall factor 1/2M. It is then easy to find the retarded and advanced
Green functions of the operators (t)
and (t
):
i
h

h
h R

G (t, t ) GR
(t t ) 2i sin (t t ), (2.525)
(t, t ) =
2M
2M
i
h
h A
h

G (t, t ) =
G (t, t ) G (t, t ) =
(t t ) 2i sin (t t). (2.526)
2M
2M

GR (t, t ) =

From the spectral representation (2.483), we obtain for the wrong commutator
A(t, t ) = h[(t),

(t
)] i =

coth1
2 cos (t t ).
2M
2kB T

(2.527)

The relation with (2.523) is again a manifestation of the fluctuation-dissipation


theorem (2.483).
The average of these two functions yields the time-dependent correlation function
at finite temperature containing only the product of the operators
GP (t, t ) h(t)
(t
)i =

[(1 2f ) cos (t t ) i sin (t t )] , (2.528)


2M

with the average particle number f of (2.466). In the limit of zero temperature
where f 0, this reduces to
GP (t, t ) = h(t)
(t
)i =

h
i(tt )
e
.
2M

(2.529)

The time-ordered Green function is obtained from this by the obvious relation
1
[A(t, t ) + (t t )C(t, t )] ,
2
(2.530)

where (t t ) is the step function of Eq. (1.322). Explicitly, the time-ordered


Green function is
G(t, t ) = (t t )GP (t, t ) (t t)GP (t , t) =

G(t, t ) hT (t)
(t
)i =

[(1 2f ) cos |t t | i sin |t t |] ,


2M

(2.531)

which reduces for T 0 to


G(t, t ) = hT (t)
(t
)i =

h
i|tt |
e
.
2M

(2.532)

Thus, as a mnemonic rule, a finite temperature is introduced into a zerotemperature Green function by simply multiplying the real part of the exponential
function by a factor 12f . This is another way of stating the fluctuation-dissipation
theorem.
There is another way of writing the time-ordered Green function (2.531) in the
bosonic case:

164

2 Field Formulation of Many-Body Quantum Physics

(h i|t t |)
cosh
h

2
.
(2.533)
G(t, t ) hT (t)
(t
)i =
h

2M
sinh
2

For t t > 0, this coincides precisely with the periodic Green function Gpe (, ) =
Gpe ( ) at imaginary-times > [see (2.428)], if and are continued analytically to it and it , respectively. Decomposing (2.531) into real and imaginary parts
we see by comparison with (2.530) that anticommutator and commutator functions
are the doubled real and imaginary parts of the time-ordered Green function:


A(t, t ) = 2 Re G(t, t ),

C(t, t ) = 2i Im G(t, t ).

(2.534)

In the fermionic case, the hyperbolic functions cosh and sinh in numerator and
denominator are simply interchanged, and the result coincides with the analytically
continued antiperiodic imaginary-time Green function [see again (2.428)].
The time-reversal properties (2.491)(2.494) of the Green functions become for
real fields (t):

GA (t, t )
A(t, t )
C(t, t )
G(t, t )

Appendix 2A

=
=
=
=

GR (t , t),
A(t , t),
C(t , t),
G(t , t).

(2.535)
(2.536)
(2.537)
(2.538)

Permutation Group and Representations on


n-Particle Wave Functions

A permutation of n particles is given by




1
2
...
n
P =
,
p(1) p(2) . . . p(n)

(2A.1)

where p(i) are all possible one-to-one mappings of the integers 1, 2, 3, . . . , n onto themselves. In the
notation (2A.1) the order of the columns is irrelevant, i.e., the same permutation can be written
in any other form in which the columns are interchanged with each other, for example,


2
1
...
n
P =
.
(2A.2)
p(2) p(1) . . . p(n)
Given n particles at positions x1 , x2 , . . . , xn , the permutation P may be taken to change the
position x1 to xp(1) , x2 to xp(2) , etc., i.e. we define P to act directly on the indices:
P xi xp(i) .

(2A.3)

Given an n-particle wave function (x1 , x2 , . . . , xn ; t), it behaves under P as follows:


P (x1 , x2 , . . ., xn ; t) =
=

(P x1 , P x2 , . . ., P xn ; t)
(xp(1) , xp(2) , . . ., xp(n) ; t).

(2A.4)

There exists a different but equivalent definition of permutations, call them P , in which the
variables x1 , x2 , . . ., xn of the wave functions are taken from their places 1, 2, . . ., n in the list of
arguments of (x1 , x2 , . . ., xn ; t) and moved to the positions p(1), p(2), . . ., p(n) in this list, i.e.,
P (x1 , x2 , . . ., xn ; t) (. . ., x1 , . . ., x2 , . . .; t)

(2A.5)

Appendix 2A

165

Permutation Group and Representations

where x1 is now at position p(1), x2 at position p(2), etc. The difference between the two definitions
is seen in the following example:


1 2 3
(x1 , x2 , x3 ; t) = (x2 , x3 , x1 ; t),
(2A.6)
2 3 1
to be compared with


1
2

2 3
3 1

(x1 , x2 , x3 ; t) = (x3 , x1 , x2 ; t).

(2A.7)

In the following we shall use only the first definition but all statements to be derived would hold
as well if we were to use the second one throughout the remainder of this appendix.
For n elements there are n! different permutations P . Given any two permutations


1
...
n
P =
p(1) . . . p(n)
and
Q=
a product is defined by rewriting Q as

Q=

1
...
n
q(1) . . . q(n)

p(1)
...
p(n)
q(p(1)) . . . q(p(n))

and setting:
QP

p(1)
q(p(1))

...
p(n)
. . . q(p(n))

1
q(p(1))

...
n
. . . q(p(n))




1
...
n
p(1) . . . p(n)

(2A.8)

Every element has an inverse since if we apply first P and multiply it by




p(1) . . . p(n)
1
P
,
1
...
n
the operation P 1 P returns all elements to their original places:


 
p(1) . . . p(n)
1
...
n
1
P 1 P =
=
1
...
n
p(1) . . . p(n)
1

... n
... n

(2A.9)

I.

(2A.10)

The right-hand side is defined as the identity permutation I.


It can easily be checked that for three permutations P QR, the product is associative:
P (QR) = (P Q)R

(2A.11)

Thus the n! permutations of n elements form a group, also called the symmetric group Sn .
If P is such that only two elements p(i) are different from i it can be written as


1 2 ...
i
...
j
... n
Tij =
(2A.12)
1 2 . . . p(i) . . . p(j) . . . n
and is called a transposition, also denoted in short by (i, j). Only the elements i and j are
interchanged. Every permutation can be decomposed into a product of transpositions. There are
many ways of doing this. However, each permutation decomposes either into an even or an odd

166

2 Field Formulation of Many-Body Quantum Physics

number of transpositions. Therefore each permutation can be characterized by this property. As


mentioned in the main text on p. 86, it is called the parity of the permutation. It is useful to
introduce a function




1
P =even
P =
for
(2A.13)
1
P =odd
which indicates the parity. This function satisfies the identity
P Q P Q ,

(2A.14)

since, if P and Q are decomposed into transpositions,


Y
T(ij) ,
n factors
P =

(2A.15)

(ij)

Q=

T(i j ) ,

m factors,

(i j )

then the product


PQ =

Y Y

T(ij) T(i j )

(2A.16)

(ij) (i j )

contains n m transpositions. This number is even for n, m both even or odd, and odd if one
of them is odd and the other even. Since the identity is trivially even, the inverse P 1 of a
permutation has the same parity as P itself, i.e.,
P 1 = P .

(2A.17)

Let us find the irreducible representations of the permutation group on the Hilbert space of
n-particle wave functions. The permutation on this space is defined by (2A.4). The irreducible
representations can be classified with the help of the so-called Young tableaux. These are arrays of
n boxes of the form
1 2 3 4 5 6
7 8 9

(2A.18)

n.
The boxes are filled successively with the numbers 1 to n. The numbers of boxes in the rows
m1 , m2 , m3 , . . . are ordered as
m1 m2 m3 . . .

n
X

mk = n.

k=1

Each tableaux defines a symmetry type of the wave functions on which the group of permutations is
represented irreducibly. The symmetry types are constructed as follows. Let pi be all permutations,
including the identity, of the numbers in the row i, and qj the corresponding operations in the
column j. Then define
Y X
P =
(
pi ),
Yi X
Q =
(
qj qj ),
j

where the product is taken over all rows i or all columns j, respectively, and q is the parity function
(2A.13). Now apply the operation QP to the indices of the wave function (x1 , x2 , . . ., xn ).

Appendix 2A

167

Permutation Group and Representations

As an example, take the Hilbert space of three-particle wave functions. There are three different
Yang tableaux corresponding to the following irreducible representations of S3 :
X
1 2 3
P =
p (sum over all 6 elements of S3 ), Q = 1
1 2
3
1
2
3

P = 1 + (1, 2),

P = 1, Q =

Q = 1 (1, 3)

q q (sum over all 6 elements of S3 )

The wave functions associated with these have the form


1 2 3 (x1 , x2 , x3 ) = (x1 , x2 , x3 ) + (x2 , x3 , x1 )
+ (x3 , x1 , x2 ) + (x1 , x3 , x2 )
+ (x2 , x1 , x3 ) + (x3 , x2 , x1 ),

(2A.19)

1 2
(x1 , x2 , x3 ) = (x1 , x2 , x3 ) + (x2 , x1 , x3 )
3
(x3 , x2 , x1 ) (x2 , x3 , x1 ),

(2A.20)

1
2 (x1 , x2 , x3 ) = (x1 , x2 , x3 ) + (x2 , x3 , x1 )
3

+ (x3 , x1 , x2 ) (x1 , x3 , x2 )

(x2 , x1 , x3 ) (x3 , x2 , x1 ).

(2A.21)

These wave functions are easily normalized

by dividing them by the square root of the number


of terms in each expression, i.e., 6, 2, 6, in the three cases. The horizontal array leads to the
completely symmetrized, the vertical array to the completely antisymmetrized wave function. The
second tableau results in a mixed symmetry.
Four-particle wave functions are classified with the following tableaux

1 2 3 4 +

1 2 3
+
4 ,

1 2
+
3 4

1 2
+
3
4

1
2
.
3
4

(2A.22)

Take, for instance,


1 2
3 4
which amounts to the operator QP with
P = (1 + (3, 4)) (1 + (1, 2)) ,

Q = (1 (2, 4)) (1 (1, 3)) ,

so that
P (x1 , x2 , x3 , x4 ) =

(x1 , x2 , x3 , x4 ) + (x2 , x1 , x3 , x4 )
+(x1 , x2 , x4 , x3 ) + (x2 , x1 , x4 , x3 ),

QP (x1 , x2 , x3 , x4 ) = (x1 , x2 , x3 , x4 ) (x3 , x2 , x1 , x4 )


(x1 , x4 , x3 , x2 ) + (x3 , x4 , x1 , x2 )

(2A.23)

168

2 Field Formulation of Many-Body Quantum Physics


+(x2 , x1 , x3 , x4 ) (x2 , x3 , x1 , x4 )
(x4 , x1 , x3 , x2 ) + (x4 , x3 , x1 , x2 )

+(x1 , x2 , x4 , x3 ) (x3 , x2 , x4 , x1 )
(x1 , x4 , x2 , x3 ) + (x3 , x4 , x2 , x1 )

+(x2 , x1 , x4 , x3 ) (x2 , x3 , x4 , x1 )
(x4 , x1 , x2 , x3 ) + (x4 , x3 , x2 , x1 ).
There exists an alternative but mathematically equivalent prescription of forming the wave functions of different symmetry types based on the permutations P introduced in (2A.5) which, instead
of performing the permutations on the indices exchanges the positions of the arguments in the wave
functions (x1 , . . ., xi , . . ., xj , . . ., xn ). As an example, take the tableaux
1 2
3
which gives with the alternative wave function
1 2
3

(x1 , x2 , x3 ) =

(x1 , x2 , x3 ) + (x2 , x1 , x3 )
(x3 , x2 , x1 ) (x3 , x1 , x2 )

rather than (2A.20).


The dimensionality with which the group elements of the permutation group are represented
for these symmetry classes are given by the formula
n!
,
i,j hij

d= Q

(2A.24)

where hij is the number of boxes to the right of the position ij plus the number of boxes below
the position ij in the tableaux, plus 1 for the box on the position ij itself.
As a useful check for the calculated dimensions we may use the dimensionality theorem, by
which the squares of the dimensions d of all inequivalent, irreducible, unitary representations add
up to the order of the group, here n!:
X
d2 = n! .
(2A.25)

This is a direct consequence of the great orthogonality theorem of such representations for any finite
group.
For the permutation group at hand, one has the additional property that the defining representation contains each irreducible representation with a multiplicity equal to its dimension.
For three particles, the symmetric and antisymmetric representations

and

are 3!/3 2 1 = 1-dimensional. The mixed representation

is 3!/3 1 1 = 2-dimensional. These dimensions fulfill the dimensionality formula (2A.25):


12 + 11 + 22 = 3! .

(2A.26)

Appendix 2B

Treatment of Singularities in Zeta-Function

169

Similarly one has for four particles the dimensionalities


4!
= 1,
4321
4!
= 3,
d=
4211

1 2 3 4

d=

1 2 3
4

4!
= 2,
3221

1 2
3 4

d=

1 2
3
4

4!
d=
= 3,
4121

1
2
3
4

d=

(2A.27)

4!
= 1.
4321

Again we check that these dimensions fulfill the dimensionality formula (2A.25):
12 + 32 + 22 + 32 + 12 = 4! .

Appendix 2B

(2A.28)

Treatment of Singularities in Zeta-Function

Here we show how to evaluate the sums which determine the would-be critical temperatures of a
Bose gas in a box and in a harmonic trap.

2B.1

Finite Box

According to Eqs. (2.363), (2.368), and (2.370), the relation between temperature T =
2 2 /bM L2kB and the fugacity zD at a fixed particle number N in a finite D-dimensional box is
h
determined by the equation
zD
N = Nn (T ) + Ncond (T ) = SD (zD ) +
,
(2B.1)
1 zD
where SD (zD ) is the subtracted infinite sum
SD (zD )

w
[Z1D (wb)ewDb/2 1]zD
,

(2B.2)

w=1

P
bk2 /2
containing the Dth power of one-particle partition function in the box Z1 (b) =
. The
k=1 e
would-be critical temperature is found by equating this sum at zD = 1 with the total particle
number N . We shall rewrite Z1 (b) as
h
i
(2B.3)
Z1 (b) = eb/2 1 + e3b/2 1 (b) ,
where 1 (b) is related to the elliptic theta function (2.365) by
1 (b)

k=2

e(k

4)b/2

i
e2b h
3 (0, eb/2 ) 1 2eb/2 .
2

According to Eq. (2.367), this has the small-b behavior


r
2b
1
1 (b) =
e e3b/2 e2b + . . . .
2b
2

(2B.4)

(2B.5)

170

2 Field Formulation of Many-Body Quantum Physics

The omitted terms are exponentially small as long as b < 1 [see the sum over m in Eq. (2.366)]. For
large b, these terms become important to ensure an exponentially fast falloff like e3b/2 . Inserting
(2B.3) into (2B.2), we find


X
D1 2
(D1)(D2) 3
1 (wb)e3wb/2 +
1 (wb)e6wb/2 +
1 (wb)e9wb/2 .
SD (1) D
2
6
w=1
(2B.6)
Inserting here the small-b expression (2B.5), we obtain
r


X
wb
wb 1 wb
e
e + e 1 + ... ,
(2B.7)
S2 (1)
2wb
2wb
4
w=1
!
r
r
3

3wb/2 1 3wb/2
3 3wb/2 3
3wb/2
S3 (1)
e
+
e
e
1 + . . . , (2B.8)
e

2wb
2 2wb
4 2wb
8
w=1
the dots indicating again exponentially small terms. The sums are convergent only for negative b,
this being a consequence of the approximate nature of these expressions. If we evaluate them in
this regime, the sums produce Polylogarithmic functions
(z)
and we find, using the property at the origin
(0) = 1/2,

X
zw
,
w
w=1

14

1
(0) = log 2,
2

P
1

1 = (0) = 1/2,
w=1 log w = (0) = 2 log 2, that
r

1
b
S2 (1) = 1 (e )
1/2 (eb ) + 0 (eb ) (0) + . . . .
2b
4
r
r 3
3
1
3

3b/2
3b/2
1 (e
)+
1 (e3b/2 ) 0 (e3b/2 ) (0) + . . . .
3/2 (e
)
S3 (1) =
2b
2 2b
4 2b
8

which imply

(2B.9)

(2B.10)

w=1

(2B.11)
(2B.12)

These expressions can now be expanded in powers of b with the help of the Robinson expansion
given in the footnote on p. 9. Afterwards, b is continued analytically to positive values and we
obtain
r

1
S2 (1) = log(C2 b)
(1/2) + (3 2) + O(b1/2 ),
(2B.13)
2b
2b
8
r
r 3
3
9
3

log(C3 b) +
(1/2)(1 + ) + (1 + ) + O(b1/2 ). (2B.14)
(3/2) +
S3 (1) =
2b
4b
4 2b
16

The constants C2,3 , C2,3


inside the logarithms turn out to be complex, implying that the limiting
expressions (2B.7) and (2B.8) cannot be used reliably. A proper way to proceed goes as follows.
We subtract from SD (1) terms which remove the small-b singularities by means of modifications
of (2B.7) and (2B.8) which have the same small-b expansion up to b0 :
r


X

4 3 wb

e
+ ... ,
(2B.15)

+
S2 (1)
2wb
2wb
4
w=1
"r
#
r
3

3
9
3

+ (1 + 2)
(1 + 2) e3wb/2 + . . . .
(2B.16)

S3 (1)
2wb
2
2wb
4
2wb
8
w=1
14

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 9.541.4.

Appendix 2B

Treatment of Singularities in Zeta-Function

171

In these expressions, the sums over w can be performed for positive b yielding
r

4 3
b

S2 (1)
1 (e )
1/2 (eb ) +
0 (eb ) + . . . ,
(2B.17)
2b
2b
4
r
r 3
3
9
3

1 (e3b/2 ) + (1+2)
1/2 (e3b/2 ) (1+2)0 (e3b/2 ) + . . . .
3/2 (e3b/2 )
S3 (1)
2b
2 2b
4
2b
8
(2B.18)
Inserting again the Robinson expansion in the footnote on p. 9, we obtain once more the above
expansions (2B.13) and (2B.14), but now with the well-determined real constants
C2 = e3/22+

0.8973,

3
C3 = e2+1/ 31/ 0.2630.
2

(2B.19)

The subtracted expressions SD (1) SD (1) are smooth near the origin, so that the leading small-b
behavior of the sums over these can simply be obtained from a numeric integral over w:
Z
Z
1.1050938
dw[S2 (1) S2 (1)] =
dw[S3 (1) S3 (1)] = 3.0441.
(2B.20)
,
b
0
0
These modify the constants C2,3 to
C2 = 1.8134,

C3 = 0.9574.

(2B.21)

The corrections to the sums over SD (1) SD (1) are of order b0 an higher and already included in
the expansions (2B.13) and (2B.14), which were only unreliable as far as C2,3 os concerned.
Let us calculate from (2B.14) the finite-size correction to the critical temperature by equat(0)
ing S3 (1) with N . Expressing this in terms of bc via (2.372), and introducing the ratio
bc bc /b(0)
which is close to unity, we obtain the expansion in powers of the small quantity
c
(0)
2/3
2bc / = [(3/2)/N ] :
s
p
(0)
(0)
bc
3
3
2bc
2b
c
3/2
(0)
b
=
1
+
log(C
b
b
)
+
(1/2)(1 + )bc + . . . .
(2B.22)
3 c
c
c
2(3/2)
4(3/2)
To lowest order, the solution is simply
s
(0)
1
bc = 1 + 2bc
log(C3 b(0)
c ) + ... ,
(3/2)

(2B.23)

yielding the would-be critical temperature to first in 1/N 1/3 as stated in (2.374). To next order,
we insert into the last term the zero-order solution bc 1, and in the second term the first-order
solution (2B.23) to find
s
(0)
1
b3/2 = 1 + 3 2bc
log(C3 b(0)
c )
c
2
(3/2)

(0)
i
2bc
3
1 h
2
(0)
(0)
+
(1/2)(1 + ) +
2 log(C3 bc ) + log (C3 bc ) + . . . .
(2B.24)
4(3/2)
(3/2)
(0)

2/3

(0)

Replacing bc by (2/) [(3/2)/N ] , this gives us the ratio (Tc /Tc )3/2 between finite- and
(0)
infinite size critical temperatures Tc and Tc . The first and second-order corrections are plotted
in Fig. 2.9, together with precise results from numeric solution of the equation N = S3 (1).

172

2 Field Formulation of Many-Body Quantum Physics


1.4
(0)

Tc /Tc

1.3
1.2
1.1

0.05

0.1

0.15 1/N 1/3

Figure 2.9 Finite-size corrections to critical temperature for N = 300 to infinity calculated once from the formula N = S3 (1) (solid curve) and once from the expansion (2B.24)
(0)
(0)
(short-dashed up to order [bc ]1/2 1/N 1/3 , long-dashed up to the order bc 1/N 2/3 ).
The fat dots show the peaks in the second derivative d2 Ncond (T )/dT 2 . The small dots
show the corresponding values for canonical ensembles, for comparison.

2B.2

Harmonic Trap

The sum relevant for the would-be phase transition in a harmonic trap is
#
"

X
1
w
1 zD
,
SD (b, zD ) =
D
wb
)
w=1 (1e

(2B.25)

which determines the number of normal particles in the harmonic trap via
Nn (T ) = Z ()D (h; zD ) SD (h, zD ).

(2B.26)

We consider only the point zD = 1 which determines the critical temperature by the condition
Nn = N . Restricting ourselves to the physical cases D = 1, 2, 3, we rewrite the sum as



X
1
(D 1) 2wb (D 1)(D 2) 3wb
wb
.
(2B.27)
e
+
e
SD (b, 1) =
D e

2
6
(1 ewb )D
w=1
According to the method developed in the evaluation of the Robinson expansion in the footnote
on p. 9, we obtain such a sum in two steps. First we go to small b where the sum reduces to an
integral over w. After this we calculate the difference between sum and integral by a naive power
series expansion.
As it stands, the sum (2B.27) cannot be converted into an integral due to singularities at w = 0.
These have to be first removed by subtractions. Thus we decompose SD (b, 1) into a subtracted
sum plus a remainder as
D
D(3D 1)
SD (b, 1) = SD (b, 1) + D SD (b, 1) + b D1 SD (b, 1) + b2
D2 SD (b, 1),(2B.28)
2
24
where
SD (b, 1) =

w=1

D e

wb

D 1 2wb (D 1)(D 2) 3wb


e
+
e

2
6

1
D
D(3D 1)
1
D D

(1 ewb )D
w b
2wD1 bD1
24wD2 bD2

(2B.29)

is the subtracted sum and




D
D1
(D 1)(D 2)
b
2b
3b
D SD (b, 1) D D (e )
D (e ) +
D (e )
b
2
6

(2B.30)

Appendix 2B

Treatment of Singularities in Zeta-Function

173

collects the remainders. The subtracted sum can now be done in the limit of small b as an integral
over w, using the well-known integral formula for the Beta function:
Z
eax
(a)(1 b)
dx
= B(a, 1 b) =
.
(2B.31)
x )b
(1

e
(1 + a b)
0
This yields the small-b contributions to the subtracted sums


7
1

s1 ,
S1 (b, 1)
b0
b
12


1
9
S2 (b, 1)
+log 2
s2 ,
b0
b
8


19
1
+log 3
s3 ,
S3 (b, 1)
b0
b
24

(2B.32)

where = 0.5772 . . . is the Euler-Mascheroni number . The remaining sum-minus-integral is obtained by a series expansion of 1/(1 ewb )D in powers of b and performing the sums over w using
formula
Z ! nh

X
X
e
1

=
(h)k ( k) (eh ).
(2B.33)

n
k!
0
n=1
k=1

However, due to the subtractions, the corrections are all small of order (1/bD )O(b3 ), and will be
ignored here. Thus we obtain
SD (b, 1) =

1
sD
+ SD (b, 1) + D O(b3 ).
D
b
b

(2B.34)

We now expand D SD (b, 1) using Robinsons formula in the footnote on p. 9 up to b2 /bD and
find
1
D S1 (b, 1) =
D (eb ),
b

1 
D S2 (b, 1) =
2D (eb ) D (e2b ) ,
(2B.35)
b2

1 
(2B.36)
3D (eb ) 3D (e2b ) + D (e3b ) ,
D S3 (b, 1) =
b3
where

b
b2
1 (eb ) = log 1 eb = log b +
+ ... ,
2 24
2
b
+ ... ,
(2B.37)
2 (eb ) = (2) + b(log b 1)
4


b
3
b2
3 (eb ) = (3) (2)
log b
+ ... ,
(2B.38)
6
2
2
The results are
S1 (b, 1) =
S2 (b, 1) =
S3 (b, 1) =

1
b
1
b2
1
b3


b
b2
( log b + ) +
+ ... ,
4 144




7b2
1
+
(2) b log b +
+ ... ,
2
24




3b
19
2
+ ... .
(3) + (2) b log b +
2
24

(2B.39)

Note that the calculation cannot be shortened by simply expanding the factor 1/(1 ewb )D
in the unsubtracted sum (2B.27) in powers of w, which would yield the result (2B.28) without the
first term S1 (b, 1), and thus without the integrals (2B.32).

174

2 Field Formulation of Many-Body Quantum Physics

Notes and References


For the second quantization in many-body physics and applications see
A.A. Abrikosov, L.P. Gorkov, and I.E. Dzyaloshinski, Methods of Quantum Field Theory in Statistical Physics, Dover, 1963; A. Fetter and J.D. Walecka, Quantum Theory of Many Particle
Systems, McGraw-Hill, New York, 1971.
Good textbooks on statistical physics are
L.D. Landau and E.M. Lifshitz, Statistical Physics, Pergamon, New York, 1958;
R. Kubo, Statistical Mechanics, North-Holland, Amsterdam, 1971;
K. Huang, Statistical Mechanics, Wiley, New York, 1987.
For representations of finite groups and, in particular, the permutation group see
H. Boerner, Darstellungen von Gruppen, mit Ber
ucksichtigung der Bed
urfnisse der modernen
Physik , Springer, Berlin, 1055;
A.O. Barut and R. Raczka, Theory of Group Representations and Applications, World Scientific,
Singapore, 1986;
M. Hamermesh, Group Theory and its Application to Physical Problems, Dover, New York, 1989;
The individual citations refer to:
[1] H. Kleinert, Path Integrals in Quantum Mechanics Statistics, Polymer Physics, and Financial Markets, World Scientific, Singapore 2008 (www.physik.fu-berlin.de/~kleinert/b5).
[2] H. Kleinert and V. Schulte-Frohlinde, Critical Phenomena in 4 -Theory, World Scientific,
Singapore, 2001 (ibid.http/b8).
[3] T.D. Lee,K. Huang, and C.N. Yang, Phys. Rev. 106, 1135 (1957).
[4] U.C. T
auber and D.R. Nelson, Phys. Rep. 289, 157 (1997).
[5] O. Penrose and L. Onsager, Phys. Rev. 104, 576 (1956).
[6] P. Nozi`eres and D. Pines, The Theory of Quantum Liquids (Addison-Wesley, New York,
1990), Vol. II.
[7] In Wolframs program MATHEMATICA, this function denoted by PolyLog[n, z] or Lin (z).
See http://mathworld.wolfram.com/Polylogarithm.html. We prefer the notation n (z)
to emphasize that n (1) = (n) is Riemanns zeta-function. The properties of n (z) are
discussed in detail in Section 7.2 of the textbook Ref. [1].
[8] H. Kleinert, Gauge Fields in Condensed Matter , Vol. I, World Scientific, 1989
(http://www.physik.fu-berlin.de/~kleinert/b1).
[9] The first observation was made at JILA with 87 Ru:
M.H. Anderson, J.R. Ensher, M.R. Matthews, C.E. Wieman, and E.A. Cornell, Science
269, 198 (1995).
It was followed by a condensate of 7 Li at Rice University:
C.C. Bradley, C.A. Sackett, J.J. Tollet, and R.G. Hulet, Phys. Rev. Lett. 75, 1687 (1995),
and in 30 Na at MIT:
K.B. Davis, M.-O. Mewes, M.R. Andrews, and N.J. van Druten, D.S. Durfee, D.M. Kurn,
W. Ketterle, Phys. Rev. Lett. 75, 3969 (1995).
[10] J.R. Abo-Shaeer, C. Raman, J.M. Vogels, and W. Ketterle, Science 292, 476 (2001).
[11] V. Bretin, V.S. Stock, Y. Seurin, F. Chevy, and J. Dalibard, Fast Rotation of a BoseEinstein Condensate, Phys. Rev. Lett. 92, 050403 (2004).

Notes and References

175

[12] Some authors define G12 ( ) as having an extra minus sign and the retarded Green function
with a factor i, so that the relation is more direct: GR
12 () = G12 (m = i + ). See
A.A. Abrikosov, L.P. Gorkov, and I.E. Dzyaloshinski, Sov. Phys. JETP 9, 636 (1959); or
Methods of Quantum Field Theory in Statistical Physics, Dover, New York, 1975; also
A.L. Fetter and J.D. Walecka, Quantum Theory of Many-Particle Systems, McGraw-Hill,
New York, 1971.
[13] E.S. Fradkin, The Greens Function Method in Quantum Statistics, Sov. Phys. JETP 9, 912
(1959).
[14] G. Baym and D. Mermin, J. Math. Phys. 2, 232 (1961).
One extrapolation uses Pade approximations:
H.J. Vidberg and J.W. Serene, J. Low Temp. Phys. 29, 179 (1977); W.H. Press, S.A.
Teukolsky, W.T. Vetterling, and B.P. Flannery, Numerical Recipes in Fortran, Cambridge
Univ. Press (1992), Chapter 12.5.
Since the thermal Green function are usually known only approximately, the continuation
is not unique. A maximal-entropy method by
R.N. Silver, D.S. Sivia, and J.E. Gubernatis, Phys. Rev. B 41, 2380 (1990)
selects the most reliable result.

Every existing thing is born without reason,


prolongs itself out of weakness,
and dies by chance
Jean-Paul Sartre (1905-1980)

3
Interacting Non-Relativistic Particles
If the particles in a nonrelativistic many-body system are no longer free, quantum
field theory turns from a trivial to a very difficult subject. Solvable models exist only
in certain limiting situations, in particular in a reduced number 1 + 1 of spacetime
dimensions. Some of them have given useful insights into the behavior of certain
quasi-two-dimensional statistical systems.
In general, the physical properties of an interacting field system can be calculated
only on the basis of some simple and crude approximations. Among these, the
most common consists in an expansion of the interacting theory in powers of the
interaction strength, say g (see Chapter 8). Such a weak-coupling expansion has
at first many mathematical problems since the underlying function is usually not
analytic at g = 0 but possesses a branch cut starting at this point. The radius of
convergence is therefore zero. Thus the expansion can at best be of an asymptotic
nature, in which only first few terms can yield useful approximations to the result,
and higher orders improve the result only if the interaction strength is very small.
If carried to high orders, an asymptotic power series always diverges.
There exist, however, s simple procedure to reconstruct approximately such functions with a divergent weak-coupling expansion. This is done by variational perturbation theory, which turns divergent weak-coupling expansions into convergent
strong-coupling expansions.1 where the powers are not g n but (1/g )n . These can
be evaluated with any desired accuracy.
Even before such mathematical methods were invented, the perturbative approach has led to remarkable results in quantum electrodynamics (QED), the quantum field theory of electrons and phonons. There, the expansion parameter is the
fine-structure constant = e2 /hc = 1/137 and numerical results obtained to third
order in do agree with experiment up to an amazing accuracy of one part in a
million.
Before developing the general perturbation expansion of quantum field theory
it is useful to treat two important weakly interacting quantum systems in a more
conventional way, using only the methods of old fashioned quantum mechanical
perturbation theory. In this way we shall appreciate more the beautiful systematics
which the general quantum field theoretic methods will have to offer in treating such
problems.
1

For details see Chapter 5 of Ref. [4] and Chapter 19 of Ref. [5].

176

177

3.1 Weakly Interacting Bose Gas

3.1

Weakly Interacting Bose Gas

If the particles in a Bose Gas interact with each other, the full power of quantum
field theory is needed to understand its behavior. For weak and short-range interactions, the Bose condensation in the limit of low temperature and particle, low
density can be studied by a very simple procedure due to Bogoliubov, Huang, Yang
and Luttinger, and by Brueckner and Sawada. Until now, the condensation process
in this limit has not yet been observed experimentally but in recent years, proposals
have been advanced how to do that. There is now a high certainty that a gas of
hydrogen atoms with parallel polarized electronic spins has only repulsive interactions and remains a gas down to zero temperature. Since the atoms are bosons they
should show Bose condensation.
For a while it was hoped that superfluid 4 He poured into a container filled with
a very fine glass powder could act like a weakly interacting Bose gas, due to the
dilution effect of the powder. The grains of the powder have to be smaller than
the scale over which the wave function of the condensate acts coherently. But the
condensation process observed in such a system is now generally explained by the
superfluid transition in the ensemble of very small micrograin fluid elements of 4 He.
The condition of being dilute is that the scattering length of the particles should
be much smaller than the average Compton wave length h
/hpi associated with the
momentum distribution in the degenerate Bose gas calculated before. Then it is
possible to include only the s-wave in the scattering process. Accordingly, the
interaction can be approximated by a -function repulsion which acts in an s-wave.
The time-independent second quantized grand-canonical energy has a freeparticle term
H0 =

x ax

1 2
ax .
p
2M


(3.1)

The interactions has the form (2.128):


Hint =

1Z 3 Z 3
d x d x ax ax V (x x )ax ax .
2

(3.2)

We have seen in the discussion of the degenerate non-interacting Bose gas that the
state of momentum p = 0 plays a special role, and we expect the same thing to
happen here.
Consider the system in a finite volume V , such that the momentum states are
discrete pi = (2/L)ni , and we may decompose as in momentum space as
1 X ipx
e ap
ax =
V p

(3.3)

The the free-particle Hamiltonian (3.1) becomes


H0

X
p

ap (p )ap =

X
p

ap p ap ,

(3.4)

178

3 Interacting Non-Relativistic Particles

where
p

p2
,
2M

(3.5)

are the single-particle energies and p p the relevant energies (2.255) in a


grand-canonical ensemble. Now we use the Fourier representation of the potential
V (x) =

1 X iqx/h
d3 q iqx/h
e
V
(q)
=
e
V (q),
(2h)3
V q

(3.6)

to rewrite the interaction Hamiltonian as


Z
X
g
3
Hint =
d
x
eiqx/h eip3 x/h eip4 x/h eip2 x/h eip1 x/h V (q)ap4 ap3 ap2 ap1. (3.7)
2V 3
p1 ,p2 ,p3 ,p4 ,q
The integral over x yields a (3) -function for overall momentum conservation
(2)3 (3) (p3 + p4 p1 p2 ), which for the discrete momenta is equivalent to
V p3 +p4 ,p1 +p2 . This can easily be verified by recalling relation (1.188) and summing
the latter over p1 , say, while integrating the former with the phase-space measure
R 3
d pV /(2)3 . Hence we obtain
Hint =

1 X
V (q)ap+q ap q ap ap .
2V p,p ,q

(3.8)

If we want to include only s-wave scattering, we restrict the potential to the simple
repulsive -function potential
V (x x ) = g (3) (x x ).

(3.9)

Then we can write Hint as follows


Hint =

g X
ap+q ap q ap ap .
2V p,p ,q

(3.10)

In scattering experiments at low momentum, one measures the phase shift as of


s-wave scattering. This shows up in a limiting scattering amplitude
f = a.

(3.11)

Its square is the differential cross section, making the total cross section in the limit
of small momentum equal to 4a2 . The phase shift is related to the renormalized
version gR of the coupling contstant g by
as =

M
gR .
4h2

(3.12)

This will be dicussed in detail in Chapter 9. The renormalized coupling constant gR


will be defined in Eq. (3.58).

179

3.1 Weakly Interacting Bose Gas

At very low temperature, the state with p = 0 will contain a macroscopic number
of particles N0 which is much larger than the number of excited particles N N0 .
We separate out the terms containing the zero modes and rewrite the interaction
Hamiltonian as
0

Hint = Hint
+ Hint
+ Hint
,
(3.13)
with
0
Hint
=

g
2V

2
a2
0 a0

X

2ap a0 ap a0

p6=0

2ap a0 ap a0

ap ap a0 a0

a0 a0 ap ap



(3.14)

and the interaction Hamiltonians


X
g

Hint
=
ap+q ap q ap ap ,
2V p6=0,p 6=0,q6=0

(3.15)

and

Hint
=


g X
ap+p ap ap + c.c. .
a0
V p,p

(3.16)

0
In the Hamiltonian Hint
we include the excited particles with p 6= 0 only up to
quadratic order and keep the full quartic order only for the operators a0 , a0 . The
cubic terms in a0 , a0 which are linear in ap , ap with p 6= 0 vanish by momentum

conservation. The second interaction Hamiltonian Hint


does not conserve the particle
number and contributes only in second-order perturbation theory.
Since the particle number N0 is very large, the harmonic oscillator associated with the term a0 , a0 in the Hamilton operator behaves almost classically,
so that
can consider a0 , a0 approximately as c-numbers and replace them by
we i/2
0
A0 N0 e
with some unknown phase . Then Hint
becomes approximately
0
Hint



X
X
g 2 2
ap ap + ap ap +
A0 A0 + 2A0 A0
=
A20 ap ap + h.c. .(3.17)
2V
p6=0
p6=0

The total particle number N is given by the operator

N = A0 A0 + Nu = N0 + Nu ,
where
Nu =


X
1 X
ap ap + ap ap =
ap ap .
2 p6=0
p6=0

(3.18)
(3.19)

counts the uncondensed particles in the nonzero momentum states. Following Bogoliubov we consider a system with a fixed total number of particles and forget about
the chemical potential. Thus we work with the free Hamiltonian
H0 =

X
p

p ap ap ,

(3.20)

180

3 Interacting Non-Relativistic Particles

and rewrite the interaction Hamiltonian (3.17) as


0
Hint

g 2
A20 ap ap
ap ap +ap ap +
N + A0 A0
2V
p6=0
p6=0
X

X


+ h.c. + H

int .

(3.21)

The last term

g
Nu 2 ,
(3.22)
2V
is of third order in g, and is neglected in Bogoliubovs weak-coupling theory. Thus
we have to deal with the approximate free Hamiltonian

Hint
=

H0

g
1X g 2
g 2 X
p + A0 A0 ap ap +
N +
A a a + h.c. .

2V
V
2 p6=0 V 0 p p
p6=0


(3.23)

At this point, Bogoliubov introduces particle and condensate densities and


0 :

N
,
V

Ns
.
V

(3.24)

Then he chooses A0 to have a real phase, so that he may identify


A0 = A0 =

V 0 ,

(3.25)

and bring (3.23) to the form


H0 gV


X
2 X
1
+ (p + g0 ) ap ap + g0
ap ap + h.c. ,
2 p6=0
2
p6=0

(3.26)

This operator is quadratic in the creation and annihilation operators, and can therefore be diagonalized by means of a canonical transformation. Following Bogoliubovs
original treatment, we introduces the new creation and annihilation operators p
and p as follows

ap = cosh p p + sinh p p
,

ap = cosh p p + sinh p p .

(3.27)

More compactly, we shall write

ap = up p + vp p
,

ap = up p + vp p ,

with

u2p vp2 = 1.

(3.28)

For the original particle operators this implies

ap ap = (upp + vp p )(upp + vp p
)

= vp2 + u2p p p + vp2 p


p + up vp (p p + p p )

(3.29)

181

3.1 Weakly Interacting Bose Gas

and

ap ap = (up p + vp p )(up p
+ vp p )

= up vp + u2p p p
+ vp2 p p + up vp (p p + p
p ),

(3.30)

which have the vacuum expectation values


hap ap i = up vp .

hap ap i = vp2 ,

(3.31)

Using these we find from (3.26) the approximate Hamiltonian


H0

i
2 X h
gV +
(p + g0 ) vp2 + g0 up vp
2 p6=0
(


i


1h

p (3.32)
(p + g0 ) u2p + vp2 + 2g0 up vp p p + p
2
)
i 


1h

+ (p + g0 ) 2up vp + g0 u2p + vp2


p p + p p .
2

We now see that the off-diagonal terms in the last line of (3.32) vanish if we set
2up vp
g0
=
.
2
2
up + vp
p + g0

(3.33)

In terms of the angles p in (3.27), this reads


tanh 2p =

g0
.
p + g0
q

(3.34)

Using the trigonometric relation cos 2p = 1/ 1 tanh2 2p , this implies


cosh 2p =

p + g0
,
Ep

sin 2p =

g0
,
Ep

(3.35)

where
Ep

(p + g0 )2 g 20 2 .

(3.36)

are the quasiparticle energies. They go to zero for p 0, in accordance NambuGoldstone theorem. Using further
cosh p =

(cosh 2p + 1)/2,

sinh p = (cosh 2p 1)/2,

(3.37)

we obtain
2

cosh p =

u2p

1
p + g0
1
p + g0
=
1+
, sinh2 p = vp2 =
1
, (3.38)
2
Ep
2
Ep

182

3 Interacting Non-Relativistic Particles

with
up vp =

g0
,
2Ep

(3.39)

and
H0 gV

i
X
2 X h
(p + g0 ) vp2 + g0 up vp +
Ep p p ,
+
2
p6=0
p6=0

(3.40)

can be rewritten as
H0 gV

X
2 1 X
+
[Ep (p + g0 )] +
Ep p p ,
2
2 p6=0
p6=0

(3.41)

Note that instead of fixing the Bogoliubov coefficients to be (3.38) by the requirement of vanishing of off-diagonal elements in the last line of Eq. (3.32), we
could also have obtained the same coefficients by extremizing (3.42) with respect to
variations in vp and up with the constraint u2p vp2 = 1, so that up = vp vp /up.
Indeed, this would have lead to
2 (p + g0 ) vp vp + g0 (u2p + vp2 )vp /up = 0,

(3.42)

which is solved again by (3.38).


At very low temperatures, the quasiparticles are frozen out and the ground state
energy is given by the first three terms:
E0 gV

2 1 X
+
[Ep (p + g0 )] .
2
2 p6=0

(3.43)

The particles with nonzero momentum p lie outside the condensate and constitute what is called the uncondensed part of the liquid . Their number is given by
the expectation value in the vacuum state of quasiparticles |0 i. This state has the
property p |0 i = 0, so that h0|p = 0, so that

hap ap i h0 |(up p + vp p )(up p + vp p


)|0 i = vp2 .

and

(3.44)

hap ap i h0 |(up p + vp p )(up p


+ vp p )|0 i = up vp .

(3.45)

hap ap i V u

(3.46)

The total number of uncondensed particles is found from the sum


Nu =

p6=0

vp2 .

p6=0

Inserting vp2 from (3.38) yields


Z
Z
1
d3 p
d3 p 2
1 X 2
v
=
vp =
u =
V p
(2h)3 p 2 (2h)3

p + g0
1 .
Ep

(3.47)

183

3.1 Weakly Interacting Bose Gas

To evaluate this, we replace the sum over p by an integral,


X

p6=0

d3 p
,
(2h)2

and measure all momenta in units of


kg0 =

2Mg0
.
h

(3.48)

Expressing the coupling constant g in terms og the experimentally observable s-wave


scattering length as via (3.12), this is equal to
q

kg0 = 2 2as 0 .

(3.49)

This is a characteristic wavenumber associated with the s-wave scattering length


(3.12). Thus we substitute
p=h
kg0 ,
(3.50)
and find for the density of uncondensed particles (3.47) the equation
u =

Nu
3 Iu
= kg
,
0
V
4 2

(3.51)

where Iu is the integral


Iu

2 + 1

2
d q
1 =
.
3
(2 + 1)2 1
2

(3.52)

Hence
3
u = kg
0

8 q 3
Iu

=
a 0 .
4 2
3 s

(3.53)

The lesson of Eq. (3.51) is that, even at zero temperature, the repulsive interaction scatters a small set of particles out of the condensate, thus causing a depletion
to a fraction [6, 7]
u
8 q 3
0

=1
1
a 0 .
(3.54)

3 s
In the strongly interacting superfluid 4 He, only about 8% of the particles condenses
in the zero-temperature state [8]. Nevertheless, all the particles participate in the
superfluid motion [9] and the superfluid density is equal to the total density. The
condensate and the normal fluid move together.
Let us also calculate the energy of the Bose system. There we encounter a typical
divergence of the present local quantum field theory: The momentum sum in the

184

3 Interacting Non-Relativistic Particles

second term of (3.41) does not converge. Indeed, the energy Ep of (3.38) behaves,
at large momenta, like
Ep (p + g0

v
u
u
) t1

g 2 20
1 g 2 20

+
g

+ ... .
p
0
2 p
(p + g0 )2

(3.55)

The integral
V

d3 p
[Ep (p + g0 )] V
(2h)3

d3 p g 2 20
(2h)3 p

(3.56)

diverges linearly. The integral can be made convergent. First we rearrange the terms
in (3.43) to
E0

g 220
2
1 X g 220 1 X
Ep p g0 +
,
gV 0
+
2
4 p6=0 p
2 p6=0
2p
"

(3.57)

The last sum is now convergent, but the sum before it is still divergent. However,
the divergence disappers if the energy is expressed in terms of the renormalized
coupling constant gR defined by
1
1 X 1
1 Z d3 p 1
1

=
.
gR
g V p6=0 2p
g
(2)3 2p

(3.58)

This is precisely the quantity that is measurable by the s-wave scattering length to
which it is related by Eq. (3.12). In terms of the experimentally measurable gR , we
find the renormalized energy
E0

2 1 X
g 2 2
gR V
Ep (p + gR ) + R 0
+
2
2 p6=0
2p
"

(3.59)

The above renormalization procedure of going from an unobservable divergent


bare coupling g to a finite observable coupling gR is typical for all quantum field
theories. Here is the first place where it appears. This may surprise the reader since
we are, after all, dealing with Schrodinger system which should have no infinities.
The puzzle is resolved by the observation, that the three-dimensional Schrodinger
equation has no proper solution if the potential is a -function. Only in one dimension it does. After the above renormalization, on the other hand, the problem of
finding the quantum behavior of a particle in a point-like potential becomes meaningful and solvable. A meaningful limit can, of course, be obtained from a study of
a completely finite problem in an extended potential hole and taking at the end the
limit of zero range while letting the depth of the potential go to infinity.
With the substitution (3.50) we can rewrite the energy as
E0 = gR V

2
IE
3
+ V g0 kg
,
0
2
4 2

(3.60)

185

3.1 Weakly Interacting Bose Gas

where IE is the integral


IE

d 2


8
2
1
.
(2 + 1)2 1 2 1 + 2 =
2
15

q

(3.61)

Inserting gR = 4h2 as /M, this can be written as [1]


E0

2h2 as 2
128 q 3
a 0 .
=V
1+
M
15 s
!

(3.62)

The operators p applied to the vacuum state generate quasiparticles. They


have the Bogoliubov energies Ep instead of the original energies p = p2 /2M. For
small momenta, the quasiparticle energy (3.36) starts out like
Ep

g0
|p| +
M

M p3
+ ... .
g0 8M 2

(3.63)

The initial linear behavior defines the velocity of second sound:


c

g0
.
M

(3.64)

This behavior is a consequence of the famous Nambu-Goldstone theorem, according


to which a spontanoeus breakdown of a continuous symmtery leads to excitation with
such an energy-momentum relation. The continuous symmetry of the present Bose
system is an invariance of the Hamiltonian under a transformation of all operators by
a U(1) phase factor ap ei ap . This symmetry is vialated by the condensate which
is charecterized by a ground-state expectation value A0 6= 0 with a definite phase,
that was chosen to be real. For more details on this theorem see Section 17.2.1.
For large momenta, the expansion is
Ep

p2
c4 M 3
+ Mc2
+ ... ,
2M
p2

(3.65)

and the particles behave like free particles, but with an apparent rest energy Mc2 ,
similar to the low-momentum behavior of relativistic particles. In Fig. 3.1 we have
plotted the p for M = 1 and g0 = 1 and compared it with the free-particle curve
p2 /2 as well as the experimental determination of the excitation spectrum in bulk
superfluid 4 He. We see that the simple approximation (3.36) yields initially a similar
behavior as in the superfluid, but is not quite capable of giving the pronounced
minimum in the experimental data. The energy of minimum is referred to as the
roton energy and an interpretation of these states was first given by Feynman in
1955.
In the neighborhood of the minimum, the experimental energy is roughly
(p p0 )2
Ep +

(3.66)

186

3 Interacting Non-Relativistic Particles


40

5
4
3

p
(K)
kB

p2
2M

p2
2M

30

Ep

20

Ep

10

0.5

1.5

2.5

|p| (
h
A

|p|

Figure 3.1 Plot of the quasiparticle energies as a function of the momenta in an interacting Bose gas. Dashed curve indicates free-particle energy. Right-hand side shows
experimental curve for superfluid helium at 1.12 K measured by neutrons of wavelength
4.04
A. Slope of dashed straight line indicates sound velocity 2.37 104 cm/sec.

with
1

p0 1.9
A h
,

8.96KkB ,

0.5me

(3.67)

In Eq. (3.64) we deduced the velocity c of second sound from the slope of the
quasiparticle spectrum at small momenta. We may also use a classical formula relating the sound velocity to the compressibility of a material. At zero temperature,
the compressibility follows from the ground state energy (3.62). Using only the
leading first term we may write
E0

gN 2
.
2V

(3.68)

From this we obtain


c

V 2 2E

MN V 2

gR
,
MV

(3.69)

in agreement with (3.64).


In bulk superfluid 4 He, the sound velocity at atmospheric pressure and zero
temperature is2
c

3 137

cm
kB
m
A K.
2.373
17.2
sec
sec
h

(3.70)

This number is obtained from measurements of the velocity c2 of the so-called second
sound, which can be measured more accurately than c. Second sound arises if the
2

In the superfluid, it is easier to measure very accurately the velocity c2 of second sound, where
the superfluid component oscillates out of phase against
the normal component (at a constant
total density). This velocity c2 is precisely by a factor 3 smaller than c, as shown first by L.D.
Landau, J. Phys, USSR, 5, 71 (1941) (see also L.D. Landau, Z. M. Lifshitz, Fluid Mechanics,
Addison Wesley, Reading, Mass. 1959, Ch. XIV, and Statistical Physics, ibid, Ch. VI, pp. 66-67).

3.2

187

Weakly Interacting Fermi Gas

superfluid component oscillates out of phase against the


normal component (at a
constant total density). The velocity c is by a factor 3 larger than c2 , as shown
first by Landau.3 The result (3.70) agrees very well with the slope in Fig. 3.1.
Note that the interaction between the Bose particles has to be repulsive, otherwise the sound velocity would be imaginary!

3.2

Weakly Interacting Fermi Gas

Two field theoretic models of interacting fermions are of great theoretical and practical importance. One helps to understand the contribution of electrons to the binding
energy of metals. The second illustrates the mechanism by which electrons can become superconductive. In particle physics, the second mechanism has inspired all
models for explaining the properties of the pion and thus of the most important
nuclear forces. It underlies all present attempts at constructing a unified theory of
strong electromagnetic and weak interactions.

3.2.1

Electrons in a Metal

When electrons run through a metal with crystalline order it is possible to define
a momentum of the particles (since there is still a translational symmetry). For
small momenta the energy behaves like p2 /2M, only that m is no longer the true
electronic mass but some effective mass, whose size depends on the interactions with
the positive ions.
Since the ions are very heavy, they take little part in the dynamics of the electrons. This is well known for single atoms, where the motion of the electron hardly
influences the nucleus. One is therefore justified in approximating the ions simply
by a uniform time-independent positive background charge. The Hamiltonian of the
electrons is
Hel =

N
X

p2i
1 X
1
+ e2
.
2
|xi xj |
i=1 2M

(3.71)

The Hamiltonian of the uniform positive background is


e2
Hb =
2

d3 xd3 x

n(x)n(x )
|x x |

(3.72)

with n(x) N/V . The electron background interaction, finally, reads


2

Hint = e
3

XZ
i

d3 x

n(x)
.
|x xi |

(3.73)

L.D. Landau, J. Phys, USSR, 5, 71 (1941) (see also L.D. Landau, Z. M. Lifshitz, Fluid Mechanics, Addison Wesley, Reading, Mass. 1959, Ch. XIV, and Statistical Physics, ibid, Ch. VI,
pp. 66-67).

188

3 Interacting Non-Relativistic Particles

In order to keep control over divergencies at long distances it is useful to insert an

exponential screening factor e(xx ) into the Coulomb potential 1/|x x |. Since
otherwise each of the terms diverges in the limit N , U , = N/V = const
(the so-called thermodynamic limit).
In the Hamiltonian
H = Hel + Hb + Heb

(3.74)

the limit 0 can be taken at the end, together with the limit V . One
has
1
3
to let the screening length grow to infinity in the same way as the size L =
V
of the system. In order to be able to make use of translational invariance we shall
assume the finite system to be a cube of periodically repeated volume V = L3 in all
directions, i.e., we assume periodic boundary conditions in all wave functions. Then
all Fourier components have the discrete momenta
2
ni .
L
Let us first integrate the energy of the positive background:
ki =

Hb

1 2 N 2
e
=
2
V
1 2 N 2 4
=
e
.
2 V 2


 Z

|xx |
3 e
xd x
|x x |

1
N
= e2
2
V


2 Z

(3.75)

d3 x

4
2
(3.76)

We see the need of using the length 1 to keep this expression finite.
The electron background interaction energy can be calculated just as easily
He,b = e2

N
X

N
i=1 V

d3 x

e|xxi |
N 2 4
= e2
|x xi |
V 2

(3.77)

We may therefore confine our attention purely to the electrons Hamiltonian remembering only to subtract, at the end, the constant energy e2 /2N 2 V 1 42 before
taking the limit 0.
The kinetic energy of the electron gas is easily written down
Eel,kin =

p ap, ap, ,

(3.78)

p,

where = 1/2 distinguishes the two spin states. Alternatively, we shall use the
spin-up and spin-down symbols = and =.
The potential energy is given by
Hel,pot =

1 X
e2
ax , ax, .
ax, ax ,
2 x,x ,,
|x x |

(3.79)

Expanding
1 X ipx
ax, =
e ap,
N p,

(3.80)

3.2

189

Weakly Interacting Fermi Gas

and

1 e|xx |
1 X
1
=
eiqx ,

2
4 |x x |
V q q + 2

(3.81)

we find
Hel,pot =

e2
2V

p,p ,q,,

q2

4
ap+q , ap q , ap , ap, .
2
+

(3.82)

Consider the terms with q = 0


q=0
Hel,pot
=

e2 4
2V 2

ap, ap , ap , ap, .

(3.83)

p,p ,,

Using the commutator


Xh

p ,

ap, , ap , ap , = ap ,

(3.84)

they can be written as


q=0
Hel,pot
=

e2 4
2V 2
2

p,p ,,

ap , ap , ap, ap, pp ap ap

e 4 2

(N N).
2V 2

(3.85)

For a fixed particle number N, this is equal to


Eel,pot =

e2 4 2
(N N)
2V 2

(3.86)

The first term cancels with the constant background term. The second term grows
linearly with the total particle number. It is of the order
1
L

1
L

!2

(3.87)

where L V 1/3 is the linear size of the system, and disappears for L as long
as L stays fixed.
The remaining terms in the energy (3.82) have a smooth limit 0 so that we
can continue the calculation with = 0 and study the Hamiltonian operator

H= Hel,kin + Hel,pot
=

X
p,

X
e2
p2
4
ap, ap, +
ap+q , ap q , ap , ap, . (3.88)
2
2M
2Vp,p ,q6=0,, q

We now make the important observation that although the potential is not weak and
of long range, there exists a physical regime where the interaction can be treated as a

190

3 Interacting Non-Relativistic Particles

small perturbation. This is in the regime of high density. Indeed, the momenta in the
first term has the dimension L2 , the second term has the dimension 1/V q 2 L1 .
The remaining calculations involve only dimensionless sums over integer numbers.
Thus, in a very small box, the second term is negligible compared to the first.
Let us calculate the first term. For simplicity, we shall do this in the limit of
large L. At zero temperature, the electrons fill all states up to the Fermi momentum.
Then
E (0) =

1 X 2
p (pF p)
2M p,

V 2 p F d3 p 2
p
2M 3 0 (2)3
V
V 4h2 5
p5F
=
=
k .
(2)3 5m F
2Mh3 10 2
Z

(3.89)

and pF is determined from


N
2 Z p F d3 p
2 p3F
= 3
=
,
V
h
0 (2)3
h
3 6 2

(3.90)

3 p2F
E (0)
=
.
N
5 2M

(3.91)

so that

Using the Fermi energy


F

p2F
,
2M

(3.92)

this can be written as


3
E (0)
= F .
N
5

(3.93)

and (3.90) becomes




N
V

2mpF
F .
3 2h
3

(3.94)

The result is often stated in terms of the length rs , the spherical volume per
particle
4 3
V
rs = .
3
N

(3.95)

From (3.90) and (3.95) we obtain the relation


9
pF = h
kF =
4


1/3

.
rs

(3.96)

3.2

191

Weakly Interacting Fermi Gas

Remembering the Bohr radius


aH = 4

h
2
,
me2

(3.97)

the energy density at zero temperature can then be written as


E (0)
1 e2 3 9 2/3 aH
=
N
4 2aH 5 4
rs


aH 2
Ry,
2.21
rs


2

(3.98)

where
Ry

1 e2
13.5eV
4 2aH

(3.99)

is the Rydberg energy, the binding energy of the hydrogen atom. Note that we are
using rationalized electrostatic unity in which the Coulomb potential is e2 /4r.
Consider now the electron-electron interaction. To lowest order, the energy is
shifted by
E (1) =

e2 X X 4
h0 |ap+q , ap q , ap , ap, |0 i
2
2V q6=0 p,p ,, q

(3.100)

where |0 i denotes the ground state of the free electron gas. In order to give a
non-zero contribution, the state |0 i must contain an electron in the states p ,
and p, as well as in the states p + q, and p q, . Both p and p must therefore
lie inside the Fermi sphere p2 < p2F , p2 < p2F . The annihilation operators create two
holes. These must subsequently be filled up by ap+q , ap , so that also p + q, p q
are inside the Fermi sphere. Thus one must either have
(p + q, ) = (p, );

(p q, ) = (p , ),

(3.101)

(p + q, ) = (p , ),

(p q, ) = (p, ).

(3.102)

or

The first pairing would imply q = 0, but this momentum is not included in the sum.
So we are only left with the second case which contributes an expectation value
E (1) =

e2 X X h
2
p ,p+q h0 |ap , ap, ap , ap, |0 i
2V q6=0, p,p q 2

2 X
e2 X h
2
(pF |p|)(pF |p + q|).
2V q6=0 q 2 p

(3.103)

192

3 Interacting Non-Relativistic Particles

The factor 2 in the last expression comes from the sum over . Substituting p
p q/2, p + q p + q/2, and approximating the sum over p by an integral, this
becomes
V
(2h)3

d p pF


q
q

pF p + .
2
2


(3.104)

This expression represents the common volume of two spheres of radius pF in momentum space whose origins are separated by a distance q (see Fig. 3.2).

Figure 3.2 Common volume of two spheres at a distance q in momentum space. Righthand picture illustrates intermediate integration (3.105).

In units of p3F , the volume is found by calculating the integral over the spherical
shell x2 + (y 1)2 = 1 (illustrated on the right-hand side of Fig. 3.2):
2

xm
0

dxx y = 2

ym

y3
y2
dyy(1 y) = 2 m m .
2
3
!

(3.105)

2
3
Subtracting this from the volume x2m ym = (2ym
ym
) of a disk of thickness ym ,
we find

2
ym

1 3
,
ym
3


ym = 1

q
.
2pF

(3.106)

This has to be multiplied by a factor 2 p3F to obtain the volume of overlap in


Fig. 3.2:
4p3F
3

3
1
1 xq + x3q (1 xq ),
2
2

xq

q
.
2pF

(3.107)

Inserting this for the the momenta q in Eq. (3.103), we arrive at


E

(1)

2
3
1
e2 1 4p3F X h
2
1 xq + x3q
=
3
2
2 (2h) 3
2
2
q6=0 q




Z 1
e2 V 4p3F
3
1 3
2
2

4
h
2p
dx
1

x
+
x
F
2 (2h)6 3
2
2
0
2
V e
2k 4 .
=
(2)3 4 F

(3.108)

3.2

193

Weakly Interacting Fermi Gas

Replacing kF3 by 3 2 N/V via (3.90) yields


E (1)
1 e2 3
=
kF .
N
4 2 2

(3.109)

Inserting here pF from (3.96), we obtain


1 e2 3
E (1)
=
N
4 2 2

9
4

1/3

,
rs

(3.110)

or, in terms of Rydberg units,


E (1)
9
=
N
4


1/3

3 aH
aH
Ry 0.916
Ry.
2 rs
rs

(3.111)

The total energy per volume is then


E
1
=
V
(2)3

e2 4
4 5
kF
2k + . . . ,
5m
4 F
!

(3.112)

and the energy per particle has the high-density expansion


E
N

1 e2 3 h

9 1/3 3 aH
+ ...
=

4
2 rs
4 2 2 rs
"
#

2
aH
aH
2.21
+ . . . Ry.
0.916
rs
rs
"

3 9
5 4


2/3 

aH
rs

2

(3.113)

The leading term shows repulsion due to the exclusion principle. The second term
represents the attraction due to the exchange interaction between the electrons. The
curve is plotted in Fig. 3.3. It has a minimum at
4
rs
=
aH
5
where

9
4

1/3

4.823,

Emin
45
=
0.095 Ry.
N
48 2

(3.114)

(3.115)

It is interesting to see that the minimum at E/N 0.095 Ry 1.29 eV lies in


the neighborhood of the values found in metallic sodium:
rs
3.90,
aH

E
1.13Na.
N

(3.116)

This suggests that the main part of the metallic binding energy is due to the electrons
in the lattice.

194

3 Interacting Non-Relativistic Particles

0.1

E/N

0.05
2

10

r0 /a0

-0.05
-0.1

Figure 3.3 Energy density in units Ry of electron gas in uniform background of positive
charge.

Improvements of this lowest-order result were given by Gell-Mann and Brueckner


1957.4 They derived the next term in the expansion, which Wigner5 called the
correlation energy,
E

(2)

rs
rs
2
+ C Ry 0.0622 log
0.096 Ry.
= 2 (1 log 2) log

aH
aH


(3.117)

At metallic electron densities, the ratio rs /aH lies between 2 and 5. This regime can
be described approximately by the above energies if E (2) is modified to6
E

(2)

rs
0.0622 log
0.230 Ry.
aH


(3.118)

The further corrections vanish for rs 0. For more details see the textbook by
C. Kittel (1963)7. Note that in the high-density limit, Eq. (3.113) approaches the
exact result.
Let us also remark that Wigner8 pointed out the possibility of an electron lattice
(Wigner lattice) forming in the limit of low density. Such a gas has meanwhile been
observed experimentally.

3.3

Superconducting Electrons

In a metal, the Coulomb repulsion does not always dominate the phenomena of
the electrons. At low temperatures, the electron-phonon interactions may play an
4

M. Gell-Mann and K. Br
uckner, Phys. Rev. 106, 364 (1957), M. Gell-Mann, Phys. Rev. 106,
369, 367 (1957).
5
E.P. Wigner, Phys. Rev. 46, 1002 (1934). For the calculation see W. Macke, Z. Naturforsch.
5a, 192 (1950), also A.L. Fetter and J.D. Walcka, Quantum Theory of Many-Particle Systems,
McGraw-Hill, 1971, New York; Section 12.
6
P. Noziers and D. Pines, Phys. Rev. 111, 442 (1958).
7
C. Kittel, Quantum Theory of Solids, John Wiley & son, Inc., New York, 1963, p. 115
8
E.P. Wigner, Phys. Rev. 46, 1002 (1934)

3.3 Superconducting Electrons

195

important role and lead to a completely new physical phenomenon, that of superconductivity.
Superconductors are materials without resistance to the flow of electricity. The
phenomenon was first observed in 1911 by the Dutch physicist Heike Kamerlingh
Onnes at Leiden University. When he cooled mercury down to the temperature of
liquid helium, which appears at about 4 degrees Kelvin (-269C), its resistance suddenly disappeared. Onnes won the Nobel Prize in physics in 1913 for this discovery.
Superconductors have an important property which distinguishes them from ordinary conductors of extremely low resistance: They are perfect diamagnets implying that they do not tolerate any magnetic fields on their inside. This is the so-called
Meissner-Ochsenfeld effect discovered in 1933. This effect causes superconductive
materials lying on top of a sufficiently strong magnetic field to be lifted if they are
cooled below the critical temperature (levitation). A perfect conductor would only
hover above the magnet if brought in from the outside due to induction generating
a current with a magnetic moment opposite to the external field.
For the purpose of energy conservation, it is a challenging problem to find superconductive materials which can transport high currents without loss at room
temperature. Since 1941, the record was held for a long time by niobium-nitride,
which becomes superconducting at 16 K, surpassed in 1953 by vanadium-silicon
with a critical temperature of 17.5 K. In 1962, a first commercial superconducting
wire was manufactured from an alloy of niobium and titanium. First applications
were made in 1987 in the Fermilab high-energy particle-accelerator Tevatron where
the magnetic field were produced by supercurrents in copper-clad niobium-titanium.
The magnets had been developed in 1960 at the Rutherford-Appleton Laboratory
in the UK.
The first satisfactory theory of superconductivity was developed in 1957 by N.N.
Bogoliubov and, independently, by J. Bardeen, L. Cooper, and J. Schrieffer [11],
now called BCS which won them the Nobel prize in 1972. This theory explains the
early forms of superconductivity at temperatures close to absolute zero observed for
elements and simple alloys.
New advances were made in the 1980 when the first organic superconductor was
synthesized by the Danish researcher Klaus Bechgaard of the University of Copenhagen and three French team members, which turned out to become superconductive
at 1.2K transition. The possibility that this could happen had been pointed out in
1964 by Bill Little at Stanford University.
The latest major breakthrough was made in 1986 by Alex M
uller and Georg
Bednorz at the IBM Research Laboratory in R
uschlikon, Switzerland [49]. They
synthesized brittle ceramic compound that superconducted at the record temperature of 30 K. What made this discovery so remarkable was that ceramics are normally insulators. They dont conduct electricity well at all. So, researchers had
not considered them as possible high-temperature superconductor candidates. The
Lanthanum, Barium, Copper and Oxygen compound that Mler and Bednorz synthesized, behaved in a not-as-yet-understood way. Their discovery won them the

196

3 Interacting Non-Relativistic Particles

Nobel Prize in 1987. It was later found that tiny amounts of this material were
actually superconducting at 58 K.
Since then there has been a great deal of activity trying to find ceramics of mayn
combinations with higher and higher critical temperatures. In 1987 superconductivity was reached in a material called YBCO at 92 K, a temperature which can
simply be reached by cooling with liquid nitrogen.
The present world record was reached in 1994 at Tc = 138 K by a thallium-doped,
mercuric-cuprate comprised of the elements Mercury, Thallium, Barium, Calcium,
Copper and Oxygen. Under extreme pressure of 300,000 atmospheres, this critical
temperature an be raised by 25 to 30 more degrees (see Fig. 3.4, from Wikipedia
page on Superconductivity).

Figure 3.4 Historical evolution of critical temperatures of onset of superconductivity (in


units of Kelvin).

The phenomenon is based on an attraction between the electrons that moce


in momentum space close to the Fermi surface. An electron running through the
lattice attracts the positive ions in its neighborhood. These, in turn, attract the
other electrons. The relevant phonons are those with the highest frequency, which
in the Debye model of the specific heat of the solid is called Debye frequency D :
This corresponds typically to a temperature of about 10000 K. If the electrons are
near the Fermi surface, the phonon attraction can overcome the Coulomb repulsion
as was first suggested by Frohlich in 1950 and brought to the level of a theory by
Bardeen, Cooper, and Schrieffer (1957) (BCS) [11].
The effective electron Hamiltonian that emerges from an interaction like (3.10),
rewritten in terms of the momentum transfer q q3 q1 for the scattering process
caused by the potential as
X
g
Hint =
ap+q, ap q, ap , ap, .
(3.119)
2V p,p ,q,,
Assuming that there is an attractive -function interaction only between electrons
with opposite spins, the total Hamiltonian in a grand-canonical ensemble coupled to

197

3.3 Superconducting Electrons

a reservoir of chemical potential is written after the sum over the two spin states
as
X
g X
H = H0 + Hint =
p ap, ap,
ap+q , ap q , ap , ap, ,
(3.120)
V
p
p,p ,q
where
p p

(3.121)

is the energy difference with respect to the chemical potential. If the potential is
not V (x x ) = g (3) (x x ) but
V (x x ) =

d3 q iqx
e V (q),
(2)3

(3.122)

then the sum in the interaction contains and extra factor V (q.
Since in the ground state, the main effect to be discussed will come from the attraction where also the momenta are opposite, BCS went further and approximated
the interaction by the terms with p = p:
g X
Hint Hpair
ap+q , apq , ap, ap,
(3.123)
V p,q
The sum H0 +Hpair is the BCS pairing Hamiltonian. The kinetic energy p is initially
the free-particle kinetic energy p2 /2M of the electron in a metal, with an effective
mass m different from the electron mass due to the effect of the crystal lattice. In
order to find the phenomenon of superconductivity we must, however, work with a
grand canonical ensemble of fluctuating total particle numbers and study the system
in contact with a particle reservoir of chemical potential . This will be close to pF ,
the Fermi surface of the free force electron gas at T = 0. Instead of studying it we
N.
The N
term components to replace p by p p .
shall investigate H
Moreover, since the attractive interaction can be derived only for electrons near
the Fermi surface, we can replace p approximately by p (pF /m)(|p|
pF ) + p2F /2M (see Fig. 3.5). After this, we treat the Hamiltonian according to
Bogoliubovs method9 , which goes as follows.
Let us introduce two arbitrary real momentum-dependent parameters p and
p into the free Hamiltonian, changing it to
0
H

X
p

(p p ) ap, ap, ap, ap,

X
p

p ap, ap,

X
p

p ap, ap, + p .
(3.124)

The additional terms are subtracted from the interaction Hamiltonian, which becomes


X
X
int g
H
ap+q , ap q , ap , ap, +
p ap, ap, ap, ap,
V p,p ,q
p
+

X
p

p ap, ap, +

X
p

p ap, ap, . (3.125)

N.N. Bogoliubov, Sov.Phys.-JETP 7 , 41(1958). J.G. Valatin, Nuovo Cimento 7 , 843 (1958).

198

3 Interacting Non-Relativistic Particles


2

p =

1.5

p2
2m

2
1.5

app
p

app
p

0.5

0.5

app
p
0.5

1.5

|p|

0.5

1.5

|p|

Figure 3.5 Dashed straight line indicates the approximate energies p (pF /m)(|p|
pF )+p2F /2M of a free electron near the Fermi surface in a grand canonical ensemble. Righthand side shows energies if the states below the Fermi surface are filled up in which case the
hole excitations below the Fermi surface have approximate energies p (pF /m)(|p|
pF ) + p2F /2M .

The total Hamiltonian


0 + H
int
H =H

(3.126)

0 is bilinear in the creation


has remained unchanged. The modified Hamiltonian H
and annihilation operators and can be written as
0 =
H

X
p

Ep

ap, , ap,

with

hp

ap,
ap,

+ p ,

(3.127)

(p p )2 + 2p

Ep

and the matrix

(3.128)

p
1 p p
.
hp
Ep p
p + p

(3.129)

This matrix can be diagonalized by a similarity transformation due to Bogoliubov


1
0
0 1

hp = Bp

Bp ,

Bp

up vp
vp up

(3.130)

with matrix elements, fixed up to some phase factor ei ,


up =

v
u
u Ep
t

+ p p
,
2Ep

vp =

Choosing the positive square root, these satisfy


u2p

1
p p
=
1+
,
2
Ep

vp2

v
u
u Ep
t
!

p + p
.
2Ep

1
p p
=
1
,
2
Ep

2up vp =

(3.131)

p
.
Ep

(3.132)

Due to their frequent occurance we shall abbreviate the subtracted energy p p


by
p p p .
(3.133)

199

3.3 Superconducting Electrons

The similarity transformation of hp amounts to a canonical transformation of the


electrons to new quasiparticles whose creation and annihilation operators are given
by
p = up ap, vp ap, ,

p = vp ap, + upap, .

(3.134)

In terms of these, the diagonalized Hamiltonian (3.127) reads


0 =
H

X
p

p =

1.5

p
p + p .
E p p p

p2
2m

1
0.5

-1

Ep

1.5

0.5
-2

(3.135)

|p|

-2

Ep

-1

|p|

Figure 3.6 Gap in the energy spectrum caused by attraction of pairs of electrons with
opposite spin and momenta.

The energy spectrum has a gap in comparison to the original energies in the first
term in (3.124) as indicated in Fig. 3.6. The creation and annihilation operators p ,

p and p ,
p of the quasiparticles have the same anticommutation rules as the
original particle operators ap and ap . The operators
p create obviously states of
negative energy. These can be turned into hole states of positive energy by replacing

p ,

p .

(3.136)

This brings the Hamiltonian (3.135) to the form


0 =
H

Ep p p + p p +

X
p

X
p

(p Ep ) .

(3.137)

The energies of quasiparticles and holes are shown in Fig. 3.7. The creation operators
of the hole states have the same anticommutation rules as those of the particles.
Altogether we have the anticommutation rules:
n

p , p = pp ,

p , p = p,p ,

The inverse transformation of (3.134) reads

p , p

= 0.

(3.138)

,
ap, = up p + vp p

ap, = vp p + up p .

(3.139)

200

3 Interacting Non-Relativistic Particles


2

2
1.75
1.5
1.25
1
0.75
0.5
0.25

Ep

1.5

0.5
-2

Ep

-1

|p|

-2

-1

Ep

|p|

Figure 3.7 Exciting a hole in the Fermi sea in which all quasiparticle states below the
Fermi surface are filled costs positive energy with the spectrum shown in the right-hand
figure.
1 E=
0.8

p
2 + 2

1.5

0.6

0.4
0.5

0.2
-1

-0.5

0.5

0.5

1.5

Figure 3.8 Gap in energy spectrum. The straight dashed line is the same as in Fig. 3.5
with a shifted origin. The right-hand side shows the position of these energy curves with
respect to the free-particle energy in momentum space.

3.3.1

Zero Temperature

For zero temperature, we now use the vacuum state with respect to the quasiparticles
0 has zero
as a trial ground state of the system. In it, the free part of the energy H
expectation. Consider now the expectation value of the interaction (3.125). For the
pair terms, these are immediately calculated
D 

hap, ap, i = 0 up p + vp p



 E

vp p + up p
0 = up vp =

p
,
2Ep

p
.
2Ep
Their momentum integrals will be denoted by and respectively, i.e.,
D 

hap, ap, i = 0 vp p + up p



 E

up p + vp p
0 = up vp =

1 X
1 X p
hap, ap, i =
V p
V p 2Ep
X
1 X p
1
.
hap, ap, i =
V p
V p 2Ep

(3.140)
(3.141)

(3.142)
(3.143)

Note that the diagonal terms have the expectation values


hap, ap, i

p
1
,
2
Ep

 E
D 

1

0 up p +vp p up p +vp p
0 = vp2 =

(3.144)

201

3.3 Superconducting Electrons

hap, ap, i

p
1+
. (3.145)
2
Ep

 E
D 

1

0 vp p +upp vp p +upp
0 = u2p =

The difference is compensated by the extra last term in (3.127). If we denote the
expectation values (3.144) by
hap, ap, i np, V p, ,

hap, ap, i 1 np, 1 V p, ,

(3.146)

and perform a sum over all momenta of these, we find


X
p

hap, ap, i =

vp2 = V

X
p

1
p, = V = N.
2

(3.147)

The expectation value of the interaction is more complicated to obtain. It requires evaluating
E
D


0 ap+q , ap q , ap , ap, 0 .
(3.148)
This can be done by using the anticommuatation rules10 (2.210) to expand (3.148)
into the three pair terms
E
D


0 ap+q , ap q , ap , ap, 0

=
+

ED
E
D




0 ap+q , ap q , 0 0 ap , ap, 0
E
ED
D




0 ap+q , ap, 0 0 ap q , ap , 0
ED
E
D




0 ap+q , ap , 0 0 ap q , ap, 0 .

(3.149)

Using (3.140)(3.145), the first two terms on the right-hand sides contribute
E
D


0 ap+q , ap q , ap , ap, 0

= p ,p up vp up+q vp+q + q,0 vp2 vp2 .

(3.150)

The last term in (3.149) gives no contribution due to the opposite spins, such that
we obtain for the interaction in (3.120):
h0|Hint|0i =



g X
g X 
up vp up+q vp+q + q,0 vp2 vp2 =
up vp up vp + vp2 vp2
V p,p ,q
V p,p

(3.151)

and for the interaction (3.125):


int |0i =
h0|H

 X

 X
g X
up vp up vp + vp2 vp2 + p 2vp2 1 + p 2up vp . (3.152)
V p,p
p
p

0 +H
int is from (3.137), (3.151)
Then the expectation value of the total energy H = H
h0|H|0i =
10

X
p

(p p Ep + p 2up vp ) +

X
p

p 2vp2


g X
up vp up vp + vp2 vp2 .
V p,p

More elegantly we can use Wicks theorem to be derived in Section 7.14.1.

(3.153)

202

3 Interacting Non-Relativistic Particles

The first term can be simplified using Eq. (3.132), leading to


h0|H|0i =

X
p

2 p vp2


g X
up vp up vp + vp2 vp2 .
V p,p

(3.154)

Let us vary this equation with respect to vp under the constraint u2p + vp2 = 1, so
that up /vp = vp /up. This yields for each momentum an equation
2 p vp



g X 2
g X
up vp up vp2 /up 2
v vp = 0.
V p,p
V p,p p

(3.155)

Multiplying this with up and remembering the momentum sums (3.142) and (3.147).
Then Eq. (3.155) becomes


2 (p g ) up vp g u2p vp2 = 0,

(3.156)

or
0 (p g ) p = g(p p ).

(3.157)

This is solved by the constant variational parameters


p = g ,

p = g.

(3.158)

Inserting these into (3.143) and (3.145), we find the self-consistent gap equations
= g =

g X
,
V p 2Ep

(3.159)

and the self-consistent particle number equation


!

p g
1 X
,
1
=
2V p
Ep

(3.160)

Inserting this into the energy (3.154) we find


h0|H|0i =

X
p

2 (p g ) np,

V 2
.
g

(3.161)

or, more generally,


h0|H|0i =

X
p

V
p (np, + np, ) 2 .
g

(3.162)

The quantity is calculated from the gap equation which is solved the help of the
modified chemical potential

= + g ,

(3.163)

203

3.3 Superconducting Electrons

which in turn is is determined from the particle number equation (3.160). The result
is simply the Hartreee-Fock-Bogoliubov result.
Let us calculate this explicitly. First for repulsive g < 0 where the solution of
the first gap equation in Eq. (3.159) is trivial: = 0. Then the momentum sum
is carried out from zero to the place where p changes sign and p, vanishes. This
happens at the -dependent modified Fermi momentum pF determined by
pF
The sum yields
E (0) =

2M
.

p (p, + p, ) = 2V

(3.164)

p<pF

d3 p p2
.
(2)3 2M

(3.165)

Recalling Eq. (3.89) this becomes


4 5
V
p .
3
(2h) 5m F

E (0) =

(3.166)

so that
g N
V
4 5
pF V
h0|H|0i =
3
(2h) 5m
2 V


2

(3.167)

We now determine the size of from the particle number by performing the sum
N =V

(p, + p, ) = 2V

p<pF

d3 p
.
(2)3

(3.168)

Recalling Eq. (3.90), this is equal to


N
1 pF 3
= 3 2,
V
h
3

(3.169)

The result (3.170) can also be expressed in a dimensionless equation with the help
of (3.94) as
h0|H|0i = N

3 pF 2 g 1 pF 3

5 2m 2 h
3 3 2

= NF

3 g M pF

.
5 2h
3 3 2
!

(3.170)

Expressing the attractive coupling strength g in terms of the s-wave phase shift by
a formula g = 4h2 as /M to be derived in (9.264), this becomes
h0|H|0i = NF

3
2pF
+ 2 as .
5 3 h

(3.171)

For a repulsive interaction with as > 0, the total energy is increased above the
free-particle energy. For an attractive interaction the chemical potential increases:

= + g/2.

(3.172)

204

3 Interacting Non-Relativistic Particles

From here on we shall assume that such a shift has been performed without
exhibiting it in the notation.
For an attractive interaction with g > 0 the gap equation (3.159) has a nontrivial
solution. It determines the gap by solving
Z
1
1 X 1
d3 p
1
q
=
=
.
g
V p 2Ep
(2h)3 2 p2 + 20

(3.173)

We have added a subscript 0 to the energy gap to emphasize the that this is the
gap at zero temperature.
The integral diverges. In an old-fashioned weak-coupling superconductor, the
attraction is caused by phonons whose energies are restricted to a small range
(D , D ). to a thin layer around the surface of the Fermi sphere where restricts
the momentum integral |pF |
p| pF |/m h
D . If we change variables from p to
and write
Z D
d3 p

N
(0)
d
(2h)3
D

(3.174)

with
N (0) =

3 N
mpF
3 =
2
4F V
2 h

(3.175)

being the density of states at the surface of the Fermi sea (the right-hand part of
the equation being valid only for spin 1/2), we arrive at the gap equation
1
N (0)
g

2D
1
= N (0) ln
d q
.
2
0
2 + 0

(3.176)

Thus the energy gap is given by


0 2D e1/gN (0) .

(3.177)

This expression is non-perturbative in an essential way. It cannot be expanded in a


power series in g.

3.4

Renormalized Theory at Strong Interactions

In modern high-temperature superconductors where the attraction is strong, the


restriction to the thin layer around the Fermi sphere is no longer valid, and the
integral has to be done over the entire momentum space. Then the gap equation
(3.159) reads
=

g X
.
V p 2Ep

(3.178)

205

3.4 Renormalized Theory at Strong Interactions

This is a divergent quantity. The divergent part is


gdiv =

g X
.
V p 2p

(3.179)

It can be removed from the theory by adding a divergent so-called mass counterterm
to the interaction Hamiltonian (3.125):
Hdiv = gdiv

X
p

ap, ap, + ap, ap, .

(3.180)

This term must be added to Eq. (3.154) changing it to


h0|H|0i =

X
p

2 p vp2


X
g X
up vp up vp + vp2 vp2 + gdiv
up vp .
V p,p
p

(3.181)

Variation replaces the former equation (3.155) by


2p vp



g X
g X 2
up vp up vp2 /up 2
vp vp +gdiv (up vp2 /up) = 0. (3.182)
V p,p
V p,p

If this is multiplied by up and summed over p , it turns into




0 2 (p g ) up vp g( div ) u2p vp2 = 0,

(3.183)

and (3.156) becomes


0 (p g ) p = g( div )(p p ).

(3.184)

Thus Eqs. (3.158) turn into the renormalized equations


p = g ,

p = g( div ) = gR .

(3.185)

and the self-consistent gap equation (3.159) turns into the renomalized gap equation
!

g X 1
1
= gR =
,

V
2Ep 2p
p

(3.186)

so that the (3.173) is renormalized to a finite equation:

1
1
1 Z d3 p
q

=
.
g
(2h)3 2 p2 + 20 22p

(3.187)

The total energy (3.181) becomes


h0|H|0i =

X
p

2
2 (p g ) vp2 gR2 + gdiv
.

(3.188)

206

3 Interacting Non-Relativistic Particles

The last term must be removed at the end by an additive renormalization of the
vacuum energy. See the discussion of this subtraction in Section 11.7.
It is worth emphasizing that the equation for p needs no mass counterterm
since the momentum sum in (3.147) has no divergence.
The subtraction of a divergence as in (3.187) was encounterd before in the Bose
gas in Eq. (3.58), and it was pointed out that the renormalaized coupling constant
gR can be obeserved exprimentally via the s-wave scattering length as in the twobody scattering process, the relation being 1/gR = M/4h2 as where as < 0 in the
attractive case. This brings the subtracted gap equation (3.187) to the useful form

Z
M
1
d3 p
1
q

=
.

2
(2h)3 2 p2 + 20 22p
4h as

3.4.1

(3.189)

Finite Temperature

It is easy to do the same calculation for finite temperature T . At finite T , the


expectation values (3.140)(3.145) become

i = vp2 + hu2pp p vp2 p


p i,
hap, ap, i = hu2p p p + vp2 p p

(3.190)

p 1i, (3.191)
i = up vp hp p + p
hap, ap, i = up vp hp p p p

hap, ap, i = up vp hp p
+ p p i = up vp hp p + p
p 1i. (3.192)

The thermal expectation values of the quasiparticle densities p p and p p are


given by the Bose occupation numbers
hp p i = hp p i = f f (Ep ) =

1
eEp /T

(3.193)

(3.194)

+1

so that


hap, ap, i = hap, ap, i = vp2 + u2p vp2 f f (Ep ),

hap, ap, i

= hap, ap, i = up vp [1 2f f (Ep )].

(3.195)

The momentum sums (3.160) and (3.159) are now given by



i
1 Xh 2  2
vp + up vp2 f f (Ep ) ,
V p
1 X
up vp [1 2f f (Ep )].
=
V p

= =

(3.196)
(3.197)

207

3.4 Renormalized Theory at Strong Interactions

Inserting (3.193), these turn into the following T 6= 0 versions of and [compare
(3.160) and (3.159)]
p g
E
d3 p
tanh
1
=
3
(2h)
Ep
2T


Z
3
1
E
dp 1
= g
tanh
.
3
2 (2h) Ep
2T
1
=
2

"

#

(3.198)
(3.199)

With these, the energy (3.153) takes the form [respecting (3.158)]
hHiT =

X
p

2Ep
p (np, + np, ) V 2 .
+

eEp /T + 1
g

(3.200)

For old-fashined weak-coupling superconductors the gap equation (3.173) receives at nonzero temperature an extra factor 1 f f (Ep) so that it becomes
1
=N
g

d
E
.
tanh
2T
2 + 2

(3.201)

It is solved by writing it as
"Z
#

Z D
D
d
1
E
d
2
2
= N (0)
1 .
tanh
+
g
2T
0
+ 2
+ 2
0

(3.202)

The first integral gives


2D
2D
0
N(0) ln
,
= N (0) log
+ log

(3.203)

where 0 is the zero-temperature gap and the gap at finite temperature. In the
second integral we expand
tanh

X
E
eE/T
()n+1 enE/T .
1 = 2 E/T
= 2
T
e
+1
n=1

(3.204)

The integrals over the terms in the series are all rapidly convergent so that we can
take the upper limit D to infinity in each of them. Then we may use the integral
representation of the associated Bessel function K0 (z):
Z

d
eE/T = K0 (/T ),
+ 2

(3.205)

to obtain the gap equation at all temperatures

X
n(T )
0
ln
.
= 2 ()n+1 K0
(T )
T
1

(3.206)

208

3 Interacting Non-Relativistic Particles

For large arguments


2K0

0
T

1 q
2T 0 e0 /T ,
(0)

(3.207)

so that the first correction to the low-temperature value of the gap is exponentially
small
(T ) = 0

1 q
2T 0 e0 /T .
(0)

(3.208)

For increasing T , the gap decreases. It vanishes at the critical temperature Tc .


Near Tc , the gap equation is most conveniently studied by expanding
2
X
1
+ 2
1
2
=
T
tanh
2
2
2
2T
2 + 2
m m + +
"
#
X
1
2
4
=T
2
+ 2
+ . . . , (3.209)
2 + 2
m
(m + )2 (m
+ 2 )3
m
where
m = 2T m.

(3.210)

are the Matsubara frequencies for the temperature T . Performing now the integrals
over in (3.202) gives the gap equation
1

1
1
d

2 X
3 4 X
tanh
=
+
+ . . . . (3.211)

2T
(T )2 m (2m+1)3 8 (T )4 m (2m+1)5

The integral is calculated as follows. A partial integration yields


Z

D 1 Z


d log cosh2
tanh
= log tanh
.

2T
T
2T 0
2 0
T
T
2T

(3.212)

Since D /T 1, the first term is equal to log(D /2T ), with exponentially small
corrections which can be ignored. In the second integral, we have taken the upper
limit of integration to infinity since it converges. We may use the integral formula11
Z

dx


x1
4 
2
()( 1),
=
1

2
(2a)
cosh2 (ax)

(3.213)

set = 1 + , expand the formula to order , and insert the special values
1
(1) = , (0) = log(2) log(4e /),
2
11

(3.214)

See, for instance, I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series, and Products,
Academic Press, New York, 1980, Formula 3.527.3.

209

3.4 Renormalized Theory at Strong Interactions

to find from the linear terms in :


Z

dx

log x
= 2 log(2e /),
2
cosh (x/2)

(3.215)

so that we obtain
Z

D 2e
d
.
tanh
= log

2T
T


(3.216)

The sums over m in (3.211) can be expressed in terms of Riemanns zeta function
(2.276) as
X
m

1
2z 1
=
(z)
(2m + 1)z
2z

(3.217)

We now use the zero-temperature gap equation


1
2D
= N (0)
g
0

(3.218)

to eliminate 1/g in (3.211). This brings the left-hand side of (3.211) to the form
log(T /Tc ) so that the gap equation reads
log

1
1
2 X
3 4 X
T
=
+
+ ... .
2
3
4
Tc
(T ) m (2m+1)
8 (T ) m (2m+1)5

(3.219)

where we have set


Tc =

e
0
0
,

1.76

(3.220)

which is the critical temperature at which the gap vanishes. For T closely below the
critical temperature Tc , we may keep only the first term on the right-hand side of
(3.219) and find that the gap behaves like
8
T
.
(T )
2 Tc2 1
7(3)
Tc
2

(3.221)

The full solution is plotted in Fig. 3.9. The existence of an energy gap in superconductors was first confirmed experimentally by millimeter-microwave absorption
in lead films below Tc 7.20 K. The result was 20 3.4Tc which is in good
agreement with the theoretical values 2 1.76. The energy gap is the reason
behind the extraordinary stability of electric currents in a superconductor. In practical situations, the lifetime is often larger than 105 years. Superconductivity has
not yet been exploited for long-distance transport of industrial electric currents because of the need to maintain wires at extremely low temperatures, for instance
via liquid 4 He. Until a few years ago, the superconductor with the highest transition temperatures were alloys of Niob and Zirkonion, or the compound Niobstannid

210

3 Interacting Non-Relativistic Particles

(T )2 /20

T /Tc

Figure 3.9 Solution of gap equation for weak attraction between electrons.

Nb3 Sn, which has Tc 18.20 K, and remains superconductive up to magnetic fields
of 20 Tesla. Superconductive coils have been used with great advantage in highenergy particle accelerators and in nuclear magnetic resonance devices, which have
found important application in medicine (tomography). With the recent discovery of superconductivity at temperatures up to 1000 K, and with the possibility of
more economic cooling via liquid nitrogene, superconductivity will becomes useful
for broader industrial applications.

3.5

Crossover to Strong Couplings

If the coupling strength increases, the above approximation of integrating only over
a thin shell around the surface of the Fermi sphere is no longer applicable, but the
momentum integral must run over the entire momentum space. Then the subtracted
gap equation (3.189) becomes
"

2
2
1
=

kF as
3J2 (
)
where

/ and
J1 (z)

J2 (z)

#1/3

J1 (
),

(3.222)

1
dx x2 q
2,
(x2 z)2 + 1 x

x2 z

.
dx x2 1 q
(x2 z)2 + 1

(3.223)
(3.224)

The particle number equation (3.160) which fixes the chemical potential
reads
3
1=
2 F


3/2

J2 (
).

(3.225)

211

3.5 Crossover to Strong Couplings

The two equations can be solved numerically yielding the gap and the chemical potential
= +g/2 as a function of the inverse s-wave scattering length 1/kF as from
the weak-coupling (left) to the strong-coupling (BEC) limit as shown in Fig. 3.10.
1.0
0.5

/F
2.0

/F

-2

-1

1.5
-0.5
1.0
-1.0
0.5
-1.5

-2

BSC

-1

1/kF as

-2.0

BEC

BSC

1/kF as

BEC

Figure 3.10 Plot of the gap function and of the chemical potential against the inverse
s-wave scattering length.

At finite temperature, we define the natural temperature TF F /kB and the


reduced temperature t T /TF , and the gap euquation becomes
"

1
2
2

=
kF as
3J2 (
, t)

#1/3

J1 (
, t),

(3.226)

where
J1 (z, t)

J2 (z, t)

x2
1
tanh
2,
dx x q
2t x
(x2 z)2 + 1
2

2 z

2
d 2 1 q
tanh .
2t
(2 z)2 + 1

(3.227)
(3.228)

while the the particle number equation (3.160) that fixes the chemical potential

reads
3
1=
2 F


3.5.1

3/2

J2 (
, t).

(3.229)

Bogoliubov Bose Gas at Finite Temperature

It is easy to extend Bogoliubovs weak-coupling theory to finite temperature T .


There the expectation values (3.31) become

hap ap i = vp2 + hu2p p p + vp2 p


p i,

(3.230)

and

hap ap i = up vp + up vp hp p + p
p i.

(3.231)

212

3 Interacting Non-Relativistic Particles

The thermal expectation values of the quasiparticle densities p p are given by


the Bose occupation numbers
1

hp p i = f b (Ep ) =

eEp /T

so that


(3.232)

hap ap i = vp2 + u2p + vp2 f b (Ep ),

(3.233)

hap ap i = up vp [1 + 2f b (Ep )].

(3.234)

The momentum sum of the first expectation value gives the temperature-dependent
density of uncondensed particles
u =


i
1 Xh 2  2
vp + up + vp2 f f (Ep ) .
V p

(3.235)

Inserting (3.232), this becomes


u

1
=
2

d3 p
p + g0
Ep
coth
3
(2h)
Ep
2T
"

1 .

(3.236)

Hence the density of uncondensed particles (3.47) is

q
2Mg0
3 Iu (t)
u (t) = kg0
, kg0 =
= 2 2as 0 ,
4 2
h

(3.237)

where Iu (t) is the integral


Iu (t)

+1

d 2 q
coth
2
2
( + 1) 1

(2 + 1)2 1
2t

where t is the reduced temperature

1 =

2
h (t), (3.238)
3 u

t kB T /g0 ,

(3.239)

and
g0

2
h
2 kg
4h2
0
=
(as 0 )2 .
2M
M

(3.240)

Hence
3
u (t) = u (0)hu (t) = kg
0

Iu
80 q 3 1/2

a 0 hu (t).
(t)
=
h

4 2 u
3 s

(3.241)

We may also write


3
hu (t) =
2

2 + 1

1+ 2 2
d 2 q
1 . (3.242)
(
+1)
1/t
2
2
( + 1) 1
e
1

213

3.5 Crossover to Strong Couplings

so that
3
hu (t) 1 + hu (t) = 1 +
2

2
2 + 1
2 2
. (3.243)
d 2 q
(2 + 1)2 1 e ( +1) 1/t 1

After introducing some new variables y = 2 /t and x =


q

(2 + 1)2 1/t, yt =

(xt)2 + 1 1, so that d 2 = tdy (ty)1/2 /2 and dy = dx xt/(yt + 1), we may


rewrite hu (t) in the following form:
q
Z

(xt)2 +11)1/2
(
3
t
dx
.
(3.244)
hu (t) = 1 +
2 0
ex 1

The corresponding formula for the full uncondensed density (3.237) is


q

Z
2 +11)1/2
3
3
(xt)
(
k
kg
2
g0 2
0
. (3.245)
hu (t) =
+t
dx
u (t) =
2
4 3
4 2
3
ex 1
0
For weak coupling, g0 becomes very small, so that the integral (3.243) is dominated by the large- regime, and we find
6
hu (t)
2

3
=
d 2 /t
e
1
2
2

3/2
t (3/2),
2

(3.246)

so that
 3/2
3
kg
80
T
3/2
0
u (t) = u (0) + 2 Iu hu (t) =
,
as 0 +
T 0
4
3
Tc

(3.247)

with [recall (2.333) for zero spin]


Tc [(3/2)]2/3 4Ta ,

(3.248)

where
Ta a /kB ,

a h
2 /2Ma2 .

(3.249)

The ratio T T /Ta is the temperature measured in units of the fundamental


B T = 1/t.
temperature Ta , so that s/T = /k
The temperature Tc at zero coupling strength, agrees with the condensation
temperature of free Bose gases determined in Eq. (2.337).
To plot the full temperature behavior it is convenient to introduce the dimensionless parameter s g0 /a . Now we rescale the momentum integration, and
obtain
hu (t) =

3/2
s
2

2
2 + s
2 2 2
d 2 q
,
(2 + s)2 s2 e ( +s) s /T 1

(3.250)

where T T /Ta is the temperature measured in units of the fundamental temperature Ta . The plots as shown in Fig. 3.11.

214

3 Interacting Non-Relativistic Particles


1.0

0.3 0.2 0.1 0.05

0.01

s=0

0.8

u /

0.6

0.4

0.2

0.0
0.0

0.2

0.4

0.6

0.8

1.0

T /Tc

Figure 3.11 Temperature dependence of the uncondensed fraction u / in Bose gas.


The temperature T is measured in units of the natural temperature Ta , where Ta is the
fundamental temperature Ta = a /kB = h
2 /2M a2 kB . The critical temperature of the free
Bose gas lies at Tc = Tc /Ta = [(3/2)]2/3 4. The parameter s on theqvarious curves is the

dimensionless ratio s g0 /a of the quasiparticel energies Ep = (p + g0 )2 g2 20 .


The transition to the normal phase takes place when the curves reach unity. Note the
nonzero intercepts on the vertical axis, showing the depletion of the condensate at zero
temperature.

3.6

Bose Gas at Strong Interactions

We are now prepared to carry also the Bogoliubov theory of a weakly interacting
spinless Bose gase to strong couplings. For this we proceed as Eq. (3.124) and add
to the free Hamiltonian (3.4) variational energy terms with real p


i
Xh
trial 1
H
p ap ap + ap ap +p ap ap + p ap ap .
2 p

(3.251)

In contrast to the Bogoliubov theory we work in a grand-canonical ensemble and


allow for a chemical potential 6= 0 and rewrite the free part of the Hamiltonian as
i


Xh
0 1
(p +p) ap ap + ap ap +p ap, ap, +p ap ap .
H
2 p

(3.252)

To this we add the zero-momentum part (3.17) of the free Hamiltonian (3.17), and
are faced with
H0 V


X
X

V
1
0 + 20 + (p +20 +p )ap ap + (0 +p )
ap ap +h.c. .
g
2g
2
p6=0
p6=0

(3.253)

An important difference with respect to Bogoliubovs theory, the zero-momentum

operator a0 is not approximated by a scalar A0 that is identified as V 0 , where 0


is the condensate density. Instead, we shall take the expectation value A0 h(x)i
P
of the field (x) = p eipx/h ap and introduce a parameter
0 gA20 /V,

(3.254)

215

3.6 Bose Gas at Strong Interactions

leaving its physical meaning open, and considering it as a variational parameter to


be determined in each order in perturbation theory by optimization. Optimization
will mostly proceed by extremizing the energy with respect to 0 . Only after that,
0 will acquire a physical meaning. To lowest order, 0 will turn out that to be
equal to g times the condensate density 0 , as in Bogoliubovs theory, but to higher
orders, there will be corrections.
After a Bogoliubov transformation we arrive at the Hamiltonian
H0 = V

o
X n

V
[p + 20 + p ] vp2 + (0 + p )up vp
(3.255)
0 + 20 +
g
2g
p6=0



o

1n

p
[p + 20 + p ] u2p + vp2 + (0 + p )2upvp p p + p
2
!
o
o 

1n

2
2
+
p p + p p
[p + 20 + p ] 2upvp + (0 + p ) up + vp
,
2

At zero temperature when all quasiparticles are frozen out only the first line of
this energy survives. This will be denoted by W0 and provides us with the lowestorder variational energy of the system. The energy W0 is now extremized with
respect to the Bogoliubov coefficients up and vp while maintaining the constraints
u2p vp2 = 1, so that up = vp vp /up , as in the previous treatment of Eq. (3.41).
To this we must add the expectation value of the interaction energy (3.15):

hHint
i=

g
2V

p6=0,p 6=0,q6=0

hap+q ap q ap ap i.

(3.256)

As in Eq. (3.149) we can use the commutation rules12 (2.210) to expand the righthand side into the three pair terms
E
D


0 ap+q ap q ap ap 0

=
+
+

E
ED
D




0 ap+q ap q 0 0 ap ap 0
ED
E
D




0 ap+q ap 0 0 ap q ap 0
E
ED
D
0 ap+q ap 0 0 ap q ap 0 .

(3.257)

Note the opposite sign of the last term with respect to the fermion expression (3.152).
Inserting the expectation values (3.31), this becomes
E
D


0 ap+q ap q ap ap 0

= p,p up+q vp+q up vp + 2q,0 vp2 vp2

(3.258)

and we find

hHint
i=
12

g
2V

p6=0,p 6=0

up vp up vp + 2vp2 vp2 .

More elegantly we can use Wicks theorem to be derived in Section 7.14.1.

(3.259)

216

3 Interacting Non-Relativistic Particles

trial . The
Finally we must subtract the expectation value of the trial Hamiltonian H

total variational ground state energy Etot = E0 + hHint i hHtrial i is now

Etot
= V

g
2V

X
V

0 + 20 +
[(p + 20 +p ) vp2 + (0 + p )up vp ]
g
2g
p6=0
X

p6=0,p 6=0

up vp up vp + 2vp2 vp2

(p vp2 + p up vp ).

(3.260)

p6=0

According to the rules of Variational Perturbation Theory we shall write this as


W1 = W0 + W11 + W12 , where W0 denotes in the first line of (3.260) that constitutes
the zeroth-order variational energy, and W11 +W12 are the two term in the the second
line. It is important to treat at least the energy W12 as a first-order perturbation.
Otherwise the variational parameters p and p cancel completely and do not
improve the ground-state results.13
By differentiating the energy with respect to and setting the result equal to
N we find for the density = N/V :
=

0 X 2
+
vp .
g
p6=0

(3.261)

As in Eq. (3.47), the second term is the density of excited, uncondensed particles
u , so that (3.261) reveals 0 /g as the condensate density 0 ,
0
= 0 = u ,
g

(3.262)

as before in Bogoliubovs theory.


Next we extremize W1 with respect to the variational parameter 0 which yields
the equation
X
0 X 2
=
(2vp +upvp ) = 2u + up vp = 2u +,
g
p6=0
p6=0

(3.263)

= + u + .
g

(3.264)

so that

We are now able to fix the size of the Bogoliubov coefficients up and vp . The
original way of doing this is algebraic, based on the elimination of the off-diagonal
elements of the transformed Hamiltonian operator. In the framework of our variational approach it is more natural to use the equivalent procedure of extremizing
13

The first term in W11 can be split into two parts 0 W11 + (1 0 )W11 considering the first
part as belonging to W0 , the second part to the perturbation. The results of this splitiing will be
independent of 0 .

217

3.6 Bose Gas at Strong Interactions

the energy W0 with respect to up and vp under the constraint u2p vp2 = 1, so that
up /vp = vp /up. This yields for each nonzero momentum the equation


2 (p +20 + p ) vp + (0 + p ) up +vp2 /up = 0.

(3.265)

In order to solve this we assume p to be independent of p so that we introduce


a constant
+ 20 + .

(3.266)

In the prefactor in the second term of (3.265) we assume that also p is independent
of p which we may set equal to , so that we may define the constant
0 + .

(3.267)

and rewrite (3.265) in the simple form




vp +
up + vp2 /up = 0,
2 p +

(3.268)

which is solved for all p by the new Bogoliubov transformation coefficients


u2p

p +
1
p +
1
1+
, vp2 =
1
,
=
2
Ep
2
Ep
!

(3.269)

with up vp = /2E
p , and the new quasiparticle energies
Ep =

r


2

p +

2,

(3.270)

According to the Nambu-Goldstone theorem, these have to vanish linearly for p 0.


= ,
or
This forces us to set
0
=

(3.271)

thus avoiding the main hurdle in previous attempts to go beyond the Bogoliubov
theory [21].
Having determined the Boliubov coefficients we may now calulate the momentum
sums in Eq. (3.264). First, there is the density of uncondensed particles which is
evaluated in the same way as before expression (3.47) of Bogoliubovs theory. We
insert vp2 of (3.269) into the momentum sum and arrive at
u =

hap ap i
p6=0
X

Z
Z
1 X 2
1
d3 p
d3 p 2
=
v
=
vp =
V p6=0
(2h)3 p 2 (2h)3

p +
1 . (3.272)
Ep
!

Performing variable substitutions as before in (3.47) we obtain


u = k3

Iu
,
4 2

(3.273)

218

3 Interacting Non-Relativistic Particles

where Iu = 2/3 is the same number as before in (3.52), whereas k is defined as

in (3.49), but the energy g0 replaced by the new energy :

2M
k =
.
(3.274)
h

The other momentum sum in Eq. (3.264) becomes, due to (3.269),


=

hap ap i =

p6=0

Z
Z d3 p 1

1 X
d3 p
.
u
v
=

up vp =
p
p
V p
(2h)3
2 (2h)3 Ep

(3.275)

This is a divergent quantity. The divergence can be removed in the same way as
before the divergence in the energy (3.57), by absorbing it into the inverse coupling constant of the model using Formula (3.58). Thus we introduce the finite
renormalized quantity
Z
Z d3 p
1 X

d3 p
R =
up vp =
up vp =
V p
(2h)3
2 (2h)3

1
1

,
E p p

(3.276)

and write
= R + div

(3.277)

where the infinity is contained in the momentum sum


div

Z d3 p 1
X 1

=
V p 2p
2 (2h)3 p

(3.278)

Let us denote the divergence by


1
=
Vv

d3 p 1
.
(2h)3 p

(3.279)

then we have
=

= R + div = R
,
g
2V v

.
gR

(3.280)

Inserting (3.277) together with (3.277) into (3.264), we find


0

= 2u + R + div ,
g
g

(3.281)

Recalling (3.262), this implies

= 0 + 2u + R + div = + u + R + div .
g

(3.282)

We evaluate the convergent momentum sum (3.276) in the same way as in (3.272)
and find
R = k3

I
.
4 2

(3.283)

219

3.6 Bose Gas at Strong Interactions

where I is given by an integral similar to (3.52):


I

1
1
2 = 2.
d 2 q
(2 + 1)2 1

(3.284)

Hence we see that


R = 3u .

(3.285)

Before we continue, we may simplify all diverging expressions by making use of


the renormalizability of the theory. Since we only have added and subtracted only
terms in the Hamiltonian (3.260) which are quadratic in the fields, and contain at
most two spatial derivatives (this only if p and p contain a nonzero term p
and p ), all infinties can be absorbed in the initial parameters of the Hamiltonian.
This property will be discussed in more detail in Chapter 11 and 21. In such theries,
we may eliminate all divergencies simply by simply using Veltmans rule (11A.1),
which amounts here to setting
Z

d3 p

1
= 0,
p

(3.286)

and thus div = 0. After this we may identify all quantities directly with the renormalized, observable ones.
It is further useful to introduce a natural length scale, the average distance per
particle a, that makes the particle density equal to = 1/a3 . Associated with it we
introduce the natural energy scale
a

h
2
,
2Ma2

(3.287)

and the reduced s-wave scattering length


a
s 8

as
,
a

(3.288)

in terms of which the renormalized coupling constant is


gR =

4h2
as = 8a a2 as = a a3 a
s .
M

(3.289)

After this, we re-express all equations in a dimensionless form, for instance


s a ,

(3.290)

s
s
=
= 3 .
2
gR
8a as
aa
s

(3.291)

which leads to
k =

s
,
a

220

3 Interacting Non-Relativistic Particles

=
we see that the parameter

From the quasiparticle spectrum (3.292) with


determines the slope in the gapless energy spectrum, which we may write as
Ep =

p + 2 .
2
p

(3.292)

This fixes the velocity of second sound to


c

,
M

(3.293)

thereby generalizing (3.64). Inserting (3.290) and (3.287), we find its dimensionless
form
c=

s
va ,
2

va

pa
h

.
M
aM

(3.294)

In these units, the uncondensed particle density (3.273) reads


u
I
= s3/2 u2 .

(3.295)

As a result we can write Bogoliubov-transformed energy (3.260) as


W1 = V



2
V

1 + gV 2u 2 + 2 V (u + ),(3.296)
0 + 20 + V w()
g
2g
4Vv
2

where we have abbreviated



w()



2 
1 X 
+ .
Ep p
2V p6=0
2p

(3.297)

The subtracted term is the analogue of the last term in the original Bogoliubov
energy (3.57), which was performed to make the momentum sum convergent. It can
be evaluated in the same way as in (3.59) yielding a similar result as in (3.298)
= k
3 IE ,
w()

4 2

(3.298)

where IE = 8 2/15 is the previous integral (3.61).


The parameter still needs to be determined. For this we must extremize W1
+ 20 from
with respect to . This is done most easily by inserting
0 from (3.271) into (3.296) to write W1 as
(3.266), and
W1 = V
+

2
V

3 IE 1
0 + 20 + V k

g
2g
4 4 4 V v


h
i
gV  2
+ 20 )u +(
0 ) .
2u + 2 V (
2

(3.299)

221

3.6 Bose Gas at Strong Interactions

To do this, we use the fact that at zero


and extremizing this with respect to .
temperature
3 IE = k 3 Iu + I
= k
w()

4 2
4 2 4 2


= u + ,

(3.300)

denoted by a prime:
so that we obtain the derivative with respect to ,
i

+ 20 ) +(
0 ) ,
W1 = gV (2u u + ) V (
u
Setting W1 equal zero this yields

+ 20 2gu )u +(
0 g) = 0,
(

(3.301)

(3.302)

and inserting 0 and from Eqs. (3.262) and (3.282) (the latter after having applied
at T = 0:
Veltmans rule), we obtain the equation for
g( u + )]u +[
g( u )] = 0,
[

(3.303)

which is solved by


= u u
= 0.
g
u +

(3.304)

Inserting u from (3.273) and from (3.283), and using (3.285), this reduces to

1
= + u
g
2

(3.305)

In natural units, this amounts to the equation


s
1 Iu
=1+
,
a
s
2 4 2

(3.306)

which is solved explicitly by


a
s =

s
1 + 21 s3/2

2 1
3 4 2

(3.307)

For a plot of this relation see Fig. 3.12a.


Equation (3.306) can be solved for s as a function of the reduced s-wave scattering
length a
s 8as /a by
s=a
s +

1
1
7
a

a11/2
+ O(a7s ) .
5/2
a4s +
s
s +
4
192
12 2
9216 2 6

(3.308)

Inserting this into (3.295) yields


1 3
u
1
7

+
a
9/2 + O(a6s ).
= 2 a3/2
a

+
s
s
4
6

96
6 2
4608 2

(3.309)

222

3 Interacting Non-Relativistic Particles


4
15

u /

10

a
s = /
5

a)

200

400

600

800

a
s 8as /a

b)

10

15

a
s 8as /a

a and uncondensed particle fraction as functions of


Figure 3.12 Reduced gap s /
the reduced s-wave scattering length a
s = 8as /a = 8as 1/3 . At as where u / = 1, the
gas becomes normal.

The first term agree with Bogoliubovs weak-coupling calculation (3.54) of the depletion of the condensate due to interactions. The general behavior up to strong
couplings is shown in Fig. 3.12b.
We now calculate the total energy (3.299) at T = 0. We go to natural units
and intoduce the reduced quantities u u / = s3/2 Iu and R / = s3/2 I , to
obtain the reduced variational energy w1 W1 /Na in the form:
a
s

w1 = as (1+ u + )(1
u )+ (1 u )2 +s5/2 IE
2
a
s

(2
2u + 2 ) a
s ( u + ),
+
2

(3.310)

and
with s/as 1 + 3
u + and s/as 1 + u . Inserting u and ,
going from the grand-canonical to the true proper energies by adding N to W1 .
and forming W e = W1 + V , we obtain the reduced energy

a
s
2
2
a
s 3
e
3/2
w1 =
+ 2a
s s 2 s5/2 +
s.
(3.311)
2
3
5
72 4
Inserting here the expansion (3.308), we find that up to the term a
4s the energy
has the expansion

as
2 2 5/2
1 4
e
w1 =
+
a
s +
a
+ . . . ..
(3.312)
2
2
15
72 4 s
The first two terms agree with the corresponding terms in Bogoliubovs weakcoupling result (3.62):

2 2 5/2
a
s
+
a
.
(3.313)
wBog =
2
15 2 s
The result is plotted in Fig. 3.13.
The accuracy can be increased to any desired level, with an exponentially fast
convergence, as was demonstrated by the calculation of critical exponents in all

223

3.6 Bose Gas at Strong Interactions

150

100

w1e = W1 /N a
50

wBog
5

10
15
a
s 8as /a

20

25

Figure 3.13 Reduced energy per particle w1e = W1 /N a as a function of the reduced
s-wave scattering length a
s = 8as /a compared with Bogoliubovs weak-coupling result
(3.313). There is a continuous phase transition at (
as , s) (16.08, 48.23), where the
condensate becomes depleted and the ratio u / reaches unity.

Euclidean 4 theories with N components in D dimensions [5]. The fact that the
theory is renormalizable, so that all divergencies can be removed by Veltmans rule,
is an essential advantage of the present theory over any previous strong-coupling
scheme.14
Let us also study the temperature dependence of the depletion equation (3.295)
and of Eq. (3.272). For this we proceed as in Eqs. (3.236), and introduce the
temperature-dependent integral (3.238) to find, instead of (3.237) the result
u (t)

I (t)
= k3 u 2 ,

k =

2M
s
=
,
h

(3.314)

and t is now the reduced temperature


= kB T /sa .
t kB T /

(3.315)

It will be useful to rewrite Iu (t) in a form analogous to (3.238), (3.242) as

Iu (t) =

2
2
hu (t) =
+t
3
3

dx

q

(xt)2 +11
ex 1

1/2

(3.316)

and
su (t)
14

u
s3/2
=
I (t).

4 2 u

(3.317)

Our results can be made more reliable by calculating the contribution of the still-missing second
two-loop diagram, the second in Eq. (3.741) of the textbook [13]. Its contribution would be the
3 + 1-dimensional version of the last term in Eq. (3.767), is essential in the X 6= 0 phase. Without
this term, the slope of the quantum-mechanical energy as a function of the coupling constant is
missed by 25%, as discussed in the heading of Fig. 5.24.

224

3 Interacting Non-Relativistic Particles

In addition we need a nonzero temperature version of the integral I of (3.276)


and (3.284):
Z

I (t)

1
q

(2 + 1)2 1

coth

(2 + 1)2 1
2t

1
.
2

(3.318)

to find
R s3/2
= 2 I (t).
Rs (t)

(3.319)

Here we write alternatively

I (t) 2h (t) =

or
I (t)

1
1+ 2 2
d 2 q
2 , (3.320)

(2 + 1)2 1
e ( +1) 1/t 1

2 + I (t) =

2
1
2 2
. (3.321)
d 2 q
(2 + 1)2 1 e ( +1) 1/t 1

A convenient formula of the type (3.316) is


I (t) =

2t

( (xt)2 + 1 1)1/2

.
dx q
(xt)2 + 1(ex 1)

(3.322)

from which we obtain


3/2
R (t) = s I (t).
(t)

4 2

(3.323)

Finally we must calculate the nonzero temperature version of the energy (3.297),
which replaces (3.298) by
t) V k
3 IE (t) ,
w(,

4 2

(3.324)

which amounts to the reduced energy


w s (s, t)

t)
w(,
s5/2
= 2 IE (t),
a
4

(3.325)

where IE (t) is the integral


IE (t)

d 2 (2 +1)2 1 coth

(2 +1)2 1
2t

1
2 1+
, (3.326)
22

225

3.6 Bose Gas at Strong Interactions

with

Z
q
8 2
2
8 2
d 2 (2 + 1)2 1 2 2
+ IE (t) =
+
IE (t)
.(3.327)
15
15
0
e ( +1) 1/t 1
A formula of the type (3.316) for IE (t) is:

8 2
+t
15

IE (t) =

( (xt)2 + 1 1)1/2
2

dx (xt) q
,
(xt)2 + 1(ex 1)

(3.328)

Now we can derive the finite-temperature version of Eq. (3.304). For this we

must calculate the -derivatives


of (3.314) and (3.319), and we find using (3.316)
and (3.322) the expressions
u (s, t)

s1/2 3
I (t) tIu (t) ,
=
a 4 2 2 u


(3.329)

where
d
Iu (t) Iu (t) =
dt
and

3(xt)2 + 2 2 (xt)2 + 1

q
.
dx q
2 (xt)2 + 1( (xt)2 + 11)1/2 (ex 1)

R (s, t) =

(3.330)

s1/2 3
I (t) + tI (t) ,
a 4 2 2


(3.331)

where
d
I (t) I (t) =
dt

Similarly we find

[(xt)2 + 2] (xt)2 + 1 2

dx q
.
3 q
2 (xt)2 + 1 ( (xt)2 + 11)1/2 (ex 1)

w (s, t) =

(3.332)

1 s3/2 5
IE (t) tIE (t) ,
a 4 2 2


(3.333)

with
d
IE (t) IE (t) =
dt

dx

3(xt)2 ( (xt)2 + 11) + (xt)4


((xt)2 + 1)3/2

q

 q
5
2

(xt)2 + 11

(xt)2 + 12

1/2

(ex 1)

. (3.334)

At finite temperature, the relations (3.285) and (3.300) is no longer true. The latter
relation must be replaced by the frequency sum
1
3 IE (t) = k 3 Iu (t)+I (t) (s,t) = u (s,t)+(s,t)(s,t),(3.335)
t)= k
w(,

V
4 2
4 2

226

3 Interacting Non-Relativistic Particles

where is yet another frequency sum


Z
1
4
1 X p
s3/2
q
 . (3.336)
d
(s, t) =
=

4V p kB T sinh2 (Ep /2kB T )


2t
sinh2
(2 +1)2 1/2t

Using these we calculate, instead of (3.304), the finite-temperature relation be and g:


tween

1
u
+
=

.
u
g
g u +
u +

(3.337)

The correponding dimensionless relation between s and a


s is
"

a
s = s +

u +

where

#,"


1 u u
,
u +
#

(3.338)

I (t)

= s3/2
,

4 2

(3.339)

with
I (t) =

1
xt
1
q
d
=
dx q
2
2t
sinh ( (2 +1)2 1/2t) 4
(xt)2 + 1
Z

q

(xt)2 + 11
sinh2 (x/2)

3/2

.(3.340)

The primes in (3.338) denote now the derivatives with respect to s.


Let us evaluate this for small s where t = kB T /sa is large so that (3.340)
becomes15

t3/2
x3/2
6 2
3/2 3
(3/2)
=
I (t)
=
t
dx
4
4
s3/2
sinh2 (x/2)
Z

Ta
Tc

3/2

(3.341)

Here we have used the ideal-gas critical temperature Tc [(3/2)]2/3 4Ta [recall
(3.248)] to equate t = s1 (T /Tc )[(3/2)]2/3 4 . Hence we find that for s 0, that

is equal to

I (t)
s3/2
4 2

kB T
a

!3/2

3 Ta
2 Tc


3/2

3 T
2 Tc


3/2

(3.342)

The phase transition lies at the temperature where u = . For weak couplings,
this can be calculated analytically. The calculation is somewhat subtle since the
small-s region of the integral cannot simply be obtained by expanding
the integrand
in powers of s. Instead, the first correction starts with the power s. To see this,
15

I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series, and Products, Academic Press,
New York, 1980, Formula 3.527.

227

3.6 Bose Gas at Strong Interactions

we must proceed as in the derivation of the Robinson expansion of the Bose-Einstein


integral function [13]. Recalling (3.250), we study the integral
1
u
= 2

2 + s

(2

s)2

s2

2
2 2 2
( +s) s /T
e
1

(3.343)

for small s. The integral can be done immediately for s = 0 where we find, as in
(3.247),
u 0
u
T
=
=
s0

Tc


3/2

(3.344)

For small s we are left with the subtracted exression


1
u
= 2

2 + s

, (3.345)
2 2 2
d 2 q
2
(2 + s)2 s2 e ( +s) s /T 1 e /T 1

The first term can be expanded in powers of s, but the integral takes its leading behavior from linear momentum behavior of the sound a la Nambu-Goldstone, leaving
!
r
Z
u
T
T s
1
2 + s
2
2

,
d

2 0
4 + 2s2 2
4 2

so that

r
 3/2
u
s3/2 2
T
T s
= 2
+
+ ... .

4 3
Tc
4 2

Thus we obtain for small s the leading terms

r
 3/2
s3/2 2
T
u
T s
= 2
+
+ ....

4 3
Tc
4 2
The second frequency sum has the small-s-behavior

s3/2
2
3/2 I (t)
=s
=
2 + 2 s3/2 h (t),
2
2

4
4
4
where the correction term is [compare (3.243)]

Z
2 3/2
s
2
1
2 2 2
s h (t) 2
d 2 q
,
2
4
4 0
(2 + s)2 s2 e ( +s) s /T 1

(3.346)

(3.347)

(3.348)

(3.349)

(3.350)

so that
r
s3/2
T s

=
+ ...
2

4 2
4 2

(3.351)

228

3 Interacting Non-Relativistic Particles

We can now evaluate Eq. (3.337) at small s near the phase transition. Since

3/2
constant for s = 0, the last term in Eq. (3.337) is of the order of s /as and can be
neglected in comparison with the left-hand side, so that we obtain
s
T
1
a
s
Tc


3/2

r
T s
+ O(s3/2 ) + . . . .
4 2

(3.352)

The integral is once more treated with care (3.250), approximating it by


Z

T
2T

=
,
2 + 2s
4 2s

(3.353)

and we find
I (t)
s3/2 2
4

=s

3/2

r
2
T s

.
4 2 4 2

(3.354)

The last terms in (3.348) and (3.354) are dominant for small s. They have a
dramatic effect upon the phase diagram, with the consequence observed in earlier
publications that for small coupling constant, the critical temperature increases
above the free Bose gas value [14]. Let us see how this happens. We insert (3.348)
and (3.354) into (3.338) and find for small s:
s
T
=1
a
s
Tc


3/2

r
T s
+ O(s3/2 ) + . . . ,
+
4 2

(3.355)

the last term in (3.338) being exponentially small for small s.


Hence we obtain for small s at the phase transition the relation between s and
a
s
r

s
T
= as
+ O(a2s ).
2
8

(3.356)

Inserting this into (3.348), we obtain on the phase transition curve where, u = ,
the relation
T
1
Tc


3/2

r
T s

,
4 2

(3.357)

or
T
2 Tc2
1+
a
s + . . . .
Tc
3 32 2

(3.358)

Inserting Tc = [(3/2)]2/34 from (3.248), this becomes


1
T
=1+
a
s + . . . .
Tc
3(3/2)4/3

(3.359)

229

3.6 Bose Gas at Strong Interactions

This is the surprising initial increase of the critical temperature for small repulsion
between the bosons observed in [14].
Numerically, the prefactor c of the initially linear relation
as
T
= 1+ c + ...
Tc
a

(3.360)

loop
is c 3(3/2)4/3 1.03. This agrees resonably well with the values c5VPT
=
0.93 0.13 predicted from 5-loop variational perturbation theory in Ref. [15], and
with its extension to seven-loops [16] which gave c = 1.27 0.11. It happens to be
exactly equal to the value derived by Baym et al. from large-N calculations [17].
Let us now calculate the superfluid density of the condensate. Recalling the
remarks made about superfluid 4 He on p. 183, this must be distinguished from the
condensate density. The general formula in D dimensions is [18]

s = n = +
=

2 X
1
p p Ep /k T
B 1
DV p
e

X p eEp /kB T
2
.
DkB T V p (eEp /kB T 1)2

(3.361)

Using the surface of the unit sphere in D dimensions SD = 2 D/2 /(D/2), this can
be written as
n =

2 SD
D (2)D

dk k D1

yex
.
(ex 1)2

(3.362)

where y p /kB T and x Ep /kB T . Setting k = k this is equal to


2 SD D
k
D (2)D

n =
Since y = 2 /t and x =
that we obtain

d D1

yex
.
(ex 1)2

(3.363)

(2 + 1)2 1/t, we have d D1 = tD/2 dy y D/21 /2 so

1 SD D D/2
=
k t
D (2)D

dy y

D/2

ex
,
(ex 1)2

(3.364)

where dy = dx x/(ty + 1) = dx tx/ (xt)2 + 1, implying that


q

( (xt)2 +11)D/2
ex
1 SD D Z
q
k
t
dx
x
,
n =

D (2)D 0
(ex 1)2
(xt)2 +1

(3.365)

For large t, the integral can be approximated by


tD/2

dx

xD/2 ex
D/2 D
=
t
(D/2)(D/2),
(ex 1)2
2

(3.366)

230

3 Interacting Non-Relativistic Particles


1.0
0.8
0.6
h(t)

0.4
0.2
0

=
Figure 3.14 Temperature dependence of the normal particle density, where t kB T /
0
0
0

kB T /sa = Tc T /Tc s = Tc /s is the reduced temperature.

so that n has the large-t behavior


n =

D/2 (D/2) D D/2


k t ,
(2)D

(3.367)

which yields in D = 3 dimensions


(3/2)k3 T
n =

(4 2 )3/2


3/2

T
=
Tc0

!3/2

= 3/2 .

(3.368)

At finite t, the right-hand side is multiplied by the function


2t

h(t)
D(D/2)(D/2)tD/2

dx x

( (xt)2 +11)D/2
q

(xt)2 +1

ex
. (3.369)
(ex 1)2

The function h(t) is plotted in Fig. 3.14.


The result may be compared with an analogous expression for the uncondensed
particle density (3.272), which can be written in a form like (3.362) as
u =

k3
4 2

+t
3

dx

( (xt)2 +11)1/2
ex 1

(3.370)

(3.371)

Similarly we can write (3.276) as


R =

k3
4 2

2t

( (xt)2 +11)1/2

.
dx q
2
x
(xt) +1 (e 1)

There is no problem to increase the accuracy to any desired level, with exponentially fast convergence, as was demonstrated by the calulation of critical exponents
in all Euclidean 4 theories with N components in D dimensions [19]. The procedural rules were explained in the paper [20]. We merely have to calculate higher-order
diagrams using the harmonic Hamiltonian (3.253) as the free theory that determines
the Feynman diagrams, and
var

trial .
Hint
= Hint
H

(3.372)

3.7 Corrections Due to Omitted Interaction Hamiltonian

231

as the interaction Hamiltonian that determines the vertices. At any given order,
the results are optimized in the variational parameters 0 , , and . The theory
is renormalizable, so that all divergencies can be absorbed in a redefintion of the
parameters of the orginal action, order by order. This is the essential advantage of
the present theory over any previous strong-coupling scheme published so far in the
literature, in particular over those based on Hubbard-Stratonovic transformations
of the interaction, which are applicable only in some large-N limit as explained in
[20], and for which no higher-loop calculations are renormalizable.

3.7

Corrections Due to Omitted Interaction Hamiltonian

Let us now calculate the changes of this result caused by the omitted interaction

Hamiltonian Hint
in Eq. (3.16). It contributes to the energy a term16
1 Z tb
ih

2 E = lim
dtdt hT Hint
(t)Hint
(t )i,
tb ta tb ta 2
h2 ta

(3.373)

where Hint
(t) is the time-dependent operator

Hint
(t) eiH0 t/h Hint
(t)eiH0 t/h ,

(3.374)

and T is the time-ordering operator (2.231).


In order to calculate this, we split the creation and annihilation operators into
a hermitian and an antihermitian part:
1
ap (p + ip ),
2

1
ap (p ip ).
2

(3.375)

Using Eq. (3.28), we can express p and p in terms of the quasiparticle operators
p and p as follows:
p =

up + vp

(p + p
),
2

p =

up vp

(p p
)
2

(3.376)

Using the hyperbolic angles (3.38) for the Bogoliubov coefficients, this becomes
ep

),
p = (p + p
2

ep

p = (p p
)
2

(3.377)

Transforming these to arbitrary times via (3.374), we find


ep
ep

p(t) = (p eiEp t/h + p


eiEp t/h ), p (t) = (p eiEp t/h p
eiEp t/h ),(3.378)
2
2i
16

See Formula (1.304) in Chapter 1.

232

3 Interacting Non-Relativistic Particles

where E(q) are the quasiparticle energies (3.292). These operators have the timeordered correlation functions
iEp |tt |/
h
2p e

hp (t)p (t )i = e

iEp |tt |/
h
2p e

, hp (t)p (t )i = e

hp(t)p (t )i = i

eiEp |tt |/h


,
2

hp (t)p (t )i = i

eiEp |tt |/h


.
2

(3.379)

We now express the hyperbolic angles, remembering Eqs. (3.36) and (3.269), as
cos 2p = u2p + vp2 =

p +
, sin 2p = 2up vp = .
Ep
Ep

(3.380)

so that
e2p

v
u

u p + 2
p + 2
=t
.
=
Ep
p

(3.381)

gA0 X

p (t)[pp (t)p (t) + pp (t)p (t)],


V 2 p,p

(3.382)

p
=
=
Ep

p
,
p + 2

e2p

Now we transform the interaction Hint


of Eq. (3.16) to the time-dependent
operator

(t) =
Hint

and use Wicks theorem17 the expand the correlation function into two-point correlations functions:
hT (t)[ 2(t)+ 2(t) ](t )[ 2 (t )+ 2(t )]i = 6hT(t)(t )i3 +12hT (t)(t)ihT(t)(t )i2
+2hT (t)(t)ihT(t)(t )i2 + 4hT (t)(t )ihT (t)(t )i2
= 6I1,1,1 12I1,0,0 + 2I1,1,1 + 4I1,0,0 .
(3.383)
They are pictured in Fig. 3.15.

Figure 3.15 Diagrams picturing the Wick contractions in Eq. (3.383). A solid line represents a contraction hT(t)(t )i, a dashed line hT(t)(t )i, and a mixed line hT(t)(t )i
or hT(t)(t )i.

Inserting this into (3.373), and allowing for all intermediate three-quasiparticle
states, we find the extra energy
2 E =
17

g 2 A20
g 2 A20
M(3I1,1,1 6I1,0,0 + I1,1,1 + 2I1,0,0 ) =
MI,
4
4

See to be derived in Section 7.14.1.

(3.384)

233

3.7 Corrections Due to Omitted Interaction Hamiltonian

where Il,m,n denotes the momentum integrals


Il,m,n

d3 q1 d3 q2
(2h)6

p1
E q1

!l

p2
E q2

!m

p12
Eq12

!n

2M(Eq1

1
,
+ Eq2 + Eq12 )

(3.385)

and q12 = q1 + q2 . They are simplified by a vector version of the substitution (3.50)
to dimensionless wave vectors:
ph
k ,

(3.386)

so that they become


Il,m,n h
2 k4

d k1 d k2
(2)6

21
2
1 +2s

l/2 

22
2
2 +2s

m/2 

12
212 +2s

A(1 ) + A(2 ) + A(12 )

n/2

(3.387)

with the dimensionless quasiparticle energies


A() 2MEp /h2 k2 =

2 (2 + 2s).

(3.388)

The integrals have a quadratic divergence that vanishes in dimensional regularization. There is, further, a logarithmic divergence which in dimensional regularization
appears as a pole term proportional to 1/(D 3). This is omitted if we treat the
cubic interaction term by a minimal subtraction. The remaining finite contribution
was calculated in Ref. [24], yielding18
2 c3 = 16V
2 M 3 c3 ,
I = 3I1,1,1 6I1,0,0 + I1,1,1 + 2I1,0,0 = (4M )

(3.389)

where c3 was calculated in Appendix 3A [see Eq. (3A.12)]. Inserting this into (3.384),
replacing A20 by V 0 /g according to Eq. (3.254), and adding 2 E to the previous
2 c3 . Minimizing
energy W1 in Eq. (3.299), which changes it by 2 W1 = 4V gM 3 0
changes (3.304) to the following equation
W1 + 2 W1 with respect to

0
u
1 8g 2M 3 ( 12 + c3 )
=

.
0
g
u +
u +
"

(3.390)

Inserting now the relation = 3u , this becomes

3
1

0
1 8g 2 M 3 ( 21 + c3 ) = 0 + u = + u .
g
4u
2
2
"

(3.391)

Going now to reduced variables, the modified relation (3.306) becomes


"

1 u
1
s
1 8a2s ( 12 + c3 )
=
1
+
u ,
a
s
4
u
2
18

See Ref. [24], Eq. (A.21).

(3.392)

234

3 Interacting Non-Relativistic Particles

or
s
s
1
= 1 + u + 2as ( 12 + c3 ) (1 u ) ,
a
s
2
u
from which we find the relation
3/2 !
3/2 !

2s
2s
2 1
2
2s.
+
c

+
log
s)
1

+
a

(
6
s=a
s 1 +
3
s 2
24 2
12 2

(3.393)

(3.394)

It is interesting to observe that the extra contribution


2/3 vanishes on the transition
2
line where (u /)cr = 1, i.e., at s = scr = (12 / 2) 19.14.

Appendix 3A

Two-Loop Momentum Integrals

Here we calculate the integrals appearing in Eq. (3.384) following Ref. [24]. Let us first define the
4
integrals in (3.387) asR In,l,m = h2 k
RJn,l,m and study the integral Jn,l,m which are all proportional
2
3
3
to s . Abbreviating d /(2) by  we must evaluate
Jl,m,n =

1 2

n

l

m
1 / 1 2 + 2s 2 / 2 2 + 2s
r/ r2 + 2s

1 1 2 + 2s + 2 2 2 + 2s + 12 r2 + 2s

(3A.1)

where 12 = |1 + 2 |. In D = 3, these integrals have quartic and quadratic ultraviolet divergences
that cancel in the combination of integrals
J = 6J0,0,1 J1,1,1 3J1,1,1 2J1,0,0 .

(3A.2)

needed in Eq. (3.389). The expression for J can be written


J

1

"

p
2 21 + s
61
p

1
21 + s
#
p
p
31 2 12
1 22 + s 212 + s
p
p
p
p 2

. (3A.3)
1 + s 22 + s 212 + s
2 12 22 + s
1

p
p
p
2 1 21 + s + 2 22 + s + 12 212 + s

This integral still has linear and logarithmic ultraviolet divergences. By subtracting and adding
appropriate terms in the integrand of J, we can isolate the linear and logarithmic divergences into
separate terms:
J = Jlin + Jlog + Jnum .
(3A.4)
The term containing the linear ultraviolet divergence is
#Z
p
Z "
21 + 2s
1
1
.
2 p 2

Jlin = 2
2
1
1 + 2s
1
2 2

The term in (3A.4) containing the logarithmic ultraviolet divergence is


Z Z (
2
Jlog = s2
2 + 4s)( 2 + 2s)(2 + 2 + 2 + 2s)
(
2
1 2
1
1
2
12
)


1
1
1
+ 2
.

1 + 22 + 212 + 2s 2(22 + 2s) (21 + 2s)2

(3A.5)

(3A.6)

Appendix 3A

235

Two-Loop Momentum Integrals

The integral Jnum obtained by subtracting (3A.5) and (3A.6) from (3A.3) is convergent in D = 3
dimensions and can be evaluated numerically. It is convenient to symmetrize the integrand over
the six permutations of 1 , 2 , and 12 , in order to avoid cancellations between different regions
of momentum space. The resulting expression is
"
(
p
Z Z
X
2 21 + 2s
61
1
1
p
p
p

Jnum =

1
21 + 2s
1 2 6 (1 ,2 ,12 ) 1 21 + 2s + 2 2 2 + 2s + 12 212 + 2s
#
p

31 2 12
p 22 + 2s r2 + s
p
p
p
p 2

2
2
1 + 2s 2 + 2s 12 + 2s
2 12 21 + 2s
"
#
p
21 + 2s
1
2
+ 2 2 p 2

2
1
1 + 2s
8s2
+ +
+ 4s)(p2 + s)(2 2 + 2s)
)


4s2
1
1
.

+
21 + 2 2 + 212 + 4s 2(22 + 2s) (21 + 2s)2
+

(21

22

212

(3A.7)

Since s is the only scale in the integrand, dimensional analysis implies that the integral is proportional to s2 . Evaluating the coefficient of s2 numerically, Braaten and Nieto obtain
Jnum = 2.10 103 s2 .

(3A.8)

Because of the severe cancellations between the various terms in the integral, they were only able
to calculate it to 3 significant figures.
The ultraviolet divergent integrals Jlin in (3A.5) and Jlog in (3A.6) are evaluated using dimensional regularization. The integral over 2 in (3A.5) vanishes since there is no scale in the
integrand, and therefore Jlin = 0. The integral (3A.6) is evaluated in the limit D 3, and the
result is


4 3 3
1
Jlog =

1.13459
(2s)D1 .
(3A.9)
192 3
D3
Adding (3A.8) and (3A.9), we obtain the complete result for J using dimensional regularization:


1
4 3 3
(2s)D1 + c3 (2s)D1 = J1 + c3 (2s)D1 .
(3A.10)
J=
192 3
D3

with c3 = 0.57(4 3 3/192 3 ) 0.001411. The first term J1 still has logarithmic divergence.
Since s is dimensionless, it may be rewritten



1
4 3 3
4 3 3 (2s)2 (D3) log(2s)D3
2
(2s)
e

+ log (2s) + . . . .
(3A.11)
J=
192 3 D 3
192 3
D3
The first term is removed by a pole counter term in the total energy. The remainder modifies c3
to
c3 (s) = c3 + log 2 + log s c3 + log s,

c3 0.695.

(3A.12)

Let us finally evaluate the logarithmically divergent term (3A.6). It can be written
Jlog = s2 (2K1 + K2 ) ,
where K1 and K2 are the following integrals:
Z Z
1
K1 =
2 + 2 + 2 + 4s)(2 + 2s)(2 + 2s) ,
(
 
1
2
12
1
2

Z 1Z 2
1
1
1
K2 =
2 + 2 + 2 + 4s 2(2 + 2s) (2 + 2s)2 .

1 2 1
2
12
2
1

(3A.13)

(3A.14)
(3A.15)

236

3 Interacting Non-Relativistic Particles

We first consider the integral K1 . Setting 12 = 21 + 22 + 21 2 and introducing Feynman
parameters, the integral becomes
Z 1 Z 1x Z Z
1
(3A.16)
K1 =
dx
dy
2 + (1 x)2 + z   + s]3 ,
[(1

y)
1
2
1 2
0
0
1
2
where z = 1 x y. Dimensional regularization allows us to shift and rescale the momentum
variables. We eliminate the scalar product in the denominator by performing the shift 1 1


2 z/[2(1 y)]. After rescaling 1 by (1 y)1/2 and 2 by [(1 x)(1 y) z 2 /4]/(1 y) 1/2 ,
and factor the integral into a Feynman parameter integral and an integral over the momenta:
Z 1 Z 1x
Z Z
1
dy[(1 x)(1 y) z 2 /4]D/2
dx
K1 =
(3A.17)
2 + 2 + 2s)3 .
(
0
0
1 2 1
2
The integral over the momenta can be done analytically:
Z Z
1
(3 D)
(2s)D3 ,
2 + 2 + 2s)3 =
D
(
2(4)
1 2 1
2

(3A.18)

with a pole at D = 3 from the gamma function. To obtain K1 for D 3, we expand the Feynman
parameter integral in powers of D 3:
(Z
3/2

Z 1x
1
z2
(3 D)
D3
dx
dy
(1

x)(1

y)

(2s)
K1 =
2(4)D
4
0
0

)


Z
Z 1x
3/2
z2
D3 1
z2
+ . . . , (3A.19)
ln (1 x)(1 y)

dx
dy (1 x)(1 y)
2
4
4
0
0
where z = 1 x y. The first integral in (3A.19) is equal to 4/3. The second integral has to be
computed numerically and has the value 9.43698. Extracting the pole in D 3 from the gamma
function in (3A.19) and keeping all terms that survive in the limit D 3, we obtain


1
1
D3
K1 =
(2s)
+
1.12646
+

ln(4)
,
(3A.20)
96 2
D3
where is Eulers constant. We next consider the integral K2 in (3A.15). By introducing a
Feynman parameter, it can be written

Z 1
Z Z 
1
1

. (3A.21)
K2 =
dx (1x)
2
2
3
[(1 x)21 + x22 + 2s]3
1 2 [1 + x2 + x1 2 + 2s]
0
By shifting and rescaling the momentum variables, we can reduce the integral over the momenta
to (3A.18). In the first term of (3A.21), we shift 1 1 x2 /2 and then rescale 2 by
(x(4 x)/4)1/2 . In the second term, we rescale 1 by (1 x)1/2 and 2 by x1/2 . After
integrating over 1 and 2 , we obtain
Z 1
h
i
(3 D)
D3
K2 =
(2s)
dx (1 x)xD/2 (1 x/4)D/2 (1 x)D/2 .
(3A.22)
D
2(4)
0
To obtain K2 in the limit D 3, we need to expand the integrand in (3A.22) in powers of D 3:
(Z
!
1
1
8(1

x)
(3 D)
(2s)D3
dx p
p
K2 =
2(4)D
x3 (4 x)3
x3 (1 x)
0
!)
Z
D3 1
1
8(1 x)
x(4 x)

p
dx p
ln
ln[x(1 x)]
. (3A.23)
2
4
x3 (4 x)3
x3 (1 x)
0

237

Notes and References

The integrals can be evaluated analytically. Extracting the pole in D 3 from the gamma function
and keeping all terms that survive in the limit D 3, we obtain



1 3
3
1
4
D3

K2 =
(2s)
+
ln

1
+

ln(4)
.
(3A.24)
64 3
D3 3 3 2 4

Notes and References


The theory of the weak-coupling Bose gas is due to
N.N. Bogoliubov, On the Theory of Superfluidity , Izv. Akad. Nauk SSSR (Ser. Fiz.) 11, 77
(1947); Sov.Phys.-JETP 7, 41(1958).
See also
J.G. Valatin, Nuovo Cimento 7, 843 (1958);
K. Huang, C.N. Yang, J.M. Luttinger, Phys. Rev. 105, 776 (1957);
K. Huang, C.N. Yang, Phys. Rev. 105, 767 (1957). The theory of rotons in superfluid 4 He is
described in the textbook
R.P. Feynman, Statistical Mechanics , W.A. Benjamin, Reading Mass, 1972.
K.A. Bruckner, K. Sawada, Phys. Rev. 106, 1117(1957)
The individual citations refer to:
3/2

[1] The term proportional to as was calculated by Lee and Yang in


T.D. Lee, C.N. Yang, Phys. Rev. 105, 1119 (1957).
[2] The first observation was made at JILA with 87 Ru:
M.H. Anderson, J.R. Ensher, M.R. Matthews, C.E. Wieman, and E.A. Cornell, Science
269, 198 (1995).
This was followed by a condensate of 7 Li at Rice University:
C.C. Bradley, C.A. Sackett, J.J. Tollet, and R.G. Hulet, Phys. Rev. Lett. 75, 1687 (1995),
and in 30 Na at MIT:
K.B. Davis, M.-O. Mewes, M.R. Andrews, and N.J. van Druten. D.S. Durfee, D.M. Kurn,
W. Ketterle, Phys. Rev. Lett. 75, 3969 (1995).
[3] J.R. Abo-Shaeer, C. Raman, J.M. Vogels, and W. Ketterle, Science 292, 476 (2001).
[4] H. Kleinert, Path Integrals in Quantum Mechanics Statistics, Polymer Physics, and Financial Markets, World Scientific, Singapore 2008 (www.physik.fu-berlin.de/~kleinert/b5).
[5] H. Kleinert and V. Schulte-Frohlinde, Critical Phenomena in 4 -Theory, World Scientific,
Singapore, 2001 (ibid.http/b8).
[6] T.D. Lee, K. Huang, and C.N. Yang, Phys. Rev. 106, 1135 (1957).
[7] U.C. T
auber and D.R. Nelson, Phys. Rep. 289, 157 (1997).
[8] O. Penrose and L. Onsager, Phys. Rev. 104, 576 (1956).
[9] P. Nozi`eres and D. Pines, The Theory of Quantum Liquids, Addison-Wesley, New York,
1990, Vol. II.
[10] J.G. Bednorz and K.G. Mueller, Z. Phys. B 64, 198 (1986).
[11] H. Frohlich, Phys. Rev. 79, 845 (1950);
J. Bardeen, L.N. Cooper, J.R. Schrieffer, Phys. Rev. 106 , 126 (1957).
[12] H. Kleinert, Mod. Phys. Lett. B 17, 1011 (2003) (cond-mat/0210162).
[13] J.E. Robinson, Phys. Rev. 83, 678 (1951). See also the textbook [4], p. 172.

238

3 Interacting Non-Relativistic Particles

[14] H. Kleinert, S. Schmidt, and A. Pelster, Phys. Rev. Lett. 93, 160402 (2004).
A recent discussion and comparison of various data is found in
K. Morawetz, M. M
annel, and M. Schreiber, Phys. Rev. B 76, 075116 (2007).
[15] H. Kleinert, Mod. Phys. Lett. B 17, 1011 (2003) (klnrt.de/320)
[16] B. Kastening, Phys.Rev. A 69, 043613 (2004).
[17] G. Baym, J.-P. Blaizot, and J. Zinn-Justin, Europhys. Lett. 49, 150 (2000).
[18] P. B. Weichmann, Phys. Rev. 38, 8739 (1988).
[19] H. Kleinert and Schulte-Frohlinde, Critical Properties of P hi4 -Theories, World Scientific,
Singapore 2001 (klnrt.de/b8).
[20] H. Kleinert, EJTP 8, 57 (2011) (www.ejtp.com/articles/ejtpv8i25p57.pdf).
[21] V. I. Yukalov and H. Kleinert, Phys. Rev. A 73, 063612 (2006)
[22] V. I. Yukalov and E. P. Yukalova, Phys. Rev. A 76, 013602 (2007) .
[23] M.H. Kalos, D. Levesque, and L. Verlet, Phys. Rev. 9, 2178 (1974).
[24] E. Braaten and A. Nieto, Euro. Phys. J. B 11, 143 (1999).

Hope not without despair, despair not without hope


Seneca (4 BC65)

4
Relativistic Free Fields
Having learned how many-particle Schrodinger theory can be reformulated as a
quantum field theory, we shall now try to find possible field theories for the description of relativistic many-particle systems. This will first be done classically. The
fields will be quantized in Chapter 7.

4.1

Relativistic Particles

The nonrelativistic energy-momentum relation used in the Schrodinger theory


(p) =

p2
2M

(4.1)

is valid only for massive particles which move very much slower than the velocity of
light [1]
c = 2.99792458 1010 cm/sec.
(4.2)
If particles are accelerated to large velocities close to c this condition is no longer
fulfilled. Instead of (4.1), the energy follows the relativistic law
(p) =

c2 p2 + c4 M 2 .

(4.3)

In particular, the light particles themselves, the photons, follow this law with the
mass M = 0. It will be convenient to replace the energy by the new variable
p0 (p)/c.

(4.4)

Then the relation (4.3) can be expressed as


2

p0 p2 = M 2 c2 .

(4.5)

Thus, energy and momentum of a particle of mass are always such that the fourvector
p = (p0 , pi )
(4.6)
is situated on the upper, p0 > 0, hyperboloid in a four-dimensional energymomentum space. This is called the mass shell of the particle of mass M. If
239

240

4 Relativistic Free Particles and Fields

the particles are massless, the hyperboloid degenerates into a cone, the so-called
light cone.
Since a free particle remains free when seen from any rotated or uniformly moving
coordinate frame, energy and momentum must transform in such a way as to remain
always on the same mass shell. For a simple rotation of the frame this is obvious.
The energy remains the same while the momentum p changes only its direction. For
example, p may appear rotated around the z-axis by a transformation
pi = R3 ()i j pj ,

(4.7)

where R3 () is the matrix

cos sin 0

cos 0
R3 () sin
.
0
0 1

(4.8)

The angle is defined in such a way that in the new rotated frame the momentum
of the same particle appears rotated in the xy-plane in the anticlockwise direction,
i.e., the coordinate axis have been rotated clockwise with respect to the original
frame. We speak of a passive rotation of the system. The effect is the same as if the
observer had remained in the same frame but rotated the experimental apparatus,
and thereby the particle orbit, in the anticlockwise sense. The transformations
defined in this way are called active transformations. There are two equivalent
ways of formulating all invariance principles either in the active or in the passive
way. In this text we shall use the passive way. The reader should be aware that
different texts use different conventions and the formulas calculated in one cannot
always be compared directly with those in the other but may require changes, which
fortunately are quite straightforward.
For a general rotation by an angle with an axis pointing in the direction of the
unit vector1 , the transformation has the matrix form

'

pi = R' ()i j pj .

(4.9)

We shall also write, with slightly shorter notation for the rotation matrix,

'

pi = Ri j ( ) pj .

(4.10)

Explicitly, this transformation reads

'

p = cos p + sin ( p) + p|| .

(4.11)

'

Here p|| , p are the projections parallel and orthogonal to the rotation axis :

''

p|| (p ) ,

p p p|| ,

respectively. The set of all rotations form a group called the rotation group.
1

Hats on vectors in this section denote unit vectors, not Schrodinger operators.

(4.12)

241

4.1 Relativistic Particles

Consider now another set of transformations in which the second frame moves
with velocity v into the z-direction of the first. Then in the new frame the z
momentum of the particle will appear increased. The particle appears boosted in
the z-direction with respect to the original observer. The momenta in x- and ydirections are unaffected. Since the total four-momentum still satisfies the mass
2
2
shell condition (4.3), (4.5), the combination p0 p3 has to remain invariant. This
implies that there must be a hyperbolic transformation mixing p0 and p3 which may
be parametrized by a hyperbolic angle , called rapidity: :
p 0 = cosh p0 + sinh p3 ,
p 3 = sinh p0 + cosh p3 .

(4.13)

This is called a pure Lorentz transformation. We may write this transformation in


a 4 4 matrix form as

p =

cosh
0
0
sinh

0
1
0
0

0 sinh
0
0
1
0
0 cosh

p B3 () p .

(4.14)

The subscript 3 of B3 indicates that the particle is boosted into the z-direction. A
similar matrix can be written down for x and y-directions. In an arbitrary direction
, the matrix elements are

i sinh

cosh

.
B () B( ) =
i sinh ij + i j (cosh 1)

(4.15)

The spatial velocity of a particle is given by


v (p)/p.

(4.16)

In Schrodinger theory this is the velocity of a wave packet. In terms of v |v|, one
defines the Einstein parameter
1
= cosh .
q
1 v 2 /c2

(4.17)

In terms of these quantities, we can rewrite (4.15) as

B( ) =

v i /c

v i /c ij + ( 1)v i v j /v 2

(4.18)

where ( 1)v i v j /v 2 is often conveniently rewritten as 2 v i v j /c2 ( 1).


By combining rotations and boosts one obtains a 6-parameter manifold of matrices
(4.19)
( , ) = B( )R( ).

'

 '

242

4 Relativistic Free Particles and Fields

These are called proper Lorentz transformations. For all these, the combination
2

p0 p2 = p0 p2 = M 2 c2

(4.20)

is an invariant. These matrices form a group, the proper Lorentz group. We can
easily see that the Lorentz group allows reaching every momentum p on the mass
shell by applying an appropriate group element to some fixed reference momentum
pR . For example, if the particle has a mass M we may choose for pR the so-called
rest momentum
(4.21)
pR = (Mc, 0, 0, 0),
-direction
and apply the boost in the p

( ) = B( ),

(4.22)

with the rapidity given by


cosh =

p0
,
Mc

sinh =

|p|
.
Mc

(4.23)

With this, we can rewrite the general boost matrix (4.15) in the pure momentum
form

B( ) =

p0 /M

|p|pi /M 2 c2

pi |p|/M 2 c2 ij + pi pj (p0 /M 1)

(4.24)

Instead of (4.22), we may choose the following more general boost in the p
direction (p) = B(( )R( ), where R( ) is an arbitrary rotation, since these leave
the rest momentum pR invariant. In fact, the rotations form the largest subgroup
of the group of all proper Lorentz transformations which leaves the rest momentum
pR invariant. It is referred to as the little group or Wigner group of a massive particle. It has an important physical significance since it serves to specify the intrinsic
rotational degrees of freedom of the particle. If the particle is at rest it carries no
orbital angular momentum. If it happens that its quantum mechanical state remains
completely invariant under the little group R, the particle must also have zero intrinsic angular momentum or zero spin. Besides this trivial representation, the little
group being a rotation group can have representations of any angular momentum
s = 21 , 1, 32 , . . . . In these cases, the state at rest has 2s + 1 components which are
linearly recombined with each other upon rotations.
The situation is quite different in the case of massless particles. They move with
the speed of light and p cannot be brought by a Lorentz transformation from the
light cone to a rest frame. There is, however, another standard reference momentum
from which one can generate all other momenta on the light cone. It is given by

 '

'

pR = (1, 0, 0, 1)|p|,

(4.25)

243

4.1 Relativistic Particles

with an arbitrary size of the spatial momentum |p|. It remains invariant under
a different little group, which is again a three-parameter subgroup of the Lorentz
group. The little groups will be discussed in detail in Section 4.16.3.
It is useful to write the invariant expression (4.20) as a square of a four-vector

p formed with the metric

1
1

g =

(4.26)

namely
p2 = g p p .

(4.27)

In general, we define a scalar product between any two vectors as


pp g p p = p0 p0 pp .

(4.28)

Following Einsteins summation convention, repeated greek indices are summed from
zero to 3 [recall (2.100)]. A space with this scalar product is called Minkowski space.
It is useful to introduce the covariant components of any vector v as
v g v .

(4.29)

Then the scalar product can also be written as


pp = p p .

(4.30)

With this notation the mass shell properties (4.20) for a particle before and after a
Lorentz transformation reads simply
2

p = p 2 = M 2 c2 .

(4.31)

Note that, apart from the minus signs in the metric (4.26), the mass shell condition
2
2
2
2
p2 = p0 p1 p2 p3 = M 2 c2 that is left invariant by Lorentz transformations,
2
2
2
2
is completely analogous to the spherical condition p4 + p1 + p2 + p3 = M 2 c2
left invariant by the rotation group in a four-dimensional Euclidean space. Both
groups are parametrized by six parameters which are associated with linear transformations in the six planes, the Lorentz group in the planes 12, 23, 31; 10, 20, 30, the
rotation group in the planes 12, 23, 31; 14, 24, 34. In the case of the four-dimensional
Euclidean space these are all rotations which form the group of special orthogonal
matrices called SO(4). The letter S indicates the property special . A group of
matrices is called special if all matrices have a unit determinant. By analogy, the
proper Lorentz group is denoted by SO(1,3). The numbers 1,2 indicate that in the
Minkowski metric (4.26), one diagonal element is equal to +1 and three are equal
to 1.

244

4 Relativistic Free Particles and Fields

The fact that all group elements are special follows from a direct calculation
of the determinant of the matrices in (4.9) [with (4.11)] and (4.15).
How do we have to describe the quantum mechanics of a free relativistic particle
in Minkowski space? Energy and momentum p0 and p must be related to the time
and space derivatives of particle waves in the usual way
p0 = ih

ih 0 ,
ct
x

pi = ih

.
xi

(4.32)

In relativistic notation these read


p = ih

.
x

(4.33)

Together with the coordinates, they satisfy the canonical commutation rules
[
p , p ] = 0,
[x , x ] = 0,
[
p , x ] = ihg .

(4.34)

We expect a free particle with momentum p be described by (x) of the plane-wave


type analogous to nonrelativistic wave (2.211):
p (x) = N ei(p

0 x0 pi xi )/
h

= N eipx/h .

(4.35)

where N is some normalization factor. Since the zeroth component p0 is fixed by


the mass shell condition (4.5), only the spatial momentum needs to be specified
just as in the nonrelativistic plane wave solutions (2.211). However, in contrast
to those, there
now two solutions for each momentuma
one with energy
2are
2 p,
0
0
2
4
2
2
p = (p) = c p + c M , and one with p = (p) = c p + c4 M 2 . Thus we
have two plane-wave solutions
i((p)x
(+)
p (x) = N e

p() (x) = N e

0 pi xi )/
h

i((p)x0 +pi xi )/
h

= N eipx/h N fp (x),

= N fp
(x),

(4.36)
(4.37)

of positive and negative energy, respectively. For later convenience, we have


introduced the notation fp (x) for the positive-energy solution (+)
p (x), so that
()

p (x) = fp (x).

The energies of fp (x) and fp


(x) are seen by applying p0 to these wave functions:
i0 fp (x) = (p)fp (x),

i0 fp
(x) = (p)fp
(x).

(4.38)

The latter equation hold, of course, also for fp (x). The solutions fp (x) and fp (x)
will also be called positive- and negative-frequency wave functions, respectively. If
not stated differently, the zeroth component p0 will from now on always denote the
positive energy (p).

245

4.1 Relativistic Particles

At this point we do not yet normalize the wave functions since we must first
find a proper scalar product for calculating physical observables from these wave
functions. This scalar product will be given in (4.174).
We have stated previously that permissible energy-momentum states of a free
particle can be realized by considering one and the same particle in different coordinate frames connected by a transformation . Suppose that we change the
coordinates of the same spacetime point as follows:
x x = x.

(4.39)

Under this transformation the scalar product of any two vectors remains invariant:
x y = xy.

(4.40)

For rotations, this is obvious since xy = x0 y 0 xy. For Lorentz transformations the
invariance is a direct consequence of the fact that the boost matrix (4.14) satisfies
the relation

(4.41)
g B3 () B3 () = g ,
or in matrix notation
B3T ()gB3() = g.

(4.42)

The same relation holds obviously for the arbitrary boost matrix (4.15), and after
a combination with all rotations for the general Lorentz transformation (4.19):
T g = g,

(4.43)

g = g ,

(4.44)

or
The invariance (4.40) of the scalar product is then verified in matrix notation as
follows:
x y xT gy = (x)T g(y) = xT T gy = xT gy = xy.
(4.45)
This holds also for scalar products between momentum and coordinate vectors
p x = px.

(4.46)

If the metric were Euclidean, this would be the definition of orthogonal matrices.
In fact, in the notation (4.45) of scalar products in which the metric is suppressed,
we may write
(p)(x) = p1 x = px,
(4.47)
so that there is no difference between the manipulation of orthogonal and Lorentz
matrices.
When changing spacetime coordinates from x to x = x, the plane wave function
of a particle behaves like
p (x) = N eip

1 x /
h

= N ei(p)x /h = p (x ).

(4.48)

246

4 Relativistic Free Particles and Fields

This shows that in the new coordinates the same particle appears with a different
momentum and energy:
p = p.
(4.49)
Consider now an observable field (x) describing a particle which does not possess
any intrinsic orientational degree of freedom, i.e., no spin. The field can be an
arbitrary superposition of different plane wave functions. After a coordinate transformation it will still have the same value at the same spacetime point. Thus (x )
as seen in the new frame must be equal to (x) in the old frame, i.e.,
(x ) = (x).

(4.50)

A field with this property is called a scalar field or, for historical reasons, a KleinGordon field [2]

4.2

Differential Operators for Lorentz Transformations

Equation (4.50) contains the same point of the physical system on both sides, labeled
by different coordinates x and x . For the derivation of consequences of symmetries
(see Chapter 8) it is preferable to formulate the property (4.50) in the form of a
transformation law at the same spacetime coordinates x (corresponding to different
points of the physical system). Thus we shall express the transformation property
(4.50) of a scalar field in the following form

(x)
(x) = (1x).

(4.51)

For clarity, we have marked the transformation producing (x) by a subscript . It


is useful to realize that the inverse Lorentz transformation of the coordinates inside
the field argument can also be achieved with the help of a differential operator. To
find it we observe that the finite transformation matrices (4.9) and (4.18) can all
be written in a convenient exponential form. We begin with the rotations. The
four-dimensional form of the rotation (4.8) of the coordinate frame by an angle
clockwise around the z-axis brings a point with coordinates x to new coordinates
x = R3 () x with the 4 4 -matrix

R3 () =

1
0
0
0 cos sin
0 sin
cos
0
0
0

0
0
0
1

(4.52)

This can be written in the exponential form

R3 () =

exp

0
0
0
0

0
0 0
0 1 0
1
0 0
0
0 0

eiL3 .

(4.53)

247

4.2 Differential Operators for Lorentz Transformations

The matrix

L3 = i

0
0 0
0
0 1
0 1 0
0
0 0

0
0
0
0

(4.54)

is called the generator of this rotation within the Lorentz group. There are similar
generators for rotations around x- and y-directions

0
0
0
0

0
0
0
0

L1 = i

L2 = i

0
0
0
0
0
0
0 1

0
0
0
1

0
0
1
0

0
0
0 1

.
0
0
0
0

(4.55)

(4.56)

For all three cases we may write the generators as


Li i

0 0
0 ijk

(4.57)

where ijk is the completely antisymmetric Levi-Civita tensor with 123 = 1. The
pure rotation matrix (4.9) is given by the exponential
= ei'L ,

(4.58)

as can also be verified by expanding the exponential in a power series.


Let us now find the generators of the pure Lorentz transformations. First in the
z-direction where we see from (4.14) that the boost matrix is

B3 () =

exp

0
0
0
1

0
0
0
0

0
0
0
0

1
0
0
0

= eiM3

(4.59)

with the generator

M3 = i

0
0
0
1

0
0
0
0

0
0
0
0

1
0
0
0

(4.60)

248

4 Relativistic Free Particles and Fields

Similarly we have

M1 = i

M2 = i

0
1
0
0

1
0
0
0

0
0
0
0

0
0
0
0

(4.61)

0
0
1
0

0
0
0
0

1
0
0
0

0
0
0
0

(4.62)

The general Lorentz transformation matrix (4.15) is given by the exponential


= eiM ,

(4.63)

as can be verified by expanding the exponential in a power series.


The full Lorentz group is therefore generated by the six matrices Li , Mi , to be
collectively denoted by Ga (a = 1, . . . , 6). Every element of the group can be written
as
(4.64)
= ei('L+M) eia Ga .

'

If the exponential expanded in a power series, one reobtains for = 0 or = 0 the


general transformation matrices (4.9) or (4.15), respectively.
There exists a Lorentz-covariant way of specifying the generators of the Lorentz
group. We introduce the 4 4-matrices
(L ) = i(g g g g ),

(4.65)

labeled by the antisymmetric pair of indices , i.e.,


L = L .

(4.66)

There are 6 independent matrices which coincide with the generators of rotations
and boosts as follows:
1
ijk Ljk ,
2
= L0i .

Li =

(4.67)

Mi

(4.68)

With the generators (4.65), we can write every element (4.456) of the Lorentz group
as follows
1

= ei 2 L ,
(4.69)
where the antisymmetric angular matrix = collects both, rotation angles
and rapidities:
ij = ijk k ,
(4.70)
0i = i .

(4.71)

249

4.2 Differential Operators for Lorentz Transformations

Summarizing the notation we have set


= ei('L+M) = ei( 2 ijk L
1

jk + i L0i )

= ei( 2 ij L

ij +

0i L

0i )

= ei 2 L .

(4.72)

Note that if the metric were Euclidean

g=

1
1

(4.73)

the situation would be well familiar from basic matrix theorems: Then would
comprise all real orthogonal matrices which could be written as an exponential of
all real antisymmetric matrices. In our case only iLs are antisymmetric while iM
are symmetric, a consequence of the minus signs in the Minkowski metric (4.26).
The reason for writing the group elements in this exponential form in terms of six
generators is that by this the multiplication rules of infinitely many group elements
can be completely reduced to the knowledge of the commutation rules among the six
generators Li , Mi . This is a consequence of the Baker-Campbell-Hausdorff formula
written in the form (see Appendix 4A)
1

eA eB = eA+B+ 2 [A,B]+ 12 [AB,[A,B]]+....

(4.74)

According to this formula we have with the general notation Ga = (Li , Mi ) for the
six generators in Eqs. (4.57) and (4.60)(4.62),
1

1 2 = eia Ga eib Gb
1
= exp ia1 Ga ib2 Gb + [ia1 Ga , ib2 Gb ]
2

1
1
2
1
2
+ [i(c c )Gc , [ia Ga , ib Gb ]] + . . . .
12


(4.75)

The exponent involves only commutators among Ga s. For the Lorentz group these
can be calculated from the explicit 4 4 -matrices (4.54)(4.56) and (4.60)(4.62).
The result is
[Li , Lj ] = iijk Lk ,
[Li , Mj ] = iijk Mk ,
[Mi , Mj ] = iijk Lk .

(4.76)
(4.77)
(4.78)

This algebra of generators is called the Lie algebra of the group. The number of
linearly independent matrices Ga (here 6) is called the rank r of the Lie algebra. In
the general notation with the generators Ga , the algebra reads
[Ga , Gb ] = ifabc Gc .

(4.79)

250

4 Relativistic Free Particles and Fields

The commutator of two generators is a linear combination of generators. The coefficients fabc are called structure constantsThey are completely antisymmetric in
a, b, c, and satisfy the relation
fabd fdcf + fbcd fdaf + fcad fdbf = 0,

(4.80)

which guarantees that the generators obey the Jacobi identity


[[Ga , Gb ], Gc ] + [[Gb , Gc ], Ga ] + [[Gc , Ga ], Gb ] = 0,

(4.81)

which is the law of associativity for Lie Algebras. The relation (4.80) can easily be
verified for the structure constants (4.76)(4.78) using the identity for the -tensor
ijl lkm + jkl lim + kil ljm = 0.

(4.82)

The Jacobi identity implies that the r matrices with r r elements


(Fc )ab ifcab

(4.83)

satisfy the commutation rules (4.79). They form the so-called adjoint representation
of the Lie algebra. The matrix (4.57) for Li is precisely of this type.
In terms of the matrices Fa of the adjoint representation, the commutation rules
can also be written as
[Ga , Gb ] = (Fc )ab Gc .
(4.84)

Continuing the expansion in terms of commutators in the exponent of (4.75),


the commutators can be executed successively and one remains at the end with an
expression
12
1 2
12 = eia ( , )Ga ,
(4.85)

with the parameters of the product a12 being completely determined from a1 , a2 for
any given structure constants fabc .
In the tensor notation L for Li , Mi of Eqs. (4.67), (4.68), and with multiplication performed covariantly, so that products L L have the matrix elements
(L ) (L ) , the commutators read
[L , L ] = i(g L g L + g L g L ).

(4.86)

Due to the antisymmetry in and it is sufficient to specify only the


simpler commutators
[L , L ] = ig L ,

no sum over ,

(4.87)

thereby omitting vanishing components in (4.86) in which none of the indices is


equal to one of the indices .
After these preparations we are ready to derive a differential operator which
achieves the transformation of the spacetime argument in (4.51). First we consider
infinitesimal Lorentz matrices
1
1 i L .
(4.88)
2

4.2 Differential Operators for Lorentz Transformations

The transformation

x
x = x

251

(4.89)

can be written as an infinitesimal symmetry transformation


1
s x = x x = i L x.
2

(4.90)

Inserting the 4 4 matrix generators (4.65), this becomes more explicitly


s x = x .

(4.91)

With this, the argument on the right-hand side of the field transformation law (4.51)
reads x x , so that the infinitesimal transformation takes the form

(x)
(x) + s (x),

with s (x) = x (x).

(4.92)

Alternatively we can write


1
(x),
s (x) = i L
2

(4.93)

are the differential operators


where L
i(x x ).
L

(4.94)

In terms of the quantum mechanical canonical momentum operators (4.33), this


becomes
1 (x p x p ).
(4.95)
L
h

The object in parentheses is the tensor version of the four-dimensional operator of


angular momentum.
The operators (4.95) satisfy the same commutation relations (4.86), (4.87) as
the 4 4 -generators L of the Lorentz group. They form a representation of the
Lie algebra (4.86), (4.87) in terms of differential operators generating the Lorentz
transformations on the spacetime argument of any field.
one
In working out the commutation rules among the differential operators L

and x as well as p :
conveniently uses the commutation rules between L
, x ] = i(g x g x ) = (L ) x ,
[L
, p ] = i(g p g p ) = (L ) p .
[L

(4.96)
(4.97)

These commutation rules identify x and p as vector operators [recall the definition
in (2.112)]. In general, an operator t1 ,,n is said to be a tensor operator of rank n
if it is transformed by L like x or p in each tensor index:
, t1 ,...,n ] = i[(g 1 t,...,n g 1 t,...,n ) + . . . + (g n t1 ,..., g n t1 ,..., )]
[L
= (L )1 t2 ,...,n (L )2 t1 ,...,n (L )n t1 2 ,..., . (4.98)

252

4 Relativistic Free Particles and Fields

The simplest examples for such tensor operators are t1 ,...,n = x1 xn or


t1 ,...,n = p1 pn .
The commutators (4.86) between the generators imply that these, themselves,
are tensor operators of rank 2.
can also be expressed in
It is worth observing that the differential operators L
terms of the 4 4 matrix generators (4.65) as
= i (L ) x p = i xT L p = i
L
pT L x.
h

(4.99)

follow the same algebraic construction rules as the opIn this form, the operators L
erators a
Mi a in Section 2.5. There we showed that sandwich constructions between
creation and annihilation operators a
L a
carry the commutation rules between the

matrices L into a larger Hilbert space without changing the commutation rules.
Since i
p and x commute in the same way as a
and a
, the same argument applies
to the sandwich construction (4.99), where the matrix generators L of Eq. (4.65)
between i
p and x produces an infinite-dimensional representation in terms of
differential operators acting on the Hilbert space of square-integrable functions.
The commutation relations between the group generators (4.95) and vector and
tensor operators can be used to calculate the effect of finite group transformations
upon these operators. They can also be used to express the transformation property
(4.50) in another way. A finite Lorentz transformation of a scalar field is obtained by
exponentiating the generators just as in the 4 4 -representation (4.69) and (4.72).
Thus we define the differential operator representation of finite group elements (4.69)
as
1

(4.100)
D()
ei 2 L .
commute in the same way among each other as the 4 4
Since the generators L

-matrix generators (4.65), the operators D()


obey the same group multiplication
rules as 4 4 -matrices . This follows directly from the expansion (4.75) of the
product in terms of commutators.
Let us apply such a finite transformation to the vector x and form
1

D()x
D ().

(4.101)

We do this separately for rotations and Lorentz transformations, first for rotations.
An arbitrary three-vector (x1 , x2 , x3 ) is rotated around the 3-axis by the operator
3 ()) = eiL 3 with L
3 = i(x1 2 x2 1 ) by the operation
D(R
3 ())xi D
1 (R3 ()) = eiL 3 xi eiL 3 .
D(R

(4.102)

The right-hand side is evaluated with the help of Lies expansion formula
2
iA
B]
+ i [A,
[A,
B]]
+ ... .
eiA B
e = 1 i[A,
2!

(4.103)

253

4.2 Differential Operators for Lorentz Transformations

3 commutes with x3 , this component is unchanged by the operation (4.111):


Since L
3 ())x3 D
1 (R3 ()) = eiL 3 x3 eiL 3 = x3 .
D(R

(4.104)

For x1 and x2 , the Lie expansion of (4.101) contains the commutators


i[L3 , x1 ] = x2 ,

i[L3 , x2 ] = x1 .

(4.105)

Thus, the first-order expansion term transforms the two-dimensional vector (x1 , x2 )
into (x2 , x1 ). The second-order terms are obtained by commuting the operator
3 with these components, yielding (x1 , x2 ). To third-order, this is again transiL
formed into (x2 , x1 ), and so on. Obviously, all even orders reproduce the initial
two-dimensional vector (x1 , x2 ) with an alternating sign, while all odd powers are
proportional to (x2 , x1 ):
3
iL

1
1
= 1 2 + 4 + . . . (x1 , x2 )
2!
4!


1 3 1 5
+ + + . . . (x2 , x1 ).
3!
5!


3
iL

(x , x )e

(4.106)

The even and odd powers can be summed up, and we obtain

eiL3 (x1 , x2 )eiL3 = cos (x1 , x2 ) + sin (x2 , x1 ).

(4.107)

Together with (4.104), the right-hand side corresponds precisely to the inverse of
the rotation (4.52). Thus
i

j
jx

i

j
jx

3 ())xi D
1 (R3 ()) = eiL 3 xi eiL 3 = eiL3
D(R


= R31 ()i j xj .

(4.108)

By performing successively rotations around the three axes we can generate in this
way any rotation.
1 (R' ()) = ei'L xi ei'L = ei'L
' ())xi D
D(R


= R'1 ()i j xj ,

(4.109)

this being the finite rotation form of the commutation relation for the vector operator
xi :
i , xk ] = xj (Li )jk ,
[L
(4.110)
which holds for any vector operator vi instead of xi [recall again the definition in
(2.112)].
3 commutes
The time component x0 is obviously unchanged by a rotation since L
0
with x .
A similar calculation may be done for pure Lorentz transformations. Here we

3 = L
03 =
first consider a boost in the 3-direction B3 () = ei M3 generated by M
i(x0 3 + x3 0 ) [recall (4.72), (4.68), and (8.85)]. Note the positive relative sign of
03 caused by the fact that i = i in contrast to
the two terms in the generator L
0
0 = . Thus we form
3 ())xi D
1 (B3 ()) = ei M 3 xi ei M 3 .
D(B

(4.111)

254

4 Relativistic Free Particles and Fields

The Lie expansion of the right-hand side involves the commutators


i[M3 , x0 ] = x3 , i[M3 , x3 ] = x0 , i[M3 , x1 ] = 0, i[M3 , x2 ] = 0. (4.112)
Thus the two-vector (x1 , x2 ) is unchanged, while the two-vector (x0 , x3 ) is transformed into (x3 , x0 ). In the second expansion term, the latter becomes (x0 , x3 ),
and so on, yielding
3
i M

3
i M

(x , x )e

1
= 1 + 2 +
2!

1
+ 3 +
3!


1 4
+ . . . (x0 , x3 )
4!

1 5
+ . . . (x3 , x0 ),
5!


(4.113)

which can be summed up

ei M3 (x0 , x3 )ei M3 = cosh (x0 , x3 ) sinh (x3 , x0 ).

(4.114)

Together with the invariance of (x1 , x2 ), this corresponds precisely to the inverse of
the boost transformation (4.14). Thus:
3 ())x D
1 (B3 ()) = ei M 3 x ei M 3 = eiM3
D(B


= B31 () x .

(4.115)

By performing successively rotations and boosts in all directions we find for the
entire Lorentz group
1
1
1

D()x
D () = ei 2 L x ei 2 L = (ei 2 L ) x = (1 ) x , (4.116)

where are the parameters (4.70) and (4.71). In the last term on the right-hand
side we have expressed the 4 4 -matrix as an exponential of its generators as
well, to emphasize the one-to-one correspondence between the generators L and
.
their differential-operator representation L
At first it may seem surprising that the group transformations appearing as a lefthand factor of the two sides of these equations are inverse to each other. However,
we may easily convince ourselves this is necessary to guarantee the correct group
multiplication law. Indeed, if we perform two transformations after each other they
appear in opposite order on the right- and left-hand sides:
2 1 )x D
1 (2 1 ) = D(
2 )D(
1 )x D
1 (1 )D
1 (2 )
D(

1

1
1
= [(2 1 )1 ] x . (4.117)
= (1
1 ) D(2 )x D (2 ) = (1 ) (2 ) x
If the right-hand side of (4.116) would contain instead of 1 , the order of the
factors in 2 1 on the right-hand side of (4.117) would be opposite to the order in
2 1 ) on the left-hand side.
D(
A straightforward extension of the operation (4.116) yields the transformation
law for a tensor t1 ,...,n = x1 xn :

1 () = ei 21 L t1 ,...,n ei 21 L
D()
t1 ,...,n D

= (1 )1 1 (1 )n n t1 ,...,n
1

= (ei 2 L )1 1 (ei 2 L )n n t1 ,...,n .

(4.118)

255

4.3 Space Inversion and Time Reversal

This follows directly by inserting in the product x1 xn , an auxiliary unit factor

1 () = ei 21 L ei 21 L between neighboring factors xi and per1 = D()


D
forming the operation (4.118) on each of them. The last term in (4.118) can also be
written as
h

ei 2 (L

111

i
+ ... + 1L 11) 1 ...n

1 ...n

t1 ...n .

Since the commutation relations (4.98) determine the result completely, the transformation formula (4.118) is true for any tensor operator t1 ,...,n not only those
composed from a product of vectors xi .
For a field (x) which can be expanded into a power series, the transformation
law (4.118) generalizes immediately to

1 () = ei 12 L (x)ei 12 L = (1x) = (ei 21 L x).


D()(x)
D

(4.119)

The finite transformation law (4.51) of a scalar field can therefore be expressed with
1

the help of the differential operator D()


= ei 2 L as

1 () ei 2 L (x)ei 2 L . (4.120)
(x)
(x) = (1 x) = D()(x)
D
The last factor on the right-hand side can, of course, be omitted if there are no
x-dependent functions behind it.
If a particle has spin degrees of freedom its field transforms differently from
(4.120). Then the wave function has several components to account for the spin
orientations. The transformation law must be such that the spin orientation in space
remains the same at the same space point. This implies that the field components
which specify the orientation with respect to the different coordinate axes will have to
be transformed by certain matrices. How this is done for relativistic fields has first
been understood for electromagnetic and gravitational fields which exhibit vector
and tensor transformation properties, respectively. These will be recalled below,
before generalizing them to the case of arbitrary spin.

4.3

Space Inversion and Time Reversal

In addition to the continuous Lorentz transformations, there are also two important
discrete transformations which leave scalar products p x invariant. First there is
the space inversion, also called space reflection, or parity transformation

P =

1
1

(4.121)

which reverses the direction of the spatial vectors, x x. Note that a space
inversion differs from a mirror reflection by a rotation. The space inversion maps
the generators Li , Mi of the Lorentz group into parity transformed generators
P

Li
LPi P Li P 1 = Li ,

Mi
MiP P Mi P 1 = Mi .

(4.122)

256

4 Relativistic Free Particles and Fields

This behavior is obvious in the tensor form of the generators (Li , Mi ) =


( 21 ijk Lik , L0i ): each spacelike index gives rise to a factor 1. The transformation
preserves the commutation rules (4.77)(4.78):
[LPi , LPj ] = iijk LPk ,

(4.123)

[LPi , MjP ] = iijk MkP ,

(4.124)

[MiP , MjP ]

iijk LPk .

(4.125)

In general, a mapping of the generators into linear combinations of generators which


have the same commutation rules is called an automorphism of the Lie algebra.
Second there is the time inversion or time reversal transformation

T =

1
1
1

(4.126)

which changes the sign of x0 . When applied to the generators of the defining representation (Li , Mi ), the time reversal transformation produces the same automorphism of the Lie algebra as the parity transformation (4.122). This, however, is
a special feature of the reality of the Lorentz transformations . This makes the
4 4-matrices of the generators (4.54)(4.56) and (4.60)(4.62) accidentally purely
imaginary.
Physically, a process is invariant under time reversal if we are unable to judge
whether a movie of the process runs forward or backward. Running it backward
amounts to changing momentum and angular momentum Li . Since momentum is
generated by boost transformations, time reversal must change the direction of Mi .
For hermitian matrices it is only necessary to change the eigenvalues, such that
we can also require Li going to Li . In fact, there is a natural automorphism of
the Lie algebra (4.77)(4.78) in which we simply take the complex conjugate of the
commutation rules, bringing them to
[Li , Lj ] = iijk (Lk ),
[Li , Mj ] = iijk (Mk ),
[Mi , Mj ] = iijk (Lk ).

(4.127)
(4.128)
(4.129)

As we shall see in detail below when discussing the time reversal properties of the various fields, this automorphism has precisely the desired observational consequences
which we would like to associate with a time reversal transformation. Explicitly,
time reversal transforms the generators as follows:
T

Li
LTi T Li T 1 = Li ,

Mi
MiT T Mi T 1 = Mi .

(4.130)

If the operations P and T are incorporated into the special Lorentz group SO(1,3)
one speaks of the full Lorentz group.
Note that the determinant of (4.121) and (4.126) is negative so that the full
Lorentz group no longer deserves the letter S in its name. It is called O(1, 3).

257

4.4 Relativistic Free Scalar Fields

4.4

Relativistic Free Scalar Fields

The question now arises as to how the nonrelativistic free-field action


A=

h
2 2
dtd3 x (x, t) iht +
x (x, t)
2M
#

"

(4.131)

introduced in (2.201) has to be modified in order to permit a quantum mechanical


description of arbitrary relativistic n-particle states. According to the definition
in (2.160), this is a local action. A field theory based on a local action will be
called local field theory. All field theories which explain successfully the properties
of elementary particle have so far turned out to be local. The locality property seems
to be extremely fundamental. All fundamental forces which historically have been
discovered as bilocal forces (they used to be called action at a distance forces) have
eventually been found to be the result of a local action involving an extra field that
mediates the interaction. In the case of electromagnetism, the Coulomb interaction
which initially is described by a bilocal term in the action (see Section 2.8) can be
reduced to a local interaction involving an extra potential field, which turns out to
be a component of a vector potential describing the particles of light, the photons
(see Chapter 12). The same thing holds for gravitational forces and the particles
called gravitons. In nuclear physics, the locality principle has led Yukawa to the
discovery of the fundamental particle called -meson [see Section 24.3, in particular
Eqs. (2.157) and (2.138) for V2 (x x ) 1/|x x |].
In order to accommodate the kinematic features discussed in the last section
we shall require the action to be invariant under Lorentz transformations (4.39),
(4.50). Hence, space and time derivatives have to appear on equal footing, i.e., both
must appear linearly or quadratically and with no higher power than that if we
want to maintain the usual principle of classical mechanics in which all differential
equations are of second order in time. Depending on the possible internal spin
degrees of freedom there are different ways of making the action relativistic. These
will now be discussed separately.
Consider first a field associated with a relativistic point particle which carries no
spin degree of freedom, thus avoiding a nontrivial behavior under space rotations.
Such a field is called a scalar field and will be denoted by (x). As in the nonrelativistic case, the action of this field must contain the square of the spatial derivatives i
to guarantee rotational invariance. Since there must be Lorentz symmetry between
spatial and time derivatives, we are led to a classical local action
A=

dx0 L =
=

dx0 d3 x (x, t) c1h


2 (02 x2 ) c2 (x, t)
h

d4 x (x) c1 h
2 c2 (x),

(4.132)

where c1h
2 , c2 are two arbitrary real constants.
It is easy to see that this action is indeed Lorentz invariant: Under the transformation (4.39), the four-volume element does not change
dx0 d3 x d4 x d4 x = d4 x.

(4.133)

258

4 Relativistic Free Particles and Fields

This follows directly from Eq. (4.43) which implies that det = 1 If we therefore
take the action in the new frame
A=

(4.134)

(4.135)

d4 x (x ) c1 h
2 c2 (x ),

we can use (4.50) and (4.132) to rewrite


A=

d4 x (x) c1h
2 c2 (x).

But since
= ,

(4.136)

with
g g ,

(4.137)

2 = 2 ,

(4.138)

we see that

and the transformed action coincides with the original action (4.132).
As in (2.160), we call the integrand of the action a Lagrangian density:
h

L(x) = (x) c1h


2 c2 (x).

(4.139)

Then the invariance of the action under Lorentz transformations is a direct consequence of the Lagrangian density being a scalar field satisfying the transformation
law (4.50):
L (x ) = L(x),
(4.140)
which follows directly from the invariance (4.138) and (x ) = (x).
The free-field equation of motion is derived by varying the action (4.132) with
respect to the fields (x), (x) independently. The independence of these variables
is expressed by the functional differentiation rules
(x)
= (4) (x x ),
(x )
(x)
= 0,
(x )

(x)
= (4) (x x ),
(x )
(x)
= 0.
(x )

(4.141)

Applying these rules to (4.132) we obtain directly


A
=
d4 x (4) (x x)(c1 h
2 2 + c2 )(x )
(x)
= (c1 h
2 2 + c2 )(x) = 0.
Z

(4.142)

259

4.4 Relativistic Free Scalar Fields

Similarly
Z
A
=
d4 x (x )(c1 h
2 2 + c2 ) (4) (x x)
(x)

= (x)(c1 h
2 2 +c2 ) = 0,

(4.143)

where the arrow on top of the last derivative indicates that it acts on the field to
the left. The second equation is just the complex conjugate of the previous one.
The field equations (4.142) and (4.143) can be derived directly from the Lagrangian density (4.139) by forming ordinary partial derivatives of L with respect
to all fields and their derivatives. Indeed, a functional derivative of a local action
can be expanded in terms of derivatives of the Lagrangian density according to the
general rule
A
L(x)
L(x)
L(x)
=

+
+ ... ,
(x)
(x)
[ (x)]
[ (x)]

(4.144)

and a similar expansion holds for the derivative with respect to (x). These expansions follow directly from the defining relations (4.141). Applying them to the
Lagrangian density (4.139), the field equation for (x) is particularly simple:
L(x)
A
=
= (c1h
2 2 + c2 )(x) = 0.

(x)
(x)

(4.145)

For (x), on the other hand, all derivatives written out in (4.144) have to be
evaluated to obtain
L(x)
L(x)
L(x)
A
=

+
= (c1 h
2 2 + c2 ) (x) = 0.
(x)
(x)
[ (x)]
[ (x)]
(4.146)
The equation
L(x)
L(x)
L(x)

+
+ ... = 0
(x)
(x)
(x)

(4.147)

is the Euler-Lagrange equation of a general local field theory.R This expression is


invariant under partial integrations within the local action A = d4 x L(x). Take for
example a Lagrangian density which is equivalent to (4.139) by a partial integration
in the action (4.135):
L = c1 h
2 (x)(x) c2 (x)(x).

(4.148)

Inserted into (4.147), they produce once more the same field equations.
The field equations (4.145) and (4.146) are solved by the quantum mechanical
plane waves (4.36) and (4.37) of positive and negative energies, respectively:
fp (x) = N eipx/h ,

fp (x) = N eipx/h .

(4.149)

260

4 Relativistic Free Particles and Fields

These form a complete set of plane-wave solutions.


The field equations (4.145) and (4.146) require the four-momenta to satisfy the
condition
c1 p p c2 = 0.
(4.150)
This has precisely the form of the mass shell relation (4.31) if we choose
c2
= M 2 c2 .
c1

(4.151)

A positive sign of c1 is necessary for the field fluctuations to be stable. The size can
be brought to unity by a multiplicative renormalization of the field. This makes the
field normalization different from the nonrelativistic one in the action (4.131).
After this, the mass shell condition fixes the free-field action to the standard
form
Z
i
h
(4.152)
A = d4 x (x) h2 M 2 c2 (x).

The nonrelativistic limit of the action (4.152) is obtained by removing from the
positive frequency part of the field (x) a fast trivial oscillating factor corresponding
to the rest energy Mc2 , replacing
(x) eiM c

t/
h

1
(x, t).
2M

(4.153)

For
wave fp (x) of Eq. (4.149), the field (x, t) becomes p (x, t) =
a plane
0
2
2M N ei(p cM c )t/h eipx/h . In the limit of large c, the first exponential becomes
2
eip t/2M , such that the field p (x, t) coincides with the nonrelativistic plane wave
(2.211) which extremizes the nonrelativistic action (4.131). The negative-frequency
plane wave fp (x) in (4.149), on the other hand, does not contribute in this limit

0
2
since it is equal to 2M N ei(p c+M c )t/h eipx/h . This contains a temporal prefactor
2
e2iM c t/h that oscillates infinitely rapidly for c , and is therefore equivalent to
zero by the Riemann-Lebesgue Lemma [3].
The appearance of the constants h
and c in all future formulas can be avoided
if we work from now on with new fundamental units l0 , m0 , t0 , E0 different from the
ordinary cgs units. They are chosen to give h
and c the value 1. Expressed in terms
of the conventional length, time, mass, and energy, these new natural units are given
by
h

,
Mc
= M,

l0 =
m0

,
Mc2
E0 = Mc2 .
t0 =

(4.154)
(4.155)

If, for example, the particle is a proton with mass mp , these units are
l0 = 2.103138 1011 cm
= Compton wavelength of proton,

(4.156)

t0 = l0 /c = 7.0153141 1022 sec

(4.157)

261

4.4 Relativistic Free Scalar Fields

= time it takes light to cross the Compton wavelength,


m0 = mp = 1.6726141 1024 g,
E0 = 938.2592 MeV.

(4.158)
(4.159)

For any other mass, they can easily be rescaled.


With these natural units we can drop c and h
in all formulas and write the action
simply as
Z
A=

d4 x (x)( 2 M 2 )(x).

(4.160)

The Lagrange density of the complex scalar field may be taken either as
L(x) = (x)( 2 M 2 )(x),

(4.161)

or
L(x) = (x) (x) M 2 (x)(x),

(4.162)

with an obvious modification for real fields (x). The surface term by which the
associated actions differ from each other after a partial integration does not change
the Euler-Lagrange equation (4.147).
Actually, since we are dealing with relativistic particles there is no fundamental
reason to assume (x) to be a complex field. In the nonrelativistic theory this was
necessary in order to construct a term linear in the time derivative
Z

dt it .

(4.163)

For a real field (x) this would have been a pure surface term and thus not influenced
the dynamics of the system. For second-order time derivatives as in (4.160) this is
no longer necessary.
Thus we shall also study the real scalar field with an action
1
A=
2

d4 x (x)( 2 M 2 )(x).

(4.164)

In this case it is customary to use a prefactor 1/2 to normalize the field. Here the
Lagrange density may be taken either as
1
L(x) = (x)( 2 M 2 )(x),
2

(4.165)

or as

1
L(x) = {[(x)]2 M 2 2 (x)}.
2
For either field we obtain the Klein-Gordon equation
( 2 M 2 )(x) = 0.

(4.166)

(4.167)

262

4 Relativistic Free Particles and Fields

For a complex field, there exists an important local conservation law, which
generalizes Eq. (1.108) of Schrodinger theory to relativistic fields. We define the
four-vector of probability current density

j (x) = i ,

(4.168)

which described the probability flow of the Klein-Gordon particle. It is easy to verify
that due to (4.167), this satisfies current conservation law
j (x) = 0.

(4.169)

This conservation law will permit us in Chapter 18 to couple electromagnetism to


the field and identify j (x) as the electromagnetic current (if we choose natural units
in which the electric charge e is equal to unity).
The deeper reason for the existence of a conserved current will be understood
in Subsection 8.11.1, where we shall see that it is intimately connected with an
invariance of the action (4.160) of a free complex scalar field under arbitrary changes
of the phase of the field
(x) ei (x).
(4.170)
It is this invariance which gives rise to a conserved current density [see (8.270), also
(18.67)].
The zeroth component of j (x) the the probability current density
(x) = j 0 (x).

(4.171)

describes the charge density. The spatial integral over (x) is the total probability
equal to the total charge in natural units:
Q(t) =

d3 x j 0 (x).

(4.172)

Because of the local conservation law (4.169), the charge does not depend on time.
This is seen by rewriting

Q(t)
=

d x 0 j (x) =

d x j (x)

d x i j (x) =

d3 x i j i (x)

(4.173)

and applying to the right-hand side Gausss theorem as in (1.109), assuming the
currents to vanish at spatial infinity.
By removing from the positive-frequency solutions of the Klein-Gordon field (x)
the fast oscillation as in (4.153), we can take the nonrelativistic limit and find that
the nonrelativistic limit of the spatial part of the current density (4.168) satisfies
the local conservation law (1.108) of Schrodinger theory.
Since the current conservation law (4.169) is the direct relativistic generalization
of the nonrelativistic probability conservation law (1.108), it is suggestive to define
the matrix elements of the charge Q(t) as the scalar product between relativistic

263

4.4 Relativistic Free Scalar Fields

wave functions such as the plane waves (4.149). For states of momenta p and p we
define the scalar product
(fp , fp )

d3 x fp (x, t)i 0 fp (x, t).

(4.174)

It is formed as a spatial integral at any fixed time. It is unnecessary to record this


time in the notation, since the result is really time-independent due to charge conservation. Analogous scalar products exist between positive- and negative-frequency
solutions fp (x) and fp (x), and between two negative-frequency solutions fp (x) of
different momenta. Both sets of wave functions fp (x) and fp (x) are needed to span
the space of all solutions of the Klein-Gordon equation. Within the scalar product
(4.174), we choose to normalize the plane wave functions so that they satisfy the
orthogonality relations
(fp , fp ) = p ,p ,

(fp , fp ) = p ,p ,

(fp , fp ) = 0.

(4.175)

The spatial integrals ensure the spatial momenta to be equal or opposite to each
other. Then the energies p0 are equal to each other, so that the time derivative in

the scalar products produces either zero (between fp (t) and fp


(t), or a nonzero
between equal wave functions. In a finite volume V , these values are the norms of
the wave functions fp (x, t) and fp (x, t). They coincide with the matrix elements
of the charge (4.172), if the appropriate plane wave is inserted for the field (x) in
(4.168).
The charge of the plane waves with negative frequency fp (x),
(fp , fp )

d3 x fp (x, t)i 0 fp (x, t),

(4.176)

is negative, so that the scalar product is not positive definite. Historically, this
was an obstacle for the Klein-Gordon theory to being a direct generalization of
Schrodinger theory to relativistic particles; rightfully so, as we shall see in Chapter 7.
In a finite total spatial volume V , the properly normalized wave functions (4.149)
are explicitly
fp (x, t) =

1
eipx ,
2V p0

fp (x, t) =

1
eipx .
2V p0

(4.177)

where p0 is the particle energy p0 = p = p2 + M 2 . The norms are 1.


In an infinite volume, a convenient normalization is
fp (x, t) = eipx ,

fp (x, t) = eipx ,

(4.178)

and the orthogonality relations become


(3)
(fp , fp ) = 2p0 - (p p),
(3)
(f , f ) = 2p0 - (p p),
p

(fp , fp ) = 0,

(4.179)

264

4 Relativistic Free Particles and Fields

where - (p p) = (2)3 (3) (p p) as defined in Eq. (1.195). The convenience in


having a factor 2p0 accompany the -function
is that this combination has pleasant
transformation properties under the Lorentz group. It yields unity when integrated
over the Lorentz-invariant volume element in momentum space
(3)

d3 p
.
(2)3 2p0

(4.180)

The Lorentz invariance is obvious by rewriting this as


Z

4.5

d4 p
(p0 )(p2 M 2 ).
(2)3

(4.181)

Other Symmetries of Scalar Action

The actions (4.160) and (4.164) of a real or complex scalar field are invariant under
more than just the Lorentz group.

4.5.1

Translations

First, the actions are invariant under space as well as time translations of the coordinate system
x = x + a .
(4.182)
Recall that under Lorentz transformations, a scalar field at the same spacetime point
remains unchanged by the change of coordinates

(x)
(x) = (1x).

(4.183)

The same is true for translations:


a

(x)
a (x) = (x a).

(4.184)

Inserting this into the Lagrangian density (4.161), we see that it transforms like a
scalar field
L (x) = L(x a).
(4.185)
Together with the the trivial translational invariance of the volume integral, the
action is indeed invariant.
The combinations of Lorentz transformations and translations
x = x + a

(4.186)

form a group called the inhomogeneous Lorentz group or Poincare group. Under it,
the scalar field transforms as
(x)
(x) = (1 (x a)).

(4.187)

265

4.5 Other Symmetries of Scalar Action

Thus, a free scalar field theory is not only Lorentz-invariant but also Poincareinvariant. This holds also for real and complex scalar fields (x).
Translations can be generated by a differential operator in just the same way
as Lorentz transformations in Eq. (4.120). Obviously we can write the translation
(4.184) as

(x)
a (x) = (x a) = D(a)(x)
eia p /h (x),

(4.188)

where p = ih is the differential operator of momentum (4.33). This is proved by


applying Lies expansion formula (4.103) to the coordinates x :
p
/
h

eia

p
/
h

x eia

= ea

x ea

= x a ,

(4.189)

Poincare transformations are then obtained from operations


1


D()(x)

(x)
a (x) = (1 (x a)) = D(a)
eia p /h ei 2 L (x),
(4.190)
with the parameters specified in (4.70), (4.71). This follows from the behavior
of the coordinate vector:
p
/
h

eia

p
/
h

ei 2 L x ei 2 L eia

x = (1 ) (x a) ,

(4.191)

thus extending (4.116) to the Poincare group. The last equation states in a global
way the vector properties of x under Poincare transformations.

4.5.2

Space Inversion

Second, the scalar actions (4.160) and (4.164) are invariant under the operation of
space inversion [see (4.121)], under which the coordinates go into
P

x
xP = x (x0 , xi ),

(4.192)

whereas the scalar field is transformed as follows


P

(x)
P (x) = (
x).

(4.193)

Note that this transformation behavior is not the only possible one. Since the
parity operation is not related continuously to the identity, there is no reason why
the field (x) should go into itself as it does for the continuous group of Lorentz
transformations. The parity operation forms, together with the unit element, a
group called the group of space reflections. . The group multiplication table reads
as follows:
1 P
1 1 P
(4.194)
P P 1
The only requirement which the transformation of the field must satisfy is to be
consistent with the group multiplication law in the table. This is assured if the

266

4 Relativistic Free Particles and Fields

successive application of two parity operations, which is the identity operation, leads
back to the original field (cyclicity of order 2). It is therefore possible to choose any
transformation law
P

(x)
P (x) = P (
x),

(4.195)

as long as P satisfies P2 = 1. This allows for two solutions, the above trivial one
in (4.195) with P = 1 and the alternating one with P = 1, i.e.,
P = 1.

(4.196)

Thus the scalar field could also pick up a negative sign upon space reflection.
If the interactions which generate a particle are invariant under the parity operation, the value of P is a characteristic property of the particle. It is called the
intrinsic parity of the particle. States with positive or negative intrinsic parities are
familiar in quantum mechanics where they appear as bound states with even or odd
orbital angular momentum, respectively. Only a particle with P = 1 is called a
proper scalar particle, while P = 1 is called a pseudoscalar particle.
The most important fundamental particles of odd parity are the -mesons which
are the source of the long-range part of nuclear interactions. The pseudoscalar nature
is most simply seen in the decay of 0 into two photons, which is the main reason
for the finite lifetime = (8.4 0.6) 1017 sec of this particle (branching ratio of
two-photon with respect to all decay channels is 98.798%). In the rest frame of the
pion, the two final photons emerge in opposite directions and polarized parallel to
each other. Under a space inversion, the two-photon state is transformed into itself,
but with reversed polarization directions. This corresponds to a negative parity.
The negative parity of a charged pion , whose lifetime is much longer [ =
(2.6030 0.0024) 108sec], can be deduced from the existence of the absorption of
a -meson at rest by a deuteron. The deuteron is a bound state of a neutron and
a proton in an s-wave with parallel spins, so that the total angular momentum of
a deuteron at rest is J = 1. For an s-wave, the parity of the orbital wave function
is positive. An additional pion at rest does not change J. The final state consists
of two neutrons flying apart in opposite directions. By the Pauli principle, their
wave function has to be antisymmetric. Thus it can only be in spin-singlet states
for even orbital angular momenta l, or in spin-triplet states for odd l. Since the
final total angular momentum must be J = 1, only the spin-triplet l = 1 state is
allowed. This, however, has a negative parity, which can only be caused by the
additional -meson being a pseudoscalar. Note that the intrinsic parities of proton
and neutron do not matter in this argument since these particles are present before
and after the absorption process.
If particles with definite intrinsic parity interact with each other and the interactions are invariant under space reflections, the intrinsic parity supplies characteristic
selection rules in scattering and decay processes. In quantum mechanics for example, decays of atomic states in the dipole approximation have to change the parity
of the state since the dipole operator itself has a negative parity.

267

4.5 Other Symmetries of Scalar Action

4.5.3

Time Reversal

As a second extension of the Lorentz invariance of the scalar actions (4.160), (4.164)
we can reverse the sign of the time axis via the time reversal transformation
T

x
xT =
x

(4.197)

which has the same multiplication table as the parity transformation:


1 T
1 1 T
T T 1

(4.198)

(x)
T (x) = T (xT ),

(4.199)

T = 1.

(4.200)

The field transforms like


T

where again
Note that this transformation law holds for both real and complex fields. The
field transformation law (4.199) should not be confused with the corresponding
transformation law of wave functions. In order to clarify this point consider nonrelativistic Schrodinger theory. There, given a solution (x, t) of the free-particle
Schrodinger equation
!
h
2 2
(4.201)
x (x, t) = 0,
it +
2M
the wave function (x, t) is also a solution with the same energy. The presence of
only a single time derivative necessitates the complex conjugation. A plane-wave
solution of momentum p
p (x, t) = eiEt/h+ipx/h ,
(4.202)
transforms like

p (x, t)
p T (x, t) = T p (x, t).

(4.203)

This satisfies the Schrodinger equation


h
2 2
x p T (x, t) = 0,
it +
2M
!

(4.204)

From (4.202) we see that the right-hand side is equal to


T p (x, t).

(4.205)

The particle momentum in the transformed wave function is reversed the particle
runs backwards in time. antilinear operator. Thus it has the property

268

4 Relativistic Free Particles and Fields

Since the transformation law (4.203) involves complex conjugation,


scalar prodR 3

ucts between two arbitrary wave functions, h2 (t)|1 (t)i = d x 2 (x, t)1 (x, t) go
over into their complex conjugates at the negative time:
T

h2 (t)|1 (t)i
T h2 (t)|1 (t)iT = h2 (t)|1 (t)i .h1 (t)|2 (t)i. (4.206)
This property guarantees the preservation of probabilities under this transformation.
In general, any transformation which carries all scalar products into their hermitian
conjugates is referred to as antiunitary. The antiunitarity implies that the time
reversal operation is necessarily antilinear . A transformation is antilinear if the
coefficients of a linear combination of wave functions go over into their complex
conjugates.
At the level of the Schrodinger differential operators, the antiunitarity produces
a sign reversal in the transformation properties of energy and momentum. The
defining representation
T 1 tT = t, T 1 xT = x,
(4.207)
implies that

T 1 it T = it ,

T 1 ix T = ix .

(4.208)

The antiunitary representation of this operation is


T

it T = it ,

ix T = ix .

(4.209)

It leaves the energy invariant, while reversing the direction of the momentum:
T

E
E,

p
p.

(4.210)

Thus particles keep their positive energy but run backwards in time. A unitary
representation would have the opposite effect and produce a state which cannot be
found in nature.
In contrast to the wave functions, the Schrodinger field operator, which was the
result of second quantization in Chapter 2, transforms under time reversal like
T
t)
t).
(x,
T (x, t) = T (x,

(4.211)

At the operator level, the effects of complex conjugation are brought about by the
antilinearity of the time reversal operator T in the second-quantized Hilbert space.
In fact, the transformation law (4.203) for the wave functions can be derived from
(4.211). In terms of T , the transformation (4.211) reads
T
t).
t)T = T (x,
t)
(x,
T 1 (x,

(4.212)

Expanding the field operator terms of creation and annihilation operators as in


Eq. (2.214),
Z
t) = d-3 p p (x, t) a
(x,
p ,
(4.213)

269

4.5 Other Symmetries of Scalar Action

and using the antilinearity of T , we obtain


T

t)T =
(x,

d-3 p p (x, t) T 1 a
p T =

d-3 p eiEt/hipx/h T 1 a
p T .

(4.214)

According to (4.212), the right-hand side has to be equal to


T

d-3 p eiEt/h+ipx/h a
p ,

(4.215)

which implies that the annihilation operators transform like


T 1 a
p T = T a
p .

(4.216)

A particle of momentum p goes over into a particle of momentum p, which is the


correct transformation law.
Thus, in spite of their contradictory appearance, the two transformation laws
(4.203) for wave functions and (4.211) for Schrodinger fields, with the antiunitary
operator implementation (4.212) after field quantization, are completely consistent
with each other. Thus we need no longer be astonished about the absence of a
complex conjugation on the right-hand side of the field transformation law (4.199).
Another feature of the antiunitarity which has to be kept in mind is that the
time-reversed field operator does not satisfy the Schrodinger equation of the original
field
!
h
2 2
iht +
(4.217)
x (x, t) = 0,
2M
but

h
2 2
iht +
x
2M

T (x, t) = 0.

(4.218)

This follows immediately by multiplying (4.217) with the operator T from the left
and passing T on to the right-hand side of the differential operator. The antilinearity causes the complex conjugation of the differential operator in (4.219). This is
necessary to produce the correct Schrodinger equation (4.204) for the time-reversed
wave functions.
At first sight the field equation appears to be in contradiction with the correspondence principle from which one might expect equations for operators to go directly
over into those for classical objects in the limit of small h
. Properly, however,
this limit must be taken on equations for measurable amplitudes, not the operators themselves, and these do follow the Schrodinger equation.2 Take for instance
the single-particle amplitude of an arbitrary state |i in the Heisenberg picture
t)|i, which satisfies
h0|(x,
h
2 2
t)|i = 0.
iht +
x h0|(x,
2M
!

(4.219)

This is a manifestation of Ehrenfests theorem for the semiclassical limit of field equations.

270

4 Relativistic Free Particles and Fields

For the time-reversed field operator (4.211), the equation is


h
2 2
x
iht +
2M

t)|i = 0.
h0|(x,

(4.220)

The time-reversed amplitude on the right-hand side is the amplitude for the time
reversed state T |i (since h0|T 1 = h0|):
t)|i = h0|T
h0|(x,

t)T |i = h0|(x,
t)T |i.
(x,

(4.221)

This obeys the Schrodinger equation


h
2 2
t)T |i,
iht +
x h0|(x,
2M
!

(4.222)

as it should.
The operator implementation of the time reversal transformation will be discussed in detail in the Chapter 7 for fields and particles of different spin and the
reader is referred to Subsecs. 7.1.6, 7.3.4, 7.4.3, and 7.5.4. An important observation should, however, be made right here: As a consequence of the antiunitarity, the
phase factor T appearing in the time reversal transformation of a complex field is
arbitrary. It cannot be fixed in the usual way by applying the transformations T
twice. The antilinearity will change T arising in the first transformation into T ,
so that the combined phase factor is T T = 1. This is fulfilled for any phase factor
T = ei , not just 1.
Since the phase factor T is arbitrary, it may be chosen arbitrarily, for instance
T 1.

4.5.4

(4.223)

Charge Conjugation

At the level of a relativistic scalar field (x) there is one more discrete symmetry.
We can change (x) into (x) without changing the action (4.160). This operation
is called charge conjugation and denoted by C. Since C 2 = 1, a complex field can
transform in two possible ways
C

(x)
C (x) = C (x),

(4.224)

where the phase factor can take the values


C = 1.

(4.225)

For a real field with the action (4.164), we may simply drop the complex conjugation
on the right-hand side of (4.224).
An explanation is necessary for the name of this operation. In Eq. (4.168) we
have seen that there exists a conserved current j (x) which can be used to couple

271

4.6 Electromagnetic Field

the complex scalar field to electromagnetism.


If this is done, j (x) becomes an elecR 3 0
tromagnetic current density, and Q(t) = d x j (x) the charge of the field system.
Now, under the transformation (4.224) this electromagnetic current density changes
its sign:
C

j (x)
jC (x) = j (x),

(4.226)

Thus the transformation reverses the charge of the field, and this is why the discrete
operation C is called charge conjugation
As we shall discuss later in Chapter 24, electromagnetic and strong interactions
are invariant under charge conjugation, implying that the phase factor C is a fixed
measurable property of the particle, called charge parity.
Take for instance the neutral -meson. As mentioned above, it decays in (8.4
0.6) 1017 sec mostly into two photons. Since these are also neutral particles, they
have a charge parity on their own. Whatever it is (we shall see in Subsec. 4.7.2
that photons have a negative charge parity C ), the two-photon state must have
a positive charge parity, and this must consequently be the charge parity of the
-meson.

4.6

Electromagnetic Field

Electromagnetic fields move with light velocity and have no mass term.3 The fields
have two polarization degrees of freedom (right and left polarized for circular polarization) and are described by the usual electromagnetic action. Historically, this
was the very first example of a relativistic classical field theory. Thus it could also
have served as a guideline for the previous construction of the action (4.164) of a
real scalar field (x).

4.6.1

Action and Field Equations

The action may be given in terms of a real auxiliary four-vector potential A (x)
from which the physical electric and magnetic fields can be derived as follows
E i = ( 0 Ai i A0 ) = t Ai i A0 ,
1
1
B i = ijk ( j Ak k Ai ) = ijk (j Ak k Aj ).
2
2

(4.227)
(4.228)

It is useful to introduce the so-called four-curl of the vector potential, the tensor
F = A A .
3

(4.229)

The best upper limit for the mass of the electromagnetic field M deduced under terrestrial
conditions, from the shape of the earths magnetic field, is M < 4 1048 g corresponding to
a Compton wavelength
= h/M c > 1010 cm, which is larger than the diameter of the sun.
Astrophysical considerations (wisps in the crab nebula) give > 1016 cm. If metagalactic
magnetic fields could be discovered, the Compton wavelength would be larger than 1024 1025 cm,
quite close to the ultimate limit set by the horizon of the universe = c age of the universe
1028 cm. See G.V. Chibisov, Sov. Phys. Usp. 19 , 624 (1976).

272

4 Relativistic Free Particles and Fields

Its six components are directly the electric and magnetic field strengths
F0i = F 0i = 0 Ai + i A0 = 0 Ai i A0 = E i ,
Fij = F ij = i Aj j Ai = i Aj + j Ai = ijk B k ,

(4.230)
(4.231)

or, in a more conventional notation


1
A0 (x),
E(x) = A(x)
c
B(x) = A(x).

(4.232)
(4.233)

For this reason the tensor F is also called the field tensor. Note that F is
related to the fields Bi and Ei in the same way as the generators L of the Lorentz
group were related to Li and Mi in Eq. (4.67).
The electromagnetic action reads
A=

d4 x L =

1
4

d4 x F F =

1
d4 x (E2 B2 ).
2

(4.234)

The four-curl F satisfies a so-called Bianchi identity for any smooth A :


F = 0,

(4.235)

where

1
(4.236)
F = F
2
is the dual field tensor, with being the four-dimensional Levi-Civita tensor
with 0123 = 1. Note that
F F = 4E B
(4.237)
is a pseudoscalar.
Equation (4.235) can be rewritten as

1
1

F = ( )A (x) = 0.
2
2

(4.238)

Multiplying this by another -tensor and using the tensor identity


= + +

(4.239)

we obtain the integrability condition of Schwarz lemma, according to which the


derivatives of an integrable function commute:
( )A (x) = 0.

(4.240)

The equations of motion which extremize the action are


A
L(x)
1
=
= F (x) = 0,

A (x)
[ A (x)]
2

(4.241)

273

4.6 Electromagnetic Field

or more explicitly
(g 2 )A (x) = 0.

(4.242)

Separating the equations (4.235) and (4.241) into space and time components
they are seen to coincide with the four Maxwells equations in empty space:
F = 0 :
F = 0 :

B = 0,
E = 0,

E + t B = 0,
B t E = 0.

(4.243)
(4.244)

The first equation in (4.243) states that there can be no magnetic monopoles. The
second equation is Faradays law of induction. The first equation in (4.244) is
Coulombs law in the absence of charges, the second is Amp`eres law in the absence
of currents (including, however, Maxwells displacement current caused by the time
derivative of the electric field).
In terms of the vector field A (x), the action reads explicitly
A=

1
d4 x [ A (x) A (x) A (x) A (x)]
2Z
1
d4 x A (x)(g 2 )A (x).
(4.245)
2

d4 x L(x) =
=

The latter form is very similar to the scalar action (4.164). The first piece is the same
as in (4.164) for each of the spatial components A1 , A2 , A3 . The time component
A0 , however, appears with an opposite sign. A field with this property is called a
ghost field. When trying to quantize such a field the associated particle states turn
out to have a negative norm. If the theory is to be physically consistent, such states
must never appear in any scattering process. The second term A A in the
action (4.245) is novel with respect to the scalar case. It exists here as an additional
possible Lorentz invariant since A is a vector field under Lorentz transformations.
It is instructive to insert the individual components A0 and A = (A1 , A2 , A3 )
into the action and find
1
A=
2

d4 x A0 (x)(2 )A0 (x) 2A0 (x)0 i Ai (x)


i

A(x)(02 2 )A(x)Ai (x)i i Aj (x) .

(4.246)

This shows that the field component A0 appears without a time derivative. This
will have the consequence that this component remains classical when going over to
quantum field theory in Chapter 7. The component A0 will be fully determined the
classical field equation.

4.6.2

Gauge Invariance

The field tensor (4.229) is invariant under local gauge transformations


A (x)
A (x) = A (x) + (x),

(4.247)

274

4 Relativistic Free Particles and Fields

where (x) is any smooth field which satisfies the integrability condition
( )(x) = 0.

(4.248)

Gauge invariance implies that one scalar field degree of freedom in A (x) does not
contribute to the physically observable electromagnetic fields E(x) and B(x). This
degree of freedom can be removed by fixing a gauge. One way to do that is to
require the vector potential to satisfy the Lorentz gauge condition of having no fourdivergence:
A (x) = 0.
(4.249)
For a vector field satisfying this condition, the field equations (4.242) decouple end
become simply four massless Klein-Gordon equations:
2 A (x) = 0.

(4.250)

If a vector potential A (x) does not satisfy this condition, one may always perform a gauge transformation (4.247) to a new field A (x) that has no four divergence. We merely have to choose a gauge function (x) solving differential equation
2 (x) = A (x),

(4.251)

and A (x) will satisfy A (x) = 0.


There are infinitely many solutions to equation (4.251). Given one solution (x)
which leads to the Lorentz gauge, one can add any solution of the homogenous KleinGordon equation without changing the four-divergence of A (x). The associated
gauge transformations
A (x)
A (x) + (x),

2 (x) = 0,

(4.252)

are called restricted gauge transformations or gauge transformation of the second


kind, or on-shell gauge transformations . If a vector potential A (x) in the Lorentz
gauge solves the field equations (4.242), the gauge transformations of the second
kind can be used to remove the spatial divergence A(x, t). Under (4.252), the
components A0 (x, t) and A(x, t) go over into
A0 (x) A0 (x, t) = A0 (x, t) + 0 (x, t),
A(x) A (x, t) = A(x, t) (x, t).

(4.253)

Thus, if we choose the gauge function


(x, t) =

d3 x

1
A(x , t),
4|x x |

(4.254)

then
2 (x, t) = A(x, t)

(4.255)

275

4.6 Electromagnetic Field

and the gauge-transformed field A (x, t) has no spatial divergence: it is completely


transverse:
A (x, t) = [A(x, t) (x, t)] = 0.
(4.256)
The condition

A (x, t) = 0

(4.257)

2 (x, t) = 0.

(4.258)

is known as the Coulomb- or radiation gauge.


The solution (4.254) to the differential equation (4.255) is undetermined up to
an arbitrary solution (x) of the homogeneous Poisson equation

Together with the property 2 (x, t) = 0 from (4.252), one also has
t2 (x, t) = 0.

(4.259)

This leaves only trivial linear functions (x, t) of x and t which do not describe
propagating waves.
Another possible gauge is obtained by removing the zeroth component of the
vector potential A (x) satisfying the field equations (4.242). We form again
A (x) = A (x) + (x),

(4.260)

but now with a gauge function


(x, t) =

dt A0 (x, t ).

(4.261)

Then A (x) will satisfy the field equations (4.242), while having the property
A0 (x) = 0.

(4.262)

This is called the axial gauge. The solutions of Eqs. (4.261) are determined up to
a trivial constant, so that no more gauge freedom is left comparable to the second
kind in (4.252).
For free fields, the Coulomb gauge and the axial gauge coincide. This is a
consequence of Coulombs law E = 0 in Eq. (4.244). By expressing E(x)
explicitly in terms of the spatial and time-like components of the vector potential
E(x) = 0 A(x) A0 (x),

(4.263)

2 A0 (x, t) = A(x,
t).

(4.264)

Coulombs law reads

This shows that if A(x) = 0, also A0 (x) = 0, and vice versa.


The differential equation (4.264) can be integrated to
A0 (x, t) =

1
4

d3 x



1
(x , t).

A
|x x|

(4.265)

276

4 Relativistic Free Particles and Fields

In an infinite volume with asymptotically vanishing fields there is no freedom of


adding to the left-hand side a nontrivial solution of the homogenous Poisson equation
2 A0 (x, t) = 0,

(4.266)

which in principle would be possible.


In the presence of charges, Coulombs law will have a source term and read [see
Eq. (12.52)]
E(x, t) = (x, t).
(4.267)
where (x, t) is the electric charge density. Now the divergence of (4.263) yields the
equation

2 A0 (x, t) = A(x,
t) (x, t),
(4.268)
which is solved by
A0 (x, t) =

1
4

d3 x



1
(x , t).

A
|x x|

(4.269)

In contrast to the previous (4.265), the vanishing of A(x, t) no longer implies


A0 (x, t) 0, but determines it to be the instantaneous Coulomb potential around
the charge distribution (x , t):
1
A (x, t) =
4
0

d3 x

|x

1
(x , t).
x|

(4.270)

Remarkably, there is no retardation. This is an apparent violation of the relativity


priciple. The contradiction will be resolved due to gauge invariance in Chapters 7
and 12.
Note that the fields A can still be modified by adding and one has the
possibility of choosing (x, t) either to satisfy the Coulomb gauge
A(x, t) 0,

(4.271)

or any other gauge, such as the axial gauge


A0 (x, t) 0.

(4.272)

Only for free fields the two gauges coincide.

4.6.3

Lorentz Transformation Properties of Electromagnetic


Fields

The Lorentz transformation properties of the electromagnetic fields were understood


a long time ago within classical electrodynamics. They are the origin of the famous
Lorentz force acting on moving charged particles. The experimentally observed
electric and magnetic forces can be derived by going from the laboratory frame

277

4.6 Electromagnetic Field

with fields E, B to the reference frame of the moving particle with fields E , B and
transforming
v
B ,
c


v
= B E ,
c


E|| = E|| ,

E = E +

B|| = B|| ,

(4.273)
(4.274)

with v being the velocity of the particle and 1/ 1 v 2 /c2 the Einstein parameter (4.17). These equations can also be written without separating transverse and
longitudinal components as
v
2 v v
B
E ,
c
+1 c c




2 v v
v
B .
B = B E
c
+1 c c


E = E +

(4.275)
(4.276)

The transformed fields exert the observed electric and magnetic forces eE + gB.
The subscripts || and indicate projections of the fields parallel and orthogonal to
v. From this transformation one may derive the transformation law of the vector
field A under Lorentz transformations. Let the frame in which the moving particle
is at rest be related to the laboratory frame by

x = B( )x,

(4.277)

where B( ) is a boost in v-direction with a rapidity


cosh = ,

v
sinh = ,
c

v
tanh = .
c

(4.278)

Then the transformation law (4.274) is equivalent to

A (x ) = B ( )A (x),

(4.279)

apart from an arbitrary gauge transformation. An analogous transformation law


holds for rotations, so that the transformations (4.274) and their rotated forms
correspond to the Lorentz transformations
A (x ) = A (x),

(4.280)

plus possible gauge transformations. In the notation (4.51), we write the transformation law as

A (x)
A (x) = A (1 x).
(4.281)
By analogy with Eq. (4.120) for the scalar field, and recalling (4.69), this transformation can be generated as follows:

i 1 J

A (x)
A (x) = D()
A (x),
A (x) e 2

(4.282)

278

4 Relativistic Free Particles and Fields

with the parameters specified in (4.70), (4.71), and



J L 1 + 1 L

(4.283)

being the generators of Lorentz transformations of both the vector index and spacetime argument The former are generated by the 4 4 matrices L of Eq. (4.65),
the latter by the differential operators (8.85). The combined generators (4.283) are
called the generators of the total four-dimensional angular momentum. The two
generators on the right-hand side of Eq. (4.283) on different spaces, transforming
once the vector index and once the spacetime coordinate x. One therefore writes
often shorter and somewhat sloppily
,
J L + L

(4.284)

with the tacit understanding that what is meant is the direct sum (4.283) of spin
and orbital generators.
Since the two terms in (4.283) act independently on space and spin indices, the
:
operators J satisfy the same commutation rules as L and L
[J , J ] = i(g J g J + g J g J ).

(4.285)

The transformation laws (4.280) and (4.282) differ from those of a scalar field
in Eqs. (4.50) and (4.120) in the way discussed above for particles with nonzero
intrinsic angular momentum. The field has several components. It points in the
same spatial direction before and after the change of coordinates. This is ensured
by its components changing in the same way as the coordinates of the point x .
Note that the four-divergence A (x) is a scalar field in the sense defined in
(4.50). Indeed
A (x ) = ( ) A (x) = A (x).

(4.286)

For this reason the second term in the action (4.245) is Lorentz-invariant, just as
the mass term in (4.164). The invariance of the first term is shown similarly
A (x ) 2 A (x ) = A (x) 2 A (x ) = A (x) 2 A (x) = A (x) 2 A (x).
(4.287)
Hence the action (4.245) does not change under Lorentz transformations, as it
should.

4.7

Other Symmetries of Electromagnetic Action

Just as the scalar action, also the electromagnetic action (4.234) is invariant under
more symmetry transformations than those of the Lorentz group.

279

4.7 Other Symmetries of Electromagnetic Action

4.7.1

Translations

Under translations (4.182),


x = x + a ,

(4.288)

A (x) = A (x a).

(4.289)

the vector field transforms like

The combination of these with Lorentz transformations forms the Poincare group
(4.186),
x = x + a ,
(4.290)
under which the field A (x) transforms like
A (x) = A (1 (x a)),

(4.291)

leaving the action (4.245) invariant.


As in the scalar equation (4.190), we can generate all Poincare transformations
of the vector potential A (x) with the help of differential operators as

ia
D()

A (x)
A (x) = D(a)
A (x) e

p
/
h

ei 2 J A (x),

(4.292)

with the parameters specified in (4.70), (4.71).

4.7.2

Space Inversion, Time Reversal, and Charge Conjugation

Under space inversion, the four-vector A (x) behaves as follows:


P
x).
A (x)
A P (x) = A (

(4.293)

Under time reversal one has


T

where

x),
A (x)
A T (x) = A (

(4.294)

A = (A0 , Ai ).

(4.295)

In principle, there is the possibility of a vector field V (x) transforming like


P

x),
V (x)
V P (x) = P V (

(4.296)

with P = 1. For P = 1 the field V (x) is called an axial vector field. The
electromagnetic gauge field A (x), however, is definitely a vector field. This follows
from the vector nature of the electric and the axial vector nature of the magnetic
field which are observed in the laboratory.
Similarly, the behavior of a physically observable vector field V with respect to
time reversal is given by
T

V (x)
V T (x) = T V (xT )

(4.297)

280

4 Relativistic Free Particles and Fields

with an arbitrary phase factor T . If the vector field is real and physically observable,
then the only alternatives are T = 1. For the electromagnetic vector potential
A (x), the phase factor T is as specified in (4.294) since under time reversal, all
spatial currents change their directions whereas the zeroth component stays the
same. This reverses the direction of the B-field but has no influence on the E-field
generated by flowing charges.
The complex conjugation on the right-hand side of (4.297) has the same origin
as in the transformation law (4.203) for the complex scalar field.
The operation of charge conjugation is performed by exchanging the sign of all
charges without changing their direction of flow. Then both E and B change their
directions. Hence
C
(4.298)
A (x)
A C (x) = A (x).
In general, the vector field could have transformed as
C

A (x)
A C (x) = C A (x)

(4.299)

with C = 1. The fact that the electromagnetic field has C = 1 means that it
is odd under charge conjugation.

4.8

Plane-Wave Solutions of Maxwells Equations

The plane-wave solutions of the field equations (4.242) are direct extensions of
Eqs. (4.177) and (4.178):
fk (x, t) =
or

1
(k, )eikx ,
2V k 0

fk (x, t) = (k, )eikx ,

fk =

1
(k, )eikx ,
2V k 0

fk = (k, )eikx ,

(4.300)
(4.301)

with the momentum on the mass shell with M = 0, the so-called light cone. The
momentum-dependent four-vectors (k, ) specify the polarization of the plane electromagnetic wave. The label counts the different polarization states. In the
Lorentz gauge, the vector potential must have a vanishing four divergence (4.249),
implying the condition
k (k, ) = 0.
(4.302)
Being solutions of the wave equations, there is a further restricted gauge freedom
(4.252). We may add to the solutions fk (x, t) or fk (x, t) the total gradient of a
function (x) which is itself a plane wave (x) = eikx with k 2 =0, thus solving also
the Klein-Gordon equation. The total gradient adds to the polarization vector a
term proportional of the four-momentum k :
(k, ) (k, ) = (k, ) + k (k, ).
By choosing
(k, ) =

1
k (k, ),
k2

(4.303)

(4.304)

281

4.8 Plane-Wave Solutions of Maxwells Equations

the spatial part of the polarization vector (k, ) acquires the property

k (k, ) = 0,

(4.305)

which is the Coulomb gauge (4.257) for the polarization vector.


We can also choose
1
(k, ) = 0 0 (k, ),
k

and (k, ) will satisfy


0 (k, ) = 0.

(4.306)
(4.307)

which is the axial gauge (4.262) for the polarization vector. In Section 4.6.2 we
showed that for free fields, the two gauges coincide. we see this here explicitly. The
Lorentz gauge (4.302) implies for the spatial and time-like components of 0 (k, ):

k 0 0 (k, ) = k (k, ),

(4.308)

so that the two conditions (4.305) and (4.307) are indeed the same.
Since the four-component vectors 0 (k, ) are restricted by two conditions, only
two of them can be independent. These will be labeled by = 1 and are constructed as follows.
Because of the axial property (4.307) we set

(k, ) (0, (k, )),

(4.309)

and impose on the spatial part the Coulomb gauge property (4.305). This equation
is solved by two polarization vectors orthogonal to the spatial momentum k. These
are defined uniquely by the following consideration. If k points in the z-direction,
then the two vectors (k, 1) coincide with the eigenvectors (1) of the 3 3 matrix L3 of the rotation group in Eq. (4.54), with eigenvalues 1. There are three
eigenvectors () ( = 1, 0, 1) which are determined by the equations

= 1, 0, 1.

L3 () = (),

(4.310)

The result is

1
1

(1) = i ,
2
0

0
1
(0) = 0 ,
2 1

(4.311)

The opposite signs of (1) are chosen to comply with the so-called Condon-Shortley
phase convention to be discussed in detail in Subsection 4.18.3 (see in particular
Footnote 18 and Fig. 4.3). They ensure that the 3 3 raising and lowering matrices
formed from the 3 3 spatial submatrices of the generators (4.55) and (4.56),

0 0 1

L = L1 iL2 = 0 0 i ,
1 i
0

(4.312)

282

4 Relativistic Free Particles and Fields

have the positive matrix elements:

L+ (1) = 2 (0),

L (+1) = 2 (0),








2 (+1),

L (0) = 2 (1).

L+ (0) =

(4.313)
(4.314)

The vectors (1) and (0) are the so-called spherical components of the threedimensional unit vectors

1
related by

= 0 ,
0

2

= 1 ,
0

3

= 0 ,
1

(4.315)

(1) 12 (1 i2),

(4.316)

Together with the unit vector

 (0) = 3,

(4.317)

they form a basis of the unitary spin-1 representation of the rotation group.
In order to obtain the polarization vectors (k, 1) for momenta in an arbitrary
We shall use the rotation
direction, we must rotate (1) into the direction k.
4
matrix

R(, ) = eiL3 eiL2


with the spherical angles

cos cos sin sin cos


cos sin sin
=

cos sin
sin
0
cos

[0, ),

[0, 2),

(4.318)

(4.319)

to arrive at a momentum direction

with the polarization vectors

sin cos

=
k
sin sin ,
cos

(4.320)

cos cos i sin


1
(k, 1) = cos sin i cos
.
2
sin

(4.321)

Together with the third vector

 (k, 0) k,
4

Some authors prefer the matrix R(, ) = eiL3 eiL2 eiL3 .

(4.322)

283

4.8 Plane-Wave Solutions of Maxwells Equations

for which there is no electromagnetic plane wave, they form a representation of


L the matrices L from
spin 1 diagonalizing L2 and the helicity matrix H = k
Eq. (4.57). The labels = 1 specify the two helicities of a light wave of momentum
k. Experimentally, these waves are observed as right and left circularly polarized
light.
Since the spatial polarization vectors are orthogonal to the momentum k they are
also referred to as transverse polarization vectors, and the Coulomb gauge condition
(4.305) as transverse gauge condition.
The spatial polarization vectors (k, ) are orthonormal:

 (k, ) (k, ) = ,

, = 0, 1,

(4.323)

and transversely complete:


i (k, )

=1,1

j (k, ) = PTij (k) ij kkk2 .


i j

(4.324)

The matrix PTij (k) is a projection to the direction transverse to of k since it satifies
PTij (k)PTik (k) = PTik (k).

(4.325)

The property (4.323) goes over to the four-dimensional polarization vectors


(4.309) as follows:
, = 0, 1.

(k, ) (k, ) = ,

(4.326)

These vectors have the reflection property


(k, ) = (k, ),

(4.327)

which follows from the fact that k k corresponds to , +


(mod 2).
In four dimensions, the polarization sum over (4.324) over two vectors gives the
following polarization tensor:
PT (k)

=1

(k, ) (k, ) =
0

0 ij k i k j /k2

0 PTij (k)

= g

0 k i k j /k2

. (4.328)

This is a 4 4 projection matrix into the two-dimensional purely spatial subspace


transverse to k. It contains only purely spatial nonzero components PTij (k).
The projection is obviously noncovariant since it lends a special significance to the
zeroth component of the electromagnetic vector field. To exhibit the noncovariance,
it is useful to introduce a fixed timelike unit vector

1
0

(4.329)

0
0

284

4 Relativistic Free Particles and Fields

We also define a purely spacelike unit vector orthogonal to it pointing in the direction
of k:
!
0

(4.330)
k .
k
This can be rewritten in terms of and the momentum vector k as follows
k (k)
k q
.
(k)2 k 2

(4.331)

The fixed vector eliminates the zeroth component of k , no matter whether it is on


shell or off shell.
We readily show that the polarization sum (4.328) can be rewritten as
PT (k) g + k k
k + k

k k
2
+
k

k
, (4.332)
= g
(k)2 k 2
(k)2 k 2
(k)2 k 2
where the zeroth component k 0 can have any value.
One may define scalar products between the solutions by complete analogy with
those for the scalar field in Eqs. (4.174), only that those are multiplied with an extra
scalar product between polarization vectors:
(fp , fp )

d3 x fp (x, t)i 0 fp (x, t),

(4.333)

with obvious definitions between positive- and negative-frequency solutions fp (t)


and fp (t), and between two negative-frequency solutions fp (t).

4.9

Massive Vector Fields

In order to understand weak interactions and some strongly interacting particles,


we must also learn to describe massive vector fields (see Section 7.5). They can be
electrically neutral or with electric charges 1.

4.9.1

Action and Field Equations

The action of a neutral massive vector field V (x) can be obtained by writing down
an action like the electromagnetic one in (4.234), and simply adding a mass term:
A=

d xL(x) =

1
1
d x F F + M 2 V V ,
4
2
4

(4.334)

where the field tensor is now


F (x) V (x) V (x).

(4.335)

285

4.9 Massive Vector Fields

Charged vector fields are described by the action


A=

d xL(x) =

1
F + M 2 V V .
d x F
2
4

(4.336)

In either case, the equation of motion reads


F + M 2 V = 0,

(4.337)

or more explicitly
[( 2 M 2 )g + ]V (x) = 0.

(4.338)

As for electromagnetic fields, the Euler-Lagrange equation for the zeroth component
V 0 (x) does not involve the time derivative of V 0 (x) and is therefore not a dynamical equation, but relates V 0 (x) to the spatial components V i (x) and their time
derivatives via
V 0 (x) = F 0 /M 2 =

i
1 h
2 0

V(x)
+

V
(x)
.
M2

(4.339)

In the limit M 0, this gives rise to Coulombs law (4.264). Taking the fourdivergence of (4.337), we see that the vector field V (x) has no four-divergence
scalar content:
V (x) = 0.
(4.340)
Physically, this eliminates any scalar content s(x) from the vector potential. In
contrast to the electromagnetic vector field A (x), the zero four-divergence is not
a matter of choice as in the Lorentz condition (4.249), but follows here from the
Euler-Lagrange equations. Inserting (4.340) back into (4.338), we find that the four
components of V (x) satisfy the Klein-Gordon equation
( 2 M 2 )V (x) = 0,

(4.341)

which is the massive version of the electromagnetic field equations (4.250) in the
Lorentz gauge.

4.9.2

Plane Wave Solutions for Massive Vector Fields

The plane-wave solutions look the same as in Eqs. (4.300) and (4.301) for the electromagnetic vector potential. The mass of the vector field modifies only the possible
polarization vectors (k, ). As in the electromagnetic case, the zero divergence
property (4.340) eliminates one degree of freedom in the polarization vectors. It is,
however, impossible to eliminate more, since there exists no gauge invariance, and
thus no analog of the restricted gauge transformations (4.252). Adding to V (x) a
gradient (x) and inserting the new field into the field equation (4.337) produces
the condition M 2 (x) = 0 admitting only a trivial constant for (x). The polarization vector has therefore three independent components. Physically this reflects

286

4 Relativistic Free Particles and Fields

the fact that a massive vector particle can be studied in its rest frame. There the
third component of angular momentum L3 , and the square L2 have three eigenstates
corresponding to three linear combinations of the spatial vector components.
With the restriction (4.340) the polarization vectors satisfy k (k, s3 ) = 0,
which leaves at rest the three polarization vectors

1
(0, 1) =

0
1
i
0

(0, 0) =

0
0
0
1

(4.342)

These are obviously eigenstates of the 4 4 angular momentum matrices L3 and L2


in the defining representation (4.54)(4.56). The relative phases have been chosen
as in (4.311) to comply with the Condon-Shortley convention (see Footnote 18 and
Fig. 4.3) specified in Eqs. (4.313) and (4.314) that applying
L+ to (0, ) leads to

the states with = 1 with positive matrix elements 2.


The polarization vectors of momentum k are obtained from those at rest by
applying the boost matrices (4.24):

(k, s3 ) = B( ) (0, s3 ).

(4.343)

The zero-helicity polarization vector (k, 0) is also called longitudinal polarization


vector, which is allowed for massive vector particles.
The three boosted polarization vectors satisfy the orthogonality relations
(k, s3 ) (k, s3 ) = s3 s3 .

(4.344)

In order to find the completeness relations, we boost the polarization vectors at rest
(4.342) to a final momentum k (k , 0, 0, k 3) in the z-direction using the matrix
(4.15) and obtain

H (k 3 z, 1) =

0
1
i
0

H (k 3 z, 0) =

1
M

where z is the unit vector in the z-direction.


We can now calculate the completeness sum
X

(k 3 z, s3 ) (k 3 z, s3 ) =

s3

1
M2

By rewriting the right-hand side in the form


g +

1
M2

k0k0
0
0
0

0
1
0
0

k3
0
0
0

0
0
0
0
1
0
3 3
0 k k

0
1
0
0

0
0
1
0

k3
0

,
0
k

0
0
0
k2

(4.345)

(4.346)

(4.347)

287

4.9 Massive Vector Fields

we recognize the general covariant form for any direction of the momentum k
X

(k, s3 ) (k, s3 )

= P (k) g

s3

k k

M2

(4.348)

If the vectors are first boosted into the z-direction and then rotated into the
direction of their momentum
= (sin cos , sin sin , cos ),
k

(4.349)

with the polar angles , , we obtain the polarization vectors in the helicity basis.
The subsequent rotation into the direction by the trivial four-dimensional extension
of the matrix (4.318) leads to the polarization vectors of helicities = 1 and 0:
H (k, 1)

0
(k, 1)

H (k, 0)

|k|

k k

1
=
M

(4.350)

where (k, 1) in the transverse four-vectors with helicities = 1 are the threedimensional polarization vectors (4.321). The polarization four-vectors H (k, 1)
agree with (4.321), the vector H (k, 0) is an additional longitudinal polarization
vector.
The covariant completeness sum (4.348) can, of course, be derived directly from
the general polarization vectors (4.350), whose completeness sum gives
X

s3 ) (k,
s3 )
(k,

|k|2 /M 2 k 0 k i
k i k 0 /M 2 ij + k i k j /M 2

s3

(4.351)

which agrees with (7.584) and is the obvious rotated form of (4.347).
Let us verify that the polarization vectors (k, s3 ) and its complex-conjugate
(k, s3 ) lead to the correct Wigner rotations for the annihilation and creation
operators ak,s3 and bk,s3 . These have to be the spin-1 generalizations of (4.730) and
(4.736), or (4.731) and (4.732). The simplest check are the analogs of Eqs. (4.737).
For this we form the 4 3 -matrices corresponding to u(0) and v(0) in (4.737) [recall
(4.663) and (4.664)]

(k)

0
1
i
0

0 0
0 1
0 i
0 0

(k)

0
1
i
0

0
0
0
0

0
1
i
0

(4.352)

Multiplying these by the 4 4 -generators [recall of (4.54)(4.56)]

L3 = i

0
0 0
0
0 1
0 1 0
0
0 0

0
0
0
0

L1 = i

0
0
0
0

0
0
0
0
0
0
0 1

0
0
1
0

L2 = i

0
0
0
0

0
0
0
1

0
0
0 1

.
0
0
0
0
(4.353)

288

4 Relativistic Free Particles and Fields

from the left, we find the same result as by applying the 3 3 spin-1 representation
matrices [recall (4.820)(4.821)]:

1 0 0
0 1 0
0 1 0
1
i

j
j
j

D (L3 ) = 0 1 0 , D (L1 ) =
1 0 1 , D (L2 ) =
1 0 1
2 0 1 0
2 0 1 0
0 0 1
(4.354)
from the right.
For finite Lorentz transformation, the polarization vectors undergo the Wigner
rotations

(k, s3 )
(k, s3 ) = (k, s3 ) =

1
X

(1)

(p, s3 )Ws ,s3 (p , , p),


3

s3 =1

(k, s3 )
(k, s3 ) = (k, s3 ) =

1
X

(4.355)

(1)

(p, s3 )Ws ,s3 (p , , p). (4.356)


3

s3 =1

(1)

where Ws ,s3 is the spin-1 representation of the 2 2 Wigner rotations (4.730). The
3
second transformation law follows from the first as a consequence of the reality of
the Lorentz transformations .
In closer analogy with (4.672), we can obtain the charge-conjugate polarization
vectors by forming
(1)

c (k, s3 ) (k, s3 )cs s3 ,


3

(4.357)

(s)

where cs s3 denotes the spin-s representation of the rotation matrix c = ei2 /2


3
introduced in Eq. (4.872). It is easy to verify that c (k, s3 ) coincides with (k, s3 )
(s)
The matrix cs s3 turns the spin into the opposite direction:
3

(s)

cs3 ,s = hs, s3 |eiL2 /2 |s, s3 i = dss3,s3 () = ()s+s3 s3 ,s3 .


3

(4.358)

and has the property [recall (4.873)]:


D j (R)c(j) = c(j) D j(R),

(4.359)

which confirms that c (k, s3 ) transforms with the complex-conjugate of the Wigner
rotation of (k, s3 ).
In principle, there exists also a fourth vector which may be called scalar polarization vector to be denoted by
H (k, s) k .

(4.360)

This polarization vector will be of use later in Subsecs. 4.10.6 and 7.4.2. It corresponds to a pure gauge degree of freedom since in x-space it has the form .

289

4.10 Gravitational Field

As such, it transforms under an extra independent and irreducible representation of


the Lorentz group describing a scalar particle degree of freedom and no longer forms
part of the vector particle. It certainly does not contribute to the gauge-invariant
electromagnetic action.
Observe that for a very small mass where k |k|, the longitudinal polarization
vector has the limit
!
M 0
1

MH (k, 0)
|k| = k ,
(4.361)
k
i.e., it goes over into the unphysical scalar degree of freedom. This is why the longitudinal polarization does not contribute to the action af a massless vector particle.
In this limit, the actions (4.334) and (4.336) become invariant under local gauge

transformations of the vector field V (x) V (x) = V (x) + (x).

4.10

Gravitational Field

The gravitational field is carried by a varying metric g (x) in spacetime. Its presence manifests itself in a local dependence of the invariant distance between events
(ds)2 = g (x)dx dx .

(4.362)

The distances in such a spacetime no longer satisfy the axioms of Minkowski geometry since the spacetime can have a local curvature. According to the equivalence
principle of general relativity, the motion of point particles is independent of their
mass. All point particles follow the lines of shortest distance in this geometry, the
so-called geodesics.
The generation of the gravitational field is governed by a complicated nonlinear
theory and deserves a detailed treatment on its own not to be elaborated in this
text. In the present context we merely state a few relevant facts about a very weak
gravitational field running through empty spacetime. It may be described by a
small deviation of the metric g (x) from the Minkowski metric, which in this case
is denoted by , for better distinction. The deviation is
h (x) g (x) .

4.10.1

(4.363)

Action and Field Equations

The action of the field h (x) is obtained from the famous Einstein-Hilbert action
of the gravitational field
f

A =

1
2

d4 x gR,

(4.364)

where R is the the curvature scalar of spacetime. It is formed from the Riemann
curvature tensor R by the contraction g R with the Ricci tensor R = R .
The symbol g denotes the determinant of the metric tensor g which makes the

290

4 Relativistic Free Particles and Fields

volume element gd4 x invariant under coordinate transformations. The curvature


tensor is the covariant curl5
R [ , ]

(4.365)

of the Christoffel connection


1
= ( ) g ( g + g g ).
2

(4.366)

The integrand of (4.364) can be replaced by



i

h


gR = (g g) + gg . (4.367)

The first term is a pure divergence, and may be omitted in the action (4.364),
keeping only the second term. After inserting g = + h , g h + . . .
from (4.363), we calculate

a
b
a
g = e(1/2)tr log(ab hab ) = e(1/2)tr log(ab ) e(1/2)ha (1/4)ha hb +...
1
1
1
(4.368)
= 1 + ha a ha b hba + (ha a )2 + . . . ,
2
4
8
and the Christoffel symbols to linear order in h ,


1
h + h h ,
2

(4.369)

Inserting this into (4.365), we find for the curvature tensor (4.365) the linear contributions
R

i
1h
h h ( ) + . . . ,
2

(4.370)

which for the Ricci tensor


R g R

(4.371)

1
R ( h + h h 2 h ) + . . . .
2

(4.372)

amounts to

The ensuing scalar curvature R R starts out like


R ( 2 h h ) + . . . .
5

(4.373)

This definition is completely analogous to the definition in nonabelian gauge theories [see
(27.11)]. The relation to the standard Riemann tensor in gravitational textbooks is R = R
This makes our Ricci tensor (4.371) equal to minus the Ricci tensor of gravitational textbooks,
and the negative sign carries over to the Einstein tensor (4.378), leading to the Einstein equation
(5.56).

291

4.10 Gravitational Field

The quadratic part of the action is found by inserting (4.369) into the 2 -terms of
Eq. (4.367):
f


1 Z 4 
.
d xg

2

(2)

(4.374)

Inserting everything into (4.364), we find the quadratic part


pression for g R
of the action governing free gravitational waves:
f

1
8

d4 x

n

h + h h ( h + h h )


h + h h ( h + h h ) .

The right-hand side can be rearranged, using the symmetry of h , to

1 Z 4
d x( h h 2 h h + 2 h h h h). (4.375)
A =
8
f

where h is defined to be the trace of the tensor h , i.e., h h . Equivalently we


may write
f

A =

1
8

d4 x h h .

(4.376)

The equivalence can be verified with the help of the identity


= + + + .
(4.377)
A further useful way of writing the field action (4.376) is obtained in terms of the
Einstein tensor
1
G R g R
2

(4.378)

whose linear approximation reads


1
G = R g R
2
1
1 2
( h + h h h ) + ( 2 h h ), (4.379)
2
2
with a trace G G = R = 2 h h . This may be written as a fourdimensional double curl:
1
G h ,
4

(4.380)

from which we see that the action (4.364) becomes simply


f

A =

d4 x L(x) =

1
2

d4 x h G ,

(4.381)

292

4 Relativistic Free Particles and Fields

where the constant is proportional to Newtons gravitational constant


GN = 6.672 108 cm3 g 1sec2 ,

(4.382)

in the combination

8GN
.
(4.383)
c3
It specifies the force between two planets of masses M, M and distance r being
=

F = GN MM /r 2 .
Instead of one also uses the so-called Planck mass
MP
or the Planck length

h
c/GN = 2.1737 105 g = 1.22 1019 GeV,

(4.384)

lP h
/MP c = 1.616 1033 cm,

(4.385)

M 2 c2
h

1
= P
=
.
2
16 h

16lP2

(4.386)

so that the prefactor 1/2 can be written as

The Einstein tensor plays a similar role in gavity as the dual field tensor F does
in electromagnetism [recall (4.235)]. First, being a double-curl (4.380), it trivially
satisfies a Bianchi identity
G = 0,
(4.387)
for any smooth single-valued field h (x), i.e., any field h (x) which satisfies the
integrability condition ( )h (x) = 0. Second, it is invariant under local
gauge transformations [just as (4.236) is under (4.247)]:
h (x) h (x) + (x) + (x).

(4.388)

These are the linearized versions of Einsteins general coordinate transformations:


x x + (x).

(4.389)

As a consequence of this invariance, the symmetric tensor h (x) carries only 6


instead of 10 independent physical components.
f

The free-field equations are obtained by variation of A with respect h (x):


f

1
A
= G (x) = 0.
(4.390)
h (x)

Thus, for a free gravitational field, the Einstein tensor vanishes, and so does the
Ricci tensor R = G :
R (x) = 0.
(4.391)
In the presence of masses the field equation (4.390) will be modified by being equal
to the energy-momentum tensor T [see Eq. (5.56)], which is simply the total
symmetric energy-momentum tensor of the material particles.

293

4.10 Gravitational Field

4.10.2

Lorentz Transformation Properties of Gravitational Field

Under Lorentz transformations, h (x) behaves of course like a tensor. In a straightforward generalization of the transformation law (4.280) we may immediately write
h (x) = h (1 x).

(4.392)

This transformation can be generated by analogy with (4.282) as follows:

i J

h (x)
h (x) = D()
h (x),
h (x) e 2

(4.393)

with the parameters specified in (4.70), (4.71), and


,
J L 1 1 + 1 L 1 + 1 1 L

(4.394)

this being a direct generalization of the total angular momentum operator (4.283).
The commutation rules between the generators J are of course given by (4.285)
as in the case of the vector potential.

4.10.3

Other Symmetries of Gravitational Action

Just as the scalar and electromagnetic actions, the gravitational action (4.381) is
invariant under extensions of the Lorentz group.

4.10.4

Translations

Under translations (4.182),


x = x + a ,

(4.395)

the gravitational field transforms like


h (x) = h (x a).

(4.396)

The combination of these with Lorentz transformations forms the Poincare group
(4.186),
x = x + a ,
(4.397)
under which the fields transform like
h (x) = h (1 (x a)),

(4.398)

leaving the action (4.381) invariant.


As in the scalar and vector cases (4.190) and (4.292), we can of course generate
all Poincare transformations on the field with the help of differential operators:

ia p /
h i 12 J
D()

h (x)
h (x) = D(a)
e
h (x),
h (x) e
(4.399)
with the parameters specified in (4.70), (4.71).

294

4.10.5

4 Relativistic Free Particles and Fields

Space Inversion, Time Reversal, and Charge Conjugation

Since h (x) determines the invariant distances in space via (4.362) and (4.363),
it transforms like dx dx under space inversion and time reversal. Under charge
conjugation, it is invariant.
Thus we have under space inversion
P

x),
h (x)
h
P (x) = h (

(4.400)

where the tilde reverses the sign of h (x) for each spatial index, whereas under time
reversal:
T

(4.401)
h (x)
h
T (x) = h (xT ).
In principle, a tensor field t (x) has two possible transformation behaviors under
space inversion:
P
x),
(4.402)
t (x)
t
P (x) = P t (

with P = 1 and in the case P = 1 the field t (x) would have been called
a pseudotensor field. The gravitational field h (x), however, is definitely a tensor
field. This follows from its metric nature and the distance definition in (4.362),
(4.363).
Similarly, the phase factor of a tensor field t (x) arising from time reversal could
in principle be
T

(4.403)
t (x)
t
T (x) = T t (xT ),
with an arbitrary phase factor T . For an observable real field, however, only T =
1 are admissible. The gravitational field has T = 1 to preserve the definition of
the distance in (4.362), (4.363) under space inversion.
Under charge conjugation, the gravitational interactions are invariant, so that
C

h (x)
h
C (x) = h (x).

(4.404)

In general, a tensor field t (x) could transform like


C

t (x)
t
C (x) = C t (x)

(4.405)

with C = 1.

4.10.6

Gravitational Plane Waves

To discuss the properties of gravitational waves we remove the gauge freedom, we


remove the gauge degree of freedom (4.388) by fixing a specific gauge, the so called
Hilbert gauge:
1
h(x) = h (x).
(4.406)
h (x) = h(x),
2
It corresponds to the Lorentz gauge of electromagnetism [recall (4.249)].

295

4.10 Gravitational Field

The Hilbert gauge can always be achieved by a gauge transformation. If h (x)


is not in this gauge, we simply perform the transformation (4.388) and determine
(x) from the differential equation
2 (x) = h (x) 21 h(x).

(4.407)

The gauge transformation (4.388) can be used to fix another propert of h . By


applying it to the trace h = h with the special gauge function = we find
h h = h + 2 2 .

(4.408)

This can be used to arrive at a traceless field h . If h is nonzero, h vanished if we


choose
= (1/2 2 )h.

(4.409)

It is useful to introduce the field


1
h h,
2

(4.410)

With this, the Einstein tensor (4.379) reads


1
1
G ( 2 ) .
2
2

(4.411)

with the trace G = G = 12 ( 2 + 2 ). Under the gauge transformations


(4.388), the field (x) changes like
(x) (x) + (x) + (x) (x),

(4.412)

and the Einstein tensor is invariant.


Imposing now the Hilbert condition (4.406), i.e.,
(x) = 0,

(4.413)

the Einstein tensor (4.411) reduces to


1
G = 2 ,
2

(4.414)

and the free equation of motion (4.390) implies a massless Klein-Gordon equation
for each field component6
2 (x) = 0.
(4.415)
Thus must be solved in the Hilbert (4.413).
Since the graviton field obeys the zero-mass Klein-Gordon equation (4.415), the
plane waves in the field (x) are proportional to eikx and eikx , with k 0 lying
6

Compare Eq. (4.250) for the electromagnetic field in the Lorentz gauge

296

4 Relativistic Free Particles and Fields

on the light cone k 0 = k = k2 . These waves are accompanied by symmetric


polarization tensors (k, ) [compare (4.300)]:
fk (x, t) =

1
(k, )eikx ,
0
2V k

fk =

1
(k, )eikx .
0
2V k

(4.416)

In the Hilbert gauge, the polarization tensors (k, ) satisfy transversality condition
k (k, ) = 0.
(4.417)
In principle, there exists 10 possibilities of forming symmetric polarization tensors (k, ). The gauge condition (4.417) eliminates four of them. The remaining
six components can be constructed from symmetrized tensor products of the transverse polarization vectors in (4.350), and the scalar one in (4.360):

H (k, 2) H (k, 1)H (k, 1),


1

H (k, 1) [H (k, 1)H (k, 0) + ( )],


2
s
2
1

(k, 0)H (k, 0),

H (k, 0) [H (k, 1)H (k, 1) + ( )] +


3 H
6
1

H (k, 1) [H (k, 1)H (k, s) + ( )],


2

H (k, 2) H (k, 1)H (k, 1),

H (k, s)

1
[H (k, 1)H (k, 1) + ( )]
3

1
H (k, s)H (k, 0). (4.418)
3

The combinations are formed with the Clebsch-Gordan coefficients [see Appendix
Appendix 4C and Table 4.2] which couple two vectors symmetrically to five components of spin 2 and one component of spin zero.
|2 2i = |1 1i|1 1i,

1 
|2 1i = |1 1i|1 0i + |1 0i|1 1i
2
s

2
1 
|1 0i|1 0i
|2 0i = |1 1i|1 1i + |1 1i|1 1i +
3
6

1 
|2 1i = |1 1i|1 0i + |1 0i|1 1i
2
|2 2i = |1 1i|1 1i,

1
1 
|0 0i = |1 1i|1 1i + |1 1i|1 1i |1 0i|1 0i.
3
3

(4.419)

Since the three polarization vectors satisfy k H (k, ) = 0, the tensors satisfy
the Hilbert gauge condition (4.417). In addition, they are traceless
H (k, ) = 0,

(4.420)

297

4.10 Gravitational Field

as follows directly from the expressions (4.350), (4.360) and the properties
H (k, 1)H (k, 1) = 0, H (k, 1)H (k, s) = 0, H (k, s)H (k, s) = 0. The tracelessness of the polarization tensors is, of course, a consequence of the invariance
under the gauge transformation (4.408).
Because of gauge invariance, not all of these tensors correspond to physical degrees of freedom. Under a gauge transformation (4.412), the Fourier transform of
the field (k) receives an additional term
k + k k ,

(4.421)

which cannot contribute to any observable quantities. If we do this transformation


for physical energy-momenta on the mass shell, four more components of the polarization tensors (4.419) can be eliminated as being unobservable. This corresponds
to the second type of gauge transformations that were discussed for electrodynamics
in Eq. (4.252). Thus, of the six degrees of freedom (4.419) only two remain physical.
Setting = H (k, 1), H (k, 1), and H (k, s), and using (4.417), we see that
the three combinations (4.421) become precisely the polarization tensors
H (k, 1)
and
(k,
s),
respectively.
H
But also the polarization tensor H (k, 0) is unphysical. This is seen by introducing a forth linearly independent polarization vector
H (k, s) = (k 0 , k).

(4.422)

The name vector is set in quotation marks since it indicates here merely a fourcomponent object. It does not mean that H (k, s) behaves like a vector under
Lorentz transformations. This can be seen most easily by forming the product
2

H (k, s)H (k, s) = k 0 + k2 = 2k2 ,


which is not Lorentz-invariant. We shall call the vector H (k, s) an antiscalar polarization vector. The four vectors H (k, 1), H (k, s), H (k, s) form a complete
basis in the space of polarization vectors. This is expressed in the completeness
relation
H (k, 1)H (k,1)+H (k,1)H (k, 1)+

1
[ (k, s)H (k, s)+( )] = .(4.423)
2k2 H

The first two terms on the left-hand side can obviously be replaced by the polarization tensor
H (k, 0), so that (4.423) may be rewritten as

H (k, 0) =

1
[ (k, s)H (k, s) + ( )].
2k2 H

(4.424)

Inserting on the right-hand side the explicit form H (k, s) = k for the scalar polarization vector (4.360), and setting H (k, s)/2k2 = , the right-hand side can
be rewritten as k k k , thus demonstrating that the polarization
tensor
H (k, 0) is of the pure gauge form (4.421), and thus unphysical.

298

4 Relativistic Free Particles and Fields

Thus we remain with only two physical polarization tensors


H (k, 2). These
describe gravitational waves with helicities = 2. An analysis of the temporal
behavior of the fields shows that the tensors
H (k, 2) give the gravitational waves
a circular polarization, anticlockwise or clockwise around the momentum direction,
respectively, the analoga of the circularly polarized light described by plane electromagnetic waves with polarization vectors H (k, 1) discussed in Section 4.8.
In electromagnetism, one often describes plane wave with real polarization vectors
H (k, 2) = i[H (k, +1) H (k, 1)],
(4.425)
corresponding to linearly polarized light waves whose field vectors oscillate along
the directions orthogonal to the momentum k. By analogy, we introduce the real
combinations of the two physical polarization tensors
H (k, 1) = [H (k, +1) + H (k, 1)],

H+ (k) [H (k, +2) + H (k, 2)]


2
1

H (k) = [H (k, 2) H (k, 2)].


2i

(4.426)
(4.427)

An analysis of the motion of a circular ring of mass points in a plane gravitational


field g (x) = + h (x) with the polarization tensors (4.427) reveals that
H+ (k)
corresponds to waves distorting the circular ring periodically into an ellipsoidal one,
with vertical and horizontal axes, whereas
H (k) does the same thing with the axes
rotated by 45 degrees. For more details see Section 5.6.
The two polarization tensors satisfy the completeness relation

H+ (k)H+ (k) + H (k)H (k) = H (k, 2)[H+ (k, 2)] + H (k, 2)[H (k, 2)]

(4.428)
= P , (k),
TT

is the projection matrix


where the spatial part of PT, (k)
i
h
1 P (k)P
(k)
, (4.429)
(k)
P (k)P
(k)
+ P (k)P
(
k)
PT,
T
T
T
T
T
T
2 T

is the transverse projection matrix (4.328) of the electromagnetic


where PT (k)
waves. It is easy to verify the projection property:
, (k)
= P , (k).

PT,
T (k)PT T
TT

(4.430)

are, if
The explicit form of the nonzero purely spatial components of PT,
(k)
T
applied to a symmetric tensor,
il jm

PTij,lm
21 ij lm 2 il k j km + 21 ij kl km + 21 lm k i k j + 12 ki kj k l k m . (4.431)
T (k) =

This projection produces always a traceless tensor since

PTii,lm
T (k) = 0,

PTij,ll
T (k) = 0,

(4.432)

299

4.11 Relativistic Free Fermi Fields

as a reflection of the gauge condition (4.420) following from the invariance (4.408).

If PTij,lm
T (k) is applied to a traceless symmetric tensors, it can be replaced by
il jm

PTij,lm
2 il k j km + 21 ki k j kl km .
T (k) =

(4.433)

After the on-shell gauge fixing of the fields h in the Hilbert gauge, they have
the gauge properties h0i = 0, h0 0 = 0, h = 0. Referring to the projection (4.429),
they are referred to as T T -waves.
can, of course, be ob(k)
The competely covariant tensor expression for PT,
T
tained by inserting the projection matrix for spin-1 modes (4.332) into Eq. (4.429).

4.11

Relativistic Free Fermi Fields

For Fermi fields, the situation is technically more involved. Experimentally, fermions
always have an even number of spin degrees of freedom. We shall denote the associated field by a , where the index a labels the different spin components. Under
rotations, these spin components are transformed into each other, as observed experimentally in the Stern-Gerlach experiment. We shall see below that Lorentz
transformations also lead to certain well-defined mixtures of different spin components.
The question arises whether we can construct a Lorentz-invariant action involving
(2s + 1) spinor field components. To see the basic construction principle we use the
known transformation law (4.280) for the four-vector field A as a guide. For an
arbitrary spinor field we postulate the transformation law

a (x)
a (x ) = Dab ()b (x),

(4.434)

with an appropriate (2s + 1) (2s + 1) spinor transformation matrix Dab () which


we have to construct. This can be done by using purely mathematical arguments.
The construction is the subject of the so-called group representation theory. First
of all, we perform two successive Lorentz transformations,
x = x = 2 x = 2 1 x.

(4.435)

Since the Lorentz transformations 1 , 2 are elements of a group, the product


2 1 is again a Lorentz transformation. Under the individual factors 2 and 1 ,
the field transforms as
1

(x)
(x ) = D(1 )(x),
2

(x )
(x ) = D(2 ) (x ),

(4.436)

so that under = 2 1 ,
2 1

(x)
(x ) = D(2 )D(1 )(x).

(4.437)

300

4 Relativistic Free Particles and Fields

For the combined , the transformation matrix is D(), bringing the field to
(x ) = D(2 1 )(x).

(4.438)

Comparing this with (4.437) shows that the matrices D() which mix the spinor
field components under the Lorentz group must follow a group multiplication law
which has to be compatible with that of the group itself. The mapping
D()

(4.439)

is a homomorphism, and the D()s form a matrix representation of the group.


Note that the transformation law (4.280) for the vector field A (x) follows the
same rule with
1

D() = ei 2 L
(4.440)
being the defining 4 4 -representation of the Lorentz group. The parameters
are specified in (4.70), (4.71).
The scalar field transformation law (4.50) follows trivially the same rule, where
D() 1 is the identity representation.
The group laws for and D() are sufficiently stringent to allow only for a countable set of fundamental 7 finite-dimensional transformation matrices D(). Below
we shall see that these are characterized by two quantum numbers, s1 and s2 , either
one of these taking all possible half-integer or integer values 0, 12 , 1, 32 , . . . . The
representation spaces associated with D (s1 ,s2 ) () will turn out to harbor particles of
spins |s1 s2 | to s1 + s2 . Hence, particles with a single fixed spin s can only follow
the D (s,0) () or D (0,s) () transformation laws.
From the development in Sections 4.18 and 4.19 we know how the representation
matrices D (s1 ,s2 ) () are most efficiently constructed. We have to determine all possible sets of six matrices which satisfy the commutation relations (4.86). Any set of
such matrices forms a representation of the Lie algebra defined by (4.86).
In this general framework, the scalar particles studied so far transform with the
trivial representation D (0,0) () 1 of the Lorentz group. They are said to have spin
zero.
Let us now study the smallest nontrivial representation.

4.12

Spin-1/2 Fields

The smallest matrices satisfying the subalgebra (4.76) associated with the rotation
subgroup of the Lorentz group are
Li =
7

i
,
2

(4.441)

Mathematically, fundamental means that the representation is irreducible. Any arbitrary representation is equivalent to a direct sum of irreducible ones.

4.12

301

Spin-1/2 Fields

where i are the Pauli matrices


1

0 1
1 0

0 i
i 0

1
0
0 1

(4.442)

The two-component basis on which L2 and L3 are diagonal, with eigenvalues 3/4
and 1/2, respectively, are the Pauli spinors
1
0

( ) =
1
2

12

0
1

)=

(4.443)

The full Lie algebra (4.76)(4.78) can be satisfied in two inequivalent ways. Either
Mi = i i /2,

(4.444)

Mi = i i /2.

(4.445)

or
The first representation
(Li , Mi ) =

i
i
, i
2
2

(4.446)

is denoted by ( 21 , 0), the second one


i i
,i
2 2

(Li , Mi ) =

(4.447)

by (0, 12 ). Exponentiating these generators we obtain the global representations of


the Lorentz group
1
D ( 2 ,0) () = ei('/2i/2) ,
(4.448)
D (0, 2 ) () = ei('/2+i/2) .
1

(4.449)

An alternatively decomposition which will later be useful is a factorization into


rotations and subsequent boost:
D ( 2 ,0) () = e/2 ei'/2 ,

(4.450)

1
D (0, 2 ) () = e/2 ei'/2 .

(4.451)

The right-hand sides can easily be calculated explicitly by expanding them in a


Taylor series. The expansion terms separate naturally into even and odd powers
due to the normalization property of the -matrices ( )2 = 1, ( )2 = 1, where
, are the directions of rotation axis and Lorentz boost, respectively. For rotations,
this implies that for integer k: ( )2k = 2k , ( )2k+1 = ( ) 2k+1 , so that

'

i'/2

'

'

(i)n
=
n!
n=0

'

'

!n
 

X
X

(1)k 2k
'

=
i'

k=0

(2k)!



(1)k

2
k=0 (2k + 1)!


2k+1

. (4.452)

302

4 Relativistic Free Particles and Fields

Summing up even and odd powers of the parameters separately, this becomes
R' () = ei'/2 = cos

 '

i sin .
2
2

(4.453)

Similarly:

B( ) = e/2 = cosh

  sinh 2 .

(4.454)

When applying these representations to a particle of mass M at rest, energy and


momentum are boosted from p = (M, 0) to

p = (p0 , p) = M(cosh , sinh ).

(4.455)

Using the relations

=
cosh
2

1
(cosh + 1) =
2

p0 + M
,
2M

sinh
=
2

1
(cosh 1) =
2

p0 M
,
2M

(4.456)

we can express the pure Lorentz transformations of the ( 21 , 0)-representation in terms


of energy and momentum of the boosted particle as follows:

1
B( ) = e/2 = q
(p0 + M
2M(p0 + M)

 p).

(4.457)

We may also use the Einstein parameter to write this as

1
B( ) = q
( + 1 v/c).
2( + 1)

(4.458)

It is convenient to introduce an extra 2 2 Pauli matrix


0

1 0
0 1

(4.459)

and define a four-vector of matrices


( 0 , i ).

(4.460)

Note that the four Pauli matrices satisfy a multiplicative algebra:


0 0 = 0 , i 0 = 0 i = i , i j = iijk k + ij 0 .

(4.461)

With , we can form Lorentz-covariant matrices


p p = p .

(4.462)

4.12

303

Spin-1/2 Fields

This notation allows us to write the boosts in the ( 12 , 0)-representation of the Lorentz
group as
M + p
e/2 = q
,
(4.463)
2M(p0 + M)

For a corresponding expression in the (0, 21 )-representation we define, by analogy


with x (x0 , xi ) in Eq. (4.192), the space inverted vectors:
p (p0 , pi ),

= ( 0 , i ).

(4.464)

Then:
M + p

M + p
=q
.
e/2 = q
2M(p0 + M)
2M(p0 + M)

(4.465)

For many explicit calculations to follow it is useful to realize that the boost
matrices (4.463) and (4.465) may be considered as the square root of the same
expression with twice the rapidity
e = cosh

  sinh .

(4.466)

Because of (4.455), the right-hand sides have simple momentum representations


e =

p
,
M

e =
.
M

(4.467)

Using these we may write the boost matrices shorter as


e/2 =

p
,
M

e/2 =

,
M

(4.468)

a notation which will be convenient in later calculations.


Having thus succeeded in finding the smallest dimensional representations of the
Lorentz group, we define fields (x) and (x) to have the corresponding transformation laws

(x)

(x ) = D ( 2 ,0) () (x),
1

(x)
(x ) = D (0, 2 ) () (x).

(4.469)

The different transformation behavior of the two kinds of spinors is exhibited by the
lower undotted and upper dotted indices. These spinors were introduced by Weyl,8
and are referred to as Weyl spinors of type ( 21 , 0) and (0, 21 ), respectively.
Let us now see whether we can construct a Lorentz-invariant free-field action
from Weyl spinors which can contain only quadratic terms in the fields and their
8

H. Weyl, J. Phys. 50, 330 (1929).

304

4 Relativistic Free Particles and Fields

first derivatives. First we look for invariant quadratic terms without spacetime
derivatives, which are needed to describe particles with a mass. Let us begin by
looking for suitable combinations of and . which are invariant only under the
1
1
rotation subgroup. For this, the representation matrices D ( 2 ,0) and D (0, 2 ) are both
equal to the same unitary matrix U = ei/2 to be applied to both and . Due
to the unitarity of U, all quadratic expressions , , , are rotationally
invariant quadratic field combinations, for example,
(x ) (x ) = (x)U U(x)
= (x)(x).

(4.470)

1
1
Consider now pure Lorentz transformations. Then D ( 2 ,0) = e/2 and D (0, 2 ) =
e/2 are both nonunitary but hermitian. The quadratic expressions and
are no longer invariant. However, since the two representation matrices are inverse
to each other,

1
1 1
1
(4.471)
D ( 2 ,0) = D ( 2 ,0) = D (0, 2 ) ,

the mixed quadratic expressions and are invariant field combinations. Thus
we conclude: If the action is to contain a quadratic field combination without space

time derivatives then both spinors , , i.e. four field components 1 , 2, 1 , 2 , are
necessary to form an invariant.
Let us now construct a Lorentz invariant term involving spacetime derivatives
. It is necessary if an action is to describe a particle which can move through
spacetime. Consider first the spatial derivatives. It is easy to see that
i i ,
i i ,

i i ,
i i ,

(4.472)

are all rotationally invariant. Take for example


(x ) i i (x ) = (x)U i i U(x).

(4.473)

From the commutation rules among the Pauli matrices i it is easy to derive the
transformation law
U 1 i U = R' ()i j j .
(4.474)
The proof of this proceeds in the same way as in the finite transformation of the
spatial vector xi in (4.104) with the help of Lies expansion formula (4.103). The
commutation rules between i 3 /2 and i are precisely the same as those between
3 and xi , so that we find for a rotation around the 3-axis,
L
3

ei /2 3 ei /2 = 3 ,
3
3
ei /2 1 ei /2 = cos 1 sin 2 ,
3
3
ei /2 2 ei /2 = sin 1 + cos 2 ,

(4.475)

corresponding precisely to (4.474) with the matrix (4.8). The 2 2 representation


has the advantage used already in (4.453) that the exponentials can be expanded

4.12

305

Spin-1/2 Fields

into linear combinations of 0 and i . If we do this on the left-hand sides, we can


calculate the right-hand sides also using products of i rather than commutators as
in Lies expansion formula (4.103).
Now, since derivatives transform like a vector,
R

i
i = R' ()ij j ,

(4.476)

the expression (4.473) is indeed a rotationally invariant field combination. For the
other terms in (4.472), the proof is the same.
How can we extend the expressions (4.472) to form relativistic invariants? For
this we remember that in the boost matrices (4.468), the four-vector generalizations
,
of the Pauli matrices appeared naturally contracted with p . This suggests
studying
,
,

,
,

(4.477)

and once more the same combinations but with replaced by


.
The additional time derivatives in (4.477) are trivially invariant under rotations
and thus do not destroy the rotational invariance of the spatial parts shown in
(4.473). Consequently we have to study only the behavior under pure Lorentz
transformations in, say, the z-direction. Under these, the x- and y-components do
not change. It is easy to verify that
0 e

3 /2

= e

( 3 )e

3 /2

= e ( 3 ) = sinh 0 + cosh ( 3 ),

e
e

3 /2

3 /2

= cosh 0 + sinh ( 3 ),

e /2 0 e /2 = e = cosh 0 + sinh 3 ,
3
3
3
e /2 3 e /2 = e 3 = sinh 0 + cosh 3 .

(4.478)

In contrast to the transformation laws (4.475), this cannot be derived with the help
of Lies expansion formula (4.103), since the exponentials on the left-hand sides have
the same exponents. However, as before in the calculation of (4.454), we may use
the multiplicative algebra of i -matrices rather than the commutation rules between
them to expand the exponentials to expand the exponentials into linear combinations
of 0 and i . and calculate (4.478).
The first two lines in (4.478) show that the matrices transform like a fourvector under ( 21 , 0)-boosts, the remaining lines that do the same under (0, 21 )boosts. In combination with rotations, we thus have proved the Lorentz transformation behavior:
1

D ( 2 ,0) ()
D ( 2 ,0) () =
,
1

D (0, 2 ) () D (0, 2 ) () = ,

(4.479)

This allows us to conclude that the quadratic field terms with a derivative

,

(4.480)

306

4 Relativistic Free Particles and Fields

are Lorentz-invariant. For instance


1

(x ) (x ) = (x)D (0, 2 ) () D (0, 2 ) ()(x) = (x) (x)


= (x) (x).

(4.481)

It is easy to see that the other quadratic combinations in (4.477) are not invariant.
If we allow only for these lowest-order in the derivative terms, the most general
Lorentz-invariant action reads
A=

d x L(x) =

d4 x (
i + i M1 M2 ).

(4.482)

Observe that this expression involves necessarily both two-component spinors and
. Only for zero parameters M1,2 , a single species of two-component spinors or
possesses an invariant action.
The equations of motion are obviously
i
(x) = M1 (x),
i (x) = M2 (x).

(4.483)

Combining the second equation with the first and vice versa we find the two secondorder field equations
( 2 M1 M2 )(x) = 0,
( 2 M1 M2 )(x) = 0.

(4.484)
(4.485)

In deriving these, we have used the relation

+
= 2g ,

(4.486)

= 2 .

(4.487)

by which
which is easily proved by direct evaluation. In momentum space, equations
(4.484) are solved by particles of mass
M=

M1 M2 .

(4.488)

For M 6= 0, it is useful to combine the two spinors and into a single fourcomponent object, called bispinor or Dirac spinor:
(x)
(x)

(x) =

(4.489)

In terms of this, the action may be written as


A=

d x L(x) =

"

d x (x)
i
4

M2 0
0 M1

!#

(x),

(4.490)

4.12

307

Spin-1/2 Fields

where are the 4 4 Dirac matrices:


0

(4.491)

and (x)
is the adjoint Dirac spinor defined by

(x)
(x) 0 = ( (x), (x)).

(4.492)

The Dirac matrices satisfy the anticommutation rules


{ , } = 2g .

(4.493)

It has become customary to abbreviate the contraction of with any vector v


by
v/ v .

(4.494)

In this notation, the derivative terms in (4.490) becomes simply / .


For the Dirac spinor, the equations of motion (4.483) take the form
"

M2 0
0 M1

i/

!#

(x) = 0.

(4.495)

This is almost, but not quite, the wave equation postulated first by Dirac for the
electron. He assumed a diagonal 4 4 -mass matrix
M2 0
0 M1

=M

1 0
0 1

(4.496)

and proposed the Dirac equation


(i/
M) (x) = 0

(4.497)

corresponding to an action
A=

d4 x L(x) =

(i/
d4 x (x)
M) (x).

(4.498)

Just as in the case of a complex scalar field, there exists a four-vector quantity

j (x) = (x)
(x),

(4.499)

which by virtue of the Dirac equation satisfies a local conservation law [recall (4.168)
and (4.169)]
j (x) = 0.
(4.500)
This four-vector will be used in 12 to couple Dirac fields to electromagnetism, thus
becoming the electromagnetic current density of the Dirac field. Later in Subsec.
8.11.1 we shall see that the local conservation law (4.500) is a consequence of the

308

4 Relativistic Free Particles and Fields

invariance of the Dirac action (4.498) under arbitrary changes of the phase of the
field
(x) ei (x).
(4.501)
It is this invariance which gives rise to a conserved current density [see (8.272), also
(12.48)].
By construction, the actions (4.490) and (4.498) are invariant under the bispinor
or Dirac-Lorentz transformations

a (x)
a (x ) = Da b ()b (x),

(4.502)

where the 44 -matrices D() consists of rotations and pure Lorentz transformations
D(R) =

ei'/2 0
0
ei'/2

e/2 0
0
e/2

D(B) =

(4.503)

With the 2 2 matrices (4.450) and (4.451) we can write D() in the form
1

0
D ( 2 ,0) ()
(0, 21 )
0
D
()

D() =

(4.504)

The invariance can be seen most directly by combining the transformation laws
(4.479) to the 4 4 -form

D()
D() = .

(4.505)

where we define
1

D()

0
D ( 2 ,0) ()
(0, 12 )
0
D
()

= D 1 ().

(4.506)

From (4.505), the invariance of the derivative term in (4.498) follows at once, using
(4.136), (4.137), and (4.44):


D()

i (x ) (x ) = i(x)
D() (x) = i(x)
(x)

= i(x)
(x).
(4.507)

In terms of Dirac matrices (4.491), the pure Lorentz transformations can also be
written as
D(B) =

M + p/ 0
q

2M(p0 + M)

(4.508)

The representation matrices of all Lorentz transformations may be expressed in


terms of covariant generators as
1

D() = ei 2 S ,

(4.509)

4.12

309

Spin-1/2 Fields

where is the same antisymmetric matrix as in (4.69) containing both the rotation
and boost parameters as specified in (4.70) and (4.71). Taking the matrices (4.503)
to the limit of small and , we identify the 4 4 -matrix generators S as

'

ij

S = ijk

1
2

k 0
0 k

i
S =
2
0i

i 0
0 i

(4.510)

The 4 4 -generator of the rotation group on the left-hand side contains the direct
4 4 -extension of the Pauli matrices is denoted by


and the spin matrix
S

1
2

,

(4.511)

(4.512)

with components S i = 21 ijk S jk .


The generator of the pure Lorentz transformations on the right-hand side of
(4.510) is also written as S 0i = ii /2 with the matrix

(4.513)

It is customary to introduce the matrices


i
[ , ].
2

(4.514)

In terms of these, equations (4.510) can be summarized as


1
S .
2

(4.515)

It is easy to check that the matrices S satisfy the same commutation rules as the
defining generators of the Lorentz group in (4.87):
[S , S ] = ig S ,

no sum over .

(4.516)

Thus they are generators of a new 4 4 -representation of the Lorentz group. The
44 -Dirac representation matrices D() transforming bispinors are mathematically
inequivalent to the defining 44 -representation of the Lorentz group transforming
vectors.
With the gamma matrices, the generators S have the following commutation
rules
[S , ] = (L ) = i(g g ),

(4.517)

which state that is a vector operator in bispinor space [recall (4.97)]. The transformation law (4.505) is the global version of this transformation law.

310

4 Relativistic Free Particles and Fields

In terms of the generators S , we can write the field transformation law (4.502)
more explicitly as

(x)
(x) = D()(1x) = ei 2 S (1 x),

(4.518)

As in the case of the scalar field [recall (4.120)], it is useful to perform the
transformation of the spacetime argument on the right-hand side in terms of the
differential operator of four-dimensional angular momentum and rewrite (4.518) as

(x)
(x) = D()D()(x)
= ei 2 J (x),

(4.519)

.
J S 1 + 1 L

(4.520)

where

are the generators of the total four-dimensional angular momentum of a Dirac field
by analogy with (4.283). The tensor products exhibits the two separate representation spaces associated with Dirac index and spacetime coordinates. The commutation rules between the generators J are of course the same as in the case of the
vector potential (4.285), and in fact for any spin.

4.13

Other Symmetries of Dirac Action

As in the scalar case, the spin-1/2 action (4.498) is invariant under more than just
the Lorentz group.

4.13.1

Translations and Poincar


e group

First, it is automatically invariant under translations (4.182) for which

(x)
a (x) = (x a).

(4.521)

Together with the Lorentz transformations, these form the inhomogeneous Lorentz
group or the Poincare group (4.186). Under it, the spinor transforms as
,a

(x)
(x) = (1 (x a)).

(4.522)

Extending (4.519), we can generate all Poincare transformations by the operations


,a

(x)
(x) = D()D()(x)
= eia p /h ei 2 J (x).

4.13.2

(4.523)

Space Inversion

In contrast to scalar fields, the Ponicare invariance does not automatically imply
invariance under parity transformations. In the quantum theory of electrons and
photons, quantum electrodynamics (QED), it is an additional requirement confirmed

311

4.13 Other Symmetries of Dirac Action

by all experimental data. The action of free electrons must therefore be invariant
under parity transformations. To achieve this, we have to first define an appropriate
way how to transform a bispinor under space reflections P . The bispinor must form
a representation of the Lorentz group extended by P
P

(x)
P (x) = D(P )(
x),

(4.524)

where xP = x as in (4.192), and the representation matrices D(P ) must combine


with the representation matrices of the Lorentz group D() in the same way as the
4 4 matrix of space reflections

P =

(4.525)

combines with the Lorentz transformations . From the explicit matrices (4.54)
(4.56) and (4.60)(4.62) we find
P 1LP = L,

P 1 MP = M.

(4.526)

Thus rotations commute with reflections, which is intuitively obvious since x and
x rotate both with the same 3 3 matrices. Pure Lorentz transformations, on the
other hand, are space-inverted into those for the opposite direction. Since upper and
lower components in a Dirac spinor contain the boost matrices (4.468) in opposite
directions, we can immediately write down the transformation law for a Dirac spinor
under space inversion as
P

(x)

P (x)

= P

0 1
1 0

(
x).

(4.527)

The phase P is the intrinsic parity of the field (x). The representation matrix of
the reflection P in Dirac space is denoted by
D(P ) = P

0 1
1 0

= P 0 .

(4.528)

The property P 2 = 1 must be reproduced by the representation matrix D(P ). By


applying two successive space inversions on (x) we can conclude that the intrinsic
parity can only have the values P = 1.
We easily check that the matrix D(P ) transforms the 4 4 -bispinor representation (4.510) of the generators Li and Mi as in the defining representation (4.526):
D 1 (P )D(Li )D(P ) = D(Li ),

D 1 (P )D(Mi )D(P ) = D(Mi ).

(4.529)

We now postulate Lorentz invariance of the action (4.490) under space reflection.
Since
!
!
M2 0
M1 0
1
D (P )
D(P ) =
,
(4.530)
0 M1
0 M2

312

4 Relativistic Free Particles and Fields

this is only possible if M1 = M2 , so that the action takes the Dirac form (4.498).
Also, since space inversion transforms (x) into (x), a parity-invariant theory necessarily contains both fields and thus the full bispinor (x).
Having set M1 = M2 , the mass term of the Dirac action is parity-invariant. In
order to ensure the invariance of the derivative term, we observe that the representation matrix (4.528) of P satisfies
D 1 (P ) D(P ) = .

(4.531)

Hence we calculate
1

(x )i (x ) = (
x)i (
x) = (x)D
(P )i D(P )(x)

= (x)i
(x),
(4.532)

which proves the invariance of the derivative term, and thus of the full Dirac action.
The Dirac equation
(i M)(x) = 0
can trivially be rewritten as
(i M)(
x) = 0.
Using (4.527), (4.528) we can replace (
x) by P 0 P (x), and take the matrix 0 to
the left of the Dirac operator with the help of (4.531). The result is
(i M)P (x) = 0,
i.e. the Dirac equation for the mirror-reflected bispinor.
As we shall see in Chapter 12, the interactions of electrons with electromagnetism
are described with extreme accuracy by a parity-invariant theory called quantum
electrodynamics. The electrons in this theory are described by a Dirac action (4.498).
A theory based only on a single two-component spinor field or is necessarily
massless and violates parity. Such a theory was used successfully to describe
neutrino processes. There exist several neutrinos in nature, one associated
with every charged lepton e , , whose masses are me = 0.510 MeV, m =
105.66 MeV, m = 1777.03 0.30 MeV, the latter two having finite lifetimes
= (2.19703 0.00004) 106 sec and = (290.6 1.1) 1015 sec. The corresponding neutrinos are denoted by e , , . The six leptons seem to exist in
nature completely parallel to six quarks which are the elementary building blocks of
strongly interacting particles. The parallel configurations are illustrated in Fig. 4.1.
The electron neutrino was postulated in 1931 by Pauli in order to explain an
apparent violation of energy conservation in the final state of the -decay of the
neutron. Energy conservation, would have been violated if only the observed particles proton and electron emerged from the decay. From the energy spectrum of the
electron one can deduce that the mass of the electron neutrino is extremely small,

313

4.13 Other Symmetries of Dirac Action

Figure 4.1 Six leptons and quarks

e
e

u c t
d s b

less than 2 eV/c2 . The most precise value is expected from an ongoing experiment
performed in Karlsruhe. There one studies the beta radiation of tritium which decay in 12.3 years with a total energy release of 18.6 keV shared by elektron and
neutrino. The energy of the electrons is measured in KATRIN, a giant electrostatic
Spektrometer of 7 m diameter and 20 m length, after which their identity is proved in
a semiconducting detector. The experiment is sensitive to electron neutrino masses
0.4 eV/c2 . The masses of the other two neutrinos have presently the bounds
m < 0.19 MeV and m < 0.18.3 MeV.
In summary, in a mirror-symmetric Lorentz-invariant theory of particles with
the lowest nontrivial spin 1/2, the action is given by Diracs expression (4.498). If
parity is allowed to be violated, there are two simpler Lorentz-invariant actions of
massless particles
Z
Z
A = d4 x L(x) = d4 x (x)i
(x),
(4.533)
or

A=

d4 x L(x) =

d4 x (x)i (x).

(4.534)

The parity-invariant action (4.498) is the correct one for electrons, the action (4.533)
describes neutrinos. For neutrinos, it has become customary to work also with fourcomponent bispinors (x), but make use only of the two upper components of it.
The upper and lower components are extracted from (x) by the projection matrices
Pu,l

1
(1 5 )
2

1 0
0 0

0 0
0 1

(4.535)

5.

(4.536)

where 5 denotes the 4 4 -matrix


5

0 1 2 3

5 = i =

1 0
0 1

This is a Lorentz-invariant matrix, since it may also be expressed in the contracted


form
i
(4.537)
5 = 5 ,
4!
where is the completely antisymmetric tensor with 0123 = 1. Using 5 , the
actions (4.533), (4.534) can be written as
A=

d4 x L(x) =

1
2

d4 x (x)i
(1 5 )(x).

(4.538)

314

4 Relativistic Free Particles and Fields

60
27 Co

Figure 4.2 Asymmetry observed in the distribution of electrons from the -decay of
polarized 60
27 Co.

From (4.531) we see that under space inversion, 5 transforms as follows:


D 1 (P )5D(P ) = 5 ,

(4.539)

thereby interchanging the two actions (4.538) with each other.


The parity-violating actions (4.533) and (4.534), or (4.538), have an interesting
history. They were first proposed by H. Weyl9 in 1929 to describe massless spin-1/2
fields, but rejected on theoretical grounds, since at that time all interactions were
firmly believed to be invariant under space reflections. Electromagnetic and nuclear
interactions had definitely displayed this property, and it was suggestive to assume
that nature should follow the same principle in all its interactions. In 1956, however,
Lee and Yang pointed out that the decay modes of the heavy mesons K 0 and K + ,
the first decaying into 0 0 with relative s-waves, the second into + + with both
+ + and + in relative s-waves, suggested a violation of parity. In 1957, the same
authors stressed the relevance of measuring the -decay from a polarized nucleus.10
If parity was an invariance of the weak interactions, the distribution of electrons
would have to be symmetric with respect to the direction of spin. Indeed, since the
scalar product between the spin, which is an axial vector, and the momentum vector
is a pseudoscalar operator, its expectation value should vanish. In 1957, Madame Wu
and collaborators11 performed the historic experiment observing a nonzero up-down
asymmetry in the distribution of electrons coming from polarized 60
27 Co (see Fig.
12
4.2). The polarization of the sample was done by placing it into a strong magnetic
field. By going to a very low temperature, sufficient population difference between
spin down and spin up was achieved. The experiment showed a clear violation of
parity. In later experiments it was found that the violation is even maximal, in the
sense that the unobserved neutrino emitted in the decay process can only have one
9

H. Weyl, J. Phys. 50, 330 (1929).


T.D. Lee and C.N. Yang, Phys. Rev. 104, 254 (1956); 105, 167(1957).
11
C.S. Wu, E. Ambler, R.W. Hayward, D.D. Hoppes, and R.P. Hudson, Phys. Rev. 105, 1413
(1957).
12
In nuclear physics the customary notation for a nucleus X is Z XN , where A is the nucleon
number, Z the number of protons or atomic number number,atomic(also the charge number ), and
N the number of neutrons. The latter two labels are not necessary for uniqueness since the name
of the nucleus is associated with a unique Z, and since N = A Z. They are, however, convenient
reminders.
10

315

4.13 Other Symmetries of Dirac Action

polarization along its momentum direction, the other being completely forbidden. A
massless neutrino possesses no mirror image in nature and can be described by a
pure Weyl action with only a 21 (1 5 ) field. We shall see in the Chapter 7 that
also a massless antineutrino is described by the field 12 (1 5 ).
The initial 60
27 Co-state has a spin s = 5 and intrinsic parity P = +1, thus
being a sP = 5+ -state. The -decay transforms it into an excited state of 60
28 Ni
P
+
with spin-parity s = 4 . As such, it is a so-called Gamow-Teller transition. In
this transition it can be shown that only combinations of tensor and axialvector
couplings contribute (T-A). The details will be explained in Chapter 26.
A year later a crucial hypothesis was made by several authors13 that the weak
interactions are mediated by a specific combination of vector and axial vector couplings. This is the famous V-A hypothesis which eventually led to the present standard model of weak and electromagnetic interactions (see Chapter 26).
The bispinors 21 (1 5 ) are eigenvalues of the matrix 5 with eigenvalue 1.
They are called states of left and right chirality, respectively.
There exist, of course, many equivalent representations of the Lorentz group
including space inversion on spin-1/2 fields including space inversion. Instead of
the bispinors (x) transforming with the 4 4 -matrices (4.504), in which h parity
exchanges upper and lower two-component spinors in (x) via the matrix D(P ) of
(4.528), consider symmetric and antisymmetric combinations
D (x) SD (x),

(4.540)

where SD is the similarity transformation matrix


1
SD
2

1 1
1 1

(4.541)

In the bispinor D (x), upper and lower components are eigenstates of opposite parity.
These fields transform according to the 4 4 representation
DD () = SD D()SD1

D ( 2 ,0) + D (0, 2 ) () 12 D ( 2 ,0) D (0, 2 ) ()



 1
 1
=
. (4.542)
1
1
( 2 ,0)
(0, 12 )
()
D
+
D
21 D ( 2 ,0) D (0, 2 ) ()
2
1
2

When boosting a massive particle from rest to momentum p this matrix becomes
explicitly, with (4.468) and (4.465),

13

DD (B( )) =

1
2

e/2 + e/2

21 e/2 e/2







21 e/2 e/2
1
2

e/2 + e/2

E.C.G. Sudarshan and R.E. Marshak, Phys. Rev. 109 , 1860 (1958); R.P. Feynman and M.
Gell-Mann, ibid. 193 (1958); J.J. Sakurai, Nuovo Cimento 7 , 649 (1958); W.R. Theis, Z. Phys.
150 , 590 (1958); Fortschr. Physik 7 , 559 (1959).

316

4 Relativistic Free Particles and Fields

1
2
1
2

q

p/M +

q

p/M

= q
2M(M + p0 )

p
/M

p
/M

12

1
2

q

p/M

q

M + p0 p
p
M + p0

p/M +

p
/M

p
/M




(4.543)

The Dirac matrices which ensure in this case the invariance of the action (4.490) are
now
D

= SD

SD1

1
0
0 1

0 1
1 0

D5 = SD 5 SD1 = iD0 D1 D2 D3 =

 !0

!)

(4.544)
(4.545)

In terms of these, the boost transformation (4.543) takes the form


M + p/ D D0
,
DD (B) = q
2M(p0 + M)

(4.546)

which is the same as the similarity-transformed SD D(B)SD1 of the boost matrix


(4.508) in the chiral representation. The generators are
SDij

= SD S

ij

SD1

= ijk

1
2

k 0
0 k

SD0i

= SD S

0i

SD1

i
=
2

0 i
i 0

,(4.547)

which are equal to


i

D
SD SD1 = [D , D ],
2

(4.548)

as in (4.514). The generator of rotations SDij are the same as in the chiral representation (4.510). Indeed, writing SDij = ijk 12 iD , we see that the 4 4 -generalization
of the Pauli matrices (4.511) is invariant under the similarity transformation SD :

D = SDSD1 =

,

(4.549)

so that we can write SDij = ijk 21 i , as before in Eq. (4.512).


For small momenta, the boost matrix (4.543) has the limit

DD (B( ))

p /2
p /2
1
1

(4.550)

This shows the spinors in Diracs representation of the gamma matrices have small
lower (or upper) components for slow particles (or antiparticles). The Dirac representation is therefore useful for studying the nonrelativistic limit of Dirac particles.

317

4.13 Other Symmetries of Dirac Action

For such calculations it is useful to state the Dirac matrices in a direct-product


form
2
(4.551)
D0 = 3 1,
D = i ,

in which D5 = 1 1 and the generators of the Lorentz group (4.548) take the form
ij
D
= ijk k = ijk 1 k ,

0i
D
= i 1 i .

(4.552)

In each case, the second matrix acts on the up and down components of the spin,
the first matrix mixes upper and lower components.
Actually, it is this representation of the Dirac matrices which was found by Dirac
himself in his original paper.14 This is why it is used in most textbooks. It is also referred to as the standard representation. In the standard representation, the chirality
matrix 5 is not diagonal, as in the previous representation (4.536). To emphasize
this property, the 4 4 -matrices (4.491) are called the chiral representation of the
gamma matrices (also the Weyl representation).
As a historical note we mention that Dirac did not find his matrices from grouptheoretic considerations but by searching for a relativistic Schrodinger equation for
an electron which, in contrast to the Klein-Gordon equation contains only a single
time derivative, so that there would be no negative-energy solutions. He wanted the
time-independent electron field to satisfy the wave equation

H(x)
=

2 + M 2 (x) = E(x)
p

(4.553)

with only the positive square root. Since in relativistic theories energy and momentum appear on equal footing, he searched for a way to take an explicit square
root if the result was allowed to be a matrix. Then (x, t) could consist of several
components representing the spin degrees of freedom of the electron. So he made
the ansatz
D (x) = (i pi + M)(x) = E(x),
H
(4.554)
D twice
with i , being unknown matrices. Then he required that by applying H
2
2
2
D (x) = (
upon (x) should give H
p + M )(x) = E (x). This led him to the
algebraic relations
n

i , j

i ,

= ij ,
= 0,

(4.555)

2 = 1.

He solved these anticommutators with the matrices


= D
14

D0

1
0
0 1

= D

D0 D

P.A.M. Dirac, Proc. Roy. Soc. A 117 , 610 (1928), A 118 , 351 (1928).


0

(4.556)

318

4 Relativistic Free Particles and Fields

He could of course have solved them just as well in the chiral representation by
0

(4.557)

By multiplying Eq. (4.554) with and going over to a time-dependent equation


by replacing E by p0 = ix0 , he obtained the Dirac equation in the form
(D p M)(x) = 0.

(4.558)

with the matrices


D0 ,

Di i .

(4.559)

The anticommutation relations (4.555) go over into the anticommutation relations


(4.493) for the Dirac matrices
{D , D } = 2g .

(4.560)

Inserting into the Dirac equation (4.558) the decomposition (4.489) of (x) into
the two-component spinors (x) and (x), we find the Dirac equation in hte form




it (x) + i (x) =
M (x),
it (x) + i (x) = M (x),

(4.561)

which can also be written as


(t +

D + iMD )(x) = 0,

(4.562)

It interesting to note that the Mawell equations (4.243) and (4.244) can be brought
to a similar form using a spin-1 version of the matrix D :

0 L
L 0

(4.563)

where (Li )jk = iijk are the generators (4.57) of rotation for vectors. The Maxwell
bivector is
M

E
iB

(4.564)

The two separate vector equations read


t E + L (iB) = 0,
t (iB) + L E = 0,

(4.565)

which are indeed the same as the Maxwell equations (4.243) and (4.244).
Note, however, that the bivectors (4.564) cannot be used to set up an action
analog to Diracs (4.490) for zero mass. That must involve the vector potential A

319

4.13 Other Symmetries of Dirac Action

to be local. A formulation that incorporates the dual symmetry between electricity


an magnetism is possible by defining the bivector components B1 = E, B2 = B, the
associated vector potentials AA whose curls are these fields, i.e.,
B A = AA ,

(4.566)

and the Levi-Civita tensor AB = BA with 12 = 1. Then we can write the Maxwell
action as
A=


1Z 4 A
d x B AB t AB AB BB .
2

(4.567)

Indeed, a variation of this action yields the Maxwell equations since


A =

d4 x BA AB t AB AB BB ,

(4.568)

which after an integration by parts becomes


A =

d4 x AA AB t BB AB BB ,

(4.569)

implying the Maxwell equations (4.243) and (4.244) [6]:


t B + E = 0,

t E B = 0.

(4.570)

The equations is invariant under the duality transformation


E = B,

B = E,

(4.571)

and so is the action (4.569).


The second of the transformations (4.571) corresponds to a nonlocal transformation of the vector potential
A = (2 )1 E.

(4.572)

In the chiral representation, the equations (4.561) read, using (4.483) with M1 =
M2 = M:




it (x) i (x) = M (x),


it (x) + i (x) = M (x).

(4.573)

We have noted before when calculating the small-momentum limit (4.550) that
the Dirac representation is most convenient for studying the nonrelativistic limit.
This limit, in which Mc2 , corresponds in natural units to letting M .
The energies of slowly moving particle are very close to M, so that (it M)(x) is
much smaller than (it +M)(x), which can be approximated by 2M(x). The lower
equation in (4.561) can therefore be solved approximately by the relation between
lower and upper spinor

(x).
(4.574)
(x) i
2M

320

4 Relativistic Free Particles and Fields

We can also remove the fast temporal oscillations as in (4.153) and replace
(x)
(x)

iM c2

1
t/
h

2M

(x)
(x)

(4.575)

If we solve again the lower equation by a relation like (4.574), the upper equation
reduces to the Schrodinger equation for each spinor component:
it (x) =
Using

1
( )2 (x).
2M

(4.576)

( )2 = i j i j = ij + iijk k i j = 2 ,
this becomes

(4.577)

1
2 (x).
(4.578)
2M
In the presence of electromagnetic interactions, the last step will be nontrivial.
yielding the nonrelativistic Pauli equation (6.115) with the correct magnetic moment
of a Dirac particle.
In both representations, we can insert one equation into the other and find that
(x) and (x) satisfy the Klein-Gordon equations (4.484) with M1 = M2 = M:
it (x) =

( 2 M 2 )(x) = 0,
( 2 M 2 )(x) = 0.

(4.579)

This follows of course simply from the Dirac equation (4.497) upon multiplying it
by (i/
+ M)and working out


(i/
+ M) (i/
M) (x) = /
2 M 2 (x) = 2 M 2 (x).

(4.580)

In the massless case, the Dirac equations (4.561) have a very similar structure
to Maxwells equations (4.243) and (4.244):
t B + E = 0,
t E B = 0.

(4.581)

To see this similarity we rewrite the cross product with the help of the 3 3 generators (4.57) of the rotation group, (Li )jk = iijk , where they read
it E + i(L ) (iB) = 0,
it (iB) + i(L ) E = 0,

(4.582)

thus becoming quite similar to the Dirac equations (4.561) derived from Diracs representation of -matrices. The reader is encouraged to discuss the analogy between
the transformation properties of (, ) in Diracs representation and (E, iB) and the
generators and L.

321

4.13 Other Symmetries of Dirac Action

It is worth mentioning that Diracs procedure of looking for a matrix version of


the wave equation (4.553) has other solutions, for example
2
FV = p (3 + i2 ) + M3
H
2M

(4.583)

The subscript indicates that this Hamiltonian was found by Feshbach and Villars.
Here Pauli matrices have no relation to spin and serve only to specify the 2 2
It is easy to verify, using the multiplication rules (4.461), that just as
matrix H.
2 has the
the 4 4 Dirac matrix HD of Eq. (4.554), also the 2 2 -matrix H
FV
2
FV
FV (x) = E(x) have
2 + M 2 . Thus the solutions of the equation H
square H
=p
again the proper relativistic energy-momentum relation. However, contrary to the
solutions of the Dirac equation (4.554), they carry no spin. In fact, a field theory
based on the Lagrangian density


(x, t)
L = (x, t) it H

(4.584)

is completely equivalent to the Klein-Gordon theory of scalar particles.

4.13.3

Charge Conjugation

In Section 4.5.4 we observed that the action of a scalar field was invariant under
an extra discrete symmetry not related to the Lorentz group, charge conjugation.
It consisted in a simple exchange of the scalar field by its complex conjugate. A
similar invariance can be found for the action of the Dirac field. There is only one
complication: We must make sure that this operation commutes with the Lorentz
group. Thus we must form linear combinations of the components of the conjugate
bispinor (x) which transform again like the original bispinor (x). Let us call this
new bispinor
c (x) C T (x),
(4.585)
where the superscript T on the right-hand side indicates a transposition of the row
which makes c . a column vector. The operation of charge conjugation is
vector ,
then defined by
C

with a phase

(x)
C (x) = C c (x),

(4.586)

C = 1.

(4.587)

The matrix C is determined by the requirement that C (x) must satisfy the Dirac
equation
(i/
M)C (x) = 0.
(4.588)
Inserting the right-hand side of (4.586) this reads

or

(i/
M)C T (x) = 0,

(4.589)

[iC 1 C M]T (x) = 0.

(4.590)

322

4 Relativistic Free Particles and Fields

The transpose of this is

(x)[i(C
C)T M] = 0.

(4.591)

Consider, on the another hand, the adjoint of the Dirac equation


(i M)(x) = 0,
which is

(4.592)

(x)(i M) = 0.

(4.593)

Multiplying this by 0 from the right and using the fact that
(0)1 0 = ,
we see that

(4.594)

M) = 0.
(x)(i

(4.595)

Comparing this with (4.591) we conclude that T (x) satisfies the Dirac equation if
the matrix C satisfies
C 1 C = T .

(4.596)

In both the chiral and the Dirac representation, the transposition of changes only
the sign of 2 . A matrix C with this property in the chiral representation is given
by
c
0
0 c

C=

where c is the 2 2 matrix


2

c = i =

0 1
1
0

(4.597)

(4.598)

This matrix is the two-dimensional representation of rotation around the 2-axis by


an angle :
2
c = ei /2 ,
(4.599)
as can easily be verified using (4.453) [or by a direct power series expansion as
in (4.452)]. From this rotation property it follows directly [or via Lies expansion
formula (4.103) as in (4.475)] that
1
1

1
2
2
c c =
,
3
3

and we find
c1
0
0 c1

c
0
0 c

c1 c
1
c
c
0
0

(4.600)

= ( 0 , 1 , 2 , 3 ) = T ,

(4.601)

323

4.13 Other Symmetries of Dirac Action

so that (4.596) is fulfilled.


Note that the 2 2 -matrix c satisfies the identities
c = cT = c1 = c ,

(4.602)

which also hold for the 4 4 -matrices C:


C = C = C T = C 1 = C .

(4.603)

Using these properties, we find that the conjugate Dirac field behaves under the
transformation (4.586) as
C

(x)

C (x) = C c (x),

with

(4.604)

c (x) T (x)C.

(4.605)

This follows from the simple calculation:

= T 0 (C T )T 0 = (C 0T )T 0 = T 0 C T 0 = T C = c . (4.606)
Note that

c (x) = c (x),

(4.607)

since
c = c 0 = (C T ) 0 = (T ) C 0 = ( 0 )T C 0 = T 0T C 0 = T C.

(4.608)

The minus sign on the right-hand side of Eq. (4.604) will be seen in Chapter 7 to
have the important consequence that antiparticles have the opposite intrinsic parity
of particles.
By writing the charge conjugation matrix (4.597) as
C = i 0 2 ,

(4.609)

we can take the result directly to the Dirac representation (4.544) where15
CD

iD0 D2

c
0
0 c

= SD

SD1

0 c
c 0

(4.610)

The reason for the name charge conjugation is the same as in the scalar case.
In contrast to the scalar case, however, this cannot simply be seen by studying the
effect of charge conjugation upon the conserved particle current. In contrast to
Eq. (4.226) which shows that the current reverses its sign under charge conjugation,
the operation (4.585) with C satisfying (4.596) leaves the current density (4.499)
unchanged:
C

j (x)
j (x) = j (x).
15

The minus sign is added to agree with Diracs sign convention for CD .

(4.611)

324

4 Relativistic Free Particles and Fields

This follows directly from


C

j (x) = (x)
(x)
T (x)C 1 C T (x)

= T (x) T T (x) = (x)


(x) = j (x)

(4.612)

The proper physical effect will only be reached after field quantization. This makes
the fields anticommuting fermion operators which produce a sign change in the last
step of the transformation (4.612), thus justifying the name charge conjugation for
the operation (4.585).
It is possible to imitate this effect of quantization at the cloassical level by imagining the classical fields to be anticommuting or Grassmann variables. Such fields
will be introduced in Chapter 13 and used in Chapter 25.

4.13.4

Time Reversal

Let us now see how time reversal acts upon the Dirac field. Under time reversal,
the direction of a particle momentum and angular momentum are both reversed,
and the generators of the Lorentz group are subject to an automorphism (4.130).
The same automorphism is now applied to the 4 4 bispinor representation D(T ).
Writing D(T ) = T DT we must have
DT1 D(Li )DT = D(Li ) ,

DT1 D(Mi )DT = D(Mi ) .

(4.613)

The explicit form of the transformation matrix DT is now determined by the


requirement that the time-reversed field T (x) defined by
T

(x)
T (x) = D(T ) (xT ),

(4.614)

with xT =
x, satisfies the Dirac equation
(i M)T (x) = 0.

(4.615)

The reason for the complex conjugation of the field on the right-hand side of (4.614)
was discussed in Subsec. 4.5.3, where it was shown that the Schrodinger equation for the time-reversed Schrodinger operator carries a complex conjugation [see
Eq. (4.219)]. This is needed to keep the energy in the time-dependent phase factor
0
eip t positive for t t.
Inserting (4.614) into (4.615) we obtain
D 1 (T )(i M)D(T ) (
x) = 0.

(4.616)

From the original Dirac equation we know that


(i
M)(
x) = 0,

(4.617)

(i
M) (
x) = 0.

(4.618)

or

4.13 Other Symmetries of Dirac Action

325

To be compatible with (4.616), the matrix D(T ) has to satisfy


D 1 (T ) D(T ) = .

(4.619)

In both the chiral and the Dirac representation, the -matrices have the property
= T .

(4.620)

Using the property (4.596), of the matrix C, we can substitute


T = C C 1 ,

(4.621)

D 1 (T )C C 1 D(T ) =
.

(4.622)

D(T ) = T C5 .

(4.623)

and the condition (4.619) becomes


This is satisfied by

It is easy to verify that this matrix transforms the generators of the Lorentz group
for the Dirac spinors (4.515) as required by (4.613):
D 1 (T )S D(T ) = S .

4.13.5

(4.624)

Transformation Properties of Currents

An important role in interacting field theory is played by bilinear combinations of


the Dirac field formed with 16 combination of Dirac matrices, collectively called ,
which all are selfadjoint under the Dirac conjugation (4.492):
= 0 0 .

(4.625)
These are the scalar, vector, tensor, axialvector, and pseudoscalar matrices
i

S 1, V ,
= [ , ], A 5 , P i5 , (4.626)
T
2
which form a so-called Clifford algebra. They are used to define corresponding
current densities. The most important of these is vector current density

j (x) = (x)
(x),
(4.627)
V

which turns out to be the source of electromagnetism (see Chapter 12). By sandwiching the other -matrices between two Dirac fields, one obtains fields which
transform under the Lorentz group as scalar, tensor, axialvector, and pseudoscalar
fields. For instance,

jA (x) = (x)
5 (x)
(4.628)
is an which, together with the vector current, is responsible for weak interactions.
The combination
i

jT (x) = (x)[
, ](x)
(4.629)
2
is a tensor current related to the current spin density to be introduced in Section
8.6.2. The different possible current densities are shown in Table 4.1, which also
lists the behavior of these currents under the discrete transformations T, C, P , and
their various combinations.

326

4.14

4 Relativistic Free Particles and Fields

Majorana Fields

In the chiral and Dirac representations of -matrices used so far, the bispinor
fields (x) are necessarily complex since only 2 and thus 2 is imaginary, whereas
0 , 1 , 3 and thus 0 , 1 , 3 are real. One may then wonder whether the Dirac equa
tion (i/
M)(x) = 0, and thus the Lorentz transformations ei S , necessarily
mix real and imaginary parts of a spin-1/2 field. It can easily be seen that this is not
so. The complex conjugate Dirac fields are transformed by the 4 4 -representation
matrices
1

D () =

D ( 2 ,0)
0
(0, 12 )
0
D
1

(4.630)

As far as rotations are concerned, D ( 2 ,0) and D (0, 2 ) are equivalent to the original
representations by a similarity transformation
1

D ( 2 ,0) = c1 D ( 2 ,0) c,
1
1
D (0, 2 ) = c1 D (0, 2 ) c,

(4.631)

with c = i 2 . This follows directly by writing the 2 2 rotation matrices in the


explicit form:


ei'/2

= ei'

/2

(4.632)

The complex conjugation reverses the 1- and 3-components in the exponent, since
1 , 3 are real, while preserving the 1-component, since 2 is imaginary. Using
(4.600) we see that
c1 c = ,
(4.633)
so that the right-hand side of (4.632) becomes

ei' = c1 ei'/2 c,

(4.634)

Table 4.1 The transformation properties of various composite fields. We have omitted the
j =
, j = i [ , ], j =
phase factors P , C , T . The notation is jS = ,
V
T
A
2
5 , P = i
5 . The wiggle on the tensor field in the table denotes the parity

transformation on both indices. In the transformation law for charge conjugation we have
inserted the minus sign which will arise only after a second quantization of the Dirac field,
as explained after Eq. (4.612).

P
C
T
P CT

jS (x)
jV (x)

jV (
x)
jS (
x)
jS (x)
jV (x)
jV (
x)
jS (
x)
jS (x) jV (x)

jT (x)
i

2 [ , ]
jT (
x)
jT (x)
jT (
x)
jT (x)

jA (x)
5

jA (
x)
jA (x)
jA (
x)
jA (x)

jP (x)
5
i
jP (
x)
jP (x)
jP (
x)
jP (x)

327

4.14 Majorana Fields

which is the same as (4.631). Therefore, the charge conjugate bispinor


T

(x) C (x) = C

0T

(x) =

0 c
c
0

(x)

(4.635)

transforms under rotations just as (x) itself.


Consider now pure Lorentz transformations of the complex-conjugate bispinor
(x):

(x)
(x ) =

e
0

/2

e /2

(x).

(4.636)

With (4.633), the right-hand side becomes


!
c1 e/2 c
0
(x).
0
c1 e/2 c

(4.637)

Writing (x) as in (4.489) we see that the upper complex conjugate components
c (x) transform like the lower components , whereas the lower components c (x)
transform like (x). Hence also under Lorentz transformations, c (x) behaves like
, and we can write for the entire proper Lorentz group the transformation law
c (x) = D() c (x),

(4.638)

with the transformation matrix D() satisfying the relation


1

C 1 0 D () 0 C = C 1 D()C
= D()

(4.639)

[recalling the definition of Dirac-adjoint matrices (4.492)]. Since and c both


transform in the same way under D() we may form the combinations
1
( + c ),
2
1
( c ),
2i

(4.640)

which are separately irreducible representations of the Lorentz group and eigenstates
of charge conjugation with charge parity C . Since the original field had 4 complex
degrees of freedom, these combinations only can have half as many degrees of freedom, i.e. four real degrees of freedom. Explicitly, the components of the bispinors
(4.640) satisfy in the chiral representation the relations
1 = 4 ,
1 = 4 ,

2 = 3 ,
2 = 3 .

(4.641)
(4.642)

We may now ask whether there are -matrices which make these real degrees of
freedom explicit. This would be the case if we would find a representation of the

328

4 Relativistic Free Particles and Fields

-matrices in which C 0 is the unit matrix. Then c would be equal to and the
fields , would be purely real. Such a representation does indeed exist. It is given
by the -matrices in the so-called Majorana representation:
0
M
2
M

0 2
2 0

0 2
2
0

1
M

3
M

i 3 0
0 i 3

i 1
0
0
i 1

,
!

(4.643)

They are obtained from in the choral representation (4.491) by a similarity transformation

1
M
= SM SM
,
(4.644)
with the transformation matrix
1
SM =
4

1 2
1 + 2
1 + 2 1 + 2

(4.645)

The action written with the Majorana matrices M


is invariant under Lorentz transformations,
1
DM () = SM D()SM
.
(4.646)

In the Majorana representation (4.643), all -matrices are purely imaginary, so that
the Dirac equation

(iM
M)(x) = 0
(4.647)
is purely real. The complex conjugate field satisfies the same equation as itself.

M) (x) = 0.
(iM

(4.648)

A matrix C of complex-conjugation satisfying (4.596) is now given by


CM =

0
M

0 2
2 0

=i

0 c
c 0

(4.649)

2
In contrast to the other two representations, this is normalized to CM
= 1 rather
2
than C = 1, satisfying

C = C T = C 1 = C .

(4.650)

rather than (4.603). This is more convenient here since we want two successive
applications of the operations (4.635) to be the identity operation, ( c )c = . This
T
requires (C 0 )2 = 1. In the other two representations of the gamma matrices where
C anticommutes with 0 , this implies C 2 = 1. In the Majorana representation
where C and 0 commute, one has C 2 = 1. Note that up to a factor i, the matrix
CM happens coincide with CD of Eq. (4.610).
It should be pointed out that CM is not related to C by a similarity transfor
1
0 2
does not have the same sign changes under
mation SM CSM
= iM
M , since M

329

4.14 Majorana Fields

transposition as and D : Whereas T = ( 0 , 1 , 2 , 3 ) holds also for the


T
Dirac matrices D , the Majorana matrices satisfy M
= ( 0 , 1 , 2 , 3 ).
According to (4.908), C 0 is equal to the unit matrix, so that
C = ,

(4.651)

The bispinors (4.640) reduce to


1
( + ),
2

1
( ),
2i

(4.652)

which are now real fields transforming irreducibly under the Lorentz group. They
are called Majorana spinors.
Under the operation of charge conjugation they transform into themselves
C

(x)
C (x) = C (x)

(4.653)

with a charge parity C = 1.

Note that between Majorana spinors (x), the quadratic expressions


M
and

M
M are identically zero.
The 5 -matrix (4.536) has now the form
5 M =

4.14.1

0 1 2 3
iM
M M M

2 0
0 2

=i

c
0
0 c

5
M
.

(4.654)

Plane-Wave Solutions of Dirac Equation

By analogy with the scalar case we now seek for all plane-wave solutions of the Dirac
equation (4.497):
(i/
M)(x) = 0.
(4.655)
We make an ansatz
eipx
,
fp s3 (x) u(p, s3) q
V p0 /M

eipx
fpc s3 (x) v(p, s3 ) q
,
V p0 /M

(4.656)

thereby distinguishing as in (4.177) waves with positive and negative frequencies,


and allowing for a spin orientation index s3 . Due to the presence of Dirac indices,
the solutions will no longer be merely the complex-conjugates of each other, as
in (4.149). The superscript of fpc s3 (x) indicates the appropriate generalization of
complex conjugation.
If the wave functions in (4.656) are to solve the Dirac equation (4.655), the
bispinors u(p, s3 ) and v(p, s3 ) in momentum space have to satisfy the Dirac equations in momentum space
(/
p M)u(p, s3 ) = 0,

(/
p + M)v(p, s3 ) = 0.

(4.657)

330

4 Relativistic Free Particles and Fields

The normalization of these wave functions will be chosen as in the scalar case by
requiring the charge (4.172) of these solutions to be of unit size, with the charge
density j 0 (x) of Eq. (4.499).
In Section 4.4 we have introduced scalar products for solutions of the KleinGordon equation (4.174) with the help of the zeroth component of the conserved
particle current. This is generalized to the Dirac case by introducing the scalar
products
(fp s3 , fp s3 )
(fpc s3 , fpc s3 )

(fpc s3 , fp s3 )
(fp s3 , fpc s3 )

d3 x fp s3 (x, t) 0 fp s3 (x, t) = p ,p s3 ,s3 ,

d3 x fpc s3 (x, t) 0 fp s3 (x, t) = 0,

d3 x fpc s3 (x, t) 0 fpc s3 (x, t) = p ,p s3 ,s3 ,

d3 x fp s3 (x, t) 0 fpc s3 (x, t) = 0.

(4.658)

From these we deduce the orthonormality conditions for the bispinors:


p0
p0
s3 ,s3 ,
v(p, s3 ) 0 v(p, s3 ) =
s ,s ,
M
M 3 3
u(p, s3 ) 0 v(p, s3 ) = 0,
v(p, s3 ) 0 u(p, s3 ) = 0.
(4.659)
u(p, s3 ) 0 u(p, s3) =

The reversal of the momentum in v(p, s3 ) appears in the second line since the
spatial integrals in (4.659) enforce opposite momenta in scalar products between
solutions of positive and negative frequency. According to this, vanishing scalar
products in bispinor space are necessary to produce orthogonality.
In contrast to the scalar product (4.174) for Klein-Gordon wave functions, both
positive- and negative-frequency solutions have now a positive charge, since for any
0

spinor (x)
(x) = (x)(x) is positive definite.
The explicit form of the bispinors u(p) and v(p) depends on the representation
employed for the matrices , and we shall discuss the different cases separately.
In an infinite volume we shall use plane wave functions analogous to (4.178):
fp s3 (x) u(p, s3 )eipx ,

fpc s3 (x) v(p, s3 )eipx ,

(4.660)

which satisfy the Lorentz-invariant orthonormality conditions


(f

p s3

, fp s 3 )

(fpc s3 , fpc s3 )
(fpc s3 , fp s3 )

(fp s3 , fpc s3 )

- p) ,
d3 x fp s3 (x, t) 0 fp s3 (x, t) = 2p0 (p
s3 ,s3

d3 x fpc s3 (x, t) 0 fp s3 (x, t) = 0,

- p) ,
d3 x fpc s3 (x, t) 0 fpc s3 (x, t) = 2p0 (p
s3 ,s3

d3 x fp s3 (x, t) 0 fpc s3 (x, t) = 0.

(4.661)

331

4.14 Majorana Fields

Spinors in Chiral Representation


Using the chiral representation (4.491) for , Eqs. (4.657) take the form
0 p
p
0

0 p
p
0

u(p, s3) = Mu(p, s3 ),

v(p, s3 ) = Mv(p, s3 ). (4.662)

We can immediately write down 4 2 -matrices solving these equations. The first
is solved by

1 q p
M

u(p) =
,
p

2
M

(4.663)

the second by

1
v(p) =
2

p
qM
p

(4.664)

This follows from the matrix identities


s

p
M
M

p 2 p

=M
M
M

p
M

p p

MM

(4.665)

and
pp
=

1
(p p
+ p p
) = p p = M 2 ,
2

(4.666)

the latter being a direct consequence of (4.486).


The 4 2 -matrices (4.663) and (4.664) can be multiplied by an arbitrary 2 2 matrix from the right, and they will still solve the equations (4.662). There are
several convenient choices for such a matrix with different advantages, as we shall
see below. The two-column vectors in the 4 2 -matrices form independent bispinor
solutions of Eqs. (4.662). The projection into these is accomplished by multiplication
from the right with two real two-component real unit spinors, the Pauli spinors
(4.443),
( ) =
1
2

1
0

( ) =
1
2

0
1

(4.667)

By taking the 4 2 -matrices (4.663) as they stand and multiplying them with the
unit spinors (4.667), we obtain the canonical bispinors
q

p
1
M
u(p, s3 ) = q p
(s3 ).

2
M

(4.668)

332

4 Relativistic Free Particles and Fields

The unit spinors (4.667) are eigenvectors of the spin-1/2 generator L3 = 3 /2 of


the rotation group:
L3 ( 12 ) = 21 ( 21 ),

L3 ( 12 ) = 21 ( 12 ).

(4.669)

The associated bispinors at rest


0

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

u(0, 21 ) =

,
u(0,

.
=
=

)
=
2
1
1

2 ( 2 )
2
2 ( 2 )
2 0
1
(4.670)
are eigenstates of the 4 4 bispinor representation of the generator of rotations
around the z-axis [recall (4.511)(4.515)]:

S =S

12

1
0
1
0

3 0
0 3

1
1
1
= 12 = 3 =
2
2
2

(4.671)

In order to construct explicit bispinors v(p, s3 ) we do not directly multiply them


with the unit spinors (4.667) from the right hand side as we did to obtain u(p, s3 )
in (4.668), but use first the above-observed freedom of multiplying (4.664) by an arbitrary 2 2 -matrix from the right. This is necessary to construct a spinor v(p, s3 )
with the physically most appropriate transformation properties unter Lorentz transformations.
It is possible to find from the 42 -matrices (4.664) directly the solutions v(p, s3 )
of the second equation in (4.662). We simply define e v(p, s3 ) as the charge conjugate
spinor of v(p, s3 ) by an operation of the form (4.585), i.e.,
v(p, s3 ) = C uT (p, s3 ).

(4.672)

It is easy to verify that this v(p, s3 ) solves (4.662). To see this we take

 q
q
1
p

p
u(p, s3 ) = u (p, s3 ) 0 = (s3 )
,
M
M
2

(4.673)

and form
1
uT (p, s3 ) =
2

q
T
p

q M T (s3 ).
p
M

(4.674)

Multiplying this by the charge-conjugation matrix C of (4.597) yields

1
C uT (p, s3 ) =
2

c p
qM T
c p
M

(s3 ),

(4.675)

333

4.14 Majorana Fields

with c = i 2 of Eq. (4.598). At this place we realize that due to the hermiticity
property of the Pauli matrices (8.159) and (4.459) one has,
= T ,

(4.676)

such that relation (4.633) implies the four-component relation


c c1 = c T c1 =
,

(4.677)

With this, the charge-conjugate spinor uc (p, s3 ) = C uT (p, s3 ) goes directly over into
the bispinor

p
1
M
q
v(p, s3 ) =
c (s3 ),

2
p
M

(4.678)

Thus, while the 4 2 -solutions u(p) of Eq. (4.663) are multiplied by the Pauli
spinors (s3 ) of Eq. (4.667), the 4 2 -solutions v(p) of Eq. (4.664) are multiplied
from the right by the spinors
c (s3 ) c(s3 ) = (s3 )(1)ss3 .

(4.679)

These are called


charge-conjugate Pauli spinors. Their explicit form is
c

(s3 ) =

0
1

1
0

(4.680)

This construction is necessary to make sure that the Pauli spinors v(p, s3 ) at
rest have the same transformation behavior under rotations as the spinors u(p, s3)
at rest.
Under a rotation, the original basis spinors (s3 ) are multiplied by a 22 rotation
matrix ei/2 . Then their spin indices are linearly recombined with each other by
the rotation matrix as follows:
R

i/2

(s3 )
(s3 ) = e

(s3 ) =

1/2
X

(s3 ) ei/2


s3 =1/2

s3 ,s3

(4.681)

The last step follows from the specific form (4.667) of the unit spinors. The same
mixing occurs in the charge-conjugate spinors
c (s3 )
c (s3 ) = c (s3 ) = cei
R

/2

(s3 )

(4.682)

Using (4.633), we see that


c1 ei/2 c = ei

/2

(4.683)

334

4 Relativistic Free Particles and Fields

such that the right-hand side becomes


i/2

i/2 c

c (s3 ) = e

(s3 ) =

1/2
X

c (s3 ) ei/2


s3 =1/2

s3 ,s3

(4.684)

such that c (s3 ) is indeed rotated precisely like (s3 ).


At rest, the 4 2 -matrices (4.663) and (4.664) becomes
0
0

1
u(p) =
2

1
v(p) =
2

0
0

(4.685)

such that the rotation properties of the bispinors u(p, s3 ) = u(p)(s3 ) and v(p, s3 ) =
v(p)c (s3 ) at rest are the same as those of (s3 ) and c (s3 ).
Explicitly, the bispinors v(p, s3 ) at rest become
1
v(0, 12 ) =
2

c ( 12 )
c ( 21 )

1 1
1

= , v(0, 21 ) =

0
2
2
1

c ( 12 )
c ( 21 )

1 0

,
=

1
2
0
(4.686)

to be compared with (4.670) for u(0, s3 ).


For the bispinors at rest (4.670) and (4.686), the Dirac equations in momentum
space (4.657) take the simple form
M( 0 1)u(0, s3) = 0,

M( 0 + 1)v(0, s3 ) = 0.

(4.687)

By applying the boost matrix (4.508), we find the alternative expression for the
bispinors with momentum p:
M + p/ 0

u(p, s3 ) = q
2M(p0 + M) 2

(s3 )
(s3 )

M + p/ 0

1
c (s3 )

, v(p, s3 ) = q
.
c
2M(p0 + M) 2 (s3 )

(4.688)

Since 0 is a simple off-diagonal unit matrix, we can replace it by 1 in the left and
right equation, respectively, and write just as well
1

u(p, s3 ) = q
2M(p0 + M) 2
M + p/

(s3 )
(s3 )

1
c (s3 )

, v(p, s3 ) = q
.
c
2M(p0 + M) 2 (s3 )
M p/

(4.689)

The two sets of bispinors u(p, s3) and v(p, s3 ) satisfy the orthonormality conditions (4.659). Using (4.668) and (4.678), we find
u (p, s3 )u(p, s3) =

p
1 T
(s3 )
,
2
M
r

p
sM
p

(s3 )

335

4.14 Majorana Fields

1 p p

(s3 )
+
2 M
M
p0
p0 T
(s3 )(s3 ) =
s ,s ,
=
M
M 3 3
= T (s3 )

(p, s3 )v(p, s3 )

(4.690)

rM
p

p
1 cT

=
(s3 )
,
2
M
M

u (p, s3 )v(p, s3 ) =

(p, s3 )u(p, s3 )

p0
s s ,
M 33

c
(s3 )

(4.691)

s
r
1 T
p
p

(s3 )
,
2
M
M

= 0,
= 0.

rM
p

c
(s3 )

(4.692)
(4.693)

The reason for the appearance of the negative momenta in the bispinors v(p, s3 )
is that the plane wave solutions fpc s3 (x) in (4.656) carry negative momenta, so
c
that states of a fixed momentum p are associated with fp s3 (x) and fp
s3 (x). The
momentum reversal in the conjugate wave functions goes along with the reversal of
the spin orientation in the charge-conjugate Pauli spinors Eq. (4.679). The physical
reason for these two reversals will be understood after field quantization in Section
7.3.3.
Inserting a matrix 0 between the bispinors in (4.690)(4.693) we may also derive
orthonormality relations between bispinors u(p, s3 ) and v(p, s3 ):
u(p, s3 )u(p, s3 )
v(p, s3 )v(p, s3 )
u(p, s3 )v(p, s3 )
v(p, s3 )u(p, s3 )

=
=
=
=

s3 ,s3 ,
s3 ,s3 ,
0,
0.

(4.694)

The two sets of spinors u(p, s3 ) and v(p, s3 ) span the spinor space at a fixed
momentum. This may be expressed by a completeness relation
Xh
s3

u(p, s3)u (p, s3 ) + v(p, s3 )v (p, s3 ) =

p0
.
M

(4.695)

336

4 Relativistic Free Particles and Fields

To prove this 4 4 -matrix equation in Dirac space, we derive the separate polarization sums16 for u- and v-spinors. These can be calculated directly from (4.668)
and (4.678) as follows:

X
s3

u(p, s3 )
u(p, s3 ) =
2
=

p
sM
p

v(p, s3 )
v (p, s3 ) =

s3

2
1
2

s
r
X
p

(s3 ) (s3 )
s
M
M
3

p
1
M
p

1
Mr
p
sM
p

M
p
1
M
p

1
M

p/ + M
2M

(4.696)

s
r
X
p

c (s3 )c (s3 )
s
M
M
3

p/ M
.
2M

(4.697)

Combining the two polarization sums for p and p, respectively, and multiplying
them by 0 proves their completeness.
By subtracting the two polarization sums from each other yields another completeness relation
X
s3

[u(p, s3 )
u(p, s3 ) v(p, s3 )
v (p, s3 )] = 1.

(4.698)

There is a minus sign in the second sum which corresponds to the minus sign in the
second orthogonality relation (4.694). This sign will be important in Section 7.8 to
prove a famous theorem on the relation between spin and statistics of fundamental
particles.
Polarization sums will frequently be needed later, in particular in the process of
field quantization. We introduce the sums
P (p)

u(p, s3 )u (p, s3 ),

s3

P (p)

v(p, s3 )v (p, s3 ),

(4.699)

s3

defined only for p0 on the mass shell, p0 = p . They satisfy the relation P (p) =
P (p). Similar polarization sums exist for plane-wave solutions for any spin. In
general, the polarization sums P (p) and P (p) of positive and negative energies of
momenta p and p, respectively, fulfill the relation
P (p) = P (p),
16

Since They may be called


semi-completeness relations.

(4.700)

337

4.14 Majorana Fields

where the upper sign holds for integer spin, the lower for half-integer spin. The
matrices P (p) and P (p) are projection matrices onto solutions of momenta p and
p with energies p0 = p and p0 = p , respectively. As such they satisfy
P (p)2 = P (p).

(4.701)

It is always possible to find a single covariant expression for P (p) defined for arbitrary
off-shell values of p0 , which at the on-shell values p0 = p reduces to th above
projections P (p) and P (p). In the Dirac case, where P (p) = (/
p M) 0 /2M, we
0
verify that (4.701) is true for p = p .
Spinors in Dirac Representation
Let us also write down the bispinors in the Dirac representation (4.544) of the matrices. The rest bispinors are solutions of equations (4.687), where
D0

1
0
0 1

Thus we have
u(0, 21 ) =

( 12 )
0

and
v(0, 21 ) =

0
c 1
(2)

1
0
0
0

u(0, 21 ) =

0
0
0
1

( 12 )
0

0
c
( 12 )

v(0, 12 ) =

0
1
0
0

=
2

(4.702)

0
0
1
0

(4.703)
The bispinors at finite momentum are obtained from these by applying the 4 4
boost matrix (4.508), yielding
M + p/ D

u(p, s3 ) = q
2M(p0 + M)

(s3 )
0

M p/ D

v(p, s3 ) = q
2M(p0 + M)

0
c (s3 )

(4.704)

Since D0 has a simple diagonal form with eigenvalues 1 for upper and lower spinor
components [see Eq. (4.556)], we have replaced D0 directly by its eigenvalues when
going from (4.508) to (4.704) [as we did from (4.688) to (4.689)]. More explicitly,
we can write
s

p
p +M
0

u(p, s3 ) =

2M
p

2M(p0 + M)

(s3 ),

v(p, s3 ) =

2M(p0 + M)
s
p0 + M
2M

c
(s3 ).(4.705)

338

4 Relativistic Free Particles and Fields

In this representation, the bispinors u(p, s3 ) of slowly moving particles have large
upper and small lower spinor components. The converse is true for the bispinors
v(p, s3 ). This is what makes the original Dirac spinors useful for discussing the
nonrelativistic limit of spin-1/2 particles, as observed before in the boost matrix
(4.550) and in x-space equations (4.573).
The Dirac spinors possess of course the same polarization sums (4.699) as in the
chiral case, if the appropriate Dirac matrices D are used on the right-hand side.
Helicity Spinors
Sometimes, the choice of the spinors (4.667) with the particle spins quantized along
the z-axis will not be the most convenient basis in spinor space. Instead of the
z-axis, one may choose any quantization direction, in particular, the direction of the
momentum of the particle. This amounts to multiplying the 4 2 -matrix solutions
(4.663) by a 2 2 rotation matrix from the right:
R(
p) ei

3 /2

ei

2 /2

(4.706)

where , are the spherical angles of the momentum p. In contrast to the notation in
(4.9), the rotation matrix carries now an argument indicating that the z-direction is
.17 Equivalently,
rotated into the momentum direction indicated by the unit vector p
we may choose in the bispinors (4.668), instead of (s3 ), a basis h (
p, ) with
= 1/2, defined by
h (
p, ) R(
p) (s3 ) ei

3 /2

ei

2 /2

().

(4.707)

Explicitly:

h (
p, 21 ) =

cos ei/2
2
i/2
sin e
2

h (
p, 12 ) =

sin ei/2
2

cos ei/2
2

(4.708)

The spinors (4.707) diagonalize the projection of the angular momen =


tum in the rest frame along the direction of motion of the particle p
(sin cos , sin sin , cos ), the so-called helicity:

1
1
=
=
p
h(
p) L p
2
2

cos
sin ei
sin ei cos

(4.709)

The eigenvalues are


h(p)(
p, ) = (
p, h),
17

1
= .
2

(4.710)

As before in this chapter, hats on vectors in this section denote unit vectors, not Schrodinger
operates.

339

4.14 Majorana Fields

The eigenstates h (
p, ) are called helicity spinors. The associated bispinors are, as
in (4.668) and (4.678),
p
sM
p

1
1

(
p, ), vh (p, ) = r M
uh (p, ) =

p
2
2

M
with the charge-conjugate helicity spinors

c
(
p, ), (4.711)

c (
p, ) = c (
p, ).

(4.712)

These diagonalize the 4 4 bispinor representation of the helicity:


1
=
H(
p) S p
2

1
=
p
2


0

.
p

(4.713)

This is a direct consequence of the fact that H(


p) commutes with the boost matrix
(4.508).
Alternatively, we can obtain the helicity bispinors (4.711) by boosting the
bispinors at rest (4.707), (4.708), first into the z-direction, and rotating them after direction
wards into the p

1
uh (
p, ) =

and

1
vh (
p, ) =

p0 0 |p| 3
(h)
R(
p)
M
s
p0 0 + |p| 3
R(
p)
(h)
M
s

p0 0 + |p| 3 c
(h)
R(
p)
M
s
p0 0 |p| 3 c
R(
p)
(h)
M

(4.714)

(4.715)

The equality with (4.711) follows from the transformation law (4.474), according to
which
p).
(4.716)
R(
p)|p| 3 = p R(

One of the important advantages of the helicity spinor is that it has a smooth
limit as the particle mass M tends to zero. Indeed, by expanding
p0 =

p2 + M 2 = |p| +

M2
+ ...
2|p|

(4.717)

we see that
p0 0 |p| 3 M 0 2

M
M
0 0
3
M 0
p + |p|
2

M
M

0 0
0 |p|
|p| 0
0 0

(4.718)

340

4 Relativistic Free Particles and Fields

Thus the massless helicity spinors uh (p, ) and vh (p, ) have only two nonzero components. We shall normalize them to
uh (p, )u(p, ) = 2p0 = 2|p|,

vh (p, )v(p, ) = 2p0 = 2|p|,

(4.719)

as opposed the normalization to p0 /M of the massive spinors (4.711). The explicit


form is

uh (p, 21 ) =

uh (p, ) =
1
2

vh (p, ) =

vh (p, 12 ) =

|p|

and

1
2

|p|

R(
p)

1
0

R(
p)

0
1

uR (p),

uL (p),

0
0

0
0

R(
p)

0
1

R(
p)

1
0

|p|

0
0

|p|

0
0

= vR (p)

= vL (p).

(4.720)

(4.721)

The helicity bispinors uh (p, ) and vh (p, ) are eigenstates of the chirality matrix
5 =

1 0
0 1

with the eigenvalue 2. By applying the 44 projection matrix


uh (p, ), forming
1 5
uh (p, ),
2

(4.722)
15
2

to the bispinors

(4.723)

we obtain a negative helicity state. Such projected bispinors are used for the description of neutrinos which only exist with negative helicity (left-handed neutrinos).
As we shall see later in Section 26, weak interactions involve also the orthogonally
projected bispinors
1 + 5
vh (p, )
2

(4.724)

which describe antineutrinos. These exist only with positive helicity (right-handed
antineutrinos).

341

4.15 Lorentz Transformation of Spinors

4.15

Lorentz Transformation of Spinors

Let us study the behavior of the bispinors u(p, s3 ) and v(p, s3 ) under Lorentz transformations. This will be most straight-forward in the chiral representation, where
we may focus attention upon the upper components only which we shall denote
by (p, s3 ). The properties of the lower components, to be denoted by (p, s3 ),
can be obtained by a simple change in the direction of the momentum. The upper
components can be written explicitly as

(p, s3) = B( )(s3 ) = e/2 (s3 ).

(4.725)

Applying a general Lorentz transformation D ( 2 ,0) () to this, the momentum p is


changed to some other vector p , which is the spatial part of the four-vector p .
The transformation can be done in three steps: First deboost the particle by applying
a boost opposite to the particles momentum which brings it to rest, with the fourmomentum p becoming pR = (M, 0). Second we perform a rotation, and third we
boost the particle to its final four-momentum p . Thus we can write the general
Lorentz transformations as

D ( 2 ,0) () = B( )W (p , , p)B1() e /2 ei /2 e/2 .

(4.726)

The rotation W (p, , p) in the middle is called a Wigner rotation. It is an element


of the little group of the massive particle acting only in its rest frame [see the earlier
short discussion on p. 35].
Let p be the momentum reached from the momentum p after a Lorentz transformation . Then the spinor (4.725) changes as follows:
(p, s3)
(p , s3 ) = D ( 2 ,0) ()(p, s3) = e /2 ei /2 (s3 )

1/2
X

= e /2

(s3 ) ei /2

s3 =1/2

1/2
X

s3 ,s3

(p , s3 )Ws3 ,s3 (p , , p),

(4.727)

s3 =1/2

where we have used the rotation property (4.681) of the spinors (s3 ), which amounts
here to

W (p , , p)(s3 ) =

1/2
X

(s3 )Ws3 ,s3 (p , , p).

(4.728)

s3 =1/2

By analogy, the spinor (p, s3 ) transforms like

(p, s3 )
(p , s3 ) = D (0, 2 ) ()(p, s3) =

1/2
X

s3 =1/2

(p , s3 )Ws3 ,s3 (p , , p),

(4.729)

342

4 Relativistic Free Particles and Fields

implying for the Dirac spinor u(p, s3 ) the transformation law

u(p, s3 )
u (p , s3 ) = D()u(p, s3) =

1/2
X

u(p , s3 )Ws3 ,s3 (p , , p).

(4.730)

s3 =1/2

The result can be expressed most compactly in terms of the 4 2 -matrix form
(4.664) for the bispinor solutions as

u(p)
u(p ) = D()u(p) = u(p )W (p , , p),

(4.731)

We are now prepared to understand the group-theoretic reason for the occurrence of the rotation matrix c = ei2 /2 in the charge conjugate bispinor v(p) of
Eq. (4.678). The 4 2 solutions v(p) of (4.664) transform in the same way as u(p)
of (4.663):

v(p)
v (p ) = D()v(p) = v(p )W (p, , p).

(4.732)

The behavior of the Dirac spinors v(p, s3 ) is found from this by multiplying this
equation from the right-hand side with c (s3 ) [recall (4.678)]. This yields

v(p, s3 )
v (p , s3 ) = D()v(p, s3) = v(p )W (p, , p)c (s3 ),

(4.733)

Now we use the fact that the 22 -Wigner rotation can be written as ei /2 , which
satisfies the relation (4.683), so that
c1 W (p , , p)c = W (p , , p),

(4.734)

to rewrite on the right-hand side


W (p, , p)c (s3 ) = cW (p , , p)(s3 ) = c (s3 )Ws3 s3 ,

(4.735)

and we obtain the transformation law for the bispinors v(p, s3 ):

v(p, s3 )
v (p , s3 ) = D()v(p, s3) =

1/2
X

s3 =1/2

v(p , s3 )Ws3 s3 (p , , p), (4.736)

Thus we find that under Lorentz transformations, the spin orientations of the
bispinors v(p, s3 ) are linearly recombined with each other by the complex-conjugate
Wigner rotations Ws s3 (p , , p). This is a consequence of the presence of the matrix
3
c in the 4 2 -matrices (4.664) for v(p), which has reversed canonical spin indices.
Had we used the bispinors v(p, s3 ) in the form (4.678), the same result would
have been obtained from the observation (4.682), that the spin indices of the
charge-conjugate Pauli spinors c (s3 ) are linearly recombined with each other by
the complex-conjugate rotation matrix.

343

4.16 Precession

The transformation properties (4.731) and (4.732) can be verified most easily in
an infinitesimal form for spinors at rest. If is an infinitesimal rotation R with
the 4 4 -matrix R = i L , the left-hand sides must be multiplied by
D() = R = 1 i /2, where i is given by (4.511). This produces the same
infinitesimal Wigner rotation on the right-hand sides 1 i /2. Thus we have the
relations

'

' 

'

ii u(0) = u(0)i i ,

ii v(0) = v(0)i i .

(4.737)

Using (4.663) and (4.664), these become explicitly


i

0
0

1
=
2

0
0

i ,

0
0

1
=
2

0
0

c(i i ) . (4.738)

Let us also write down the Wigner rotations for the helicity spinors (4.711).
Since they arise from the 4 2 -matrix solutions (4.663) by a multiplication from
the right with the 22 rotation matrix (4.706), a Lorentz transformation of uh (p, )
yields obviously
1/2
X

uh (p, )
uh (p , ) = D()uh(p, ) =

uh (p , )Wh (p , , p). (4.739)

=1/2

with the helicity form of the Wigner rotation


Wh (p , , p) = R1 (
p )W (p , , p)R(
p).

(4.740)

Similarly,

vh (p, )
vh (p , ) = D()vh (p, ) =

1/2
X

vh (p , )WH
, (p , , p). (4.741)

=1/2

with
W H (p , , p) = R1 (
p )W (p , , p)R (
p).

4.16

(4.742)

Precession

The properties of relativistic spinors under Lorentz transformations are crucial for a
phenomenon known in atomic physics as Thomas precession. The Thomas precession
is a direct consequence of what may be called Wigner precession.

4.16.1

Wigner Precession

Consider an electron moving around an atomic nucleus. In each time interval t, it


receives a small centripetal Lorentz boost changing its momentum. Let us see what
happens to the upper two components of the canonical bispinors u(p, s3 ) which are
explicitly
(4.743)
(p, s3) = B( )(s3 ) e/2 (s3 ).

344

4 Relativistic Free Particles and Fields

At an instant of time t, the electron moves with a certain velocity through


space. Its state can be described by the two-component spinor (p, s3 ) defined in
Eq. (4.725). As the atomic force acts on the electron, it is accelerated towards
the nucleus. Thus, after a small time interval dt, the electron will have a new
momentum and a spinor (p , s3 ) which can be obtained from the first by applying
a small Lorentz boost
B(d ) = ed/2 .
(4.744)

to the spinor (p, s3 ), which changes its momentum p to p . The resulting transformation is split into three factors, as in (4.726). The first is a pure boost in
the p direction, which brings the four-momentum p to its rest frame where it is
pR = (M, 0). The second factor is a rotation, and the third is a boost into the
final four-momentum p . In this process, the spin indices of the spinor (p, s3 ) are
linearly recombined with each other by a Wigner rotation according to Eq, (4.727).
Let us calculate this, taking advantage of the fact that d is very small. Then
differs very little from , say, = + d , where d is another small rapidity of
the order of d . To indicate the smallness of the associated rotation vector in the
Wigner rotation we shall denote it by d W . Its size is calculated from the equation

eid W /2 = e(+d)/2 ed/2 e/2 .

(4.745)

Before calculating d W exactly, let us estimate it for slowly moving particles where
d , d , and are all of the same order. Then we may expand both sides of (4.745)
up to the second order in all quantities as follows

 

 
 
ih
i
h
(4.746)
1 d /2 + (d)2 /4 1  /2 + 2 /4 .
In the product on the right-hand side, we set d d to cancel the first-order terms.
The second-order terms decompose into hermitian and antihermitian parts. Since
we are interested only in d , we must extract the antihermitian part. Using the
identity
(4.747)
a  b  = a b + i (a b) ,
1 id

W /2

1 + ( + d ) /2 + ( + d )2 /4

we find
1 id

W /2

  
  

 

  

1 i {[( + d ) d ] + [( + d ) ] [d ]} /4
1 + i (d ) /4.
(4.748)

so that we obtain the rotation vector for small :


d

W 12  d.

(4.749)

The spin matrix is rotated under a Wigner rotation as follows:

  = W 1W.

(4.750)

345

4.16 Precession

W /2 this yields


d = i [d W , ] = d W .

For an infinitesimal W 1 id

(4.751)

An accelerated point particle receives a small boost d in each small time interval
dt. In this time interval, the spin precesses at an angular velocity d W /dt. In the
limit dt 0, Eq. (4.749) implies an angular velocity of Wigner rotations

W 21   2c12 v v.
4.16.2

(4.752)

Thomas Precession

A relative of the Wigner precession is observable in atomic physics as a Thomas


precession. In an atom, the small additional boost acts on the electron moving with
momentum p in its rest frame. This implies that the small boost (4.744) is to be
replaced by
) = B( )B(d )B 1 ( ) = e/2 ed/2 e/2 ,
B(d
(4.753)

and (4.745) becomes


eid T /2 = e(+d)/2 e/2 ed/2

(4.754)

The small-velocity calculation (4.746) becomes now


1 id

T /2

1 + ( + d ) /2 + ( + d )2 /4

 
1 id R /2

 
 
ih
h
i
1  /2 + 2 /4 1 d /2 + (d)2 /4

leading with d d to

 
 

(4.755)

 

1 i {[( + d ) ] + [( + d ) d ] [ d ]} /4
1 + i ( d ) /4..
(4.756)

The resulting small Thomas rotation vector


d

T 12  d.

(4.757)

is the exactly the opposite of the Wigner rotation vector in Eq. (4.749). The same
thing is, of course, true for the rate of Thomas precession

T 12   2c12 v v.

(4.758)

For the spin vector S, which is the total angular momentum in the electrons
rest frame, this amounts to the equation of motion
dS
=
dt

T S 2c12 (v v) S .

(4.759)

346

4 Relativistic Free Particles and Fields

For finite , this equation will acquires relativistic correction factors and becomes

T = c12 + 1 v v.
2

(4.760)

The derivation of this expression is somewhat tedious and will therefore be given in
Appendix 4B.
The angular velocity of Wigner rotation has observable consequences in atomic
physics, where it is seen as a Thomas precession, which will be discussed in more
detail in Subsection 6.1.3. It is a purely kinematic effect, caused entirely by the
structure of the Lorentz group. Mathematically speaking, it is due to the fact that
pure Lorentz transformations do not form a subgroup of the full Lorentz group.
When performing pure Lorentz transformations one after another in such a way
that the final frame is again at rest with respect to the initial one, the result is
always a Wigner rotation.

4.16.3

Spin Four-Vector and Little Group

The working of the Wigner rotations in the little group found in Section 4.15 can
be understood independently of the particular spinors. For any massive elementary
of composite physical system we introduce a quantity called total spin four-vector .
+ S with the total
It is a combination of the total angular momentum J = L
momentum operator p into a vector
1
S = J p
2

(4.761)

where is the totally antisymmetric unit matrix with 0123 = 1.


For massive elementary particles of momentum p , the time and space components of the spin four-vector become explicitly

S0 = p J,

= p0 J
p K,

(4.762)

= (J23 , J31 , J12 ) and K = (J01 , J02 , J03 ). Studying particles at fixed mowhere J
menta, we have dropped operator hats on the momenta in (4.762), and the gen become differential operators in
erators of orbital angular momentum orbital L
momentum space and read explicitly [compare (8.85)]
i(p p )
L

(4.763)

with /p .
Using the commutation relations (4.285) and (4.97), the components of S in
(4.761) can be shown to satisfy the commutation rules
[S , S ] = i S p .

(4.764)

347

4.16 Precession

The proof makes use of the tensor identity (4.239) which, after taking advantage of

the antisymmetry of J in the indices and and the symmetry of p p in


and , leads to
1

J p p = J p p ,
2

(4.765)

and thus to the right-hand side of (4.764).


The same result can of course be derived without the lengthy identity (4.239) by
considering time and space components in (4.762) separately using the commutators
(4.76), (4.78), and (4.97). Then we find
i , p0 ] = 0,
[L
i , p0 ] = ipi ,
[M

i , pj ] = iijk pk ,
[L
i , pj ] = i ij p0 .
[M

(4.766)

For a free particle, p is independent of time, and so is S . By definition, the


spin four-vector S is orthogonal to the four-momentum:
S p = 0.

(4.767)

The physical significance of S becomes clear by going into the rest frame of a
massive particle where the system has no velocity, so that [recall (4.21) ]
pR = Mc(1, 0, 0, 0).

(4.768)

Then
SR0 = 0,

1
SRi = Mc ijk Jik Mc Ji .
2

(4.769)

Removing an overall factor Mc, we define the operators of Wigner rotations


i Si /Mc,
W
R

(4.770)

satisfying he commutation relations


i, W
j ] = iijk W
k.
[W

(4.771)

Thus the total spin four-vector has the property that its spatial components coincide
in the rest frame with the total angular momentum of the system. This is certainly
time independent due to angular momentum conservation.
Moreover, at zero momentum, the orbital part of Ji vanishes, so that only the
spin part S i of Ji survives, and we can drop the hats on top of SRi which indicate
the presence of differential operators. Then we obtain pure spin matrices for the
i
operators W
W i SRi /Mc,
(4.772)
satisfying the same commutation relations as the operators in (4.771). They will be
called spin three-vectors.

348

4 Relativistic Free Particles and Fields

The relation between the spin three-vector and the spin four-vector is obtained
by applying the pure Lorentz transformation matrix (4.18) to (4.769), yielding
2
=S
R + 1 (S
R v)v,
S
+ 1 c2

1
S0 = S v.
c

(4.773)

The inverse relations are


R = S

1 0
1 (S
v)v,
S v= S
+1c
+ 1 c2

SR0 = 0,

(4.774)

as can be verified with the help of the relation


v2
2 1
=
,
c2
2

(4.775)

which implies that /( 1)c2 = ( 1)/v 2 . Note that


1

S0 = S
v= S
R v.
c
c

(4.776)

For massless particles, the Wigner rotations have quite different properties than
for massive particles. In the special reference frame in which the massless particle
runs along the z-axis with a reference momentum pR = (1, 0, 0, 1)p introduced in
(4.25), these components become
S0 = p J3 ,

S1 = p (J 1 + K 2 ),

1 ),
S3 = p (J2 K

S3 = p J3 .

(4.777)

The three independent components


3 S0 = S3 ,
W

W 1 S1 ,

W 2 S1

(4.778)

1, W
2 ] = 0.
[W

(4.779)

satisfy the commutation relations


3, W
1 ] = iW
2,
[W

3, W
2 ] = iW
1,
[W

These generate a Euclidean group in a plane. Recall the definition of this group. In
D dimensions, it consists of D(D 1) generators Lij of the D-dimensional rotation
group:
[Lij , Lik ] = iLjk ,

(4.780)

and D generators of translation pi , which commute with each other and are vector
operators under rotations:
[pi , pj ] = 0,

[Lij , pk ] = i (ik pj jk pi ) .

(4.781)

These commutation rules can be obtained from those of the Lorentz group in Ddimensions [see Eqs. (4.76)(4.78)] by setting pi = Mi /c and letting c go to infinity.
This construction is called group contraction.

349

4.17 Weyl Spinor Calculus

1 and W
2 can be diagonalized simultaneously like
The commuting generators W
commuting momentum operators in a plane with arbitrary continuous eigenvalues
3 generates rotations in this plane
w 1 and w 2 , respectively. The third generator W
3
with discrete eigenvalues w = , where are azimuthal quantum numbers. which
can be equal to an integer of a half-integer number. In mirror-symmetric theories,
0 is given, according to
both signs have to occur. In an arbitrary reference frame, W
0 = p J/p. This shows that the eigenvalues measure
(4.762), by the operator W
the angular momentum around the momentum direction, i.e., the helicity of the
particle.
It turns out that in nature, all massless particles happen to follow a representation of the Wigner algebra which have only trivial eigenvalues w 1 = w 2 = 0. They
are characterizes completely by the helicity, which is unchanged under Wigner rotations, but merely receives a pure phase factor which multiplies the helicity spinors
or massless polarization vectors.
The occurrence of only such a subset of all possible zero-mass representations can
be understood by a limiting process such as the one performed in the derivation of
the massless spinors (4.720) and (4.721). We imagine for a moment that all massless
particles carry a small mass which we let go to zero. It can then be verified that the
limiting spinors change under a Lorentz transformation merely by a phase factor
associated with helicity = 1/2.:

D()uR(p) = uR (p)ei(p ,,p)/2,

D()vR (p) = uR (p)ei(p ,,p)/2 ,

D()uL (p) = uL (p)ei(p ,,p)/2,

D()uL (p) = uL (p)ei(p ,,p)/2 .

(4.782)

For the polarization vectors of electromagnetism (k, ) in (4.309), and the tensors
(k, ) of gravity in (4.418) we have, similarly:

(p, ) = (p, )ei(p ,,p),

(p, ) = (p, )ei(p ,,p). (4.783)

The reader is invited to derive this directly from the explicit expressions for these
objects.

4.17

Weyl Spinor Calculus

Weyl has devised a simple calculus for constructing spinor invariants of the Lorentz
group. It is very similar to the tensor calculus. The spinor in the upper two components of the Dirac field, which transforms under boosts via the matrix e/2 ,
was previously denoted by , the spinor in the lower component, transforming via

e/2 , by . Complex conjugation brings into a spinor ( ) which transforms

via e /2 . Such a spinor is given a lower dotted index, i.e., we write
( ) .

(4.784)

Similarly, we define ( ) which transforms via e /2 as

( ) .

(4.785)

350

4 Relativistic Free Particles and Fields

From the earlier discussion in Section 4.12 we know that

( ) ,

( )

(4.786)

are Lorentz invariants. With the above notation, these can be viewed as
,

(4.787)

Thus the invariants arise by simple contractions of equal upper and lower indices.
A further invariant can be constructed from two spinors which both have lower
indices and , namely
c ,
(4.788)
with the 2 2 -charge conjugation matrix c = i 2 of (4.598). Writing (4.788) in
matrix notation as T c , it goes under rotations over into
T ei'

T /2

cei'/2 .

(4.789)

Since T = [see (8.159)] and c = c from (4.633), this is obviously invariant. A similar manipulation shows invariance under boosts. Thus the matrix c
constitutes an antisymmetric (or symplectic) metric in spinor space. Accordingly,
we define
c .
(4.790)
Then the invariant (4.788) arises by a contraction of equal upper and lower indices,
just as in the notation in Minkowski space:
c = .

(4.791)

Similarly, we can form an invariant from two -spinors

(c1 ) ,

(4.792)

(c1 ) ,

(4.793)

defining

which makes the contraction a Lorentz invariant.

The Lorentz transformation matrices associated with and are D ( 2 ,0) ()


1
1
1

and D (0, 2 ) (), respectively. They carry Weyl indices D ( 2 ,0) () and D (0, 2 ) () .
It is possible to combine Weyl spinors to vectors rather than scalars with the
help of the -matrices (8.157) and of
. They may be thought of as carrying Weyl
labels
( ) ,
(4.794)
and

(
) .

(4.795)

351

4.18 Higher Spin Representations

Then the indices show directly which spinors are required to form vectors:


= (
) ,

= ( ) .

(4.796)

The vector nature of these combinations is proved by rewriting the transformation


law (4.505) out in 2 2 -form:
1

D ( 2 ,0) ()1 D (0, 2 ) () = ,


1

D (0, 2 ) ()1
D ( 2 ,0) () =
.

(4.797)

Written with Weyl indices, this reads


1

[D ( 2 ,0) ()1 ] ( ) D (0, 2 ) () = ( ) ,


1

) D ( 2 ,0) ()
[D (0, 2 ) ()1 ] (

(4.798)

= 0.

(4.799)

= (
)
.

In Weyls calculus, the Dirac equation reads

(i/
M)(x) =

4.18

M
(i )

(i
)
M

Higher Spin Representations

Given the fundamental spin-1/2 field it is very simple to generalize the transformation matrices to higher spins. A system with two spin-1/2 particles can have
spin 1 or 0. Similarly, n spin-1/2 particles can couple to spin n/2, n 1/2, . . . down
to 1/2 or 0. Thus in order to build an arbitrary spin s, all we have to do is put
2s spin-1/2 representations together in an appropriate fashion. The problem is
completely analogous to the previous extension in Section 2.5 of the one-particle
Schrodinger equation to an arbitrary nparticle equation. Thus we shall construct
representations of arbitrary spin by a second quantization of spin, and further of
the generators of the entire Lorentz group.

4.18.1

Rotations

If the particle is at rest, spin is defined by the rotation subgroup. The 22 hermitian
generators
1
L=
(4.800)
2
may be considered as the analog of the single-particle Schrodinger operator
h2

x 2 + V1 (x; t) in Eq. (2.92), or as the matrix Mi in the commutation rules


2M
(2.98). According to (2.100), the second quantized version of L reads

i = 1 a
L
i a.
2

(4.801)

352

4 Relativistic Free Particles and Fields

where a
and a
with = 1, 2 are two bosonic creation and annihilation operators.
As proved in general in (2.101), these operators satisfy the same commutation rules
as the Pauli matrices i :
i, L
j ] = iijk L
k.
[L
The states

| 21 , 21 i a
1 |0i,

(4.802)

| 12 , 12 i a2 |0i

(4.803)

may be identified as the two basis states of the fundamental spin-1/2 representation.
i is, as in the general Eq. (2.102):
The effect of the three operators L
1
= a
i ,
2
i
h
1
i =
a
, L
i a
.
2
h

i, a
L

(4.804)

On these states, the second-quantized operators (4.801) have eigenvalues


s3 i = s(s + 1)|s, s3 i,
L|s,

3 |s, s3i = s3 |s, s3i.


L

(4.805)

In the present restriction to the rotation group we shall use only lower indices 1 and
2 rather than Weyl indices of the previous section.
As in Section 2.5, we may now compose all higher representations of the rotation
group by combining many of these fundamental representations, and forming states
such as
2s
Y

(ai )|0i.

(4.806)

i=1

It is easy to see that linear combinations of such states with a fixed number of a
form an invariant representation space of the rotation group. The reason is that the
i commute with the particle number operator
three generators L
N =a
a
.

(4.807)

2
In fact, a little algebra shows that the Casimir operator of the rotation group L
which characterizes an irreducible representation is equal to
1 1
2 = 1 (a i a
L
N +1 ,
)2 = N
4
2
2


(4.808)

showing explicitly that under rotations, states built from a fixed number 2s of spin
creation operators a always maintains the same number 2s, and that the eigenvalue
of the Casimir operator in the space of 2s particles is s(s + 1). Thus the space is
indeed invariant under rotations.
But it is also an irreducible representation of spin s. To see this take the complete
set of basis states in the space
|n1 , n2 ) =

1
(a1 )n1 (a2 )n2 |0i,
n1 !n2 !

(4.809)

353

4.18 Higher Spin Representations

which diagonalize the occupation numbers of the two operators a1 and a2 , and which
are normalized to unity. Applying to these the operator
3
3 = a a = 1 (a1 a
L
1 a2 a2 ),
2
2

(4.810)

we see that it measures the difference in the number of spin up and spin down
particles created by a1 and a1 , respectively:
1
L3 |n1 , n2 ) = (n1 n2 )|n1 , n2 ) m|n1 , n2 ).
2

(4.811)

Thus the state has the azimuthal (magnetic) quantum number m = n1 n2 . The
operators
+ L
1 + iL
2, L
L
1 iL
2,
L
(4.812)
on the other hand, have the form
+ = a
L

0 1
0 0

a
=a
1 a
2

= a
L

0 0
1 0

a
=a
2 a
1 .

(4.813)

3 they satisfy the commutation rules


Among each other and with L
, L
+ ] = 2L
3
[L
3, L
+] = L
+
[L
3, L
] = L
.
[L

(4.814)

They remove a spin down while adding a spin up and vice versa. Their matrix
elements are
+ |n1 , n2 ) =
L

|n1 , n2 ) =
L

(n1 + 1)n2 |n1 + 1, n2 1),

n1 (n2 + 1)|n1 1, n2 + 1).

(4.815)

It is now obvious that starting from an arbitrary state, say |n, 0) with no number
3 , we can reach every other state
of spin-down particles and eigenvalue n/2 of L
. The process ends at
|n k, k) with k = 1, . . . , n by repeated application of L

|0, n), where L |n, 0) = 0. This proves the irreducibility of the representation.
3 is referred to as the value s of angular
Conventionally, the highest eigenvalue of L
momentum:
s = n/2.
(4.816)
2 as
It appears in the eigenvalue (4.808) of L
2 |n k, k) = s(s + 1)|n k, k).
L

(4.817)

354

4 Relativistic Free Particles and Fields

+ and L
upon the states |s, mi.
Figure 4.3 Effect of raising and lowering operators L

Hence the states may be reexpressed in terms of the numbers s and m as


1
|s, mi = |s + m, s m) = q
(a1 )s+m (a2 )sm |0i,
(s m)!(s + m)!

(4.818)

and the matrix elements (4.811) and (4.815) read


3 |s, mi = m|s, mi,
L
q
+ |s, mi =
L
(s m)(s + m + 1)|s, m + 1i,
|s, mi =
L

(s + m)(s m + 1)|s, m 1i.

(4.819)
(4.820)
(4.821)

+ and L

For smaller spin values the effect of the raising and lowering operators L
upon the states |s, mi is illustrated in Fig. 4.3.
Note that we could have defined the states (4.818) with arbitrary phase factors
i1
e and ei2 accompanying a1 and a2 . Then the application of L+ and L would
produce phase factors ei(1 2 ) and ei(1 2 ) in (4.820) and (4.821), respectively. It
is easy to verify that these phases drop out in the commutation rules (4.814) of the
rotation group, and thus in (4.802).
The choice of a positive-square root in (4.820) and (4.821) without such extra
phase factors is known as the Condon-Shortley phase convention.18

4.18.2

Extension to Lorentz Group

It is now quite simple to extend this construction of spin representations of the


rotations to the entire Lorentz group. For this we deduce from the commutation
rules between the generators Li and Mi in Eq. (4.76)(4.78) that the combinations
1 = (L
+ iM)/2,

J
2 = (L
iM)/2

J
18

(4.822)

E.U. Condon and G.H. Shortley, Theory of Atomic Spectra, Cambridge University Press, New
York, 1935.

355

4.18 Higher Spin Representations

have the commutation rules


h

Ji1 , Jj2
Ji1 , Jj1
Ji2 , Jj2

= 0,

= iijk Jk2 .

= iijk Jk1 ,

(4.823)

Therefore they generate two independent sets of rotations. Extending the previous
construction, we now form the second-quantized operators




(1) = a a
J
,
2
(2) = b b,
J
2
and obtain

(4.824)




= J
(1) + J
(2) = a
L
a
+ b b,
2  2

1
1 (1) (2)

a a
(J J ) =
b b .
M =
i
i
2
2

(4.825)

(4.826)

These operators correspond to forming the second-quantized operators


= 1 (a , b )
L
2

= 1 (a , b ) i
M
2

(4.827)

(4.828)

with the 4 4 -representation matrices (4.511) and (4.513) of the Lorentz group.
For a clearer display of the Lorentz transformation properties one may take
advantage of the Weyl notation. From (4.825) and (4.826) we see that the operators
1
1
a and b transform according to the fundamental representations D ( 2 ,0) and D (0, 2 ) ,
since the 2 2 -matrices between them are for rotations /2, i /2 and for Lorentz
boosts /2, i /2,respectively. Thus we may write them in the Weyl notation as a

and b . By applying the same arguments as for the rotation group to the operators
(1) and J
(2) , we can now easily see that the set of states
J

1
a
a
b
b
(a1 )n1 (a2 )n2 (b 1 )n1 (b 2 )n2 |0i
|na1 , na2 , nb1 , nb2 i = q
na1 !na2 !nb1 !nb2 !

(4.829)

for fixed numbers na = na1 + na2 and nb = nb1 + nb1 are irreducible representation
spaces of the whole Lorentz group. They are denoted by
(s1 , s2 )

na nb
.
,
2 2
!

(4.830)

356

4 Relativistic Free Particles and Fields

(1) and J
(2) .
The states (4.829) live in the direct-product space of the two operators J
They may be relabeled by the quantum numbers of the two rotation subgroups as
in (4.818):
1
|s1 , m1 ; s2 , m2 i = q
(s1 m1 )!(s1 + m1 )!(s2 m2 )!(s2 + m2 )!

(a1 )s1 +m1 (a2 )ns1 m1 (b 1 )s2 +m2 (b 2 )s2 m2 |0i.

(4.831)

is the direct sum of those of


Since by (4.825) the operator of angular momentum L
the two rotation subgroups, an irreducible representation D (s1 ,s2 ) () of the Lorentz
group contains different irreducible representations of the rotation subgroup gen They are obtained from the rules of addition of angular momenta.
erated by L.
2 and the third component L
3 can be diagonalized with eigenvalues
The operator L
s(s + 1) and m by forming linear combination
|s, mi =

m1 ,m2

|s1 , m1 ; s2 , m2 ihs1 , m1 ; s2 , m2 |s, mi

(4.832)

where hs1 , m1 ; s2 , m2 |s, mi are Clebsch-Gordan coefficients [5]. Their calculation and
properties are recalled in Appendix 4C. The values of total angular momentum s
occurring in the decomposition (4.832) are
|s1 s2 | s s1 + s2 .

(4.833)

The Clebsch-Gordan coefficients are orthogonal and complete, so that (4.832) can
be inverted to
|s1 , m1 ; s2 , m2 i =

4.18.3

X
s,m

|s, mihs, m|s1, m1 ; s2 , m2 i.

(4.834)

Finite Representation Matrices

To complete this discussion let us calculate the finite representation matrices. Due
to the decompositions (4.825) and (4.826), we only need those of the rotation group.
This is done as follows.
Rotation Group
We first observe that every 3 3 -rotation matrix R' () in Eq. (4.9) can be decomposed into Euler angles
R' () = ei'L = eiL3 eiL2 eiL3 R(, , ),

(4.835)

and so can the general rotation operator

, ).
ei'L = eiL3 ei L2 ei L3 R(,

(4.836)

357

4.18 Higher Spin Representations

3 is diagonal on the states |jmi, the finite rotation ei'L acts on the states
Since L
|jmi as
j
X

, )|jmi =
R(,

m =j

j
X

m =j

|jm iei(m +m) hjm |ei L2 |jmi


j
|jm iDm
,m (, , ),

(4.837)

so that the only nontrivial matrix elements are

djm m () = hjm |ei L2 |jmi.

(4.838)

For a single creation operator a , we have from (4.804):


1
i
[a1 a
,
2 a
2 a
1 , a
1 ] = a
2i
2 2
i
1
2, a
[a1 a
,
2 a
2 a
1 , a
2 ] = a
[L
2 ] =
2i
2 1
2, a
[L
1 ] =

(4.839)

and therefore

+a
2 sin ,
2
2

= a1 sin + a
2 cos .
2
2

ei L2 a
1 ei L2 = a1 cos
ei L2 a
2 ei L2

(4.840)

Of course, this is just the statement that a


|0i transforms according to the spin1/2-representation of the rotation group [recall (4.453)]

cos 2 sin 2

ei L2 a ei L2 = a

=

cos 2

sin 2

i2 /2

(4.841)

But then an arbitrary state (4.818) goes over into


1
eL2 |jmi = q
(j m)!(j + m)!
a
1

cos + a2 sin
2
2

!j+m

a1

sin + a
2 cos
2
2

!jm

|0i. (4.842)

We now expand each factor in a binomial series, reorder it, and write
X
m

1
q

(j m )!(j + m )!

(a1 )j+m (a2 )jm |0idjm m (),

(4.843)

358

4 Relativistic Free Particles and Fields

with
djm m ()

v
u
u (j
t

+ m )!(j m )! X
(j + m)!(j m)! k=0

(1)

jmk

cos
2

j+m
j m k

!2k+m +m

sin
2

jm
k

!2j2km m

(4.844)

The sum can be expressed in terms of hypergeometric functions


F (a, b, c; z) 1 +

a(a + 1) b(b + 1) z 2
ab
z+
+ ...
c
c(c + 1)
2!

(4.845)

as
v
u

(1)m m u
t (j m)!(j + m )! cos
djm m () =
(m m)! (j + m)!(j m )!
2

!m +m

sin
2

!mm

.
F j + m , j + m + 1; m m + 1; sin
2

(4.846)

This formula is best used only for m m where the hypergeometric function is
regular at the origin. For m < m one may simply use the property
djm m () = (1)m m djmm () = (1)m m djm m ().

(4.847)

Additional useful relations are


djm m ( ) = (1)jm djm m ()

(4.848)

djm m () = (1)j+m m ,m .

(4.849)

and

The hypergeometric functions can also be expressed as a function of Jacobi


polynomials
!

F j + m , j + m + 1; m m + 1; sin
2

(j m )!
(m m,m +m)
=
(cos ).
(m m)!Pjm
(j m)!

(4.850)

The matrix elements dl00 () coincide with the Legendre polynomials,


dl00 () = Pl (cos ),

(4.851)

and the matrix elements dlm0 ()ei are proportional to the spherical harmonics
Ylm (, ):
s
4
l
i
m l
i
Ylm (, ).
(4.852)
dm0 ()e = (1) d0m ()e =
2l + 1

359

4.18 Higher Spin Representations

For j =

1
2

we reobtain the spinor representation of the rotation group


1/2
dm m ()

cos /2 sin /2
sin /2
cos /2

= ei

2 /2

while for j = 1 we find the vector representation


1

1
(1
+
cos
)

sin
/
(1

cos
)
2
2


2
d1m m () = sin / 2
cos

sin
/
2
1
1
(1 cos )
sin / 2 2 (1 + cos )
2

(4.853)

(4.854)

The indices have the order +1/2, 1/2 and +1, 0, 1, respectively.
The representation functions (4.837) of all rotations
j
i(m+m ) j
dm,m ()
Dm,m
(, , ) = e

(4.855)

have the following orthonormality properties:


2j + 1
8 2

Z2
0

Z
0

Z2

j1
j2
d Dm
(, , )Dm ,m (, , ) = j1 ,j2 m1 ,m2 m ,m . (4.856)
2
1 ,m
1
2
1

At fixed m1 , m2 , one has relations


2
Z
2j + 1 Z
j1
j2
d d Dm
(, , 0)Dm ,m (, , 0) = j1 ,j2 m1 ,m2 .
1 ,m1
2
2
4
0

(4.857)

j
The representation matrices Dm,m
(, , ) with j = 1 are related to the original
3 3 -rotations R(, , ) of Eq. (4.835) as by a similarity transformation. It
is the same transformation which relates the three spherical components () in
Eq. (4.311) to the unit vectors (4.315). The eigenvectors (4.311) supply us with the
matrix elements of the desired similarity transformation. Setting

"

hi|1, mi i (m),

(4.858)

we can write, using the 3 3 matrices (4.54),


(L3 )jk =

1
X

hj|1, mimh1, m|ki,

m=1

(L )jk = hj|1, 1i 2h1, 0|ki + hj|1, 0i 2h1, 1|ki.

(4.859)

In Diracs bracket notation, the original 3 3-matrices R(, , ) in (4.835) may be

considered as matrix elements of the general rotation operator R(,


, ) in (4.836)
between the basis states |ii:
, )|ji.
Rij (, , ) = hi|R(,

(4.860)

360

4 Relativistic Free Particles and Fields

From the manipulation rules of Dirac brackets it is then obvious that the matrix
elements transform under finite rotations as
3
X

Rij hj|1, mi =

j=1

3
X

, )|jihj|1, mi
hi|R(,

j=1

3
1
X
X

hi|1, m ih1, m|R(,


, )|jihj|1, mi

(4.861)

j=1 m =1

1
X

hi|1, m ih1, m|R(,


, )|1, mi =

m =1

1
X

1
hi|1, m iDm
,m (R),

m =1

which may also be written in matrix form as as19

R (m) =

1
X

(m)Dm1 ,m(R).

m =1

(4.862)

In Eq. (4.110) we stated the transformation law of a vector operator [see also
Eq. (2.112)]:
i , vk ] = vj (Li )jk ,
[L
(4.863)
With the help of the above similarity transformation, we find the spherical components of the vector operator vi :
v(m)
or, explicitly,

3
X
i=1

vi hi|1, mi,

1
v1 i
v2 ),
v(1) (
2

(4.864)

v(0) v3 .

(4.865)

For these components, the commutation rules (4.863) become


i , v(m)] =
[L

1
X

m =1

i |1, mi.
v(m )h1, m |L

(4.866)

They may be generalized to an arbitrary spherical tensor operator v(j, m) of spin j:


i , v(j, m)] =
[L

1
X

m =1

i |j, mi.
v(j, m )hj, m |L

(4.867)

3 and L
, these commutation relations become
For L
q

3 , v(j, m)] = v(j, m) m, [L


, v(j, m)] = v(j, m1) (j m)(j m + 1). (4.868)
[L
19

Note that the spinor transformation laws (4.739) and (4.741) are a generalization of this
relation.

361

4.18 Higher Spin Representations

They are in one-to-one correspondence with the relations (4.819), (4.820) and (4.821)
for the states |j, mi.
For finite rotations, they give rise to the transformation behavior
v(j, m)R
1 = v(j, m )D j (R).
R
mm

(4.869)

The use of defining such spherical tensor operators lies in the fact that all their
matrix elements are related to each other by Clebsch-Gordan coefficients (4.832) via
the so-called Wigner-Eckart theorem. Applying v(j, m) to a state |j , m i, we obtain
a state v(j, m)|j , m i, which transforms by a direct product of the representation
matrices (4.837) and (4.869) like a state |j, m; j , m i. Its irreducible contents can be
obtained with the help of the Clebsch-Gordan series (4.832). If we therefore expand
v(j, m)|j , m i =

j ,m

|j , m ihj , m |
v (j, m)|j , m i,

(4.870)

the matrix elements hj , m |


v (j, m)|j , m i must be proportional to the associated
Clebsch-Gordan coefficients:
hj , m |
v(j, m)|j , m i = hj ||v(j)||j ihj , m |j, m; j , m i.

(4.871)

The proportionality constants hj ||v(j)||j i are independent of the azimuthal quantum numbers m, m , m , and are called the reduced matrix elements of the spherical
tensor operator v(j, m). They vanish if j does not satisfy the vector coupling condition |j j | j j + j .
2
For j = 1/2, the matrix djm m () is equal to ei /2 = i 2 . It is therefore the
spin-j representation of the matrix c of (4.598), and will therefore be denoted by
c(j) djm ,m (c) = (1)j+m m ,m .

(4.872)

When applied as a similarity transformation to the representation matrix


j
Dm
m (, , ), we find a spin-j generalization of the important 22 -relation (4.683):

[c(j) ]1 D j (, , )c(j) = D j (, , ).

(4.873)

The matrix c(j) gives rise to an invariant bilinear form for any pair of spherical tensor
operators v(j, m) and v (j, m):

(
v, v )

j
X

m,m =j

(j)
v(j, m)cmm v (j, m )

j
X

m=j

(1)j+m v(j, m)
v (j, m).

(4.874)

This product remains invariant under rotations, since


D jT (, , )c(j)D j (, , ) = c(j) .

(4.875)

The invariance of (4.874) is a generalization to spin-j operators of the Weyl invariance of the spinor product c of (4.788).

362

4 Relativistic Free Particles and Fields

For j = 1, the invariant product (4.874) is equivalent to the ordinary scalar


product. This is seen by replacing on the right-hand side of (4.874) the spherical
components v(1, m) by the cartesian ones according to (4.865), yielding
(
v , v ) = ij vi vj .

(4.876)

Also for j = 1, the spherical relation (4.875) is equivalent to the invariance of


the Kronecker symbol ij under rotations in the 3 3 defining representation:
Ri1 i1 Ri2 i2 i1 i2 = i1 i2 .

(4.877)

The invariance of scalar products (4.874) formed with c(j) can be used to extend
the Weyl calculus to spin-j objects as follows: The spherical tensor operator v(j, m)
is written as v( mj ), and we introduce the contravariant spherical tensor operator
v( mj )

j
X

(j)

j
cm,m v( mj ) = (1)j+m v( m
).

(4.878)

m =j

Then the invariant form (4.874) can simply be written as


(
v , v ) = v( mj ) v ( mj ),

(4.879)

with the convention that pairs of upper and lower indices m are being summed.
The relation between the axis-angle and the Euler representation of the rotations
on the two sides of (4.835) is easily found using the explicit 2 2 -representation
(4.466) in the axis-angle form, and in the Euler form
ei

3 /2

ei /2 ei /2
!



3
2
3
= cos i sin
cos i sin
cos i sin
2
2
2
2
2
2
+
1

+ sin
i
= cos cos
2
2
2 2
2

+ 3

i sin sin
i .
(4.880)
sin cos
2
2
2
2


Comparing the coefficients gives

+
= cos cos
,
2
2
2



sin
sin .
= sin
, cos
, sin
2
2
2
2
2
cos

'

(4.881)

More details on the rotation group can be found in the textbook Ref. [5].

363

4.19 Higher Spin Fields

Lorentz Group
To extend these results to the Lorentz group we make use of the fact that due to
the decompositions (4.825) and (4.826), pure rotations can be decomposed as
ei'L = ei'J ei'J ,
(1)

(2)

(4.882)

where J(1) and J(2) are the matrices (4.822). The pure Lorentz transformations are

(1)
(2)
eiM = eJ eJ ,

(4.883)

(1)
(2)
with eJ eJ having again matrix elements of rotations, as calculated above.
Thus, given the parameters
and of the Lorentz transformation in question,
we merely have to find the corresponding Euler angles and take the corresponding
rotation matrices from (4.837) and (4.846).
Note that for pure Lorentz transformations the rotation parameters are imaginary so that the trigonometric functions become hyperbolic. For pure Lorentz
transformations with imaginary angles, the relation between the axis-angle and the
Euler representations corresponding to the two sides of (4.835) is then given by relations like (4.881), but with cosine and sine functions continued to the corresponding
hyperbolic forms.

'

4.19

Higher Spin Fields

The construction of invariant actions can be generalized to fields of arbitrary spin.


If we restrict ourselves to those representations which contain only one spin, the
situation is very similar to the spin-1/2 case: There are two spinor fields of the Weyl

type and , transforming according to the D (s,0) and D (0,s) -representations,


respectively

(x)
(x ) = D (s,0) () (x),

(x)
(x ) = D (0,s) () (x).

(4.884)

Now, according to the last section, the matrices D (s,0) () are just the symmetrized
1
direct products of 2s representations D ( 2 ,0) (). They satisfy the same relation as
1
D ( 2 ,0) itself:
D (s,0) () = D (0,s) ()1 .
(4.885)
Hence
,

(4.886)

are the only Lorentz-invariant invariant bilinear combinations of the spinor fields.
What about invariants involving derivatives? For this we recall that in the spin- 12
case
i
, i
(4.887)

364

4 Relativistic Free Particles and Fields

were invariant due to the property (4.479). The invariance remains true for a product
of 2s factors whose right and left indices are symmetrized. Therefore
(i2s
1 . . .
2s 1 2s ) (
i){2s}

(4.888)

(i2s 1 . . . 2s 1 2s ) (i){2s}

(4.889)

and

are invariants, where the curly brackets indicate the symmetrization of the indices.
We may therefore write the action as
A=

d4 x (
i){2s} + (i){2s} M12s M22s .

(4.890)

In the absence of the mass terms, each of the derivative pieces gives by itself an
invariant action which maximally violates parity. This fact will be needed to explain
the maximal parity violation in weak interactions briefly discussed on p. 314. The
equations of motion (4.484) and (4.485) become
(
i){2s} (x) = M12s (x),
(i){2s} (x) = M22s (x).

(4.891)
(4.892)

They can be inserted into each other to give


(
i){2s} (i){2s}

(x)
(x)

= M12s M22s

(x)
(x)

(4.893)

The left-hand side contains a product of two symmetrized products. Since each 2s
factors are symmetric under simultaneous exchange of left and right indices, we can
omit the symmetrization in the contracted indices, and use in each of them relation
(4.487) to derive
[(
i)(i)]{2s} = ( 2 1){2s} = ( 2 )2s (1){2s} ,

(4.894)

with (1){2s} being the unit matrix in the symmetrized subspace. In momentum
space, this amounts to the mass shell relation
(p2 )2s = M12s M22s = (M 2 )2s .

(4.895)

As in the spin-1/2 case, space inversion changes and the representation


matrices
D (s,0) () D (0,s) ().
(4.896)

If one wants to have a representation space of the Lorentz group including space
inversions, one must combine the two spinors (x) and (x) into a bispinor with
2 (2s + 1) components
!
(x)
(x) =
.
(4.897)
(x)

365

4.19 Higher Spin Fields

On this space, parity is represented as in (4.527) by


P

(x)
P (x) = D(P )(
x),

(4.898)

with a representation matrix which looks like (4.528), but contains now four blocks
of (2s + 1) (2s + 1) -matrices:
D(P ) = P

0 1
1 0

= P 0 .

(4.899)

It is obvious that this matrix changes the generators L and M of the Lorentz group
in the spin-s representation as in Eqs. (4.526).
Invariance under space inversion requires
M1 = M2 = M 2s ,

(4.900)

and thus the presence of both derivative terms in (4.890). There are now generalized
-matrices with
(i){2s} =

0
(i){2s}
{2s}
(
i)
0

(4.901)

and we can write down a parity-invariant action for the bispinors (x) as
A=

d4 x (x)
(i){2s} M 2s (x),
i

(4.902)

where

(x)

The field equation is

= (x)( 0 ){2s} .

(i){2s} M 2s (x) = 0.

4.19.1

(4.903)

(4.904)

Plane-Wave Solutions

One can easily write down plane wave solutions of the spin-s wave equation (4.904):
eipx
fp s3 (x) u(p, s3) q
,
V p0 /M

eipx
fpc s3 (x) v(p, s3 ) q
,
V p0 /M

(4.905)

where u(p, s3 ) and v(p, s3 ) are the positive- and negative-energy solutions of momentum p and p, respectively. satisfying the generalized Dirac equations in momentum
space
(/
p {2s} M)u(p, s3 ) = 0,
(/
p {2s} + M)u(p, s3 ) = 0.
(4.906)

366

4 Relativistic Free Particles and Fields

The second can be obtained from the first via a relation like (4.672):
v(p, s3 ) = C uT (p, s3 ).

(4.907)

Here C is the charge conjugation matrix for arbitrary spin


C=

{2s} 0

c{2s} 0
0
c{2s}

(4.908)

The matrices c{2s} 0 are equivalent to the matrices c(s) introduced in Eq. (4.872).
They have the important property that
c{2s} = (1)2s .

(4.909)

The rest spinors {2s} (s3 ) have symmetrized labels 1/2 and 1/2. These are uniquely
specified by the number n1 of up-spins and n2 of down-spins, which are the labels of
the basis vectors |n1 , n2 ) in Eq. (4.809). Thus we may write {2s} (s3 ) more explicitly
as n{2s}
(s3 ). The label s3 specifies the eigenvalues of the third component of angular
1 ,n2
momentum, and corresponds to the label m of the basis vectors (4.818). Hence
n{2s}
(s3 ) = s3 ,(n1 n2 )/2 ,
1 ,n2

n1 + n2 = 2s.

(4.910)

These spinors satisfy the obvious completeness relation


X
s3

{2s}

n{2s}
(s3 )n ,n (s3 ) = n1 ,n2
1 ,n2
1

(4.911)

Using this, we find for the spinors u{2s} (p, s3 ) a polarization sum [compare (4.699)]
P (p)

u{2s} (p, s3 )u{2s} (p, s3 ) =

s3

M + p {2s}
2M

!{2s}

(4.912)

It is a straightforward generalization of the Dirac case. The polarization sums for


the spinors v {2s} (p, s3 ) can be calculated similarly.
From the spinors w 2s (p, s3 ) we form the mirror-reflected spinors w 2s (p, s3 )

{2s}

(
p, s3 ) =

{2s}
Bp (){2s} (s3 )

p
M

{2s}

(4.913)

and after multiplication with the generalized charge-conjugation matrix c2s we combining both spinors to bispinors u2s (p, s3 ) and v 2s (p, s3 ) of particle and antiparticles
of spin s.

367

4.20 Vector Field as Higher Spin Field

4.20

Vector Field as Higher Spin Field

Some remarks are in order concerning the fields transforming according to the representation D (s1 ,s2 ) () with both s1 , s2 6= 0 which were omitted in the above construction. The most prominent among them is the spinor field with s1 = s2 = 12

which we denote by (x). It is this field which is equivalent to a vector field


A (x), which was discussed before separately in Section (4.6) in the massless case
(for the massive case see Section 7.5). To see this equivalence we observe first of

all, that both representations have a spin content 0 and 1. For the spinor (x)
this follows from the addition rule of angular momenta (4.833). In the vector field
A (x), the zeroth component transforms according to the spin-0 representation, the
spatial components according to the spin-1 representation of the rotation group.
There exists asimple
 relation between the two fields. The spinor field transforms
1 1
according to the 2 , 2 representation of the Lorentz group as follows:

(x )

= D ( 2 ,0) () D (0, 2 ) () (x)


=

D ( 2 ,0) ()(x)D (0, 2 )T ()

(4.914)

The 2 2 components of the spinor can be mapped into the four components of a
vector by forming

(x) c
(x) = tr[c
(x)].

(4.915)

Using the Lorentz transformation rules of Section 4.17 it is easy to verify that (x)
transforms indeed like the vector field A (x) in Eq. (4.280):

(x)
(x ) = tr c
D ( 2 ,0) ()(x)D (0, 2 )T ()
1

= tr c(c1 D (0, 2 )T ()c)


D ( 2 ,0) ()(x) .

Now we make use of the relation (4.677) to set


1

c1 T c = ,

(4.916)

c1 D (0, 2 )T ()c = D ( 2 ,0) (),

(4.917)

(4.918)

and to rewrite (4.916) as




(x ) = tr cD ( 2 ,0) ()
D ( 2 ,0) ()(x) .

With the help of (4.479) we obtain now the vector property of the composite field
m (x):
(x ) = tr [c
(x)] = (x),

(4.919)

so that (x) transforms indeed like A (x) in (4.280).

A special feature of all representations D (s,s ) with s = s is that they are invariant under space inversions since this interchanges s and s . Thus no doubling
of

fields is needed to accommodate space inversions. In the vector form of the 21 , 21
representation this was seen before in the transformation law (4.296).

368

4.21

4 Relativistic Free Particles and Fields

Rarita-Schwinger field for Spin 3/2

Another frequently-encountered form of higher spin fields which is not of the (s, 0) +
(0, s) type is due to Rarita and Schwinger and describes spin-3/2 particles. [4] It
combines vector and bispinor properties and is written as a (x), thus transforming
according to
1

(x )
a (x)
a

0
D ( 2 ,0) ()
(0, 21 )
0
D
()

b (x).

(4.920)

ab

Group-theoretically speaking, this is a direct product of the representations




1 1
,
2 2

(for the indices , ) and 21 , 0 + 0, 21 (for the indices a, b).


We can employ the usual rules for the addition of angular momentum and apply
them to J(1) and J(2) in (4.825). Then the direct product of two representations

D (s1 ,s2) D (s1 ,s2 )

(4.921)

must have the following irreducible contents

D (|s1 s2 |,|s1s2 |) + D (|s1 s2 |,|s1s2 |+1)


+ . . . + D (|s1 s2 |,s1+s2 )

|s1 s2 |+1,|s1 s2 |)
(|s1 s2 |+1,|s1 s2 |+1)
+D
+ D
+ . . . + D (|s1 s2 |+1,s1+s2 )
+ ...

+ D (s1 +s2 ,|s1 s2 |) + D (s1 +s2 ,|s1 s2 |+1)


+ . . . + D (s1 +s2 ,s1 +s2 ) .
In this expansion, the spins s1 and s2 of of J1,2 combine to all spins from |s1 s2 | to
s1 + s2 [recall (4.833)], and similarly s1 and s2 to all spins from |s1 s2 | to s1 + s2 .
Therefore a is equivalent to a sum of D (s1 ,s2 ) representations:
1
1
1
1
0,
+ 1,
+
,0 +
,1 .
2
2
2
2

(4.922)

Remember that the symmetry with respect to the interchange s1 s2 is necessary


for a parity-invariant Lagrangian.
Now,
if we want to describe only a spin-3/2 par


1
1
ticle, the representations 0, 2 and 2 , 0 are superfluous and have to be projected
out. This can be done by a constraint analogous to the Lorentz condition for the
electromagnetic field:
a (x) = 0.
(4.923)


Obviously, this derivative transforms like 0, 21 + 21 , 0 and has only a spin-1/2


content, which is therefore
from (4.922). It remains to make sure that
 removed


1
1
the representation 1, 2 + 2 , 1 in a describes only a spin-3/2 particle. This is
achieved by another condition on the field:
ab b = 0.

(4.924)

Appendix 4A

369

Derivation of Baker-Campbell-Hausdorff Formula




This projection of the field transforms once more like 0, 12 + 21 , 0 , and setting
it equal to zero eliminates one more spin- 12 degree of freedom thus ensuring the
survival of only the spin-3/2 content in a .
It remains to construct an invariant action with the property that the equations
of motion automatically ensure the constraints (4.923), (4.924). There are now
several possible invariants which can be used. If we allow at most for one derivative
we may combine
i ,

(4.925)

The most general combination which leads to a hermitian action of a pure spin-3/2
particle can be shown to be20
A=

d4 x L(x) =

d4 x (x)L (i) (x),

(4.926)

where L (i) is the differential operator


L (i) = (i/
M)g + w i + w
1
+ (3ww + w + w + 1) i/

2 

3
+M 3ww + (w + w ) + 1 ,
2

(4.927)

where w is an arbitrary complex number. The equations of motion are given by


L (i) (x) = 0.

(4.928)

It can easily be verified that a field a (x) which satisfies the constraints (4.923),
(4.924) solves (4.928) if and only if the Dirac equation is fulfilled separately for each
vector index :

(4.929)
(i/
M)a a a (x) = 0,
showing that the particle has mass M. Some algebra is necessary to show that the
constraints (4.923), (4.924) follow from (4.928). For this we go to momentum space
and contract L (p) (p) = 0 once with and once with p , using the relations
p/ = /
p + 2p and = 4. The two contractions yield = 0 and

p = 0, which are a direct consequence of the anticommutation rules (4.493).

Appendix 4A

Derivation of Baker-Campbell-Hausdorff
Formula

The standard Baker-Campbell-Hausdorff formula, from which our formula (4.74) can be derived,
reads

eA eB = eC ,
(4A.1)
20

See Notes and References for literature.

370

4 Relativistic Free Particles and Fields

where
+
C = B

dtg(eadA t eadB )[A].

(4A.2)

Here g(z) is the function


g(z)

X
log z
(1 z)n
=
z 1 n=0 n + 1

(4A.3)

in the so-called adjoint representation, which is defined


and adB is the operator associated with B
by
[B,
A].

adB[A]
(4A.4)
0

One also defines the trivial adjoint operator (adB) [A] = 1[A] A. By expanding the exponentials
in Eq. (4A.2) and using the power series (2A.3), one finds the explicit formula
+ A +
C = B

X
(1)n
n+1
n=1

pi ,qi ;pi +qi 1


q1

(adA)p1 (adB)
p1 !
q1 !

1+

1
Pn

i=1

pi

(adA)pn (adB)qn
[A].
pn !
qn !

(4A.5)

The lowest expansion terms are




+ A
1 1 adA + adB + 1 (adA)2 + 1 adA adB + 1 (adB)2 +. . . [A]

C = B
2
6
2
2
2

1

+ 31 (adA)2 + 12 adA adB + 12 adBadA + (adB)2 + . . . [A]


3
A]
+ 1 ([A,
[A,
B]]
+ [B,
[B,
A]])
+ ... .
1 [B,
= A + B
2
12

(4A.6)

To prove formula (4A.2) and thus the expansion (4A.5), we proceed similar to the derivation
of the interaction formula (1.302) by deriving and solving a differential equation for the operator
function
B

= log(eAt
C(t)
e ),
(4A.7)

and obtain the desired operator C from its value at t = 1, C = C(1).


The starting point is the
,
observation that for any M

eC(t)
],
eC(t) M
= eadC(t) [M
(4A.8)
by definition of adC. The left-hand side can also be rewritten as

B
],
e eAt = eadA t eadB [M
eAt eB M

so that we have

(4A.9)

eadC(t) = eadA t eadB .

(4A.10)

d C(t)

e
= A.
dt

(4A.11)

Differentiation of (4A.7) shows that

eC(t)

The left-hand side, on the other hand, can be rewritten in general as

eC(t)

d C(t)

e
= f (adC(t))[C(t)],
dt

where
f (z)

ez 1
.
z

(4A.12)

(4A.13)

Wigner Rotations and Lobatschewksi Geometry of Rapidities 371

Appendix 4B

This will be verified below. It implies that

= A.
f (adC(t))[C(t)]

(4A.14)

We now define the function g(z) as in (4A.3) and see that it satisfies
g(ez )f (z) 1.

(4A.15)

C(t)
= g(eadC(t) )f (adC(t))[C(t)].

(4A.16)

We therefore have the trivial identity

Using (4A.14) and (4A.10), this turns into the differential equation

= eadA t eadB [A],

C(t)
= g(eadC(t) )[A]

(4A.17)

from which we find directly the result (4A.2).


To complete the proof we must verify (4A.12). For this consider the operator
d C(t)s

t) eC(t)s
e
.
O(s,
dt

(4A.18)

Differentiating this with respect to s gives


t) =
s O(s,
=
=

eC(t)s C(t)



d 
d  C(t)s

C(t)eC(t)s
e
eC(t)s
dt
dt

C(t)s

eC(t)s C(t)e

eadC(t)s [C(t)].

(4A.19)

Hence
t) O(0,
t) =
O(s,

, t)
ds s O(s

=
from which we obtain

sn+1

(adC(t))n [C(t)],
(n
+
1)!
n=0

d C(t)

t) = eC(t)
e
= f (adC(t))[C(t)],
O(1,
dt

(4A.20)

(4A.21)

which is what we wanted to prove.


Note that the final form of the series for C in (4A.6) can be rearranged in many different ways,
using the Jacobi identity for the commutators. It is a nontrivial task to find a form involving the
smallest number of terms.21
The derivation is an excerpt of the textbook cited Ref. [1] on p. 80.

Appendix 4B

Wigner Rotations and Lobatschewksi


Geometry of Rapidities

Here we calculate the full rate of Wigner precession and the related Thomas precession.
21

For a discussion see J.A. Oteo, J. Math. Phys. 32 , 419 (1991).

372

4 Relativistic Free Particles and Fields

4B.1

Wigner Precession

For brevity, we denote the small rotation (4.745) by

 

R(t) eid  dt/2 = e(t+dt)/2 ed/2 e/2 = B 1 ( (t + dt))B(d )B( ).

(4B.1)

The pure rotation character of the product on the right-hand side is obvious, since a particle in
its rest frame is transformed by three boost transformations back to the rest frame. Being a small
rotation, the left-hand side has necessarily the form

where

R=1i

dt/2,

(4B.2)

is an angular velocity describing the Wigner precession rate with a Heisenberg equation

 = U (R) U (R)
1

.

(4B.3)

The parameter d of the infinitesimal Lorentz transformation in (4B.1) must be chosen such that
the final laboratory rapidity is + d . As every Lorentz transformation, the transformation

 

 

B 1 ( )B(d )B( )

(4B.4)

can be decomposed into a product of a Lorentz transformation and rotation:

  
)B(d ),
(4B.5)
where both parameters d and d are small. Then we can expand (4B.1) up to first order in d
B 1 ( )B(d )B( ) R(d

as follows:

= B 1 ( (t + dt))B( )R(d )B(d )

)B( )dt + [R(d ) 1] + [B(d ) 1].


= 1 + B(

 

(4B.6)

It is straightforward to calculate the second term in the notation (4.458), in natural units with
c = 1, where
1
(4B.7)
B( ) = p
( + 1 v ).
2( + 1)

and

)=
B(

1
B( ) + p
[ (1 + v ) + v ]
2( + 1)
2( + 1)

(4B.8)

) we obtain
After multiplying this with B(

 

)B( ) =
B(

+
[ (1 + v
2( + 1) 2( + 1)

The sum of all terms without any factor

) + v ] ( + 1 v ).

(4B.9)

 cancel each other since they are equal to




v 2 + 2 v v ,
+
2( + 1) 2 2( + 1)

(4B.10)

which can be shown to vanish: to to




1
2

v + 2 = 0,
2( + 1) 2 2( + 1)

using the trivial identities


= v v 3 ,

v2 =

2 1
,
2

(4B.11)

(4B.12)

Wigner Rotations and Lobatschewksi Geometry of Rapidities 373

Appendix 4B

With the help of Formula (4.747), the remaining terms can be decomposed as follows:

 

/2 i /2,

)B( ) =
B(

(4B.13)

where

= ( +1 1) [ v ( + 1)v] = v
2

v ,

(4B.14)

is found via Eq. (4B.12). The vectors v k and v denote the projections of v parallel and orthogonal
to v, respectively. The parameter in (4B.13) is found to be


= ( + 1) v v.
2

(4B.15)

The first, hermitian term in (4B.13) corresponds in (4B.6) to a pure infinitesimal Lorentz transformation. Since the final result (4B.6) must be a pure rotation, this term must be canceled by
the last term in (4B.6) which is also hermitian. Thus we conclude that

B(d ) = 1


1 2
v k + v
2

 dt.

(4B.16)

The remaining antihermitian term is of the type (4B.2) and gives a first contribution to the angular
velocity of Wigner rotations (after reinserting here the omitted light velocity c):

W1

1 2

(v v).
c2 + 1

(4B.17)

In order to find the second contribution to the angular velocity we must calculate the term

) 1 in (4B.6). Thus we transform B(d) = 1 d /2 according to (4B.4) and (4B.5):


R(d )B(d ) = e  e   e   = B ()B(d)B()
1
1
d 
p
=1 p
( + 1 + v )
( + 1 v ).
(4B.18)
2
2( + 1)
2( + 1)

R(d

/2 d /2 /2

We now use the two rules

 

[a , b ] = 2i(ab) , a

 b  c  = a b c  a c b  + b c a  + i(ab) c, (4B.19)

to calculate

)B(d )1 = 4(1+1) ( + 1) + v d 2 vd v  +2i(vd)  .

R(d

2 2

(4B.20)

Expressing v 2 via (4B.12), the bracket simplifies to 2( + 1). After separating d into parallel
and orthogonal projections with respect to v, we obtain:


1
i
(v d )
.
(4B.21)
R(d )B(d ) 1 = d k + d +
2
+1

   

 

Comparison with (4B.16) we identify from the hermitian terms

= v k dt,

= v dt, ,

d = vdt,

) in the decomposition (4B.6) is


/2,
R(d ) 1 i

and find that the extra extra rotation R(d

(4B.22)

W2

(4B.23)

374

4 Relativistic Free Particles and Fields

with an angular velocity

which is twice the negative of (4B.17):

W2

W2

2 2

(v v).
c2 + 1

(4B.24)

The total angular velocity of Wigner rotation is therefore

W1

W2

1 2

(v v),
c2 + 1

(4B.25)

which generalizes the small-velocity result in Eq. (4.752) to (4.760).

4B.2

Thomas Precession

It is now easy to modify the calculation to obtain the corresponding generalization of the Thomas
frequency (4.758). We simply replace the small Lorentz transformation B(d ) in (4B.1) by the
small Lorentz transformation (4.753) in the rest frame of the moving electron. As a consequence,

the transformation (4B.5) is simply B(d


) and the only difference with respect to the previous

calculation is that the rotation R(d ) is absent. For this reason the over-compensating rotation
by W 2 is absent and we find again that the rate of Thomas precession T is equal to the first
contribution W 1 in Eq. (4B.17) to the Wigner precession and thus precisely the opposite of the
total rate of Wigner rotation:

T = W =

4B.3

1 2

(v v).
c2 + 1

(4B.26)

Calculation in Four-Dimensional Representation

The above calculations can certainly be also performed in the 4 4 -representation of the Lorentz
group. As an illustration, let us rederive the 4 4 -version of Eq. (4B.13). We denote the 4 4

-representation B  ()B () by (v)(v).


Differentiating (4.18) with respect to time we see
that

v i
vi

c
c

(4B.27)
(v)
=
i j
i j .
i j
2
i
i

v v + v v
v
v
( + 2) v v
+

c
c
( + 1)2 c2
+1
c2

Multiplying this by (v) from the right yields for the first row [(v)(v)]
i of the product
 





vi
v2
vv
2 v2 vi
v i
3 vv

1 2 2 2 , 1
.
(4B.28)

c
c
+ 1 c2 c
c + 1 c2 c

Using again the relations vv/c


= /
3 , and v 2 /c2 = ( 2 1)/ 2 [compare (4B.12)], the first entry
and the -terms

in the second entry disappear, and we remain with




vi
3 vv
v i
0

.
(4B.29)
[(v)(v)] = 0,
c
+ 1 c2 c

Introducing the components v k and v of the acceleration parallel and orthogonal to v, such

that (vv)v
= v 2 v k = ( 2 1)v k / 2 , this can be written as in terms of the 4 4 matrices (4.60)(4.62) generating pure Lorentz transformations. Their first rows are (M1 )0 i =
i(0, 1, 0, 0), (M2 )0 i = i(0, 0, 1, 0), (M3 )0 i = i(0, 0, 0, 1), such that we can write with the vector
notation M (M1 , M2 , M3 ):
0

[(v)(v)]
i =


i 2
v k + v M.
c

(4B.30)

Wigner Rotations and Lobatschewksi Geometry of Rapidities 375

Appendix 4B

If we replacing the 4 4 -generators M in this equation by the 2 2 -generators of pure Lorentz


transformations i /2 we obtain

1
(4B.31)
2 v k + v ,
2c
which agrees with the previous result (4B.16), apart from the factor 1/c omitted there.
A third way of deriving this result makes use of the spin four-vector introduced in Eq. (4.761).
We may calculate the precession rate by comparing the spin at the time t, where the velocity
is v, with the spin at t + dt, where the velocity is v(t + dt), after the particle has been slightly
accelerated. During this time, the spin four-vector changes from S to S S + S dt. The initial
spin is obtained by bringing the electron to its rest frame via a deboosting Lorentz transformation
(v). Using Eq. (4.18) and v 2 from (4B.12), we have

vi 0
2 vi vj j
S .
S + Si +
c
+ 1 c2

SR i = i (v)S =

(4B.32)

The final spin is obtained by a similar Lorentz transformation with a slightly different velocity
v(t + dt). The result is
i
i
SR
+ dSR

= i (v(t + dt))(S + S dt)


= i (v)S + [ i (v)S + i (v)S ]dt

(4B.33)

We now use the fact that for an acceleration by a pure boost, which does not change the total
angular momentum, the change of the spin four-vector S is parallel to the direction of u =
(, v i ). This will be shown in Eq. (6.59). Moreover, we can easily verify that i (v)u = 0 by
substituting u for S in (4B.32). Hence we obtain
i
S R

i
dSR
= i (v)S
dt

 i
( + 2) v i v j j
2 v i v j + v i v j j
v i
v
S 0 +
S
+
S .
= +
c
c
( + 1)2 c2
+1
c2

(4B.34)

Expressing S i with the help of (4.773), and using vv/c


= /
3 , the last term becomes
 i j

2 v i v j j
v v j
2 v i v j j
2
2 vv v i v j j

vi vj j
=
S
=
S
+
S
S
+

S .
R
R
R
+ 1 c2
+1
c2
+1 c2 c2
+1 c2
( +1)2 c2 R
i
i
We now use (4.773) to substitute v i S i /c = v i SR
/c and S 0 = v i S i /c = v i SR
/c. Then all terms
containing cancel each other and we arrive at the formula for the temporal change of the spin
vector:

1 2
1 2
j
i
S R ]i
S R
= 2
(v i v j v i v j )SR
= 2
[(v v)
c +1
c +1

SR ,

(4B.35)

with the vector of angular velocity (4B.17). As a rate of change of a three-vector, it corresponds
to a pure rotation. With the help of the generators (Li )jk = iijk of the rotation group, we may
also write

S R = i(

4B.4

L)SR .

(4B.36)

Hyperbolic Geometry

Such kinematic calculations can, incidentally, be done quite elegantly in a geometric approach.
One may exploit the fact that the four-velocities u p /M can be written as

u = c(cosh , sinh ),

(4B.37)

376

4 Relativistic Free Particles and Fields

which shows that up to a factor c they are vectors on a unit hyperbola. These are hyperbolic
analogs of Euclidean vectors
(4B.38)
uE = c(cos , sin )

on a unit sphere. As such, relative rapidities follow the hyperbolic version of spherical trigonometry,
called Lobatschewski Geometry. The product of three pure Lorentz transformations B( ) = e/2
with rapidities a , b , c can be represented as a triangle in hyperbolic space. The angles of the
triangle a , b , c indicate the relative angles between the corresponding vectors, i.e., cos a =
, etc. (see Fig. 4.4). The angles and sides of the triangle are then related by the cosine and
b
c

  

 

Figure 4.4 Triangle formed by rapidities in a hyperbolic space. The sum of angles is
smaller than 1800 . The angular defect yields the angle of Thomas precession.
sine theorems
cosh a
cos a

and

= cosh b cosh c sinh a sinh b cos c ,

= cos b cos c + sin a sin b cosh c ,

sinh a
sinh b
sinh c
=
=
,
sin a
sin b
sin c

(4B.39)
(4B.40)
(4B.41)

(4B.42)

respectively.
Given two sides and the enclosed angle, say a , b , c , we can use Napiers analogies
sinh 21 (a b )
sin 12 (a + b )

tan 12 (a b )
cot 21 c

cosh 12 (a b )
cosh 12 (a + b )

tan 21 (a + b )
cot 21 c

(4B.43)

to calculate the other two angles a , b . After that, either of the two analogous formulas
sin 21 (a b )
sin 12 (b + b )

tanh 12 (a b )
tanh 21 c

cos 12 (a b )
cos 12 (a + b )

tanh 12 (a + b )
tanh 21 c

(4B.44)

serves to calculate the third side c .


Since the hyperboloid has a negative unit radius, the sum of the angles is less than . The
angular defect, also called excess
E = a b c
(4B.45)

Appendix 4C

Clebsch-Gordan Coefficients

377

determines the area A of the triangle. For a hyperbola of radius R:


A = R2 E.

(4B.46)

In spherical geometry this formula is known as Girards theorem. The angular defect is given
in terms of the three sides by the hyperbolic version of the LHuilliers formula in spherical
trigonometry:22
r
s
s a
s b
s c
E
tanh
tanh
,
(4B.47)
tan = tanh tanh
4
2
2
2
2
where
s = (a + b + c )/2
For R it reduces to Herons formula.
p
A = s(s a)(s b)(s c),

Anor formula is

cos

s = (a + b + c)/2 = semiparameter.

E
1 + cosh a + cosh b + cosh c
.
=
2
4 cosh2 (a /2) cosh2 (b /2) cosh2 (c /2)

(4B.48)

(4B.49)

(4B.50)

A pure Lorentz transformation of a particle is a parallel transport along a side of a triangle.


When doing three successive parallel transports around a triangle, a particle which was initially
at rest comes again to rest. Its spin, however, changes the direction by the angular defect which is
determined by the area integral. Since the radius is here equal to 1, formula (4B.50) determines
directly the total angle of Thomas precession.
The reader is encouraged to derive the rate of Thomas precession once more using this geometric calculus.23

Appendix 4C

Clebsch-Gordan Coefficients

A direct product of irreducible representation states |s1 , m1 i and |s2 , m2 i can be decomposed into a
sum of irreducible representation states |smi with total angular momentum s = |s1 s2 |, , (s1 +
s2 ). This is done with the help of Clebsch-Gordan coefficients. For this we multiply any product
state with the completeness relation of all irreducible representation states
s
X X

s m=s

|s, mihs, m| = 1,

(4C.1)

and obtain
|s1 , m1 ; s2 , m2 i =

sX
1 +s2

s
X

s=|s1 s2 | s,m=s

|s, mihs, m|s1 , m1 ; s2 , m2 i.

(4C.2)

The expansion coefficients on the right-hand side are the desired Clebsch-Gordan coefficients.
The expansion (4C.2) can be inverted by means of a similar completeness relation in the
product space
s2
s1
X
X
|s1 , m1 ; s2 , m2 ihs1 , m1 ; s2 , m2 | = 1,
(4C.3)
m1 =s1 m2 =s2

22

See, for example, D.D. Ballow and F.H. Steen, Plane and Spherical Trigonometry with Tables,
Ginn, New York , 1943
23
For details see J.A. Smorodinskij, Fortschr. Phys. 13, 157 (1965); A. Sommerfeld, Electrodynamics, Academic, New York, 1949.

378

4 Relativistic Free Particles and Fields

yielding the expansion


|s, mi =

s2
X

s1
X

m1 =s1 m2 =s2

|s1 , m1 ; s2 , m2 ihs1 , m1 ; s2 , m2 |s, mi.

(4C.4)

By subjecting these relations to an arbitrary rotation (4.835), and using (4.837), we find the
transformation behavior of the Clebsch-Gordan coefficients:
s

s 1
s 1
Dm,m
hs, m |s1 , m1 ; s2 , m2 i(D )
m ,m1 (D )m ,m2 = hs, m|s1 , m1 ; s2 , m2 i,
1

(4C.5)

or, because of unitarity of the representation matrices,


s
s

Dm,m
(Dm ,m ) (Dm ,m ) hs, m |s1 , m1 ; s2 , m2 i = hs, m|s1 , m1 ; s2 , m2 i,
1
2
1

(4C.6)

and since the Clebsch-Gordan coefficients are real following the Condon-Shortley convention we
also have
s
s
s

(Dm,m
(4C.7)
) Dm ,m Dm ,m hs, m |s1 , m1 ; s2 , m2 i = hs, m|s1 , m1 ; s2 , m2 i,
1
2
1

The Clebsch-Gordan coefficients are related in a simple way to the more symmetric Wigner
3j-symbols defined as follows:


s1 s2 s3
1/2
s1 s2 m3
.
(4C.8)
(2s3 + 1)
hs3 , m3 |s1 , m1 ; s2 , m2 i = (1)
m1 m2 m3
As a consequence of relation (4.873), this has the invariance property
 

s1 s2
s1 s2 s3
s
s
s
=
Dm
D
D

m2 ,m2 m3 ,m3
1 ,m1
m1 m2
m1 m2 m3

s3
m3

(4C.9)

The Levi-Civita symbol ijk is a cartesian version of the Wigner 3j-symbol for s1 = s2 = s3 = 1.
It exhibits the invariance (4C.9) with respect to the 3 3 defining representation matrices of the
rotation group:
(4C.10)
Ri1 i1 Ri2 i2 Ri3 i3 i1 i2 i3 = i1 i2 i3 .
Under even permutations of columns, the 3j-symbols are invariant while under odd permutations they pick up a phase factor (1)s1 +s2 +s3 . Also:




s1
s2
s3
s1 s2 s3
s1 +s2 +s3
.
(4C.11)
= (1)
m1 m2 m3
m1 m2 m3
In Eq. (4.878) we introduced a contravariant notation for spin-j objects. This is also done in
relation (4C.8), writing it as


s 1 s 2 m3
1/2
s1 s2 s3
.
(4C.12)
(2s3 + 1)
hs3 , m3 |s1 , m1 ; s2 , m2 i (1)
m1 m2 s 3
The simplest 3j-symbol is


j
m

j
m

0
0

= (1)jm (2j + 1)1/2 m,m

(4C.13)

(j)

this being also equal to (1)2j cm,m . In the contravariant notation, one has



j m 0
= (1)2j (2j + 1)1/2 m m .
m j
0

(4C.14)

In order to calculate the Clebsch-Gordan coefficients we observe that the state of maximal
quantum numbers |s1 , s1 ; s2 , s2 i is a state |s, mi of the irreducible representation with the maximal

Appendix 4C

Clebsch-Gordan Coefficients

379

Table 4.2 The lowest Clebsch-Gordan


coefficients hs, m|s1 , m1 ; s2 , m2 i. The table en
tries CG are to be read as CG. The coefficients are all real, i.e. equal to
hs1 , m1 ; s2 , m2 |s, mi. For more symmetry properties see Eqs. (4C.22).

hs1 , m1 ; s2 , m2 |s, mi
= (1)ss1 s2 hs2 , m2 ; s1 , m1 |s, mi
= (1)ss1 s2 hs2 , m2 ; s1 m1 |s, mi

380

4 Relativistic Free Particles and Fields

angular momentum s = m = s1 + s2 . By repeatedly applying the lowering operator of angular


momentum to it according to the general relation (4.821), we obtain the matrix elements (compare
Fig. 4.3)
p
L |s, si =
(2s) 1|s, s 1i,
p
L |s, s 1i =
(2s 1) 2|s, s 1i,
..
.
(4C.15)
p
(2s 1) 2|s, s + 1i,
L |s, s + 2i =
p
L |s, s + 1i =
(2s) 1|s, si.
In the direct-product space, an application of the lowering operator L 1 + 1 L to the state
|s1 , s1 ; s2 , s2 i yields, with the same rules as in (4C.15),
p
(2s1 ) 1|s1 , s1 1; s2 , s2 i
(L 1 + 1 L )|s1 , s1 ; s2 , s2 i =
p
+
(2s2 ) 1|s1 , s1 ; s2 , s2 1i.
(4C.16)

Continuing this with the help of the general relation


p
(s1 + m1 )(s1 m1 + 1)|s1 , m1 1; s2 , m2 i
(L 1 + 1 L )|s1 , m1 ; s2 , m2 i =
p
+
(s2 + m2 )(s2 m2 + 1)|s1 , m1 ; s2 , m2 1i,

(4C.17)

we find all other states |s, mi of the irreducible representation with s = s1 + s2 .


The state of the lower total angular momentum s1 + s2 1 with maximal magnetic quantum
number m = s is obtained from the orthogonal combination of (4C.17):
p
p
|s1 +s2 1, s1 +s2 1i = (2s1 ) 1|s1 , s1 1; s2 , s2 i (2s2 ) 1|s1 s1 ; s2 s2 1i.
(4C.18)
This can be verified by applying to it the raising operator (L+ 1 + 1 L+ ): generalizing (4.821)
to the direct-product space:
p
(s1 m1 )(s1 + m1 + 1)|s1 , m1 + 1; s2 , m2 i
(L+ 1 + 1 L+ )|s1 , m1 ; s2 , m2 i =
p
(s2 m2 )(s2 + m2 + 1)|s1 , m1 ; s2 , m2 + 1i,
+
(4C.19)
and finding that it is annihilated.
By applying the lowering operator to the state (4C.18). we generate all states of the irreducible
representation |s1 + s2 1, mi with m = s1 s2 + 1, . . . , s1 + s2 1.
Multiplying (4C.17) from the left by hs, m| and using the hermitian adjoint of relation (4.821),
we obtain the recursion relation
p
(s + m)(s m + 1)hs1 , m1 ; s2 , m2 |s, mi
p
= (s1 m1 + 1)(s1 + m1 )hs1 , m1 1; s2 , m2 |s, m 1i
p
(4C.20)
+ (s2 m2 + 1)(s2 + m2 )hs1 , m1 ; s2 , m2 1|s, m 1i.
Similarly we can multiply the raising operator relation (4C.19) in the direct-product space multiply
it by hs, m| from the left and the adjoint of (4.820) to find
p
(s m)(s + m + 1)hs1 , m1 ; s2 , m2 |s, mi
p
= (s1 + m1 + 1)(s1 m1 )hs1 , m1 + 1; s2 , m2 |s, m + 1i
p
+ (s2 + m2 + 1)(s2 m2 )hs1 , m1 ; s2 , m2 + 1|s, m + 1i.
(4C.21)

Appendix 4D

381

Spherical Harmonics

The Clebsch-Gordan coefficients have the following important symmetry properties:


(1)js1 s2 hs, m|s2 , m2 ; s1 , m1 i
hs, m|s2 , m2 ; s1 , m1 i

hs, m|s1 , m1 ; s2 , m2 i =
=

(1)js1 s2 hs, m|s1 , m1 ; s2 , m2 i


r
2s + 1
s1 m1
hs2 , m2 |s1 , m1 ; s, mi
(1)
2s2 + 1
r
2s + 1
(1)s2 +m2
hs1 , m1 |s, m; , s2 , m3 i.
2s1 + 1

=
=
=

(4C.22)

Some frequently-needed values are listed in Table 4.2.

Appendix 4D

Spherical Harmonics

The spherical harmonics are defined as


Ylm (, ) (1)m

2l + 1 (l m)!
4 (l + m)!

where Plm (z) are the associated Legendre polynomials

1/2

Plm (cos )eim ,

(4D.1)

1
dl+m
(4D.2)
(1 z 2 )m/2 l+m (z 2 1)l .
l
2 l!
dx
The spherical harmonics are orthonormal with respect to the rotation-invariant scalar product
Z
Z 2

d sin
d Ylm
(, )Yl m (, ) = ll mm .
(4D.3)
Plm (z) =

Explicitly, they read for the lowest few angular momenta:


Y0,0 (, )

Y11 (, )

Y10 (, )

Y22 (, )

Y21 (, )

Y20 (, )

Y33 (, )

Y32 (, )

Y31 (, )

Y30 (, )

1
,
2
r
1
3
sin ei ,

4
2
r
3
cos ,
4
r
1 15
sin2 e2 i ,
4 2
r
15

sin cos ei ,
8
r


3
1
5
2
,
cos
4 2
2
r
1 35
sin3 e3 i ,

4 4
r
1
105 2
sin cos e2 i ,
4
2
r

21
1

sin 5 cos2 1 ei ,
4 4
r


7
5
3
3
cos cos .
4 2
2

(4D.4)

382

4 Relativistic Free Particles and Fields

The spherical harmonics with negative magnetic quantum number m are obtained via the relation

Ylm (, ) = (1)m Yl,m


(, ).

(4D.5)

For m = 0, the spherical harmonics reduce to


Ylm (, ) =

2l + 1
Pl (cos ),
4

where
Pl (z) Pl0 (z) =

(4D.6)

1 dl 2
(z 1)l
2l l! dz l

(4D.7)

are the Legendre polynomials.


For integer j = l, the rotation functions djm,m () can be derived recursively from dlm,0 () =
Pl (cos ) with the help of the recursion relation
p
p
j1
2 (j + m )(j + m 1)djm,m () = (j + m)(j + m 1)(1 + cos ) dm1,m
1 ()
p
p
+2 j 2 m2 sin dj1
(j m)(j m 1)(1 cos ) dj1
m,m 1 () +
m+1,m 1 ().
(4D.8)
For an iterative determination of the rotation functions djm,m () with half-integer j we use the
recursion relation
s
s
j m j1/2
j + m j1/2

j
dm +1/2,m+1/2 cos +
dm 1/2,m+1/2 sin .
(4D.9)
dm ,m () =
jm
2
jm
2
Inserting m = 1/2 and using (4.848), we deduce
s
s
j + m j1/2
j m j1/2

j
dm 1/2 () =
dm1/2,0 cos
dm+1/2,0 sin .
j + 1/2
2
j + 1/2
2

(4D.10)

For j = l + 1/2, the right-hand side contains only Legendre polynomials. Starting from these
djm,1/2 (), we find all other rotation functions with half-integer j the recursion relation (4D.8).
For j = 1/2 and j = 1, the explicit results were given in Eqs. (4.853) and (4.854). For j = 3/2
we obtain
3/2

d3/2,1/2 ()

3/2

3/2

3/2

d3/2,3/2 ()

d3/2,1/2 ()
d3/2,3/2 ()
3/2

d1/2,1/2 ()

3/2

d1/2,1/2 ()

(1 + cos ) cos ,
2
2

3
(1 + cos ) sin ,

2
2

3
(1 cos ) cos ,
2
2

1
(1 cos ) sin ,
2
2

1
(3 cos 1) cos ,
2
2
1

(3 cos + 1) sin .
2
2

(4D.11)

The remaining matrix elements are obtained via relations (4.848). Similarly for j = 2:
d22,2 ()

d22,1 ()

1
(1 + cos )2 ,
4
1
(1 + cos ) sin ,
2

383

Notes and References


d22,0 ()

d22,1 ()

d22,2 ()

d21,1 ()

d21,0 ()

d21,1 ()

d20,0 ()

6
sin2 ,
4
1
(1 cos ) sin
2
1
(1 cos )2 ,
4
1
(1 + cos )(2 cos 1),
2r
3

sin cos ,
2
1
(1 cos )(2 cos + 1),
2
1
(3 cos2 1).
2

(4D.12)

Notes and References


The particular citations in this chapter refer to:
[1] The light velocity c has, by definition, the value stated in Eq. (4.2). This has been so since
1983, when the previous meter has been redefined in the Conference Generale des Poids et
Mesures in Paris to make this value exact. For a discussion see
B.W. Petley, Nature 303, 373 (1983).
[2] O. Klein, Z. Phys. 37, 895 (1926);
V. Fock, ibid. 38, 242; 39, 226 (1926);
W. Gordon, ibid. 40, 117 (1926).
Note that the name Klein-Gordon equation does injustice to E. Schrodinger. He actually
invented the Klein-Gordon equation first , from which he derived his famous nonrelativistic
wave equation in the limit of large c, although his papers in Ann. Phys. 79, 361, 489; 80,
437; 81, 109 (1926), suggest the opposite order. This was pointed out by
P.A.M. Dirac in The Development of Quantum Theory, Gordon and Breach, N.Y., 1971.
See also Diracs popular articles in Nature 189, 335 (1961) and in Scientific American 208,
45 (1963).
[3] This statement holds in the sense of distributions. Statements about distributions must
always be integrated with an arbirary smooth testfunction as a factor to test their validity.
The Riemann-Lebesgue Lemma states that an integral over an infinitely rapidly oscillating
function multiplied by a smooth function yields zero. See Chapter 1 of E.C. Titchmarsh,
Introduction to the Theory of the Fourier Integral , Oxford University Press, Oxford, 1937.
[4] The Rarita-Schwinger action and its generalizations have been investigated in great detail
in attempts to understand the pion nucleon scattering amplitude near the first resonance at
around 1240 GeV. Among the many references see:
R.D. Peccei, Phys. Rev. 176, 1812 (1968);
L.S. Brown, W.J. Pardee, and R.D. Peccei, Phys. Rev. D 4, 2801 (1971);
V. Bernard, U-G. Meissner Phys. Lett. B 309, 421 (1993), Phys. Rev. C 52 2185 (1995);
V. Bernard et al., Int. J. Mod. Phys. E 4, 193 (1995).
Details and many references can be found in the comprehensive review by
G. H
ohler, Elasic and Charge Exchange Scattering of Elementary Particles, in LandoltBornstein Vol. I/9b2, Springer, Berlin 1983.
[5] A thorough discussion of the rotation group is found in the textbook
A.R. Edmonds, Angular Momentum in Quantum Mechanics, Princeton Univ. Press, Princeton, 1962.

384

4 Relativistic Free Particles and Fields

[6] S. Deser and C. Teitelboim, Phys. Rev. D 13, 1592 (1976);


C. Bunster and M. Henneaux, Phys. Rev. D 83, 045031 (2011), (arXiv:1011.5889);
S. Deser, Class. Quant. Grav. 28 085009 (2011), (arXiv:1012.5109).

All power corrupts


Chinese Wisdom

5
Classical Radiation
5.1

Electromagnetic Waves

Changes in the local or temporal charge distribution lead to changes in the electromagntic field
created by them. If the change is sufficiently rapid, the electromagnetic field will start propagating
with light-velocity through spacemine. The standard example is the radiation emitted by an
antenna. A mass distribution lead to changes of the gravitational field. Since the adjustment to a
new field configuration can propagate at most with the speed of light, the universe must be filled
with gravitational waves. The collapse of stars, supernovae explosion, birth of neutron stars, and
similar dramatic events in the universe must all be accompanied by burst of such waves whose
general properties will now be studied.

5.2

Wave Equation

In the presence of a classical source j (x), the electromagnetic action (4.234) contains
an extra term
1Z
A =
dx A (x) j (x)
c
j

(5.1)

so that the Euler-Lagrange equation for the electromagnetic field becomes


1
F = 2 A + A = j .
c

(5.2)

In order to solve this we have to choose a specific gauge, say the Lorentz gauge
A (x) = 0,

(5.3)

in which case the equation reduces to


1
2 A = j .
c

(5.4)

(s) gives rise to a current


A charge point moving along the trajectory x
j (x) = ec

ds x (s) (4) (x x(s)) = ec


385

x (s) (3)
(s)),
(x x
d
x0 /ds

(5.5)

386

5 Classical Radiation

where s is an invariant length parameter, such that


d
x0
=
ds

2 (t)
v
= (t),
c2

(5.6)

(t) is the velocity along the trajectory x


(t):
where v
(t) x
(t).
v

(5.7)

The current components are


(t)),
j 0 (x) = ec (3) (x x

(t)).
j(x) = e
v(t) (3) (x x

(5.8)

To find the electromagnetic field emerging from this current we solve the field equation (5.4) by
Z
i
d4 y GR (x y) j (y),
(5.9)
A (x) =
c
where GR (x y) is the retarded Green function
GR (x x ) = i(x0 x0 )

i
1 h 0 0
(x x R) (x0 +x0 + R) .
4R

(5.10)

Thus we can write Eq. (5.9) as


A (x) =

1
4c

d3 x

1
j (x )|t =t|x x|/c .
|x x |

(5.11)

(t) from the charge, the components


Introducing the distance vector R(t) x x

of A (x) are the famous Lienard-Wiechert potentials:


1
(t t R(t )),
4R
Z
1
(t )(t t R(t )).
v
A(x) = e dt (t t )
4R
0

A (x) = ec

dt (t t )

(5.12)
(5.13)

We now simplify the -functions as follows:


1
1
(t tR ) =
(t tR ),

(tR )
+ R(t )]/dt |t =tR
1 n(tR ) v
(5.14)
where n(t) is the direction of the distance vector R(t)
(t t R(t )) =

|d[t

n(t)

R(t)
.
|R(t)|

(5.15)

Inserting this into (5.12) and (5.13), we find the vector potential
(t) x
(t).
v

(5.16)

387

5.2 Wave Equation


"

e
1
A (x) =
/c)R
4 (1 n v
0

ret

"

1
e
v/c
A(x) =
/c)R
4 (1 n v

(5.17)

ret

The brackets with the subscript ret indicate that the time argument is retarded
with respect to the time on the left-hand side of the equation:
t tR t R(tR )/c.

(5.18)

By forming the combinations


1
E(x) = A(x)
A0 (x),
c
B(x) = A(x),

(5.19)
(5.20)

from the Lienard-Wiechert potential (5.17), we find the electromagnetic fields


/c) 1
/c) v
] 1
(n v
n [(n v
e
+
E(x, t) =
2
2
3
/c) R
/c)
4 (1n v
(1n v
R
B(x, t) = [n E]ret .
"

(5.21)

ret

(5.22)

The two terms in E(x, t) have different falloff behaviors in 1/R with increasing
distance from the source. The first is a velocity field which falls off like 1/R2 . It
is essentially the moving static field around the particle. The second term is an
acceleration field, which has a slower fall-off proportional to 1/R. For this reason
it can carry off radiation energy to infinity. Indeed, the energy flux through a solid
angle d is given by the scalar product of the Poynting vector1 E B with the area
element dS = r 2 d:
=
E = dS (E B) = r 2 dr 2 (E B) k

1
k)|
2 = r 2 d cE2 , (5.23)
| k (k
3
8c

For a radiating electron at small velocities near the coordinate origin, the acceleration field simplifies to
E(x, t) =

e
)],
[
x (
xx
4rc2

B(x, t) =

e
).
(
xx
4rc2

(5.24)

with r = |x|. The radiated power per solid angle is then


e2
e2
dE
2

2 sin2 ,
=
(
x

x
)
=
x
d
(4)2 c3
(4)2 c3

(5.25)

where is the angle between the oscillating dipole end the direction of emission.
By integrating over all solid angles, we obtain the total radiated power
E =
1

e2 Z
2 e2 2
2

.
d
(
x

x
)
=
x
(4)2 c3
3c3 4

(5.26)

Note that the Poynting vector coincides with the components T 0i = (E B)i of the energymomentum tensor of the electromagnetic field, whose-component T 00 = 12 (E2 + B2 ) is the energy
density.

388

5 Classical Radiation

This is the famous Larmor formula of classical electrodynamics.


For a harmonically oscillating charge at position
x(t) = x0 eit + x0 eit = 2|x0 | cos(t + ),

(5.27)

equation (5.25) yields the temporal average power


dE
e2 4
e2 4
= 2 4 |
x x0 |2 = 2 4 |x0 |2 sin2 ,
d
8 c
8 c

(5.28)

and the total radiated power is given by the antenna formula


e2 4 4
E =
|x0 |2 .
4 3 c4

(5.29)

Note that with respect to standard classical electrodynamics textbooks where the
electromagnetic Lagrangian carries a prefactor 1/4, the square of the charge has an
extra factor 1/4, and e2 is related to the finestructure constant by e2 = 4hc.
A more general radiation formula valid for any charge configuration is obtained
by rewriting the vector potential (5.11) as
A (x) =

1
4c

d3 x

1
|x x |

where

j (x , ) =

j (x , )eit+i|xx |/c ,
2

dt eit j (x , t)

(5.30)

(5.31)

are the Fourier components of the energy-momentum tensor. Approximating for


large r = |x|:

ei|xx |/c
eir/c inx /c

e
(5.32)
|x x |
r
leaving an x -dependence only in the sensitive phase factor, caused by the exact
splitting (5.21) into a velocity and an acceleration field which carries off anergy.
At a point x far away from the source, the spherically radiated field (5.30) looks
like a passing plane wave with eir/c eikx . We therefore set nx /c = kx , and
find
A (x, t) =

1 eir/c
4c r

d it
e
j (k, ),
2

(5.33)

where j (k, ) is the Fourier transform in spacetime:

j (k, ) =

dt

d3 x eitikx j (x, t).

(5.34)

At a fixed k and , we can also write


A (x, t) =

1 eir/c
j (k, t).
4c r

(5.35)

389

5.2 Wave Equation

We now calculate the energy flux from formula (5.23). Expressing on the right-hand
side the electric field in terms of the vector potential via (5.19), we find
d2 E
1
+ A0
= r2 c A
d
c


2

(5.36)

In momentum space, the Lorentz gauge (5.3) implies


A(k, t),
A0 (k, t) = k
such that we can rewrite
i
1
1 h
0 (k, t) = 1 t AT (k, t),
A(k, t) + ikA0 (k, t) = t A(k, t) kA
c
c
c

(5.37)

(5.38)

where

AT (k, t) PT (k)A(k,
t)

(5.39)

is the transverse part of the vector field A(k, t) defined by the projection matrix
(4.324),
ij k i kj ,
(5.40)
PTij (k)

which satisfies the property (4.325).


We now express AT in terms of jT PT j using Eq. (5.35) and find from Formula
(5.36) the radiated energy per unit time and solid angle
1
dE
=
[t jT (k, t)] [t jT (k, t)].
d
16 2 c3

(5.41)

An equivalent expression can be obtained by rewriting (5.36) directly as


1
dE
2
=
[t j(k, t)] [t j(k, t) c2 k2 j 0 ].
2
3
d
16 c

(5.42)

2
2
Since the radiated waves are on the mass shell, we can replace c2 k2 j 0 by j 0 , so
that (5.43) can be written in a fully covariant form as

dE
1
=
[t j (k, t)] [t j (k, t)|.
d
16 2c3

(5.43)

Of course we could also have used current conservation in momentum space to reexpress in (5.41):
jT (k, t)jT (k, t) = j l (k, t)(lm kl km )j m (k, t) = j2 (k, t)j 02 (k, t) = j (k, t)j (k, t).
(5.44)
For long wavelengths, the spatial components
of current density have a negligible
R
dependence on k: j i (k, t) j i (k = 0, t) = d3 xj i (x, t). Then we can perform the
integral over all angles in (5.41) using the angular averages
1
hki kj i = ij ,
3

= 2 ij ,
hPijT (k)i
3

(5.45)

390

5 Classical Radiation

to find, for long wavelenths,


E

1 2
t
4c3 3


d x j(x, t)

2

(5.46)

We may now use a partial integration and current conservation to calculate for the
integral on the right-hand side the simple expression
Z

d x j(x, t) =

where

d x x j(x, t) =

d(t)

d3 x x(x,
t) = d(t),

d3 x x (x, t)

(5.47)

(5.48)

is the dipole moment of the charge distribution. Hence (5.46) becomes


2 1
2
E = 4 [d(t)]
.
3c 4

(5.49)

For a single nonrelativistic point particle moving along the orbit x(t), the spatial
current density is
(t) (3) (x x
(t)),
j(x, t) = e x
(5.50)
and Eq. (5.50) becomes

2 e2
(t)]2 ,
E = 3 [x
3c 4

(5.51)

in agreement with with the Larmor formula (5.26).

5.3

Gravitational Waves

By analogy with the generation of electromagnetic waves, changes in mass distributions lead to changes of the gravitational field. Since the adjustment to a new field
configuration can propagate with the speed of light, the universe must be filled with
gravitational waves. The collapse of stars, supernovae explosion, birth of neutron
stars, and similar dramatic events in the universe must all be accompanied by burst
of such waves whose general properties will now be studied.

5.3.1

Gravitational Field of Matter Source

The gravitational field is determined by Einsteins equation. That is derived by


extending Einstein-Hilbert action (4.364) for the gravitational field by the action of
m
all matter. Its action will be denoted by A and consists of a sum of the actions of
worldlines of point particles or of various fields. Under a variation of the metric,
m

using the property g = 21 gg g = 12 g g , we find from A the energymomentum tensor of all matter:
Z
m
m

A = d4 x gg (x) T (x).
(5.52)

391

5.3 Gravitational Waves

Varying the field action (4.364) yields




Z
1
1
4

d x g g g R + g R + g R
A =
2
2


Z

1
1
=
d4 x g g (R g R) + g R .
2
2
f

(5.53)

The last term vanishes in spaces without torsion2 , so that we obtain from (5.53) the
Einstein tensor G (x) defined in Eq. (4.378):
f

A =

d4 x g g (x)G (x).

(5.54)

If we finally extremize the total gravitational action


grav

A =A +A

(5.55)

with respect to g , we obtain the Einstein equation for the gravitational field in
the presence of matter:
m

G = T

(5.56)

This corresponds to
m

G = R = R = T (x).

(5.57)

We have seen in Eq. (4.415) that the free gravitational field equations is simplest
if written down in terms of the field = h 12 h in the Hilbert gauge (4.413).
Since the Einstein tensor in this gauge is given by (4.414), the linearized Einstein
equation (5.56) reduces to
m
1 2
(x) = T (x).
2

(5.58)

Note that this field equation (5.56) follows directly from the linearized field action
(4.375) by adding to it the linearized interaction following from (5.52):
A

int

d4 x h (x) T (x),

(5.59)

this being the analog of the interaction (5.1).


The solution of Eq. (5.58) involves again the retarded Coulomb Green function:
2
(x) =
4

m

1

dx
.
T (x )

|x x |
t =t(x x)/c
3

(5.60)

Far away from the source, in the radiation zone, we may approximate 1/|x x |
1/|x| 1/r and (x, t) behaves locally like an outgoing plane wave. Thus it must
2

See Section 15.2 in the textbook Ref. [13].

392

5 Classical Radiation

be purely transverse, i.e., some linear combination of the polarization tensors


H+
and
of
Eqs.
(4.426)
and
(4.427).
These
have
no
00-components
and
are
traceless
H
[recall (4.420)]. Thus only the spatially traceless part of will contribute and we
can approximate them by
ij ij

2
4r

d3 x [T

ij

(x , t |x x |/c) 13 i j T kk (x , t |x x |/c)].(5.61)

In this way we have merely modified the spatial trace of ij which cannot carry
away radiation energy since for large r the trace of ij can be gauged to zero due to
(4.408).
Note that there is also here a splitting into a velocity and an acceleration field
which carries off energy discussed for electromagnetism in Eq. (5.21).
In linear approximation, the energy-momentum term of the source is conserved
(i.e., neglecting to linear order the energy-momentum tensor of the gravitational
field)
m

= 0.

(5.62)

The neglected terms are of order O(). Hence


m

0 T

0 T

i0
00

(x, t) = j T

(x, t) = j T

ij

(x, t),

0j

(5.63)

(x, t).

(5.64)

From these equations we find that


Z

d x T

0j

(x , t) =

d3 x xi k T

= 0

0k

d3 x xi T

(x , t)

00

(x , t)

(5.65)

and further
Z

dx T

ij

(x , t) =
=
=

3 j

d x x k T
m

d3 x xi k T

ik

jk

(x , t) =

(x , t) =

d3 x xj 0 T
m

d3 x xi 0 T

i0

j0

(x , t)

(x , t)

m
1 Z 3 i m j0
0 d x [x T (x , t) + xj T i0 (x , t)]
2

(5.66)

and
Z

d3 x [xj T

i0

(x , t) + xi T

j0

(x , t)] =
= 0

d3 x xi xj k T
m

d3 x xi xi T

00

k0

(x , t)

(x , t).

(5.67)

Hence
Z

d3 x T

ij

1
(x , t ) = 0 2
2

d3 x xi xj T

00

(x , t ).

(5.68)

393

5.3 Gravitational Waves

5.3.2

Quadrupole Moment

This equation permits us to express the right-hand side of (5.61) in terms of the
quadrupole moment
1 Z 3 i j 1 2 ij m 00
Q (t)
d x x x r
T (x , t),
c
3


ij

of the mass distribution (x, t) = T

00

(5.69)

(x , t |x x |/c)/c as:

1 ij
Q (t )
.
(x, t) =
4r c
t =t rc

ij

(5.70)

The field components (x, t) determine the energy carried away by the wave.
For this we need the analog of the Poynting vector for linearized gravity. This is supplied by symmetric energy-momentum tensor that can be derived from the quadratic
action (4.375) by the techniques to be developed in Section 8. In Eq. (8.263) we
shall find the symmetric energy-momentum tensor in the Hilbert gauge = 0:
1
1
2
=
T
8
2
f
0i

If the T is multiplied with the unit vector of the outgoing wave k,


f

c2 T

0i i

k =



i
c2 h 0 i
2 0 i k i .
8

(5.71)

(5.72)

we obtain the energy current density along the k-direction.


Separating into
space and time parts, we can write
f
dE
= c2 T
2
r d

0i i

k =

c2 h 0 kl i kl
2 4 0 k0 i k0 + 0 00 i 00
8
i
+ 0 kk i 00 + 0 00 i kk 0 kk i kk k i .

(5.73)

This is to be integrated over the surface of a large sphere at r . Before doing


this we note that in the asymptotic fields we can forget all space derivatives of
the prefactor 1/r since they give nonleading 1/r 2, 1/r 3, . . . contributions. We do
have to keep, however, derivatives with respect to r arising from the retarded time
argument [recall the approximation (5.32)]. Hence we can approximate, to leading
order in 1/r,
i 0 i r = 0 k i .

(5.74)

Applying this to the components kl brings the first term in (5.73) to the form
2
2 kl , while i 00 becomes 0 00 ki . It is furthermore possible to express the time
derivatives on the right-hand side of (5.73) in terms of time derivatives of the purely
spatial field components (5.60), in the Hilbert gauge (4.413):
0 k0 = i ki = 0 ki k i ,

(5.75)

394

5 Classical Radiation

and
0 00 = i i0 = 0 i0 ki = j ij kj = ij ki kj .

(5.76)

Inserting for ij the asymptotic traceless expressions ij of Eq. (5.61), we finally


obtain the energy current becomes, to leading order in 1/r,
f
dE
2
=
c
T
r 2 d

0i i

k =

1
1
kl 2 2 kl km kl km + kl mr kk kl km kr .
4
2


(5.77)

The terms in brackets contain the transverse projection matrix (4.433), and can
therefore be written as contractions with the polarization tensors (4.426), (4.427),
as
kl |2 + |kl (k)
kl |2 = kl P kl,mn (k)
mn ,
|kl
(5.78)
TT
+ (k)

Inserting here the field (5.70), we obtain the rate of energy emitted per unit solid
angle d:
f
dE
= r2c T
d

0i i

k =

1 1 kl,mn
Q P
(k) Qmn .
4 (4)2 c2 kl T T

(5.79)

This has a complicated angular dependence. However, the integral over all directions
is easily found using the angular averages of products of all direction vectors
1
hki kj i = ij ,
3


1  ij kl
+ ik jl + il jk .
hki kj kk km i =
15

(5.80)

The tensor structure on the right-hand sides follow directly from the rotational
symmetry. The normalizations are found by a contracting of the indices, using
ni ni = 1. Inserting (5.80) into (4.429) yields

hPTij,kl
T (k)i


1
2 1  ik jl
+ il jk ij kl .
=
5 2
3


(5.81)

Now the angular integral is straightforward, and we find, recalling from (4.383),
1 1 2 kl 2
G kl 2
E =
(Q ) = 5(Q
),
2
4 4 c 5
5c

G
1

.
5
c
3.6 1052 W

(5.82)

For comparison, recall the analogous electromagnetic radiation formula (5.23) where
the direction-dependent energy loss due to dipole radiation is
dE
1
d)|
2,
=
| k (k
3
d
8c

(5.83)

with the total radiated power:


1 2
E = 3 d
3c

(5.84)

395

5.3 Gravitational Waves

The reader may wonder about the translational invariance of the gravitational
rediation formulas (5.82), since a quadrupole moment possesses a reference point.
Consider the second moment of the energy-momentum tensor after a translation
m

d3 x (xi ai )(xj aj ) T
ai

00

d x xj T

00

(x, t) =

(x, t) aj

d3 x xi xj T
m

d x xi T

00

00

(x, t)

(x, t) + ai aj

d3 x T

00

(x, t). (5.85)

We now observe that the last term is time-independent because of energy conservation. The other two terms in the second line may be rewritten using the conservation
law T (x, t) and a partial integration as follows:
02

d x xj T

00

(x, t) = 0
=

d x xj i T
m

d3 x k T

jk

i0

(x, t) =

(x, t) = 0

d3 x xj i k T
m

d3 x T

j0

ik

(x, t)

(x, t) = 0,

(5.86)

the last zero following from momentum conservation. Thus Qij (t) changes at most
by a linear function of t, and formula (5.84) is indeed independent of the choice of
the reference point for calculating the quadrupule moment.3
There exists another way of stating the radiation formula (5.79) that is more
similar to the electromagnetic formula (5.41). We use the relation (5.60) and the
transverse projection matrix (4.431) directly to replace (5.79) by
m
m

d2 E
kl,mn
=
[
(k,
t)]
P
(k)
T
T
t
kl
mn (k, t) .
T
T
d
16 2

(5.87)

Now we make use of the energy-momentum conservation law (5.63) in Fourier space
m
m
m
m
ki T i (k, ) = T 0 (k, ), ki kj T ij (k, ) = T 00 (k, ), and rewrite this in a fully
covariant form analogous to (5.43) as
m
2
m
2

dE
t T (k, t) 1 t T (k, t) .
=
2




d
16 2
(

(5.88)

For long wavelengths, the spatial components of the energy-momentum tensor


m
m
m
R
have a negligible dependence on k: T ij (k, t) T ij (k = 0, t) = d3 x T ij (x, t).
Then we can perform the integral over all angles in (5.88) using the angular averages
(5.80), yielding
2
E =
4 5

(

d xT

j (x, t)

2

Z
m
1
t d3 x T i i (x, t)

3


2 )

(5.89)

The right-hand side can be rewitten as


 



m
m
m
m
2 Z 3
d xt T i j (x, t) 13 i j T k k (x, t) t T j i (x, t) 13 j i T k k (x, t) .(5.90)
E =
4 5
3

For more subtle aspects of gravitational radiation see F.I. Cooperstock and P.H. Lim, Phys.
Rev. Lett. 55, 265 (1985).

396

5 Classical Radiation

As before we use (5.68) and (5.70) to replace


Z

dx T

ij

ij

(x, t) T
1
3

k (x, t)

1 ij
Q ,
2c

(5.91)

and obtain once more the flux equations (5.82).

5.3.3

Average Radiated Energy

If we express the time-dependent energy-momentum tensors in Formula (5.88) by


their Fourier integrals

dE
=
f (t, , )
d 16 2



(k, )

1
2

(k, )

 

(k,

)
1
2

(k,



) .

(5.92)

where

f (t, , )

d d i( )t
e
2 2

(5.93)

Over a long time, the oscillation of the difference frequency = cancel each
other and only the average rate of the integral
Z

t2

t1

dtei( )t 2( )

(5.94)

survives, implying that


hf(t; , )i =

1
2( ).
t2 t1

(5.95)

Integrating this over all times from t1 to t2 gives therefore the total radiated energy
per unit solid angle
dE

=
d 8

5.3.4



(k, )

1
2

(k, )

 

(k, )

1
2

(k, )



(5.96)

Relation to Gravitational Interaction Energy

Note that if we apply the projection matrix (4.433) to the energy-momentum tensor
T lm (k), we find the transverse part of the energy-momentum tensor
m ij
T TT

Tm lk
PTij,lm(k)

k i k m m jm 1 ij k l k m m lm 1 k i k i m ll
1 ij m ll
+
+
= T T 2 2 T
T
T
2
k
2
k2
2 k2
1 k i k j k l k m m lm
+
(5.97)
T .
2
k4
m

ij

397

5.3 Gravitational Waves


m
T

Its only nonzero components are the helicity components T


[recall (4.428)] where
m

(2)

(k) = [
H (k, 2)] T

=
H (k, 2) T

(k),

(2)

(5.98)

which for k pointing in the z-direction are equal to


m
m
m
1 m
(2)
(k) = [T 11 (k) T 22 2i T 12 (k)].
(5.99)
T
2
Let us now calculate the transverse interaction

X m ij
X
X  m
m
m
m
m ij
(2)
(2)
(2)
(2)
int
2
ij
(, 5.100)
(T ) T
+ (T
) T
AT =
( ) =
T TT =
2 TT
2
2
k k
k k
k k

where we have set /2, and


z-direction, this is equal to
Aint
T =

X
k

VT

dk 4 /(2)4. For k pointing in the

m
1 m11 m 22 2
(T T ) + 2(T 12 )2 .
2
k 2

(5.101)

Consider now the covariant interaction



X  m
X m m
X m
m
m
m
m
00 2
0i

0i
ij 2
(
[(T 00 )2 + (T 33 )2
)
2
=
Aint =
+(
=
)
T T
T
T T
T
2
2
2
k
k
k
k
k
k
m

2(T

01 2

) 2(T

02 2

) 2(T

03 2

) +2(T

12 2

) +2(T

23 2

31 2

) +2(T

) +(T

11 2

) +(T
m

22 2

) ].(5.102)
m

If k points in the z-direction, we use the conservation laws k 0 T 00 = k i T i0 to


m
m
m
m
m
2
2
replace 2(T 03 )2 by 2(k 0 /k2 )(T 00 )2 , 2(T 13 )2 by 2(k 0 /k2 )(T 01 )2 , and (T 33 )2 by
m
4
(k 0 /k4 )(T 00 )2 so that (5.102) becomes
int

A =

X
k

2 !2

k0
m 00 2
1

)
[(
T
k2
k2
m

+ 2(T

12 2

2[(T

) +(T

11 2

01 2

) + (T

) +(T

2!

k0
) ] 1 2
k

02 2

22 2

) ].

(5.103)

02

Now we replace 1/k 2 by 1/k2 (1 k /k2 ) and obtain


int

A =

X
k

2!

k0
1 m
2 (T 00 )2 1 2
k
k
"

2(T

01

m
m
1 m
+ 2 (T 11 )2 + (T 22 )2 + 2(T 12 )2
k

)2(T

)

02 2

(5.104)

If we now observe that for k pointing in the z-direction,


we can
use the relation
 m

m
m
m
m
m
33
02
2
00

02
2
00
1 k /k T 11 T 22 , and
T = (k /k ) T to replace the trace T by T
see that the transverse interaction (5.102) is is the same as
A

int

2!

m
m
m
m
m
k0
1 1 m 00 2
01
02 2
00
11
22
(
1

2(
)
)2(
)
+
(
+
)
=
T
T
T
T
T
T
2
2
2
k
k k

X 1 m m
1m m

+
(5.105)
T T ,
2 T T
2
k k

"

398

5 Classical Radiation

The first line is due to the instantanous gravitational energy of the fields h00 and
h0i which do not have a time derivative in the field action (4.381). If we recall
R
R
P
( /k2 )f 2 (k) = (/4 2 ) dk f 2 (k) = (/8) d3 xd3 x f (x)f (x )/|x x | =
that
R 3 k3
G d xd x f (x)f (x )/|x x |, we see that the second line, for a fixed direction of k
and on shell, i.e., for k 0 = |k|, is directly related to the radiated energy (5.96).

5.4

Simple Models for Sources of Gravitational Radiation

Let us calculate the radiated power for a few typical radiating systems. We shall
distinguish oscillating and bursting systems [1].
Consider a single nonrelativistic point particle moving along the orbit x(t), the
energy-momentum tensor has the spatial components
m

ij

(x, t) =

M i
j
(t)).
x (t)x (t) (3) (x x
c

(5.106)

Then Eq. (5.89) becomes


2 M 2
1
k
i j
E =
t x x ij x x k
2
4 5 c
3


2

(5.107)

Together with the energy, also the total angular momentum


Li (t) = ijk

d3 x xj T

0k

(x, t)

(5.108)

of the gravitational system decreases in time. This happens at a rate

il (t) Q
im (t) = 2G klm Q
il (t) Q
im (t).
Lk = 2 1 klm Q
2
5
8 5c
5c

(5.109)

For checking the dimensions of the above equations the following list of dimensions is useful:

cm2
,
h

[] =
Note that

5.4.1

, [E] =
, [p] =
, [Qij ] = h
sec,
4
cm
sec
cm
cm5
h

5
[G] =
,
[c
/G]
=
3.6 1052 W.
h
sec3
sec2

[T ] =

[A] = h
,

ij ] = h
[Q
,
sec2
(5.110)

ij has the dimension of a power.


Q

Vibrating Quadrupole

Imagine two equal masses M oscillating at the ends of a spring (see Fig. 5.1). The
masses have the time-dependent positive
!

d
z=
+ a sin t .
2

(5.111)

399

5.4 Simple Models for Sources of Gravitational Radiation

Figure 5.1 Two equal masses M oscillating at the ends of a spring as a source of gravitational radiation.

Assuming a d, we may approximate


z2

d2
+ ad sin t.
4

(5.112)

The quadrupole moment is therefore


4a
sin t qij
Qij (t) = 1 +
d


where
2

Qij (0) =
and

Md

(5.113)

(5.114)

22 (t) = 1 Q
22 (t) = 2Mad 3 sin t .
Q
(5.115)
2
3
According to Eq. (5.79), the angular distribution for a diagonal quadrupole moment
is
11 (t) =
Q

n 2
i
h
1
dE
+ Q
2 + Q
2 2 ( Q
11 n1 )2 + ( Q
22 n2 )2 + ( Q
33 n3 )2
=
Q
11
22
33
d
4 (4)2 c2

1
2

+ (Q
.
(5.116)
n
n
+
Q
n
n
+
Q
n
n
)
1 1
22 2 2
33 3 3
2 11

Introducing spherical angles


n1 = sin cos , n2 = sin sin , n3 = cos .

(5.117)

as illustrated in Fig. 5.1, the curly brackets in (5.116) become for Q11 = Q22


2 + Q
2 2( Q
2 sin2 + Q
2 cos2 ) + 1 [ Q
sin2 + Q
33 cos2 ]2 . (5.118)
2Q
11
33
11
33
2 11


400

5 Classical Radiation

For Q33 = 2Q11 , this reduces to


2 9 sin4
Q
11
2

(5.119)

The rate of energy radiation per solid angle d is then


1
dE
= 2
M 2 [ 3 ad sin t]2 .
d
2c (4)2

(5.120)

The radiation is maximal in the direction of the equation and vanishes in the pole
directions of oscillator.
Integrating (5.120) over all angles gives the total emitted power
1
32
M 2 [ 3 ad sin t]2
2
2
2c (4)
15
8G 2 3
M [ ad sin t]2
=
5
15c

E =

(5.121)

whose temporal average is

8G 2 6 2 2
E =
M ad .
15c5
The rate of radiation damping is defined as
rad

1
trad

(5.122)

1 dE
,
E dt

(5.123)

where trad is the damping time. Since the kinetic energy of each mass is M 2 a2 /2,
we obtain
46
rad =
Md2 4 .
(5.124)
15c5
The formula estimates the damping rate of any linearly oscillating system of two
masses. The linear character of the oscillation is important since for a spherically
symmetric pulsating star, there is no gravitational radiation at all.
Vibrational radiation may emerge from nova exploitations at an early stage.
These arise if a star circles around a white dwarf and transfers matter to him. After
some time, the matter becomes large enough to explode.
This explosion causes vibrations in the white dwarf with frequencies 0.01 to 1 Hz.
The energy released in a nova explosion is typically 1045 erg, of which 10% could be
deposited in vibrations, which send out gravitational radiation.

5.4.2

Two Rotating Masses

If the two masses in the previous example rotate around the z-axis (see Fig. 5.2)
the quadrupole moment (5.114) in the xy-plane becomes
Md2
Qij (t) =
4

1 3 cos 2t 3 sin 2t
3 sin 2
1 + 3 cos 2t

(5.125)

401

5.4 Simple Models for Sources of Gravitational Radiation

Figure 5.2 Two spherical masses in circular orbits about their center of mass.

where M is the reduced mass


M1 M2
,
M1 + M2

(5.126)

G(M1 + M2 )
.
r3

(5.127)

M
and is given by the third Kepler law:
=

The third time derivatives of the quadrupole moments are therefore

sin 2t cos 2t

2 3

0
0
Qij (t) = 6Md
.
cos 2t sin 2t

(5.128)

Inserting these into (5.116), integrating over all angles, and averaging over all times
yields the total emitted power
8G
E = 5 M 2 d4 6 .
5c

(5.129)

Using of Eq. (5.127), this becomes


32G4
E = 5 5 (M1 M2 )2 (M1 + M2 ).
5c d

(5.130)

The total energy of the binary system is


E=

1 M1 M2 2 2 1 GM1 M2
d =
,
2 M1 + M2
2
d

(5.131)

implying a rate of radiation loss is


=

64G4
M1 M2 (M1 + M2 )
5c5 d4

(5.132)

Since E is inverse proportional to the distance d by (5.131), the distance between


the masses decreases as follows:
64G3
d
= 5 4 M1 M2 (M1 + M2 ).
d
5c d

(5.133)

402

5 Classical Radiation

By Eq. (5.127) this implies that the frequency increases at a rate

3 d
= .

2d

(5.134)

Due to the smallness of the gravitational constant, the power radiated by planetary
systems is extremely small. The earth orbiting around the sun emits only 200 W,
the Jupiter 5300 W. For narrow double stars, the power can increase to 1030 W and
more, for doubleneutron stars for which d can be quite small, it can reach to 1045 W.
Table 5.1 shows various astronomical objects and the gravitational amplitudes
which can arrive from them at the earth.
Table 5.1 Binary systems as sources of gravitational radiation (from D. H. Douglass
and V. B. Braginsky Gravitational-radiation Experiments, in S. W. Hawking and W. Israel General Relativity (Cambridge University Press, Cambridge, 1979). The binary PSR
1913+16 emits radiation at multiples of 70 106 Hz due to the large eccentricity of the
orbit.

System

Masses

Dist.

(M )
(pc)
Eclipsing binaries
Boo
1.0, 0.5
11.7
Sco
12, 12
109
V Pup
16.5, 9.7 520
Cataclysmic binaries (novas)
AM CVn
1.0, 0.041 100
WZ Sge
1,5 0.12
75
SS Cyg
0.97, 0.83 30
Binary X-ray sources (black holes
Cyg X1
30,6
2500
PSR 1913+16 1.4, 1.4
5000

Wave
frequ.
(106 Hz)

Luminosity Flux at Amplitude


at earth
earth at earth
30
(10 erg/s) (1022 )

86
16
16

1.1
51
59

68.0
38.0
1.9

51
210
46

300
24
2
stars)
1.0
0.6
2.9
5.8

240
37
20

5
8
30

1
0.2
1.1
2.1

4
0.12
0.14
0.12

1900
410
84
or neutron
4.1
70
140
210

Formula (5.134) has been used as an indirect evidence for the existence of gravitational radiation. In 1974, Hulse and Tayler searched for pulsars (rotating neutron
stars emitting radio pulses) with the Arecibo telescope and found an object whose
emitted radio frequency is periodically modulated. The modulation is attributed to
the Doppler shift caused by the orbital motion around an undetected companion.
A careful analysis of the modulation allowed to derive the eccentricity and the rate
of the perihelion precession of the binary object (see Table 5.2 for details). A particularly interesting effect was the slow-down of the orbital motion. If denotes the
period, one finds

403

5.4 Simple Models for Sources of Gravitational Radiation

Table 5.2 Some observed parameters of PSR 1913+16 (Table taken from Ref. [6]).

Distance
Pulsar period (nominal)
Semi-major axis
Eccentricity
Orbital period
Rate of precession of periastron
Amplitude of time-dilation factor
Rate of decrease of orbital period
Rate of decrease of semimajor axis
Calculated lifetime (to final inspiral)
Diameter of each neutron star
Periastron separation
Apastron separation
Velocity of stars at periastron in CM frame
Velocity of stars at apastro in CM frame
Rate of precession of spin axis

21 000 ly
59.02999792988 ms
1 950 100 km
0.617131 0.000003
27907 0.00002 s
(4.22263 0.0003)o/y
0.0044 0.0001
2.4184(9)1012 s/s=0.0000765s/y
3.5 m/y
300 000 000 y
20 km
746 600 km
3 153 600 km
450 km/sec
110 km/sec
?

Figure 5.3 Gravitational Amplitudes arriving at the earth from possible sources. LIGO
denotes the Laser Interferometer Gravitational Wave Observatory (see http://www.ligo.
caltech.edu). LISA is the Laser Interferometer Space Antenna, a joint three-spacecraft
mission of ESA and NASA http://www.esa.int/esaSC/120376 index 0 m.html), where
the time-dependence of the distance of two objects 5 106 km apart will be monitored.
WDB denotes the regime of white dwarf binaries.

2.4184(9) 1012 sec per sec.

(5.135)

The observed shift of the time of periastron passage is plotted in Fig. 5.5. The values
of the masses were deduced from the perihel precession and the time delay of the

404

5 Classical Radiation

signal passing the companion. The first depends only on M1 + M2 , the M1 and M2
in a different combination. From these data, one deduces
M1 (1.4414 0.0002)M
M2 (1.3867 0.0002)M

for pulsar,
for companion.

(5.136)
(5.137)

Using these and the projected semi-major axis in Table 5.2 one finds a rate of change
of the period
theor 2.38 1012 sec per sec
(5.138)
in good agreement with the observed rate in Table 5.2.
The properties of binary objects containing a pulsar can be studied so well
that the approximation (5.130) for the radiated power need several corrections [7].
Consider two masses Figure 5.6 orbiting around the common center-of-mass. If
d denotes the distance of the two masses, the distances from the center-of-mass
are d1 = M2 d/(M1 + M2 ), d2 = M1 d/(M1 + M2 ). Denoting the reduced mass
M1 M2 /(M1 + M2 ). as before by M, the components of the quadrupole moment are
Qij (t) = Md

cos2
sin cos
sin cos
sin2

(5.139)

The orbit of a Kepler ellipse has the general form


d=

a(1 e2 )
,
1 + e cos

(5.140)
q

where a is the semi-major axis of the ellipse, e 1 b2 /a2 is the eccentricity (b


denotes the semi-minor axis), and the angular velocity is given by
q

(t)
= d2 (t) (M1 + M2 )a(1 e2 )

Figure 5.4 Two pulsars orbiting around each other.

(5.141)

405

5.4 Simple Models for Sources of Gravitational Radiation

Figure 5.5 Shift of time of periastron passage of PSR 1913+16 fore each orbit caused
by the shrinking of the Kepler orbits as a consequence of formula (5.147). Curve is from
theory (see Ref. [6]).

Figure 5.6 Two masses in a Keplerian orbit around the common center-of-mass.

From this we derive immediately


11 = P (1 + e cos )2 (2 sin 2 + 3e sin cos2 )
Q
22 = P (1 + e cos )2 [2 sin 2 + e sin (1 + 3 cos2 )]
Q
12 = P (1 + e cos )2 [2 sin 2 e sin (1 3 cos2 )] =
Q

21 ,
Q

(5.142)
(5.143)
(5.144)

where P is a power factor


P

v
u 3 2 2
u G M1 M2 (M1
2t

a5 (1

+ M2 )

e2 )5

(5.145)

Inserting (5.142)(5.144) into (5.116), integrating over all angles, and averaging over
all times yields the total emitted power
8G4 M12 M22 (M1 + M2 )
E =
[1+e cos (t)]4 {12[1+e cos (t)]2 +e2 sin2 (t)}, (5.146)
15c5
a5 (1 e2 )5

406

5 Classical Radiation

where t on the right-hand side is retarded with respect to the left-hand side. For
circular orbits where e = 0, a = d/2, this reduces properly to (5.130). Averaging
over one period of the elliptical orbit yields
73
37
32G4 M12 M22 (M1 + M2 )
1 + e2 + e4 .
E =
5
5
2
7/2
5c
a (1 e )
24
96


D E

(5.147)

With respect to a circular orbit of equal total energy, the power is enhanced by a
factor
73 2
37 4
1 + 24
e + 96
e
f=
,
(5.148)
(1 e2 )7/2
which grows rapidly from f (e) = 1 at e = 0 to infinity at e = 1. For a full multipole
analysis of the radiation see [7].
Due to the shrinking of the Kepler orbits implied by formula (5.147), the orbital
time shrinks according to the elliptic generalization of Eq. (5.134), implying that the
periastron is reached a few seconds earlier for each orbit. The time shift is plotted
in Fig. 5.5.
The radiation properties of binary objects can be studied especially well if both
stars are pulsars. Such an astronomical object has recently been found. One of the
two pulsars rotates with a period of 23 milliseconds (PSR J0737-3039A) around its
axis, the other with a period of 2.8 seconds (PSR J0737-3039B) [8] (see Fig. 5.4).

5.4.3

Particle Falling into Star

Among the bursting sources of gravitational radiation, the simplest one consists of
a mass falling into a star as shown in Fig. 5.7. If the mass starts at z = , its
velocity is
1 2 GmM
mz =
,
(5.149)
2
z
so that
1
GM
(2GM)3/2
z = 1/2 (2GM)1/2 , z = 2 ,
z =
.
(5.150)
z
z
z 7/2
The triple time derivative of the quadrupole moment
z 2 0
0

2
0
Qij = m 0 z

0
0 2z 2

is

(5.151)

1 0 0

Qij = m(6z
z + 2z z) 0 1 0
,
0
0 2

(5.152)

and thus, because of (5.150):

1 0 0
3/2

ij = m (2GM)
Q
0 1 0 ,
5/2
z
0
0 2

(5.153)

5.4 Simple Models for Sources of Gravitational Radiation

407

Inserted into (5.82), this leads to the energy loss per second
2Gm2
E =
(6z
z + 2z z )2
15c5
Combining this this with (5.150) implies
dE/dz =

1 2Gm2
(2GM)5/2 dz.
z 9/2 15c2

(5.154)

(5.155)

The radiated energy from z = to z = R is


E=

1 4Gm2
(2GM)5/2 .
R7/2 105c5

(5.156)

Obviously, the radiated energy increases with decreasing R. Suppose the large object
is a black hole. If we let the mass m fall down to the Schwarzschild radius R
Rs = 2

GM
,
c2

(5.157)

we obtain

2
m
m
mc2
0.019mc2 .
(5.158)
105
M
M
If one takes relativistic effects into account, as the particle approaches Rs , the number 0.019 changes to 0.01044.
For a black hole of radius M 10M with Schwarzschild Rs 30km, the total
radiated energy is
E 2 1051 erg.
(5.159)
E=

The radiated energy is mostly emitted at the end (see Fig. 5.7 for the curve resulting
from a detailed analysis).

5.4.4

Cloud of Colliding Stars

Let us treat the stars in a cloud approximately as point-like objects moving with
velocities vn , where they have an energy-momentum tensor [?]
T =

X
n

p p (3)
(x x(t)).
En

(5.160)

Suppose the stars all collide at the origin at t = 0 and run away with changed
n . Then the time-dependent energy-momentum tensor is
velocities v
T (x, t) =

X
n

X p
pn pn (3)
n (3)
np
n t)(t),
(x vn t)(t)
(x v
En
En
n

M. Paves, R. Ruffini, W. H. Press, R. H. Price, Phys. Rev. Lett. 27, 1466 (1971).

(5.161)

408

5 Classical Radiation

Figure 5.7 Spectrum of the gravitational radiation emited by a particle of mass m falling
radially into a black hole of mass M . The quantity dE/d gives the amount of energy
radiated per unit frequency interval. The curve marked l = 2 corresponds to quadrupole
radiation; the other curves (l = 3, l = 4) correspond to multipole radiation of higher order.
Note that most of the radiation is emitted with frequencies below 0.5c3 /GM [9].

Figure 5.8 Particle falling radially toward a mass.

409

5.4 Simple Models for Sources of Gravitational Radiation

Table 5.3 Typical Astrophysical Sources of Gravitational Radiation. Distances have been
selected large enough to yield approximately three events per year (from K. S. Thorne,
Gravitational Radiation, in S. W. Hawking, and W. Israel, eds., Three Hundred Years of
Gravitation, (Cambridge University Press, Cambridge, 1987). ).

Source
Periodic sources
Binaries
Nova
Spinning neutron star (Crab)
Bursting sources
Coalescence of binary
Infall of star into 10M b.h.
Supernova
Gravitational collapse of 104 M star

Frequency

Distance

Amplitude
(A)

104 Hz
102 to 1
60

10 pc
500 pc
2 kpc

1020
1022
< 1024

10 to 103
104
103
101

100 MPC
10 Mpc
10 kpc
3 Gpc

1021
1021
1018
1019

where (t) is the Heaviside function defined in Eq. (1.310). Representing this as a
Fourier integral
Z
d eit
,
(5.162)
(t) = i
2 i
and the -function similarly as
(3)

(x v) =

d3 k ikx
e ,
(2)3

(5.163)

we see that the energy momentum tensor (5.161) has the Fourier components

T (k, ) =

X
n

pn pn
i
p p
i
.
n n
En vn k i
En vn k + i
!

(5.164)

Now we observe that


En (vn k ) = pn k,

(5.165)

where k is the four-momentum of the emerging gravitational wave. Since the stars
are nonrelativistic objects, we can further drop the is. Thus
T

(k, ) = i

X
n

pn pn pn pn
.

pn k
pn k
!

(5.166)

This energy-momentum tensor satisfies k T = n (pn pn ) = 0, as it should


by momentum conservation. Inserting (5.166) into (5.96) yields the total radiated
energy per solid angle and frequency insterval d:
P

dE
1
2X
n n
(pn pn )2 Mn2 Mn2 ,
=

dd
16
2
nn (pn k)(pn k)


(5.167)

410

5 Classical Radiation

where the index n runs over the particles before and n is equal to 1 before, and 1
after the collisions. Integrating thos over all directions of kn yields
2
dE
1 + nn
2X
1 + nn
1

M
M
=

log
n
n
n
n
2
1/2
dd
16 2 nn
(1 nn
1 nn
)
nn

(5.168)

where nn is the relative velocity divided by c:


nn

Mn Mn
1
(pn pn )2

!1/2

(5.169)

The integral over all frequencies diverges. This is due to the fact that we have
assumes an instantaneous change of momenta during the collisions. In actual collisions, the change takes place over some finite collision time t and the integral
only contain frequencies up to 1/t. For nonrelativistic two-body scattering, the
radiated energy is
2 2
dE
=

d
5

d 4 2
v sin ,
2

(5.170)

where is the reduced mass, v are the relative velocities, and the scattering angles
in the center-of-mass frame.

5.5

Orders of Magnitude of Different Radiation Sources

In order to have an idea as to how much energy can be radiated in various precesses,
consider a massive steel rod of radius 1m, length d=20m, mass M 4.9108 g (=490
tons, using the steel density 7.89g/cm3 ). If the maximal quadropole radiation can
be obtained by rotating it around an axis orthogonal to the rod with an angular
velocity . The formula (5.82) yields the total emitted power
E =

2
M 2 l4 6 .
45c5

(5.171)

The angular velocity is limited by the tensile strength, which is t 3 10g dyn/cm2 .
Thus one can maximally use an angular velocity
=

8t
l2

!1/2

28

1
.
sec

(5.172)

This gives a radiated power


E 1023 erg/sec.

(5.173)

In order to have an idea how small this is we note that a single photon in the visible
range has an energy
h
c

2
2
1027 1010
erg 1.5 1012 erg.

4 105
4000A

(5.174)

5.5 Orders of Magnitude of Different Radiation Sources

411

Thus the radiated power corresponds to the emission of one visible photon in 10
seconds or 3000 years.
If one wants to have any observable effects it is therefore necessary to look for
radiation emitted by large stellar objects. Consider a bunch of stars of total mass M
distributed over a region of size R. their velocity is of the order R/T where T is the
time it takes for the masses to move from our side to the other. Their quadrupole
moment is of the order of
Q R2 M.

(5.175)

Hence we can estimate


2
R M M R
Q
T3
T

2

1
.
T

(5.176)

The right-hand side has the dimension energy per time. It gives an estimate for the
internal power flow in the system. The radiated power is equal to
1
E 5
c /G

R2 M
T3

!2

(5.177)

The denominator term c5 /G is itself a power [compare (5.110)]


erg
c5
3.63 1059
.
G
sec

(5.178)

It is called the Planck power . Thus, radiated power is of the order of the internal
power square divided by the Planck power c5 /G.
Note that if a system is to experience sizable radiation damping, it has to have
internal power flow of the order of c5 /G. This is an immense power. It corresponds
to a kinetic energy which is generated by burning up 2 105 solar masses per second
via nuclear precesses.
A sizable radiation can reach us only from stellar catastrophies. A mass m falling
into a black hole of mass M radiates off a total energy
m
E 0.0104
mc2 .
(5.179)
M
As an example, consider a star with solar mass falling into a black hole whose mass
is 10 times the solar mass. This gives
E 2 1051 erg.

(5.180)

This energy is radiated during a time [11]


t 103sec.

(5.181)

Hence the radiated power is of the order


m2 c5
E 2 .
M G

(5.182)

412

5 Classical Radiation

The most likely sources of detectable gravitational radiation are supernova explosions. They have been estimated to emit up to 1054 erg/sec.
The youngest known pulsar of the Crab Nebuluae is believed to be a neutron star
in fast rotation. Its gravitational radiation has been estimated to be of the order
of 1038 erg/sec, if it is deformed by 0.001 from a spherical shape. This radiation
would be necessary to explain the observed slowdown of the rotation frequency by
gravitational radiation damping. It is not clear, however, whether a neutron star
can be deformed to such an extent.
Compare this with the radiation emitted from binary stars. For a circular orbit
their kinetic energy is
Ekin =

1 M1 M2
.
2 R

(5.183)

The power output is


32 2 M 3
E =
5 R5

(5.184)

where = M1 m2 /(M1 + m2 ), M = M1 + M2 . The energy loss due to gravitational


radiation will lead to the two stars spiraling together within a time
tspiral =

5 R4
.
256 M 2

(5.185)

For the sun-jupiter system, this time is 2.51023 years. But for the binary system
PSR 1913+16, the spiral time shrinks to 3 108 years, as shown in Table 5.2, which
makes it observable, as we have seen in Fig. 5.5.
This makes it almost impossible to observe this radiation damping of the Kepler
orbits.
For neutron stars a few thousand km apart, the spiral times could shrink to the
order of years or days, so it could be observable at least, in principle. Here the
problem lies in the identification of the source.

5.6

Detection of Gravitational Waves

In order to detect gravitational waves we consider a test particle whose equation of


motion reads
1
du
= b h hbc u u.
d
2


(5.186)

If it is initially at rest or moving very slowly, the right-hand side reduces to


1
0 h 0 h00 c2 .
2


(5.187)

413

5.6 Detection of Gravitational Waves

Since the two physical polarization tensors have neither a 00 nor 0i components,
this vanishes. Thus two particles retain their relative positions as seen from the
background Minkowski frame.
This does not mean that their physical distance remains unchanged. This distance is evaluated not via the Minkowski metric ab , but via the proper slightly
distorted metric gab = ab + hab . If a wave with a linear polarization + (k) runs
along the z-axis
1
hab = [ (1)b (1) (2)b (2)] a+ ei(kzt) + c.c.,
2

(5.188)

and hits two particles at (0, d0/2, 0, 0) and (0, d0 /2, 0, 0), which initially do not
move due to the vanishing of du /d , their spatial distance changes as follows

d2 = (0, d0 , 0, 0)(ab + hab )


= d0

"

a
1 cos(t )
2

0
d0
0
0

(5.189)

where is the phase of the wave amplitude a+ .


For particles which are separated by d0 along the y axis, the sign in front of a+
is the opposite. In general, a pair of mass points at positions
d0
d0
0, cos , sin , 0
2
2

(5.190)

has a distance
2

d =

d20

"

a
1 cos(2) cos(t ) .
2

(5.191)

We can picture the change in the distance by imagining a circular necklace of mass
points placed in the gravitational beam. If the momentum points orthogonal to
the paper plane, the circle distorts into vertical and horizontal ellipses, as shown in
Fig. 5.9. This is why we have chosen the notation t for the polarization tensor
of this wave. The distortions are of quadrupole character, and the area within the
necklace remains invariant, due to the tracelessness of
+ (k) (since gab = ab + hab

2
has the determinant g = 1 + O(a ) implying the volume element d3 x = gd3 x
is invariant, to lowest
order in a). For a wave with polarization tensor (kzab ) =
[ (1)b (2) + (a b)]/ 2, the distance changes with time as follows
d2 (t) = d0 2

"

a
1 sin(2) cos(t ) .
2

(5.192)

The situation is the same as before, except for a rotation /4. Thus
the necklace undergoes the same quadruple distortions, except that the principal

414

5 Classical Radiation

Figure 5.9 Distortions of a circular array of mass points by the passage of a gravitational

quandrupole wave. The left is caused by a polarization tensor


+ , the second from .

axes of the ellipses lie along the diagonals as indicated by the subscript of the
polarization tensor. The rotation by 450 can also be displayed more directly by
writing the polarization tensor
H of Eq. (4.427) as a combination
1
1

H = [ (1) (2) + (a b)] = [ () () () ()] .


2
2

(5.193)

where
1
() = [ (1) + (2)],
2

1
() = [ (1) (2)]
2

(5.194)

are the diagonal polarization vectors. Similarly we rewrite


H+ (kz ) of (4.426) as
1
1

H+ (kz ) = [ (1)b (1) (2)b (2)] = [ () () + () ()] . (5.195)


2
2

The right-hand sides of (5.193) and (5.195) have the same forms as
H+ and H
expressed in terms of (1) and b (2), but with (1) and b (2) exchanged by the
450 -rotated diagonal polarization vectors () and ().
The acceleration

1
a
d2 d
= d0 2
2
dt
d
2

cos 2
sin 2

cos(t )

(5.196)

implies the presence of tidal forces acting upon the necklace. Their field lines are
shown in Fig. 5.10. If gravitational waves of helicity 2 hit the necklace, it is
deformed in to an ellipse with a fixed shape which rotates clockwise or counterclockwise around the direction of the wave. The wave is circularly polarized. This
is seen by taking again two particles at positions
!

d
d
0, cos , sin , 0 ,
2
2

(5.197)

415

5.6 Detection of Gravitational Waves

Figure 5.10 Field lines of tidal forces of a gravitational wave in the z direction with
ab
polarization tensor ab
+ and , respectively. The field lines change direction with cos t.

and measuring their distances in the metric


h

g = + (+) (+)ei(kzt) + c.c.


which gives
n

(5.198)

d2 (t) = d0 2 1 (cos2 sin2 ) + i2 cos sin aeit + c.c.


= d0 2 [1 |a| cos(t 2 )] ,

(5.199)

implying the above described rotations of the neclace, the azimuthal angle acting
merely as a phase shift.
The time-dependent of the length measured by the metric g = + h has
a direct experimental equivalence. If we take a piezoelectric crystal, then it shows
a pulsating voltage due to the distance changes between the atoms, even through
its atoms remain at rest in the Minkowski coordinates. The distance is given by
the minima of the interatomic potentials. Since the atomic interactions are due to
electromagnetism which spreads through a space with the metric g , the distances
of these minima change according to changes in g , and this gives rise to the
piezoelectric voltage.
How large are the distortions caused by a gravitational wave? If we assume a
typical astrophysical source (for the emission mechanism see the next section) with
an energy flux of 1010 erg/(cm2 sec) at 104 /sec, we calculate the distortion of
the metric to be of the order
h 107 .

(5.200)

This will make the distance earth-moon, which is 3.8 1010 cm, oscillate by
107 cm = 10
A. This distance is of the same order as the distance between
the atoms in matter which is of the same order.
At present, a laser pulse ranging to the moon can at most detect length changes
of the order of 10 cm.

416

5 Classical Radiation

Other possible observable effects are quadrupole vibrations which can be excited
by incoming gravitational waves within the earth or the moon itself. Their natural
frequencies are 54 minutes and 15 minutes, respectively. By studying seismometer
data of the earths vibrations, Weber found in 1967 an upper limit for the flux of
gravitational waves
energy flux
erg
< 3 107 2
frequency
cm sec Hertz

(5.201)

at a frequency 3.1 104 Hertz.


In 1972, J. Weber5 built a gravitational detector consisting of a cylindrical aluminum block of 1.53 m length and 0.66 m diameter (weight 1.41 106 g). The
block has an eigenfrequency of 1.66 Hz. By a piezoelectric strain transducer, Weber measured length changes in the material of the block. Setting up on block at
the University of Maryland and another one at Argonne Mational Laboratory and
looking at coincidences between the two, he eliminated random vibrations caused
by the daily activities in the neighborhood of each detector.
In 1972, he observed two sudden simultaneous excitations which he interpreted
as a possible response to a gravitational wave passing through.
Unfortunately, his observations have, until now, not found any recurrence in
spite of collective efforts of several laboratories. A new approach seems necessary
to become sensitive to such small effects.

5.7

Attractive Gravity versus Repulsive Electric Forces


between Like Charges

The energy of gravitational waves gives a simple insight why the fields lead to an
attraction between masses (which are always positive) while electromagnetism is
repulsive between line charges. The physical components of gravitational waves are
the purely spatial ones ij . Their energy has to be positive and hence the field
energy carries plus sign when expressed in momentum space
Egrav k2 (ij )2 .

(5.202)

In electromagnetism, the same is true for the spatial part of the electromagnetic
field Ai
2

Eelm k2 (Ai ) .

(5.203)

As a simple consequence of Lorentz invariance, the components 0 and A0 have to


appear with opposite signs in E:
Egrav k2 00

2
2

Eem k2 A0 .

See http://en.wikipedia.org/wiki/Joseph Weber.

(5.204)

417

5.8 Nonlinear Gravitational Waves

But these components are the relevant ones coupling to the mass density T 00 or the
charge density j 0 , respectively, thereby giving rise to Newtons or Coulombs law.
The opposite signs cause the forces in the opposite direction.

5.8

Nonlinear Gravitational Waves

Plane gravitational waves are such simple phenomena that they can easily be found
as solutions to the full nonlinear Einstein equations in the vacuum
G = 0.

(5.205)

If the wave runs in z-direction we can make the 2 dimensional Ansatz for the invariant distance
(ds)2 = (dt)2 L2 [e2 (dx)2 + e2 (dy)2] (dz)2

(5.206)

where L and are functions depend only on z and t. It is useful to go to so-called


light cone coordinates
u = t z,
v = t + z,

(5.207)

which sit on wave crests moving along the positive and negative z-directions. It can
easily be verified that the only non-zero component of the Ricci tensor is
Ruu = 2L1 (L + 2 L)

(5.208)

so that the nonlinear wave equation in empty space reads


L + 2 L = 0.

(5.209)

Our previous linear waves correspond to the equation


L 0,

(5.210)

L 1.

(5.211)

ds2 (1 + 2)dx2 (1 2)dy 2 dz 2

(5.212)

0 0
0 2
2 0
0

(5.213)

which can be solved by

In this limit,

corresponding to

hab =

0
0
0
0 0

418

5 Classical Radiation

i.e., a wave with a polarization tensor + (kz )ab .


Going back to the nonlinear case, we assume the space to be flat before the wave
arrives, say at
t z = T

(5.214)

L = 1 for u < T.

(5.215)

i.e., we take
= 0,

We shall assume that the wave comes as a pulse of width 2T and assume that,
after the pulse has passed, space is left again flat. While the pulse passes, i.e., for
T > u > T , we allow for an arbitrary (u) but assume, for simplicity, that the
pulse is not sharp:
| (u)|

1
.
T

(5.216)

Then we do not have to solve the full differential equation. While the pulse is
passing, we have
(u) = arbitrary, with | |
L(u) = 1

du

1
T

(5.217)

du [ (u )]2 + O(( T )4 ) foru (T, T ).

After it has passed, (u) is again zero by assumption, and


L(u) = 1

u
a

a RT

1
+ O(( T )2 )
2
du

for u > T.

(5.218)

The change in L from 1 to u/a is physically not observable. The space is flat after
the pulse has passed. This is seen by going to new coordinates X, Y, U, V defined by
x=

X
Y
1 X2 + Y 2
, y=
, u = U, v = V +
1 U/a
1 U/a
a 1 U/a

(5.219)

which brings the invariant distance to the Minkowski form


ds2 = dUdV d2 (dY )2 .

Appendix 5A

(5.220)

Nonexistence of Gravitational Waves in


D=3 and D=2 Spacetime Dimensions

The counting procedure of physical degrees of freedom has an immediate consequence for the
existence of gravitational waves in a hypothetical lower-dimensional world. In three dimensions,
the symmetric tensor has six independent components, three of which are eliminated by the
Hilbert condition = 0. This leaves only three components. These, however, are just as many
as there are gauge degrees of freedom
+ + .

(5A.1)

Appendix 5A

Nonexistence of Gravitational Waves for D = 3 and D = 2

419

In fact, all field degrees of freedom are gauge degrees of freedom. We see this by choosing k in the
z-direction, where polarization vectors can be written as
(l) =
(l ) =
(1) =

1
1 1

k = (1, 0, 1),
2 |k|
2
1
(1, 0, 1),
2
(0, 1, 0).

(5A.2)

They allow us to form the three symmetric polarization tensor which satisfy the Hilbert condition
(l) (l)
(1) (l) + (a b)
(1) (1).

(5A.3)

The first is proportional to k k , i.e., to a pure gauge of the form (5A.1) with = k . The second
is a pure gauge with = (1). The third, finally, is equal to
[ (l) (l ) + (l ) (l)],
(5A.4)

which has the pure gauge form (5A.1) with = (l )/ 2|k|.


This implies that the three-dimensional gravitational field has no dynamical degrees of freedom.
There exist no freely propagating gravitational waves in three spacetime dimensions.
A three-dimensional Einstein theory would have a further disease. It would not even possess
a Newtonian weak-field limit. To see this, suppose we had found a field from the equation
m

2 = 2 T .

(5A.5)

For the solution , we calculate in D dimensions


h =

1
.
D2

(5A.6)

In order for the weak-field limit to satisfy Newtons equation of motion it is necessary that for a
massive point particle of the origin, h00 satisfies the Poisson differential equation
2 h00 = GM (3) (x).

(5A.7)

Moreover, T has only a T 00 -component appearing in (5A.7). Hence has only a single
nonvanishing component 00 . In three dimensions this implies that
h00 = 00

1
D 3 00
00 =
.
D2
D2

(5A.8)

Bot 00 must satisfy (5A.5)! Hence the Newton potential vanishes identically.
Another equivalent place where this disease manifests itself is in the coupling of the gravitational field to the energy-momentum tensor of a massive particle. Since this coupling is
H T

(5A.9)

we see that
h T =






ab

T = T
T .
D2
D2

(5A.10)

With the particle being at rest and Tab having only a 00-component, the interaction becomes


00
00
T00
(5A.11)

T00 ,
D2

420

5 Classical Radiation

which vanishes for D = 3.


Only by a technical trick can we obtain a non-zero limit. The theory has to be defined for
continuous dimensions D in the neighborhood of 3 with a coupling constant which diverges at
D 3:
= 0 /(D 3).
The above-described diseases are not a consequence of the linear approximation to gravity.
We have mentioned before that in three dimensions, the full curvature tensor R is completely
determined in terms of the Ricci tensor
R = G

1
g G,
D2

(5A.12)
m

and thus in terms of the Einstein tensor G = T . Due to Einsteins equation G = T , R


and thus R vanishes identically everywhere, except right at the mass point. This implies, in
particular, that the empty-space field equation
G = 0,

(5A.13)

which we have used in four dimensions to find the Schwarzschild metric, can have only the trivial
solution g = , up to a trivial reparametrization of space.6
It is curious to note that this disease makes it possible to develop a quantum theory of gravity,
in contrast to four dimensions where such a theory does not yet exist. It is possible to define a
wave function for the universe. In the absence of matter, the entire Hilbert space consists only of
one state, the vacuum |0i.7
Let us take a look at gravitational waves in two dimensions. There the only polarization tensor
satisfying Hilberts constraint is
(k) =

1
k k ,
|k|2

(5A.14)

and this is obviously a pure gauge. As far as the equation (5A.6) is concerned, the situation is
even worse than in three dimensions. It is impossible to recover from the metric gab since the
equation
hab = ab

ab

D2

(5A.15)

is meaningless for D = 2. In order to see the origin of the problem let us choose another than the
Hilbert gauge, and use the gauge freedom
h00
h01

h00 + 20 0
h01 + 0 1 + 1 0

(5A.16)

to make h00 and h01 vanish. For the remaining component h11 h we find from (4.375) the
Lagrangian density:

f
1
(h)2 2(1 h)2 + 2(1 h)2 (h)2 = 0
(5A.17)
L =
4
which vanishes identically. This property could, in fact, have been anticipated. It is well known in
differential geometry that in two dimensions the Einstein action
Z
f

1
(5A.18)
d2 gR
=
A
2
6

For a detailed discussion see S. Giddings, J. Abbott, and K. Kuchar, Rel. Grav. 16, 751 (1986);
see also S. Deser, R. Jackson, and S. Templeton, Am. Phys. (NY) 140, 372 (1982).
7
See H. Leutwyler, Nuovo Cimento 55, (1960); Phys. Rev. B 134, 1755 (1964).

Appendix 5A

Nonexistence of Gravitational Waves for D = 3 and D = 2

421

is a pure surface term. By the Gauss-Bonnet theorem, it is entirely determined by the global
topological properties of the space. For closed surfaces, it is equal to (4/2)(1 h), where h is
the number of handles of the surface. This makes it impossible to derive equations of motion from
such an action.
For completeness, let us compare the situation with the electromagnetism case. In three
dimensions, the vector potential has three components minus one, due to the Lorentz condition.
This leaves two degrees of freedom. One of them is a pure gauge mode, the other is physical.
Hence there exists a freely propagating photon in three dimensions.
In two dimensions, the Lorentz gauge allows A to be written in the form of a two-dimensional
curl:
A = .

(5A.19)

In empty space, it satisfies the free field equation


2 A = 0.

(5A.20)

In terms of the field , the field strength is


F 01 = 0 A1 1 A0 = 2 .

(5A.21)

Note the relation with the theory of complex functions. For zero field, is a harmonic function.
It can therefore be considered as the real part of an analytic function
(x) = Re f (z),

(5A.22)

z = x + iy.

(5A.23)

where

The vector A is the real part of the gradient of this function.


A = Re f (z).

(5A.24)

Now, the real and imaginary part of f satisfy the Cauchy-Riemann differential equations,
1 Re f

2 Im f

1 Im f

2 Re f.

(5A.25)

Hence we can also write


A = Im f.

(5A.26)

This shows that A is a pure gauge which, therefore, carries no electromagnetic field, in agreement
with the assumption we started out from.
It is important to point out that the nonexistence of propagating electromagnetic waves does
not rule out Coulomb forces. They do not need the exchange of physical electromagnetic waves. The
exchange of virtual electromagnetic fields with zero momentum is sufficient. The inhomogeneous
equation
1
2 A = j
c

(5A.27)

has, outside the charge distribution, only the solution 2 A = 0, and we have seen before that the
only field satisfying this is a pure gauge field. Still, the Maxwell equation
1 F 01 =

1 0
j
2

(5A.28)

422

5 Classical Radiation

allows for a nonvanishing constant field


F 01 = const.

(5A.29)

It is carried by the k = 0-components of A to which the previous polarization discussions do not


apply. It is this component which gives rise to the Coulomb force in two space-time dimensions.
Another way to see this is by considering the Maxwell action
Z
Z
1
1
2

A=
d x Fab F + 2
d2 x j A ,
(5A.30)
2c
c
and going to the Lorentz gauge,
A = b

(5A.31)

in which it becomes (up to surface terms)


Z
Z
1
1
2
2 2
A=
d x( ) + 2 d2 x ( jb ).
2c
c

(5A.32)

Extremizing this in the field gives the field equation


( 2 )2 =

1
jb .
c

(5A.33)

Reinserting this into the action, the extremum is found to be


Z

1
1
Aextr = 2
d2 x j 2 2 j .
c
( )

(5A.34)

Using

= ,

(5A.35)

performing a partial integration, and taking advantage of current conservation j = 0, this


becomes
Z
1
1
Aextr = 2
(5A.36)
d2 x j 2 j
c

which is precisely the Biot-Savad interaction law between currents. Using once more current
conservation in the form


Z
Z
1
1
2
02
12
2
02
1
1 j
d x(j j ) =
d x j 1 j
(1 )2


Z
1
1
2
0
0
0
0
=
d x 1 j
(5A.37)
1 j 0 j
0 j
(1 )2
(1 )2
Z
2
j0
=
d2 xj 0
(1 )2
we can rewrite (5A.36) as
Aextr

1
= 2
2c

d2 xj 0

1
j0.
(1 )2

(5A.38)

This is an instantaneous linear potential between the charges carried by j0 . It is due to the constant
electric field F allowed in empty space.
In this respect, the situation is quite different from the gravitational case in three dimensions.
There the absence of R outside a mass distribution implies a vanishing of the gauge invariant
curvature tensor R and hence the vanishing of all tidal forces.

Notes and References

423

Notes and References


[1] A thorough discussion of various radiating systems can be found in the textbook
H. Ohanian and R. Ruffini. Gravitation and Spacetime, W.W. Norton, N. Y., 1994.
[2] For the scalar tensor theory see
C. Brans and R.H. Dicke, Phys. Rev. 126, 925 (1961).
[3] R.A. Porto, Post-Newtonian corrections to the motion of spinning bodies in NRGR, grqc/0511061.
[4] W. Goldberger and I. Rothstein, hep-th/0409156. For a recent review see I. Rothstein, TASI
lectures on effective field theories, hep-ph/0308266.
[5] R.V. Wagoner and C.M. Will, Astrophys. J. 210, 764 (1976) [Erratum-ibid. 215, 984
(1977)];
M. Walker and C.M. Will, Phys. Rev. Lett. 45, 1741 (1980);
L. Blanchet, T. Damour, and G. Schafer, Mon. Not. Roy. Astron. Soc. 242, 289 (1990).
[6] J.M. Weisberg and J.H. Taylor, Relativistic Binary Pulsar b1913+16: Thirty Years of Observations and Analysis, in Proceedings of Aspen Winter Conference on Astrophysics: Binary Radio Pulsars, Aspen, Colorado, 11-17 Jan 2004. ASP Conf.Ser. (to be published)
(astro-ph/0407149).
[7] P.C. Peters and J. Mathews, Gravitational Radiation from Point Masses in a Keplerian
Orbit , Phys. Rev. 131, 435 (1963).
[8] M. Burgay, N. DAmico, A. Possenti, R.N. Manchester, A.G. Lyne, B.C. Joshi, M.A.
McLaughlin, M. Kramer, J.M. Sarkissian, F. Camilo, V. Kalogera, C. Kim, D.R. Lorimer,
The Highly Relativistic Binary Pulsar PSR J0737-3039A: Discovery and Implications,
(astro-ph/0405194),
A.G. Lyne, M. Burgay, M. Kramer, A. Possenti, R.N. Manchester, F. Camilo, M.A.
McLaughlin, D.R. Lorimer, N. DAmico, B.C. Joshi, J. Reynolds, and P.C.C. Freire. A
Double-Pulsar System - A Rare Laboratory for Relativistic Gravity and Plasma Physics,
Science 8, January 2004.
For animations see the internet page http://www.jb.man.ac.uk/news/doublepulsar.
[9] Figure is taken from
M. Davis R. Ruffini, W.H. Press, and R.H. Price, Phys. Rev. Lett. 27 , 1466 (1971).
[10] See more details in
S. Weinberg Gravitation and Cosmology, John Wiley and Sons, New York, 1972.
The gravitational red shift of the sun has been claimed to be 1.5 .05 the theoretical value
by J. Brault, Bull. Am. Phys. Soc. 8, 28 (1963)
J.E. Blamont, F. Roddier, Phys. Rev. Lett. 7, 437 (1961).
The terrestial red shift was measured by R.V. Pound and G.A. Rebka, Phys. Rev. Lett. 4,
332 (1960), R.V. Pound and J.L. Snider, Phys. Rev. Lett. 13, 539 (1964).
[11] R. Ruffini, W.H. Press, and R.H. Price, Phys. Rev. Lett. 27, 1446 (1971).
[12] S. Deser, Gen. Rel. Grav. 1, 9 (1970).
[13] H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation,
World Scientific, Singapore 2009, pp. 1497 (kl/b11).

The most tragic word in the English language is potential


Arthur Lotti

6
Relativistic Particles and Fields in
External Electromagnetic Potential
Having found classical fields to describe relativistic particles, we may ask what corrections the relativistic motion causes to the Schrodinger description of atoms. We
shall therefore study Klein-Gordon and the Dirac equation in an external field. Let
A (x) be the associated four-vector potential, accounting for electric and magnetic
field strengths via (4.227) and (4.228). For classical relativistic point particles, an
interaction with these external fields is introduced via the so-called minimal substitution rule, which we are now going to discuss. Throughout this Chapter, the
presence of an external electromagnetic field will be assumed, without asking for its
origin.
An important property of the electromagnetic field is its description in terms of a
vector potential A (x) and the gauge invariance of this description. In Eqs. (4.230)
and (4.231) we have expressed electric and magnetic field strength as components
of a four-curl F = A A of a vector potential A (x). This four-curl is
invariant under gauge transformations
A (x) A (x) + (x).

(6.1)

The gauge invariance restricts strongly the possibilities of introducing electromagnetic interactions into particle dynamics and the Lagrange densities (6.94) and (6.95)
of charged scalar and Dirac fields.
The prescription for coupling electromagnetism has been known for a long time
in the classical electrodynamics of point particles.

6.1

Charged Point Particles

A free relativistic particle moving along an arbitrarily parametrized path x ( ) in


four-space is described by an action
A = Mc

d q ( )q ( ).
424

(6.2)

425

6.1 Charged Point Particles

The physical time along the path is given by q 0 ( ) = ct, and the physical velocity
by v(t) dq(t)/dt. In terms of these, the action reads:
A=

6.1.1

dt L(t) Mc

v2 (t)
dt 1 2
c
"

#1/2

(6.3)

Coupling to Electromagnetism

If the particle has a charge e (in our convention, e has a negative value for electrons)
and lies at rest at position x, its electric potential energy is
V (x, t) = e(x, t)

(6.4)

(x, t) = A0 (x, t).

(6.5)

where

In our convention, the charge of the electron e has a negative value to have agreement
with the sign in the historic form of the Maxwell equations
E(x) = 2 (x) = (x),

B(x) E(x)
= A(x) E(x)

h
i
1

= 2 A(x) A(x) E(x)


= j(x).
c

(6.6)

If the electron moves along a trajectory q(t), its potential energy is


V (t) = e (q(t), t) .

(6.7)

In the Lagrangian L = T V , this contributes with the opposite sign


Lint (t) = eA0 (q(t), t)

(6.8)

giving a potential part of the interaction


Aint
pot = e

dt A0 (q(t), t) .

(6.9)

Since the time t coincides with q 0 ( )/c of the trajectory, this can be expressed as
Aint
pot =

eZ
dq 0 A0 .
c

(6.10)

In this form it is now quite simple to write down the complete electromagnetic interaction purely on the basis of relativistic invariance. The direct invariant extension
of (6.11) is obviously
Aint =

e
c

dq A (q).

(6.11)

426

6 Relativistic Particles and Fields in External Electromagnetic Potential

Thus, the full action of a point particle can be written in covariant form as
A = Mc

d q ( )q ( )

eZ
dq A (q),
c

(6.12)

or more explicitly as
A=

dt L(t) = Mc

v2
dt 1 2
c
"

#1/2

1
dt A v A .
c


(6.13)

The canonical formalism supplies us with the canonically conjugate momenta


P=

v
e
e
L
= Mq
+ A p + A.
v
c
1 v2 /c2 c

(6.14)

The Euler-Lagrange equation obtained by extremizing this equation is


L
d L
=
,
dt v(t)
q(t)

(6.15)

or

ed
e
d
p(t) =
A(q(t), t) eA0 (q(t), t) + v i Ai (q(t), t).
dt
c dt
c
We now split
d

A(q(t), t) = (v(t) )A(q(t), t) + A(q(t), t),


dt
t

(6.16)

(6.17)

and obtain
e
e
e
d
p(t) = (v(t) )A(q(t), t)
A(q(t), t) eA0 (q(t), t) + v i Ai (q(t), t).
dt
c
c t
c
(6.18)
The right-hand side contains the electric and magnetic fields (4.232) and (4.233), in
terms of which it takes the well-known form
d
v
p=e E+ B .
dt
c


(6.19)

This can be rewritten in terms of the proper time t/ as



d
e  0
Ep +p B ,
p=
d
Mc

(6.20)

Recalling Eqs. (4.230) and (4.231), this is recognized as the spatial part of the
covariant equation
d
e
p =
F p .
(6.21)
d
Mc
The temporal component of this equation
e
d 0
p =
Ep
d
Mc

(6.22)

427

6.1 Charged Point Particles

gives the energy increase of a particle running through an electromagnetic field. In


real time this is
v
d 0
p = eE .
(6.23)
dt
c
Combining this with (6.19), we find the acceleration
d
v
d p
e
v v
E+ B
v(t) = c
=
E
0
dt
dt p
M
c
c c




(6.24)

The velocity is related to the canonical momenta and external field via
e
P A
v
c
.
= s
2
c
e
P A + m2 c2
c

(6.25)

This can be used to calculate the Hamiltonian via the Legendre transform
H=

L
vL= PvL
v

(6.26)

giving
H=c

s


e
P A
c

2

+ m2 c2 + eA0 .

(6.27)

In the non-relativistic limit this has the expansion


1
e
H = mc +
P A
2m
c
2

2

+ eA0 + . . .

(6.28)

Thus, the free theory goes over into the interacting theory by the minimal substitution rule
e
e
p p A,
H H A0 .
(6.29)
c
c
or, in relativistic notation:
e
p p A .
c

6.1.2

(6.30)

Spin Precession in Atom

In 1926, Uhlenbeck and Goudsmit noticed that the observed Zeeman splitting of
atomic levels could be explained by an electron of spin 21 . Its magnetic moment is
usually expressed in terms of the combination of fundamental constants which have
the dimension of a magnetic moment, the Bohr magneton B = eh/Mc as

 = gB Sh ,

eh
,
2Mc

(6.31)

428

6 Relativistic Particles and Fields in External Electromagnetic Potential

where S = /2 is the spin matrix which has the commutation rules


[Si , Sj ] = ihijk Sk ,

(6.32)

and g is a dimensionless number called the gyromagnetic ratio or Lande factor.


If an electron moves in an orbit under the influence of a torque-free central force,
as an electron does in the Coulomb field of an atomic nucleus, the total angular
momentum is conserved. The spin, however, shows a precession just like a spinning
top. This precession has two main contributions: one is due to the magnetic coupling
of the magnetic moment of the spin to the magnetic field of the electron orbit, called
spin-orbit coupling. The other part is purely kinematical, it is the Thomas precession
discussed in Section (4.16), caused by the slightly relativistic nature of the electron
orbit.
The spin-orbit splitting of the atomic energy levels (to be pictured and discussed
further in Fig. 6.1) is caused by a magnetic interaction energy
H LS (r) =

g
1 dV (r)
SL
,
2
2
2M c
r dr

(6.33)

where V (r) is the atomic potential depending only on r = |x|. To derive this H LS (r),
we note that the spin precession of the electron at rest in a given magnetic field B
is given by the Heisenberg equation
dS
=
dt

 B,

(6.34)

where is the magnetic moment of the electron. In an atom, the magnetic field in
the rest frame of the electron is entirely due to the electric field in the rest frame of
the atom. A Lorentz transformation (4.276) to the electron velocity produces the
magnetic field in the electron rest frame:
B = Bel =

v
E,
c

1
=q
.
1 v 2 /c2

(6.35)

Since an atomic electron has a small velocity ratio v/c which is of the order of the
fine-structure constant 1/137, we can write approximately
Bel

v
E.
c

(6.36)

The electric field gives the electron an acceleration


v =

e
E,
M

(6.37)

Mc 1

v v,
e c2

(6.38)

so that we may also write


Bel

429

6.1 Charged Point Particles

and the precession equation (6.34) as


g
dS
S.
2 (v v)
dt
2c
This can be written as

(6.39)

dS
= LS S,
(6.40)
dt
where LS is the angular velocity of spin precession caused by the orbital magnetic
field at in the rest frame of the electron:
g

v v.
(6.41)
LS
2c2

In the rest frame of the atom where the electron is accelerated towards the center
along its orbit, this result receives a relativistic correction. To lowest order in 1/c,
we must add to LS the angular velocity T of Thomas precession, such that the
total angular velocity of precession becomes

=
LS +
T g 2 1 v v.

(6.42)

Since g is very close to 2, the Thomas precession explains why the spin-orbit splitting
was initially found to be in agreement with a normal gyromagnetic ratio g = 1, the
characteristic value for a rotating charged sphere.
If there is also an external magnetic field, this is transformed to the electron
rest frame by a Lorentz transformation (4.273), where it leads to an approximate
equation of motion for the spin
dS
=
dt

 B 

v
E .
c


(6.43)

 via Eq. (6.31), this becomes




v
eg
dS
S
em
S B E .
(6.44)
dt
2Mc
c
This equation defines the frequency
em of precession due to the magnetic end elec-

Expressing

tric fields in the rest frame of the electron. Expressing E in terms of the acceleration
via Eq. (6.37), this becomes
g
M
dS
.

S eB
(v v)
dt
2Mc
c


(6.45)

The acceleration can be expressed in terms of the central Coulomb potential V (r)
as
x 1 dV
v =
.
(6.46)
r M dr
the spin precession rate in the electrons rest frame is
dS
g
x 1 dV
=
S eB + v
dt
2Mc
r c dr

g
1 dV
=
S eB
L
2Mc
Mc dr

. (6.47)

430

6 Relativistic Particles and Fields in External Electromagnetic Potential

There exists a simple Hamiltonian operator for spin-orbit interaction H LS (t)], from
which this equation can be derived via Heisenbergs equation (1.279)
i

S(t)
= [S(t), H LS (t)].
h

(6.48)

The operator is

1
dV
L
H LS (r) = B
Mc e dr
ge
g
1 dV
=
SB+
SL
.
2
2
2Mc
2M c
r dr

(6.49)

Indeed, using the commutation rules (6.32), we find immediately (6.46). Historically,
the interaction energy (6.49) was used to explain the experimental level splittings
assuming a gyromagnetic ratio g 1 for the electron.
Without the external magnetic field, the angular velocity of precession caused
by spin-orbit coupling is is
g
1 V
L
.
(6.50)
LS =
2
2
2M c
r r
It was realized by Thomas in 1927 that the relativistic motion of the electron change
the factor g as in (6.42) to g 1, so that the true precession frequency is

g 1 1 V
L
.

=
LS +
T = 2M
2 c2
r r

(6.51)

This implied that the experimental data gave g 1 1, so that g is really twice
as large as expected for a rotating charged sphere. Indeed, the value g 2. was
indeed predicted by the Dirac theory of the electron.
In Section 12.15 we shall find that the magnetic moment of the electron has a g
factor slightly larger than the Dirac value 2, the relative deviations a (g 2)/2
being defined as the anomalous magnetic moments. From measurements of the
above precession rate, experimentalists have deduced the values
a(e ) = (115 965.77 0.35) 108 ,
a(e+ ) = (116 030 120) 108 ,
a( ) = (116 616 31) 108 .

(6.52)
(6.53)
(6.54)

In quantum electrodynamics, the gyromagnetic ratio will receive further small


corrections, as will be discussed in detail in Chapter 12.

6.1.3

Relativistic Equation of Motion for Spin Vector and


Thomas Precession

If an electron moves in an orbit under the influence of a torque-free central force,


such as an electron in the Coulomb field of an atomic nucleus, the total angular

431

6.1 Charged Point Particles

momentum is conserved. The spin, however, performs a Thomas precession as


discussed in the previous section. There exists a covariant equation of motion for the
spin four-vector introduced in Eq. (4.761) which describes this precession. Along a
particle orbit parametrized by a parameter , for instance the proper time, we form
the derivative with respect to , assuming that the motion proceeds at fixed total
angular momentum:
dS
dp
= J
.
d
d

(6.55)

The right-hand side can be simplified by multiplying it with the trivial expression
1
g p p = 1.
M 2 c2

(6.56)

Now we use the identity for the -tensor


g = g + g + g + g ,

(6.57)

which can easily be verified using the antisymmetry of the -tensor and considering
= 0123. Then the right-hand side becomes a sum of the four terms

1 

, (6.58)

J
p
p
p
+
J
p
p
p
+
J
p
p
p
+
J
p
p
p

M 2 c2

where p dp /d . The first term vanishes, since p p = (1/2)dp2 /d =


(1/2)dM 2 c2 /d = 0. The last term is equal to S p p /M 2 c2 . Inserting the identity (6.57) into the second and third terms, we obtain twice the left-hand side of
(6.55). Taking this to the left-hand side, we find the equation of motion
dS
1
dp
= 2 2 S
p .
d
M c
d

(6.59)

Note that on account of this equation, the time derivative dS /d points in the
direction of p .
Let us verify that this equation yields indeed the Thomas precession. Denoting
the derivatives with respect to the time t = by a dot, we can rewrite (6.59) as
dS
S
=
dt
dS0
=
S 0
dt


1 dS
1 
2
v,
= 2 2 S 0 p0 + S p p = 2 (S v)
d
M c
c
1d
2
.
(S v) = 2 (S v)
c dt
c

(6.60)
(6.61)

We now differentiate Eq. (4.774) with respect to the time using the relation =
2

3 vv/c
, and find
S R = S

1 0
3
1 0
1
S 0 v.

S
v

S
(v v)
v

+ 1 c2
+ 1 c2
( + 1)2 c4

(6.62)

432

6 Relativistic Particles and Fields in External Electromagnetic Potential

Inserting here Eqs. (6.60) and (6.61), we obtain


S R =

1 0
3
2 1

S 0 v.

(S

v)v

S
(v v)
v

+ 1 c2
+ 1 c2
( + 1)2

(6.63)

On the right-hand side we return to the spin vector SR using Eqs. (4.773) and
(4.776), and find
S R =

2 1
(SR v)v]
=
[(SR v)v
+ 1 c2

T SR,

(6.64)

with the Thomas precession frequency

T = ( + 1) c12 v v,
2

(6.65)

which agrees with the result (4B.26) derived from purely group-theoretic considerations.
In an external electromagnetic field, there is an additional precession. For slow
particles, it is given by Eq. (6.45). If the electron moves fast, we transform the
electromagnetic field to the electron rest frame by a Lorentz transformation (4.273),
and obtain an equation of motion for the spin becomes

v
2 v
S R = B = B E
c
+1c
Expressing

" 

v
B
c

#

(6.66)

 via Eq. (6.31), this becomes

S R SR

em

eg
=
SR
2Mc

"

v
v
B E
c
+1c


v
B
c

#

(6.67)

which is the relativistic generalization of Eq. (6.44). It is easy to see that the
associated fully covariant equation is
1 d
eg
1
g
dS
eF S +
F S + 2 2 p S F p .
=
p S p =
d
2Mc
Mc
d
2Mc
M c
"

(6.68)

On the right-hand side we have inserted the relativistic equation of motion (6.21)
of a point particle in an external electromagnetic field.
If we add to this the relativistic Thomas precession rate (6.59), we obtain the
covariant Bargmann-Michel-Telegdi equation1
dS
g2 d
e
g2
1
egF S +
gF S + 2 2 p S F p .
=
p S p =
d
2Mc
Mc
d
2Mc
M c
(6.69)
"

V. Bargmann, L. Michel, and V.L. Telegdi, Phys. Rev. Lett. 2 , 435 (1959).

433

6.2 Charged Particle in Schr


odinger Theory

For the spin vector SR in the electron rest frame this implies a change in the
electromagnetic precession rate in Eq. (6.67) to2
dS
=
dt

em T S (
em +
T ) S

(6.70)

with a frequency given by the Thomas equation

e
=
Mc

"

g
g
v
1

v
g
B
1 +
1
B

2
+1 c
c
2 +1

v
E .
em T
c
(6.71)
The contribution of the Thomas precession is the part without the gyromagnetic
retio g:

"

e
1
1
1
=
1
B+
(v

B)
v
+
vE .
Mc

+1 c2
+1 c

(6.72)

This is agrees with the Thomas frequency (6.65) after inserting the acceleration
(6.24).
The Thomas equation (6.71) can be used to calculate the time dependence of
of an electron, i.e., its component of the spin in the direction
the helicity h SR v
of motion. Using the chain rule of differentiation,
1
d
d
) = S R v
+ [SR (
(SR v
v SR )
v] v
dt
v
dt

(6.73)

and inserting (6.70) as well as the equation for the acceleration (6.24), we obtain
dh
e
=
SR
dt
Mc



c
g
gv
B+
E .
1 v

2
2c v


(6.74)

where SR is the component of the spin vector orthogonal to v. This equation shows
that for a Dirac electron which has g = 2 the helicity remains constant in a purely
magnetic field. Moreover, if the electron moves ultrarelativistically (v c), the
value g = 2 makes the last term extremely small, (e/Mc) 2 SR E, so that the
helicity is almost unaffected by an electric field. The anomalous magnetic moment
of the electron, however, changes this a finite value (e/Mc)aSR E. This drastic
effect was used to measure the experimental values listed in Eqs. (6.52)(6.54).

6.2

Charged Particle in Schr


odinger Theory

When going over from quantum mechanics to second quantized field theories we
found the rule that a non-relativistic Hamiltonian
H=
2

L.T. Thomas, Phil. Mag. 3 , 1 (1927).

p2
+ V (x)
2m

(6.75)

434

6 Relativistic Particles and Fields in External Electromagnetic Potential

became an operator
H=

2
+ V (x) (x, t),
d x (x, t)
2m
3

"

(6.76)

where we have omitted the operator hats, for brevity. With the same rules we see
that the second quantized form of the interacting nonrelativistic Hamiltonian in a
static A(x) field,
H=

(p eA)2 e 0
+ A ,
2m
c

(6.77)

is given by
H=

"

e
1
i A
d x (x, t)
2m
c
3

2

+ eA (x) (x, t).

(6.78)

When going to the action of this theory we find


A=

dtL =

dt

d x (x, t) it + eA0 (x, t)


+

e
1
(x, t) i A
2m
c


2

(x, t) .

(6.79)

It is easy to verify that (6.78) reemerges from the Legendre transform


H=

L
(x, t) L.

(x, t)

(6.80)

The action (6.79) holds also for time-dependent A (x) fields.


We can now deduce the second quantized form of the minimal substitution rule
(6.29) which is
e
i A(x, t),
c
t t + ieA0 (x, t),

(6.81)

or covariantly:

e
+ i A (x).
(6.82)
c
This substitution rule has the important property that the gauge invariance of
the free photon action is preserved by the interacting theory: If we perform the
gauge transformation
A (x) A (x) + (x),

(6.83)

A0 (x, t) A0 (x, t) + t (x, t)


A(x, t) A(x, t) (x, t),

(6.84)

i.e.,

435

6.2 Charged Particle in Schr


odinger Theory

the action remains invariant if we simultaneously change the fields (x, t) of the
charged particles by a spacetime-dependent phase
e

(x, t) ei c (x,t) (x, t).

(6.85)

Under this transformation, the derivatives of the field change like


e
i (x, t) ,
(x, t) e
c
i ec (x,t)
(t iet ) (x, t).
t e
i ec (x,t)

(6.86)

The modified derivatives appearing in the action have therefore the following simple
transformation law:
e
e
e
i A (x, t) ei c (x,t) i A (x, t),
c 
c



e 0
e
t + i A (x, t) ei c (x,t) t + ieA0 (x, t).
c

(6.87)

These combinations of derivatives and gauge fields are called covariant derivatives
and written as
e
i A (x, t)
c 

Dt (x, t) t + ieA0 (x, t)
D(x, t)

(6.88)

or, in four-vector notation,


D (x) =

e
+ i A (x).
c

(6.89)

Here the adjective covariant does not refer to the Lorentz group but to the gauge
group. It records the fact that D transforms under local gauge changes (6.81) of
in the same way as itself in (6.85):
e

D (x) ei c (x) D (x).

(6.90)

With the help of this covariant derivative any action, which is invariant under global
phase changes by a constant phase angle
(x) ei (x),

(6.91)

can easily be made invariant under local gauge transformations (6.83). We merely
have to replace all derivatives by covariant derivatives (6.89) and add the gauge
invariant photon action (4.234).

436

6.3
6.3.1

6 Relativistic Particles and Fields in External Electromagnetic Potential

Charged Relativistic Fields


Scalar Field

The Lagrangian density of a free charged scalar field is from Eqs (4.162)
L = (x) (x) M 2 (x)(x)

(6.92)

For a relativistic scalar field of charge e, the interacting Lagrange density is a


straightforward generalization of the Schrodinger case (6.79):
L = [D (x)] D (x) M 2 (x)(x)




e
e
= i A (x) (x) + i A (x) (x) M 2 (x)(x)
c
c

6.3.2

(6.93)

Dirac Field

The Lagrangian densities of a free charged spin-1/2 field is from Eq. (4.498):
L = (x) (x) M 2 (x)(x)

(6.94)

(i/
m) (x).
L(x) = (x)

(6.95)

and

If the particle carries a charge e, we replace in this Lagrangian:

/ =

e
e
+ i A = / + i A
/
c
c

D
/.

(6.96)

In this way we arrive at the Lagrangian of quantum electrodynamics (QED)


1 2
(i/
L(x) = (x)
D m) (x) F
.
4

(6.97)

The classical field equations can easily be found by extremizing the action and
variation with respect to all fields which gives
A
= (i/
D M) (x) = 0

(x)
1
A
= F (x) j (x) = 0.
A (x)
c
with the current density

(6.98)
(6.99)

j (x) e c (x)
(x).

(6.100)

1
F (x) = j (x).
c

(6.101)

Equation (6.99) is the Maxwell equation for the electromagnetic field around a classical four-dimensional vector current j (x):

437

6.4 Pauli Equation from Dirac Theory

In the Lorentz gauge A (x) = 0, this equation reads simply


1
2 A (x) = j (x).
c

(6.102)

The current j combines the charge density (x) and the current density j of
particles of charge e in a four-vector
j = (c, j) .

(6.103)

In terms of electric and magnetic fields E i = F i0 , B i = F jk , the field equations


(6.101) turn into the Maxwell equations
0 = e
E = = e
= 1 j = e .
BE
c
c

(6.104)

The first is Coulombs law, the second Amperes law in the presence of charges and
currents.
Note that the physical units employed here differ from those used in many books
of classical electrodynamics3 by the absence of a factor 1/4 on the right-hand side.
The Lagrangian used in those books is
1
1 2
F (x) j (x)A (x)
8
c


i
1 h 2
1
2
=
E B (x) j A (x),
4
c

L(x) =

(6.105)

which leads to Maxwells field equations


E = 4
4
j.
B =
c

(6.106)

The form employed


conventionally

in quantum field theory arises from this by replacing A 4A and e 4e. The charge of the electron in our units has
therefore the numerical value
q

(6.107)
e = 4 4/137

rather than e = .

6.4

Pauli Equation from Dirac Theory

It is instructive to take the Dirac equation (6.98) to a two-component form corresponding to (4.561) and (4.573), and further to (4.579). Due to the fundamental
3

See for example J.D. Jackson, Classical Electrodynamics, Wiley and Sons, New York, 1967.

438

6 Relativistic Particles and Fields in External Electromagnetic Potential

nature of the equations to be derived we shall not work with natural units in this section but carry all fundamental constants along explicitly. As in (4.580), we multiply
(6.98) by (ihD
/ Mc) and work out the product
e
+ i A + Mc
(ihD
/ Mc) (ihD
/ + Mc) = i h
c




e
+ i A Mc
i h
c
(6.108)


We now use the relation


1
1
= ( + ) + ( ) = g i ,
2
2
with from (4.514), and find

(6.109)

e
+ i A
c



e
e
i
e
e

h
+ i A + i A [h + i A , h
=g
+ i A ]
c
c
2
c
c

e
h
+ i A
c



e
= h
+ i A
c


2

1 eh
F .
2 c

(6.110)

Thus we obtain as a generalization of Eqs. (4.579) the Pauli equation


"

e
h
+ i A
c


2

1 eh
F M 2 c2 (x) = 0,

2 c

(6.111)

and the same equation once more for the other two-component spinor field (x).
Note that in this equation, electromagnetism is not coupled minimally. In fact,
there is a non-minimal coupling of the spin via a term

1
F = H + i E.
2
where in the chiral and Dirac representations

D = 
0

(6.112)


0

(6.113)

respectively. Thus, in the chiral representation, Eq. (6.111) decomposes into two
separate two-component equations for the upper and lower spinor components (x)
and (x) in (x):
"

e
h
+ i A
c


2

 (H iE) M

2 2

#(

(x)
(x)

= 0.

(6.114)

In the nonrelativistic limit where c , we remove the fast oscillations from


2
(x), setting (x) eiM c t/h (x, t)/ 2M as in (4.153), and find for (x, t) the
nonrelativistic equation
h
2
e
it +
i A
2M
ch

"

2

e
H eA0 (x) (x, t) = 0.
+
2Mc

(6.115)

439

6.5 Relativistic Wave Equations in Coulomb Potential

This corresponds to a magnetic interaction energy


(6.116)
For a small magnet with a magnetic moment , the magnetic interaction energy is
Hmag =  H.
(6.117)
Hmag =

eh
H.
2Mc

A Dirac particle with has therefore a magnetic moment


e

e h
=2
S,
 = Mc
2
2Mc

(6.118)

where S = /2 is the spin matrix. Experimentally, one parametrizes a magnetic


moment of a fundamental particles as
e
S
 = g 2Mc

(6.119)

with the gyromagnetic ratio g, which is equal to unity for a uniformly charged sphere.
According to the Dirac theory, an electron has a gyromagnetic ratio
ge |Dirac = 2.

(6.120)

The experimental value is very close to this. The small deviation called anomalous
magnetic moment will be explained in Chapter 12.
The nonrelativistic Pauli equation (6.115) could also have been obtained by introducing a minimal coupling directly into the nonrelativistic two-component equation
(4.576). This changes it it eA0 and ( )2 [ ( ieA)]2 . The latter
is worked out as in (6.121), yielding

[ ( ieA)]2 = ij + iijk k (i ieAi )(j ieAj )

= ( ieA)2 + iijk k (i ieAi )(j ieAj ) + e H, (6.121)

so that Eq. (4.578) goes over into (6.115), after reinserting all fundamental constants,

6.5

Relativistic Wave Equations in Coulomb Potential

It is now easy to write down field equations for a Klein-Gordon and a Dirac field
in the presence of an external Coulomb potential of charge Ze. In natural units we
have

Z
VC (x) =
,
r = x2 ,
(6.122)
r
corresponding to a four-vector potential
eA (x) = (VC (x, 0), 0).

(6.123)

440

6 Relativistic Particles and Fields in External Electromagnetic Potential

Since this does not depend on time, we can consider the wave equations for wave
functions (x) = eiEt E (x) and (x) = eiEt E (x), and find the time-independent
equations
(E 2 + 2 M)E (x) = 0

(6.124)

and

( 0 E + i M)E (x) = 0.

(6.125)

In these equations we simply perform the minimal substitution


EE+

Z
.
r

(6.126)

The energy-eigenvalues obtained from the resulting equations can be compared with
those of hydrogen-like atoms. The velocity of an electron in the ground state is of
the order Zc. Thus for rather high Z, the electron has a relativistic velocity and
there must be corrections significant deviations from Schrodinger theory. We shall
see that the Dirac equation in an external field reproduces quite well a number of
features resulting from the relativistic motion.

6.5.1

Reminder of Schr
odinger Equation with Coulomb
Potential

The time-independent Schrodinger equation reads




1
Z
2
E E (x) = 0.
2M
r


(6.127)

The Laplacian may be decomposed into radial and angular parts by writing
2 =

2
2
2
L
+

,
r 2 r r
r2

(6.128)

= xp
are the differential operators for the generators of angular momenwhere L
tum [the spatial part of Li = L23 of (4.95)]. Then (6.127) reads
2 2ZM
2
L
2
2
+ 2
2ME E (x) = 0.
r
r r
r
r
!

(6.129)

2 are the spherical harmonics Ylm (, ), which diagonalize also


The eigenstates of L
with the eigenvalues
the third component of L,
2 Ylm (, ) = l(l + 1)Ylm (, ),
L
3 Ylm (, ) = mYlm (, ).
L

(6.130)

6.5 Relativistic Wave Equations in Coulomb Potential

441

The wave functions may be factorized into a radial wave function Rnl (r) and a
spherical harmonic:
nlm (x) = Rnl (r)Ylm(, ).
(6.131)
Explicitly,
Rnl (r) =

1
1/2
aB n (2l

1
+ 1)!

v
u
u
t

(n + l)!
(n l 1)!

(6.132)

(2r/naB )l+1 er/naB M(n + l + 1, 2l + 2, 2r/naB )


v
u

1 u (n l 1)! r/naB
e
(2r/naB )l+1 L2l+1
= 1/2 t
nl1 (2r/naB ),
(n
+
l)!
aB n
where aB is the Bohr radius, which in natural units h
= c = 1 is equal to
aB =

1
,
ZM

(6.133)

For a hydrogen atom with Z = 1, this is about 1/137 times the Compton wavelength
of the electron e h
/Me c. The classical velocity of the electron on the lowest Bohr
orbit is vB = c. Thus it is almost nonrelativistic, which is the reason what the
Schrodinger equation explains the hydrogen spectrum quite well. The functions
M(a, b, z) are confluent hypergeometric functions or Kummer functions, defined by
the power series
a
a(a + 1) z
M(a, b, z) F1,1 (a, b, z) = 1 + z +
+ ... .
b
b(b + 1) z!

(6.134)

For b = n they are polynomial related to the Laguerre polynomials4 Ln (z) by


Ln (z)

(n + )!
M(n, + 1, z),
n!!

(6.135)

The radial wave functions are normalized to


Z

drRnr l (r)Rnr l (r) = nr nr .

(6.136)

They have an asymptotic behavior Pnl (r/n)er/naB , where Pnl (r/n) is a polynomial
of degree nr = n l 1, which defines the radial quantum number. The energy
eigenvalues depend on are n in the well-known way:
En = Z 2
4

M2
.
2n2

(6.137)

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 8.970 (our definition differs from that in
L.D. Landau and E.M. Lifshitz, Quantum Mechanics, Pergamon Press, New York, 1965, Eq. (d.13):
Our Ln = () /(n + )!Ln+ |L.L. ).

442

6 Relativistic Particles and Fields in External Electromagnetic Potential

The number 2 M/2 is the Rydberg-constant:


Ry =

2 M
27.21

eV 3.288 1015 Hz.


2
2

(6.138)

Later in Section 12.21) we shall need the value of the wave function at the origin.
It is non-zero only for s-waves where it is equal to
1
|n00 (0)| =

6.5.2

ZM
1
=
n

1
.
naB

(6.139)

Klein-Gordon Field in Coulomb Potential

After the substitution (6.126) into (6.124), we find the Klein-Gordon equation in
the Coulomb potential (6.122):
"

Z
E+
r

2

+ M

E (x) = 0.

(6.140)

With the angular decomposition (6.128), this becomes


2 Z 2 2 2ZE
2
L
2
2
+

(E 2 M 2 ) E (x) = 0.
r
r r
r2
r

"

(6.141)

The solutions of this equation can be obtained from those of the nonrelativistic
Schrodinger equation (6.129) by replacing
2 L
2 Z 2 2 ,
L
E
,
M
E2 M 2
E
.
2M

(6.142)
(6.143)
(6.144)

The replacement (6.142) is done most efficiently by defining the eigenvalues l(l +
2 Z 2 2 by analogy with those of L
2 as ( + 1):
1) Z 2 2 of the operator L
l(l + 1) Z 2 2 ( + 1).

(6.145)

Then the quantum number l of the Schrodinger wave functions is simply replaced
by l = l l , where
l

"

2

1
1
= l+ l+
2
2
2 2
Z
+ O(4 ).
=
2l + 1

#1/2

(6.146)

The other solution of relation (6.145) with the opposite sign in front of the square
root is unphysical since the associated wave functions are too singular at the origin

443

6.5 Relativistic Wave Equations in Coulomb Potential

to be normalizable. As before, the radial quantum number nr which determines the


degree of the polynomial Pnl (r/n) in the wave functions must be an integer. This
is no longer true for the combination of quantum numbers which determines the
energy, which is now given by
nr + + 1 = nr + l + 1 l = n l .

(6.147)

This leads to the equation for the energy eigenvalues


Enl 2 M 2
Z 2 M2 Enl 2
1
=
,
2M
2
M 2 (n l )2

(6.148)

with the solution


M
Enl = q
2
1 + Z 2 /(n l )2

(Z)2 3 (Z)4
(Z)4
= M 1
+

+ O(Z 6 6 ) .
2n2
8 n4
n3 (2l + 1)
"

(6.149)

The first two terms correspond to the Schrodinger energies (6.137) (plus rest energy),
the next two are relativistic corrections. The first of these breaks the degeneracy
between the levels of the same n and different l, generated in the Schrodinger theory
by the Lentz-Runge vector [O(4)-invariance].
The correction terms become large for large central charge Z. In particular, the
lowest energy and successively the higher ones become complex for central charges
Z > 137/2. The physical reason for this is that the large potential gradient near the
origin can create pairs of particles from the vacuum. This phenomenon can only be
properly understood after quantizing the field theory.
As for the free Klein-Gordon field, the energy appears with both signs.

6.5.3

Dirac Field in Coulomb Potential

After the substitution (6.126) into (6.125), we find the Dirac equation in the
Coulomb potential (6.122):

Z 0
E+
+ i M E (x) = 0.
r



(6.150)

In order to find the energy spectrum it is useful to establish contact with the
Klein-Gordon case. Multiplying (6.150) by the operator


we obtain

"

Z
E+
r

2

Z 0
+ i + M,
E+
r


+ i

M 2 E (x) = 0.,
r

(6.151)

(6.152)

444

6 Relativistic Particles and Fields in External Electromagnetic Potential

In the chiral representation, 44 -matrix 0 = has a diagonal block form (4.557).


We therefore decompose
!
E (x)
E (x) =
,
(6.153)
E (x)
and find the equation for the upper two-component spinors
"

Z
E+
r

2

Z
M 2 E (x) = 0,
+ +i
r
2

(6.154)

and the same equation for E (x) with i replaced by i. Expressing 2 via (6.128)
and writing 1/r =
x/r 2 , we obtain the differential equation

2 Z 2 2 + iZ x
2ZE
L
+

(E 2 M 2 ) E (x) = 0,
r2
r
(6.155)
and a corresponding equation for E (x).
, the total angular momentum
Due to the rotation invariance of x
"

2
2

+
r 2 r r

=L
+S=L
+
J

(6.156)

2 and
commutes with the differential operator in (6.155). Thus we can diagonalize J
1
3
J3 with eigenvalues j(j + 1) and m. For a fixed value of j = 2 , 1, 2 , . . . , the orbital
angular momentum can have the value l+ = j + 21 and l = j 1/2. The two states
is a pseudoscalar and must necessarily
have opposite parities. The operator x
is the unit matrix, its eigenvalues must
change the parities. Since the square of x
be 1. Moreover, the unit vector x changes necessarily the parity and thus l. Thus,
in the two-component Hilbert space of fixed quantum numbers j and m with orbital
angular momenta l = l = j 1/2, the diagonal matrix elements vanish




(6.157)

|jm, +i = 1.
hjm, | x

(6.158)

|jm, +i = 0,
hjm, +| x

|jm, i = 0.
hjm, | x

For the off-diagonal elements we easily calculate

|jm, i = 1,
hjm, +| x

The central parentheses in (6.155) have therefore the matrix elements




L Z iZ
x =

(j + 21 )(j + 23 ) Z 2 2
iZ
iZ
(j 12 )(j + 12 ) Z 2 2

(6.159)
In analogy with the Klein-Gordon case, we denote the eigenvalues of this matrix by
( + 1). The corresponding values of are found to be
j + =

"

1
j+
2

2

#1/2

j = j+ 1.

(6.160)

445

6.5 Relativistic Wave Equations in Coulomb Potential

These may be written as


j = j

1
j l j ,
2

(6.161)

where l 21 is the orbital angular momentum, and


1
j j +
2

"

1
j+
2

2

#1/2

Z 2 2
+ O(Z 4 4 ).
2j + 1

(6.162)

When solving Eq. (6.155), the solutions consist, as in the nonrelativistic hydrogen
atom, of an exponential factor multiplied by a polynomial of degree nr which is the
radial quantum number. It is related to the quantum numbers of spin and orbital
angular momentum, and to the principal quantum number n, by
nr + j + 1 = nr + l + 1 l = n j .

(6.163)

In terms of j , the energies obey the same equation as in (6.149), so that we obtain
Enj = q

M
1 + Z 2 2 /(n j )2

(Z)4
(Z)2 3 (Z)4
+

+ O(Z 6 6 ) .
= M 1
2n2
8 n4
n3 (2j + 1)
#

"

(6.164)

The condition nr 0 implies that


j n

j n

3
2
1
2

for

j+ = j + 21 j ,

j = j 12 j .

(6.165)

For n = 1, 2, 3, . . . , the total angular momentum runs through j = 12 , 23 , . . . , n 21 .


The spectrum of the hydrogen atom according to the Dirac theory is shown in
Fig. 6.1. As a remnant of the O(4)-degeneracy of the levels with l = 0, 1, 2, . . . , n 1
and fixed n in the Schrodinger spectrum, there is now a twofold degeneracy of levels
of equal n and j, with adjacent l-values, which are levels of opposite parity. An
exception is the highest total angular momentum j = n 1/2 at each n which
occurs only once. The lowest degenerate pair consists of the levels 2S1/2 and 2P1/2 .5
It was an important experimental discovery to find that this prediction is wrong.
There is a splitting of about 10% of the fine-structure splitting. This is called the
Lamb shift. Its explanation is one of the early triumphs of quantum electrodynamics,
which will be discussed in detail in Chapter .
As in the Klein-Gordon case, there are complex energies, here for Z > 137, with
S1/2 being the first level to become complex.
5

Recall the notation in atomic physics for an electronic state: n2S+1 LJ , where n is the principal
quantum number, L the orbital angular momentum, J the total angular momentum, and S the
total spin. In a one-electron system such as the hydrogen atom, the trivial superscript 2S + 1 = 2
may be omitted.

446

6 Relativistic Particles and Fields in External Electromagnetic Potential

Figure 6.1 Hydrogen spectrum according to Diracs theory. The splittings are shown
only schematically. The fine-structure splitting of the 2P -levels is about 10 times as big
as the hyperfine splitting and Lamb shift.

An important correct prediction of the Dirac theory is the presence of fine structure. States with the same n and l but with different j are split apart by the forth
term in Eq. (6.164) MZ 4 4 n3 /(2j+1). For the states 2P1/2 and 2P3/2 , the splitting
is
Z 4 2 2
fine E2P =
M.
(6.166)
32
In a hydrogen atom, this is equal to
fine E2P = 3.10.95 GHz.

(6.167)

Thus it is roughly of the order of the splitting caused by the interaction of the magnetic moment of the electron with that of the proton, the so-called hyperfine-splitting.
This is for 2S 1/2 , 2P 1/2 , and 2P 3/2 levels approximately equal to 1, 1/8, 1/24, 1/60
times 1 420 MHz.6
In a hydrogen atom, the electronic motion is only slightly relativistic, the velocities being of the order c, i.e., only about 1% of the light velocity. If one is not
only interested in the spectrum but also in the wave functions it is advantageous
to solve directly the Dirac equation (6.150) with the gamma matrices in the Dirac
6

See H.A. Bethe and E.E. Salpeter in Encyclopedia of Physics (Handbuch der Physik) 335 ,
Springer, Berlin, 1957, p. 196.

447

6.5 Relativistic Wave Equations in Coulomb Potential

representation (4.544). Multiplying (6.150) by 0 and inserting for 0


Dirac matrix (4.556), we obtain

Z
i

EM +

Z
E+M +
i
r

E (x)
E (x)

= 0.

= the
(6.168)

This is of course just the time-independent version of (4.573) extended by the


Coulomb potential according to the minimal substitution rule (6.126). To lowest
order in , the lower spinor is related to the upper by
E (x) i

 E (x).

(6.169)

2M

We may take care of rotational symmetry of the system by splitting the spinor wave
functions into radial and angular parts
Gjl (r) l
yj,m(, )
i

r
,
E (x) =
F (r)

jl
l
yj,m(, )
x
r

(6.170)

l
where yj,m
(, ) denote the spinor spherical harmonics. They are composed from
the ordinary spherical harmonics Ylm (, ) and the basis spinors (s3 ) of (4.443) via
Clebsch-Gordan coefficients (see Appendix 4C):
l
yj,m
(, ) = hj, m|l, m ; 12 , s3 iYlm (, )(s3 ).

(6.171)

The derivation is given in Appendix 6A.


The explicit form of the spinor spherical harmonics (6.171) is for l = l :

!
l+ m + 12 Yl+ ,m 1 (, )
1
l+
2

yj,m (, ) =
,
(6.172)
2l+ + 1 l+ + m + 21 Yl+ ,m+ 1 (, )
2

l
yj,m
(, )

1
2l + 1

!
l + m + 21 Yl ,m 1 (, )
2

.
l m + 12 Yl ,m+ 1 (, )
2

On these eigenfunctions, the operator L


L

(6.173)

 has the eigenvalues

l
l
 yj,m
(, ) = (1 + )yj,m(, ),

(6.174)

with

1
1
= (j + ), j = l .
(6.175)
2
2
We can now go from Eqs. (6.168) to radial differential equations by using the
trivial identity,

f (r) l+
y

r l,m

 x ( x) i ( ) f (r) yl

l,m ,
+

r2

(6.176)

448

6 Relativistic Particles and Fields in External Electromagnetic Potential

and the algebraic relation Eq. (4.461) in the form

( a)( b) = i (a b) + i(a b),

(6.177)

to bring the right-hand side to

 x (irr i L) f (r) yl

r2

l,m

"

f (r)
f (r)
= ir
i (1 + ) 2
r
r

l
 x yl,m
.
+

(6.178)

In this way we find the radial differential equations for the functions Fjl (r) and
Gjl (r):
Z
d
1
EM +
Gjl (r) = Fjl (r) (j + 1/2) Fjl (r),
(6.179)
r
dr
r


d
1
Z
Fjl (r) =
Gjl (r) (j + 1/2) Gjl (r).
(6.180)
E+M +
r
dr
r
Tosolve these, dimensionless variables 2r/ are introduced, with =
1/ M 2 E 2 , writing


F (r) =

1 E/Me/2 (F1 F2 )(), G(r) =

1 + E/Me/2 (F1 + F2 )(). (6.181)

The functions F1,2 () satisfy a degenerate hypergeometric differential equation of


the form
#
"
d
d2
(6.182)
2 + (b ) a F (a, b; ) = 0,
d
d
and the solutions are
F2 () = l F ( ZE, 2 + 1; ),
ZE
F1 () = l
F ( + 1 ZE, 2 + 1; ).
1/ + ZE

(6.183)

The constant is Einsteins gamma parameter = 1 v 2 /c2 for the atomic unit
velocity v = Zc. It has the expansion = 1 Z 2 2 /2,
As an example, we write down explicitly the ground state wave functions of the
1/2
1S state:

1
0
v
u

u (2MZ)3

0
1
1+
emZr
t

1S 1/2 , 1 =
1
i .
i 1 cos
1
i
sin
e
2
4
2(1 + 2) (2MZ)

Z
Z
1
1
i
i Z sin e i Z cos
(6.184)
The first column is for m = 1/2, the second for m = 1/2. For small , Einsteins
gamma parameter has the expansion = 1 Z 2 2 /2, and we see that in the
0, the upper componentsqof the spinor wave functions tend to the nonrelativistic
Schrodinger wave function 2 (ZM)3 /4e multiplied by Pauli spinors (4.443).
In general,
l
(6.185)
j,m
(x) = hj, m|l, m; 12 , s3 inlm (x)(s3 ).
The lower, small components vanish.

449

6.6 Green Function in External Electromagnetic Field

6.6

Green Function in External Electromagnetic Field

An important physical quantity of a propagating field is the Green function, defined


as the solution of the equation of motion with a -function source term [recall (1.314)
and (2.403)]. For external electromagnetic fields which are constant or plane waves,
this Green function can be calculated exactly.

6.6.1

Scalar Field in Constant Electromagnetic Field

For a scalar field, the Green function G(x, x ) is defined by the inhomogeneous
differential equation
( 2 M 2 )G(x, x ) = i (4) (x x ),

(6.186)

whose solution can immediately be expressed as a Fourier integral:


Z Z
d4 p
d4 p ip(xx )+i (p2 M 2 +i)
i
ip(xx )
d
e
=
e
.
(2)4 p2 M 2 + i
(2)4
0
(6.187)
A detailed discussion of this function will be given in Subsection 7.2.2.
Here we shall address the problem of calculating the corresponding Green function in the presence of a static electromagnetic field, which obeys the more complicated differential equation

G(xx ) =

[i eA(x)]2 M 2 G(x, x ) = i (4) (x x ),

(6.188)

for which a Fourier decomposition is no longer helpful. For a constant and an


oscillating electromagnetic field, however, this equation can be solved by an elegant
method due to Fock and Schwinger [1].
Generalizing the right-hand side of (6.187), we find the representation
G(x x ) =

d hx|ei [(ieA)

M 2 +i]

|x i.

(6.189)

The integrand contains the time-evolution operator associated with the Hamiltonian
operator

H(x,
i) (i eA)2 + M 2 .
(6.190)
This is the Schroedinger representation of the operator
= H(
H
x, p) = P 2 + M 2 ,

(6.191)

where P p eA (
x) is the canonical momentum in the presence of electromagnetism.
We shall calculate the evolution operator in (6.189) by introducing timedependent Heisenberg position and momentum operators obeying the HeisenbergEhrenfest equations of motion [recall (1.276)]:
h

d
x ( )
x )] = 2P ( )
= i H,
d
h
i

dP ( )
P ( ) = 2eF (
= i H,
x( ))P ( ) + ie F (
x( )).
d

(6.192)
(6.193)

450

6 Relativistic Particles and Fields in External Electromagnetic Potential

In a constant field where F (


x( )) is a constant matrix F , the last term in
the second equation is absent and we find directly the solution


P ( ) = e2eF


where the matrix e2eF




e2eF

P (0).

(6.194)

is defined by its formal power series expansion

= + 2eF + 4e2 F F

2
+ ... .
2

(6.195)

Inserting (6.194) into Eq. (6.192), we find the time-dependent operator x ( ):

x ( ) x (0) =

e2eF 1
eF

P (0).

(6.196)

where the matrix on the right-hand side is again defined by its formal power series
e2eF 1
eF

(2 )3
= 2 + e F F
+ ... .
3!
2

(6.197)

Note that division by eF is not a matrix multiplication by the inverse of the matrix
eF but indicates the reduction of the expansion powers of eF by one unit. This is
defined also if eF does not have an inverse.
We can invert Eq. (6.196) to find
eeF
1
eF
P (0) =
2
sinh eF
"

[
x( ) x(0)] ,

(6.198)

and, using (6.194),


P ( ) = L (eF ) [
x( ) x(0)] ,
with the matrix

1
eeF
L (eF )
eF
2
sinh eF
"

(6.199)

(6.200)

By squaring (6.199) we obtain


P 2 ( ) = [
x( ) x(0)] K (eF ) [
x( ) x(0)] ,

(6.201)

where
K (eF ) = L (eF )L (eF ).

(6.202)

Using the antisymmetry of the matrix F , we can rewrite this as


1
e2 F 2
K (eF ) = L (eF )L (eF ) =
4 sinh2 eF

"

(6.203)

451

6.6 Green Function in External Electromagnetic Field

The commutator between two operators x( ) at different times is


e2eF 1
[
x ( ), x (0)] = i
eF

(6.204)

and
!

e2eF 1
x ( ), x (0) + x ( ), x (0) = i
eF
!
#
"
e2eF e2eF
sinh 2eF
= i
= 2i
,

eF
eF

e2eF 1
+i
eF T

(6.205)

With the help of this commutator, we can expand (6.201) in such a way in powers
of operators x( ) and x(0), that the later operators x( ) come to lie to the left of
the earlier operators x(0) as follows:
H(
x( ), x(0); ) =
x ( )K (eF )
x ( ) x (0)K (eF )
x (0)
i
+ 2
x ( )K (eF )
x (0) tr [eF coth eF ] + M 2 .
2

(6.206)

Given this form of the Hamiltonian operator it is easy to calculate the time evolution
amplitude in Eq. (6.189):

hx, |x 0i hx|eiH |x i.
(6.207)
It satisfies the differential equation

eiH |x i
eiH |x i = hx|eiH eiH H
i hx, |x 0i hx|H
x( ), P ( ))|x , 0i.
= hx, |H(

(6.208)

Replacing the operator H(


x( ), P ( )) by H(
x( ), x(0); ) of Eq. (6.206), the matrix
elements on the right-hand side can immediately be evaluated using the property
hx, |
x( ) = xhx, |,

x(0)|x , 0i = x |x , 0i,

(6.209)

and the differential equation (6.210) becomes


i hx, |x 0i H(x, x ; )hx, |x 0i,
or
hx, |x 0i = C(x, x )E(x, x ; ) C(x, x )ei

d H(x,x ; )

(6.210)
.

(6.211)

The prefactor C(x, x ) contains a possible constant of integration in the exponent


which may have an arbitrary dependence on x and x . The following integrals are
needed:
Z

1
d K(eF ) =
4

1
e2 F 2
= eF coth eF ,
2
4
sinh eF

(6.212)

452

6 Relativistic Particles and Fields in External Electromagnetic Potential

and
Z

sinh eF
sinh eF
= tr log
+ 4 log .
eF
eF

d tr [eF coth eF ] = tr log

(6.213)

these results following again from a Taylor expansion of both sides. The exponential
factor E(x, x ; ) in (6.211) becomes, therefore,
)

1
i
1
sinh eF
E(x, x ; ) = 2 exp (xx ) [eF coth eF ] (xx ) iM 2 tr log
.

4
2
eF
(6.214)
The last term produces a prefactor

det

sinh eF
eF

1/2

(6.215)

The time-independent integration constant is fixed by the differential equation


with respect to x:

[i eA (x)] hx, |x 0i = hx|P eiH |x i = hx|eiH eiH P eiH |x i

= hx, |P ( )|x 0i,

(6.216)

which becomes, after inserting (6.199):


[i eA (x)] hx, |x 0i = L (eF )(x x ) hx, |x 0i,

(6.217)

Calculating the partial derivative we find


i hx, |x 0i = [i C(x, x )]E(x, x ; ) + C(x, x )[i E(x, x ; )]
1
= [i C(x, x )]E(x, x ; ) + C(x, x ) [eF coth eF ] (x x ) E(x, x ; ).
2
Subtracting from this eA (x)hx, |x 0i, and inserting (6.211), the right-hand side of
(6.217) is equal to [i C(x, x )]E(x, x ; ) plus


1
L (eF )(x x ) [eF coth eF ] (x x ) C(x, x )E(x, x ; ). (6.218)
2

Inserting Eq. (6.200), this simplifies to


e
F (x x ) C(x, x )E(x, x ; ),
2

(6.219)

so that C(x, x ) satisfies the time-independent differential equation


e
i eA (x) F (x x ) C(x, x ) = 0.
2

(6.220)

This is solved by

C(x, x ) = C exp ie

x
x

1
A () + F ( x )
2



(6.221)

453

6.6 Green Function in External Electromagnetic Field

The contour of integration is arbitrary since A () A () + 12 F ( x ) has a


vanishing curl:
A (x) A (x) = 0.

(6.222)

We can therefore choose the contour to be a straight line connecting x and x, in


which case the F -term does not contribute in (6.221), since d points in the same
direction of x x as x and F is antisymmetric. Hence we may write for
a straight-line connection


C(x, x ) = C exp ie

x
x

d A () .

(6.223)

The normalization constant C is finally fixed by the initial condition


lim hx, |x 0i = (4) (x x ),

(6.224)

which requires
C=

i
.
(4)2

(6.225)

Collecting all terms we obtain


x
i
1/2 sinh eF

hx, |x 0i =
d
A
()
det
exp
ie

(4 )2
eF
x


i
exp (xx ) [eF coth eF ] (xx ) iM 2 .
4

(6.226)

For zero field, this reduces to the relativistic free-particle amplitude


i (x x )2
i
exp

iM 2 .
hx, |x 0i =
2
(4 )
2
2
"

(6.227)

According to relation (6.189), the Green function of the scalar field is given by
the integral
Z

d hx, |x 0i.
(6.228)
G(x, x ) =
0

The functional trace of (6.226)


Trhx, |x 0i = V t

i
eE
2
(4 ) sinh eE

(6.229)

will be needed below. Due to tranlation invariance in spacetime, it carries a factor


total spatial volume V total time t of the universe.
The result (6.229) can be checked by a more elementary derivation [96]. We
let the constant electric field point in the z-direction, and represent it by a vector
potential to have only a zeroth component
A3 (x) = Ex0 .

(6.230)

454

6 Relativistic Particles and Fields in External Electromagnetic Potential

Then the Hamiltonian (6.191) becomes


=
2 + (
H
p20 + p
p3 + eEx0 )2 + M 2 ,

(6.231)

where p are the two-dimensional momenta in the xy-plane. Using the commutation
rule [p0 , x0 ] = i, this can be rewritten as
= eip0 p3 /eE H
eip0 p3 /eE
H

(6.232)

is the sum of two commuting Hamiltonians:


where H
= (
+ H
.
H
p20 e2 E 2 x20 ) + p2 + M 2 H
E

(6.233)

The first is a harmonic Hamiltonian with imaginary frequency E = ieE and an


energy spectrum 2(n + 1/2)ieE. The second describes a free particle in the xyplane. This makes it easy to calculate the functional trace. We insert a comlete set
of momentum states on either side of (6.207), so that the functional trace becomes
Trhx, |x 0i =

d x

d4 p
(2)4

d4 p i(pp )x

e
hp|ei (HE +H ) |p i
4
(2)

(6.234)

The matrix elements are

hp|ei H |p i = eip0 (x0 +p /eE) hp0 |eisHE |p0 iei (p +M


(2)2 (2) (p p )(2)(p3 p3 ).

2 i)

eip0 (x0 +p

3 /eE)

(6.235)

Inserting this into (6.234) and performing the integrals over the spatial parts of p
appearing in the -functions of (6.235) yields
d2 p i (p2 +M 2 i)

e
(2)2
Z
dp0 dp3 dp0 i(p0 p0 )(x0 +p3 /eE)

e
hp0 |eisHE |p0 i,

3
(2)

Trhx, |x 0i = V

dx0

(6.236)

which can be reduced to


i i (M 2 i) eE
e
Trhx, |x 0i = V t
4
2

"Z

dp0

hp0 |ei HE |p0 i .


2

(6.237)

The expression in brackets is the trace of ei HE , which is conveniently calculated


in the eigenstates |ni of the harmonic oscillator with eigenvalues 2(n + 1/2)E :

i H
E

Tre

ei 2(n+1/2)eE =

n=0

i
1
=
.
2 sin E
2 sinh eE

(6.238)

Thus we obtain
Trhx, |x 0i = V t

i
eE
,
4(2)2 2 sinh eE

(6.239)

455

6.6 Green Function in External Electromagnetic Field

6.6.2

Dirac Field in Constant Electromagnetic Field

For a Dirac field we have to solve the inhomogeneous differential equation


{i [ eA (x)] M} S(x, x ) = i (4) (x x ),

(6.240)

rather than (6.188). The solution can formally be written as


x ) = i (4) (x x ),
S(x, x ) = {i [ eA (x)] + M} G(x,

(6.241)

x ) solves a slight generalization of Eq. (6.188):


where G(x,

e
x ) = i (4) (x x ),
[i eA(x)]2 F M 2 G(x,
2

(6.242)

which is the Green function of the Pauli equation (6.111), in natural units. For
a constant field, the extra term enters the final result (6.241) in a trivial way if
we recall the relations to the Green function (6.189) and (6.228), which imply that
x ) contains the fields as follows:
G(x,
x ) =
G(x,

e
d exp i F hx, |x 0i.
2


(6.243)

Constant Electric Background Field


For a constant electric field in the z-direction, we choose the vector potential to have
only a zeroth component
A3 (x) = Ex0 .
(6.244)
Then, since F 30 = E, we have F3 0 = E and F0 3 = E, the field tensor F is
given by the matrix

F = E

0
0
0
1

0
0
0
0

0
0
0
0

1
0
0
0

= iE M3 ,

(6.245)

where M3 is the generator (4.60) of pure Lorentz transformations in the z-direction.


The exponential eeF is therefore equal to the boost transformation (4.59) B3 () =
eiM3 with a rapidity = E . Then we find from (4.14) the explicit matrices

cosh eE
0
0
sinh eE

0
1
0
0

0 sinh eE
0
0
1
0
0 cosh eE

(6.246)

0
0
0
sinh eE

0
0
0
0

0 sinh eE
0
0
0
0
0
0

(6.247)

eeF =
and hence

sinh eF =

456

6 Relativistic Particles and Fields in External Electromagnetic Potential

sinh eF
=
eF

and

sinh eE
eE
0
0

1
0

0
1

0
0
sinh eE
eE

0
1
0
0

0
0
0
0
1
0
0 coth eE

eF coth eF = eE
Thus we obtain

coth eE
0
0
0

(6.248)

(6.249)

x
i
eE
hx, |x 0i =
d A ()
(6.250)
exp
ie
2

(4 ) sinh eE
x
i
2
0
0
T 1
T
3
3
e 4 [(xx ) eE coth eE (xx ) +(xx ) (xx ) +(xx ) eE coth eE (xx ) ]iM ,

where the
superscript T indicates transverse directions to E. The prefactor
Rx
exp [ie x d A ()] is found by inserting (6.244) and integrating along the straight
line
= x + s(x x ), s [0, 1],
(6.251)
to be

exp ie

x
x

d A () = eieE(x0 x0 )

R1
0

ds[z +s(zz )]

= eieE(x0 x0 )(z+z ) . (6.252)

The exponential prefactor in the fermionic Green function (6.243) is calculated


in the chiral representation of the Dirac algebra where, due to (6.112) and (6.113)
e
exp i F
2


0
eeE
0 eeE

= exp (e E ) =

(6.253)

which is equal to
e
exp i F =
2


cosh eE sinh eE E
0

0
.
cosh eE +sinh eE E
(6.254)

Comparison with (4.503) shows that this is the Dirac representation of a Lorentz
boost with in the direction of E with rapidity = 2e|E| . The Dirac trace of the
evolution amplitude for Dirac fields is then simply
trhx, |x 0i =

i
eE
4 cosh eE,
2
(4 ) sinh eE

(6.255)

and the functional trace of this carries simply a total spacetime volume factor V t
as in the scalar expression (6.229).

6.6 Green Function in External Electromagnetic Field

457

Note that the Lorentz-transformation (6.254) has twice the rapidity of the transformation (6.246) in the defining representation, this being a manifestation of the gyromagnetic ratio of the electron in Diracs theory being equal to two [recall (6.120)].
The process of pair creation in a space- and time-dependent electromagetic field
is discussed in Ref. [3].
The above discussion becomes especially simple in 1+1 spacetime dimensions,
the so-called massive Schwinger model [4].

6.6.3

Dirac Field in Electromagnetic Plane-Wave Field

The constant-background field results in the last subsection simplify drastically if


electric and magnetic fields have the same size and are orthogonal to each other.
This is the case for a traveling plane wave of arbitrary shape [5] running along some
direction n with n2 = 0. If denotes the spatial coordinate along n, we may write
the vector potential as
A (x) = f (), nx.
(6.256)

where is some polarization vector with the normalization 2 = 1 in the gauge


n = 0. The field tensor is
F = f (),

n n ,

(6.257)

where the constant tensor satisfies


n = 0,

= 0,

= n n .

(6.258)

The Heisenberg equations of motion (6.192), (6.194) take the form


h

d
x ( )
x )] = 2P ( )
= i H,
(6.259)
d
h
i
dP ( )
)). (6.260)
)) + e n f ((
P ( ) = 2e P ( )f ((
= i H,
d
2

Note that the last term in (6.194) vanishes for a sourceless plane wave: F = 0.
Multiplying these equations by n we see that
n

d ( )
= 2n P ( ),
d

dP ( )
= 0.
d

(6.261)

Hence
nP ( ) = nP (0) = const,

) (0)
= n
(
x( ) n
x(0) = 2 nP ( ).

(6.262)

Whereas the components of P ( ) parallel to n are time independent, those orthogonal to n have a nontrivial time dependence. To find it we multiply (6.260) by
and find

d
= en f ()(2n

d = en df () , (6.263)
P ( ) = 2e P f ()
P ) = en f ()
d
d
d

458

6 Relativistic Particles and Fields in External Electromagnetic Potential

which is integrated to

+ C ,
P ( ) = en f ()

(6.264)

with an operator integration constant C which commutes with the constant nP


and satisfies n C = 0 and and
) (0)

(
C = n (nP ) = n
.
2

(6.265)

Inserting this into (6.260) and integrating that equation yields


1
+ e2 n f 2 () + e n f ()
+D
,
2eC f ()
P ( ) =
2n
2


(6.266)

is again an interaction constant commuting with nP . With this, we can


where D
P and find
integrate the equation of motion (6.259) over d = d/2n
1
1
[
x( ) x(0)] =
2
(2nP )2

)
(

+D
+ e2 n f 2 () + e n f ()
.
d 2eC f ()

2
(0)
(6.267)
, which is reinserted into (6.266) to yield
This determines D
Z

1
P ( ) =
[
x ( ) x (0)]
2


Z (
)

e
2
2

h
d 2eC f () + e n f () + n f ()
i2

2
(0)
) (0)

)) . (6.268)
)) + e2 n f 2 ((
)) + e n f ((
+
2eC f ((

2
( ) (0)


Multiplying this by , and recalling (6.258) and (6.265), we find


1
[
x ( ) x (0)] +
2
Z (
)
en
+ en f ((
)).

d f ()

(0)
( ) (0)

P ( ) =

(6.269)

Inserting this into (6.264) determines the integration constant C :


Z (
)
en
1

d f ().
[
x ( ) x (0)]
C =

2
(0)
( ) (0)

(6.270)

It is useful to introduce the notation


1
hf i

( ) (0)
and

)
(

(0)

d f ().

h (f )2 i h (f hf i)2 i = h f 2 i h f i2 .

(6.271)

(6.272)

459

6.6 Green Function in External Electromagnetic Field

In order to calculate the matrix elements


0i = x P 2 + e F + M 2 x 0 ,
hx |H|x


2

(6.273)

we must time-order the operators x( ), x(0). For this we need the commutator
[
x ( ), x (0)] = 2i g .

(6.274)

This is deduced from Eq. (6.268) by commuting it with x( ) and using the trivially
), x ( )] = 0 as well as the nonequal-time
vanishing equal-time commutator [(

), x (0)] = 0, which
commutator [(0),
x ( )] = 2in . From the latter follows [(
is also needed for time-ordering. The result is



0i = 1 (x x )2 2 i + M 2 + e2 h()2 i 2 + e f () f ( ) .
hx |H|x
4 2


(6.275)

Integrating this in we obtain the exponential factor of the time-evolution amplitude


(6.211):
(


2
1
i
f () f ( )
E(x, x ; ) = 2 exp (xx )2 + M 2 +e2 h(f )2i i e
.

4

(6.276)

The time-independent prefactor C(x, x ) is again determined by the differential equation Eq. (6.216), which reduces here to
(x x )
h f i f () hx, |x 0i,
[i eA (x)] hx, |x 0i =

#

"

(6.277)

and is solved by
i
C(x, x ) =
exp ie
(4)2

dy

(xx )
A (y)

"Z

ny

#)!

f (y )
f (ny)
.
dy
ny
(6.278)

For a straight-line integration contour, the second term does not contribute, as
before.
Observe that in Eq. (6.276), the mass term M 2 is replaced by
2
Meff
= M 2 + e2 h(f )2 i,

(6.279)

implying that in an electromagnetic wave, a particle acquires a larger effective mass.


If the wave is periodic with frequency and wavelength = 2c/, the right-hand
side becomes M 2 + e2 h f 2 i. If the photon number density is , their energy density
is (in units with h
= 1), and we can calculate
e2 h f 2i = 4

h E 2i

= 4 .
2

(6.280)

460

6 Relativistic Particles and Fields in External Electromagnetic Potential

Hence we find a relative mass shift:


M 2
= 4
2e ,
M2

(6.281)

is the Compton wavelength of the


where e h
/Me c = 3.861592642(28) 103 A
electron. For visible light, the right-hand side is of the order of
A3 /100. Present
lasers achieve energy densities of 109 W/sec corresponding to a photon density
=

W
1 eV
1
109
2.082 107
.

sec

A3 h

(6.282)

which is too small to make M 2 /M 2 observable.

Appendix 6A

Spinor Spherical Harmonics

Equation (6.171) defines spinor spherical harmonics. In these, an orbital wave function of angular momentum l is coupled with spin 1/2 to a total angular momentum
j = l 1/2. For the configurations j = l + 1/2 with m2 = 1/2 the recursion
relation (4C.20) for the Clebsch-Gordan coefficients hs1 m1 ; s2 m2 |smi, simplifies by
having no second term. Inserting s1 = l , s2 = 1/2 and s = j = l + 1/2, we find
hl , m + 21 ; 21 , 12 |l + 12 , mi =

v
u
u l
t

m + 1/2
hl , m 12 ; 21 , 12 |l + 12 , m1i.
l m + 3/2

(6A.1)

This has to be iterated with the initial condition


hl , l ; 21 , 21 |l + 12 , l 21 i = 1,

(6A.2)

which follows from the fact that the state hl , l ; 21 , 12 i carries a unique magnetic
quantum number m = l 1/2 of the irreducible representation of total angular
momentum s = j = l + 1/2. The result of the iteration is
hl+ , m 21 ; 21 , 12 |l+ 12 , mi =

v
u
u l+
t

m + 1/2
.
2l+ + 1

(6A.3)

Similarly we may simplify the recursion relation (4C.21) for the configurations j =
l+ 1/2 with m2 = 1/2 to
hl , m 21 ; 21 , 21 |l + 12 , mi =

v
u
u l
t

+ m + 1/2
hl , m + 12 ; 21 , 21 |l + 12 , m+1i,
l + m + 3/2

(6A.4)

and iterating this with the initial condition


hl , l ; 12 n 21 |l + 12 , l + 21 i = 1,

(6A.5)

461

Notes and References

which expresses the fact that the state hl l ; 12 12 i is the state of maximal magnetic
quantum number m = l + 1/2 of the irreducible representation of total angular
momentum s = j = l + 1/2. The result of the iteration is
hl , m 12 ; 21 , 21 |l+ + 12 , mi =

v
u
u l+
t

+ m + 1/2
.
2l + 1

(6A.6)

Inserting (6A.3) and (6A.6) the expression (6.171) for the spinor spherical harmonic
of total angular momentum j = l + 1/2, which now reads
l

yj,m
(, ) = hl , m 12 ; 12 , 21 |l + 12 , mi Yl m1/2 (, )( 21 )
+ hl m + 21 ; 21 12 |l + 12 , mi Yl m+1/2 (, )( 12 ),

(6A.7)

and separating the spin-up and spin-down components, we obtain precisely (6.173).
In order to find corresponding result for j = l+ 1/2, we use the orthogonality
relation for states with the same l but different j = l 1/2:
hl + 21 , m|l 21 , mi = 0.

(6A.8)

Inserting a complete set of states in the direct product space yields


hl + 21 , m|l, m 12 ; 21 , 21 ihlm 12 ; 21 12 |l 12 , mi
+hl + 21 , m|l, m + 21 ; 21 , 12 ihl, m + 21 ; 12 , 12 |l 12 , mi = 0.

(6A.9)

Together with (6A.3) and (6A.6) we find


hl+ .m 12 ; 21 , 21 |l+ 12 , mi =
hl+ , m + 12 ; 21 , 12 |l+ 12 , mi =

v
u
u l+
t

+ m + 1/2
,
2l+ + 1

v
u
u l+
t

m + 1/2
,
2l+ + 1
(6A.10)

With this, the expression (6.171) for the spinor spherical harmonics written as
l

+
yj,m
(, ) = hl+ , m 12 ; 21 , 12 |l+ 12 , mi Yl,m1/2 (, )( 12 )
+ hl+ , m + 21 ; 21 , 12 |l+ 12 , mi Yl,m+1/2 (, )( 12 )

has the components given in (6.172).

Notes and References


[1] J. Schwinger, Phys. Rev. 82, 664 (1951); 93, 615 (1954); 94, 1362 (1954).
[2] C. Itzykson and J.B. Zuber, Quantum Field Theory, McGraw-Hill (1985).
[3] H. Kleinert, R. Ruffini, and X. SheShengPhys. Rev. D 78, 025011 (2008);
A. Chervyakov and H. Kleinert, Phys. Rev. D 80, 065010 (2009).
[4] M.P. Fry, Phys. Rev. D 45, 682 (1992).
[5] C. Itzykson and E. Brezin, Phys. Rev. D 2, 1191 (1970).

(6A.11)

Life can only be understood backwards,


but it must be lived forwards
S. Kierkegaard (1813-1855)

7
Quantization of Relativistic Free Fields
In Chapter 2 we have shown that the quantum mechanics of the N-body Schrodinger
equation can be replaced by a Schrodinger field theory, where the fields are operators satisfying canonical equal-time commutation rules. They were given in Sections
2.8 and 2.10 for bosons and fermions, respectively. The great advantage of this
reformulation of Schrodinger theory was that the field quantization makes the symmetrization or antisymmetrization of the N-body wave function automatic, which
in Schrodinger theory must be imposed from the outside in order to explain atomic
spectra and Bose-Einstein condensation. Here we shall generalize the procedure to
relativistic particles by quantizing the free relativistic fields of the previous section
following the above general rules. We define for each field (x, t) in a classical Lagrangian L(t) a canonical field momentum by the functional derivative (2.155) at a
given time
L(t)
.
(7.1)
(x, t) = px (t) =
t)
(x,
The fields are now made field operators by imposing the canonical equal-time commutation rules (2.142):
[(x, t), (x , t)] = 0,
[(x, t), (x , t)] = 0,
[(x, t), (x , t)] = i (3) (x x ).

(7.2)
(7.3)
(7.4)

For fermions, there will be corresponding anticommutation rules.


In these euations we have omitted the hat on top of the field operators, and we
shall do so from now on everywhere, for brevity, since most field expressions will
contain quantized fields. In the few cases where this is not the case, or where it is
not obvious that the fields are classical objects, we shall explicitly state this. We
shall also use natural units in which c = 1, h
= 1, except in some cases where cgs
units are helpful.
Before implementing the commutation rules explicitly it is useful to note that
all free Lagrangians constructed in the last chapter are local in the sense that they
are given by volume integrals over a Lagrangian density
L(t) =

d3 x L(x, t),
462

(7.5)

463

7.1 Scalar Fields

where L(x, t) is an ordinary function of the fields and their first spacetime derivatives. The functional derivative (7.1) may therefore be calculated as a partial derivative of the density L(x) = L(x, t):
(x) =

L(x)
,

(x)

(7.6)

Similarly, the Euler-Lagrange equations may be written down using only partial
derivatives
L(x)
L(x)

=
.
(7.7)

[ (x)]
(x)
Here and in the sequel we employ a four-vector notation for all spacetime objects
such as x = (x0 , x) where x0 = ct coincides with the time t in natural units.
The locality is an important property of all present-day quantum field theories
of elementary particles. The spacetime derivatives appear usually only in quadratic
terms. Since a derivative measures the difference of the field between neighboring spacetime points, the field described by a local Lagrangian propagates through
spacetime via a nearest-neighbor hopping. In field-theoretic models of solid-state
systems, such couplings are useful lowest approximations to more complicated shortrange interactions. In field theories of elementary particles, the locality is an essential
ingredient which seems to be present in all fundamental theories. All nonlocal effects
arise from higher-order perturbation expasions of local theories.

7.1

Scalar Fields

We begin by quantizing the field of a spinless particle. This is a scalar field (x)
that can be real or complex.

7.1.1

Real Case

The action of the real field was given in Eq. (4.164). As we shall see in Subsec. 7.1.6,
and even better in Subsec. 8.11.1, the quanta of the real field are neutral spinless
particles. For the canonical quantization we must use the real-field Lagrangian
density (4.166) in which only first derivatives of the field occur:
1
L(x) = {[(x)]2 M 2 2 (x)},
(7.8)
2
The associated Euler-Lagrange equation is the Klein-Gordon equation (4.167):
( 2 M 2 )(x) = 0.

7.1.2

(7.9)

Field Quantization

According to the rule (7.6), the canonical momentum of the field is


(x) =

L(x)

= 0 (x) = (x).

(x)

(7.10)

464

7 Quantization of Relativistic Free Fields

As in the case of non-relativistic fields, the quantization rules (7.2)(7.4) will eventually render a multiparticle Hilbert space. To find it we expand the field operator
(x, t) in normalized plane-wave solutions (4.177) of the field equation (7.9):
(x) =

Xh

fp (x)ap + fp (x)ap .

(7.11)

Observe that in contrast to a corresponding expansion (2.208) of the nonrelativistic Schrodinger field, the right-hand side contains operators for the creation and
annihilation of particles. This is an automatic consequence of the fact that the
quantization of a real field must lead to a hermitian field operator. This has the
physical consequence, that a quantized real field theory which can only exists in
the relativistic setting, will necessarily allow for the creation and annihilation of
particles,
Inserting the explicit wave functions (4.177), the expansion reads
(x) =

X
p

1
(eipx ap + eipx ap ),
2p0 V

(7.12)

Since the zeroth component of the four-momentum p in the exponent is on mass


shell, we should write px more explicitly as px = p t px. However, the notation
px is shorter, and there is little danger of confusion. If this is the case, we shall state
the on- or off-shell property of p0 explicitly.
The expansions (7.11) and (7.12) are complete. The first may be solved for ap
and ap with the help of the orthonormality relations (4.174) we obtain directly
ap = (p (t), (t)),

ap = (p (t), (t)).

(7.13)

More explicitly, this can be written as


(

ap
ap

ip0 x0

=e

2V p0

d3 x eipx i(x) + p0 (x) .

(7.14)

where we have inserted (x) = (x).


Using the canonical field commutation rules (7.2)(7.4) between (x) and (x),
we find for the coefficients of the plane-wave expansion (7.12) the commutation
relations
[ap , ap ] =
h

ap , ap

ap , ap = 0,

= p,p ,

(7.15)

making them creation and annihilation operators of particles of momentum p on a


vacuum state |0i defined by ap |0i = 0. Conversely, reinserting (7.15) into (7.12) we
verify the original local field commutation rules (7.2)(7.4).

465

7.1 Scalar Fields

In an infinite volume, one often uses in (7.11) the wave functions (4.178) with
continuous momenta, and work with a plane-wave decomposition in the form of a
Fourier integral
(x) =


d3 p 1  ipx
ipx
e
a
+
e
a
p
p ,
(2)3 2p0

(7.16)

with the identification [recall (2.216)]


ap =

ap =

2p0 V ap ,

2p0 V ap .

(7.17)

The inverse of the expansion is now


(

ap
ap

= eip

0x

d3 x eipx i(x) + p0 (x) .

(7.18)

This can of course be written in the same form as in (7.13) using the orthogonality
relations (4.179) for continuous-momentum wave functions (4.178):
ap = (fp , )t ,

ap = (fp , )t .

(7.19)

In the expansion (7.11), the commutators (7.15) hold with the replacement [recalling
the definition (1.195), and keeping h
, to be explicit]
(3)
p,p 2p0 - (p p ) = 2p0 (2h)3 (3) (p p ),

(7.20)

In the following we shall always prefer to work with a finite volume and sums
over p. If we want to find the large-volume limit of any formula, we can always
go over to the continuum limit by replacing the sums as in (2.256) by phase space
integrals:
X
p

p,p

d3 p

V
,
(2h)3

(2h)3 (3)

(p p ).
V

(7.21)
(7.22)

After the replacement (7.17), the volume V disappears in all physical quantities.
A single-particle state of a fixed momentum p is created by
|pi = ap |0i.

(7.23)

Its wave function is obtained from the matrix elements of the field
1
eipx = fp (x),
2V p0
1
hp|(x)|0i =
eipx = fp (x).
0
2V p
h0|(x)|pi =

(7.24)

466

7 Quantization of Relativistic Free Fields

As in the nonrelativistic case, the second-quantized Hilbert space is obtained


by applying the particle creation operators ap to the vacuum state |0i, defined by
ap |0i = 0, leading to states (2.218):
|np1 np2 . . . npk i = N S,A (ap1 )np1 (apk )npk |0i,

(7.25)

Multiparticle wave functions are obtained by matrix elements of the type (2.220).
In an infinite volume, one may use single-particle states
|pi = ap |0i.

(7.26)

with the vacuum |0) defined by ap |0) = 0. These states satisfy the orthogonality
relation
(3)
(p |p) = 2p0 - (p p),
(7.27)
and have the wave functions
(0|(x)|p) = eipx = fp (x),
(p|(x)|0) = eipx = fp (x).

(7.28)

Energy of Free Neutral Scalar Particles


Due to the local structure (7.5) of the Lagrangian, also the Hamiltonian can be
written as a volume integral
Z
H=

d3 x H(x).

(7.29)

The Hamiltonian density is the Legendre transform of the Lagrangian density


H(x) = (x) 0 (x) L(x)
1
M2 2
1 0
[ (x)]2 + [x (x)]2 +
(x).
=
2
2
2

(7.30)

Inserting the expansion (7.12), we obtain the Hamilton operator


H =

p,p




1

0 0

2
+ ap a
p p
a
a

+
p

p
+
M
p
p,p
p
p
2 p0 p0

i
h

1
0
0
0
0
+ 0 0 p,p p0 p0 p p + M 2 ap ap ei(p +p )t + ap ap ei(p +p )t
2 p p

X 
1

0
=
p ap ap +
.
(7.31)
2
p

The vacuum state |0i has an energy


E0 h0|H|0i =

1X 0 1X
p =
p ,
2 p
2 p

(7.32)

467

7.1 Scalar Fields

due to the zero-point oscillations of all the oscillator quanta in the second quantization formalism, the so-called vacuum fluctuations of the field. In the limit of large
volume, the momentum sum turns into an integral over the phase space according
to the usual rule (7.21).
This energy is infinite. In most circumstances, however, this infinity is unobservable. It is therefore often subtraced out of the Hamiltonian, replacing H by
: H : H h0|H|0i.

(7.33)

The double dots to the left-hand side define what is called the normal product form
of H. It is obtained by the following prescription on the operators a, a :
In an arbitrary product of creation and annihilation operators,
: a a a a :,
the double dots reorder this product in such a way that all creation operators stand
to the left of all annihilation operators. For example,
: ap ap + ap ap : = 2ap ap

(7.34)

and
: H :=

p0 ap ap .

(7.35)

In following this ad hoc procedure, care has to be taken that one is not dealing
with phenomena which are sensitive to the omitted zero-point oscillations. Gravitational interactions, for example, couple to the zero-point energy. The infinity creates
a problem when trying to construct quantum field theories in the presence of a classical gravitational field, since the vacuum energy gives rise to an infinite cosmologcal
constant, which has a finite value. A possible solution of this problem will appear
later when quantizing the Dirac field in Section 7.3. We shall see in Eq. (7.237) that
for the Dirac field, the vacuum energy has the same form as for the scalar field, but
with an opposite sign. In fact, this opposite sign is a consequence of the Fermi-Dirac
satatistics of the electrons. As will be shown in Section 7.8, all particles with halfinteger spins obey Fermi-Dirac statistics and require a different quantization than
that of the Klein-Gordon field. They all give a negative contribution p /2 to the
vaccum energy for each momentum and spin degree of freedom. All particle with
integer spins, however, will be quantized in a similar way as the Klein-Gordon field
and contribute a positive vacuum energy p /2 for each momentum and spin degree
of freedom.
The sum of bosonic and fermionic vacuum energies is therefore
tot
Evac
=

1 X
1 X
p
p .
2 p,bosons
2 p,fermions

(7.36)

468

7 Quantization of Relativistic Free Fields

Expanding p as
p =

p2

M2

1 M2 1 M4

+ ... ,
= |p| 1 +
2 p2
8 p4
!

(7.37)

we see that the vacuum energy can be finite if the universe contains as many Bose
fields as Fermi fields, and if the masses of the associated particles satisfy the sum
rules
X

M2 =

bosons

bosons

M 2,

fermions

M 4.

(7.38)

fermions

The higher powers in M contribute with finite momentum sums in (7.36).


If these cancellation do not occur, the sums are divergent and a cutoff in momentum space is needed to to make the sum over all vacuum energies finite. Summing
over all momenta inside a momentum sphere |p| of radius will lead to a divergent
energy proportional to 4 .
Many authors have argued that all quantum field theories may be valid only if
the momenta are smaller than the Planck momentum PP MP c = h
/P , where MP
is the Planck mass (4.384) and P the associated Compton wavelength
P =

h
G/c3 = 1.616252(81) 1033 cm

(7.39)

which is also called Planck length. Then the vaccum energy would be of the order MP4 1076 GeV4 . From present cosmological reexpansion rate one estimates
a cosmological constant of the order of 1047 GeV4 . This is smaller than the vacuum energy by a factor of roughly 10123 . The authors conclude that in the absence
of cancellations of Bose and Fermi vacuum energies, field theory gives a too large
cosmological constant by a factor 10123 . This conclusion, however, is false. As we
shall see later when treating other infinities of quantum field theories, any divergent
quantity is not given by the quantity made finite by a cutoff, but the infinity must
be included into the bare initial parameters of the theory, which has to be chosen
such that the final result is equal to the experimentally observed quantity. The
cutoff indicates that the quantity depends on ultra-short-distance physics which we
shall never know. For this reason one must always restrict one attention to theories
which do now depend on the ultra-short-distance physics. These theories are called
renormalizable theries. The quantum electrodynamics to be discussed in detail in
Chapter 12 was historically the first theory of this kind. In this theory, which is
experimentally the most accurate theory ever, the situation of the vacuum is precisely the same as for the mass of the elctron. Also here the mass emerging from
the interactions needs a cutoff to be finite. But all cutoff dependence is absorbed
into the initial bare mass parameter so that the final result is the experimentally
observed quantity.

469

7.1 Scalar Fields

If boundary conditions cause a modification of the sums over momenta, the


vacuum energy leads to finite observable effects which are independent of the cutoff.
These are the famous Casimir forces on conducting walls, or the Van der Waals
forces between different dielectric media. Both will be discussed in Section ??.

7.1.3

Propagator of Free Scalar Particles

The free-particle propagator of the scalar field is obtained from the vacuum expectation
G(x, x ) = h0|T (x)(x)|0i.

(7.40)

As for nonrelativistic fields, the propagator coincides with the Green function of the
free-field equation. This follows from the explicit form of the time-ordered product
T (x)(x ) = (x0 x0 )(x)(x ) + (x0 x0 )(x )(x).

(7.41)

Multiplying G(x, x ) by the operator 2 , we obtain


2 T (x)(x ) = T [ 2 (x)](x )
+2(x0 x0 ) [0 (x), (x )] + (x0 x0 ) [(x), (x )] .

(7.42)

The last term can be manipulated according to the usual rules of distributions: It is
multiplied by an arbitrary smooth test function of x0 , say f (x0 ), and becomes after
a partial integration:
Z

dx0 f (x0 ) (x0 x0 )[(x), (x )] =

dx0 f (x0 )(x0 x0 )[(x), (x )]

(7.43)
Z

dx0 f (x0 )(x0 x0 )[(x),


(x )].

The first term on the right-hand side does not contribute since it contains the commutator of the fields only at equal-times, [(x), (x )]x0 =x0 , where it vanishes. Thus,
dropping the test function f (x0 ), we have the equality between distributions

(x0 x0 )[(x), (x )] = (x0 x0 )[(x),


(x )],

(7.44)

and therefore with (7.4), (7.10), (7.40)(7.42):


( 2 M 2 )G(x, x ) = i(x0 x0 ) (3) (x x ) = i (4) (x x ).

(7.45)

To calculate G(x, x ) explicitly, we insert (7.12) and (7.41) into (7.40), and find
G(x, x ) = (x0 x0 )
+ (x0 x0 )
= (x0 x0 )

1 X
1

0 0 ei(pxp x ) h0|ap ap |0i
2V p,p p p
1 X
1

0 0 ei(pxp x ) h0|ap ap |0i
2V p,p p p

1 X 1 ip(xx )
1 X 1 ip(xx )

e
+
(x

x
)
e
. (7.46)
0
0
2V p p0
2V p p0

470

7 Quantization of Relativistic Free Fields

It is useful to introduce the functions


G() (x, x ) =

X
p

1 ip(xx)
e
,
2p0 V

(7.47)

They are equal to the commutators at arbitrary times of the positive- and negativefrequency parts of the field (x) which annihilate and create free single-particle
states, respectively, defined by
(+) (x) =

X
p

1
eipx ap ,
0
2p V

() (x) =

X
p

1
eipx ap .
0
2p V

(7.48)

In terms of these,
G(+) (x, x ) = [(+) (x), () (x )],

G() (x, x ) = [() (x), (+) (x )] = G() (x , x)


(7.49)

The commutators on the right-hand sides can of course also be replaced by the
corresponding expectation values. As such, they may also be rewritten as follows:
G

()

(x, x ) =

d3 p 1 ip(xx ) ()
e
G (p, t t ),
3
0
(2) 2p

with
G() (p, t t ) eip

being expectation values

(7.50)

0 (tt )

(7.51)

G(+) (p, t t ) = h0|ap H (t)ap H (t )|0i,

G(+) (p, t t ) = h0|ap H (t )ap H (t)|0i,

(7.52)

of Heisenberg creation and annihilation operators [recall the definition (1.283), and
(2.131)] of frequency p0 = p :
0

ap H (t) eiHt ap eiHt = eip t ap .

ap H (t) eiHt ap eiHt = eip t ap ,

(7.53)

In the infinite-volume limit, the functions (7.47) have the Fourier representation
G

()

(x, x ) =

d3 p 1 ip(xx )
e
.
(2)3 2p0

(7.54)

The full commutator [(x), (x )] receives contributions from both functions and is
given by
[(x), (x )] = G(+) (x x ) G() (x x ) C(x x ).

(7.55)

The right-hand side defines the commutator function, which has the Fourier representation
Z
d3 p 1 ip(xx )
e
sin[p0 (x0 x0 )].
(7.56)
C(x x ) = i
(2)3 2p0

471

7.1 Scalar Fields

This representation is convenient for verifying the canonical equal-time commutation


rules (7.2) and (7.4) which imply that
x ) = i (3) (x x ),
C(x

C(x x ) = 0,

for x0 = x0 .

(7.57)

Indeed, for x0 = x0 , the integrand is zero, so that the integral vanishes. The time
derivative of C(x x ) removes p0 in the denominator, leading at x0 = x0 directly
to the Fourier representation of the spatial -function.
Another way of writing (7.56) is via a formal off-shell momentum integral

C(x, x ) =

d4 p

2(p0 )(p2 M 2 )eip(xx )


4
(2)

(7.58)

In contrast to the Feynman propagator which satisfies the inhomogenous KleinGordon equation, the commutator function (7.55) satisfies the homogenous equation:
( 2 M 2 )C(x x ) = 0.

(7.59)

The Fourier representation (7.58) has obviously this property, since multiplication
from the left by the Klein-Gordon operator produces an integrand proportional to
(p2 M 2 )(p2 M 2 ) = 0.
The equality of the two expressions (7.56) and (7.58) follows from the property
of the -function
1
[(p0 p ) + (p0 + p )].
(7.60)
(p2 M 2 ) = (p02 p2) =
2p
which selectsfrom the integral over p0 containing (p0 ) precisely the on-shell value
p0 = p in agreement with (7.56).
In terms of the functions G() (x, x ), the propagator can be written as
G(x, x ) = G(x, x ) = (x0 x0 )G(+) (x, x ) + (x0 x0 )G() (x, x ).

(7.61)

As follows from (7.54) and (7.61), all three functions G(x, x ), G(+) (x, x ), and
G() (x, x ) depend only on x x , thus exhibiting the translational invariance of
the vacuum state. In the following, we shall therefore always write one argument
x x instead of x, x .
As in the nonrelativistic case, it is convenient to use the integral representation
(1.318) for the -functions
dE
i
0

ei(Ep )(x0 x0 ) ,
0
2 E p + i
Z
i
dE
0

ei(E+p )(x0 x0 ) ,
(x0 x0 ) =
0
2 E + p i
(x0 x0 ) =

(7.62)

with an infinitesimal parameter > 0, to reexpress G(x, x ) from (7.54), (7.51), and
(7.61) more compactly as

G(x x ) =

dE d3 p 1
2 (2)3 2p0

i
i

eiE(x0 x0 )+ip(xx ) ,

0
0
E p + i E + p i
(7.63)

472

7 Quantization of Relativistic Free Fields

where we have used the fact that p0 is an even function of p, thus permitting us to
change the integration variables p to p in the second integral. By combining the
denominators we find
Z
i
dE d3 p

eiE(x0 x0 )+ip(xx ) .
(7.64)
3
2
2
2
2 (2) E p M + i
This has a relativistic invariant form in which it is useful to rename E as p0 , and
write
Z
d4 p
i

G(x x ) =
eip(xx ) .
(7.65)
4
2
2
(2) p M + i
Note that in this expression, p0 is integrated over the entire p0 -axis. In contrast to
all earlier formulas in this section, it is no longer constrained to satisfy the mass
2
shell conditions (p0 )2 = p0 = p2 + M 2 .
The integral representation (7.65) shows very directly that G(x, x ) is the Green
function of the free-field equation (7.9). An application of the differential operator
( 2 M 2 ) cancels the denominator in the integrand and yields

i
d4 p 2
2
ip(xx )
(p

M
)
e
(2)4
p2 M 2 + i
= i (4) (x x ).
(7.66)

( 2 M 2 )G(x x ) =

This calculation gives rise to a simple mnemonic rule for finding the Green function:
for an arbitrary free-field theory: We write the Lagrangian density (7.8) as
1
(7.67)
L(x) = (x)L(i)(x),
2
with the differential operator
L(i) ( 2 M 2 ).

(7.68)

The Euler-Lagrange equation (7.9) can then be written simply as


L(i)(x) = 0.

(7.69)

The Green function is now the Fourier transform of the inverse of L(i)
G(x x ) =

d4 p i ip(xx )
e
,
(2)4 L(p)

(7.70)

with an infinitesimal i added to the mass of the particle. This rule can directly
be generalized to all other free-field theories to be discussed in the sequel.

7.1.4

Complex Case

Here we use the Lagrangian (4.162):


L(x) = (x) (x) M 2 (x)(x),

(7.71)

whose Euler-Lagrange equation is the same as in the real case, Eq. (7.9). As in
the real case, we refer to Subsecs. 7.1.6 and 8.11.1 to show that the complex field
describes charged spinless particles.

473

7.1 Scalar Fields

Field Quantization
According to the canonical rules, the complex field possess complex canonical momenta
L(x)
= 0 (x),
(x) (x)
[ 0 (x)]
L(x)
= 0 (x).
(7.72)
(x) (x)
[ 0 (x)]
The associated operators are required to satisfy the equal-time commutation rules
[(x, t), (x , t)] = i (3) (x x ),
[ (x, t), (x , t)] = i (3) (x x ).

(7.73)

All other equal-time commutators vanish. We now expand the field operator into
its Fourier components
(x) =

X
p



1
ipx
ipx
e
a
+
e
b
p
p ,
2V p0

(7.74)

where, in contrast to the real case (7.12), bp is no longer equal to ap . The reader may
wonder why we do not just use another set of annihilation operators dp rather than
creation operators bp . A negative sign in the momentum label would be appropriate

since the associated wave function fp (x) = eipx / 2V p0 has a negative momentum.
One formal reason for using bp instead of dp is that this makes the expansion (7.74)
the straightforward generalization of the expansion (7.12) of the real field. A more
physical reason will be seen below when discussing Eq. (7.83).
In analogy with (7.14), we invert (7.74) to find
(

ap
bp

ip0 x0

=e

1
2V p0

d3 x eipx i (x) + p0 (x) ,

(7.75)

and the corresponding hermitian-adjoint operators ap and bp .


From these equations we find the operators ap , ap , bp , and bp all commute with
each other, except for
h

bp , bp = p,p .

ap , ap = p,p ,

(7.76)

There are two types of particle states, created by


|pi ap |0i,
hp| h0|ap,

|
pi bp |0i,
h
p| h0|bp .

(7.77)

They have the wave functions


1
eipx = fp (x),
2V p0
1
h
p|(x)|0i =
eipx = fp (x),
0
2V p

h0|(x)|pi =

1
eipx = fp (x),
2V p0
1
hp| (x)|0i =
eipx = fp (x). (7.78)
0
2V p
h0|(x)|
pi =

474

7 Quantization of Relativistic Free Fields

We are now ready to justify why we associated with the second term in the
expansion (7.74) a creation operator bp rather than an annihilation operator dp.
First, this is in closer analogy with the real field in (7.12) which contains a creation
operator ap accompanied by the negative-frequency wave function eipx . Second, only
the above choice leads to commutation rules (7.76) for bp , bp with the correct sign.
The choice dp would have led to the commutator [dp , dp ] = p,p . For the secondquantized Hilbert space, this would imply the norm of the states created by dp to be
negative. Third, the above choice ensures that the wave functions of incoming states
|pi and |
pi are both eipx , i.e., they both oscillate in time with positive frequency
0
like eip t , thus having a positive energy. With annihilation operators dp instead of
bp , the energy of a state created by dp would have been negative. This will also
emerge in the calculation of the second-quantized energy to be performed now.
Multiparticle states are formed in the same way as in the real case in Eq. (7.25),
except that creation operators ap and bp have to be applied to the vacuum state |0i:
|np1 np2 . . . npk ; n
p1 n
p2 . . . n
pk i = N S,A (ap1 )np1 (apk )npk (bp1 )n p1 (bpk )n pk |0i.
(7.79)
As in the real-field case, we shall sometimes use in an infinite volume the singleparticle states
|p) = ap |0),
|
p) = bp |0),
(7.80)
with the vacuum |0) defined by ap |0) = 0 and bp |0) = 0. These states satisfy
the same orthogonality relation as the those in (7.27), and have wave functions
normalized as in Eq. (7.28).

7.1.5

Energy of Free Charged Scalar Particles

The second-quantized energy density reads


H(x) = (x) 0 (x) + (x) 0 (x) L(x)
= 2 0 (x) 0 (x) L(x)
= 0 (x) 0 (x) + x (x)x (x) + M 2 (x)(x).

(7.81)

Inserting the expansion (7.74), we find the Hamilton operator [analogous to (7.31)]
H=

d x H(x) =

X
p

"

p0 + p2 + M 2
(ap ap + bp bp )
0
2p

#
2

p0 + p2 + M 2  2ip0 t
0
+
+ bp ap e2ip t (7.82)
ap bp e
2p0

X
p

p0 ap ap + bp bp + 1 .

475

7.1 Scalar Fields

Note that the mixed terms in the second line appear with opposite momentum
labels, since H does not change the total momentum. Had we used dp rather than
bp in the field expansion, the energy would have been
H=

d x H(x) =

X
p

"

p0 + p2 + M 2
(ap ap + dp dp )
2p0
2

p0 + p2 + M 2
0
0
(ap dp e2ip t + dp ap e2ip t )
+
0
2p

X
p

(7.83)

p0 ap ap + dp dp .

Here the mixed terms have lables with equal momenta. Since the commutation rule
of dp , dp have the wrong sign, the eigenvalues of dp dp take negative integer values
implying the energy of the particles created by dp to be negative. For an arbitrary
number of such particles, this would imply energies without lower bound, which is
unphysical (since it would allow one to build a perpetuum mobile)
Taking the vacuum expectation value of the positive-definite energy (7.82), we
obtain as in Eq. (7.32)
E0 h0|H|0i =

X
p

p0 =

p ,

(7.84)

In comparison with Eq. (7.32), there is a factor 2 since the complex field has twice
as many degrees of freedom as a real field. As in the real case, we may subtract this
infinite vacuum energy to obtain finite expressions via the earlier explained normalordering prescription. Only the gravitational interaction is sensitive to the vacuum
energy.
What is the difference between the particle states |pi, |
pi created by ap and
bp ? We shall later see that they couple with opposite sign to the electromagnetic
field. Here we only observe that they have the same intrinsic properties under the
Poincare group, i.e. the same mass and spin. They are said to be antiparticles of
each other.
Propagator of Free Charged Scalar Particles
The propagator of the complex scalar field can be calculated in the same fashion as
for real fields with the result
G(x, x ) G(x x ) = h0|T (x) (x )|0i
Z
i
d4 p

eip(xx ) .
=
4
2
2
(2) p M + i

(7.85)

As before, the propagator is equal to the Green function of the free-field equation
(7.9). It may again be obtained following the mnemonic rule on p. 472, by Fourier
transforming the inverse of the differential operator L(i) in the field equation [recall
(7.355) and (7.70)].

476

7.1.6

7 Quantization of Relativistic Free Fields

Behavior under Discrete Symmetries

Let us now see how the discrete operations of space inversion, time reversal, and
complex conjugation of Subsecs 4.5.24.5.4 act in the Hilbert space of quantized
scalar fields.
Space Inversion
Under a space inversion
x = x = (x0 , x),

(7.86)

a real or complex scalar field is transformed according to


P

(x)
P (x) = P (
x),

(7.87)

P = 1.

(7.88)

with
For a real field operator (x), this transformation can be achieved by assigning to the
creation and annihilation operators ap , ap in the expansion (7.12) a transformation
law
ap = P ap ,

a p = P ap .

(7.89)

Indeed, by inserting (7.89) into (7.12) we find



X
1
ipx
ipx
e
a
+
e
a
p
p
2p0 V p
X
1
(eipx ap + eipx ap ) = P (
x).
= P 0
2p V p

P (x) = P

(7.90)

It is possible to find a parity operator P in the second-quantized Hilbert space


which generates the parity the transformations (7.89), (7.90), i.e., it is defined by
Pap P 1 ap = P ap ,

Pap P 1 ap = P ap .

(7.91)

An explicit representation of P is
P = eiGP /2 ,

GP =

Xh
p

ap ap P ap ap .

(7.92)

In order to prove that this operator has the desired effect, we form the operators

a+
p ap + P ap ,

a
p ap P ap ,

(7.93)

477

7.1 Scalar Fields

which are invariant under the parity operation (7.91), with eigenvalues 1:
1

Pa
= a
p P
p ,

(7.94)

Then we calculate the same property with the help of the unitary operator (7.92),
iGP /2
expanding eiGP /2 a
with the help of Lies expansion formula (4.103).
p e
This turns out to be extremely simple since the first commutators on the right-hand
side of (4.103) are

[GP , a+
p ] = 0;

[GP , a
p ] = 2ap ,

(7.95)

so that the higher ones are obvious. The Lie series yields therefore
iGP /2

eiGP /2 a+
= a+
p e
p ,

1 2
i

+ . . . = a
= a
p e
p , (7.96)
2!

showing that the operators a


p do indeed satisfy (7.94).
For a complex field operator (x), the operator GP in (7.92) has to be extended
by the same expression containing the operators bp and bp .
Let us denote the multiparticle sates of the type (7.25), (7.79) in the secondquantized Hilbert space collectively by |i. In the Hilbert space of all such states,
the operator P is unitary:
P = P 1 ,
(7.97)


iGP /2

eiGP /2 a
= a
1 i +
p e
p

i.e., for any two states |1 i and |2 i, the scalar product remains unchanged by P

P h1 |2 iP

= h1 |P P|2 i = h1 |2 i.

(7.98)

With this operator, the transformation (7.87) is generated by


P

(x)
P(x)P 1 = P (x) = P (
x).

(7.99)

P = ()l .

(7.100)

As an application, consider a state of two identical spinless particles which move


in the common center of mass frame with a relative angular momentum l. Such a
state is an eigenstate of the parity operation with eigenvalue
This follows directly by applying the parity operator P to the wave function
|lm i =

dp Rl (p)

Ylm (
d2 p
p) ap ap |0i,

(7.101)

where Rl (p) is some radial wave function of p = |p|. Using (7.91), the commutativity
of the creation operators ap and ap with each other, and the well-known fact1
differ by a phase factor (1)l , we find
that the spherical harmonic for
p and p
immediately:
P|lm i = (1)l |lmi.
(7.102)
For two different bosons of parities P1 and P2 , the right-hand side is multiplied by
P1 P2 .
1

This follows from the property of the spherical harmonic Ylm ( , ) = (1)lm Ylm (, )
together with eim(+) = (1)m eim .

478

7 Quantization of Relativistic Free Fields

Time Reversal
The operator implementation of time reversal is not so straightforward. In a timereversed state all movements are reversed, i.e., for spin zero particles all momenta
are reversed (see Subsection 4.5.3). In this respect, there is no difference to the
parity operation, and the transformation properties of the creation and annihilation
operators are just the same as in (7.91). The associated phase factor is denoted by
T :
ap = T ap ,
ap = T ap .

(7.103)

We have indicated before [see Eq. (4.223)] and shall see below that, in contrast to
the phase factor P = 1 of parity, the phase factor T will not be restricted to 1
by the group structure. It is unmeasurable and can be chosen to be equal to unity,
If we apply the operation (7.103) to the Fourier components in the expansion
(7.12) of the field we find
1
2p0 V
p
X
1
0
= T
2p V
p

T (x) = T

= T (
x),

eipx ap + eipx ap

eipx ap + eipx ap

(7.104)

which is not yet the correct transformation law (4.203):


T

(x)
T (x) = T (xT ).

(7.105)

This is achieved by requiring the time reversal transformation to be an antilinear


operator T in the multiparticle Hilbert space. It is defined by
T ap T

T ap T

1
1

ap = T ap ,

ap = T ap ,

(7.106)

and by the antilinear property


T (a + ap )T

= T ap T

+ T ap T

(7.107)

Whenever the time reversal operation is applied to a combination of creation and


annihilation operators, the coefficients have to be switched to their complex conjugates. Under this antilinear operation, we indeed obtain (7.105) rather than (7.104):
T

(x)
T (x)T

= T (x) = T (xT ).

(7.108)

479

7.1 Scalar Fields

The same transformation law is found for complex scalar fields by defining for
particles and antiparticles
T ap T

T bp T

ap = T ap ,

bp = T bp .

(7.109)

In the second-quantized Hilbert space consisting of states (7.25) and (7.79), the
antiunitarity has the consequence that all scalar products are changed into their
complex conjugates:
T

h2 |1 i
T h2 |1 iT = h1 |2 i ,

(7.110)

thus complying with the property (4.206) of scalar products of ordinary quantum
mechanics (to which the second-quantized formalism reduces in the one-particle
subspace). This makes the operator T antiunitary.
To formalize operations with the antiunitary operator T in the Dirac bra-ket
language, it is useful to introduce an antilinear unit operator 1A which has the
effect
h2 |1A |1 i 1A h2 |1 i h2 |1 i .
(7.111)
The antiunitarity of the time reversal transformation is then expressed by the equation
T = T 1 1A ,
(7.112)
the operator 1A causing all differences with respect to the unitary operator (7.97).
Due to the antilinearity, the factor T in the transformation law of the complex
field is an arbitrary phase factor. We have mentioned this before in Subsec. 4.5.3. By
applying the operator T twice to the complex field (x), we obtain from (7.107)and
(7.108):
T 2 (x)T 2 = T T (x),
(7.113)
so that the cyclic property of time reversal T

= 1 fixes

T T = 1.

(7.114)

This is in contrast to the unitary operators P and C where the phases factors must
be 1.
For real scalar fields, this still holds, since the second transformation law (7.106)
must be the complex-conjugate of the first, which requires T to be real leaving only
the choices 1.
The antilinearity has the consequence that if a Hamilton operator is invariant
under time reversal
T HT 1 = H,
(7.115)
the time evolution operator U(t, t0 ) = eiH(tt0 )/h satisfies
T U(t, t0 )T

= U(t0 , t) = U (t, t0 ).

(7.116)

480

7 Quantization of Relativistic Free Fields

i.e., it has the time order inverted.


The antiunitary nature of T makes the phase factor T of charged fields, an
unmeasurable quantity. This why we were able to choose T = 1 after Eq. (7.114)
[see also (4.223)] Time reversal invariance does not produce selection rules in the
same way as discrete unitary symmetries do, such as parity. This will be discussed
in detail in Section 9.7, after having developed the theory of the scattering matrix.
Charge Conjugation
The operator implementation of the transformation (4.224) on a complex scalar field
operator


X
1
ipx
ipx

(x) =
e
a
+
e
b
(7.117)
p
p
2V p0
p

is quite simple. We merely have to define a charge conjugation operator C by


Cap C 1 = C bp ,

Cap C 1 = C bp ,

Cbp C 1 = C ap ,

Cbp C 1 = C ap ,

(7.118)

where C = 1 is the charge parity of the field, and see that C produces the desired
transformation
C
(x)
C(x)C 1 = C (x) = C (x),
(7.119)
as in (4.224). For a real field, we may simply identify bp with ap , thus making the
transformation laws (7.118) trivial:
Cap C 1 = C ap ,

Cap C 1 = C ap ,

(7.120)

The antiparticles created by bp have the same mass as the particles created by
ap , but the opposite charge. The latter follows directly from the transformation law
(4.226) of the classical local current. In the present second-quantized formulation
it is instructive to calculate the matrix elements of the operator current density
(4.168),

j (x) = i ,

(7.121)

between single-particle states created by ap and bp . We find


1
i(pp )x
e
,
2p0 2p0
1

h
p |j (x)|
pi = h0|bp j (x)bp |0i = (p + p ) 0 0 ei(pp )x .
2p 2p
hp |j (x)|pi = h0|ap j (x)ap |0i = (p + p )

(7.122)

The charge of the particle states |pi = ap |0i Rand |


pi = bp |0i is given by the diagonal
matrix element of the charge operator Q = d3 x j 0 (x) between these states.

hp |Q|pi =

d xhp |j (x)|pi = p ,p ,

h
p |Q|
pi =

d3 xh
p |j 0 (x)|
pi = p ,p .

(7.123)

481

7.2 Spacetime Behavior of Propagators

The opposite sign of the charges of particles and antiparticles is caused by the
opposite exponentials in the field (x) in (7.117) accompanying the annihilation
and creation operators ap and bp , respectively
As far as physical particles are concerned, the states |pi = ap |0i may be identified
with mesons of momentum p, the states |
pi = bp |0i with () -mesons. The scalar
particles associated with the negative frequency solutions of the wave equation are
called antiparticles. In this nomenclature, the particles of a real field are antiparticles
of themselves.
Note that the two matrix elements in (7.122) have the same exponential factors
which is necessary for particles and antiparticles to have the same energies and
momenta.
An explicit representation of the unitary operator C is
C = eiGC /2 ,

GC =

Xh
p

bp ap + ap bp P (ap ap + bp bp ) .

(7.124)

The proof is completely analogous to the parity case in Eqs. (7.92)(7.96).


A bound state of a boson and its antiparticle in a relative angular momentum l
has a charge parity (1)l . This follows directly from the wave function
|i =

dp Rl (p)

Ylm (
d2 p
p)ap bp |0i,

(7.125)

where Rl (p) is some radial wave function of p = |p|. Applying C to this we have
Cap bp |0i = bp ap |0i

(7.126)

and must interchange the order of bp and ap , as well as the momenta p and p, to
get back to the original state. This results in a sign factor (1)l from the spherical
harmonic. An example is the -meson which is a p-wave resonance of a (+) - and a
() -meson, and has therefore a charge parity C = 1.

7.2

Spacetime Behavior of Propagators

Let us evaluate the integral representations (7.65), (7.85) for the propagators of
real and complex scalar fields. To do this we use two methods which will be applied
many times in this text. The integrand in (7.65),(7.85) has two poles in the complex
p0
-plane as shown in Fig. 7.1, one at p0 = p2 + M 2 i, the other at p0 =
p2 + M 2 +i with > 0. The first is due to the intermediate single-particle state
propagating in the positive time direction, the other to the antiparticle propagating
in the negative time direction. A propagator with such pole positions is called a
Feynman propagator.

7.2.1

Wick Rotation

The special pole positions make it possible to rotate the contour of integration in
the p0 -plane without crossing a singularity so that it runs along the imaginary axis

482

7 Quantization of Relativistic Free Fields

Figure 7.1 Pole positions in the complex p0 -plane in the integral representations (7.65)
and (7.85) Feynman propagators.

dp0 = idp4 . This procedure is called a Wick rotation, and is illustrated in Fig.
7.2. Volume elements d4 p = dp0 dp1 dp2 dp3 in Minkowski space go over into volume
elements i dp1 dp2 dp3 dp4 i d4 pE in euclidean space. By this procedure, the integral

Figure 7.2 Wick rotation of the contour of integration in the complex p0 -plane

representations (7.65), (7.85) for the propagator become

G(x x ) =

d 4 pE
1

eipE (xx )E .
2
4
2
(2) pE + M

(7.127)

where pE is the four-momentum


pE = (p1 , p2 , p3 , p4 = ip0 )

(7.128)

and p2E the square of the momentum pE calculated in the euclidean metric, i.e.
2
2
2
2
p2E p1 + p2 + p3 + p4 . The vector xE is the corresponding euclidean spacetime
vector
xE = (x1 , x2 , x3 , x4 = ix0 ),
(7.129)
so that px = pE xE . With p2E +M 2 being strictly positive, the integral (7.127) is now
well-defined. Moreover, the denominator possesses a simple integral representation

483

7.2 Spacetime Behavior of Propagators

in the form of an auxiliary integral over an auxiliary parameter which, as we shall


see later, plays the role of the proper time of the particle orbits:
1
=
2
pE + M 2

d e (pE +M ) .

(7.130)

Since this way of reexpressing denominators in propagators is extremely useful in


quantum field theory, it is referred to after its author as Schwingers proper-time
formalism [1]. The -integral has turned the integral over the denominator into a
quadratic exponential function. The exponent can therefore be quadratically completed,
2
2

2
eipE (xx )E pE e(xx )E /4 p
with pE = pE i(x x )E /2 . Since the measure of integration is translationally invariant, and the integrand has become symmetric under four-dimensional rotations,
we can replace
Z

d 4 pE =

d4 pE = 2

dpE2 pE2 ,

(7.131)

and integrate out the four-momentum pE , yielding


1 Z d (xx )2 /4 M 2
E
G(x x ) =
.
e
16 2 0 2

(7.132)

The integral is a superposition of nonrelativistic propagators of the type (2.430),


2
with Boltzmann-like weights eM , and an additonal weight factor 2 which may
be viewed as the effect of an entropy factor eS = e2 log . This representation has an
interesting statistical interpretation. After deriving Eq. (2.430) for a nonrelativistic
propagator at imaginary times we observed that this looks like the probability for
a random walk of a fixed length proportional to h
to go from x to x in three
dimensions. Here we see that the relativistic propagator looks like a random walk of
arbitrary length in four dimensions, with a length distribution ruled mainly by the
above Boltzmann factor. This leads to the possibility of describing, by relativistic
quantum field theory, ensembles of random lines of arbitrary length. Such lines
appear in many physical systems in the form of vortex lines and defect lines [2, 3, 4].
By changing the variable of integration from to = 1/4 , this becomes
1
G(x x ) = 2
4

d e(xx )E M

2 /4

(7.133)

K ( ab),

(7.134)

Now we use the integral formula


Z

b
d ab/4
e
=2

4a

!/2

484

7 Quantization of Relativistic Free Fields

where K (z) is the modified Bessel function,2 we find




2
M K1 M (x x )E

q
G(x x ) = 2
.
4
M (x x )2
2

(7.135)

The D-dimensional generalization of this is calculated in the same way. Equation


(7.132) becomes
1
G(x x ) =
(4)D/2

d (xx )2 /4 M 2
E
,
e
D/2

(7.136)

yielding via (7.134)

G(x x ) =

dD q
1

eiq(xx )
2
D
2
(2) qE + M

D/21

M
1
q

=
(2)D/2
(x x )2

7.2.2

KD/21 M (x x )2E .

(7.137)

Feynman Propagator in Minkowski Space

Let us now do the calculation in Minkowski space. There the proper-time representation of the propagator (7.65), (7.85) reads
Z Z
d4 p i[p(xx) (p2 M 2 +i)]
i
d4 p
ip(xx )
d
e
=
i
e
.
(2)4 p2 M 2 + i
(2)4
0
(7.138)
The momentum integrations can be done, after a quadratic completion, with the
help of Fresnels formulas

G(xx ) =

1
dp0 ip0 2 a
,
e
= q
2
2 a/i

1
dpi ipi 2 a
.
e
= q
2
2 a/i

(7.139)
2

In these, it does not matter (as it did before in the Gauss integrals) that p0 and
2
pi appear with opposite signs in the exponent.
In continuing the evaluation of (7.138) it is necessary to make sure that the
integrals over dp and d can be interchanged. The quadratic completion generates
2
a term ei(xx ) /4 which is integrable for 0 only if we replace (x x )2 by
(x x )2 i. Changing again to = 1/4 we arrive at the integral representation
for the Feynman propagator
G(x x ) =
2

i
4 2

2 i]+(M 2 i)/4}

d ei{[(xx )

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 3.471.9.

(7.140)

485

7.2 Spacetime Behavior of Propagators

The two small imaginary parts are necessary to make the integral convergent at
both ends. The integral can be done with the help of the formula3
Z

b
d i(a+b/4)
e
=2

4a

!/2

i ei/2 H(1) ( ab),


2

(7.141)

where H(1) (z) is the Hankel function of the first kind. Upon taking the complex
conjugate, using4
[H(1) (z)] = H(2) (z ),

(7.142)

and replacing a by a and b by b , we obtain


(2)

M H1 (M (x x )2 )

q
.
G(x x ) = i
8
M (x x )2
2

(7.143)

The Hankel functions are combinations of Bessel and Neumann functions


H(1,2) (z) = J (z) iN (z).

(7.144)

In order to check the correctness of this expression, we go to the nonrelativistic limit


by letting c in the argument of the Hankel function. Displaying the true cgs
units, for clarity, this becomes
Mq
Mq 2
Mc2 (t t ) 1 M (x x )2

.
(x x )2 =
c (t t )2 (x x )2
h

2 h
2(t t )
(7.145)
For large arguments, the Hankel functions behave like
H(1,2) (z)

2 i(z/2/2)
e
,
z

(7.146)

so that (7.143) becomes for t > t , taking account the nonrelativistic limit (4.153)
of the fields,
c

G(x x )

1
1
iM
(x x )2 .
q
3 exp
2M 2ih(t t )/M
h
2t


(7.147)

in agreement with the nonrelativistic propagator (2.241).


q
q
In the spacelike regime (x x )2 < 0, we continue (x x )2 to i (x x )2 ,
(2)

(1)

and H1 (z) = H1 (z) together with the relation5

i H(1) (iz)
K (z)
=

2 (iz)
z
3

ibid., Formula 3.471.11, together with 8.476.9.


The last two identities are from ibid., Formulas 8.476.8 and 8.476.11.
5
ibid., Formula 8.407.1.
4

(7.148)

486

7 Quantization of Relativistic Free Fields

to rewrite (7.143) as
q

M K1 (M (x x )2 )

q
G(x x ) = 2
,
4
M (x x )2

(x x )2 < 0.

(7.149)

in agreement with (7.135).


It is instructive to study the massless limit M 0, in which the asymptotic
behavior6 K(z) 1/z leads to
G(x x )

1
1
.
4 2 (x x )2E

(7.150)

To continue this back to Minkowski space where (xx )2E = (xx )2 , it is necessary
to remember the small negative imaginary part on x2 which was necessary to make
the -integral (7.140) converge. The correct massless Green function in Minkowski
space is therefore
G(x x ) =

1
1
.
2
4 (x x )2 i

(7.151)

The same result is obtained from the M 0 -limit of the Minkowski space expression
(7.143), using the limiting property7
i
1
H(1) (z) H(2) (z) ()
.

(z/2)

7.2.3

(7.152)

Retarded and Advanced Propagators

Let us contrast the spacetime behavior of the Feynman propagators (7.143) and
(7.151) from that of the retarded propagator of classical electrodynamics. A retarded
propagator is defined for an arbitrary interacting field (x) by the expectation value
GR (x x ) (x0 x0 )h0|[(x), (x)]|0i.

(7.153)

In general, one defines a commutator function C(x x ) by


C(x x ) h0|[(x), (x)]|0i,

(7.154)

GR (x x ) = (x0 x0 )C(x x ),

(7.155)

and has the relation


For a free field (x), the commutator is a c-number, so that the vacuum expectation
values can be omitted, and C(x x ) is given by (7.55), leading to a retarded
propagator
GR (x x ) = (x0 x0 )[G(+) (x x ) G() (x x )],
(7.156)
6

M. Abramowitz and I. Stegun, Handbook of Mathematical Functions, Dover, New York, 1965,
Formula 9.6.9.
7
ibid., Formula 9.1.9.

487

7.2 Spacetime Behavior of Propagators

The Heaviside function in front of the commutator function C(x x ) ensures


the causality of this propagator. It also has the effect of changing C(x x ) which
solves the homogenous Klein-Gordon equation into a solution of the inhomogenoous
equation.
In fact, by comparing the (7.156) with (7.61) and using the Fourier
representations (7.62) of the Heaviside functions, we find for the retarded propagator
a representation very similar to that of the Feynman propagator in (7.63), except
that the i in the negative-energy pole term is reversed:
GR (x x ) =

dE d3 p 1
2 (2)3 2p0

i
i

eiE(x0 x0 )+ip(xx ) ,

0
0
E p + i E + p + i
(7.157)

By combining the denominators we now find, instead of (7.64),


Z

i
dE d3 p

eiE(x0 x0 )+ip(xx ) ,
3
2
2
2
2 (2) (E + i) p M

(7.158)

which may be written as an off-shell integral as in (7.65):


GR (x x ) =

i
d4 p

eip(xx ) ,
2
4
2
(2) p+ M + i

(7.159)

where the subscript of p+ indicates that a small term i has been added to p0 .
Note the difference in the Fourier representation with respect to that of the commutator function in (7.58), which solves the homogenous Klein-Gordon equation.
It is important to realize that the retarded Green function could be derived from
a time-ordered expectation value of second-quantized field operators if the Fourier
components of the negative frequencies were associated with annihilation operators
dp rather than creation operators bp of antiparticles. Then the propagator would
vanish for x0 < x0 , implying both poles in the p0 -plane to lie below the real energy
axis.
For symmetry reasons, one also introduces an advanced propagator
GA (x x ) (x0 x0 )h0|[(x), (x)]|0i.

(7.160)

It has the same Fourier representation as GR (x x ), except that the poles lie both
above the real axis:
GA (x x ) =

d4 p
i

eip(xx ) .
2
4
2
(2) p M

(7.161)

There will be an application of this Green function in Subsec. 12.12.1.


Let us calculate the spacetime behavior of the retarded propagator in spacetime.
We shall fist look at the massless case familiar from classical electrodynamics. For
M = 0, the integral representation (7.159) reads
GR (x x )

d4 p ip(xx ) i
,
e
(2)4
p2+

(7.162)

488

7 Quantization of Relativistic Free Fields

but with the poles in the p0 -plane at


p0 = p0 = p = |p|.

(7.163)

both placed below the real axis. This is indicated by writing p2+ in the denominator
rather than p2 , the plus sign indicating an infinitesimal +i added to p0 .
In the Feynman case, the p0 -integral is evaluated with the decomposition (7.63)
and the integral representation (7.62) of the Heaviside functionas follows:
dp0 ip0 (x0 x0 ) i
1
1
(7.164)
e
0
0
2
2p p p + i p + p i
1 ip |x0 x0 |
1
0
0
0
0
[(x0 x0 )eip (x x ) + (x0 x0 )eip (x x ) ] =
e
.
=
2p
2p
!

In contrast, the retarded expression reads


Z

dp0 ip0 (x0 x0 ) i


e
2
2p

1
1
0
0
p p + i p + p + i
1 ip (x0 x0 )
0
0
= (x0 x0 )
[e
eip (x x ) ].
2p

(7.165)

Thus we may write


d3 p

0
0
eip(xx ) eip |x x | ,
(7.166)
3
(2) 2p
Z
d3 p

0
0
0
0
eip(xx ) [eip (x x ) eip (x x ) ]. (7.167)
GR (x x ) = (x0 x0 )
3
(2) 2p

G(x x ) =

In the retarded expression we recognize, after the factor (x0 x0 ), the massless
limit of the commutator function C(x, x ) = C(x x ) of Eq. (7.56), in accordance
with the general relation (7.155).
The angular parts of the spatial part of the Fourier integral
Z

d3 p ip(xx )
e
(2)3

(7.168)

are the same in both cases, producing an integral over |p|:


1
2R

d|p|
|p| sin |p|R,

(7.169)

where R |x x |. The Feynman propagator has therefore integral representation


1 Z d|p| |p|
0
0
sin |p|R eip |x x | .
G(x x ) =
4R 0
p

(7.170)

In the massless case where p = |p|, we can easily perform the |p|-integration and
recover the previous result (7.151).

489

7.2 Spacetime Behavior of Propagators

To calculate the retarded propagator GR (x x ) = (x0 x0 ) C(x x ) in


spacetime, we may focus our attention upon the commutator function, whose integral representation is now
C(x x ) =

1
4R

dp p
0
0
0
0
sin pR [eip (x x ) eip (x x ) ].
p

(7.171)

In the massless case where p = p, we decompose the trigonometric function into


exponentials, and obtain
1
C(x x ) = i
4R

o
d n i[(x0 x0 )R]
0
0
e
ei[(x x )+R] ,
2

(7.172)

which is equal to
C(x x ) = i

i
1 h 0
(x x0 R) (x0 x0 + R) .
4R

(7.173)

It is instructive to verify in this expression the canonical equal-time commutation

properties (7.57). For this we multiply C(x)


by a test function f (r) and calculate,
with r = |x|,
Z

d3 x C(x)f
(r) = i

dr r 2

i
1 h 0
0 + r) .
(x r) (x
r

(7.174)

0 r) (x
0 + r) = (d/dr)[ (x0 r) + (x0 + r)], we can perform a
Since (x
partial integration and find
i

h
i
0 + r) [rf (r)] = i d [|x0 |f (|x0 |)].
dr (x0 r) + (x
dx0

(7.175)

At x0 = 0, this is equal to if (0) at x0 = 0, as it should.


Using the property of the -function (7.60) in spacetime
(x02 r 2 ) =

1
[(x0 r) + (x0 + r)].
2r

(7.176)

we can write the commutator function also as


1
((x x )2 ).
2

(7.177)

(x0 x0 ) (x0 x0 ) (x0 + x0 ),

(7.178)

C(x x ) = i(x0 x0 )
where (x0 x0 ) is defined by

which changes the sign of the second term in the decomposition (7.176) of ((xx )2 )
to the form required in (7.176).
Note that in the form (7.177), the canonical properties (7.57) of the commutator
function cannot directly be verified since (x
0 x0 ) = 2(x0 x0 ) is too singular

490

7 Quantization of Relativistic Free Fields

to be evaluated at equal times. In fact, products of distributions are in general


undefined, and their use is forbidden in mathematics.8 We must read (x0 )(x2 ) as
the single distribution
(x0 )(x2 )

1
[(x0 r) (x0 + r)]
2r

(7.179)

which it represents to deduce its properties.


With (7.173) and (7.177), the retarded propagator becomes [recall (7.155)]
i
1 h 0
(x x0 R) (x0 x0 + R)
4R
1
= i(x0 x0 ) ((x x )2 ).
(7.180)
2

GR (x x ) = i(x0 x0 )

The Heaviside function allows only positive x0 x0 , so that only the first -function
in (7.180) contributes, and we obtain the well-known expression of classical electrodynamics
GR (x x ) = i(x0 x0 )

1
(x0 x0 R).
4R

(7.181)

This propagator exists only for a causal time order x0 > x0 , for which it is equal to
the Coulomb potential between points which can be connected by a light signal.
The retarded propagator GR (x, x ) describes the massless scalar field (x) caused
by a local spacetime event i (4) (x ). For a general source j(x ), it serves to solve the
inhomogeneous field equation
2 (x) = j(x)
by superposition:
(x) = i

d4 x GR (x, x )j(x ).

(7.182)

(7.183)

Inserting (7.181) and separating the integral into time and space parts, the time x0
can be integrated out and the result becomes
(x, t) =

d3 x

1
j(x , tR ),

4|x x |

(7.184)

where
tR = t |x x |

(7.185)

is the time at which the source has emitted the field which arrives at the spacetime
point x. Relation (7.184) is the basis for the derivation of the Lienard-Wiechert
potential, which is recapitulated in Appendix 7A.
8

An extension of distribution theory to include products is given in the textbook H. Kleinert,


Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets, World
Scientific, Singapore 2009 (klnrt.de/b5).

491

7.2 Spacetime Behavior of Propagators

7.2.4

Comparison of Singular Functions

Let us compare the spacetime behavior (7.180), (7.181) of the retarded propagator
with that of the massless Feynman propagator. The denominator in (7.151) can be
decomposed into partial fractions, and we find a form very close to (7.180):
"

1
1
1
G(x x ) = 2
.
0
0
0
0
8 R |x x | R i |x x | + R i

(7.186)

Feynman has found it useful to denote the function 1/(t i) by i+ (t). This has
a Fourier representation which differs from that of a Dirac -function by containing
only positive frequencies:
+ (t)

d i(ti)
e
.

(7.187)

The integral converges at large frequencies only due to the i -term, yielding
+ (t) =

1 i
.
t i

(7.188)

The pole term can be decomposed as9


i
t
P
1
= 2
+ 2
= i(t) + ,
2
2
t i
t +
t +
t

(7.189)

Recall that the decomposition concerns distributions which make sense only if they
are used inside integrals as multiplyers of smooth functions. The symbol P in the
second term means that the integral has to be calculated with the principal-value
prescription.10 For the function + (t), the decomposition reads
+ (t) = (t)

iP
.
t

(7.190)

An important property of this function is that it satisfies a relation like (t) in


(7.176):
+ (t2 r 2 ) =

1
[+ (t r) + (t + r)].
2r

(7.191)

Below we shall also need the complex conjugate of the function + (t):
(t) =
9

1 i
= [+ (t)] .
t + i

(7.192)

This is often referred to as Sochockis formula. It is the beginning of an expansion in powers


of > 0: 1/(x i) = P/x i(x) + [ (x) idx P/x] + O( 2 ).
10
Because of the entirely different contexts, no confusion is possible with the second-quantized
parity operator P introduced in (7.91).

492

7 Quantization of Relativistic Free Fields

From (7.189) we see that the two functions are related by


+ (t) + (t) = 2(t).

(7.193)

Because of (7.191), the Feynman propagator (7.186) can be rewritten as


G(x x ) =

i
1 h
i
0
0
0
0

(|x

x
|

R)

(|x

x
|
+
R)
= + ((x x )2 ).
+
+
2
8 R
4
(7.194)

These expressions look very similar to those for the retarded propagator in
Eqs. (7.180) and Eqs. (7.181).
It is instructive to see what becomes of the + -function in Feynman propagators
in the presence of a particle mass. According to (7.140), a mass term modifies the
integrand in
Z
d i(xx )2
2
e
,
(7.195)
+ ((x x ) ) =

0
2
2
2
(in which we omit the i term, for brevity) from ei(xx ) to ei(xx ) iM /4 .
We may therefore define a massive version of + ((x x )2 ) by
M
+
((x x )2 ) =

d i(xx )2 iM 2 /2
e

(2)

M H1 (M (x x )2 )
q
=
.
2
M (x x )2
2

(7.196)

A similar generalization of the function ((x x )2 ) in (7.177) to M ((x x )2 )


may be found by evaluating the commutator function C(x x ) in Eq. (7.177) at
a nonzero mass M. Its Fourier representation was given in Eq. (7.171) and may be
written as
i
C(x x ) = 2
2

d|p| |p|2

sin |p|r
sin p0 (x0 x0 )
2
2
|p|r
|p| + M

(7.197)

This can be expressed as a derivative


C(x x ) =

i 1 d
F (r, x0 x0 )
4 r dr

(7.198)

q
dp
p2 + M 2 t
cos
pr
sin
p2 + M 2

(7.199)

of the function
F (r, t) =

The integral yields11

J0 (M t2 r 2 ) for t > r,
0
for t (r, r),
F (r, t) =

J (M t2 r 2 ) for t < r,
0

11

(7.200)

See Appendix in J. Schwinger, Phys. Rev. 75 , 651 (1949); also ibid.76 , 790 (1949); 82 , 664,
914 (1951); 91 , 713, 728 (1953); 92 , 1283 (1953); 93 , 615 (1954); Quantum Electrodynamics,
Dover, New York, 1958.

493

7.2 Spacetime Behavior of Propagators

where J (z) are Bessel functions. By carrying out the differentiation in (7.198) using
J0 (z) = J1 (z), we may write
with

C(x x ) = i(x0 x0 ) M ((x x )2 ),

(7.201)

M 2 J1 (M x2 )

(x ) = (x ) (x )
.
2
M x2

(7.202)

The function (x2 ) enforces the vanishing of the commutator at spacelike distances,
a necessity for the causality of the theory. Using (7.202), we can write the retarded
propagator GR (x x ) = (x0 x0 ) C(x x ) as
GR (x x ) = i(x0 x0 )

1 M
((x x )2 ).
2

(7.203)

In the massless limit, the second term in (7.202) disappears since J1 (z) z for small
z, and (7.203) reduces to (7.180).
Summarizing, we may list the Fourier transforms of the various propagators as
follows:
i
= (p2 M 2 ) = (p02 p2 ),
M 2 + i

=
[ (p0 p ) + + (p0 + p )],
2p

i
=
[ (p0 p ) (p0 + p )],
retarded propagator :
2
2
p+ M
2p

i
=
[+ (p0 p ) + (p0 + p )],
advanced propagator : 2
2
p M
2p
commutator :
2(p0 )(p2 M 2 ) = 2(p0 )(p02 p2 )

=
[ (p0 p ) + (p0 + p )],
p
(7.204)
The corresponding integration contours in the complex energy plane of the representations of the spacetime propagators are indicated in Fig. 7.3.
A little exercise with these functions, is given in Appendix 7B.
We end this section by pointing out an important physical property of the euclidean Feynman propagator. When generalizing its proper-time representation to
D spacetime dimensions, it reads
Feynman propagator :

p2

G(x x ) =

1
4

(xx )2E /4 M 2
.
De

(7.205)

reducing to (7.132) for D = 4. The integrand can be interpreted as the probability12


that a random world line of length L = 2D /a which is stiff over a length scale
12

H. Kleinert, Path Integrals in Quantum Mechanics, Statistics and Polymer Physics,, World
Scientific Publishing Co., Singapore 1995, Second extended edition, Sections 5.84 and 15.7.

494

7 Quantization of Relativistic Free Fields

Figure 7.3 Integration contours in the complex p0 -plane of the Fourier integral for the
various propagators: CF for the Feynman propagator, CR for the retarded and CA for the
advanced propagators.

a = 2D, and has the end-to-end distance x x . Thus the Feynman propagator
gives then the probability of a particle worldline connecting via all possible random
shapes the points x and x . Each line is weighted with a Boltzmann probability
2
eM depending on the various lengths. In recent years this observation has been
very fruitful by giving rise to a new type of quantum field theory, the so-called
disorder field theories. 13 They permit us to study phase transitions of a variety of
different physical systems in a uniform way that is dual to Landaus famous theory.
These transitions have in common that they can be interpreted as a consequence of a
sudden proliferation of line-like excitations, caused by an overwhelming configuration
entropy. The proliferation occurs at a temperature at which the configurational
entropy outweighs the Boltzmann suppression, the latter being proportional to the
length of the line-like excitations.
Examples are polymers in solutions, vortex lines in superfluids, and defect lines
in crystals. Whereas the traditional field theoretic description of phase transitions
due to Landau is based on the introduction of an order parameter and its spacetime
version, an order field, the new description is based on a disorder field describing
random fluctuations of line-like excitations. As a result of this different way of
looking at phase transitions one obtains a field theoretic formulation of the statistical
mechanics of line ensembles, and based on it a simple explanation of the phase
transitions in superfluids and solids.

7.3

Free Dirac Fields

We now turn to the quantization of the Dirac field obeying the field equation (4.497):
(i M)(x) = 0.
13

(7.206)

H. Kleinert, Gauge Fields in Condensed Matter , Vol. I Superflow and Vortex Lines, pp.
1744, World Scientific, Singapore 1989 (http://www.physik.fu-berlin.de/~kleinert/b1).

495

7.3 Free Dirac Fields

Their plane-wave solutions were given in Subsec. 4.14.1, and just as in the case
of a scalar field, we shall introduce creation and annihiation operators for particles
associated with these solutions.
The classical Lagrangian density is [recall (4.498)]:

L(x) = (x)i
(x) M (x)(x).

7.3.1

(7.207)

Field Quantization

The canonical momentum of the (x) field is


(x) =

L(x)
0

= i(x)
= i (x).
[ 0 (x)]

(7.208)

Up to the factor i, this is equal to the complex-conjugate field , as in the nonrelativistic equation (7.4.1). As there, the field has no conjugate momentum,
since
L(x)
= 0.
(7.209)

[ 0 (x)]
This is again a mere artifact of the use of complex field variables and unrelated
to a more severe problem in Section 7.4.1, where the canonical momentum of a
component of the real electromagnetic vector field vanishes as a consequence of
gauge invariance.
If (x, t) were a Bose field, its canonical commutation rule would read
[(x, t), (x , t)] = (3) (x x ).

(7.210)

However, since electrons must obey Fermi statistics to produce the periodic system
of elements, the fields have to satisfy anticommutation rules
n

(x, t), (x , t) = (3) (x x ).

(7.211)

Whereas in the nonrelativistic case, this modifications of the commutation rules


was dictated by the Pauli principle and the implied antisymmetric electronic wave
functions, the relativistic theory can only be quantized by anticommutation rules,
and commutation rules (7.210) are incompatible either with microcausality or the
positivity of the energy. This will be shown in Section 7.8.
We now expand the field (x) into the complete set of classical plane-wave
solutions (4.656) in a large but finite volume V :
(x) =

Xh

p,s3

fp s3 (x)ap,s3 + fpc s3 (x)bp,s3 ,

(7.212)

or explicitly:
(x) =

p,s3

1
q

V p0 /M

eipx u(p, s3 )ap,s3 + eipx v(p, s3 )bp,s3 .

(7.213)

496

7 Quantization of Relativistic Free Fields

As in the scalar equation (7.12) we have associated the expansion coefficients of the
plane waves eipx with a creation operator bp,s3 rather than an annihilation operator.
one dp,s3 . for quite similar reasons [recall the discussion after Eqs. (7.74) and
(7.82)]. The reason a sign reversal of s3 is similar to that for the momentum, and
will be expained below.
Expansion (7.213) is inverted and solved for ap,s3 and bp,s3 by applying the
orthogonality relations (4.659) for the bispinors u(p, s3 ) and v(p, s3 ) to the spatial
Fourier transform

Z
i
X
V h ip0 x0
0 0
3
ipx
q
d xe
(x) =
e
u(p, s3 )ap,s3 + eip x v(p, s3 )bp,s3 .(7.214)
p,s3
p0 /M
The result is
1

ip0 x0

ap,s3 = e

p0 /M

p0 /M

ip0 x0

bp,s3

= e

d3 x eipx (x),

d3 x eipx (x),

u (p, s3 )
v (p, s3 )

(7.215)

this being the analog of the scalar equations (7.6), (7.75). The same result can of
course be obtained from (7.212) with the help of the scalar products (4.658), in
terms of which Eqs. (7.215) are simply
ap,s3 = (fp s3 , )t ,

bp,s3 = (fpc s3 , )t ,

(7.216)

as in the scalar equations (7.19).


From (7.215) we derive that all anticommutators between ap,s3 , ap ,s , bp,s3 , bp ,s
3
3
vanish, except for
n

ap,s3 , ap ,s

bp,s3 , bp ,s
3

MM
u (p, s3 )u(p , s3 )p,p = p,p s3 ,s3 ,
p0 p0

(7.217)

MM
v (p, s3 )v(p, s3 )p,p = p,p s3 ,s3 .
p0 p0

(7.218)

The single-particle states


|p, s3 i = ap,s3 |0i

(7.219)

have the wave functions


1
eipx u(p, s3) = fp,s3 (x)
h0|(x)|p, s3i = q
0
V p /M
1

hp, s3 |(x)|0i
= q
eipx u(p, s3 ) = fp,s3 (x).
0
V p /M

(7.220)

497

7.3 Free Dirac Fields

In addition, there are states


with the matrix elements

|
p, s3 i = bp,s3 |0i,

1
c
h
ps3 |(x)|0i = q
eipx v(p, s3 ) = fp,s
(x),
3
0
V p /M
1
c

eipx u(p, s3 ) = fp,s


(x).
h0|(x)|
p , s3 i = q
3
0
V p /M

(7.221)

(7.222)

As in the scalar field expansion (7.74), the association of the negative-frequncy


solution v(p, s3 )eipx of the Dirac equation with a creation operator bp,s3 rather than
a second annihilation operator dp,s3 ensures that the wave equation of the particle
|
ps3 i has the same exponential form eipx as that of |p, s3 i, both exhibiting the same
0
time dependence ep t with the positive energy p0 = p . The other assignment would
have given a negative energy.
In an infinite volume, a more convenient field expansion makes use of the planewave solutions (4.660) and uses covariant creation and annihilation operators as in
(7.16):
(x) =

i
Xh
d3 p
ipx
ipx

e
u(p,
s
)a
+
e
v(p,
s
)b
3
p,s
3
3
p,s3 .
(2)3 p0 /M s3

(7.223)

The commutation rules between ap,s3 , ap,s3 , bp,s3 , and bp,s3 are the same as in
(7.217), (7.218), but with the replacement of the Kronecker symbols by their invariant continuum version [recall (7.20)]
p0 - (3)
p0
(3)
(p p ) =
(2h)D - (p p ).
(7.224)
M
M
Note that this normalization is different from the bosonic one in (7.20), where the
factor in front of the -function
was 2p0 . This is a standard convention spread
throughout the literature.
In an infinite volume, we may again introduce single-particle states
p,p

|p, s3 ) = ap,s3 |0),

|
p, s3 ) = bp,s3 |0),

(7.225)

with the vacuum |0) defined by ap,s3 |0) = 0 and bp,s3 |0) = 0. These states will be
satisfy the orthogonality relations
p0 - (3)
(p p)s3 ,s3 ,
M

p0 - (3)
(p p)s3 ,s3 ,
M
(7.226)
in accordance with the replacement (7.224) [and in contrast to the normalization
(7.27) for scalar particles]. They have the wave functions
(p , s3 |p, s3 ) =

(0|(x)|p, s3 ) = eipx u(p, s3 ) = fp,s3 (x),


c
(
p, s3 |(x)|0) = eipx v(p, s3 ) = fp,s
(x),
3

(
p , s3 |
p, s3 ) =

(p, s3 |(x)|0)
= eipx u(p, s3 ) = fp,s3 (x).
c

h0|(x)|
p, s3 i = eipx v(p, s3 ) = fp,s
(x).
3
(7.227)

498

7 Quantization of Relativistic Free Fields

We end this section by stating also the explicit form of the quantized field of
massless left-handed neutrinos and their antiparticles, the right-handed antineutrinos:
X
1
1 5

(x) =
eipx uL (p)ap, 1 + eipx vR (p)bp, 1 ,
(x) =
0
2
2
2
2V p
p


(7.228)

where the operators ap, 1 and bp, 1 carry helicity labels 21 , and p0 = |p|. The
2
2
massless helicity bispinors are those of Eqs. (4.720) and (4.721). Remember the
normalization (4.719):
uR (p)uR (p) = 2p0 ,
uL (p)uL (p) = 2p0 ,
vL (p)vL (p) = 2p0 ,
vR (p)vR (p) = 2p0 ,
(7.229)

which is the reason for the factor 1/ 2V p0 in the expansion (7.228) [as in the
expansions (7.12) and (7.74) for the scalar mesons].

7.3.2

Energy of Free Dirac Particles

We now turn to the energy of the free quantized Dirac field. The energy density is
given by the Legendre transform of the Lagrangian density

H(x) = (x)(x)
L(x)

= i (x)(x)
L(x)
i i

= (x)i
(x) + M (x)(x).

(7.230)

Inserting the expansion (7.213) and performing the spatial integral [recall the step
from (7.30) to (7.31)], the double sum over momenta reduces to a single sum, and
we obtain the second-quantized Hamilton operator
H=

Mh
ap,s3 ap,s3 u(p, s3 )( i pi + M)u(p, s3 )
0
p
p,s3 ,s3
+ bp,s3 bp,s v(p, s3 )( i pi + M)v(p, s3 )
X

+ M)v(p, s3 )e2ip

0t

+ bp,s3 ap,s3 v(p, s3 )( i pi + M)u(p, s3 )e2ip

0t

ap,s3 bp,s u(p, s3 )( i pi


3

We now use the Dirac equation, according to which

( i pi + M)u(p, s3 ) = 0 p0 u(p, s3 ),
( i pi + M)v(p, s3 ) = 0 p0 v(p, s3 ),

(7.231)

(7.232)

and the orthogonality relations (4.690)(4.693) and simplify (7.232) to


H=

p,s3

p0 ap,s3 ap,s3 bp,s3 bp,s3 .

(7.233)

499

7.3 Free Dirac Fields

With the help of the anticommutation rule (7.218), this may be rewritten as
H=

p,s3

p0 ap,s3 ap,s3 + bp,s3 bp,s3

p0 .

(7.234)

p,s3

The total energy adds up all single-particle energies. In contrast to the secondquantized scalar field, the vacuum has now a negatively infinite energy due to the
zero-point oscillations, as announced in the discussion of Eq. (7.36).
Note that if we had used here the annihilation operators of negative energy states
dp,s3 instead of the creation operators bp,s3 , then the energy would have read
H=

p,s3

p0 ap,s3 ap,s3 dp,s3 dp,s3

(7.235)

so that the particles created by dp,s3 would again have had negative energies. By
going over from dp,s3 to bp,s3 , we have transformed missing negative energy states
into states of antiparticles, thereby obtaining a positive sign for the energy of an
antiparticle. This happens, however, at the expense of having an negatively infinite zero-point energy of the vacuum. This also explains why the spin orientation
s3 changes sign under the above replacement. A missing particle with spin down
behaves like a particle with spin up.
Actually, the above statements about the exchange dp,s3 bp,s3 are correct only
as far as the sum of all momentum states and spin indices s3 in (7.235) is concerned.
For a single state, we have to consider also the momentum operator and the spin.
The situation is the same as in the free-electron approximation to the electrons
in a metal. At zero temperature, the electrons are in the ground state, forming a
Fermi sea in which all single-particle levels below a Fermi energy EF are occupied.
If the energies are measured from that energy, all occupied levels have a negative
energy, and the ground state has a large negative value. If an electron is kicked
out from one of the negative-energy states, an electron-hole pair is observed, both
particle and hole carrying a positive energy with respect to the undisturbed Fermi
liquid. The hole appears with a positive charge relative to the Fermi sea.
Note also that unlike the Bose case, the exchange dp,s3 bp,s3 maintains the
correct sign of canonical anticommutation rules (7.218). In fact, the negative sign of
the term dp,s3 dp,s3 in the Hamilton operator (7.235) is somewhat less devastating
than in the Bose case. It can actually be avoided by a mere redefinition of the
vacuum as the state in which all momenta are occupied by a dp,s3 particle:
|0inew =

p,s3

dp,s3 |0i.

(7.236)

This state has now the same negatively infinite energy


E0

new

new h0|H|0inew =

p0

(7.237)

p,s3

which was found in the correct quantization. Counting from this ground state energy,
all other states have positive energies, obtained either by adding a particle with the

500

7 Quantization of Relativistic Free Fields

operator ap,s3 or by removing a particle with dp,s3 . Of course, this description is


just a reflection of the fact that for anticommutators, dp,s3 dp,s3 = dp,s3 dp,s3 1,
and a reinterpretation dp,s3 bp,s3 and dp,s3 bp,s3 makes dp,s3 dp,s3 a positive
operator. For commutators, the same trick is impossible since dp,s3 dp,s3 can have
any negative eigenvalue and no new zero level can be given so that all energy
differences are positive!
It was the major discovery of Dirac that the annihilation of a negative-energy
particle may be viewed as a creation of a positive-energy particle with the same
mass and spin, but reversed directions of momentum and spin. Dirac called it the
antiparticle. When dealing with electrons, the antiparticle is a positron. Dirac
imagined all negative energy states in the world as being filled, forming a sea of
negative energy states, just in the above-described Fermi sea in metals. as the
electrons below the Fermi surface in a metal. The annihilation of a negative-energy
particle in the sea would create a hole which would appear to the observer as a
particle of positive energy with the opposite charge.
As in the scalar case it is often possible to simply drop this infinite energy of the
vacuum by introducing a normal product : H : . As before, the double dots mean:
Order all operators such that all creators stand to the left of all annihilators. But in
contrast to the boson case, every transmutation of two Fermi operators to achieve
the normal order is now accompanied by a phase factor 1.

7.3.3

Lorentz Transformation Properties of Particle States

The behavior of the bispinors u(p, s3 ) and v(p, s3 ) under Lorentz transformations
determines the bahavior of the creation and annihilation operators of particles and
antiparticles, and thus of the particle states created by them, in particular the
single-particle states (7.219) and (7.221). Under a Lorentz transformation , the
field operator (x) transforms according to the law (4.518):

(x)
(x) = D()(1x) = ei 2 S (1 x),

(7.238)

On the right-hand side, we now insert the expansion (7.223), it becomes


1

D()( x) =

X h ip x
d3 p
e
u(p , s3 )Ws3 ,s3 (p , , p) ap,s3
(2)3 p0 /M s3 ,s
3

+ eip x v(p , s3 )Ws3 ,s3 (p , , p) bp,s3 .


where we have set p p and used the fact that p1 x = p x. In order to derive
the transformation laws for the creation and annihilation operators, we rewrite this
in the same form as in the original expansion (7.223) in primed quantities.
1

D()( x) =

i
Xh
d 3 p
ip x

ip x

e
u(p
,
s
)a
+
e
v(p
,
s
)b

3 p ,s3
3
p ,s3 ,
(2)3 p0 /M s3 ,s
3

(7.239)

501

7.3 Free Dirac Fields

Because of the Lorentz invariance of the integration measure in momentum space


observed in (4.181), we can replace in (7.239)
Z

d3 p

(2)3 p0 /M

d3 p
,
(2)3 p0 /M

(7.240)

and compare coefficients to find the transformation laws


a

p ,s3

bp ,s
3

ap ,s3

b p ,s
3

1/2
X

Ws3 ,s3 (p , , p) ap,s3 ,

s3 =1/2
1/2
X

s3 =1/2

Ws3 ,s3 (p , , p) bp,s3 .

(7.241)

In contrast to the Lorentz transformations (7.238) of the field, the creation and
annihilation operators are transformed unitarily under the Lorentz group. The
right-hand sides define a unitary representation of the Lorentz tranformations :
ap ,s3 = U 1 ()ap ,s3 U() =
b p ,s
3

= U

()bp ,s U()
3

1/2
X

Ws3 ,s3 (p , , p) ap,s3 ,

1/2
X

Ws3 ,s3 (p , , p) bp,s3 ,

s3 =1/2

s3 =1/2

(7.242)

with similar relations for ap,s3 and bp,s3 . For the single-particle states (7.225), this
implies the transformation laws:
U()|p, s3 i =

U()ap,s3 U 1 ()|0i

U()|
p, s3 i = U()bp,s3 U 1 ()|0i =

1/2
X

|p , s3 iWs3 ,s3 (p , , p),

1/2
X

|
p , s3 iWs3 ,s3 (p , , p),

s3 =1/2

s3 =1/2

(7.243)

where we have used the Lorentz-invariance of the vacuum state


U()|0i = |0i.

(7.244)

We are now ready to understand the reason for introducing the matrix c in the
bispinors v(p, s3 ) of Eq. (4.678), and thus of the sign reversal of the spin orientation.
In (4.736) we found that the canonical spin indices of v(p, s3 ) transformed under
rotations by the complex-conjugate 2 2 Wigner matrices with respect to those of
u(p, s3 ). This implies that the Wigner rotations mixing the spin components of the
operators ap,s3 and bp,s3 are complex conjugate to each other. As a consequence,
bp,s3 tranforms in the same way as ap,s3 , and antiparticles behave in precisely the
same way under Lorentz tansformations as particles.

502

7 Quantization of Relativistic Free Fields

For massless particles in the helicity representation, the creation and annihilation
operators transform under Lorentz transformations merely by a phase factor, as
discussed at the end of Section 4.16.3.
Under translations by a four-vector a , the particle and antiparticle states
|p, s3 ) = ap,s3 |0) and |
p, s3 ) = bp,s3 |0) transform rather trivially. Since they have a
definite momentum, the receive merely a phase factor eipa . This follows directly by
implementing the transformation law (4.521) on the Dirac field operator:
(x)
(x) = (x a).

(7.245)

Inserting into the right-hand side the operator expansion (7.223), we see that the
creation and annihilation operators transform as follows:
ap,s3
ap,s3 = eipa ap,s3 ,

bp,s3
bp,s3 = eipa bp,s3 .

(7.246)

These transformation laws define a unitary operator of translations U(a):


ap,s3 = U 1 (a)ap,s3 U(a) = eipa ap,s3 ,

bp,s3 = U 1 (a)bp,s3 U(a) = eipa bp,s3 .

(7.247)

On the single-particle states, the operator U(a) has the effect:


U(a)|p, s3 i = U(a)ap,s3 U 1 (a)|0i = |p, s3 ieipa ,

U(a)|
p, s3 i = U(a)bp,s3 U 1 ()|0i = |
p, s3 ieipa .

(7.248)

One can immediately write down an explicit expression for this operator:
U(a) = eiaP ,

(7.249)

where P 0 is the Hamilton operator (7.234), rewritten in the infinite-volume form as


0

P H=

d3 p
(2)3 p0 /M

ap,s3

p ap,s3 +

bp,s3

p0 - (3)
p bp,s3 p
(0) .
M
0

(7.250)

The last term is a formal infinite-volume expression for the finite-vlolume vacuum
energy
0
Pvac

d3 pV 0
p,
(2)3

(7.251)

R
(3)
as we see from the Fourier respresentation of - (p) = d3 x eipx which is equal to
V for p = 0. The operator of total momentum reads



d3 p

a
p
a
+
b
p
b
p,s3
p,s3 .
p,s3
(2)3 p0 /M p,s3

(7.252)

503

7.3 Free Dirac Fields

Together, they form the operator of total four-momentum P = (P 0 , P) which


satisfies the commutation rules with the particle creation and annihilation operators
[P , ap,s3 ] = p ap,s3 ,

[P , bp,s3 ] = p bp,s3 .

(7.253)

These express the fact that by adding a particle of energy and momentum p to the
system, the total energy and momentum P is increased by p .
By combining Lorentz transformations and translations as in (4.523), we cover
the entire Poincare group, and the single-particle states form an irreducible representation
space of this group. The invariants of the representation are the mass
M = p2 and the spin s. Thus the complete specification of a representation state
is
|p, s3 [M, s]i.
(7.254)
The Hilbert space of two-particle states |p, s3 [M, s]; p , s3 [M , s ]i gives rise to a
reducible representation of the Poincare group. The reducibility is obvious from the
fact that if both particles are at rest, the state |0, s3 [M, s]; 0 , s3 [M , s ]i is simply a
product of two rotational states of spins s and s , and decomposes into irreducible
representations of the rotation group with spins S = |s s |, |s s | + 1, . . . , s + s .
If the two particles carry momenta, the combined state will also have all possible
orbital angular momenta. In the center-of-mass frame of the two particles, the
momenta are of equal size and point in opposite directions. The states are then
, and may be written as
characterized by the direction of one of the momenta, say p

|
p, s3 , s3 }. The satisfy the completeness relation
s
X

Z
s
X

s3 =s s3 =s 0

2
0

d |
p, s3 , s3 }{
p, s3 , s3 | = 1,

(7.255)

.
where and are the spherical angles of the direction p
In the absence of spin, it is then simple to find the irreducible contents of the
rotation group. We merely expand the state |
p} in partial waves:
|
p} =

X
l,m

|l, m}{l, m|
p}

X
l,m

|l, m}Ylm (, ),

(7.256)

with spherical harmonics Yl,m (, ). The states |j, m} are orthonormal and complete
in this Hilbert space:
{j, m|j , m } = j,j m,m ,

X
j,m

|j, m}{j, m| = 1.

(7.257)

In the presence of spins s, s , the decomposition is simplest if the spin orientations


are specified in the helicity basis. Then the two-particle states possess an azimuthal
, and may be written as
angular momentum of size h h around the direction p
|
p, hh }. These have the same rotation properties the wave functions of a spinning

504

7 Quantization of Relativistic Free Fields

top with an azimuthal angular momentum h h around the body axis. The latter
functions are well known they are just the representation functions

j
i(m+m ) j
dm m ()
Dm
m (, , ) = e

(7.258)

of rotations of angular momentum j introduced in Eq. (4.837). These serve as


wave functions of a spinning top with angular momentum j and magnetic quantum
number m, and with an azimuthal component m of angular momentum around the
body axis. They are the basis states of irreducible representations of the rotation
group in the center-of-mass frame. Thus, extending (7.256), we can expand
|
p, h h } =

j,m

j,m

|j, m, h h }{j, m, h h |
p}

|j, m, h h }

2j + 1 j
Dm hh (, , 0),
4

(7.259)

with an orthonormal and complete set of states |j, m, h h } at a fixed h h .


The normalization factor is determined to comply with the orthonormality property
(4.857) of the rotation functions.
Now we boost this expansion from the center-of-mass frame back to the initial
frame in which the total momentum is P = p + p , the energy P 0 = p0 + p0 , and
the mass = P 2 . The direction of P may be chosen as a quantization axis for the
angular momentum j. Then the quantum number m is equal to the helicity of the
combined state. The angular momentum j in the rest frame determines the spin of
the combined state. The resulting Clebsch-Gordan-like expansion of the two-particle
state into irreducible representations of the Poincare group is [5]

|p, s3 [M, s] ; p

, s3 [M, s ] i

X
j,m

(M +M )2

d
2

d3 P
|P, m [, j] i
(2)3 P 0 /

hP, m [, j] |p, s3 [M, s] ; p , s3 [M , s ] i,

(7.260)

with the expansion coefficients


hP, m [, j] |p, h [M, s] ; p, h [M , s ] i =

(4)
- (P p p ) N (; M, M )
2j + 1 j
Dm hh (, , 0),

(7.261)

where N (; M, M ) is some normalization factor. The product representation is


not simply reducible and requires distinguishing the different irreducible biparticle
states of spin S. For this purpose, a degeneracy label has been introduced. It may
be taken as the pair of helicity indices (h.h ) of the individual particles which the
biparticle is composed of. However, to describe different processes most efficiently,
other linear combinations may be more convenient. One may, for example, combine
first the individual spins to a total intermediate spin S, then this with the orbital

505

7.3 Free Dirac Fields

angular momentum L in the center-of-mass frame to the final total angular momentum that frame j (the spin of the biparticle). This corresponds to the LS coupling
scheme in atomic physics. Explicitly, these states are
|P, m [, j] (LS)i = hj, hh |L, 0; S, hh ihS, hh |s, h; s, h i|P, m [, j] (h, h)i.
(7.262)
A sum over all h, h is implied. Since the quantization axis is the direction of p in
the center-of-mass frame, the orbital angular momentum has no L3 -component.

Figure 7.4 Different coupling schemes for two-particle states of total angular momentum
j and helicity m. The first is the LS, the second the JL scheme.

Another possibility of coupling the two particles corresponds to the multipole


radiation in electromagnetism. Here one of the particles, say [M, s], is singled out
to carry off radiation, which is analyzed according to its total angular momentum J
composed of L and s. This total angular momentum is coupled with the other spin
s to the combined total angular momentum j of the biparticle. Here the states are
|P, m [, j] (JL)i = hJ, h|s, h; L, 0ihj, hh|J, h; s , h i|P, m [, j] (h, h)i. (7.263)
As in (7.262), a sum over all h, h is implied. The (LS) and (JL) states are related
to each other by Racahs recoupling coefficients.14 .
The integral over can be performed after decomposing the -function
ensuring
the conservation of energy and momentum as


(4)
(3)
- (P p p ) = - (P p p ) -

(p +

p ) 2

P0
.

(7.264)

A suitable normalization of the irrducible states is


(3)
- )j,j m,m , .
hP, m [, j] |P, m [ , j ] ) = - (P P i (

(7.265)

For more details on this subject see Notes and References.


14

For the general recoupling theory of angular momenta see Chapter VI of the textbook by
Edmonds, quoted in Notes and References of our Chapter 4

506

7 Quantization of Relativistic Free Fields

Propagator of Free Dirac Particles


Let us now use the quantized field to calculate the propagator of the free Dirac field.
Thus we form the vacuum expectation value of the time ordered product
)|0i.
S(x, x ) = h0|T (x)(x

(7.266)

For, if we use the explicit decomposition


) = (x0 x )(x)(x
) (x x0 )(x
)(x)
T (x)(x
0
0

(7.267)

and apply the Dirac equation (4.655) we see that


)
(i M)T (x)(x

) .
= T (i M)(x)(x ) + i(x0 x0 ) 0 (x), (x
n

(7.268)

The right-hand side reduces indeed to i (4) (xx ), due to the Dirac equation (7.206)
and the canonical commutation relation (7.211).
Let us now insert the free-field expansion (7.213) into (7.266) and calculate, in
analogy to (7.46),
S (x, x ) = S (x x )
1 X
M

q
= (x0 x0 )
ei(pxp x ) u (p, s3 )
u (p , s3 )h0|ap,s3 ap ,s |0i
3
V p,s3;p ,s3 p0 p0
(x0 x0 )

1
V

p,s3 ;p ,s3

p0 p0

ei(pxp x ) v (p, s3 )
v (p , s3 )h0|bp ,s3 bp,s |0i
3

1 X M ip(xx ) X
e
u (p, s3 )
u (p, s3 )
V p,s3 p0
s3
1 X M ip(xx) X
(x0 x0 )
e
v (p, s3 )
v (p, s3 )
V p,s3 p0
s3

= (x0 x0 )

(7.269)

We now recall the polarization sums (4.696) and (4.697), and obtain from (7.269),
1 X 1 ip(xx)
e
(/
p + M)
2V p p0
1 X 1 ip(xx )
+ (x0 x0 )
e
(/
p + M).
2V p p0

S (x x ) = (x0 x0 )

(7.270)

Let us mention here that this type of decomposition can be found for particles of
any spin. In (4.699) we introduced the polarization sum
P (p) =

X
s3

u(p, s3 )
u(p, s3 ),

P (p) =

X
s3

v(p, s3 )
v(p, s3 ) = P (p).

(7.271)

507

7.3 Free Dirac Fields

In terms of these, the propagator has the general form


1 X 1 ip(xx)
e
P (p)
2V p p0
1 X 1 ip(xx)
e
P (p).
+ (x0 x0 )
2V p p0

S (x x ) = (x0 x0 )

(7.272)

This structure is found for particles of any spin, in particular also for integer-valued
ones. The alternating sign in relation (4.700) between the polarization sums of
particles and antiparticles is cancelled by the alternating sign in the definition of
the time-ordered product for bosons and fermions, if the statistics is chosen properly.
This will be discussed in more detail in Section 7.8.
In an infinite volume, the momentum sums are replaced by integrals and yield
the invariant functions G(+) (x x ), G() (x x ) [recall (7.47) and (7.54)]. The sum
containing the term p/ may be summed in this way after taking it outside in form
of a spacetime derivative, i.e., by writing
1 X 1 ip(xx)
e
(/
p + M) = (i/ + M)G() (x x ).
2V p p0

(7.273)

Note that the zeroth component inside the sum comes from the polarization sum
over spinors and thus lies on the mass shell , and that also the Green functions
G(+) (x x ) and G() (x x ) contain only wave functions with on-shell energies.
We therefore find the propagator
S (x x ) = (x0 x0 )(i/ + M)G(+) (x x ) + (x0 x0 )(i/ + M)G() (x x ).
(7.274)
This expression can be simplified further by moving the derivatives to the left of the
Heaviside function. This gives
h

S (x x ) = (i/ + M) (x0 x0 )G(+) (x x ) + (x0 x0 )G() (x x )


h

i 0 (x0 x0 ) G(+) (x x ) G() (x x )

The second term happens to vanish because of the property

G(+) (x, 0) = G() (x, 0).

(7.275)

(7.276)

The final result is therefore the simple expression


S (x x ) = (i/ + M)G(x x ),

(7.277)

i.e., the propagator of the Dirac field reduces to that of the scalar field multiplied
by the differential operator i/
+ M. Inserting on the right-hand side the Fourier
representation (7.65) of the scalar propagator, we find for Dirac particles the representation
S (x x ) =

i
d4 p
ip(xx )
(/
p
+
M)
e
.
(2)4
p2 M 2 + i

(7.278)

508

7 Quantization of Relativistic Free Fields

This is to be contrasted with the Fourier representation of the intermediate expression (7.274). If the Heaviside functions are expressed as in Eq. (7.61), we see that
(7.274) can be written as

S (x x ) =

d4 p i
(2)4 2p

p
p

eip(xx ) . (7.279)

0
0
p p + i p + p i

The difference between (7.279) and (7.278) is an integral

dp0 ip0 (x0 x0 ) i


e
2
2p

p0 p
p0 + p
,

p0 p + i p0 + p i
!

(7.280)

But the two distributions in the integrands are of the form x/(x i), and thus
equal to unity, so that (7.280) vanishes.
The integrand in (7.278) can be rewritten in a more compact way using the
product formula
(/
p M)(/
p + M) = p2 M 2 .
(7.281)
This leads to the Fourier representation of the Dirac propagator

S (x x ) =

d4 p
i

eip(xx )
4
(2) p/ M + i

(7.282)

0
It is worth noting that while
2in Eq.2 (7.270) the particle energy k in p/ lies on the
mass shell, being equal to p + M , those in the integral (7.278) lie off-shell. They
are integrated over the entire k 0 -axis and have no relation to the spatial momenta
p.
As in the case of the scalar fields [see Eqs. (7.354)], the free-particle propagator
is equal to the Green function of the free-field equation. Indeed, by writing (4.655)
as
L(i)(x) = 0
(7.283)

with the differential operator


L(i) = i/
M,

(7.284)

we see that the propagator is the Fourier transform of the inverse of L(p):

S(x x ) =

d4 p i ip(xx )
e
.
(2)4 L(p)

(7.285)

It obviously satisfies the inhomogeneous Dirac equation


(i M)S(x, x ) = i (4) (x x ).

(7.286)

For completeness, we also write down the commutator function of Dirac fields.
From the expansion (7.213) and the canonical commutation rules (7.217) and (7.218)
we find directly
)] C (x x )
[(x), (x

(7.287)

509

7.3 Free Dirac Fields

with the commutator function


1 X 1 ip(xx )
1 X 1 ip(xx)
e
(/
p
+
M)

e
(/
p + M)
2V p p0
2V p p0

C (x x ) =

= (i/
+ M)C(x x ).

7.3.4

(7.288)

Behavior under Discrete Symmetries

Let us conclude this section by studying the behavior of spin


discrete symmetries P, C, T .

1
2

particles under the

Space Inversion
The space reflection
P

(x)
(x) = P

0 1
1 0

(
x) = P 0 (
x)

(7.289)

is achieved by defining the unitary parity operator P on the creation and annihilation
operators as follows:

Pap,s3 P 1 a (p, s3 ) = P a (p, s3 ),

Pbp,s3 P 1 b (p, s3 ) = P b (p, s3 ).

(7.290)

The opposite sign in front of bp,s3 is necessary since the spinors behave under parity
as follows
0 u(p, s3 ) = u(p, s3 ),
0 v(p, s3 ) = v(p, s3 ).

(7.291)

This follows directly from the explicit representation (4.668) and (4.678):
0 1
1 0
0 1
1 0

1 q p
1 q p
M
M
(s3 ),

(s3 ) =
p
p

2
2
M
M

p
p

1
1
q M c (s3 ).
q M c (s3 ) =

2
2
p
p
M
M

(7.292)

It is easy to verify that with (7.290) and (7.291), the second-quantized field (x)
with the expansion (7.213) transforms as it should:
P(x)P 1 = P (x) = P 0 (
x).

(7.293)

The opposite phase factors of a (p, s3 ) and b (p, s3 ) under space inversion implies
that in contrast to two identical scalar particles, the bound state of a Dirac particle

510

7 Quantization of Relativistic Free Fields

and its antiparticle in a relative orbital angular momentum l carries an extra minus
sign, i.e., has a parity
P = ()l+1 .
(7.294)
Therefore, the ground state of a positronium which is in an electron and a positron
an s-wave bound state represents a pseudoscalar composite particle. The spins are
coupled to zero.
For two different Dirac particles, the combined parity (7.294) carries, of course, an
extra factor P1 P2 .
Charge Conjugation
Charge conjugation transforms particles into antiparticles. We therefore define this
operation on the creation and annihilation operators of the Dirac particles by

Cap,s3 C 1 = a p,s3 = C bp,s3 ,

Cbp,s3 C 1 = b p,s3 = C ap,s3 .

(7.295)

To find out how this operation changes the Dirac field operator we observe that the
matrix C introduces in Eq. (4.597)
0 2

C = i =

c
0
0 c

c = ei2 /2 = i 2 ,

(7.296)

has the property of changing uT (p, s3 ) into v(p, s3 ) and vT (p, s3 ) into u(p, s3 ):
u(p, s3) = C vT (p, s3 ),

v(p, s3 ) = C uT (p, s3 ).

(7.297)

The second property was proved in (4.674). The fist is proved in precicely the same
way.
It is then obvious that the transformation laws (7.295) have the following effect
upon the second-quantized field (x):
C(x)C 1 = C (x) = C C T (x).

(7.298)

Let us now check the transformation property of the second-quantized Dirac


current j (x) under charge conjugtion. In the first-quantized form we have found
in Eq. (4.611) that the current ramains invariant. After field quantization, however,
there is a minus sign arising from the need to interchange the order of the Dirac
field operators when bringing the transformed current T (x) T T (x) in Eq. (4.612)

back to the original operator ordering (x)


(x) in the current j (x). Thus we
obtain the second-quantized tranformation law
C

j (x)
j (x) = j (x).
rather than (4.611). This is the law listed in Table 4.1.

(7.299)

511

7.3 Free Dirac Fields

A bound state of a particle and its antiparticle such as the positronium in a


relative angular momentum l and with the spins coupled to S has the charge parity
C = (1)l+S .

(7.300)

In order to see this we form the state


SS3
|lm
i

dp Rl (p)

Ylm (
d2 p
p)a p,s3 bp,s |0ihS, S3 |s, s3; s, s3 i,
3

(7.301)

where hS, S3 |s, s3; s, s3 i are the Clebsch-Gordan coefficients coupling the two spins
to a total spin S with a third component S3 [recall Table 4.2].
Under charge conjugation, the product ap,s3 bp,s goes over into bp,s3 ap,s . In
3
order to bring this back to the original state we have to interchange the order of the
two operators and the spin indices s3 and s3 , and invert p into p. The first gives
a minus sign, the second a ()l sign, and the third a ()(S2s3 ) sign, since

hS, S3 |s, s3 ; s , s3 i = ()Sss hS, S3 |s , s3 ; s, s3i.

(7.302)

Altogether, this gives ()l+S .


Time Reversal
Under time reversal, the Dirac equation is invariant under the transformation
(4.614):
T

(x)
T (x) = D(T ) (xT ),

(7.303)

D(T ) = T C5 .

(7.304)

with the 4 4 -representation matrix D(T ) of Eq. (4.623):


The second-quantized version of this tranformation of the Dirac field operators reads
T

(x)
T

(x)T = T (x) = D(T )(


x),

(7.305)

The antilinearity of T generates the complex conjugation of the classical Dirac


field in (7.303). Indeed, inserting the expansion (7.213) according to creation and
annihilation operators into T 1 (x)T , the antilinearity leads to complex-conjugate
spinor wave functions:
T

(x)T

p,s3

1
q

V p0 /M

eipx u (p, s3 )T

ap,s3 T + eipx v (p, s3 )T

= D(T )(
x).

1
bp,s3 T

(7.306)

Expanding likewise the field operator (


x) on the right-hand side, and comparing
the Fourier coefficients, we find the equations
u (p, s3 )T

v (p, s3 )T

1
bp,s3 T

ap,s3 T

= D(T )u(p, s3)ap,s3 ,


= D(T )v(p, s3)bp,s3 .

(7.307)

512

7 Quantization of Relativistic Free Fields

To calculate the expressions on the right-hand sides we go to the chiral representations (4.668) where
!
c 0
D(T ) = T C5 =
,
(7.308)
0 c
and see that
q

p
1
1 cq p
M
M

C5 q p
(s
)
=
(s3 ),
3

2
2
c
M
M

(7.309)

p
c p
1
1
q M c (s3 ) =
q M c (s3 ).
C5
p

2
2
M
c p
M

We now proceed as in the treatment of Eq. (4.675), using relation (4.677) to find
q

1 q p
1
M
M

(s3 )(1)ss3 ,
(s
)
=
C5 q p
3
p

2
2
M
M

(7.310)

p
p

1
1
q M (s3 ) =
q M (s3 )(1)ss3 ,
C5

2
2
p
p
M
M

and therefore

C5 u(p, s3 ) = u (p, s3 )(1)ss3 ,


C5 v(p, s3 ) = v (p, s3 )(1)ss3 ,

(7.311)

Comparing now the two sides of Eqs. (7.306) we obtain the transformation laws for
the creation and annihilation operators
T ap,s3 T
T

bp,s3 T 1

a p,s3 = T ap,s cs3 s3 = T (1)ss3 ap,s3 ,


3

b p,s3

= T bp,s cs3 s3
3

= T (1)ss3 bp,s3 .

(7.312)

The occurrence of the c-matrix accounts for the fact that the time-reversed state
has the opposite internal rotational motion, with the phases adjusted so that the
rotation properties of the transformed state remain correct.
The operator T is antilinear and antiunitary. This has the effect that just as
the wave functions of the Schrodinger theory, the Dirac wave function of a particle
of momentum p
fp (x) = u(p, s3)eipx
(7.313)
receives a complex conjugation under time reversal:
T

fp,s3 (x)
fp,s3 T (x) = D(T )fp,s
(xT ) = D(T )u(p, s3 )eipx
3

(7.314)

The time-reversed wave function on the right-hand side satisfies again the Dirac
equation
(i/
M)fp,s3 T (x) = 0.
(7.315)

513

7.4 Free Photon Field

In momentum space this reads


(/
p M)D(T )u (p, s3 ) = 0,

(7.316)

and this follows directly from (4.619) and a complex conjugation. A similar consideration holds for a wave function
c
fp,s
(x) = v(p, s3 )eipx ,
3

(7.317)

(/
p + M)D(T )v (p, s3 ) = 0,

(7.318)

where
As discussed previously on p. 314, parity is maximally violated. On may therefore wonder why the neutron, which may be described by a Dirac spinor as far as
its Lorentz transformation properties are concerned, has an extremely small electric
dipole moment, for which one presently knows only an upper bound:
del < 11 1026 e cm.

(7.319)

It was Landau who first pointed out that this can be understood as being a consequence of the extremely small violation of time reversal invariance. The electric
dipole moment is a vector operator d = e x where x is the distance between the
positive and negative centers of charge in the particle. This operator is obviously
invariant under time reversal. Now, for a neutron at rest, d must be proportional
to the only other vector operator available, which is the spin operator , i.e.,
d = const. .

(7.320)

But under time reversal, the right-hand side changes its sign, whereas the left-hand
side does not. The constant must therefore be zero.15
Time reversal invariance ensures that a two-particle scattering amplitude for the
process
1+23+4
(7.321)

is the same as for the reversed process

3 + 4 1 + 2.

(7.322)

This will be discussed later in Section 9.7, after having developed a scattering theory.
Also other consequences of time reversal invariance can be found there.

7.4

Free Photon Field

Let us now discuss the quantization of the electromagnetic field. The classical
Lagrangian density is, according to (4.234),
1
1
L(x) = [E2 (x) B2 (x)] = F 2 (x),
2
4
15

This argument is due to L.D. Landau, Nucl. Phys. 3, 127 (1957).

(7.323)

514

7 Quantization of Relativistic Free Fields

where E(x) and B(x) are electric and magnetic fields, and the Euler-Lagrange equations are
L
L
=
,
(7.324)

[ A ]
A
i.e.,
(g 2 )A (x) = 0.

7.4.1

(7.325)

Field Quantization

The spatial components Ai (x) (i = 1, 2, 3) have the canonical momenta


i (x) =

L(x)
= F0i (x) = E i (x),
i

A (x)

(7.326)

which are just the electric field components. The zeroth component A0 (x), however,
poses a problem. Its canonical momentum vanishes identically
0 (x) =

L(x)
= 0,
A 0 (x)

(7.327)

since the action does not depend on time derivative of A0 (x). This property of
the canonical momentum is a so-called primary constraint in Diracs classification
scheme of Hamiltonian systems.16 As a consequence, the Euler-Lagrange equation
for A0 (x)
L
L
=
,
(7.328)
i
0
[i A ]
A0
is merely an equation of constraint
E(x, t) = 0.

(7.329)

which is Coulombs law for free fields, the first of the field equations (4.244). This
is a so-called secondary constraint in Diracs nomenclature. In the presence of a
charge density (x, t), the right hand side is qual to (x, t).
We have encountered a vanishing field momentum before in the nonhermitian
formulation of the complex nonrelativistic and scalar field theries. There, however,
this was an artifact of the complex formulation of the canonical formalism [recall
the remarks after Eqs. (7.4.1) and (7.209)], and the field was fully dynamical.
Let us calculate the Hamiltonian density:
H=
16

3
X

1
i A i L = (E2 + B2 ) + E A0 .
2
i=1

(7.330)

P.A.M. Dirac, The Principles of Quantum Mechanics, Oxford University Press, Oxford, Third
Edition, 1947, p.114.

515

7.4 Free Photon Field

In the Hamiltonian H = d3 x H, the


last term can be integrated by parts, and we
R 3
find after neglecting a surface term d x (E A0)(x, t), and enforcing Coulombs
law (7.329):
Z
1
(7.331)
H = d3 x (E2 + B2 ).
2
We now attempt to convert the canonical fields Ai (x, t) and E i (x, t) into field
operators by enforcing canonical commutation rules. There is, however, an immediate obstacle to this caused by Coulombs law. Proceeding straight-forwardly, we
would impose the equal-time commutators
R

Ai (x, t), Aj (x , t)

= 0,

[i (x, t), j (x , t)] = 0,

i (x, t), Aj (x , t)

(7.332)
h

= E i (x, t), Aj (x , t) = i ij (3) (x x ).

But the third equation is incompatible with Coulombs law (7.329). The reason for
this is clear: We have seen above that the zeroth component of the vector potential
A0 (x, t) does not become an operator. But then Coulombs law, written out as
2 A0 (x, t) = 0 A(x, t).

(7.333)

implies that A(x) cannot be an operator either, and must commute with the
canonical field operators Ai (x, t) and E i (x, t). The commutation with Ai (x, t) follows directly from the first relation (7.332).
In order to enforce also the comi
mutation with E (x, t), we must correct the canonical commutation rules between
E i (x, t) and Aj (x , t). Thus we replace the third commutator in (7.332) by
h

i (x, t), Aj (x , t) = E i (x, t), Aj (x , t) = iTij (x x ).

where Tij (x x ) is the transverse -function


Tij (x

x)

d3 k ik(xx ) ij k i k j
2
e
(2)3
k

(7.334)

(7.335)

By construction, leads to a vanishing commutation rule


h

E i (x, t), A(x , t) = 0,

(7.336)

which ensures that A(x , t) is a c-number field, just as A0 (x , t), thus complying
with Coulombs law (7.333).
In the second-quantized theory, the fields A0 (x, t) and A(x, t) play a rather
inert role. Both are determined by an equation of motion. The reason for this lies
in the gauge invariance of the action, which makes the theory independent of the
choice of A(x, t), or A0 (x, t). In Section 4.6.2 we have learned that we are free
to choose the Coulomb gauge in which A(x, t) vanishes identically, so that by
(7.333) also A0 (x, t) 0.

516

7 Quantization of Relativistic Free Fields

In the Coulomb gauge, the canonical field momenta of the vector potential (7.326)
are simply their time derivatives:
i (x, t) = E i (x, t) = A i (x, t),

(7.337)

just as for scalar fields [see (7.1)]


We are now ready to expand each component of the vector potential into plane
waves just as previously the scalar field
A (x) =

X
k,

h
i
1
eikx (k, )ak, + (k, )ak, ,
2V k 0

(7.338)

where (k, ) are polarization vectors (4.309), (4.321) labeled by = 1:


(k, ) (0, (k, )),

(7.339)

with the transverse spatial polarization vectors

cos cos i sin


1
 (k, 1) = cos sin i sin
.
2
sin

(7.340)

The four-dimensional polarization vectors (4.309) satisfy the othonormality condition (4.326) and having the polarization sums (4.328) and (4.332) .
Let us quantize the vector field (7.338). In order to impose the canonical commutation rules upon the creation and annihilation operators, we invert the field
expansion (7.338) as in the scalar equation (7.14), and find:17
(

ak,
ak,

ik 0 x0

=e

2V k 0

 (k, )
 (k, )

)Z

t) + k 0 A(x, t) .
d3 x eikx iA(x,
(7.341)

Using now the commutators (7.334), we find


[ak, , ak , ] = 0,
and
[ak, , ak , ]

k 0 k 0
2V
Z

[ak, , ak , ] = 0,

(7.342)

1
1
i(k 0 k 0 )x0
+
e
k 0 k 0


d3 x d3 x ei(kxk x ) i (k, )j (k , ) Tij (x x ). (7.343)

Inserting the finite-volume version of the Fourier representation (7.335) of the transverse -function,
Tij (x
17

1 X ip(xx ) ij pi pj
2 .
x)
e
V p
p

(7.344)

The reader is invites to express this in terms of a scalar product in analogy with (7.13).

517

7.4 Free Photon Field

the spatial integrations can be done with the result




i (k, )j (k , )k,k ij k i k j /k2 .


Due to the transversality condition (4.305), this reduces to
i (k, )j (k , ) ij k,k .
Using the ortonormality relation (4.323), we therefore obtain
[ak, , ak , ] = k,k .

(7.345)

Thus we have found the usual commutation rules for the creation and annihilation
operators of the two transversely polarized photons existing for each momentum k.
Energy of Free Photons
We can easily express the field energy (7.331) in terms of the field operators. In

the Coulomb gauge with A(x, t) = 0, and E(x, t) = A(x,


t), the field energy
simplifies to
Z

1
(7.346)
0 Ai 0 Ai + Ai Ai
H = d3 x
2
This is a sum over the field energies of the three components Ai (x, t), each of them
being of the same form as Eq. (7.29) for a scalar field of zero mass. The subsequent
calculation is therefore the same as there. Inserting the field expansion (7.338), we
find the Hamilton operator
H=

k 0 ak, ak, +

k,=1

1
2

(7.347)

which contains the vacuum energy


E0 h0|H|0i =

1 X 0 X
k =
k ,
2 k,=1
k

(7.348)

due to the zero-point oscillations, and counts the energy k 0 of all transverse photons.
Propagator of Free Photons
Let us now calculate the Feynman propagator of the photon field, defined by the
vacuum expectation value
G (x, x ) h0|T A (x)A (x )|0i.

(7.349)

Inserting the field expansion (7.338), we calculate the vacuum expectation separately
for t > t and t < t as in (7.46), and find
1 X 1 ik(xx ) X
e
(k, ) (k, )
2V k k

X
X 1
1

+ (x0 x0 )
eik(xx )
(k, ) (k, ).
2V k k

G (x, x ) = (x0 x0 )

(7.350)

518

7 Quantization of Relativistic Free Fields

The Heaviside functions are represented as in Eqs. (7.62). The polarization sums in
the two terms are real and therefore the same, both being given by the projection

matrices Pphys
(k) or Pphys
(k) of Eqs. (4.328) or (4.332), respectively. Since these are
independent of k 0 , we may proceed as in going from (7.46) to (7.63), leading to the
propagator for in an infinite spatial volume:

G (x x ) =

d4 k ik(xx) i

e
Pphys
(k).
4
2
(2)
k + i

(7.351)

The independence of Pphys


(k) on k 0 permits us assume k 0 in the -dependent expression (4.332) to be off mass shell, with k following the four-dimensional integration in
(7.351). Alternatively, it may be assumed to be equal to k-dependent the mass-shell

value k , i.e., we may use the polarization sum Pphys


(k , k) rather than (4.332). The
latter is, however, an unconventional option, since the off-shell form is more convenient for assessing the consequences of gauge invariance in the presence of charged
particles [see the discussion after (12.94)].
For all previous fields, the free-photon propagator was also a Green function of
the field equations. Neglecting surface terms, we rewrite the action associated with
the Lagrangian density (7.323) after a partial integration as

A=

1
2

d4 x A (x)(g 2 )A (x),

(7.352)

corresponding to a Lagrangian density


1
L(x) = A (x)L (i)A (x),
2

(7.353)

with the differential operator


L (i) g 2 .

(7.354)

The Euler-Lagrange equation (7.325) can then be written simply as


L (i)A (x) = 0.

(7.355)

Ordinarily we would define a Green function by the inhomogenous differential equation


L (i)G (x x ) = i (4) (x x ).
(7.356)
Here, however, this equation cannot be solved since the Fourier transform of the
differential operator L (i) is the 4 4 -matrix
L (k) = g k 2 + k k ,

(7.357)

and this possesses an eigenvector with eigenvalue zero, the vector k . Thus L (k)
cannot have an inverse, and Eq. (7.356) no solution. Instead, the nonzero spatial
components of the propagator (7.351) satisfy the transverse equation
Lij (i)Gjk (x x ) = iTik (x x )(x0 x0 ).

(7.358)

519

7.4 Free Photon Field

and the transversality condition


i Gij (x x ) = 0.

(7.359)

Let us also calculate the commutator from the Feynman propagator according
to the rules (7.204). It is most conveniently expressed in terms of the explicitly
k 0 -independent projection matrix (4.328) as

[A (x), A (x )] =

d4 k

(k 0 )(k 2 )eik(xx ) Pphys


(k).
4
(2)

(7.360)

Since the polarization sum depends only on the spatial momenta, it can be taken
out of the integral replacing ki by ii , yielding

[A (x), A (x )] = g +

0 /
i j

C(x x ),

(7.361)

where C(x x ) is the commutator function of the scalar field defined in (7.55) with
Fourier representations (7.56), (7.198), and (7.171). Taking the time derivative of
x ) = i (3) (x x ) for x0 = x0
one of the fields we find, using the property C(x
of Eq. (7.57), the canonical equal-time commutator

[A (x, t), A (x , t)] = i g +

0 i j / 2

(3) (x x ),

(7.362)

which was the starting point (7.334) of the quantization procedure in the Coulomb

gauge with E(x, t) = A(x,


t).
It is instructive t see how the quantization goes in the presence of a source term
where the action (7.352) reads
1
A=
2

d x A (x)(g )A (x)

d4 x A (x)j (x).

(7.363)

Going to the individual components A0 and A = (A1 , A2 , A3 ) as in (4.246), this


reads
A=

d4 x

 h
1

A0 (x)(2 )A0 (x) 2A0 (x)0 i Ai (x)


i

A(x)(02 2 )A(x)Ai (x)i i (x) (x)A0 (x) + j(x)A(x) .

(7.364)

As we have observed in (4.246), the Coulomb gauge (4.271) ensures that the field
equation for A0 has no time derivative and can be solved by the instantanous
Coulomb potential of the charge density (x) by Eq. (4.270). The field equation
for the spatial components, however, is hyperbolic is equal to Amperes law:
(02 2 )A(x) = j(x).

(7.365)

This is solved in Eq. (5.11) with the help of the retarded Coulomb potential.

520

7.4.2

7 Quantization of Relativistic Free Fields

Covariant Field Quantization

There exists an alternative formalism in which the photon propagator is manifstely


covariant. This can, however, be achieved only at the expense of extending the
Hilbert space by a nonphysical sector from which the physical subspace is obtained
by certain operator conditions. Such an approach can be followed consistently if the
interaction does not mix physical states with unphysical ones. As a function of time,
physical states must remain physical, i.e., the physical subspace must be invariant
under the dynamics of the system.
Gauge transformations modify only the unphysical states. It will turn out that
half of them have a negative or zero norm and do not allow for a probabilistic interpretation. This, however, does not cause any problems since the physical subspace
is positive definite and dynamically invariant. On it, the second-quantized time evolution operator U = eiHt is unitary and the completeness sums between physical
observables can be restricted to physical states.
The covariant quantization method is based on a modified Lagrangian in which
0
A (x) plays no longer a special role, so that every field component A (x) possesses
a canonically conjugate momentum. For this purpose one introduces an auxiliary
field D(x), and adds to the original Lagrangian a so-called gauge-fixing term18
L(x) L (x) = L(x) + LGF (x)

(7.366)

which is defined by
LGF (x) D(x) A (x) +

2
D (x),
2

0.

(7.367)

Now it is this auxiliary field D(x) which has no canonical momentum:


D (x) =

L(x)
0,

D(x)

(7.368)

so that the Euler-Lagrange equation for D(x) is not a proper equation of motion,
but merely a constraint:
D(x) = A (x).
(7.369)
In contrast to the earlier treatment, all four components of the vector potential
A (x) are now dynamical fields and obey a proper equation of motion:
F (x) = 2 A (x) A (x) = D(x).

(7.370)

Together with the constraint (7.369), we can write these field equations as
1
A (x) = 0.
A (x) 1

18

This modification was proposed by E. Fermi, Rev. Mod. Phys. 4 , 125 (1932).

(7.371)

521

7.4 Free Photon Field

From these we may derive, by one more differentiation, a field equation for D(x):
2 D(x) = 0,

(7.372)

which shows that D(x) is a massless Klein-Gordon field.


If we want to use the new extended Lagrangian for the description of electromagnetism, where the field equations are F = 0, we have to make sure at the
end that D(x) is identically zero at all times. With D(x) satisfying the free Klein
Gordon equation (7.372), this is guaranteed if D(x) satisfies the initial conditions
D(x, t) 0,

D(x,
t) 0,

t = t0 ,
t = t0 .

(7.373)
(7.374)

In other words, if the auxiliary field is zero and does not move at some initial time
t0 , for example at t0 , it will vanish everywhere at all times. With these
initial conditions we can replace the original Lagrangian density by the extended
version L (x). This has the desired property that all field components A (x) possess
a proper canonically conjugate momentum:
(x) =

L(x)
= F0 (x) g0 D(x).
A (x)

(7.375)

Thus, while the spatial field components Ai (x) have the electric fields E i (x) as
their canonical momenta as before, the new auxiliary field D(x) plays the role of a
canonical momentum to the time component A0 (x). We can now quantize the fields
A (x) and the associate field momenta (D(x), E i (x)) by the canonical equal-time
commutation rules:
[A (x, t), A (x , t)] = 0,
h

D(x, t), Ai (x , t)

E i (x, t), E j (x , t)

E i (x, t), A0 (x , t)

= 0,

= 0,

= 0,

E i (x, t), Aj (x , t)

D(x, t), A0 (x , t)

D(x, t), E i (x , t)

[D(x, t), D(x , t)]

= i ij (3) (x x ),

= i (3) (x x ),
= 0,
= 0,

(7.376)

The Hamiltonian density associated with the Lagrangian density L (x) is


D A 0 1 (E2 B2 ) + D A D 2 .
H = A L = E A
2
2

(7.377)

Expressing the electric field in terms of the vector potential, E i = F 0i = A i


i A0 , this becomes
H =

1 2
E + B2 + E A0 + D A D 2 .
2
2

(7.378)

For a vanishing field D(x), this is the same Hamiltonian density as before in (7.330).
Let us now quantize this theory in terms of particle creation and annihilation
operators.

522

7 Quantization of Relativistic Free Fields

Feynman Gauge = 1
The field equations (7.371) become simplest if we take the special case = 1 called
the Feynman gauge. Then (7.371) reduces to
2 A (x) = 0

(7.379)

so that the four components A (x) simply obey four massless Klein-Gordon equations. The fields can then be expanded into plane-wave solutions as
A (x) =

3
X

i
1 h ikx

ikx
e

()a
+
e

()a
k,
k, ,
0
k,=0 2V k

(7.380)

where () are now four momentum-independent polarization vectors


() = g ,

(7.381)

satisfying the orthogonality and completeness relations


X

() ( ) = g ,

g () ( ) = g .

(7.382)
(7.383)

Inverting the field expansion, we obtain


(

ak,
ak,

ik 0 x0

= e

3
X

=0

2V k 0

( )
( )

)Z

d3 x eikx iA (x) + k 0 A (x) ,


(7.384)

and are ready to impose canonical commutation rules. From Eqs. (7.376) we see
that A (x, t) and A (x, t) commute with each other at equal times:
[A (x, t), A (x , t)] = 0.

(7.385)

Expressing the canonical momentum (7.375) in terms of the vector potential,


i (x) = F0i (x) = E i (x) = A i (x) + i A0 (x), the canonical commutator
[i (x, t), Aj (x , t)] = i ij (3) (x x ) implies
[A i (x, t), Aj (x , t)] = iij (3) (x x ) i [A0 (x, t), Aj (x , t)].

(7.386)

But because of (7.385), we have


[A i (x, t), Aj (x , t)] = i ij (3) (x x ).

(7.387)

In contrast to (7.334), the right-hand side is not restricted to the transverse function.
To find the commutation relations for A0 (x, t), we take the canonical commutator
[D(x, t), A0 (x , t)] = i (3) (x x )

(7.388)

523

7.4 Free Photon Field

and express D in the Feynman gauge as A = 0 A0 + i Ai [recalling (7.369)]. The


commutators of Ai (x) with A0 (x) vanish so that (7.388) amounts to
[A 0 (x, t), A0 (x , t)] = i (3) (x x ).

(7.389)

Using the fact that the commutator of D(x, t) with the spatial components Ai (x , t)
vanish by the third of the canonical commutation rules (7.376), the identity (7.369)
leads to
[A 0 (x, t), Ai (x , t)] = 0.
(7.390)
But one also has

[A i (x, t), A0 (x , t)] = 0.

(7.391)

This follows from the canonical commutator [E i (x, t), A0 (x , t)] = 0 after expressing again E i = A i + i A0 , or more directly by differentiating the commutator
[Ai (x, t), A0 (x , t)] = 0 with respect to t.
For the commutators of the time derivatives of the fields among each other we
find similarly
h

A i (x, t), A j (x , t)

A 0 (x, t), A i (x , t)

A 0 (x, t), A 0 (x , t)

E i , E j + [E i , j A0 ] (ij) + [i A0 , j A0 ] = 0,

D j Aj , E i i A0 = 0,
i

D j Aj , D j Aj = 0.

(7.392)

On the right-hand sides, the equal-time arguments x, t and x , t have been omitted,
for brevity.
Summarizing, the commutators between the field components A (x , t) and their
time derivatives are given by the manifestly covariant expressions
h

A (x, t), A (x , t)

= ig (3) (x x ),

[A (x, t), A (x , t)] = 0,


h
i
A (x, t), A (x , t) = 0.

(7.393)

They have the same form as if A (x, t) were four independent Klein-Gordon fields,
except for the fact that the sign in the commutator between the temporal components of A (x, t) is opposite to that between the spatial components. This will be
the source of considerable complications in the subsequent discussion.
Using the covariant commutation rules (7.393), we find from (7.384) the commutation rules for the creation and annihilation operators:
[ak, , ak , ]

= [ak, , ak , ] = 0,

[ak, , ak , ] = k,k g .

(7.394)

It is useful to introduce the contravariant creation and annihilation operators


a
k

3
X

=0

()ak, ,

ak

3
X

=0

()ak, .

(7.395)

524

7 Quantization of Relativistic Free Fields

They satisfy the commutation rules


[ak , ak ]

= [a
k , ak ] = 0,

[ak , a
k ] = k,k g .

(7.396)

The opposite sign of the commutator [A 0 (x, t), A0 (x , t)] shows up in the commutators (7.394) and (7.396) between creation and anihilation operators with polarization
label = 0 or = 0. This has the unpleasant consequence that state created by

the operators a0
k = ak,0 have a negative norm:
h0|a0k ak0 |0i h0|[a0k , a0
k ]0i = h0|0i = 1.

(7.397)

Such states cannot be physical. By applying any odd number of these creation
operators to the vacuum, one obtains an infinite set of unphysical states.
Another infinite set of unphysical states is
3 n
|0in (a0
k ak ) |0i,

n > 1.

(7.398)

3
These have a vanishing norm, as follows from the fact that a0
k ak commutes with
its hermitian conjugate. The label 3 can of course be exchanged by 1 or 2.

Fermi-Dirac Subsidiary Condition


At first sight this seems to make the usual probabilistic interpretation of quantum mechanical amplitudes impossible. However, as announced in the beginning,
this problem can be circumvented. According to Eqs. (7.370)(7.374), the classical
equations of motion are satisfied correctly only if the auxiliary field D(x) vanishes

at some initial time, together with its velocity D(x).


In the quantized version we
have to postulate an equivalent operator property. Exactly the same condition, the
vanishing of the second-quantized field operator D(x), would be too stringent: It
would contradict the canonical quantization rule
[D(x, t), A0 (x , t)] = i (3) (x x ).

(7.399)

We can only require that the equation D(x, t) = 0 is true when applied to physical
states. Thus we postulate that only those states in the Hilbert space are physical
which satisfy the subsidiary conditions
D(x, t)|phys i = 0,

D(x,
t)|phys i = 0

(7.400)

at some fixed initial time t. Because of the equation of motion (7.372) for D(x), this
is equivalent to requiring
D(x, t)|phys i 0
(7.401)
at all times. This is called the Fermi-Dirac condition.19
19

E. Fermi, Rev. Mod. Phys. 4 , 87, (1932); P.A.M. Dirac, Lectures in Quantum Field Theory,
Academic Press, New York, 1966; W. Heisenberg and W. Pauli, Z. Phys. 59 , 168 (1930).

525

7.4 Free Photon Field

In order to discuss its consequences it is useful to introduce the vectors


ks k /k 0 ,

kl k /2k 0 ,

k (k 0 , k).

with k (k 0 , k),

(7.402)

and k 0 = |k| being on the light cone. The quotation marks are used in the same
sense as in Section 4.10.6, where we introduced the vector (4.422). Due to the
normalization factors, both ks and kl do not transform like vectors under Lorentz
transformations. Nevertheless, they do have Lorentz-invariant scalar products with
each other:
ks ks = 0, kl kl = 0, ks kl = 1.
(7.403)
In analogy with the (4.360) (4.422), we shall span the four-dimensional vector space
here by supplementing the transverse polarization vectors (4.321), (7.339) by the
four-component objects
(k, s) = ks ,

(k, s) = ks .

(7.404)

These agree with the earlier-introduced scalar and antiscalar polarization vectors
(4.360) and (4.422) in Section 4.10.6, except for a different normalization which
will be more convenient in the sequel. We shall use the same symbols for these new
objects, to prevent a proliferation of symbols, in spite of a small danger of confusion.
Together with the transverse polarization vectors (4.321), the four vectors satisfy
the orthogonality relation
(k, ) (k, ) = g ,

(7.405)

where the index runs through +1, 1, s, s, and g is the metric

g =

1
0 0 0
0 1 0 0
0
0 0 1
0
0 1 0

(7.406)

The completeness relation reads, in a slight modification of (4.423),

(k, ) (k, )g = g .

(7.407)

=1,s,
s

Multiplying the field expansion (7.380) with the left-hand side of this relation, and
defining creation and annihilation operators with polarization label = 1, s, s by
ak, = ()

3
X

()ak, ,

ak, = ()

3
X

()ak, ,

(7.408)

=0

=0

we obtain the new field expansion


A (x) =

X
k

1
2V k 0

=1,s,
s

eikx (k, )ak, + eikx (k, )ak, ,

(7.409)

526

7 Quantization of Relativistic Free Fields

This expansion may be inverted by euqations like (7.384), expressing the new creation and annihilation operators ak, and ak, in terms of the fields A (x) and

A (x). The result looks precisely the same as in (7.384), except that the metric g

is replaced by g of (7.406), and the polarization vectors ( ) by (k, ).


For the two transverse polarization labels = 1, the new creation and annihilation operators coincide with those introduced during the earlier the noncovariat
quantization in expansion (7.338), i.e., the operators ak, are equal to ak, for polarization labels = = 1. In addition, the new expansion (7.409) contains the
creation and annihilation operators:
ak,s = ks

3
X

()ak, ,

ak,s = ks

()ak, ,

(7.410)

=0

=0

ak,s = ks

3
X

3
X

()ak, ,

ak,s

= ks

3
X

()ak, .

(7.411)

=0

=0

For the spatial momenta k pointing in the z-direction, these are


ak,s = a0k a3k ,

ak,s = a0k + a3k ,

3
ak,s = a0
k ak ,

3
ak,s = a0
k + ak .

(7.412)

Note that the previous physical polarization sum (7.393) may be written as

Pphys
(k) =

=1

(k, ) (k, ) = g + ks ks + ks ks .

(7.413)

The commutation rules for the new creation and annihilation operators are
[ak, , ak , ] = [ak, , ak , ] = 0,
[ak, , ak, ] = g k,k .

(7.414)

By Fourier-transforming the field D(x), we see that the Fermi-Dirac condition


can be rewritten as
ak,s |phys i = 0,

ak,s |phys i

= 0,

(7.415)
(7.416)

i.e., the physical vacuum is annihilated by the scalar creation and annihilation
operators.
Let us calculate the energy of the free-photon system. For this we integrate the
Hamiltonian density (7.378) over all space, insert D(x) = A 0 (x) + i Ai (x), E(x) =
i A0 , B(x) = A(x), and perform some partial integrations, neglecting
A
surface terms, to obtain for = 1 the simple expression
H =

1
d3 x (A A A A ).
2

(7.417)

527

7.4 Free Photon Field

This is precisely the sum of four independent Klein-Gordon energies [compare with
(7.30) and contrast this with (7.346)], one for each spacetime component A (x).
Due to relativistic invariance, however, the energy of the component A0 (x) has an
opposite sign which is related to the above-observed opposite sign in the commutation relations (7.394) and (7.396) for the = 0 and = 0 polarizations of the
creation and annihilation operators.
Inserting now the expansion (7.380) of the field A (x) in terms of creation and
annihilation operators, we obtain by the same calculation which led to Eqs. (7.82)
and (7.347):
H =

X
k

i
i
X k0 h
k0 h

ak, ak, ak, ak, g =


ak, ak ak ak, .
2
k 2

(7.418)

This operator corresponds to an infinite set of four-dimensional oscillators, one for

every spatial momentum and polarization state. Due to the indefinite metric g ,
the Hilbert space is not of the usual type, and the subsidiary conditions (7.415) and
(7.416) are necessary to extract physically consistent results.
The expansion (7.409) in terms of transverse, scalar, and longitudinal creation
and annihilation operators yields the Hamiltonian (7.418) in the form
H =

X
k

k0
2

 X 
=1

ak, ak, + ak, ak, ak,s ak,s ak,s ak,s ak,s ak,s ak,s ak,s .
(7.419)

Let us bring this operator to normal order, i.e., to normal order with respect to the
physical vacuum. For the transverse modes we move the annihilation operators to
the right of the creation operators, as usual. For the modes s and l, the physical
normal order has the operators ak,s and ak,s on the right to make use of subsidiary
conditions (7.415) which are true also for the physical vacuum. Using the two
commutation rules
[ak,s , ak ,s ] = k,k ,
[ak,s , ak ,s ] = k,k ,

(7.420)

the Hamiltonian takes the form:

H =

X
k

 X 

ak, ak,

=1

1
+
ak,s ak,s ak,s ak,s .
2


(7.421)

Note that the ordering of the s and l components produces no constant term due
to the opposite signs on the right-hand sides of the commutation rules (7.420). The
Dirac conditions (7.415) and (7.416) have the consequance that the last two terms
vanish for all physical states. Thus the Hamilton operator contains only the energies
of transverse photons and can be replaced by

Hphys
=

X
k


k0 X 
ak, ak, + ak, ak, .
2 =1

(7.422)

528

7 Quantization of Relativistic Free Fields

Finally, we bring also the transverse creation and annihilation operators to a normal
order (for transverse photons, the ordinary and the physical normal orders are the
same), and obtain

Hphys

X 

ak, ak,

=1

1
.
+
2


(7.423)

The second term gives the vacuum energy:


1X 0 X
k =
k .
2 k
k

E0 h0|H |0i =

(7.424)

In contrast to all other particles, this divergent expression has immediate experimental consequences in the laboratory it is observable as the so-called Casimir
effect, to be discussed Section ??.
The photon number operator obtained by dropping the factor k 0 /2 in the normally ordered part of the Hamiltonian (7.419):
N=

X
k

=1

ak, ak,

ak,s ak,s

ak,s ak,s

(7.425)

With the Dirac conditions (7.415) and (7.416), the last two terms can be omitted
between physical states, so that only the transverse photons are counted:
N=

X X 

ak, ak, .

k =1

(7.426)

Its action on the physical subspace is completely analogous to that of H .


Physical Hilbert Space
In the above calculations of the energies we have assumed that the physical vacuum
|0phys i has a unit norm. This, however, cannot be true. A simple argument shows
that the Fermi-Dirac condition (7.401) is inconsistent with the canonical commutation rule (7.388). For this one considers the diagonal element
h0phys |[D(x, t), A0(x , t)]|0phys i = i (3) (x x )h0phys ||0phys i.

(7.427)

Writing out the commutator and taking D(x, t) out of the expectation values, the
left-hand side apparently vanishes whereas the right-hand side is obviously nonzero.
This problem, however, is not fatal for the quantization approach. It occurs and
can be solved in ordinary quantum mechanics. Consider the canonical commutator
[p, x] = i between localized states |xi:
not

hx|[p, x]|xi = ihx|xi.

(7.428)

The left-hand side gives zero, the right-hand side infinity, so the Heisenberg uncertainty principle seems violated. The puzzle is resolved by the fact that Eq. (7.428) is

529

7.4 Free Photon Field

meaningless since the states |xi are not normalizable. The satisfy the orthogonality
consition -function hx |xi = (x x), in which the right-hand side is a distribution.
The equation must be multiplied by a smooth test function of x or x to turn into a
finite equation, as in the treatment of Eq. (7.42). Similarly, we can derive from the
canonical commutator only the distribution equation
hx |[p, x]|xi = ihx |xi

(7.429)

with x 6= x , which gives correct finite results after such a smearing procedure.
Another way to escape the contradiction is by abandoning the use of completely
localized states |xi and working only with approximately localized states, for example the lattice states |xn i with the proper orthogonality relations (1.137) and
completeness relations (1.142). Another possibility is to use a sequence of wave
packets in the form of narrow Gaussians
1

2
hx |xi = e(x x) /2

(7.430)

which for 0 becomes more and more localized.


The same subtleties exist for the commutation rule (7.388). The physical states
in the Hilbert space are improper states. To see how this comes about let us go to
the position representation of the Hamilton operator (7.418) which reads
H=

X
k

where

k0
2
k
2

!2

2k .

2k = 0k 1k 2k 3k .

(7.431)

(7.432)

The ground state is given by the product of harmonic wave functions


h|0i =

Y
k

1
1/4

ek /2 .

(7.433)

The excited states are obtained by applying to this the creation operators
a
k

= k + g .
k
2

(7.434)

Due to the positive sign of 0k in the ground state (7.433), this procedure does not
produce a traditional oscillator Hilbert space with the scalar product
h |i =

Y Z
k,

dk h |ih|i.

(7.435)

This is not a serious problem, however, since it can be resolved by rotating the
contour of integration in 0k in the complex 0k -plane by 900 , so that it runs along
the imaginary 0k -axis.

530

7 Quantization of Relativistic Free Fields

We now impose the subsidiary conditions (7.415) and (7.416) upon this Hilbert
space. To simplify the discussion we shall consider only the problematic modes with
polarization labels 0 and 3 at a fixed momentum k in the z-direction, so that the
momentum labels can be suppressed,
Then the conditions (7.415) and (7.416) take the simple form
(a0 a3 )|phys i = 0

(7.436)

(a0 a3 )|phys i = 0.

(7.437)

and
In the position representation, these conditions amount to
!

+
h|phys i = 0,
0 3
(0 3 )h|phys i = 0.

(7.438)
(7.439)

The first condition is fulfilled by wave functions which do not depend on 0 + 3 ,


the second restricts the dependence on 0 3 to a -function. The wave function of
the physical vacuum is therefore (for particles of a fixed momentum running along
the z-axis)
1
12
22
h|0phys i = (0 3 ) 1/4 e( + )/2 .
(7.440)

The complete set of physical states is obtained by applying to this any number of
creation operators for transverse photons
!

a1 = 1 1 ,

a2 = 2
.
2
2

(7.441)

The physical vacuum (7.440) displays precisely the unpleasant feature which
caused the problems with the quantum mechanical equation (7.428). Due to the
presence of the -functions the wave function is not normalizable, but a distribution.
The normalization problem of the wave function (0 3 ) is completely analogous
to that of ordinary plane waves, and we know how to do quantum mechanics with
such generalized states. The standard Hilbert space is obtained by superimposing
plane waves to wave packets. Normalizable wave packets can be constructed to
approximate the positional wave functions (0 0 ) in various ways. One may,
for instance, discretize the field variables and work on a lattice in field space.
Alternatively one may replace the -function in (7.440) by a narrow Gaussian
(0 3 ) const e(

0 3 )2 /2

(7.442)

with an appropriate normalization factor.


When performing calculations in the physical Hilbert space of the Dirac quantization scheme, we would like to proceed as in the case of Klein-Gordon and Dirac
particles, assuming that the scalar product h0phys |0phys i is unity, as we did when evaluating the vacuum energy (7.424) (which was really illegal as we have just learned).

531

7.4 Free Photon Field

Only with a unit norm of the vacuum state can we find physical expectation values by simply bringing all creation and annihilation operators between the vacuum
states to normal order, the desired result being the sum of the c-numbers produced
by the commutators. Such a unit norm is achieved by introducing the analogs of
the wave packets (7.430), which in the this case are normalizable would-be vacuum
states to be denoted by |0phys i which approach the true unnormalizabel vacuum
state arbitrarily close for 0. These would-be vacuum states are most easily constructed algebraically.20 Considering only the problematic creation and annihilation
operators a0 , a0 and a3 , a3 , we define an auxiliary state |0aux i to be the one that
is annihilated by a0 and a3 :
a0 |0aux i = 0,

a3 |0aux i = 0.

(7.443)

From this we construct the physical vacuum state as a power series


|0phys i =

n,m=0

cn,m (a3 )n (a0 )m |0aux i.

(7.444)

The subsidiary conditions (7.415) and (7.416) are now


as |0phys i = (a0 a3 )|0phys i = 0,

as |0phys i = (a0 a3 )|0phys i = 0.

(7.445)

These are obviously fulfilled by the coherent state:


3 a0

|0phys i = ea

|0aux i.

(7.446)

To verify this we merely note that a3 acts on this state like a differential operator
/a3 producing a factor a0 , and the same thing holds with a3 and a3 replaced by
a0 and a0 , respectively.
The norm of this state is infinite:
h0phys |0phys i =

X
1
1 = .
h0aux |(a3 )n (a0 )n (a3 )m (a0 )m |0aux i =
n,m=0 n!m!
n=0

(7.447)

We now approximate this physical vacuum by a sequence of normalized states |0phys i


defined by
|0phys i

3 a0

(2 )e(1)a

(2 )

1
[(1 )a3 a0 ]n |0aux i.
n!
n=0

(7.448)

It is easy to check that these have a unit norm:

20

h0phys |0phys i = (2 )

(1 )2n = 1.

(7.449)

n=0

R. Utiyama, T. Imamura, S. Sunakawa, and T. Dodo, Progr. Theor. Phys. 6 , 587 (1951).

532

7 Quantization of Relativistic Free Fields

On this sequence of states, the subsidiary conditions (7.445) are not exactly satisfied.
Using the above derivative rules, we see that

|0phys i = (1 ) a0 |0phys i ,
a3

|0phys i = (1 ) a3 |0phys i ,
=
a0

a3 |0phys i =
a0 |0phys i

(7.450)

corresponding to the approximate subsidiary conditions


as |0phys i = a0 |0phys i ,
as |0phys i = a3 |0phys i .

(7.451)

Let us calculate the norm of the states on the rigth-hand side. Using the power
series expansion (7.448) we find

h0phys |a3 a3 |0phys i

= (2 )

1
h0aux |(1 )n (a3 )n (a0 )n a3 a3 (1 )m (a3 )m (a0 )m |0aux i
n!m!
n,m=0

= (2 )

(n + 1)(1 )2n ,

(7.452)

n=0

and thus

h0phys |a3 a3 |0phys i =

1
.
(2 )

(7.453)

h0phys |a0 a0 |0phys i =

1
,
(2 )

(7.454)

a3 a3 = 1 + a3 a3 ,

(7.455)

Similarly we derive

and via the commutation rules


a0 a0 = 1 + a0 a0 ,
the expectation values

h0phys |a0 a0 |0phys i = h0phys |a3 a3 |0phys i =

(1 )2
,
(2 )

(7.456)

These results can actually be obtained by trivial algebra directly from


Eqs. (7.451) using the commutation relations (7.455).
Equations (7.453) and
q (7.454) show that the norm of the states on the right-hand
side of Eqs. (7.451) is /(1 ), the normalized states |0phys i satisfy the desired
constraints (7.445) in the limit 0, thus converging towards the physical vacuum
state.

533

7.4 Free Photon Field

To be consistent, all calculations in the Dirac quantization scheme have to be


done in one of the would-be vacuum states |0phys i , with the limit taken at the end.
Only in this way, the earlier-observed apparent contradiction (7.427) and (7.428)
can be avoided. In the present context, it is observed when taking the expectation
value of the commutation relation [as, as ] = 1:
h0phys |asas |0phys i h0phys |as as|0phys i = h0phys |0phys i.

(7.457)

Since as annihilates the physical vacuum to its right and to its left, the left-hand
side should be zero, contradicting the right-hand side. For the normalized finite-
would-be vacuum, on the other hand, the equation becomes

h0phys |asas |0phys i h0phys |as as|0phys i = 1.

(7.458)

From (7.453)(7.454) we find

h0phys |asas |0phys i = 1,

h0phys |as as|0phys i = 1,

(7.459)

so that (7.458) is a correct equation. Thus, although the states (7.451) have a small

norm of order , the products asas and as as have a unit expectation value.
We can easily see the reason for this by calculating
as|0phys i = (2 ) a0 |0phys i ,
as|0phys i = (2 ) a3 |0phys i ,

(7.460)

and observing that the norm


and
q of the states
q on the right-hand sides is, by (7.453)

(7.454), equal to (2 )/ (2 ) = (2 )/, which diverges like 1/ for


0 [being precisely the inverse of the norm of the states (7.451)]. Thus we can
only drop safely expectation values in which as or as act upon the physical vacuum
to its right, if there is no operator as or as doing the same thing to its left. This
could be a problem with respect to the vacuum energy of the unphysical modes in
(7.419), which is proportional to
h0phys |as as + asas + asas + as as|0phys i .

(7.461)

Fortunately, this expectation value does vanish, as it should, since by (7.453)


(7.456):
h0phys |as as|0phys i = h0phys |asas |0phys i = 1/2,

h0phys |as as|0phys i = h0phys |asas |0phys i = 1/2,

(7.462)

so that the energy of the unphysical modes in the normalized would-be vacuum
state |0phys i is indeed zero for all , and the true vacuum energy in the limit 0
contains only the zero-point oscillations of the physical transverse modes.
In this context it is worth emphasizing that the normally ordered operator a0 a0
a3 a3 is not normally ordered with respect to the physical vacuum |0phys i and thus

534

7 Quantization of Relativistic Free Fields

does not yield zero when sandwiched between two such states. This is obvious by
rewriting a0 a0 a3 a3 as asas + as as, and the fact that the annihilation operator as
on the right-hand side do not annihilate the physical vacuum: ak,s |0phys i =
6 0.
In axiomatic quantum field theory, the vacuum is always postulated to be a
proper discrete state with a unit norm. Without this properties, the discussion of
symmetry operations in a Hilbert space becomes quite subtle, since infinitesimal
transformations can produce finite changes of a state. The Dirac vacuum |0phys i
is not permitted by this postulate, whereas the pseudophysical vacuum |0phys i is.
There are other problems within axiomatic quantum field theory, however, due to the
nonuniqueness of this vauum, and due to the fact that the vacuum is not separated
from all other states by an energy gap, an important additional postulate. That
latter postulate is actually unphysical, due to the masslessness of the photon and
the existence of many other massless particles in nature. But it is necessary for the
derivation of many rigorous results in that somewhat esoteric discipline.
Because of the subtleties with the limiting procedure of the vacuum state, it will
not be convenient to work with the Dirac quantization scheme but use a simplified
procedure due to Gupta and Bleuler to be introduced below. First, however, we
shall complete the present discussion by calculating the photon propagator within
the Dirac scheme.
Propagator in Dirac Quantization Scheme
Let us calculate the propagator of the A -field in the Dirac quantization scheme.
We take the expansion of the field operator (7.409) and evaluate the expectation
value
G (x, x ) = h0phys |T A (x)A (x )|0phys i ,

(7.463)

letting 0 at the end. For x0 > x0 , we obtain a contribution proportional to


(x0 x0 )

1 X 1 ik(xx )
e
V k 2k 0

(7.464)

multiplied by the sum of the expectation values of


(k, +1) (k, +1)ak,+1 ak,+1 ,

(k, 1) (k, 1)ak,1 ak,1 ,

(7.465)

and
(k, s) (k, s)ak,s ak,s ,

(k, s) (k, s)ak,s ak,s ,

(7.466)

and a contribution proportional to


(x0 x0 )

1 X 1 ik(xx )
e
V k 2k 0

(7.467)

535

7.4 Free Photon Field

multiplied by the sum of the expectation values of


(k, +1) (k, +1)ak,+1 ak,+1 ,

(k, 1) (k, 1)ak,1 ak,1 ,

(7.468)

(k, s) (k, s)ak,s ak,s .

(7.469)

and
(k, s) (k, s)ak,s ak,s ,

The first two terms containing transverse photons produce a polarization sum
(k, +1) (k, +1) + (k, 1) (k, 1)

(7.470)

The expectation values of the normally ordered transverse photon terms (7.468)
vanish. The remaining terms in (7.466) and (7.469) are evaluated with the help
of the matrix elements (7.462) of opposite signs, and cancel each other, due to
2
2
the evenness of (k, s) (k, s) = k 0 k 3 functions of k . Such a function as a
factor inside the momentum sums (7.464) and (7.467) preserves their equality, thus
ensuring the cancelation of the contributions from (7.466) and (7.469).
For x0 < x0 , we obtain once more the same expression with spacetime and
Lorentz indices interchanged: x x , .
The propagator becomes therefore
G (x, x ) =

h0phys |T A (x)A (x )|0phys i

= (x0 x0 )
+ (x0 x0 )

2
1 X 1 ik(xx ) X
(k, ) (k, )
e
V k 2k 0
=1

2
1 X 1 ik(xx ) X
(k, ) (k, ).
e
V k 2k 0
=1

(7.471)

This is the same propagator as in Eq. (7.351), where it was obtained in the manifestly
noncovariant quantization scheme, in which only the physical degrees of freedom
of the vector potential were made operators. Thus, although the field operators
have been quantized with the covariant commutation relations (7.376), the selection
procedure of the physical states has produced the same noncovariant propagator.
It is worth pointing out the difference between the propagator (7.351) and the
retarded propagator used in classical electrodynamics. The classical one is given by
same Fourier integral as G(x x ):
GR (x x )

d4 k ik(xx ) i
e
,
(2)4
k2

(7.472)

except that the poles at k 0 = |k| are both placed below the real axis. The
quantum-field-theoretic k 0 -integral
Z

dk 0 ik0 (x0 x0 ) i
1
1
(7.473)
e
0
0
2
2|k| k |k| + i k + |k| i
1
1 i|x0 x0 |
0
0
0
0
=
[(x0 x0 )ei(x x ) + (x0 x0 )ei(x x ) ] =
e
2
2
!

536

7 Quantization of Relativistic Free Fields

with on the light cone, = |k|, is to be compared with the retarded expression
Z

1
1
dk 0 ik0 (x0 x0 ) i
(7.474)
e
0
0
2
2|k| k |k| + i k + |k| + i
1
1
0
0
0
0
sin[(x0 x0 )].
= (x0 x0 ) (ei(x x ) ei(x x ) ) = 2i(x0 x0 )
2
2
!

The angular part of the spatial integral


Z

d3 k ikx
e
(2)3

(7.475)

can be done in either case in the same way leading to


1 Z
d sin(R).
2 2 R 0

(7.476)

Thus we find
GR (x x ) = i(x0 x0 )

d{ei[(x

0 x0 )R]

ei[(x

0 x0 )+R]

(7.477)

Since R and x0 x0 are both positive, the result is


GR (x x ) = i(x0 x0 )

1
(x0 x0 R).
4R

(7.478)

The retarded propagator exists only for causal times x0 > x0 and it is equal to
the Coulomb potential if the end points can be connected by a light signal. In
contrast to this, the propagator of the quantized electromagnetic field, in which the
creation operator of antiparticles accompanies the negative energy solutions of the
wave equation, is
G(x x ) = i

1
1
+ (|x0 x0 | R) + i
+ (|x0 x0 | R).
4R
4R

where

(7.479)

d it
e
(7.480)

is equal to twice the positive-frequency part of Fourier decomposition of the function. With the help of an infinitesimal imaginary part in the time argument,
the integral can be done and gives
+ (t) =

+ (t) =

1 1
.
i t i

(7.481)

This can be decomposed as follows:


1
+ (t) =
i

i
t
+
2
2
t +
t + 2

= (t) +

1 P
.
i t

(7.482)

537

7.4 Free Photon Field

The second term selects in any integral involving + (t) the principal value. Note
that + (t) shares with (t) the property
(t2 R2 ) =

1
[(t R) + (t + R)].
2R

(7.483)

Thus the propagator G(x x ) can therefore also be written as


G(x x ) = i

1
+ ((x x )2 ).
2

(7.484)

to be compared with the retarded propagator


GR (x x ) = (x0 x0 )i

1
((x x )2 ).
2

(7.485)

The massive propagators in configuration space are, incidentally, quite simply related
to the massless ones. One simply writes the -functions ((xx )2 ) and + ((xx )2 )
and in GR (x x ) and G(x x ) as Fourier integrals
2

((x x ) ) =

d i(xx )2
e
,
2

+ ((x x ) ) =

d i(xx )2
e

(7.486)

M2

and replaces ei(xx ) by ei(xx ) 2 as in (7.133). With this additional term


in the exponent, the -integrals yield Bessel functions:

2
M
J
(M
x2 )
1

,
(7.487)
(x2 ) (x2 ) + (x2 )
2
M x2

M 2 J1 (M x2 ) iN1 (M x2 )
2
2

+ (x ) (x ) + P
.
(7.488)
2
M x2
For M 0 we use the limiting behavior J1 (z) z/2, N1 (z) 2/z and see that
in the first equation the mass term disappears whereas in the second term it yields
P/ix2 . The polarization vectors have to satisfy
1
k (k, ) 1
k k (k, ) = 0.

(7.489)

The propagator of the fields A in momentum space is obtained by inverting the


inhomogeneous version of this equation: Correspondingly, the polarization vectors
in the field expansion (7.380) solutions will now, instead of (7.382)(7.383), satisfy
the completeness and orthogonality relations

g (1 )

k k
k2

(k, )

(k, ) = g

(k, ) (k, ) = g .

k k
(1 ) 2 ,
k

(7.490)
(7.491)

538

7 Quantization of Relativistic Free Fields

Gupta-Bleuler Subsidiary Condition


One may wonder whether it is possible to avoid the awesome limiting procedure
0 and find a way of working with an ordinary vacuum state. This is indeed
possible if one is only interested in physical matrix elements containing at least one
particle. Such matrix elements can be calculated using only the condition (7.415),
while discarding the condition (7.416). Thus, we select what we shall call pseudophysical states by requiring only
ak,s |phys i = 0.

(7.492)

The restriction to this condition is the basis of the Gupta-Bleuler approach21 to


quantum electrodynamics. Take, for instance, the number operator (7.425) and
insert it between two pseudo-physical states selected in this way. It yields

hphys
|N|phys i = hphys
|ak,+1 ak,+1 + ak,1 ak,1 |phys i

hphys
|[ak,s ak,s + ak,s ak,s ]|phys i.

(7.493)

The second line vanishes because due to the condition (7.438) and its Hermitian
adjoint. Thus the number operator counts only the number of transverse photons.
The same mechanism ensures correct particle energies. We bring the Hamilton operator (7.421) with the help of the first commutation rule in (7.420) to the
normally ordered form in the pseudo-physical vacuum state |0phys i

H =

X
k

 X 

ak, ak,

=1

1
+ 1 ak,s ak,s ak,s ak,s .
+
2


(7.494)

and see that the last two terms vanish because of the Gupta-Bleuler condition
(7.492). Thus the Hamilton operator counts only the energies of the transverse
photons.
So why did we have to impose the second Dirac condition (7.416) at all? It is
needed to ensure only one property of the theory: the correct vacuum energy. In
the Gupta-Bleuler approach this energy comes out wrong by a factor two showing
that the unphysical degrees of freedom have not been completely eliminated. The
normal ordering of the Hamilton operator (7.421) with respect to the pseudo-physical
vacuum state |0phys i has produced an extra vacuum energy
E0extra =

X
k

k0 =

k ,

(7.495)

which previously did not appear because of the second Dirac condition (7.416). It is
the vacuum energy of two unphysical degrees of freedom for each momentum k. This
21

K. Bleuler, Helv. Phys. cta 23 , 567 (1950), S.N. Gupta, Proc. Phys. Soc. (London) A 63 , 681
(1950); A 64 , 426 (1951); Phys. Rev. 77 , 294L (1950); 117 , 1146 (1960). For a recent discussion
see L.P. Prokhorov, Usp. Fiz. Nauk. 154 , 299 (1988) [transl. in Sov. Phys. Usp. 31 , 151 (1988).

539

7.4 Free Photon Field

wrong vacuum energy may also be seen directly by bringing the Hamilton operator
(7.418) to normal order using the commutation rule (7.396), yielding
H =

a
k ak,

1
,
+
2


(7.496)

the second term showing the vacuum energy of four polarization degrees of freedom,
rather than just the physical ones.
A heuristic remedy to the wrong vacuum energy in the Gupta-Bleuler formalism
will be discussed below, to be followed by a more satisfatory one in Section 13.13.
Photon Propagator in Gupta Bleuler Approach
We now calculate the propagator of the A -field in the Gupta-Bleuler quantization
scheme, which is given by the expectation value in the pseudo-physical vacuum state
G (x, x ) = h0phys |T A (x)A (x )|0phys i.

(7.497)

Expanding the vector field A as in (7.409), and inserting it into (7.497), we obtain
for x0 > x0 the same contributions as in Eqs. (7.464)(7.469), except that the
expectation values are to be taken in the pseudo-physical vacuum state |0phys i,
where only ak,s |0phys i = 0 and h0phys |ak,s = 0. As a consequence, there are
contibutions from the expectation values of the terms (7.466) yielding a polarization
sum
(k, s) (k, s) + (k, s) (k, s).
which, together with the transverse sum (7.470), adds up to
X

=1

(k, ) (k, ) (k, s) (k, s) (k, s) (k, s) = g .

(7.498)

For x0 < x0 , the same expression is found with spacetime and Lorentz indices
interchanged: x x , , so that the propagator becomes
G (x, x ) = h0phys |T A (x)A (x )|0phys i
(7.499)
"
#
X 1
1 X 1 ik(xx)
1

ik(xx )

(x0 x0 )
= g
e
+ (x0 x0 )
e
V k 2k 0
V k 2k 0
Up to the prefactor g this agrees with Eq. (7.46) for the propagator G(x, x )
of the Klein-Gordon field for zero mass, which we write as in (7.61) as
G (x, x ) = h0|T A (x)A (x )|0i
h

= g (x0 x0 )G(+) (x, x ) + (x0 x0 )G() (x, x )


= g G(x, x ).

(7.500)

This photon propagator is much simpler than the previous one in (7.351), making it
much easier to perform calculations, this being the main virtue of the Gupta-Bleuler

540

7 Quantization of Relativistic Free Fields

quantization scheme. A second advantage of the propagator (7.500) is that in contrast to the noncovariant propagator (7.356), it is a proper Green function associated
with the wave equation (7.379) in the = 0 -gauge, satifying the inhomogeneous
field equation
2 G (x x ) = g i (4) (x x ).
(7.501)
A further special feature of the Gupta-Bleuler quantization procedure which is
worth pointing out is that the condition ak,s |phys i = 0, does not uniquely fix
the pseudo-physical vacuum state |0phys i. The vacuum state |0i = is one possible
choice, but infinitely many other states are equally good candidates, as long as
their physical photon number are zero. Focusing attention only on states of a fixed
photon momentum parallel to the z-axis, the zero-norm states (7.398) all satisfy the
condition ak,s |phys i = 0. As a consequence, any superposition of such states
(1)

(2)

|0i = |0i + ck |0i1 + ck |0i2 + . . .

(7.502)

= 0,
has a unit norm, contains no transverse photon, satisfies the condition ak,s |0i
and is thus an equally good pseudo-vacuum state of the system.
The expectation value of the vector potential A in this vacuum is nonzero. Its
kth component is equal to the coefficient c(1) associated with this momentum, i.e.,
it is some function c(k). Because of the second of the commutation rules (7.420),
its kth Fourier component (k, )ak, has an expectation (k, s), with a coefficient
(1)
ck = ck :
h0|A |0i = ck (k, s),
(7.503)
i.e., it is proportional to the four-momentum k . In x-space, the expectation value
of such a field is equal to a pure gradient (x) and carries no electromagnetic
field.
The condition (7.492). implies that
h0|ak,0 ak,0 |0i = h0|ak,3 ak,3 |0i.

(7.504)

The two sides are equal to some function of k, say k 0 (k). In a covariant way, the
vacuum is thus characterized by the expectations
h0|ak, ak , |0i = k (k, )k (k, )(k)k,k
k, a |0i = [g + k (k, )k (k, )(k)]k,k .
h0|a

(7.505)

k ,

If we use any of these vacua and do not impose the condition (7.492), the propagator
G0 (x x ) = h0|T A (x)A (x )|0i

(7.506)

is a gauge-transformed version of the original one. Let us calculate it. Using (7.505)
we have
G0 (x x ) = h0|T A (x)A (x )|0i

541

7.4 Free Photon Field

1 X 1 ik(xx )


(k, )(k)]

(k,
)k
e
[(x

x
)g
+
k

0
V k,, 2k 0
(k, ) (k, )

1 X 1 ik(xx )

e
[(x0 x0 )g + k (k, )k (k, )(k)]
0
V k,, 2k
(k, ) (k, )

1 X 1 ik(xx )
e
[(x0 x0 )g + k k (k)]
V k 2k 0
1 X 1 ik(xx )
+
e
[(x0 x0 )g + k k (k)].
V k 2k 0

(7.507)

The g -terms are again equal to G of Eq. (7.500). The k k terms may be placed
outside the integrals where they become spacetime derivatives of a function
F (x x ) =

1 X 1 ik(xx )

[e
+ eik(xx ) ](k).
0
V k 2k

(7.508)

Thus we find
G0 (x x ) = G0 (x x ) F (x x ).

(7.509)

Physical observables must, of course, be independent of F (x x ). This will be


proved in Chapter 12 after Eq. (12.94).
Note that in contrast to G (x x ), the Fourier components of F (x x ) are
restricted to the light cone. This is seen more explicitly by rewriting the sums in
(7.508) as integrals, and introducing the -function
(k 2 ) =

1
[(k 0 k ) + (k 0 + k )],
2k

(7.510)

so that F (x x ) is equal to the four-dimensional integral

F (x x ) =

d4 k ik(xx )
e
2(k 2 )(k).
(2)4

(7.511)

The function F (x x ) is therefore a solution of the Klein-Gordon equation 2 F (x


x ) = 0. This is the same type of freedom we are used to from Green functions.
Different Green functions satisfying the same inhomogeneous differential equation
2 G = ig (4) (x x ).

(7.512)

differ from each other by solutions of the homogeneous differential equation, and
so do the propagators G0 (x x ) and G0 (x x ) coming from different vacuum
states.

542

7 Quantization of Relativistic Free Fields

Remedy to Wrong Vacuum Energy in Gupta-Bleuler Approach


In the Hamilton operators (7.494) and (7.496) we found an important failure of the
Gupta-Bleuler quantization scheme. The energy contains also the vacuum oscillations of the two unphysical degrees of freedom oscillating in 0- and the k-direction.
Due to the opposite commutation rules of the associated creation and annihilation
operators, there is no cancellation of the two contributions. The Gupta-Bleuler
approach eliminates the unphysical modes from all multiparticle states except for
the vacuum. The elimination of this vacuum energy in the covariant approach was
achieved only much later by Fadeev and Popov.22 Their method was developed in
the attempt to quantizing nonabelian gauge theories. There the elimination of the
unphysical gauge degrees of freedom can no longer be done with the help of the
Gupta-Bleuler subsidiary condition (7.492). Essential to their approach is the use
of the functional integral formulation of quantum field theory. In it, all operator
calculations are replaced by products of infinitely many integrals over classical fluctuating fields. This formulation will be introduced in Chapter 13. When fixing a
covariant gauge in the gauge-invariant quantum partition function of the electromagnetic fields, one must multiply the functional integral by a factor removing the
partition function of the unphysical modes. At zero temperature, this factor subtracts, in the abelian gauge theory of QED, precisely the vacuum energies of the
two unphysical modes. The removal requires an equal amount of negative vacuum
energy. In Eq. (7.234) we have observed that fermions have negative vacuum energies. The subtraction of the unphysical vaccum energies can therefore be achieved
by introducing two fictitious Fermi fields, a ghost field C(x) and anti-ghost field

C(x).
They are completely unphysical objects with spin zero. They never appear
in the physical Hilbert space,23 i.e., the physical state vectors satisfy the conditions
C (x)|0phys i = 0,

C (x)|0phys i = 0.

(7.513)

These fields are called Fadeev-Popov ghosts. In QED with gauge A (x) = 0, their
only contribution lies in the negative vacuum energy
E0ghosts = 2

X
k

k0
2

(7.514)

which cancels the excessive vacuum energy (7.495) in the Gupta-Bleuler approach.
In quantum electrodynamics, the effect of the ghosts is trivial so that their
introduction is superfluous for all purposes except for calculating the vacuum energy.
This is the reason why their role was recognized only late in the development.24 In
nonabelian gauge theories, the ghost fields have nontrivial interactions with the
22

L.D. Fadeev and V.N. Popov, Phys. Lett. B 25 , 29 (1967).


For this reason, the failure to display the spin-statistics connection to be derived in Section
7.8 does not cause causality problems.
24
R.P. Feynman, Acta Phys. Polonica, 24 , 697 (1963). See also G. Barnich, Gen. Rel. Grav.
43 , 2527 (2011)
23

543

7.4 Free Photon Field

physical particles, which must be included in all calculations to make the theory
consistent.
In the light of this development, the correct way of writing the Lagrangian density
of the free-photon field is
1
C. (7.515)
Ltot = L + LGF + Lghost = F 2 D(x) A (x) + D 2(x)/2 i C
4
We shall see that this Lagrangian has an interesting new global symmetry between
called supersymmetry, which will
the Bose fields A , D and the Fermi fields C, C,
be discussed in Chapter 25. In this context it is called BRS symmetry.25
The Gupta-Bleuler gauge condition plus the subsidiary ghost conditions (7.513)
eliminating the unphysical degrees of freedom are equivalent to requiring that the
charges generating all global symmetry transformations annihilate the physical vacuum. Although structurally somewhat complicated, this is the most satisfactory
way of formulating the covariant quantization procedure of the photon field.
The pseudophysical vacuum |0phys i of the Gupta-Bleuler approach has a unit
norm, which is an important advantage over the physical vacuum |0phys i of the
Dirac approach. In axiomatic quantum field theory it can be proved26 that for any
local operator O(x)
O(x)|0i = 0

if and only if

O(x) = 0.

(7.516)

This is the reason why the Gupta-Bleuler subsidiary condition involves necessarily
a nonlocal operator. Recall that the operator in (7.492) involves only the positivefrequency part of the local field D(x). The negative frequencies would be needed to
make the field local. This will become clear in proving the spin-statistic relation in
Section 7.8.
Arbitrary Gauge Parameter
A quantization of the electromagnetic field is possible also for different values of
the gauge parameter in the Lagrangian density (7.367) which so far has been set
equal to unity, for simplicity. For an arbitrary value of , the field equation (7.371)
can be written as
L (i)A (x) = 0,
(7.517)
with the differential operator
2

L (i) g
25

1
+ 1
.

(7.518)

C. Becchi, A. Rouet, and R. Stora, Ann. Phys. 98 , 287 (1976).


R.F. Streater and A.S. Wightman, PCT, Spin & Statistics, and All That , Benjamin, New
York, 1964, pp. 139, 163; F. Strocchi, Phys. Rev. 162 , 1429 (1967). See also the discussion of N.
Nakanishi, in Quantum Electrodynamics, ed. by T. Kinoshita, World Scientific, Singapore, 1990,
pp. 3680 .
26

544

7 Quantization of Relativistic Free Fields

Multiplying Eq. (7.517) by we find that the vector potential cannot have any
four-divergence:
A (x) = 0.
(7.519)
In a gauge-invariant formulation, this property can be chosen as the Lorentz gauge
conditon. Since the Lagrangian density (7.367) breaks gauge invariance, the Lorentz
condition is a consequence of the equations of motion. For a field A (x) satisfying
the Lorentz gauge condition, the equations of motion reduce to 4 Klein-Gordon
equations, just as before in the Feynman case = 1 [see (7.379)]:
2 A (x) = 0.

(7.520)

We shall not go through the entire quantization procedure and the derivation of
the propagator in this case. What can be given without much work is the Green
function defined by the inhomogeneous field equation
L (i)G (x x ) = i (4) (x x ).

(7.521)

In momentum space, it reads


L (k)G (k) = i .

(7.522)

with the 4 4 -matrix


L (k) = k 2 g

k k
2
k

k 2 k k
.
k2

(7.523)

This matrix has an eigenvector k with an eigenvalue k 2 /, and three eigenvectors


orthogonal to it with eigenvalues k 2 . associated with the eigen For a finite , the
matrix has an inverse, which is found by rewriting L (k) as
L (k) = k 2 PT (k) +

k2
PL (k).

(7.524)

with
PT (k) = g
PL (k) =

k k
.
k2

k k
,
k2
(7.525)

These matrices are projections, since both satisfy the defining relation
P P = P .

(7.526)

The inverse of (7.524) is found by inverting the coefficients in front of PT (k) and
PL (k):
1

1
k k
1
(k) = 2 PT (k) + 2 PL (k) = 2 g 2
k
k
k
k

kk
,
k2 k2

(7.527)

545

7.4 Free Photon Field

so that we can solve (7.522) by


G (k) = iL1 (k).

(7.528)

The propagator has therefore the Fourier representation

G (x x ) =

d4 k ik(xx ) i
k k

g + (1 ) 2
.
e
(2)4
k 2 + i
k + i
"

(7.529)

As usual, we have inserted imaginary parts i to regularize the integral in the same
way as in the propagator of the scalar field. For the term k k /(k 2 + i) the prescription how to place the additional poles is not immediately obvious. The best
way to derive it is by giving the photon a small mass and setting it later equal to
zero. The reader is referred to the next section where the massive vector meson is
treated. The reader may be worried about the physical meaning of the double-pole
in the integrand of the propagator (7.529). A pole in the propagator is associated
with a particle state in the Hilbert space. How does a double pole manifest itself in
the Hilbert space? Heisenberg called the associated state a dipole ghost.27 Instead
of satisfying a Schrodinger equation
(E H)|i = 0,

(7.530)

such a state satisfies the two equations


(E H)|i =
6 0,

(E H)2 |i = 0,

(7.531)

and is unphysical. In the present context. these unphysical states do not cause any
problems since they are an artifact of the gauge fixing and do not contribute to any
observable quantity, due to gauge invariance. The same will be true in the presence
of moving charges in Chapter 12.
It is possible to derive the same result for the propagator of the photon field
governed by the general Lagarangian density (7.366). This exercise is left to the
reader.
Since the D(x)-field is now equal to A (x)/ [recall (7.369)], the gauge parameter enters in various commutation rules between A (x, t) and A (x, t). Proceeding as in Eqs. (7.385)(7.393), we find that the gauge parameter enters into
the commutation rules among A (x, t) and A (x, t) in Eqs. (7.389) and (7.392) as
follows:
[A 0 (x, t), A0 (x , t)] = [D(x, t), A0 (x , t)] = i (3) (x x ),
i
A i (x, t), A j (x , t) = 0,

(7.532)
(7.534)

= (1 )ii (3) (x x ),
= 0.

(7.535)

A 0 (x, t), A i (x , t)

A 0 (x, t), A 0 (x , t)

27

W. Heisenberg, Nucl. Phys. 4 , 532 (1957).

(7.533)

546

7 Quantization of Relativistic Free Fields

The other commutators (7.387), (7.390), and (7.391) remain unchanged.


These commutators are needed to verify that the vacuum expectation of the
time-ordered product
G (x x ) h0|T A (x)A (x )|0i

(7.536)

satisfies the field equation with a -function source (7.637). The proof of this is
nontrivial:


1
G (x x ) =




1
A (x)A (x )|0i (x0 x0 )g h0|[A (x), A (x )]|0i
1

2 g 1


h0|T 2 g

1
+ 1
g0 (x0 x0 )h0|[ A (x), A (x )]|0i = i (4) (x x ),

(7.537)

the right-hand requiring the commutation relations (7.532)(7.535).


It is useful to check the consistency of the propagator with the canonical commutation rules. This is done in Appendix 7B.

7.4.3

Behavior under Discrete Symmetries

Under P,C,T, the vector potential has the transformation properties [recall Section
4.6]:
P

A (x)
AP (x) = A (
x),

A (x)
AT (x) = A (xT ),

A (x)
AC (x) = A (x).

(7.538)
(7.539)
(7.540)

The photon has a negative charge parity which coincides with that of the vector

current V (x) = (x)


(x) formed from two Dirac fields in Table 4.1, to which it
will later be coupled in Chapter 12.
These properties are implemented in the second-quantized Hilbert space by transforming the photon creation operators as follows:
Pak, P 1 ak, ,

T ak, T 1
Cak, C 1

ak, ,
ak, .

(7.541)
(7.542)
(7.543)

The annihilation operators transform by the hermitian-adjoint relations.

7.5

Massive Vector Bosons

Besides the photon, there exist also massive vector bosons. The most important
examples are the fundamental vector bosons mediating the weak interactions, as we

547

7.5 Massive Vector Bosons

shall see in Chapter 26. They can be electrically neutral or carry an electric charge
e. There exist also strongly interacting vector particles of a composite nature.
They are short-lived and only observable as resonances in scattering experiments,
the most prominent being a resonance in the two-pion scattering amplitude. These
are the famous -mesons which will be discussed in Chapter 24. They also carry
charges 0, 1. The actions of a neutral and charged massive vector field V (x) were
given in Eqs. (4.334) and (4.336):
Z

A=

d4 x L(x) =

1
1
d4 x F F + M 2 V V ,
4
2


(7.544)

1
F + M 2 V V ,
d x F
2

(7.545)

and
A=

d x L(x) =

with the equations of motion


F + M 2 V = 0,

(7.546)

or
[( 2 M 2 )g + ]V (x) = 0.

(7.547)

implying that the vector field V (x) has no four-divergence:


V (x) = 0,

(7.548)

( 2 M 2 )V (x) = 0.

(7.549)

and the Klein-Gordon equations

It will sometimes be convenient to view the photon as an M 0 -limit of a


massive vector meson. In this case we have to add a gauge fixing term to the
Lagangian to have a proper limit. Thus we consider the Lagrangian
A=

d x L(x) =

1
1
1
d x F F + M 2 V V
( V )2 ,
4
2
2

(7.550)

1
V = 0,

(7.551)

leading to a field equation


F + M 2 V +
or


( M )g

1
+ 1
V (x) = 0.

(7.552)

Multiplying (7.550) with from the left gives for the divergence V the KleinGordon equation
( 2 + M 2 ) V (x) = 0,
(7.553)
which yields the constraint (7.548) in the limit .

548

7.5.1

7 Quantization of Relativistic Free Fields

Field Quantization

The canonical field momenta are


=

= F0
,
[0 V ]

(7.554)

where the complex conjugation is of course irrelevant for a real field. As for electromagnetic fields, the zeroth component V 0 (x) has no canonical field momentum.
There is again a primary constraint 0 = 0. This is of course due to the fact the
Euler-Lagrange equation Eq. (4.339) for V 0 (x) is not a dynamical field equation.
For the spatial components V i (x), the nonvanishing canonical equal-time commutation rules are
[i (x, t), V j (x , t)] = ii j (3) (x x ),

[i (x, t), V j (x , t)] = ii j (3) (x x ).

(7.555)

The commutators involving V 0 (x, t) can be calculated from this using relation
(4.339), yielding
[V 0 (x, t), V i (x , t)] =

i
i (3) (x x ),
M2

(7.556)

and
[V 0 (x, t), V i (x , t)] = 0.

(7.557)

To exhibit the particle content in the second-quantized fields, we now expand V (x)
into the complete set of classical solutions of Eqs. (7.546) and (7.548) in a large but
finite volume V :
V (x) =

k,s3 =0,1

i
h
1
eikx (k, s3 )ak,s3 + eikx (k, s3 )bk,s3 .
2V k 0

(7.558)

For real vector fields, the same expansions hold with particles and antiparticles
identified.
Dealing with massive vector particles, we may choose to label the polarization
states by canonical spin orientations s3 , the third components of angular momentum in the particles rest frame. The polarization vectors were given explicitly in
Eq. (4.343).
As for scalar and Dirac fields, we have associated the expansion coefficients of the
plane waves of negative energy eikx with a creation operator bk,s3 rather than a second
type of annihilation operator dk,s3 . In principle, the polarization vector associated
with the antiparticles could have been some charge-conjugate polarization vector
c (k, s3 ) in analogy to the spinors v(p, s3 ) = uc (p, s3 ) [recall (4.672)]. For a vector
field , however, which transforms under real 4 4 Lorentz matrices , charge
conjugation changes the polarization vector (or tensor) simply into the complex
conjugate one. This will be proved below.

549

7.5 Massive Vector Bosons

Alternatively we may expand the field V (x) in terms of helicity polarization


states as
V (x) =

X
k

i
X h
1

eipx H (k, )ak, + eipx


(k,
)b
H
k, .
2V k 0 =1,0

(7.559)

where H (k, ) are the polarization vectors (4.350).


The quantization of the massive vector field with the Lagrangian (7.550) proceeds
most conveniently by introducing a transverse field
VT (x) V +

1
V (x),
2
M

(7.560)

which is divergenceless as a consequence of 7.553:


VT (x) = 0.

(7.561)

The full vector field is then a sum of a purely transverse and the gradient of a scalar
field V (x):
1
V (x) = VT (x)
V (x).
(7.562)
M 2
It can be expanded in terms of three cration and annihilation operators ak, , ak,
for the three physical polarization states with helicites = 0, 1, and an extra pair
of operators ak,s , ak,s for the scalar degree of freedom. The commutation rules are
[ak, , ak , ]

= [ak, , ak , ] = 0,

[ak,, ak , ] = k,k g .

(7.563)

where runs through the helicities 0, 1 and s and the metric g has a negative
signature for = 0, 1 and a positive one for = s. These commutation rules have
the same form as those in (7.394), except for the different meaning of the labels .
The properly quantized field is then
1V (x) =

X X
k

7.5.2

=0,1

h
i
1
eikx (k, )ak, + (k, )ak,
2V k 0

h
i
1
+
eikx (k, s)ak,s + (k, s)ak,s
2V k 0

(7.564)

Energy of Massive Vector Particle

The energy can be calculated in close analogy with that of photons. We shall consider
a real field, and admit an additional coupling to an external current density j (x)
to the system, with an interaction
Lint = j (x)V (x).

(7.565)

550

7 Quantization of Relativistic Free Fields

This causes only a minor additional labor but will be useful to understand a special
feature of the massive propagator to be calculated in the next section. The current
term changes the equations of motion (7.546) to
F + M 2 V j = 0,

(7.566)

implying for the non-dynamical zeroth component of V (x) the relation


V 0 (x) =

i
1 h
2 0
0

V(x)
+

V
(x)

j
(x)
.
M2

(7.567)

rather than (4.339). As in electromagnetism, we rename the negative of the canonical


field momenta (7.554) the electric field strength of the massive vector field:

V 0 (x).
E(x) (x) = V(x)

(7.568)

Combining this with (7.567), we obtain


1

V(x)
= E(x) V 0 (x) = E(x) + 2 [ E(x) j 0 (x)].
M

(7.569)

We now form the Hamiltonian density (7.330) as the Legendre transform


H=

3
X
i=1

i V i L Lint .

(7.570)

Inserting (7.569) we obtain


H = H0 + Hint ,

(7.571)

with a free Hamiltonian density


H=

1
1 2
E + B2 + 2 ( E) + M 2 V2 ,
2
M


(7.572)

and an interacting one


Hint = j V

1 0
1
2
j E + 2 j0 .
2
M
M

(7.573)

We have introduced
B(x) = V(x)

(7.574)

in analogy with the magnetic field of electromagnetism [compare (7.330)]


Discarding again the external current, we form the free Hamiltonian H0 =
R 3
V 0 , and E as M 2 V 0 , and find the Hamild x H0 , express E(x) as V
tonian
H=

dx

i
1h
V V V V M 2 V V (E V 0 ) .
2


(7.575)

551

7.5 Massive Vector Bosons

The last term is a surface term that can be omitted in an infinite volume. Inserting
the expansion (7.559) for the field operator, and proceeding as in the case of the
photon energy (7.346), we obtain the energy
H=

k 0 ak, ak, +

k,=0,1

1
2

(7.576)

This contains a vacuum energy


E0 h0|H|0i =

1 X
3X
k0 =
k ,
2 k,=01
2 k

(7.577)

due to the zero-point oscillations, the factor 3 counting the different polarization
states.

7.5.3

Propagator of Massive Vector Particle

Let us now calculate the propagator of the massive vector particle from the vacuum
expectation value of the time-ordered product of two vector fields. As in (7.350) we
find
1 X 1 ik(xx ) X
G (x, x ) = (x0 x0 )
e
(k, ) (k, )
2V k k 0

X
X 1
1

eik(xx )
(k, ) (k, ).
(7.578)
+ (x0 x0 )
2V k k 0

and insert the completeness relation (4.348) to write


kk
1 X 1 ik(xx )

g
+
e
G (x, x ) = (x0 x0 )
2V k k 0
M2
+

(x0

1 X 1 ik(xx )
e
x0 )
2V k k 0

kk
+
M2

!
!

(7.579)

This expression has precisely the general form (7.272) with the polarization sums
of particles and antiparticles being equal to one another. The common polarization
sum can be pulled in front of the Heaviside functions yielding
G (x, x )

(7.580)

= g

M2

(x0 x0 )G(+) (x x ) + (x0 x0 )G() (x x )

1
{[ , (x0 x0 )]G(+) (x x ) + [ , (x0 x0 )]G() (x x )}.
2
M
The bracket in the first expression is equal to the Feynman propagator of the scalar
field. The second term can only contribute for , = 0, 0. For = i and = j, it
vanishes trivially. For = 0 and = i, it gives
+

1 i 0
[ (x x0 )G(+) (x x ) i (x0 x0 )G() (x x )].
M2

552

7 Quantization of Relativistic Free Fields

Since
(x0 x0 )G(+) (x x ) = (x0 x0 )G() (x x ),
this vanishes. For = = 0, on the other hand, we write
[0 0 , (x0 x0 )] = 0 (x0 x0 ) + 2(x0 x0 )0 ,
[0 0 , (x0 x0 )] = 0 (x0 x0 ) 2(x0 x0 )0 ,
and
i
0 G() (x x ) = (3) (x x )
2

(7.581)

to obtain from the last terms 2i (4) (x x ). By a calculation like (7.43) we see
that the first terms remove from this the factor 2. Thus we find for the Feynman
propagator of the massive vector field

G (x, x ) =

kk
i
d4 k ik(xx )

e
(2)4
k 2 M 2 + i
M2

i
0 0 (4) (x x ).
M2

(7.582)

Only the first term is a Lorentz tensor. The second term destroys the covariance
and is called a Schwinger term.
Let us trace its origin in momentum space. If we replace the Heaviside functions
in (7.578) by the Fourier integrals (7.62), we find for the propagator a representation
of the type (7.279):

G (x x ) =

P (k , k)
P (k , k) ip(xx )
d4 k i
e
,

(2)4 2p k 0 k + i k 0 + k i
"

(7.583)

with

P (k , k) P (k



, k)

k 0 =k

k k

M2

(7.584)

k 0 =k

Expression (7.583) is a tensor only if the on-shell energy in the argument of P (k) is
replaced by the off-shell integration variable k 0 . The difference between (7.584) and
such a covariant version has two contributions: one from the linear term in k 0 , and
one from the quadratic term. The first vanishes by the same mechanism as in the
Dirac discussion of (7.279). The quadratic term contributes only for (, ) = (0, 0),
where
2
k2 k 0
00
00
P (k , k) P (k) =
.
(7.585)
M2
2

The combination k2 k 0 = M 2 k 2 cancels the denominators in (7.583), thus


producing precisely the -function term (7.582).

553

7.5 Massive Vector Bosons

The covariant part of the propagator coincides with the Green function of the
field equation (7.547). Indeed, the differential operator on the left-hand side can be
written in analogy with (7.524), as
L (i)G (x x ) = i (4) (x x ),

(7.586)

so that the calculation of the Green function requires inverting the matrix in momentum space [the analog of (7.524)]
L (k) = (k 2 M 2 )PT (k) M 2 PL (k).

(7.587)

This has inverse


1
1
k k
1

P
(k)

P
(k)
=
g

L1 (k) = 2
k M2 T
M2 L
k2 M 2
M2

(7.588)

yielding the solution of (7.586) in momentum space:


G (k) = iL1 (k).

(7.589)

This is precisely the Fourier content in the first term of (7.582). The mass contains
an infinitesimal i to ensure that particles and antiparticles decay both at positive
infinite times.
For a massive vector meson with a gauge fixing term in the action [see (7.550)],
the matrix (7.587) becomes
k2
L (k) = (k 2 M 2 )PT (k) +
M 2 PL (k),

(7.590)

and the inverse (7.588):


1

P
(k)
+
PL (k)
T
2
2
2
2
k M "
k M
#
kk
1

g (1 ) 2
,
= 2
k M2
k M 2

L1 (k) =

(7.591)

Due to the Schwinger term, the propagator is different from the Green function
associated with the field equation (7.547). In Chapters 9, 10, and 13 we shall see
that when interactions are present, there are two ways of evaluating the physical
consequences. One of them is based on the interaction picture of quantum mechanics
and leads to a Schwinger-Dyson perturbation expansion for scattering amplitudes,
the other makes use of functional integrals and leads to a similar expansion. The
first is based on a Hamiltonian approach in which the propagator plays a central
role and interactions are described by an interaction Hamiltonian, in the second the
action is relevant and covariant Green functions replace the propagators. The results
are of course the same in both cases, as we shall see. The cancelation of all effects

554

7 Quantization of Relativistic Free Fields

of the Schwinger term is caused by the last term in the interaction Hamiltonian
density derived in Eq. (7.573).
The commutator of two vector fields is
[V (x), V (x )] = C (x x ),

(7.592)

with
C (x x ) =

d4 k ik(xx )
k k
0
2
2

e
2(k
)(k

M
)
g
+
(2)4
M2
!

C(x x )..
(7.593)
g
M2
!

At equal times, we rewrite in the integrand


(k 2 M 2 ) =

1
[(k 0 k ) + (k 0 + k )],
2k

with k k2 + M 2 and see that because of the oddness of (k 0 ), the commutator


vanishes, except for (, ) = (0, i), where k 0 (k 0 ) = k . produces an even integrand
leading to the initial canonical commutation relation (7.556).
Let us end this section by justifying the earlier-used i-prescription in the propagator (7.529) of the photon field. For this we add a mass term to the electromagnetic
Lagrangian (7.366) with the gauge-fixing Lagrangian, which leads to a massive vector photon Lagrangian
1

1
L = F F + M 2 A A D(x) A (x) + D 2 (x),
4
2
2

0.

(7.594)

The Euler-Lagrange equations are now [compare (7.371)]




( 2 M 2 )g + 1

1
A (x) = 0,

(7.595)

so that the propagator in momentum space obeys the equation




(k 2 M 2 )g 1

1
k k G (k) = i .

(7.596)

The matrix on the left-hand side is decomposed into transverse and longitudinal
parts as
L (k) = (k 2 M 2 )PT (k) +


1 2
k M 2 PL (k),

(7.597)

from which we obtain the inverse

P
(k)
+
P (k),
T
2 L
k2 M 2
k2
M
!
1
k k

k k

= 2
g

+
,
k M2
k2
k 2 M 2 k 2

L1 (k) =

(7.598)

555

7.5 Massive Vector Bosons

which can be rearranged to


L1 (k) =

kk
g
+
(

1)
,
k2 M 2
(k 2 M 2 )(k 2 M 2 )

(7.599)

so that (7.537) is solved by the matrix


G (k) = iL1 (k).

(7.600)

The condition > 0 in the Lagrangian (7.594) ensures that both poles in the second
denominator lie at a physical mass square.
Adding to the mass an infinitesimal imaginary parts i with > 0, and letting
M 0 we obtain precisely the i-prescription of Eq. (7.529).

7.5.4

Behavior under Discrete Symmetry Transformations

Under the discrete transformations P, T , and C, a massive vector field transforms in


the same way as the vector potential of electromagnetism in Eqs. (7.538)(7.542),
except for different possible phases P , T , C .
In the second-quantized Hilbert space, these amount to the following transformation laws for the creation and annihilation operators of particles of helicity .
Under parity:

Pak, P 1 a k, = P ak, ,

(7.601)

with P = 1. Under time reversal:


T ak,T

a k, = T ak, ,

(7.602)

with an arbitrary phase factor T , and under charge conjugation

Cak, C 1 a k, = C ak, ,

(7.603)

with C = 1.
Vector mesons such as the -meson of mass m 759 MeV and the -meson of
mass m 782 MeV transform with the three phase factors P , T , C being equal
to those of a photon. This is a prerequisite for the theory of vector meson dominance
of electromagnetic interactions which approximate all electromagnetic interactions
of strongly interacting particles by assuming the photon to become a mixture of a
neutral - and an -meson, which then undergo strong interactions.
We omit the calculation of the Hamilton operator since the result is an obvious
generalization of the previous expression (7.576) for neutral vector bosons:
H=

X
p

p0

X 

=1,0

ap, ap, bp, bp, =

X
p

p0

s3 =1,0

ap,s3 ap,s3 bp,s3 bp,s3 .


(7.604)

556

7 Quantization of Relativistic Free Fields

7.6

Spin-3/2 Fields

The Rarita-Schwinger field of massive spin-3/2 particles is expanded just like (7.213)
as follows:
(x) =

p,S3

1
q

V p0 /M

eipx u (p, S3 )ap,S3 + eipx v (p, S3 )b (p, S3 ) .

(7.605)

where the spinors u (p, S3 ) and v (p, S3 ) are solutions of equations analogous to
(4.657):
L (p)u (p, S3 ) = 0,
L (p)v (p, S3 ) = 0,

(7.606)
(7.607)

where L (p) is the Fourier-transformed


differential operator (4.927) with positive
2
0
2
mass shell energy p = p + M . The Rarita-Schwinger spinors u (p, S3 ) and
v (p, S3 ) can be constructed explicitly from those of spin 1/2 and the polarization
vectors (4.357) of spin 1 with the help of the Clebsch-Gordan coefficients on Table
4.2:
u (p, S3 ) =

h 23 , S3 |1, s3 ; 21 , s3 i (p, s3 )u(p, s3 ),

(7.608)

h 32 , S3 |1, s3 ; 21 , s3 i (p, s3 )v(p, s3 ).

(7.609)

s3 ,s3

v (p, S3 ) =

s3 ,s3

The calculation of the propagator


S (x, x ) = h0|T (x) (x )|0i

(7.610)

is now somewhat involved. As in the case of spin-zero, spin-1/2, and massive spin-1
fields, the result can be derived most directly by inverting the matrix L (p). For
this we make use of the fact that for reasons of covariance, L1 may be expanded
into the tensors g , , p , p , p p with coefficients of the form A + B p/ . By
multiplying this expansion with L (p) and requiring the result to be equal to ,
we find L1 is equal to the 4 4 spinor matrix
1

p/ + M
1
1
2 p p
=
(7.611)
g + +
( p p ) +
2
2
p M
3
3M
3 M
)
1
w+1

[(4w + 2) p + (w + 1)/
p + 2wp 2wM ] .
6M 2 (2w + 1)2


For simplicity, we have given this matrix only for the case w = real. In analogy with
(7.278), the propagator is equal to the Fourier transform of L1 (p):
S (x, x ) =

d4 p 1
ip(xx )
L
(p)e
.

(2)4

(7.612)

557

7.6 Spin-3/2 Fields

The wave equation has a nontrivial solution where L1 (p) has a singularity. From
(7.611) we see that this happens only on the positive- and negative-energy mass
shells at
q

p = p2 + M 2 .

(7.613)

By writing the prefactor in (7.611) as

1
p/ + M
=
,
2
2
p M
p/ M

(7.614)

we see that only solutions of

(/
p M) (p) = 0,

p0 = p2 + M 2

(7.615)

can cause this singularity. These are the spinors u (p, S3 ) and v (p, S3 ), respectively. The residue matrix accompanying the singularity is independent of the parameter c. It must have the property of projecting precisely into the subspace of
(p) in which the wave equation L (p) (p) = 0 is satisfied. Let
L1 (p) =

p2

2M
P (p) + regular piece.
M2

(7.616)

After a little algebra we find


P (p) =

p/ + M
1
1
2 p p
g + +
(p p ) +
.
2M
3
3M
3 M2


(7.617)

The residue matrix


P (p) is uniquely fixed only on for p0 on the upper and lower
mass shells p0 = p2 + M 2 . There it has the form P (p), respectively. The
expression (7.617) is a common covariant off-shell extension of these two matrices.
These matrices and their extension P (p) are independent of the parameter c.
They are projection matrices, satisfying
P g P = P .

(7.618)

It is easy to verify that for p2 = M 2 , the matrix P (p) satisfies the same
equations as the Rarita-Schwinger spinors u (p, S3 )
(/
p M)P (p) = 0,
P (p)p = 0,
P (p) = 0.

(7.619)
(7.620)
(7.621)

For this reason, P (p) with p0 = p can be expanded as


P (p) =

u (p, S3 )
u (p, S3 ),

(7.622)

s3

P (p) =

X
s3

v (p, S3 )
v (p, S3 ).

(7.623)

558

7 Quantization of Relativistic Free Fields

The projection matrix is equal to the polarization sums of the spinors u (p, S3 ) and
v (p, S3 ), except for a minus sign accounting for the fact that v(p, S3 )v(p, S3 ) =

u(p, S3 )u(p, S3), just as in the Dirac case [recall the discussion after Eq. (4.699)].
The behavior under discrete symmetry transformations of the creation and annihilation operators of a Rarita-Schwinger field is the same as that of a product of
a Dirac operator and a massive vector meson operator, i.e., it is given by a combination of the transformation laws (7.290), (7.312), (7.295) and (7.601), (7.602), and
(7.603).

7.7

Graviton

In order to quantize the gravitational field we first have to rewrite the action in
terms of squares of first derivative terms. After a partial integration, it reads
A=

i
1 Z 4 h
d x h h 2 h h + 2 h h h h ,
8

(7.624)

where h h . A few further partial integrations brings this to the alternative form
A=

1Z 4
d x L(x) =
d x h ,
2
4

(7.625)

where

1
[ h h + g h g h + g h g h ] + ( )
8
(7.626)

is defined by
(x)

L(x)
.
[ h (x)]

(7.627)

It is antisymmetric in and symmetric in . The components 0 play the role


of the canonical field momenta. Similar to the electromagnetic case, four of the
six independent components 0ij vanish, as a consequence of the invariance of the
action under gauge transformations (4.388):
h (x) h (x) + (x) + (x).

(7.628)

Going over to the field (x) of Eq. (4.410) and the Hilbert gauge (4.410), we expand
the field into plane waves eikx with k 0 = |k| and two transverse polarization tensors
(4.410):
(x) =

i
X X h
1

ikx

ikx
,
a

(k,
)e
+
a

(k,
)e
k,
k,
2V p0 k =2

(7.629)

The behavior under discrete symmetry transformations of the creation and annihilation operators of a free graviton field is the same as that of a product of two
photon operators in Eqs. (7.538)(7.540).

559

7.8 Spin-Statistics Theorem

7.8

Spin-Statistics Theorem

When quantizing Klein-Gordon and Dirac fields we directly used commutation rules
for spin 0 and anticommutation rules for spin 1/2. In nonrelativistic quantum field
theory this procedure was dictated by experimental facts. It is one of the important
successes of relativistic quantum field theory that this connection between spin and
statistics is a necessity if one wants to quantize free fields canonically. Let us first
look at the real scalar field theory, and consider the field commutator expanded as
in (7.11):
i[ 0 (x, t), (x , t)]
=

p,p

(7.630)
1

2V p0 2V p0

ei(pxp x )) p0 [ap , ap ] + ei(pxp x )) (p0 )[ap , ap ] ,

where we have left out terms with vanishing commutators [ap , ap ] and [ap , ap ].
Inserting the commutators (7.15), we see that both contributions just add up correctly to give a -function, as they should to fulfill the canonical field commutation
rules (7.2)(7.4). Suppose now for a moment that the particles were to obey Fermi
statistics. Then the symbols [ , ] would stand for anticommutators, and the
two contributions would have to be subtracted from each other and give zero. Thus
a real relativistic scalar field cannot be quantized according to the wrong Fermi
statistics. The situation is similar for a complex scalar field, where we obtain
i[ 0 (x, t), (x , t)]
(7.631)


X
1


q
ei(pxp x ) p0 [ap , ap ] + ei(pxp x ) (p0 )[bp , bp ] ,
=

2V p0 2V p0
p,p
where all those commutators have been omitted which vanish. Again, the two terms
add up correctly to give a -function while the use of anticommutators would lead
to a vanishing of the right-hand side.
These observations are related to another remark made earlier when we expanded
the free complex field into the solutions of the wave equation and wrote down immediately creation operators bp for the negative energy solutions eipx as a generalization
of ap for the real field. At that place we could, in principle, have had the option of
using bp . But looking at (7.631) we realize that this would not have led to canonical
commutation rules in either statistics.
Consider now the case of spin-1/2 fields. Here we calculate the anticommutator
expanded via (7.212):
n

(x, t), (x , t) =

p,p

1
q

V p0 /M V p0 /M

s3 ,s3 =1/2

ei(pxp x ) ap,s3 , ap ,s u(p, s3 )u (p s3 )


i(pxp x )

+e

bp,s3 , bp,s3

v(p, s3 )v (p , s3 ) .

(7.632)

560

7 Quantization of Relativistic Free Fields

We now use the anticommutation rules (7.217) and (7.218)


n

ap,s3 , ap ,s

bp,s3 , bp ,s3

= p,p s3 ,s3 ,
= p,p s3 ,s3 ,

(7.633)

and the polarization sums (4.696) (4.697),


p/ + M
0 ,
2M
s3
X
p/ M
0 ,
v(p, s3 )v (p, s3 ) =
2M
s3
X

u(p, s3 )u (p, s3 ) =

(7.634)

to find
n

(x, t), (x , t) =

X
p

1
0
V p /M

/
ip(xx ) p

+M
/ M
p
0 + eip(xx )
0 . (7.635)
2M
2M

The -function arises after a cancellation of the terms M and p in the numerator
from the sum of the p0 0 -terms.
Let us see what would happen if we use commutators for quantization and thus
the wrong particle statistics. Then the [bp,s3 , bp,s3 ] -term would change its sign and
fail to produce the desired -function. At first sight, one might want to correct
this by using in the expansion of (x, t) annihilation operators dp,s3 rather than a
creators bp,s3 . Then the commutator [bp,s3 , bp,s3 ] would be replaced by [dp,s3 , dp,s3 ],
and give indeed a correct sign after all!
However, the problem would then appear at a different place. When calculating
the energy, we found in (7.604)
H=

p,s3

p0 ap,s3 ap,s3 bp,s3 bp,s3 ,

(7.636)

independent of statistics. Changing bp,s3 to dp,s3 in the field expansion would give,
instead,


X
(7.637)
H=
p0 ap,s3 ap,s3 dp,s3 dp,s3 .
p,s3

But this is an operator whose eigenvalues can take arbitrarily large negative values on
states containing a large number of creation operators dp,s3 in front of the vacuum
state |0i. Such an energy is unphyical since it would imply the existence of a
perpetuum mobile.
The spin-statistics relation can be extended in a straightforward way to vector
and tensor fields, and further to fields of any spin s, For each of these fields, quantization with the wrong statistics would imply either a failure of locality or of the
positivity of the energy.

561

7.8 Spin-Statistics Theorem

We shall sketch the general proof only for the spinor field (x) introduced in
Section 4.19. This is expanded into plane waves (4.19) as before, assigning creation
operators of antiparticles to the Fourier components of the negative-energy solutions:
(x) =

X
p

i
h
1
ipx
ipx c

e
w(p,
s
)a
+
e
w
(p,
s
)
b
3
p,s
3
3
p,s3 .
2V p0

(7.638)

Here
(s)

w c (p, s3 ) = c{2s} w (p, s3 ) = w(p, s3 )cs ,s3 = w(p, s3 )(1)ss3

(7.639)

{2s}

are the charge-conjugate spinors. The matrices c{2s} have the labels cn1 ,n2 ; n ,n , and
1

(s)

are equivalent to the matrices cs3 ,s reversing the spin direction by a rotation around
3
the y-axis by an angle [recall the defining equation (4.872)]. The relation between
their indices n1 , n2 and s3 is the same as in the rest spinors (4.910). These matrices
have the property
2

c{2s} c{2s} = 1,
c{2s} = (1)2s .
(7.640)
The commutation or anticommutation relation between two such spinor fields
has the form

i[(x, t), (x , t )] =

X
p

"

X
1
ip(xx )
e
w(p, s3 )w (p, s3 )
0
2V p
s3
ip(xx )

w (p, s3 )w (p, s3 ) ,

s3

(7.641)

Recalling the polarization sum (4.912),


P (p)

{2s}

(p, s3 )w

s3

{2s}

p
(p, s3 ) =
M


{2s}

(7.642)

and noting that the charge-conjugate spinors have the same sum, this becomes
i[(x, t), (x , t )] =

X
p

i
h
1
ip(xx )
ip(xx )
.
P
(p)
e

e
2V p0

(7.643)

The polarization sum P (p) is a homogenous polynomial in p0 and p of degree 2s.


It has therefore the symmetry
P (p) = (1)2s P (p). The energy p0 lies
2 property
2
on the mass shell p = p + M . Replacing all even powers p2n by (p2 + M 2 )n ,
and all odd powers p2n+1 by p (p2 + M 2 )n , we obtain for P (p) the following generic
dependence on the spatial momenta
P (p) = P0 (p) + p P1 (p),

(7.644)

where P0 (p) and P1 (p) are polynomials of p with the reflection properties
P0 (p) = (1)2s P0 (p),

P1 (p) = (1)2s P1 (p).

(7.645)

562

7 Quantization of Relativistic Free Fields

Thus we can write


i[(x, t), (x , t )]
(7.646)
o
h
i
i
X 1 nh
0
ip(xx )
0
ip(xx )
2s
=
.
P
(p)

p
P
(p)
e
P
(p)
+
p
P
(p)
e

(1)
0
1
0
1
0
p 2V p
After replacing p by spatial derivatives i, the momentum-dependent factors
can be taken outside the momentum sums, and we obtain
i
1 h ip(xx )
2s ip(xx )
e

(1)
e
0
p 2V p
i
X 1 h

+ P1 (i)
eip(xx ) (1)2s eip(xx ) . (7.647)
p 2V

i[(x, t), (x , t )] = P0 (i)

In the limit of infinite volume, the right-hand side becomes [recall (7.46)]
h

P0 (i) G(+) (x x ) (1)2s G(+) (x x )


h

+ P1 (i) G(+) (x x ) (1)2s G(+) (x x ) .


At equal times, this reduces to

P0 (i)[1 (1)2s ] G(+) (x x , 0)


+P1 (i)[1 (1)2s ] (3) (x x ).
Locality requires that the commutator at spacelike distances vanishes.
G(+) (x x , 0) 6= 0, the prefactor of the first term must be zero, and hence
2s

(1) =

1 for bosonic commutators,


.
1 for fermionic anticommutators.

(7.648)

(7.649)
Since

(7.650)

This proves the spin-statistics relation for spin-s spinor fields (x).
Note that the locality of the free fields always forces particles and antiparticles
to have the same mass and spin.

7.9

CPT-Theorem

All of the above local free-field theories have a universal property under the discrete symmetry operations C, P , and T . When being subjected to a product
CP T of the three operations, the Lagrangian is invariant. We shall see later
in Chapter 26 that this remains true also in the presence of interactions that
violate CP - or P -invariance, as long as these are local and Lorentz-invariant.
@@@@@@@@@@@@@@ When introducing local interactions between local fields
they will always have the property of being invariant under the operation CP T .
This is the content of the so-called CPT-theorem. This property guarantees that
the equality of masses and spins of particles and antiparticles remains true also in
the presence of interactions. If there is a mass difference in nature, it must be extremely small. Direct measurements of masses of electrons and protons and their
antiparticles are too insensitive to detect a difference with present experiments.

563

7.10 Physical Consequences of Vacuum Fluctuations. Casimir Effect

7.10

Physical Consequences of Vacuum Fluctuations


Casimir Effect

For each of the above relativistic quantum fields we have found an infinite vacuum
energy due to the vacuum fluctuations of the field. For a real scalar field, this energy
is [see (7.32)]
1X 0 1X
h
c X q 2
E0 =
k =
k =
k + c2 M 2 /h2 ,
(7.651)
2 k
2 k
2 k

where k = p/h are the wave vectors. In many calculations, this energy is irrelevant
and discarded. There are, however, important physical phenomena where this is
impossible, since their origin lies in the vacuum energy (7.651). One of them is the
cosmological constant as discussed on p. 467. The other appears in electromagnetism
and is observable in the form of van der Waals forces between different dielectric
media.
In the present context of free fields in a vacuum, the most interesting phenomenon
is the Casimir effect. Vacuum fluctuations of electromagnetic fields cause an attraction between two parallel closely spaced silver plates in an otherwise empty space.
The basic electromagnetic property of a silver plate is to enforce the vanishing of
the parallel electric field at its surface, due to the high conductivity. Let us study
a specific situation and suppose the silver plates to be parallel to the xy-plane one
at z = 0 and the other at z = d. To have discrete momenta the whole system is
imagined to be enclosed in a very large conducting box
x (Lx /2, Lx /2),

y (Ly /2, Ly /2),

, z (Lz /2, Lz /2).

(7.652)

The plates divide the box into a thin slice of volume Lx Ly d and two large pieces
of volumes Ly Ly Lz /2 below and Lx Ly (Lz /2 d) above the slice (see Fig. 7.5).
These are three typical resonant cavities of classical electrodynamics. For waves of

Figure 7.5 Geometry of the plates for the calculation of the Casimir effect.

frequency , the Maxwell equations F = 0 read, in natural units with c = 1


[recall (4.244)],
= iE,
B= E
E = 0.
(7.653)

564

7 Quantization of Relativistic Free Fields

Since the field components parallel to the conducting surface vanish, the first equation implies that the same is true for the magnetic field components normal to the
surface. Inserting the equations into each other, and using the other two Maxwell
equations [recall (4.244)] following from the Bianchi identity F = 0 [recall
(4.243)]:
E = iB,

B = 0.

(7.654)

the fields E and B are both seen to satisfy the Helmholtz equations


E(x)
B(x)

E0 (x )
B0 (x )

= 0.

(7.655)

(7.656)

They are solved by the ansatz


(

E(x)
B(x)

eikz z ,

where x (x, y) are the components of x parallel to the plates. The fields
E0 (x ), B0 (x ) satisfy the differential equation


+ kz

E0 (x )
B0 (x )

= 0,

(7.657)

where is the transverse derivative


(x , y , 0).

(7.658)

We now split the original fields E(x) and B(x) into a component parallel and orthogonal to the unit vector z along the z-axis:
E = Ez z + E
B = Bz z + B ,

(7.659)

and see that the Maxwell equations (7.653) can be solved for E , B in terms of
Ez , Bz :
1
[ z Bz + iz Ez ] ,
2 + 2
1
[ z Ez iz Bz ] .
=
2
+ 2

B =
E

(7.660)

Therefore we only have to find Ez , Bz . Let us see what boundary conditions the
fields components satisfy at the conducting surfaces, where we certainly have
E = 0,

Bz = 0.

(7.661)

7.10 Physical Consequences of Vacuum Fluctuations. Casimir Effect

565

Due to (7.660), this implies z Ez |surface = 0. The Maxwell equations admit two
types of standing-wave solutions with these conditions:
1) Transverse magnetic waves
Bz 0,

Ez = E (x ) cos

z
n
d

n = 0, 1, 2, . . . ,

(7.662)

z
n n = 1, 2, 3, . . . ,
d

(7.663)

2) Transverse electric waves


Ez 0,

Bz = B (x ) sin

where E,B (x ) solve the transverse Helmholtz equation




2 + 2 kz 2 E,B (x ) = 0.

(7.664)

If the dimensions of the box along x- and y-axes are much larger that d, the quantization of the transverse momenta becomes irrelevant. Any boundary condition
will yield a nearly continuous set of states with the transverse density of states
Lx Ly /(2)2 . Thus we may record that there are two standing waves for each momentum



n
k = kx , ky , kz = n ,
(7.665)
d
except for the case n = 0 when there is only one wave: the transverse magnetic one.
Using the discrete momenta in the sum (7.651), the energy between the plates may
therefore be written as
Ed =

h
c
Lx Ly
2

dkx dky X q 2
2
kx + ky2 + kzn 2 .
(2)2 kn

(7.666)

We have reinserted the proper fundamental constants h


and c, for convenience. The
sum carries a prime which is supposed to record the presence of an extra factor 12 for
n = 0. In the two semi-infinite regions outside the two plates also the momentum
kz may be taken as a continuous variable so that there are the additional vacuum
energies
Eoutside

h
c
Lz dkx dky dkz q 2
=
Lx Ly
2 kx + ky2 + kz2
2
2
(2)3

Z
dkx dky dkz q 2
Lz
h
c
Lx Ly
d
2 kx + ky2 + kz2 .
+
2
2
(2)3
Z

(7.667)

Let us now compare this energy with the corresponding expression in the absence
of the plates
Z
dkx dky dkz q 2
h
c
2 kx + ky2 + kz2 .
(7.668)
E0 = Lx Ly Lz
2
(2)3

566

7 Quantization of Relativistic Free Fields

By subtracting this from Ed + Eoutside we find the change of the energy due to the
presence of the plates
E = h
cLx Ly

dkx dky X

(2)2 kz =kn
z

dkz d q 2
kx + ky2 + kz2 .

(7.669)

It is caused by the difference in the particle spectra, once with discrete kzn between
the plates and once with continuous kz without plates. The evaluation of (7.669)
proceeds in two steps. First we integrate over the k = (kx , ky ) variables but
2
enforce the convergence by inserting a cutoff function f (k
+ kz2 ) which is identical
2
to unity up to some large k2 = k
+ kz2 , say k2 = 2 . For k2 2 , the cutoff
function decreases rapidly to zero. Later we shall be able to take 2 . The
cutoff function is necessary to perform the mathematical operations. In the physical
system a function of this type is provided by the finite thickness and conductivity
of the plates. This will have the effect that for wavelengths much smaller than the
thickness, the plates become transparent to the electromagnetic waves. Thus they
are no longer able to enforce the boundary conditions which are essential for the
discreteness of the spectrum. Since the energy difference comes mainly from the long
wavelengths, the precise way in which the short wavelengths are cutoff is irrelevant.
We then proceed by considering the remaining function of kz
g(kz )

2
dk

2
2
k
+ kz2 f (k
+ kz2 ),

(7.670)

in terms of which the energy difference is given by

X
1
E = h
cLx Ly 2

4 kz =kzn

dkz d
g(kz ).

(7.671)

To evaluate this expression it is convenient to change the variables from kzn to n =


2
2 2
kz d/ and from k
to 2 = k
d / 2 , and to introduce the auxiliary function
G(n)
=

d3
d3
g(n/d)
=
4
3
Z

n2

2 n2 /d2

dk 2 kf (k 2 )

d 2 f 2 2 /d2 .

Then E can be written as

(7.672)

2 X

E = h
cLx Ly 3
G(n)
dn G(n) .
4d n=0
0
"

(7.673)

The difference between a sum and an integral over the function G(n) is given by the
well-known Euler-Maclaurin formula which reads [see Eq. (7C.20) in Appendix 7C]
Z n
1
1
G(0) + G(1) + G(2) + . . . + G(n 1) + G(n)
dn G(n)
2
2
0

i
X
B2p h (2p1)
G
(n) G(2p1) (0) .
=
p=1 (2p)!

(7.674)

567

7.10 Physical Consequences of Vacuum Fluctuations. Casimir Effect

Here Bp are the Bernoulli numbers, defined by the Taylor expansion [see (7C.15)]

X
tp
t
Bp ,
=
et 1 p=0 p!

(7.675)

whose lowest values are [see (7C.6)]


1
1
1
1
B0 = 1, B1 = , B2 = , B3 = 0, B4 = , B5 = 0, B6 = , . . . . (7.676)
2
6
30
42
The primed sum in (7.666) is therefore obtained from the expansion

n=0

G(n)

B2
[G (0) G ()]
2!
B4
[G (0) G ()]
4!
B6 (5)
[G (0) G(5) ()]
6!
..
.

dn G(n) =

(7.677)

Since the cutoff function f (k2 ) vanishes exponentially fast at infinity, the function
G(n) and all its derivatives vanish in the limit n (so that the factor 1/2 at the
upper end of the primed sum is irrelevant) For n = 0, only one expansion coefficient
turns out to be nonzero. To see this, we denote f ( 2 n2 /d2 ) by f(n2 ) and calculate
G (0) = [2n2 f(n2 )]n=0 = 0,
 i
h
 
G (0) = 4nf n2 + 4n3 f n2

n=0

= 0,

G (0) = 4f n2 + 18n2 f n2 + 8n4 f (n2 )

G(l) (0) = 0,

l > 3.

n=0

= 4,
(7.678)

The higher derivatives contain an increasing number of derivatives of f( 2 ). Since


f( 2 ) starts out being unity up to large arguments, all derivatives vanish at = 0.
Thus we arrive at

n=0

G(n)

dn G(n) = 4

1
B4
=
.
4!
180

(7.679)

This yields the energy difference


E = Lx Ly h
c

2
.
720d3

(7.680)

There is an attractive force between silver plates decreasing with the inverse forth
power of the distance:
2
.
(7.681)
F = Lx Ly h
c
240d4

568

7 Quantization of Relativistic Free Fields

Between steel plates of area 1 cm2 and distance 0.5 m, the force is 0.2 dyne/cm2 .
The existence of this force was verified experimentally [6].
Experiments are often done by bringing a conducting sphere close to a plate.
Then the force (7.682) is modified by a factor 2R d/3 to
F =

2
2R d
h
c
.
3
240d4

(7.682)

In dielectric media, a similar calculation renders the important van der Waals
forces between two plates of different dielectric constants.28
It is interesting to study also the case of the force between an electrically conducting plate and a magnetically conducting plate. The transverse electric modes
have the form (7.662) but with with n running through the half-integer values
n = 12 , 23 , 52 , . . . . These values ensure that the electric field is maximal at the
second plate so that the transverse magnetic field vanishes. The transverse magnetic modes have the form (7.662) with n running through the same half-integer
values. The relevant modification of the Euler-MacLaurin formula is
Z

G( 21 ) + G( 32 ) + . . . G( n2 )
=

n
0

dx G(x)

h
i
1
Bp () G(2p1) (n) G(2p1) (0) .
p=1 p!

(7.683)

It is a special case of the general formula [see Eq. (7C.25) in Appendix 7C]
m1
X
n=0

G(a + nh + h)

dx G(x)

h
i
hp1
=
Bp () G(2p1) (b) G(2p1) (a) .
p=1 p!

(7.684)

where b = a + mh. Here Bp () are the Bernoulli functions defined by the generalization of formula (7.675):

X
tp
tet
B
()
=
,
(7.685)
p
et 1 p=0
p!
They are related to the Bernoulli numbers Bp by
Bn () =

n
X

p=1

n
Bp np .
p

(7.686)

For instance
so that

Bn ( 21 ) = (1 21n )Bn ,

B0 ( 21 ) = 1, B1 ( 12 ) = 0, B2 ( 21 ) = 121 , B3 ( 21 ) = 0, B4 ( 21 ) =
28

(7.687)

7
240

, B5 ( 21 ) = 0, . . . .
(7.688)

See H. Kleinert, Phys. Lett. A 136, 253 (1989), and references therein.

569

7.11 Zeta Function Regularization

For h = 1 and = 21 this reduced to (7.683).


The right-hand side is a modified version of (7.677):

n=1

G( n2 )

B2 ( 21 )
[G (0) G ()]
2!
B4 ( 21 )

[G (0) G ()]
4!
B6 ( 21 ) (5)
[G (0) G(5) ()]

6!
..
.

dx G(x) =

(7.689)

and the previous result (7.679) is replaced by

n=1

Since B4 ( 21 ) =
7/8:

7
240

G( n2 )

dx G(x) = 4

B4 ( 21 )
.
4!

(7.690)

= 78 B4 , the energy difference (7.680) is modified by a factor

7 2
.
(7.691)
8 720d3
The force is now repulsive, and a little weaker than the previous attraction [6].
E = Lx Ly h
c

7.11

Zeta Function Regularization

There exists a more elegant method of evaluating the energy difference (7.669) without using a cutoff function f (k2 ). This method has become popular in recent years
in the context of the field theories of critical phenomena near second-order phase
transitions. The method will be discussed in detail when calculating the properties
of interacting particles in Chapter 11. The Casimir effect presents a good opportunity for a short introduction. We observe that we can rewrite an arbitrary negative
power of a positive quantity a as an integral
z

1
=
(z)

d z a
e .

(7.692)

This follows by a simple rescaling from the integral representation for the Gamma
function29
Z
d z
(z) =
e .
(7.693)

0
which converges for Re z > 0. Then we rewrite the energy difference (7.669) as
E = h
cLx Ly
29

dkx dky X

(2)2 kz =kn
z

dkz d 2
(kx + ky2 + kz2 )z ,

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 8.310.1.

(7.694)

570

7 Quantization of Relativistic Free Fields

and evaluate this exactly for sufficiently large z, where sum and integrals converge.
At the end we continue the result analytically to z = 1/2.
If we dont worry about the convergence at each intermediate step, this procedure
is equivalent to using formula (7.695) for z = 1/2

1 Z d 1/2 a

e ,
a=
( 12 ) 0

(7.695)

and rewriting the energy difference (7.669) as


1
E = h
cLx Ly
( 12 )

d 1/2

dkx dky X

(2)2 kz =kn

dkz d (kx2 +ky2 +kz2 )


e
.

(7.696)

Being Gaussian, the integrals over kx , ky can immediately be done, yielding

Z
1
d 3/2 X Z dkz d kz2
E = h
cLx Ly

e
.
4( 12 ) 0

0
kz =k n

(7.697)

The remaining sum-minus-integral over kz is evaluated as follows. First we write

kz =kzn

dkz d kz2
1
e
=

n=

dn e

2 n2 /d2

(7.698)

Then we make use of the Poisson summation formula (1.204):

m=

2im

n=

( n).

(7.699)

This allows us to express the sum over n as an integral over n, which is restricted
to the original sum with the help of an extra sum over integers m:

1 X
1 X
n2
e
=
2 n=
2 m=

dn en

2 +2inm

(7.700)

The Gaussian integral over n can then be performed after a quadratic completion,
and gives
r

1 X
1 X
42 m2 /
n2
e
.
(7.701)
e
=
2 n=
2 m=
The two sides are said to be the dual transforms of each other. The left-hand side
converges fast for large , the right-hand side does so for small . This duality
transformation is fundamental to the theoretical description of many phase transitions where the partition functions over fundamental excitations can be brought to

571

7.11 Zeta Function Regularization

different forms by such transformations, one converging fast for low temperatures,
the other for high temperatures.30
Now, subtracting from the sum (7.701) over n the integral over dn removes from
the auxiliary sum over m on the right-hand side precisely the m = 0 term, so that
we find the difference between sum and integral:

1
2

n=

n2

dn e

m=1

42 m2 /
e
.

(7.702)

The left-hand side appears directly in (7.698), which therefore becomes

kz =kzn

dkz d kz2
d X
2 2
ed m / .
e
=

m=1

(7.703)

To obtain the energy difference (7.697), this has to be multiplied by 3/2 and
integrated over d / . In the dually transformed integral, convergence at small is
automatic. The integral can be done yielding
Z

d 3/2 d X
d (2) X
1
2 2

ed m / =

.
4

m=1
d m=1 m4

(7.704)

The sum over m involves Riemanns zeta function


(x)

1
,
x
m=1 m

(7.705)

encountered before in statistical mechanics calculations [see (2.276)], and is equal


to [recall (2.316)]
8 4
8 4 1
(4) =
|B4 | =
.
(7.706)
4!
4! 30
With this number, the energy difference (7.697)
becomes, recalling the values of the
Gamma function (2) = 1 and (1/2) = 2 ,
1
E = h
cLx Ly
4( 21 )
= h
cLx Ly

d 3/2 X

kz =k n
z

2
(2)
=
L
L
h

c
.
x y
4 3/2 ( 12 )d3
720d3

dkz d kz2
e

(7.707)

This is the same result as in Eq. (7.680). To be completely satisfactory, these


calculations should all be done for the general expression (7.716), followed by an
analytic continuation of the power z to 1/2 at the end.
30

For a comprehensive discussion see the textbook H. Kleinert, Gauge Fields in Condensed
Matter , Vol. I Superflow and Vortex Lines, pp. 1744, World Scientific, Singapore 1989
(http://www.physik.fu-berlin.de/~kleinert/b1). and Vol. II Stresses and Defects, pp. 7441443, World Scientific, Singapore 1989 (http://www.physik.fu-berlin.de/~kleinert/b2).

572

7 Quantization of Relativistic Free Fields

While being formal, with the above justification via analytic continuation, we
may be mathematically even more sloppy, and process the -integral in the energy
difference (7.697) as a shortcut via (7.698) and write
Z

d 3/2 X

kz =k n
z

dkz d kz2
1
e
=

n=

= ( 32 )

n=

!Z

dn

d 3/2 2 n2 /d2

dn n3 .

(7.708)

The divergent sum-minus-integral over n3 can formally identified with the zeta funcP
z
tion
at the negative argument z = 3. With the help of formula (2.313),
n=1 n
we find
1
B4
=
.
(7.709)
(3) =
4
120
Inserting this together ( 32 ) = 23 ( 12 ) into (7.708), and this further into (27.109)
yields
E = Lx Ly h
c

2
.
720d3

(7.710)

The correctness of this formal approach is ensured by the explicit duality transformation done before.
The formal procedure of evaluating apparently meaningless sum-minus-integrals
over powers n with > 1 using zeta functions of negative arguments is known as
the zeta function regularization method.
We shall see later in Eq. (11.127), that by a similar analytic continuation an
R
integral 0 dk k vanishes for all . For this reason, the zeta-function regularization
P

applies even to the pure sum, allowing us to replace


n=1 n by ().
Let us rederive also the repulsive result (7.691) with the help of the -function
regularization. The energy difference has again the form (7.697), but with the
discrete values of kzn runnung through all n/d with half-integer values of n = 21 , =
1 3
= 25 , . . . . Thus we must calculate (7.698) with the sum over all half-integer
2, 2,
n. Thus can be done with a slight modification of the Poisson summation formula
(1.204)

e2im (1)m =

m=

e2i( 2 )m =

m=

n=

( n 12 ).

(7.711)

As a result we obtain a modified version of (7.704):


Z

d 3/2 d X
d (2) X
(1)m
2 2

ed m / =

m=1
d4 m=1 m4

(7.712)

which contains the the modified zeta function

(x)

(1)m
,
x
m=1 m

(7.713)

573

7.12 Dimensional Regularization

It is now easy to see that

(x)

m=1

"

1
1
x +2
= (1 21x )(x).
x
m
(2m)

so that

(7.714)

1 1 (4) = 7 (4),
(7.715)
(4)
8
8
and we obtain, as before, a sign change and a factor 7/8 with respect to the previous
attractive result (7.707), as in the previous resullt (7.691). obtained from the EulerMacLaurin summation formula.

7.12

Dimensional Regularization

The above results can also be obtained without the duality transformation (7.701)
with by calculating it in an arbitrary dimension D, and continuing everything to the
physical dimension D = 3 at the end. In D spatial dimensions, the energy difference
(7.716) becomes
E = h
cVD1

dD1 k X Z dkz d 2
(kx + ky2 + kz2 )z ,

(2)D1 kz =kn

(7.716)

where is the D 1 -dimensional volume of the plates, the dimensional extension of


Lx Ly . After using the integral formula (7.695), this turns into the D-dimensional
generalization of (7.696):
1
E = h
cVD1
( 21 )

d 1/2

dD1k X

(2)D1 kz =kzn

dkz d (kx2 +ky2 +kz2 )


e
.

(7.717)

Performing now the D 1 Gaussian momentum integrals yields


1
E = h
cVD1
( 12 )

d 1/2 1

4


 D1
2

kz =kzn

dkz d kz2
e
. (7.718)

After integrating over all , this becomes


( D2 ) 1
E = h
cVD1
( 12 ) 4


 D1
2

kz =kzn

dkz d D
kz .

(7.719)

Inserting kz = kzn = n/d the last factor is seen to be equal to

kz =kzn

 D

dkz d D
kz =

X
n

dn nD .

(7.720)

574

7 Quantization of Relativistic Free Fields

We now make use of the heuristic Veltman integral rule, which states in a renomalizable quantum field theory, and any free field theory is renormalizable as will be
shown later, we may always set the following integrals equal to zero:
Z

dk k = 0.

(7.721)

This can be done for any real . The justification of this rule will be given later in
Section 11.5. Then (7.729) can be expressed in terms of the zeta function (7.705) as

kz =kzn

 D

dkz d D
kz =

(D),

(7.722)

and the energy difference (??) becomes


( D2 ) 1
E = h
cVD1
( 12 ) 4


 D
 D1
2

(D).

(7.723)

For D = 3 dimensions, this is equal to


3
( 23 ) 1
(3).
( 12 ) 4
d

Inserting ( 32 ) = 4 /3, (1/2) = 2 , this becomes




E = h
cV2

E = hcV2

 

2
(3).
6d3

(7.724)

(7.725)

The value of (3) cannot be calculated from the definition (7.705) since the sum
diverges. There exists, however, an integral reperesentation for (x) which agrees
with (7.705) for x > 1 where the sum converges, but which can be evaluated also
for x 1. From this one can derive the reflection formula
(z/2) z/2 (z) = ((1 z)/2) (1z)/2 (1 z).

(7.726)

This allows us to calculate


(3) =

3
1
(4) =
,
4
4
120

(7.727)

so that
E = hc V2

2
2 3
(4)
=
V
h

c
,
2
6d3 4 4
720d3

(7.728)

in agreement with Eq. (7.710).


Note that this result via dimensional regularization agrees with what was obtained before with the help of the sloppy treatment in Eq. (7.697). The reason

575

7.12 Dimensional Regularization

for this is the fact that the reflection formula (7.726) is a direct consequence of the
duality transformation in Eq. (7.701) which was the basis for the previous treatment.
The derivation of the factor 7/8 for the energy between the mixed electric and
magnetic conductor plates is more direct in the present approach than before. Here
the expression (7.729) becomes

kz =kzn

 D

dkz d D

kz =

n=1/2, 3/2,...

dn nD ,

(7.729)

and we may express the right-hand side in terms of the Hurwitz zeta function defined
by31
1
x
n=0 (q + n)

 D

as

kz =kzn

Now we use the property32

(x, q)

dkz d D

kz =

(7.730)

(D, 1/2).

(x, q) = (2x 1)(x),

(7.731)

(7.732)

to calculate the right-hand side of (7.731) as


(3, 1/2) = 87 (3).

(7.733)

This exhibit immediately the desired factor 7/8 with respect to (7.728).
P
At a finite temperature, the sum over the vacuum energies 2h k k /2 is modified
by the Bose occuation factor n of Eq. (2.263),
nk =

1
1
1
h
k
= coth
+ /kB T
,
2 e k
2
2kB T
1

(7.734)

and becomes 2h k k nk . How does this modify the vacuum energy, which in
Eq. (7.725) was shown to be equal to
P

E = hcV2

2
(3)?
6d3

(7.735)

For small T , the sum over n3 in (3) is replaced by a sum over 2nD kB T /n. The
subject is discussed in detail in Ref. [6], and the resulting force between a plate end
a sphere of radius R is
(3) RkB T
.
4
d2
This has apparently been observed recently.
F =

31
32

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formula 9.521


ibid., Formula 9.535.1

(7.736)

576

7.13

7 Quantization of Relativistic Free Fields

Accelerated Frame

The vacuum oscillations of a field show unusual properties when they are observed
from an accelerated frame. Such a frame is obtaind by applying successively after
each proper time interval d an infinitesimal boost eidM to a rest frame. The
rapidity increases as a function of the invariant time like ( ) = a /c, where a is
the invariant acceleration. Hence the velocity v = carctanh increases as a function
of like
v( ) = c tanh(a /c)
(7.737)
If the acceleration points into the z-direction, the space and time coordinates evolve
like
z( ) = (c/a) cosh(a /c),

t( ) = (c/a) sinh(a /c).

(7.738)

The velocity seen from the initial Minkowski frame increases therefore like
dv(t)/dt = a(1 v 2 /c2 )3/2

(7.739)

The energy of a particle such transforms into


p0 ( ) = cosh ( ) p0 + sinh ( )p3.

(7.740)

A massless particle has p3 = p0 , where this becomes simply


p0 ( ) = e( ) p0 = ea /c p0 .

(7.741)

The temporal ocillations of the wave function of the particle ei with = cp0 /h
a /c
in an accelrated frame is therefore observed to have a frequency e( ) = ee .
As a conseqeunce, a wave of a single momentum p3 in the rest frame is seen to have
an entire spectrum of freqeuncies. Their distribution can be calculated by a simple
Fourier transformation [7]:
f (, ) =

dt eit eie

at/c c/a

= i

c
a

ic/a

c
,
a


(7.742)

or
c/2a

f (, ) = e

c ic/a log(c/a) c
i
e
,
a
a


f (, ) = ec/2a i

c ic/a log(c/a) c
e
.
a
a


(7.743)

(7.744)

We see that the probability to find the frequency in f (, ) is is |f (, )|2 =


|ec/2a (ic/a)c/a|2 . Since
(ix)(ix) = /x sinh(x),

(7.745)

7.14 Free Green Functions of n Fields

577

we obtain

2c
1
2c
=
n .
(7.746)
2c/a
a e
1
a
The factor n is the thermal Bose-Einstein distribution of the frequencies [recall
Eq. (13.196)] at a temperature TU = h
a/2ckB , the so-called Unruh temperature [8].
The particles in this heat bath can be detected by suitable particle reactions (for
details see Ref. [9]).
Let us describe the situation in terms of the quantum field. For simplicity, we
focus attention upon the relevant dimensions z and t and study the field only at the
origin z = 0, where it has the expansion [compare (7.16)]

Z

dk ch  ik t
(0, t) =
e
ak + eik t ak , k c|k|.
(7.747)
2 2k
|f (, )|2 =

The creation and annihilation operators satisfy the commutation rules

ak , ak

ak , ak = 0,

[ak , ak ] =

= 2k 0 2(k k ).

(7.748)

They can be recovered from the fields by the Fourier transform [compare (7.16)] :
(

ak
ak

= 2ck

dteik t (0, t),

(7.749)

The accelerated field has the form


a (z, t) =


dk 1  ieat/c k t
e
ak + eik t ak .
2 2k

(7.750)

Inverting the Fourier transformation (7.742), we can expand


a (z, t) = c


dK  iK(ctz)
e
AK + eiK(ctz) AK ,
2

(7.751)

where
AK =

dk
2

(7.752)

C onsider the quantum statistics of this field in a thermal environent of temperature T :

7.14

Free Green Functions of n Fields

For free fields of spin 0 and 1/2 quantized in the preceding sections we have shown
that the vacuum expectation value of the time-ordered product of two field operators, the propagator, is equal to the Green function associated with the differential

578

7 Quantization of Relativistic Free Fields

equation obeyed by the field. In the case of the real Klein-Gordon field we had, for
example,
G(x1 , x2 ) = G(x1 x2 ) h0|T (x1 )(x2 )|0i

(7.753)

( 2 M 2 )G(x, x2 ) = i (4) (x x2 ).

(7.754)

satisfying

It is useful to define a generalization of the vacuum expectation values (7.753) to


any number n of free fields:
G(n) (x1 , x2 , x3 , . . . , xn ) h0|T (x1 ) . . . (xn )|0i.

(7.755)

This object will be referred to as the free n-point propagator or n-point function, or
For massive vector fields and other massive fields with spin larger than 1/2, one must
watch out for the discrepancy between propagator and Green function by Schwinger
terms. For scalar fields, there is no problem of this kind. For complex scalar fields,
the appropriate definition is
G(n,m) (x1 , . . . , xn ; x1 , . . . , xm ) h0|T (x1 ) (xn ) (x1 ) (xm )|0i.

(7.756)

For free fields, the latter is nonzero only for n = m. Corresponding expressions
are used for fields of any spin. We shall see later that the set of all free n-point
functions plays a crucial role in extracting the information contained in an arbitrary
interacting field theory.
The n-point functions may be considered as the relativistic generalizations of the
set of all many-particle Schrodinger wave functions in the nonrelativistic theory. In
the local x-basis, these would consist in the set of all scalar products
hx1 , . . . , xn |(t)i = h0|(x1 , 0) (xn , 0)|(t)i.

(7.757)

Choosing for the wave vectors |(t)i any state in the local n-particle basis
|x1 , . . . , xn ; ti = (x1 , t) (xn , t)|0i,

(7.758)

we arrive at the scalar products


h0|(x, 0) (xn , 0) (x1 , t) (xn , t)|0i.

(7.759)

These, in turn, are a special sort of propagator,


G(n,m) (x1 , 0, x2, 0, . . . , xn , 0; x1, t, . . . , xm , t),

(7.760)

where only two time arguments appear. When going to relativistic theories, the
set of such equal-time objects is no longer invariant under Lorentz transformations.
Two events which happen simultaneously in one frame do not in another frame.
This is why the propagators G(n,m) (x, . . . , xn ; x1 , . . . , xm ) with arbitrary spacetime

7.14 Free Green Functions of n Fields

579

arguments represent the relativistic generalization of the local wave functions. This
argument does not yet justify the time ordering in the relativistic propagators in
contrast to the relevance of the causal propagator in the nonrelativistic case. Its advantage will become apparent only later in the development of perturbation theory.
At the present level where we are dealing only with free quantum fields, the
n-point functions are not really independent objects. They can all be expanded into
a sum of products of the simple propagators. Before we develop the main expansion
formula it is useful to recall the definition of a normal product introduced in Subsecs.
7.1.5 and 7.3.2 for Bose and Fermi fields, respectively. There it was contrived as
a trick to eliminate a bothersome infinite energy of the vacuum to the zero-point
oscillations of the free fields, to circumvent a proper physical discussion. Here we
need this product for purely mathematical reasons to help us evaluating vacuum
expectation values of products of many field operators.
Given an arbitrary set of n free field operators 1 (x1 ) n (xn ), each of them
consists of a creation and an annihilation part:
i (xi ) = ci (xi ) + ai (xi ).

(7.761)

Some i may be commuting Bose fields, some anticommuting Fermi fields. The
normally ordered product or normal product of n of these field operators will be
1 (x)(x2 ) . . . (xn )). In the present context, a function symbol will
denoted by N(
be more convenient than the earlier double-dot notation. The normal product is a
function of a product of field operators which has the following two properties:
i) Linearity: The normal product is a linear function of all its n arguments, i.e.,
it satisfies
((1 + )2 3 n ) = N
(1 2 3 n ) + N(
2 3 n ). (7.762)
N
1
1
If every i is replaced by ci + ai , it can be expanded into a linear combination
of terms which are all pure products of creation and annihilation operators.
ii) Normal Reordering: The normal product reorders all products arising from
a complete linear expansion of all fields according to i) in such a way that
all annihilators stand to the right of all creators. If the operators i describe
fermions, the definition requires a factor 1 to be inserted for every transmutation of the order of two operators.
For example, let 1 , 2 , 3 be scalar fields, then normal ordering produces for two
field operators
c c )
N(
1 2
c a )
N(
1 2

N(a1 c2 )
(a a )
N
1 2

=
=
=
=

c1 c2 = c2 c1 ,
c1 a2 ,
c2 a1 ,
a1 a2 = a2 a1 ,

(7.763)

580

7 Quantization of Relativistic Free Fields

and for three field operators


c c a ) = c c a = c c a ,
N(
1 2 3
1 2 3
2 1 3
c a c
c c a

N(1 2 3 ) = 1 3 2 = c3 c1 a2 ,
a c c ) = c c a = c c a .
N(
1 2 3
2 3 1
3 2 1

(7.764)

If the operators i are fermions, the effect is


(c c )
N
1 2
c a

N (1 2 )
(a1 c2 )
N
(a a )
N
1 2

=
=

c1 c2 = c2 c1 ,
c1 a2 ,
c2 a1 ,
a1 a2 = a2 a1 ,

(7.765)

and
(c c a ) = c c a = c c a ,
N
1 2 3
1 2 3
2 1 3
c a c
c c a

N (1 2 3 ) = 1 3 2 = c3 c1 a2 ,
(a1 c2 c3 ) = c2 c3 a1 = c3 c2 a1 .
N

(7.766)

The normal product is uniquely defined. The remaining order of creation or annihilation parts among themselves is irrelevant, since these commute or anticommute
with each other by virtue of the canonical free-field commutation rules. In the following, the fields may be Bose or Fermi fields and the sign of the Fermi case will
be recorded underneath the Bose sign.
The advantage of defining normal products is their important property that
they have no vacuum expectation values. There is always an annihilator on the
right-hand side or a creator on the left-hand side which produce 0 when matched
between vacuum states. The method of calculating all n-point functions will consist
in expanding all time ordered products of n field operators completely into normal
products. Then only the terms with no operators will survive between vacuum
states. This will be the desired value of the n-point function.
Let us see how this works for the simplest case of a time-ordered product of two
identical field operators
T ((x1 )(x2 )) (x01 x02 )(x1 )(x2 ) (x02 x01 )(x2 )(x1 ).

(7.767)

The basic expansion formula is


((x1 )(x2 )) + h0|T((x1 )(x2 ))|0i.
T ((x1 )(x2 )) = N

(7.768)

For brevity, we shall denote the propagator of two fields as follows:


h0|T((x1 )(x2 ))|0i = (x1 )(x2 ) = G(x1 x2 ).

(7.769)

The two dots which mentally connect the two fields are referred to as a contraction
of the fields.

7.14 Free Green Functions of n Fields

581

We shall prove the basic expansion formula (7.768) by considering it separately


for the creation and annihilation parts c and a . This will be sufficient since the
time ordered product is linear in each field just as the normal product. Now, in
>
both cases x01 < x02 we have
T ( (x1 ) (x2 )) =
c

c (x1 )c (x2 )
c (x2 )c (x1 )

)
(

c (x1 )c (x2 )
= c (x1 )c (x2 ) + h0|
c (x2 )c (x1 )

|0i,

(7.770)

which is true since c (x1 )c (x2 ) commute or anticommute with each other, and
annihilate the vacuum state |0i. The same equation holds for a (x1 )a (x2 ). The
only nontrivial cases are those with a time-ordered product of c (x1 )a (x2 ) and
>
a (x1 )c (x2 ). The first becomes for x01 < x02 :
T (c (x1 )a (x2 )) =

c (x1 )a (x2 )
a (x2 )c (x1 )

)
(

c (x1 )a (x2 )
= (x1 ) (x2 ) + h0|
a (x2 )c (x1 )
c

|0i.

(7.771)

For x01 > x02 , this equation is obviously true. For x01 < x02 , the normal ordering
produces an additional term
(a (x2 )c (x1 ) c (x1 )a (x2 )) = [a (x2 ), c (x1 )] .

(7.772)

As the commutator or anticommutator of free fields is a c-number, they may equally


well be evaluated between vacuum states, so that we may replace (7.772) by
h0| [a (x2 ), c (x1 )] |0i.

(7.773)

Moreover, since a annihilates the vacuum, this reduces to


h0|a (x2 ), c (x1 )|0i.

(7.774)

The oppositely ordered operators a (x1 )c (x2 ) can be processed by complete analogy.
We shall now generalize this basic result to an arbitrary number of field operators.
In order to abbreviate the expressions let us define the concept of a contraction inside
a normal product
1 i1 i i+1 j1j j+1 n
N

(1 i1 i+1 j1 j+1 n ) .
i j N

(7.775)

Here = 1 for bosons and (1)ji1 for fermions accounting for the number of
fermion transmutations necessary to reach the final order. A normal product with

582

7 Quantization of Relativistic Free Fields

several contractions is defined by the successive execution of each of them. If only one
field is left inside the normal ordering symbol it is automatically normally ordered
so that

N()
= .
(7.776)
Similarly, if all fields inside a normal product are contracted, the result is no longer an
may be dropped using linearity and the trivial property
operator and the symbol N

N(1)
1.

(7.777)

The fully contracted normal product will be the relevant one in determining the
n-particle propagator. With these preliminaries we are now ready to prove Wicks
theorem for the expansion of a time-ordered product in terms of normally ordered
products.33

7.14.1

Wicks Theorem

The theorem may be stated as follows. An arbitrary time ordered product of free
fields can be expanded into a sum of normal products one for each different contraction between pairs of operators:
1 n )
T(1 n ) = N(

(no contraction)

1 2 3 4 n ) + . . . + N(
1 2 3 4 n )
1 2 3 4 n ) + N(
+N(
1 2 3 4 n ) + N(
1 2 4 n ) + . . .
+N(

(one contraction)

1 2 3 4 5 n ) + N(
1 2 3 4 5 n ) + N(
1 2 3 4 5 . . . n )
+N(
(1 2 4 3 5 n )
1 2 3 4 5 n ) + N
+N(

(7.778)

1 2 4 5 3 n ) + . . . (two contractions)
+ N(
(1 2 3 4 5 6 . . . n1 n ).
+...+ N

(remaining contractions).

In this particular expression we have assumed n to be even so that it is possible to


contract all operators. Otherwise each term in the last row would have contained
one uncontracted field. In either case the expansion on the right-hand side will be
abbreviated as
X
(1 n ),
N
(7.779)
allpaircontractions

i.e., we shall state Wicks theorem in the form


T (1 n ) =
33

allpaircontractions

1 n ).
N(

G.C. Wick, Phys. Rev. 80 , 268 (1950); F. Dyson, Phys. Rev. 82 , 428 (1951).

(7.780)

7.14 Free Green Functions of n Fields

583

Using Wicks theorem it is a trivial matter to calculate an arbitrary n-point function


of free fields. Since the vacuum expectation values of all normal products are zero
except for the fully contracted ones we may immediately write for an even number
of real fields:
G(n) (x1 xn ) = h0|T ((x1 ) (xn )) |0ii
X
((x1 ) (xn )) .
=
N

(7.781)

allfullpaircontractions

A simple combinatorical analysis shows that the right-hand side consists of


n!/(n/2)!2n/2 = 1 3 5 (n 1) (n 1)!!
pair terms.
In the case of complex fields, G(n,m) can only be nonzero for n = m and all
contractions and vanish. Thus there are altogether n! different contractions.
The proof of (7.780) goes by induction: For n = 2 the theorem was proved in
Eq. (7.768). Let it be true for a product of n fields 1 n , and allow for an
additional field . We may assume that it is earlier in time than the other fields
1 , . . . , n . For if it were not, we could always choose the earliest of the n fields,
move it completely to the right in the time ordered product, with a well-defined
phase factor due to Femi permutations, and proceed as we shall now do. Once the
rightmost field is the earliest, it may be removed trivially from the time ordered
product. The resulting product can be expanded according to the Wicks theorem
for a product of n field operators as
T(1 n ) = T(1 n ) =

allpaircontractions

(1 n ).
N

(7.782)

We now incorporate the extra field on the right-hand side into each of the normal
product in the expansion. When doing so we obtain from each term precisely the
contractions required by Wicks theorem for n + 1 field operators. The crucial
formula which yields these is
(1 n ) = N
(1 n ) + N
(1 2 n ) + N(
1 2 n )
N
(1 n ),
(1 n1n +N
+ N

(7.783)

which is valid as long as is earlier than the others. This formula is proved by
splitting into creation and annihilation parts and considering each part separately,
which is allowed because of the linearity of the normal products. For = a , the
formula is trivially true since all contractions vanish. Indeed, for a field a with an
earlier time arguments than that of all others, we find:
i a h0|T ((xi )a (x))|0i = h0|(xi )a (x)|0i 0.

(7.784)

584

7 Quantization of Relativistic Free Fields

For = c , let us for a moment assume that all other i s are annihilating parts ai .
If one or more creation parts ci are present, they will appear as left-hand factors
in each normally ordered term of (7.783) and participate only as spectators in the
further operations. Thus we prove
(a a c ), (7.785)
(a a )c = N(
a a c ) + N
(a a c ) + . . . + N
N
1
n
1
n
1
n
1
n
for any c earlier than the ai s. This expansion is proved by induction. To start we
show that it is true for n = 2:
(a )c = N(
a c ) + N
(a c ).
N

(7.786)

Using (7.776), the left-hand side is equal to a c , which can be rewritten as


a c = c a [a , c ] .

(7.787)

Since the commutator or anticommutator is a c-number, it can be replaced as in


(7.773) by its vacuum expectation value, and thus by a contraction (7.784):
[a , c ] = h0|[a , c ] |0i = a c .

(7.788)

a
c
(a a a )c = a N
a
N
0 1
n
0 (1 n ) .

(7.789)

(a c ) and trivially as a c , we see that (7.786) is true.


Rewriting c a as N
Now suppose (7.785) to be true for n factors and consider one more annihilation
factor a0 inside a normal product. This can immediately be taken outside:

The remainder can be expanded via (7.785):


(a a a a )c = a [N(a a a c ) + N(
a a . . . a c )
N
1 2
n
0 1 2
n
0
1 2
n
(a a a c )].
(1 a . . . c ) + . . . + N
+ N
2
1 2
n

(7.790)

The additional a0 can now be taken into all the terms where the creation operator
c appears in a contraction:
a a a a )c = a + N(a a a a c ) + N(
a a a . . . a c )
N(
0 1 2
n
0
0 1 2
n
0 1 2
n
a a a c )].
a 1 a . . . c ) + . . . + N(
+ N(
2
1 2
n
0

(7.791)

Only in the first term, work is needed to reorder the uncontracted creation operator.
Writing out the normal product explicitly this term reads
a0 c a1 an ,

(7.792)

where denotes the number of fermion transmutations necessary to bring c from


the right-hand side to its normal position. This product can now be rewritten in
normal form by using one additional commutator or anticommutator:
c a0 a1 an + [a0 , c ] a1 an .

(7.793)

7.15 Functional Form of Wicks Theorem

585

But the commutator or anticommutator is again a c-number, and since c is earlier


than a0 , it is equal to the contraction a0 c . The two terms in (7.793) can therefore
be rewritten as normal products
(a0 an c ) + N(
a0 an c ).
N

(7.794)

In both terms we have brought the field c back to its original position thereby
canceling the sign factors and in (7.793). These two are just the missing
terms in (7.790) to verify the expansion (7.785) for n + 1 operators, and thus Wicks
theorem (7.780).
Examples are
(1 2 3 ) + 1 2 N(3 ) 1 3 N(2 ) + 2 3 N(1 ),
T (1 2 3 ) = N

(7.795)

and
1 2 3 4 )
T (1 2 3 4 ) = N(
+ 1 2 N(3 4 ) 1 3 N(2 3 ) + 1 4 N(2 3 )
+ 2 3 N(1 4 ) 2 4 N(1 3 ) + 3 4 N(1 2 )
= 1 2 3 4 1 3 2 4 + 1 4 2 3 .

(7.796)

Each expansion contains (n 1)!! fully contracted terms.


There is a useful lemma to Wicks theorem which helps in evaluating vacuum
expectation values of ordinary products of field operators rather than time-ordered
ones. An ordinary product can be expanded in a sum of normal products with all
possible pair contractions just as in (7.779), except that contractions stand for commutators of the creation and annihilation parts of the fields, which are c-numbers:
(x1 )(x2 ) [a (x1 ), c (x2 )].

(7.797)

The proof proceeds by induction in the same way as before except that one does
not have to make the assumption of the additional field in the induction is the
earliest of all field operators. For the calculation of scattering amplitudes we shall
need both expansions.

7.15

Functional Form of Wicks Theorem

Although Wicks rule for expanding the time ordered product of n fields in terms of
normal products is quite transparent, it is somewhat cumbersome to state explicitly
for any n, as we can see in formula (7.779). It would be useful to find an algorithm
which specifies the expansion compactly in mathematical terms. This would greatly
simplify further manipulations of operator products. Such an algorithm can easily

586

7 Quantization of Relativistic Free Fields

be found. Let /(x) denote the functional differentiations for functions of space
and time, i.e.,
(x)
= (4) (x x ).
(7.798)

(x )
Then we see immediately that the basic formula (7.768) can be rewritten
!

d y1 d y2
N((x
G(y1 , y2 )
1 )(x2 )).
(y1)
(y2)
(7.799)
where functional differentiation treats field variables as c-numbers. The differentiation produces the single possible contraction which by (7.769) is equal to the free
propagator G(x1 , x2 ). For three operators, the Wick expansion
1
T((x1 )(x2 )) = 1
2


T((x1 )(x2 )(x3 )) = N((x
1 )(x2 )(x3 )) + N((x1 )(x2 )(x3 ))
1 )(x2 )(x
3 )) + N
2 )(x
3 ))
((x
((x1 )(x
+ N

(7.800)

can once more be obtained via the same operation


1
1
2

((x1 )(x2 )(x3 )).


d y1 d y2
N
G(y1 , y2)
(y1 )
(y2 )
4

(7.801)

If we want to recover the correct signs for fermions we have to imagine the differentiation variables (x) to be anticommuting objects, (x)(y) = (y)(x), and use
the rule G(x1 , x2 ) = G(x2 , x1 ) for fermions.
A time-ordered product of four field operators has a Wick expansion

T ((x1 )(x2 )(x3 )(x4 )) = N((x


1 )(x2 )(x3 )(x4 ))
X
((x1 )(x2 )(x3 )(x4 ))
+
N
onepaircontractions

((x1 )(x2 )(x3 )(x4 )).


N

(7.802)

twopaircontractions

Here we can obtain the same result by applying the functional differentiation
(

1Z 4 4
d y1 d y2
G(y1 , y2 )
1
2
(y1)
(y2)
1
+
8

d4 y1 d4 y2
G(y1 , y2)
(y1 )
(y2 )

!2

(7.803)

This looks like the first three pieces of a Taylor expansion of the exponential
(

1
exp
2

d y1 d y2
.
G(y1 , y2 )
(y1 )
(y2)
4

587

7.15 Functional Form of Wicks Theorem

Indeed, we shall now prove that the general Wick expansion is given by the functional
formula
1

T ((x1 ) (xn )) = e 2

d4 y1

d4 y2 (y

1)

G(y1 ,y2 ) (y

2)

((x1 ) (xn )).


N
(7.804)

To prepare the tools for deriving this result we introduce the following generating functionals of ordinary, time-ordered, and normal-ordered operator products,
respectively:
R

O[j] ei d x j(x)(x) ,
R 4
T [j] T ei d x j(x)(x) ,

i
N[j] Ne

d4 x j(x)(x)

(7.805)

Here j(x) is an arbitrary c-number local source term, usually referred to as an


external current. For Fermi fields (x), the currents have to be anticommuting
Grassmann variables which also anticommute with the fields. Under complex conjugation, products of Grassmann variables interchange their order. The product
j(x)(x) has therefore the complex-conjugation property
[j(x)(x)] = (x)j(x).

(7.806)

Expressions (7.805) are operators and have the obvious property that their nth
functional derivatives with respect to j(x) evaluated at j(x) 0 give the corresponding operator products of n fields:

(7.807)

(7.808)

(7.809)


(x1 ) (xn ) = (i)
,

O[j]

j(x1 )
j(xn )
j(x)0
n


T ((x1 ) (xn )) = (i)
,

T [j]

j(x1 )
j(xn )
j(x)0
n


((x1 ) (xn )) = (i)
.

N[j]
N

j(x1 )
j(xn )
j(x)0
n

We can now easily prove Wicks lemma for ordinary products of field operators.
We simply separate the field in the generating functional O[j] into creation and
annihilation parts:
R 4
a
c
O[j] ei d x j(x)[ (x)+ (x)] ,
(7.810)

and
apply the Baker-Hausdorff
formula (4.74) to this expression, with A
R 4
R 4
a
i d x j(x) (x) and B i d x j(x)(x)c . Because of the anticommutativity of
(x) and j(x) for fermions, the commutator [A, B] is
[A, B] =

d4 x1

d4 x2 [a (x1 ), c (x2 )] j(x1 )j(x2 ),

(7.811)

588

7 Quantization of Relativistic Free Fields

and is a c-number. Thus the expansion on the right-hand side of (4.74) terminates,
and we can use the formula as
1

eA+B = e 2 [A,B] eA eB .

(7.812)

This yields the simple result


O[j] = exp

1
R2

ei

d4 x1

d4 x2 [a (x1 ), c (x2 )] j(x1 )j(x2 )

d4 x j(x)c (x) i

d4 x j(x)a (x)

(7.813)

The operator on the right-hand side of the curly bracket is recognized to be precisely
the generating functional of the normal products:
N[j] = ei

d4 x j(x)c (x) i

d4 x j(x)a (x)

(7.814)

Now we observe that the currents j(x) in front of this operator are equivalent to a
functional differentiation with respect to the fields, j(x)=
i/(x). Wicks lemma
can therefore be rewritten in the following concise functional form
(

1
O[j] = exp
2

d x1

d x2 [ (x1 ), (x2 )]
N[j].
(x1 ) (x2 )
4

The -sign appears when commuting or anticommutating /(x) with j(x).


A Taylor expansion of the exponential prefactor renders via the functional derivatives precisely all pair contractions required by Wicks lemma, where a contraction
represent a commutator or anticommutator of the fields.
From the earlier proof in Section (7.14.1) it is now clear that exactly the same
set of pair contractions appears in the Wick expansion of the time-ordered products
of field operators, except that the pair contractions become free-field propagators
G(x1 , x2 ) h0|T ((x1 )(x2 )) |0i.

(7.815)

Thus we may immediately conclude that Wicks theorem has the functional form
(

1
T [j] = exp
2

d x1

d x2 G(x1 , x2 )
N[j],
(x1 ) (x2 )
4

(7.816)

which becomes, after reexpressing the functional derivatives in terms of the currents,
1

T [j] = e 2

d4 y1 d4 y2 j(y1 )G(y1 ,y2 )j(y2 )

[j].
N

(7.817)

Expanding both sides in powers of j(x), a comparison of the coefficients yields Wicks
expansion for products of any number of time-ordered operators.
The functional T [j] has the important property that its vacuum expectation
value
Z[j] h0|T [j]|0i
(7.818)

589

7.15 Functional Form of Wicks Theorem

collects all informations on n-point functions. Indeed, forming the nth functional
derivative as in (7.808) we see that
G(n) (x1 , . . . , xn ) = (i)n

Z[j].
j(x1 )
j(xn )

(7.819)

For this reason, the vacuum expectation value Z[j] is called the generating functional
of all n-poin functions. We can take advantage of Wicks expansion formula and
write down a compact formula for all n-point functions of free fields. We only have to
has a trivial vacuum
observe that the generating functional of normal products N[j]
expectation value

h0|N[j]|0i
1.
(7.820)
Then (7.817) leads directly to the simple expression
1

Z[j] = e 2

d4 y1 d4 y2 j(y1 )G(y1 ,y2 )j(y2 )

(7.821)

Expanding the right-hand side in powers of j(x), each term consists of a sum of combinations of propagators G(xi , xi+1 ). These are precisely the fully pair contracted
terms appearing in the Wick expansions (7.781) for the n-point functions
There exists another way of proving Eq. (7.821), by making use of the equation
of motion for the field (x). For simplicity, we shall consider a Klein-Gordon field
satisfying equation of motion
( 2 M 2 )(x) = 0.

(7.822)

This equation can be applied to the first functional derivative of Z[j]:


R 4
Z[j]
= h0|T (x)ei d yj(y)(y) |0i,
j(x)

(7.823)

to find the differential equation


( 2 M 2 )

Z[j]
= ij(x)Z[j].
j(x)

(7.824)

That this equation must be correct is seen by integrating it functionally with the
initial condition
Z[0] = 1,
(7.825)
and reproducing the known equation (7.821) with G(x, y) = i/( 2 M 2 ). A direct
proof of the differential equation (7.824) proceeds by expanding both sides in a
Taylor series in j(x) and processing every expansion term as in (7.42)(7.44). If we
were able to pass ( 2 m2 ) through the time-ordering operation, we would obtain
zero. During the passage, however, 02 encounters the Heaviside function (x0 x0 )
of the time ordering which generates additional terms. These are collected by the
right-hand side of (7.824). TheR nth power expansion term of Z[j] contributes on the
left-hand side a term nT (x)[ dzj(z)(z)]n1 /n!. Applying ( 2 m2 ) and using

590

7 Quantization of Relativistic Free Fields

the steps (7.42)(7.44), we obtain n times


a canonical commutator beween (x)
and
R
n2
(z) which can be written as j(x)T [ dzj(z)(z)] /(n 2)!. Summing these up
gives jZ[j]. These properties will find extensive use in Chapter 14.
The formulas can be generalized to complex fields by using the generating functionals
R

O[j, j ] = ei d x[j (x)(x)+j(x) (x)] ,


R 4

T [j, j ] = Tei d x[j (x)(x)+j(x) (x)] ,

i
N[j, j ] = Ne

d4 x[j (x)(x)+j(x) (x)]

(7.826)

Note that for fermions, the interaction j (x)(x) is odd under complex conjugation,
just as for real anticommuting fields in (7.805).
Now the propagators of n fields and m fields are obtained from the expectation value
Z[j, j ] = h0|T [j, j ]|0i
(7.827)
as the functional derivatives

G(n,m) (x1 , . . . , xn ; y1 . . . ym ) =


m+n
.

Z[j,
j
]

(i)

j (x1 )
j (xn ) j(y1 )
j(ym )
j(x)0

(7.828)

The explicit form of the generating functional Z[j, j ] is


Z[j, j ] = e

d4 xd4 yj (x)G(x,y)j(y)

(7.829)

The latter form is valid for fermions if one uses anticommuting currents j(x), j (x).

7.15.1

Thermodynamic Version of Wicks Theorem

In analogy with the generalization of the one-particle Green function (7.753) to the
n-point functions G(n) of (7.755) it is useful to generalize also the thermal propagator
(2.390) in an analogous way. We shall treat here explictly only the nonrelativistic
case. The generalization to relativistic fields is straight-forward.
For this we introduce the thermal average for of the imaginary-time-ordered
product of n + n fields:

G(n,n ) (x1 , 1 , . . . , xn , n ; x1 , 1 , . . . , xn , n )
=

Tr eHG /kB T T (x1 , 1 ) (xn , n ) (x1 , 1 ) (xn , n )


Tr [eHG /kB T ]

i

, (7.830)

where the free fields are expanded as in (2.404):


(x, ) =
(x, ) =

d-3 p fp (x, )ap ,

(7.831)

d-3 p fp(x, )ap ,

(7.832)

591

7.15 Functional Form of Wicks Theorem

with the wave functions


fp (x, ) ei[px(p) ]/h fp (x, ),

(p) = (p) ,

(7.833)

where is the chemical potential. We shall write all subsequent equations with an

explicit h
. An explicit evaluation of G(n,n ) is much more involved than for vacuum

expectation values. It proceeds using Wicks theorem, expanding G(n,n ) into a sum
of products of two-particle thermal propagators, the sum running over all possible
pair contractions, just as in Eq. (7.781) for the n-point function in the vacuum.
All thermal information is therefore contained in the one-particle propagators which
have the Fourier decomposition (2.417) with (2.416). The detailed form will here be
irrelevant.
To prove now Wicks expansion we first observe that the free Hamiltonian con
serves the particle number so that G(n,n ) vanishes unless n = n . As a next step we
order the product of field operators according to their imaginary time arguments
n thereby picking up some sign factor 1 depending on the number of transpositions of Fermi fields. On the right-hand side of Wicks expansion the corresponding
permutations lead to the same sign. Thus we may assume all field operators to be
time-ordered. The ordering in yields a product
in which the fields and appear
R -3
in a mixed fashion. We expand each field as d p fp p , using the same notation for
p = ap and p = ap , with the tacit understanding that in the latter case the wave
function is taken as fp rather than fp . Expanding the product of field operators in
this way we have
G(n,n) =

d-3 p1 d-3 p2n f1 f2n




= Tr eHG /kB T 1 2n /Tr eHG /kB T ,

(7.834)

where we have omitted the momentum labels on the operators , using only their
numeric subscripts. There are always equally many creators as annihilators among
the i s. In order to reduce the traces to a pure number, let us commute or anticommute 1 successively to the right via
1 2 3 . . . 2n = [1 , 2 ] 3 2n
2 [1 , 3 ] 4 2n
+
+2 3 [1 , 2n ]
2 3 2n 1 .

(7.835)

The commutators or anticommutators among 1 and i will give 0, if both are


creators or annihilators, 1 if 1 is an annihilator and i a creator, and 1 in the
opposite case. At any rate, they are c-numbers, so that they may be taken out of
the trace so that


Tr eHG /kB T 1 2 2n Tr eHG /kB T 2 2n 1

592

7 Quantization of Relativistic Free Fields




= [1 , 2 ] Tr eHG /kB T 3 4 2n [1 , 3 ] Tr eHG /kB T 2 4 2n




+ + [1 , 2n ] Tr eHG /kB T 2 3 2n1 .

(7.836)

Now we make use of the cyclic property of the trace and observe that
1 eHG /kB T = eHG /kB T 1 e(p)/kB T

(7.837)

where the sign factor = 1 depends on whether 1 is a creator or an annihilator,


respectively. Then we can rewrite the left-hand side in (7.836) as


1 e(p)/kB T Tr eHG /kB T 2 2n 1 .

Dividing the prefactor out, we obtain




Tr eHG /kB T 1 2n

(7.838)

= Tr eHG /kB T 1 2 3 2n


+ Tr eHG /kB T 1 2 3 2n
+ 

+ Tr eHG /kB T 1 2 3 2n ,

(7.839)

where a thermal contraction is defined as

[1 , i ]
,
1 e(p)/kB T

(7.840)

[ap , ap ]
= (3) (p p )(1 n(p) ).
1 e(p)/kB T

(7.841)

[ap , ap ]
= (3) (p p )n(p) .
1 e(p)/kB T

(7.842)

1 i =

with the same rules of taking these c-numbers outside the trace as in the ordinary
Wick contractions in Subsec. 7.14.1.
These contractions are just the Fourier transforms of the one-particle thermal
propagators as given in (2.407) and (2.408). For if 1 and i are both creators or
annihilators, this is trivially true: both expressions vanish identically. If 1 is an
annihilator and i a creator, we see that [recalling (2.406) and (2.409)]
1 i =
In the opposite case

1 i =

Multiplying either of these expressions by the product of wave functions fp (x)fp (x),
and integrating everything over the entire momentum space, we find that these
contractions are precisely the propagators G(x, ; x , ) = G(x x , ):

(x,
) (x , ) = (x , )(x,
) = G(x x , ).

(7.843)

In (7.839), the n-particle thermal Green function G(n,n) has been reduced to a sum of
(n1)-particle propagators, each multiplied by an ordinary propagator. Continuing
this procedure iteratively we arrive at Wicks expansion fot thermal Green functions.

593

7.15 Functional Form of Wicks Theorem

As before, this result can be expressed most concisely in a functional form. For
this we introduce


Z[j, j ] = Tr eHG /kB T T ei

d3 x d [j(x, ) (x, )+(x, )j (x, )]

/Tr eHG /kB T

(7.844)

as the generating functional of thermal Green functions. Then Wicks expansion


amounts to the statement
Z[i, j ] = e

d3 x d d3 x d j (x, )G(xx , )j(x, )

(7.845)

As before, this can be derived from the equation of motion [compare (2.400) and
(2.401)]

2
+
Z[j, j ] = Z[j, j ]j(x, ),
+

2M
j (x, )
!

+
+ = j (x, )Z[j, j ],
Z[j, j ]
j(x, )
2M

(7.846)

following the procedure to prove Eq. (7.824). Here the proof is simpler since only one
time derivative needs to be passed through the time-ordering operation T producing
a canonical equal-time commutator for every pair of fields, as in (2.237). As a check
we integrate again (7.846) with the initial condition Z[0, 0] = 1, and recover (7.845).
It is worth deriving also a related theorem concerning the thermal expectations
of normally ordered field operators. According to Wicks theorem in the operator
form (7.780), any product of time ordered operators can be expanded into normal
products, so that
Z[j, j ] = e

d4 xd4 x j (x)G(xx )j(x)

HG /kB T

Tr e

Ne

d4 x[j (x)(x)+ (x)j(x)]

Tr(eHG /kB T )

(7.847)

where we have used the four-vector notation x = (x, ) and d4 x = d3 x d , for


brevity. We can show that the expectation of the normal product can be expanded
once more in a Wick-like way. For this we introduce expectation of two normally
ordered operators
GN (x1 , x2 )

Tr eHG /kB T N (x1 )(x2 )


Tr [eHG /kB T ]

i

(7.848)

This quantity can be calculated by inserting (7.831) and (7.832), as well as intermediate states, with the result
GN (x1 , x2 ) =

d-3 p ei[p(x2 x1 )+i(p)(1 2 )]/h n(p) ,

(7.849)

594

7 Quantization of Relativistic Free Fields

where n is the particle distribution function for bosons and fermions, respectively:
n =

1
e/kB T 1

(7.850)

It can now be shown that the thrmal expectation value of an arbitrary normal
product can be expanded into expectation values of the one-particle normal products
(7.848) in the same way as the -ordered product into the expectations of oneparticle -ordered propagators. The proof is most easily given functionally. We
consider the generating functional

HG /kB T

N[j, j ] Tr e

Ne

d4 x( (x)j(x)+j (x)(x))

/Tr eHG /kB T

(7.851)

and differentiate this with respect to j (x):


R
N[j, j ]
HG /kB T
i d4 x( j+j )
=
Tr
e
N
(x)e
ij (x)




/Tr eHG /kB T .

(7.852)

Applying to this the field equation


2
+ (x) = 0,
+
2M

(7.853)

+
+
N[j, j ] = 0.

2M
j

(7.854)

j ] satisfies
we see that N[j,
!

This is solved by
[j, j ] = e
N

d4 x1 d4 x2 j (x1 )GN (x1 ,x2 )j(x2 )

(7.855)

where GN (x1 , x2 ) is a solution of the homogeneous differential equation. To obtain


the correct result for the ordinary imaginary-time propagator, we expand both sides
in j, j and see that GN (x1 , x2 ) coincides with (7.848), (7.849).

Appendix 7A

Li
enard-Wiechert Potential

(t), whose current


Let us calculate the field around a point source with a time-dependent position x
is
Z
1
(t )),
(3) (x x
(7A.1)
j(t , x ) = d (4) (x x
( )) =
d
x0 /d
p
2 (t) = (t), with v
(t) being is the velocity along
where is the proper time, and d
x0 /d = 1 v
(t):
the trajectory x
(t).
(t) x
(7A.2)
v
After performing the spatial integral in (7.183) we obtain
Z
1 1
(t t R(t )),
(x, t) = dt (t t )
4R

(7A.3)

Appendix 7B

Equal-Time Commutator from Time-Ordered Product

595

where R(t) is the distance from the moving source position at the time t:
(t )|.
R(t ) = |x x

(7A.4)

The integral over t gives a contribution only for t = tR determined from t by the retardation
condition
t tR R(tR ) = 0.
(7A.5)
At that point, the -function can be rewritten as
(t t R(t )) =

1
1
(t tR ) =
(t tR ),
(tR )
|d[t + R(t )]/dt |t =tR
1 n(tR ) v

(7A.6)

where n(t) is the direction of the distance vector from the charge.
n(t)

R(t)
|R(t)|

(7A.7)

Thus we find the Lienard-Wiechert-like potential:




1
.
(x, t) =
)R ret
(1 n v

(7A.8)

and the retarded distance from the point source


(tR )|.
Rret (t) = |x x

Appendix 7B

(7A.9)

Equal-Time Commutator from


Time-Ordered Product

As an exercise in handling the functions in Section 7.2, consider the relationship between the
commutator at equal-times and the time-ordered product. The first can be obtained from the
latter by forming the difference
[(x), (x )]x0 x0 = T((x)(x ))|x0 x0 = T((x)(x ))|x0 x0 = ,

(7B.1)

where is an infinitesimal positive time. Thus, the Fourier representation of the equal-time commutator is obtained from that of the Feynman propagator by replacing in the p0 -integral of (7.65)
0
0
0
the exponential eip (x x ) by
ip0
ip0
e
e . Thus we should get for the commutator function at equal times
Z
Z
0
dp0 ip0
i
d3 p ip(xx )
e
(e
eip ) 2
.
(7B.2)
C(x x )|x0 =x0 =
3
(2)
2
p M 2 + i
The p0 -integral can now be performed by closing the contour in the first term by a semicircle in
the lower halfplane, in the second term by one in the upper halfplane. To do this, we express the
Feynman propagator via the first of the rules (7.204) [see also (7.191)] as
i
2
= (p2 M 2 ) = (p02 p
),
p2 M 2 + i
and use the simple integrals
Z
dp0 ip0
e
(p0 p ) = 1,
2

Z
dp0 ip0
e
(p0 p ) = 0,
2

(7B.3)

dp0 ip0
e
+ (p0 + p ) = 0,
2

(7B.4)

dp0 ip0
e + (p0 + p ) = 1.
2

(7B.5)

596

7 Quantization of Relativistic Free Fields

to find a vanishing commutator function.


0
For its time derivative, These imply that exponentials eip together with (p0 p ) and
0
0
+ (p p ) have the same effect as Heaviside functions (p ) accompanied by 2 (p0 p ) and
2 (p0 +p ), respectively. We can therefore replace in the p0 -integral of the Fourier representation
(7B.2) of the commutator function
Z
Z

dp0 
i
dp0 ip0
ip0
(p0 ) (p0 ) 2(p2 M 2 ),
(7B.6)
(e
e ) 2

2
p M 2 + i
2
and this is the same as the p0 -integral in Eq. (7.198) for C(x x ) at equal times, which vanishes.
For the time derivative of the commutator function C(x x ), the integrand carries an extra
factor p0 , and (7B.6) becomes
Z
Z
0
0
dp0
dp0
i
(ip0 )(eip eip ) 2

i
p 2(p2 M 2 ) = i,
(7B.7)
2
2
p M + i
2

so that we find the correct result C(x


)x x0 = i (3) (x x ).
With the same formalism one may also check the consistency of the photon propagator (7.529)
with the commutation rules (7.532)(7.535). As in (7B.1), we write
Z
0

i
d4 k ik0
(e
eik )eik(xx ) 2
[A (x), A (x )]|x0 =x0 =
4
(2)
k + i


kk
g + (1 ) 2
(7B.8)
k + i

The contribution of g in the brackets yields the commutator function C(x x ) of the scalar
field. The second contribution proportional to k k is nontrivial. First we insert
1
1
[+ (k 0 k ) + + (k 0 + k )],
= + (k 2 ) =
k 2 + i
2k
with k = |k|. By forming the derivative and taking = 0, the first term gives
i (3) (x x ).
The second term is equal to
Z
0

d4 k ik0
1
i(1 )
(e
eik )eik(xx )
[+ (k 0 k ) + + (k 0 + k )]k 0
(2)4
2k

(7B.9)

Since k 0 changes the relative sign of the two pole contributions, the k-integration gives
i(1 ) (3) (x x )
leading to the equal-time commutator
[ A (x), A0 (x )]|x0 =x0 = i (3) (x x )

(7B.10)

in agreement with (7.532), provided that


[i Ai (x), A0 (x )]|x0 =x0 = 0.

(7B.11)

This commutator has the a spectral decomposition coming only from the second term in (7B.8):
It is equal to
Z
0

d4 k ik0
1
i(1 )
(e
eik )eik(xx ) 0 [+ (k 0 k ) + + (k 0 + k )]
(2)4
2k
1
k2 k 0
[+ (k 0 k ) + + (k 0 + k )]
(7B.12)
2k

Appendix 7C

Euler-Maclaurin formula

597

By writing
k0

1
1
[+ (k 0 k ) + + (k 0 + k )] = k 0
[+ (k 0 k ) + (k 0 + k )]
2k
2k

we see that the + -functions appear in the combination [+ (k 0 k )]2 [+ (k 0 +k )]2 . When closing
the integration contours in the upper or lower k 0 -plane, the double-poles give no contribution.
Hence
[i Ai (x), A0 (x )]|x0 =x0 = 0.

(7B.13)

For the same reason, the commutators [A i (x), A j (x )]|x0 =x0 receive only a contribution from the
first term and yield iij (3) (x x ). Other commutators are
[D(x), A i (x )]|x0 =x0

= [D(x), i A0 (x )]|x0 =x0


Z
0

d4 k ik0
i
=
(e
eik )eik(xx ) 2
k i k 0
(2)4
k + i

(7B.14)

yielding
i (3) (x x )
which shows that D(x) commutes with E i (x) at equal times. Together with the previous result
we recover the third of the canonical commutators (7.532)(7.535).

Appendix 7C

Euler-Maclaurin formula

The Euler-Maclaurin formula which serves to calculate discrete sums such as that in Eqs. (7.674),
(7.683), and (7.674) can be derived from a fundamental summation formula due to Bernoulli. Just
as a power np is integrated to
Z

dn np =

bp+1 (M ) bp+1 (N )
,
p+1

(7C.1)

where bp+1 (x) xp+1 , the corresponding sum over integers is given by sums over integer numbers
are given by Bernoulli polynomials Bp (x) as follows:
M
X

np =

n=N

Bp+1 (M + 1) Bp+1 (N )
,
p+1

(7C.2)

where Bp (x) are Bernoulli polynomials:


B0 (x)
B1 (x)
B2 (x)
B3 (x)
B4 (x)
B5 (x)
..
.

1,

1
= + x,
2
1
=
x + x2 ,
6
1
3
=
x x2 + x3 ,
2
2
1
= + x2 2x3 + x4 ,
30
5
5
1
= x + x3 x4 + x5 ,
6
3
2
(7C.3)

598

7 Quantization of Relativistic Free Fields

The more general integral


Z

dn (n + x)p =

reads for a sum:

M
X

(n + x)p =

n=N

bp+1 (M + x) bp+1 (N + x)
p+1

Bp+1 (M + 1 + x) Bp+1 (N + x)
.
p+1

(7C.4)

(7C.5)

The values of the polynomials Bp (x) at x = 0 are called Bernoulli numbers Bp :


1
1
1
1
, ... .
B0 = 1, B1 = , B2 = , B3 = 0, B4 = , B5 = 0, B6 =
2
6
30
42

(7C.6)

Except for B1 , all odd Bernoulli numbers vanish.


The close correspondence between integrals and sums has its counterpart in differential and
difference equations: By differentiating bp (x) = xp , we find:
bp (x) = pxp1 ,

(7C.7)

whereas Eqs. (7C.2) and (7C.5) follow from an analogous difference equation
Bp (x + 1) Bp (x) = pxp1 ,

(7C.8)

B0 (x) 1.

(7C.9)

bp (x) = pbp1 (x)

(7C.10)

Bp (x) = pBp1 (x).

(7C.11)

and the initial condition


The differential relation
is completely shared by Bp (x):

There is another important property of the function bp (x) = xp , the binomial expansion of (x + h)p
bp (x + h) =

p  
X
p
q=0

bq (x)hpq ,

(7C.12)

Bn (x)hpn ,

(7C.13)

which is shared by Bp (x):


Bp (x + h) =

p  
X
p

n=0

A noncommon property of the Bernoulli polynomials which is caused by the discrete nature of
the sum is
Bp (1 x) = (1)p Bp (x).
(7C.14)
There exists a simple generating function for the Bernoulli polynomials which ensures the
fundamental property (7C.8):

X
text
tp
=
B
(x)
.
(7C.15)
p
et 1 p=0
p!
The property (7C.5) can directly be used to calculate the difference between an integral and
a discrete sum over a function F (t). If an interval t (a, b) is divided into m slices of width
h = (b a)/N , we obviously have
h

N
1
X
n=0

F (a + (n + x)h)

dt F (t) =

X
hp
p=1

p!

[Bp (N + x) Bp (x)]F (p1) (a).

(7C.16)

Appendix 7C

599

Euler-Maclaurin formula

This follows from expanding the left-hand side into powers of h. After this, the binomial expansion
(7C.29) leads to

X
hp
p=1

p!

Bp (x + N + 1)F (p1) (a) =

p
X
X
hp  p 
Bl (x) N pl F (p1) (a)
p!
l
p=1
l=0

p
X
hl
l=0

l!

p
X
hl
l=0

l!

Bl (x)

X
p=l

1
(hN )pl F (p1) (a)
(p l)!

Bl (x)F (l1) (b),

(7C.17)

which brings (7C.16) to Eulers formula


h

N
1
X
n=0

F (a + (n + x)h)

dt F (t) =

X
hp
p=1

p!

Bp (x)[F (p1) (b) F (p1) (a)].

(7C.18)

Since all odd Bernoulli numbers vanish except for B1 = 1/2, it is useful to remove the p = 1
-term from the right-hand side and rewrite (7C.25) as
h

N
1
X
n=0

F (a + (n + x)h)
+

dt F (t) = (1/2 + x) [F (b) F (a)]

X
h2p
B2p (x)[F (2p1) (b) F (2p1) (a)].
(2p)!
p=1

(7C.19)

For x = 0 we may extend the sum on the left-hand side to n = N and obtain the Euler-Maclaurin
formula:34
h

N
X

n=0

F (a + nh)

dt F (t) =

1
[F (b) + F (a)]
2

X
h2p
+
B2p (x)[F (2p1) (b) F (2p1) (a)].
(2p)!
p=1

(7C.20)

If the sums (7C.5) are carried to infinity, they diverge. These infinities may, however, be
removed by considering such sums as a analytic continuations of convergent sums
(z, x)

1
,
(n
+
x)z
n=0

Re z > 1.

(7C.21)

known as Riemanns zeta functions. A continuation is possible by an appropriate deformation of


the contour in the integral representation
Z z1 xt
1
t e
(z, x) =
.
(7C.22)
(z) 0 1 et
This yields the relation:
(p, x) =
34

Bp+1 (x)
p+1

(7C.23)

See M. Abramowitz and I. Stegun, Handbook of Mathematical Functions, Dover, New York,
1965, Eqs. 23.1.30 and 23.1.32.

600

7 Quantization of Relativistic Free Fields

Thus the finite sum (7C.5) can be understood as the difference between the regularized infinite
sums

X
Bp+1 (x)
,
(n + x)p = (p, x) =
p+1
n=0
and
(7C.24)

X
Bp+1 (m + 1 + x)
p
(n + x) = (p, m + x) =
.
p+1
n=m+1
With the help of these regularized infinite sums we can derive Eulers formula (7C.25) by considering the left-hand side as a difference between the infinite sum
"

n=0

F (a + (n + x)h)

dt F (t)

(7C.25)

and a corresponding sum for b instead of a. By expanding each sum in a powers of h and using
the formula (7C.24), we find directly (7C.25).
For completeness, let us mention that alternating sums are governed quite similarly by Euler
polynomials
M
X

n=N

(1)mn (n + x)n =

En+1 (M + 1 + x) + (1)m En+1 (p + x)


,
2

(7C.26)

and En 2n En (1/2) are called Euler numbers:35


E0 = 1, E1 = 0, E2 = 1, E3 = 0, E4 = 5, E5 = 0, E6 = 61 . . . .

(7C.27)

All odd Euler numbers vanish. The Euler polynomials satisfy


Ep (x + 1) + Ep (x) = 2xp2 , Ep (x) = pEp1 (x), Ep (1 x) = (1)p Ep (x)

(7C.28)

and
Ep (x + h) =

p  
X
p

n=0

En (x)hpn ,

(7C.29)

The values Ep (0) are related to the Bernoulli numbers by


Ep (0) = 2(2p+1 1)Bp+1 /(p + 1).

(7C.30)

The generating function of the Euler polynomials is36

X
2ext
tp
=
E
(x)
.
p
et + 1 n=0
p!

35

(7C.31)

Note the difference with respect to the definition of the Bernoulli numbers (7C.6) which are
defined by the values of Bn (x) at x = 0.
36
Further properties of these polynomials can be found in M. Abramowitz and I. Stegun, Handbook of Mathematical Functions, Dover, New York, 1965, Chapter 23.

Notes and References

601

Notes and References


The particular citations in this chapter refer to:
[1] See Appendix in
J. Schwinger, Phys. Rev. 76, 790 (1949). In particular see the Appendix.
[2] See the textbook
H. Kleinert, Gauge Fields in Condensed Matter I , World Scientific, 1989 (http://www.physik.fu-berlin.de/~kleinert/b1),
[3] H. Kleinert, Proceedings of a NATO Advanced Study Institute on Formation and Interactions
of Topological Defects at the University of Cambridge, England, A-C. Davis adn R. Brandenberger, eds., Kluwer, London, 1995 (http://www.physik.fu-berlin.de/ kleinert/227).
[4] H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation,
World Scientific, Singapore 2008 (klnrt.de/b11).
(http://www.physik.fu-berlin.de/~kleinert/b11).
[5] The Clebsch-Gordan-like reduction of two-particle states into irreducible representations of
the Poincare group is discussed in detail in
H. Joos, Fortschr. Phys. 10, 65 (1962);
A. Macfarlane, Rev. Mod. Phys. 34, 41 (1962);
J. Dreitlein, in Lectures in Theoretical Physics Vol. VIIa, Proceedings of the Boulder Summer
School 1964, University of Colorado press, Boulder, 1965.
The expansion is intimately related to the Jacob-Wick expansion of helicity amplitudes of
general scattering and decay amplitudes:
M. Jacob, Nouvo Cimento 9, 826 (1958);
M. Jacob and G.C. Wick, Ann. Phys. 7, 404 (1959);
G.C. Wick, Ann. Phys. 187, 65 (1962);
L. Durand, L. De Celles, and L. Marr, Ann. Phys. 187, 65 (1962).
A review of the quantization a la Dirac is given in
L.V. Prokhorov, Sov.Phys.Uspekhi, 31, 151 (1988)
[6] M.J. Sparnay Physica 24, 751 (1958);
J. Tabor and R.H.S. Winterton, Nature 219, 1120 (1968);
T.H. Boyer, Annals of Physics 56, 474 (1970); Phys. Rev. 174, 1764 (1968);
S.K. Lamoreaux, Phys. Rev. Lett. 78 , 5 (1997).
See also
G. Esposito, A.Y. Kamenshchik, K. Kirsten, On the Zero-Point Energy of a Conducting
Spherical Shell , Int. J. Mod. Phys.A 14, 281 (1999) (hep-th/9707168);
G. Bressi, G. Carugno, R. Onofrio, and G. Ruoso, Phys. Rev. Lett. 88, 041804 (2002).
If the plates form spherical shells the energy was calculated in
A.A. Saharian, Phys. Rev. D 63, 125007 (2001).
Further aspects of the Casimir force are discussed in
M.V. Mustepanenko and N.N. Trunov, The Casimir Effect and its Applications, Clarendon
press, Oxford, 1997.
Recent measurements are discussed in
A.O. Suskov, W.J. Kim, D.A.R. Dalvit, S.K. Lamoreaux, (arXiv:1011.5219)
The temperature dependence is discussed in
I. Brevik, J.B. Aarseth, J.S. Hoeye, K.A. Milton, Phys. Rev. E 71, 056101 (2005).
[7] P.M. Alsing and P.W. Milonni, Am. J. Phys. 72, 1524 (2004) (arXiv:quant-ph/0401170).
[8] W.G. Unruh, Phys. Rev. D 14, 870 (1976).
See also

602

7 Quantization of Relativistic Free Fields


B.S. DeWitt, Phys. Rep. 19, 295 (1975);
P.C.W. Davies, J. Phys. A 8, 609 (1975).

[9] A. Higuchi, G.E.A. Matsas, and D. Sudarsky, Phys. Rev. D 45, R3308 (1992); 46, 3450
(1992); Phys. Rev. D 45 R3308 (1992); 46, 3450 (1992); D.A.T. Vanzella and G.E.A.
Matsas, Phys. Rev. D 63, 014010 (2001). J. Mod.Phys. D 11, 1573 (2002).

Nature always creates the best of all options


Aristoteles (384 BC322 BC)

8
Continuous Symmetries and Conservation Laws.
Noethers Theorem
In many physical systems, the action is invariant under some continuous set of transformations. Then there exist local and global conservation laws analogous to current
and charge conservation in electrodynamics. With the help of Poisson brackets, the
analogs of the charges can be used to generate the symmetry transformation, from
which they were derived.

8.1

Point Mechanics

Consider first a simple mechanical system with a generic action


A=

8.1.1

tb
ta

dt L(q(t), q(t),

t).

(8.1)

Continuous Symmetries and Conservation Law

Suppose A is invariant under a continuous set of transformations of the dynamical


variables:
q(t) q (t) = f (q(t), q(t)),

(8.2)
where f (q(t), q(t))

is some functional of q(t). Then these transformations are called


symmetry transformations. If the action is subjected successively to two symmetry
transformations, the result is again a symmetry transformation. Thus, symmetry
transformations form a group called the symmetry group of the system. It is important that the equations of motion are not used when testing whether the action is
invariant under (8.2).
For infinitesimal symmetry transformations (8.2), the difference
s q(t) q (t) q(t)

(8.3)

will be called a symmetry variation. It has the general form


s q(t) = (q(t), q(t),

t)
603

(8.4)

604

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Symmetry variations must not be confused with ordinary variations q(t) used in
Section 1.1 to derive the Euler-Lagrange equations (1.8). While the ordinary variations q(t) vansih at the ends, q(tb ) = q(ta ) = 0 [recall (1.6)], the symmetry
variation s q(t) are ususally nonzero at the ends.
Let us calculate the change of the action under a symmtery variation (8.4). Using
the chain rule of differentiation and a partial integration we obtain
s A =

tb

ta

"

b
L
L
L

s q(t) +
t
s q(t) .
dt

q(t)
q(t)

q(t)

ta

(8.5)

For orbits q(t) which satisfy the Euler-Lagrange equations (1.8), only boundary
terms survive, and we are left with
t

a
L

s A =
(q, q,
t) .

q
tb

(8.6)

By the symmetry assumptions, s A vanishes for any orbit q(t), implying that the
quantity
Q(t)

L
(q, q,
t)
q

(8.7)

is the same at times t = ta and t = tb . Since tb is arbitrary, Q(t) is independent of


the time t, i.e., it satisfies
Q(t) Q.

(8.8)

It is a conserved quantity, a constant of motion. The expression on the right-hand


side of (8.7) is called Noether charge
The statement can be generalized to transformations s q(t) for which the action
is not directly invariant but equal to an arbitrary boundary term
tb

s A = (q, q,
t) .

(8.9)

ta

In this case,

Q(t) =

L
(q, q,
t) (q, q,
t)
q

(8.10)

is a conserved Noether charge.


It is possible to derive the constant of motion (8.10) also without invoking the
action, but starting from the Lagrangian. For this we expand its symmetry variation
as follows:
"

"

d L
L
L
s q(t) +
t
s q(t)
s L L (q + s q, q + s q)
L(q, q)
=
q(t)
q(t)

dt q(t)

(8.11)

605

8.1 Point Mechanics

On account of the Euler-Lagrange equations (1.8), the first term on the right-hand
side vanishes as before, and only the last term survives. The assumption of invariance
of the action up to a possible surface term in Eq. (8.9) is equivalent to assuming
that the symmetry variation of the Lagrangian is purely a total time derivative of
some function (q, q,
t):
s L(q, q,
t) =

d
(q, q,
t).
dt

(8.12)

Inserting this into the left-hand side of (8.11), we find


"

d L
(q, q,
t) (q, q,
t) = 0,

dt q

(8.13)

thus recovering again the conserved Noether charge (8.8).


The existence of a conserved quantity for every continuous symmetry is the
content of Noethers theorem [1].

8.1.2

Alternative Derivation

From Eq. (8.5) we see that we may subject a classical orbit qc (t) extremizing the
action to an arbitrary variation a q(t). If this does not vanish at the boundaries,
the action changes by a pure boundary term:
t

L b
a A =
a q

q
ta

(8.14)

This observation leads to another derivation of Noethers theorem. Suppose we


subject a classical orbit to a new type of symmetry variation, to be called local
symmetry transformations, which generalizes the previous symmetry variations (8.4)
by making the parameter time-dependent:
st q(t) = (t)(q(t), q(t),

t).

(8.15)

The superscript t indicates the new time dependence in the parameter (t). These
variations may be considered as a special set of the general variations a q(t) introduced above. Thus st A must also be a pure boundary term of the type (8.14).
For the subsequent discussion it will be useful to introduce the infinitesimally
transformed orbit
q (t) q(t) + st q(t) = q(t) + (t)(q(t), q(t),

t),

(8.16)

and the associated Lagrangian:


L L(q (t), q (t)).

(8.17)

606

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Then the local symmetry variation of the action with repect to the time-dependent
parameter (t) is
st A

tb
ta

"

"

b
L
d L
d L

dt
(t) +
(t) .


(t) dt (t)

dt
ta

(8.18)

Along the classical orbits, e the action is extremal and satisfies the equation
A
= 0,
(t)

(8.19)

which translates for a local action to an Euler-Lagrange type of equation


d L
L

= 0.
(t) dt (t)

(8.20)

This can also be checked explicitly differentiating (8.17) according to the chain rule
of differentiation:
L
L
L
(q, q,
t);
=
(q, q,
t) +
(t)
q(t)
q(t)

L
L
=
(q, q,
t),
(t)

q(t)

(8.21)
(8.22)

and inserting on the right-hand side the ordinary Euler-Lagrange equations (1.8).
We now invoke the symmetry assumption, that the action is a pure surface term
under the time-independent transformations (8.15). This implies that
L
d
= .

dt

(8.23)

Combining this with (8.20), we derive a conservation law for the charge:
Q=

L
.

(8.24)

Inserting here Eq. (8.22) we find that this is the same charge as that derived by the
previous method

8.2

Displacement and Energy Conservation

As a simple but physically important example consider the case that the Lagrangian
does not depend explicitly on time, i.e., that L(q, q,
t) L(q, q).
Let us perform a
time translation on the coordinate frame
t = t .

(8.25)

In the new coordinate frame, the same orbit has the new description
q(t
) = q(t),

(8.26)

607

8.2 Displacement and Energy Conservation

i.e., the orbit q(t)


at the translated time t is precisely the same as the orbit q(t) at
the original time t. If we replace the argument of q(t)
in (8.26) by t , we describe a
time-translated orbit in terms of the original coordinates. This implies the symmetry
variation of the form (8.4) as
s q(t) = q (t) q(t) = q(t + ) q(t)
= q(t ) + q(t
) q(t) = q(t),

(8.27)

The symmetry variation of the Lagrangian is in general


s L = L(q (t), q (t)) L(q(t), q(t))

L
L
s q(t) +
s q(t).

q
q

(8.28)

Inserting s q(t) from (8.27) we find, without using the Euler-Lagrange equation,
!

L
L
d
s L =
q +
q = L.
q
q
dt

(8.29)

This has precisely the form of Eq. (8.12) with = L, as expected since time translations are symmetry transformations. The function in (8.12) happens to coincide
here with the Lagrangian.
According to Eq. (8.10), we find the Noether charge
Q=

L
q L(q, q)

(8.30)

to be a constant of motion. This is reognized as the Legendre transform of the


Lagrangian which is, of course, the Hamiltonian of the system.
Let us briefly check how this Noether charge is obtained from the alternative
formula (8.10). The time-dependent symmtry variation is here
st q(t) = (t)q(t)

(8.31)

under which the Lagragian is changed by


st L =
with

and

L
L
L
L
q +
(q +
q) =
+
,

q
q


L
L
=
q

q
L
L
d
L
=
q +

q = L.

q
q
dt

(8.32)

(8.33)

(8.34)

This shows that time translations fulfill the symmetry consition (8.23), and that the
Noether charge (8.24) coincides with the Hamiltonian found in Eq. (8.10).

608

8.3

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Momentum and Angular Momentum

While the conservation law of energy follow from the symmetry of the action under
time translations, conservation laws of momentum and angular momentum are found
if the action is invariant under translations and rotations.
Consider a Lagrangian of a point particle in a Euclidean space
L = L(xi (t), x i (t), t).

(8.35)

In contrast to the previous discussion of time translation invariance, which was


applicable to systems with arbitrary Lagrange coordinates q(t), we denote the coordinates here by xi to emphasize that we now consider cartesian coordinates. If
the Lagrangian does depends only on the velocities x i and not on the coordinates
xi themselves, the system is translationally invariant. If it depends, in additiona,
only on x 2 = x i x i , it is also rotationally invariant.
The simplest example is the Lagrangian of a point particle of mass m in Euclidean
space:
m
(8.36)
L = x 2 .
2
It exhibits both invariances, leading to conserved Noether charges of momentum
and angular momentum, as we shall now demonstrate.

8.3.1

Translational Invariance in Space

Under a spatial translation, the coordinates xi change to


xi = xi + i ,

(8.37)

where i are small numbers. The infinitesimal translations of a particle path are
[compare (8.4)]
s xi (t) = i .

(8.38)

Under these, the Lagrangian changes by


s L = L(xi (t), x i (t), t) L(xi (t), x i (t), t)
L
L
s xi = i i = 0.
=
i
x
x

(8.39)

By assumption, the Lagrangian is independent of xi , so that the right-hand side vanishes. This is to be compared with the symmetry variation of the Lagrangian around
the classical orbit, calculated via the chain rule, and using the Euler-Lagrange equation:
!

"

d L
d L
L
s L =
s xi +

s xi
i
i
i
x
dt x
dt x
"
#
d L i
=

dt x i

(8.40)

609

8.3 Momentum and Angular Momentum

This has the form (8.6), from which we extract a conserved Noether charge (8.7) for
each coordinate xi :
L
.
x i
These are simply the canonical momenta of the system.
pi =

8.3.2

(8.41)

Rotational Invariance

Under rotations, the coordinates xi change to


xi = Ri j xj

(8.42)

where Ri j is an orthogonal 3 3 -matrix. Infinitesimally, this can be written as


Ri j = i j k kij

(8.43)

where ! is an infinitesimal rotation vector. The corresponding rotation of a particle


path is
s xi (t) = xi (t) xi (t) = k kij xj ( ).

(8.44)

It is useful to introduce the antisymmetric infinitesimal rotation tensor


ij k kij ,

(8.45)

s xi = ij xj .

(8.46)

and write

Then we can write the change of the Lagrangian under s xi ,


s L = L(xi (t), x i (t), t) L(xi (t), x i (t), t)
L
L
i
=

x
+
s x i ,
s
xi
x i

(8.47)

as
!

L j L j
x + i x ij = 0.
s L =
xi
x

(8.48)

If the Lagrangian depends only on the rotational invariants x2 , x 2 , x x and powers


thereof, the right-hand side vanishes on account of the antisymmetry of ij . This
ensures the rotational symmetry.
We now calculate the symmetry variation of the Lagrangian once more via the
chain rule and find, using the Euler-Lagrange equations,
!

"

d L
d L
L
s L =
s xi +

s xi
i
i
i
x
dt x
dt x
"
"
#
#
d L j
L
1d
=
xi j (i j) ij .
x ij =
dt x i
2 dt
x

(8.49)

610

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

The right-hand side yields the conserved Noether charges of the type (8.7), one for
each antisymmetric pair i, j:
Lij = xi

L
L
xj i xi pj xj pi .
j
x
x

(8.50)

These are the antisymmetric components of angular momentum.


Had we worked with the original rotation angles k , we would have found the
angular momentum in the more common form:
1
Lk = kij Lij = (x p)k .
2

(8.51)

The quantum-mechanical operators associated with these after replacing pi


i/xi have the well-known commutation rules
i, L
j ] = iijk L
k.
[L

(8.52)

In the tensor notation (8.50), these become




ij , L
kl ] = i ik L
jl il L
jk + jl L
ik jk L
il .
[L

8.3.3

(8.53)

Center-of-Mass Theorem

Consider now the transformations corresponding to a uniform motion of the coordinate system. We shall study the behavior of a set of free massive point particles
in Euclidean space described by the Lagrangian
L(x i ) =

X
n

mn 2
x .
2 n

(8.54)

Under Galilei transformations, the spatial coordinates and the time are changed
to
x i (t) = xi (t) v i t
t = t.

(8.55)

where v i is the relative velocity along the ith axis. The infinitesimal symmetry
variations are
s xi (t) = x i (t) xi (t) = v i t,

(8.56)

which change the Lagrangian changes by


s L = L(xi v i t, x i v i ) L(xi , x i ).

(8.57)

Inserting the explicit form (8.54), we find


s L =

X
n

i
mn h i
(x n v i )2 (x n i )2 .
2

(8.58)

611

8.3 Momentum and Angular Momentum

This can be written as a total time derivative:


d
d X
v2
s L = =
mn x in v i + t
dt
dt n
2

"

(8.59)

proving that Galilei transformations are symmetry transformations in the Noether


sense. By assumption, the velocities v i in (8.55) are infinitesimal, so that the second
term can be ignored.
By calculating s L once more via the chain rule with the help of the EulerLagrange equations, and equating the result with (8.59), we find the conserved
Noether charge
X

Q =

L
s xi
x i
X

mn x in t

mn xin

vi.

(8.60)

Since the direction of the velocity v i is arbitrary, each component is separately a


constant of motion:
Ni =

mn x i t +

mn xn i = const.

(8.61)

This is the well-known center-of-mass theorem [2]. Indeed, introducing the centerof-mass coordinates
xiCM

and velocities
i
vCM

mn xn i
,
n mn

(8.62)

mn x n i
,
n mn

(8.63)

P
n

the conserved charge (8.61) can be written as


Ni =

i
mn (vCM
t + xiCM ).

(8.64)

The time-independence of N i implies that the center-of-mass moves with uniform


velocity according to the law
i
xiCM (t) = xi0 CM + vCM
t,

where

(8.65)

Ni
(8.66)
xi0 CM = P
n mn
is the position of the center of mass at t = 0.
Note that in non-relativistic physics, the center of mass theorem is a consequence
of momentum conservation since momentum mass velocity. In relativistic
physics, this is no longer true.

612

8.3.4

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Conservation Laws resulting from Lorentz Invariance

In relativistic physics, particle orbits are described by functions in spacetime


x ( )

(8.67)

where is an arbitrary Lorentz-invariant parameter. The action is an integral over


some Lagrangian:
Z
A = d L (x ( ), x ( ), ) ,
(8.68)

where x ( ) denotes the derivative with respect to the parameter . If the Lagrangian depends only on invariant scalar products x x , x x , x x , then it is
invariant under Lorentz transformations
x x = x

(8.69)

where is a 4 4 matrix satisfying


gT = g,

(8.70)

with the Minkowski metric

g =

1
1
1

(8.71)

For a free massive point particle in spacetime, the Lagrangian is


q

L(x(
)) = Mc g x x .

(8.72)

It is reparametrization invariant under f ( ), with an arbitrary function f ( ).


Under translations
s x ( ) = x ( ) ( ),
(8.73)
the Lagrangian is obviously invariant satisfying , s L. Calculating this variation
once more via the chain rule with the help of the Euler-Lagrange equations, we find
L
L
0 =
s x + s x
d

x
x

!
Z
d L
=
d
.
d x

(8.74)

From this we obtain the Noether charges


p

x ( )
L
= Mc q
= Mcu ,

x
g x x

(8.75)

613

8.3 Momentum and Angular Momentum

which satisfy the conservation law


d
p (t) = 0.
d

(8.76)

They are the conserved four-momenta of a free relativistic particle. The quantity
x
u q
g x x

(8.77)

is the dimensionless relativistic four-velocity of the particle. It has the property


u u = 1 and is reparametrization invariant. By choosing for the physical time
t = x0 /c, we can express u in terms of the physical velocities v i = dxi /dt as
u = (1, v i /c),

with

1 v 2 /c2 .

(8.78)

Note the minus sign in the definition of (8.75) of the canonical momentum with
respect to the nonrelativistic case. It is necessary to write Eq. (8.75) covariantly.
The derivative with respect to x transforms like a covariant vector with a subscript
, whereas the physical momenta are p .
For small Lorentz transformations near the identity we write
= +

(8.79)

= g

(8.80)

where
is an arbitrary infinitesimal antisymmetric matrix. An infinitesimal Lorentz transformation of the particle path is
s x ( ) = x ( ) x ( )
= x ( ).

(8.81)

Under it, the symmetry variation of a Lorentz-invariant Lagrangian vanishes:


s L =

L
L
x
+
x = 0

x
x

(8.82)

This is to be compared with the symmetry variation of the Lagrangian calculated


via the chain rule with the help of the Euler-Lagrange equation
!

"

d L
d L
L
s x +

s x
s L =

x
d x
d x
"
#
d L
=
x
d x
!
1 d
L
L
=
x
.

x
2
d
x
x

(8.83)

614

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

By equating this with (8.82) we obtain the conserved rotational Noether charges
[containing again a minus sign as in (8.75)]
L = x

L
L
+ x
= x p x p .
x
x

(8.84)

They are the four-dimensional generalizations of the angular momenta (8.50). The
quantum-mechanical operators
i(x x )
L

(8.85)

obtained after the replacement p i/x satisfy the four-dimensional spacetime


generalization of the commutation relations (8.53):


, L
] = i g L
g L
+ g L
g L
.
[L

(8.86)

The quantities Lij coincide with the earlier-introduced angular momenta (8.50).
The conserved components
L0i = x0 pi xi p0 Mi

(8.87)

yield the relativistic generalization of the center-of-mass theorem (8.61):


Mi = const.

8.4

(8.88)

Generating the Symmetries

A mentioned in the introduction to this chapter, the relation between invariances


and conservation laws has a second aspect. With the help of Poisson brackets, the
charges associated with continuous symmetry transformations be used to generate
the symmetry transformation from which they it was derived. Explicitly,
x(t)].
s x = i[Q,

(8.89)

The charge derived in Section 7.2 from the invariance of the system under time
displacement is the most famous example for this property. The charge (8.30) is by
definition the Hamiltonian,
Q H,
whose operator version generates infinitesimal time displacements by the Heisenberg
equation of motion:
x(t)].
x (t) = i[H,
(8.90)
This equation is obviously the same as (8.89).
To quantize the system canonically, we may assume the Lagrangian to have the
standard form
M 2
x V (x),
(8.91)
L(x, x)
=
2

615

8.4 Generating the Symmetries

so that the Hamiltonian operator becomes, with the canonical momentum p x:

2
= p + V (
H
x).
2M

(8.92)

Equation (8.90) is then a direct consequence of the canonical equal-time commutation rules
[
p(t), x(t)] = i, [
p(t), p(t)] = 0, [
x(t), x
(t)] = i.
(8.93)
The charges (8.41) derived Section 7.3 from translational symmetry are another
famous example. After quantization, the commutation rule (8.89) becomes, with
(8.38),
j = ii [
pi (t), xj (t)].
(8.94)
This coincides with one of the canonical commutation relations (here it appears only
for time-independent momenta, since the system is translationally invariant).
The relativistic charges (8.75) of spacetime generate translations via
s x = = i [
p (t), x
( )].

(8.95)

in agreement with the relativistic canonical commutation rules (4.34).


Similarly we find that the quantized versions of the conserved charges Li in
Eq. (8.51) generate infinitesimal rotations:
i , xj (t)],
s xj = i ijk xk (t) = i i [L

(8.96)

whereas the quantized conserved charges N i of Eq. (8.61) generate infinitesimal


Galilei transformations, and that the charges Mi of Eq. (8.87) generate pure rotational Lorentz transformations:
s xj = i x0 = ii [Mi , xj ],
s x0 = i xi = ii [Mi , x0 ].

(8.97)

Since the quantized charges generate the rotational symmetry transformations,


they form a representation of the generators of the symmetry group. They have the
same commutation rules among each other as the generators of the symmetry group.
The charges (8.51) associated with rotations, for example, have the commutation
rules
i, L
j ] = iijk L
j ,
[L
(8.98)
which are the same as those between the 3 3 generators of the three-dimensional
rotations (Li )jk = iijk .
The quantized charges of the generators (8.84) of the Lorentz group satisfy the
commutation rules (8.86) of the 4 4 generators (8.85)
, L
] = ig L
.
[L

(8.99)

This follows directly from the canonical commutation rules (8.95) [i.e., (4.34)].

616

8.5

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Field Theory

A similar relation between continuous symmetries and constants of motion holds in


field theories.

8.5.1

Continuous Symmetry and Conserved Currents

Let A be the action of an arbitrary field (x),


A=

d4 xL(, , x),

(8.100)

and suppose that a transformation of the field


s (x) = (, , x)

(8.101)

changes the Lagrangian density L merely by a total derivative


s L = ,

(8.102)

or equivalently, that it changes the action A by a surface term


s A =

d4 x .

(8.103)

Then s L is called a symmetry transformation.


Given such a symmetry transformation, we can find a current four-vector
j =

(8.104)

that has no four-divergence


j (x) = 0.

(8.105)

The expression on the right-hand side of (8.104) is called a Noether current and
(8.105) is referred to as current conservation law . It is a local conservation law .
Assuming all fields to vanish at spatial infinity, this implies a global conservation
law for the charge associated with the charge density j 0 :
Q(t) =

d3 x j 0 (x, t).

(8.106)

Indeed, writing the time derivative of the charge as an integral


d
Q(t) =
dt

d3 x 0 j 0 (x, t)

(8.107)

and adding on the right-hand side a spatial integral over a total three-divergence,
which vanished because of the boundary conditions, we find
d
Q(t) =
dt

d3 x 0 j 0 (x, t) =

d3 x [0 j 0 (x, t) + i j i (x, t)] = 0.

(8.108)

617

8.5 Field Theory

Thus the charge is conserved:


d
Q(t) = 0.
dt

(8.109)

The proof of the local conservation law (8.105) is just as easy as for the mechanical action (8.1). We calculate the symmetry variation of L under the symmetry
transformation in a similar way as in Eq. (8.11), and find
!

L
L
L
s +
s L =

s



!
!
L
L
L
=
+

.

(8.110)

Then we invoke the Euler-Lagrange equation to remove the first term. Equating the
second term with (8.102) we obtain

L
= 0.

(8.111)

The relation between continuous symmetries and conservation is called Noethers


theorem [1].

8.5.2

Alternative Derivation

There is again an alternative derivative of the conserved current analogous to


Eqs. (8.15)(8.24). It is based on a variation of the fields under symmetry transformations whose parameter is made artificially spacetime-dependent (x), thus
extending (8.15) to
sx (x) = (x)((x), (x), x).
(8.112)
As before in Eq. (8.17), let us calculate the Lagrangian density for a slightly
transformed field
(x) (x) + sx (x),
(8.113)
calling it
L L( (t), (t)).

(8.114)

say The associated action differs from the original one by


sx A =

dx

("

L
L
L
(x) +

(x)
(x)
(x)
(x)
#

"

#)

(8.115)

From this we obtain the Euler-Lagrange-like equation


L
L

= 0.
(x)
(x)

(8.116)

618

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

By assumption, the action was a pure surface term under x-independent transformations, implying that
L
= .
(8.117)
(x)
Together with (8.116) we see that
j =

sx L

(x)

(8.118)

has no four-divergence. By the chain rule of differentiation


L(x)
L
+
,

(8.119)

L
L
=
(, , x),
(x)

(8.120)

st L =
so that

i.e., the current (8.118) coincide with (8.104).


Note that (8.120) can also be written as
L
L sx
=

8.5.3

(8.121)

Local Symmetries

If we apply the alternative derivation of a conserved current to a local symmetry,


such as a local gauge symmetry, the current density (8.118) vanishes identically.
Let us here illuminate the symmetry origin of this phenomenon.
To be specific, we consider directly the field theory of electrodynamics. The
theory does have a conserved charge resulting from the global U(1)-symmetry of the
matter Lagrangian. There is a conserved current which is the source of a massless
particle, the photon. This is described by a gauge field which is minimally coupled to
the conserved current. A similar structure exists for many internal symmetries giving
rise to nonabelian versions of the photon, such as gluons which give rise to strong
interactions, and W - and Z-vector mesons, which mediate the weak interactions. It
is useful to reconsider Noethers derivation of conservation laws in such theories.
The conserved matter current in a locally gauge invariant theory cannot be found
any more by the rule (8.118) which was so useful in the globally invariant theory. For
the gauge transformation of quantum electrodynamics, the derivative with respect
to the local field transformation (x) would be simply given by
j =

L
.

(8.122)

619

8.5 Field Theory

and yield identically zero, due to local gauge invariance. We may, however, subject
just the matter field to a local gauge transformation at fixed gauge fields. Then we
obtain the correct current

L
(8.123)
.
j
em
Since the complete change under local gauge transformations sx L vanishes identically, we can alternatively vary only the gauge fields and keep the particle orbit
fixed

L
(8.124)
.
j =
m
This is done most simply by forming the functional derivative with respect to the
em
gauge field and omitting the contribution of L :
m

L
j =
.

(8.125)

An interesting consequence of local gauge invariance can be found for the gauge
field itself. If we form the variation of the pure gauge field action
em

s A =

em

d4 x tr sx A

A
A

(8.126)

and insert for sx A an infinitesimal pure gauge field configuration


sx A = i (x)

(8.127)

the right-hand side must vanish for all (x). After a partial integration this implies
the local conservation law j (x) = 0 for the current
em

A
j (x) = i
.
A

(8.128)

In contrast to the earlier conservation laws derived for matter fields which were valid
only if the matter fields obey the Euler-Lagrange equations, the current conservation
law for gauge fields is valid for all field configurations. It is an identity which we
may call Bianchi identity due to its close analogy with the Bianchi identities in
Riemannian geometry.
To verify the conservation of (12.62), we insert the Lagrangian (12.3) into (12.62)
and find j = F /2. This current is trivially conserved for any field configuration
due to the antisymmetry of F .

620

8.6

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Canonical Energy Momentum Tensor

As an important example for the field theoretic version of the theorem consider the
usual case that the Lagrangian density does not depend explicitly on the spacetime
coordinates x:
L = L(, ).

(8.129)

We then perform a translation along an arbitrary direction = 0, 1, 2, 3 of spacetime


x = x ,

(8.130)

under which field (x) transforms as


(x ) = (x).

(8.131)

This equation expresses the fact that at the same absolute point in space and time,
which in one coordinate term is labeled by the coordinates x and in the other by
x , the field has the same value.
Under an infinitesimal translation of the field configuration coordinate the Lagrangian density undergoes the following symmetry variation
s L L( (x), (x)) L((x), (x))
L
L
s (x) +
s (x),
=
(x)

(8.132)

where
s (x) = (x) (x)

(8.133)

is the symmetry variation of the fields. For the particular transformation (8.131),
the symmetry variation becomes simply
s (x) = (x).

(8.134)

The Lagrangian density (8.129) changes by


s L(x) = L(x).

(8.135)

Hence the requirement (8.103) is satisfied and s (x) is a symmetry transformation.


The function happens to coincide with the Lagrangian density
= L.

(8.136)

We can now define a set of currents j , one for each , which in the particular case
at hand are denoted by ,
=

L
L.

(8.137)

621

8.6 Canonical Energy Momentum Tensor

They have no four divergence


(x) = 0.

(8.138)

As a consequence, the total four momentum of the system defined by

P =

d3 x0 (x)

(8.139)

is independent of time.
The alternative derivation of the currents goes as follows. Introducing
sx (x) = (x) (x)

(8.140)

sx (x) = (x) (x).

(8.141)

we see that
On the other hand, the chain rule of differentiation yields
sx L =
Hence

L
L
(x) (x) +
{[ (x)] + (x)} . (8.142)
(x)
(x)
L
L
=

(x)

(8.143)

and we obtain again the energy-momentum tensor (8.137).


Note that (8.143) can also be written as
L
L sx
=
(x)
(x)

(8.144)

Since is a contravariant vector index, the set of currents u forms a Lorentz


tensor called the canonical energy momentum tensor . The component
0 0 =

L
0 L
0

(8.145)

is recognized to be the Hamiltonian density in the canonical formalism.

8.6.1

Electromagnetism

As a an important physical application of the field theoretic Noether theorem, consider the free electromagnetic field with the action
L=

1
F F ,
4c

(8.146)

where F are the field strength F A A . Under a translation in space


and time from x to x , the vector potential undergoes a similar change as in
(8.131):
A (x ) = (x).

(8.147)

622

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

As before, this equation expresses the fact that at the same absolute space time
point, which in the two coordinate frames is labeled once by x and once by x, the
field components have the same numerical values. The equation transformation law
(8.147) can be rewritten in an infinitesimal form as
s A (x ) A (x ) A (x )
= A (x + ) A (x )
= A (x ).

(8.148)
(8.149)

Under it, the field tensor changes as follows


s F = F ,

(8.150)

so that the Lagrangian density is a total four divergence:


s L =

1
F F = L
2c

(8.151)

Thus, the spacetime translations (8.149) are symmetry transformations, and the
currents
=

L
A L
A

(8.152)

are conserved:
(x) = 0.

(8.153)

Using L/ A = F , this becomes more explicitly

1
1
= F A F F .
c
4


(8.154)

This is the canonical energy-momentum tensor of the electromagnetic field.

8.6.2

Dirac Field

We now turn to the Dirac field which has the well-known action
A=

d x L(x) =

d4 x(x)(i
M)(x),

(8.155)

where are the Dirac matrices

(8.156)

Here ,
are four 2 2 matrices
( 0 , i ).
( 0 , i ).

(8.157)

623

8.6 Canonical Energy Momentum Tensor

whose zeroth component is the unit matrix


0

1 0
0 1

(8.158)

and whose spatial components consist of the Pauli spin matrices


1

0 1
1 0

0 i
i 0

1
0
0 1

(8.159)

On behalf of the algebraic properties of the Pauli matrices


i j = ij + iijk k ,

(8.160)

the Dirac matrices (8.156) satisfy the anticommutation rules


{ , } = 2g .

(8.161)

x = x ,

(8.162)

Under spacetime translations

the Dirac field transforms in the same way as the previous scalar and vector fields:
(x ) = (x),

(8.163)

s (x) = (x).

(8.164)

or infinitesimally:
The same thing is true for the Lagrangian density, where
L (x ) = L(x),

(8.165)

s L(x) = L(x).

(8.166)

and
Thus we obtain the Noether current
=

L
+ cc L

(8.167)

with the local conservation law


(x) = 0.

(8.168)

From (8.155) we see that


L
1

=

2

(8.169)

so that we obtain the canonical energy-momentum tensor of the Dirac field:


1
=
+ cc L
2

(8.170)

624

8.7

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Angular Momentum

Let us now turn to angular momentum in field theory. Consider first the case of a
scalar field (x).
Under a rotation of the coordinates,
xi = Ri j xj

(8.171)

the field does not change, if considered at the same space point, i.e.,
(xi ) = (xi ).

(8.172)

The infinitesimal symmetry variation is:


s (x) = (x) (x).

(8.173)

Using the infinitesimal form (8.46) of (8.171),


xi = ij xj ,

(8.174)

we see that
s (x) = (x0 , xi xi ) (x)
= i (x)xj ij .

(8.175)

Suppose we are dealing with a rotationally Lorentz-invariant Lagrangian density


which has no explicit x-dependence:
L = L((x), (x)).

(8.176)

Then the symmetry variation is


s L = L( (x), (x)) ((x), (x))

L
s (x) +
s (x)
=
(x)
(x)

(8.177)

For a Lorentz-invariant L, the derivative L/ is a vector proportional to .


More explicitly, Eq. (8.175) reads
"

L
L
s L =
i xj +
(i Lxj ) ij


#
"


L
j
i ij = i Lxj ij
= (i L)x +
j

(8.178)

The right-hand side is a total derivative. In arriving at this result, the antisymmetry
of ij has been used twice: once in order to drop the second term in the brackets,
which is possible since L/i is proportional to i , as a consequence of the

625

8.8 Four-Dimensional Angular Momentum

assumed rotational invariance of L,1 and once, in order to pull xj inside the last
parentheses.
Calculating s L once more using the Euler-Lagrange equations gives
L
L
s +
s
L

!
!
L
L
L
s +

s
=



!
L
j
=
i x ij .

s L =

(8.179)

Thus the Noether charges


ij,

L
i xj i L xj (i j)

(8.180)

have no four-divergence
Lij, = 0.

(8.181)

The associated charges


Lij =

d3 xLij,

(8.182)

are called the total angular momenta of the field system. Remembering the canonical
energy-momentum tensor
=

L
L,

(8.183)

we see that Lij, can also be rewritten as


Lij, = xi j xj i .

8.8

(8.184)

Four-Dimensional Angular Momentum

A similar procedure applies to pure Lorentz transformations. An infinitesimal boost


to a rapidity i produces a coordinate change
x = x = x + i i x + 0 i xi .

(8.185)

This can be written as


x = x ,
1

Recall the similar argument after Eq. (8.48)

(8.186)

626

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

where
ij = 0,
0i = i0 = i .

(8.187)

With the tensor , the restricted Lorentz transformations and the infinitesimal
rotations can be treated on the same footing. The rotations have the form (8.186)
for the particular choice
ij = ij = ijk k
0i = i0 = 0.

(8.188)

We can now identify the symmetry variations of the filed as being


s (x) = (x x ) (x)
= (x)x .

(8.189)

Just as in (8.178), the Lagrangian density transforms as the total derivative


s = (Lx )

(8.190)

and we obtain the Noether currentss


,

L
=
x L x ( )

(8.191)

The right-hand side can be expressed in terms of the canonical energy-momentum


tensor (8.137) so that we find
,

L
=
x Lx ( )

= x x .

(8.192)

These currents have no four-divergence


L, = 0.

(8.193)

The associated charges

d3 x L,0

(8.194)

are independent of time.


For the particular form of in (8.187), we find time independent components
i0
L . The components Lij coincide with the previously-derived angular momenta.
The constancy of Li0 is the relativistic version of the center-of-mass theorem
(8.65). Indeed, since
Li0 =

d3 x (xi 00 x0 i0 ),

(8.195)

627

8.9 Spin Current

we can then define the relativistic center of mass


xiCM
and the average velocity

d3 x 00 xi
= R 3
d x 00
R

i
vCM
= cR

Pi
d3 xi0
=
c
d3 x 00
P0

(8.196)

(8.197)

i
Since d3 xi0 = P i is the constant momentum of the system, also vCM
is a constant.
0i
Thus, the constancy of L implies the center of mass to move with the constant
velocity

i
xiCM (t) = xi0CM + v0CM
t

(8.198)

with xi0CM = L0i /P 0.


The quantities L are referred to as four-dimensional orbital angular momenta.
It is important to point out that the vanishing divergence of L, makes
symmetric:
L, = (x x )
= = 0.

(8.199)

Thus, translationally invariant field theories whose orbital angular momentum is


conserved have always a symmetric canonical energy-momentum tensor.
=

8.9

(8.200)

Spin Current

If the field (x) is no longer a scalar but has several spatial components, then
the derivation of the four-dimensional angular momentum becomes slightly more
involved.

8.9.1

Electromagnetic Fields

Consider first the case of electromagnetism where the relevant field is the four-vector
potential A (x). When going to a new coordinate frame
x = x

(8.201)

the vector field at the same point remains unchanged in absolute spacetime. But
since the components A refer to two different basic vectors in the different frames,
they must be transformed simultaneously with x . Since A (x) is a vector, it transforms as follows:
A (x ) = A (x).

(8.202)

628

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

For an infinitesimal transformation


s x = x

(8.203)

this implies the symmetry variation


s A (x) = A (x) A (x) = A (x x) A (x)
= A (x) x A .

(8.204)

The first term is a spin transformation, the other an orbital transformation. The
of the
orbital transformation can also be written in terms of the generators L
Lorentz group defined in (8.84) as
A(x).
sorb A (x) = i L

(8.205)

It is convenient to introduce 4 4 spin transformation matrices L with the matrix


elements:
(L ) i (g g g g ) .
(8.206)
They satisfy the same commutation relations (8.86) as the differential operators
defined in Eq. (8.85). If we add the two and form the operator of total fourL
dimensional angular momentum
+ L ,
J L

(8.207)

we can write the transformation (8.204) as


sorb A (x) = i J A(x).

(8.208)

If the Lagrangian density involves only scalar combinations of four-vectors A


and if it has no explicit x-dependence, it changes under Lorentz transformations like
a scalar field:
L (x ) L(A (x ), A (x )) = L(A(x), A(x)) L(x).

(8.209)

Infinitesimally, this amounts to


s L = ( Lx ) .

(8.210)

With the Lorentz transformations being symmetry transformations in the


Noether sense, we calculate as in (8.180) the current of total four-dimensional angular momentum:
J

L
L
A
A x Lx ( ).
=
A
A

(8.211)

The last two terms have the same form as the current L, of the fourdimensional angular momentum of the scalar field. Here they are the currents of
the four-dimensional orbital angular momentum:.
,

L
=
A x Lx ( ).
A

(8.212)

629

8.9 Spin Current

Note that this current has the form


L, = i

i
L h
L
A
+

Lx

)
.
A

(8.213)

are the differential operators of four-dimensional angular momentum in


where L
the commutation rules (8.86).
Just as the scalar case (8.192), the currents (8.212) can be expressed in terms of
the canonical energy-momentum tensor as
L, = x x .

(8.214)

The first term in (8.211),


,

"

L
A ( ) ,
=
A

(8.215)

is referred to as the spin current. It can be written in terms of the 4 4-generators


(8.206) of the Lorentz group as
, = i

L
(L ) A .

(8.216)

The two currents together


J , (x) L, (x) + , (x)

(8.217)

are conserved, J , (x) = 0. Individually, the terms are not conserved.


The total angular momentum is given by the charge
J =

d3 x J ,0 (x).

(8.218)

It is a constant of motion. Using the conservation law of the energy-momentum


tensor we find, just as in (8.199), that the orbital angular momentum satisfies
L, (x) = [ (x) (x)] .

(8.219)

From this we find the divergence of the spin current


, (x) = [ (x) (x)] .

(8.220)

For the charges associated with orbital and spin currents


L (t)

d3 xL,0 (x),

(t)

d3 x,0 (x),

(8.221)

this implies the following time dependence:


L (t) =
(t) =

d3 x [ (x) (x)] ,

d3 x [ (x) (x)] .

(8.222)

630

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Fields with spin have always have a non-symmetric energy momentum tensor.
Then the current J , becomes
J , =

sx L
Lx ( )
(x)

(8.223)

By the chain rule of differentiation, the derivative with respect to , (x) can come
only from field derivatives. For a scalar field

and for a vector field

L sx
sx L
=
,
(x)
(x)

(8.224)

sx L
L sx A
=
(x)
A

(8.225)

The alternative rule of calculating angular momenta is to introduce spacetimedependent transformations


x x = (x)x
(8.226)
under which the scalar fields transform as
s = (x)x

(8.227)

sx = L (x)x = (x L) (x)

(8.228)

and the Lagrangian density as

By separating spin and orbital transformations of sx A we find the two contributions


, and L, to the current J , of the total angular momentum, the latter
receiving a contribution from the second term in (8.223).

8.9.2

Dirac Field

We now turn to the Dirac field. Under a Lorentz transformation (8.201), this transforms according to the law

(x )
(x) = D()(x),

(8.229)

where D() are the 4 4 spinor representation matrices of the Lorentz group. Their
matrix elements can most easily be specified for infinitesimal transformations. For
an infinitesimal Lorentz transformation
= + ,

(8.230)

under which the coordinates are changed by


s x = x

(8.231)

631

8.9 Spin Current

the spin transforms under the representation matrix


D(

1
+ ) = 1 i (x),
2

(8.232)

where are the 4 4 matrices acting on the spinor space


i
= [ , ]
2

(8.233)

From the anticommutation rules (8.161) it is easy to verify that the spin matrices
S /2 satisfy the same commutation rules (8.86) as the previous orbital and
and L of Lorentz transformations.
spin-1 generators L
The field has the symmetry variation [compare (8.204)]:
s (x) = (x) (x) = D( + )(x x) (x)
1
= i (x) x (x)
2


1
(x) i 1 J (x),
= i S + L
2
2

(8.234)

the last line showing the separation into spin and orbital transformation for a Dirac
particle.
Since the Dirac Lagrangian is Lorentz-invariant, it changes under Lorentz transformations like a scalar field:
L (x ) = L(x).

(8.235)

Infinitesimally, this amounts to


s L = ( Lx ) .

(8.236)

With the Lorentz transformations being symmetry transformations in the


Noether sense, we calculate the current of total four-dimensional angular momentum extending the formulas (8.192) and (8.211) for scalar field and vector potential.
The result is
J

h
i
L
L
L + cc + Lx ( ) .
i
= i

(8.237)

As before in (8.212) and (8.192), the orbital part of (8.237) can be expressed in
terms of the canonical energy-momentum tensor as
L, = x x .

(8.238)

The first term in (8.237) is the spin current


,

1
L
=
i
+ cc .
2

(8.239)

632

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Inserting (8.169), this becomes explicitly


i
1 [ }]
1
.
, =
=
=
2
2
2

(8.240)

The spin density is completely antisymmetric in the three indices.2


The conservation properties of the three currents are the same as in Eqs. (8.218)
(8.222).
Due to the presence of spin, the energy-momentum tensor is nonsymmetric.

8.10

Symmetric Energy-Momentum Tensor

Since the presence of spin is the cause for the asymmetry of the canonical energymomentum tensor, it is suggestive that by an appropriate use of the spin current
should be possible to construct a new modified momentum tensor
T = +

(8.241)

which is symmetric,
while still having the fundamental property of , that the
R
integral P = d3 x T a0 is the total energy-momentum vector of the system. This
is be the case if a0 is a three-divergence of a spatial vector. Such a construction
was found by Belinfante in 1939 [3]. He introduced the tensor
1
T = (, , + , )
2

(8.242)

whose symmetry is manifest, due to (8.220) and the symmetry of the last two terms
under . Moreover, by the components
1
T a0 = a0 (a0,c 0, + ,0 ) = x T x T
2

(8.243)

which gives the same total angular momentum as the canonical expression (8.217):
J

d3 x J ,0 .

(8.244)

Indeed, the zeroth component of (8.243) is


x 0 x 0

i
1h
k (0,k 0k, + k,0 )x ( ) .
2

(8.245)

Integrating the second term over d3 x and performing a partial integration gives, for
= 0, = i:
1

2
2

i0,k

d x x k (

0k,i

ki,0

00,k

) x k (

0k,0

k0,0

) =

d3 x 0i,0 ,
(8.246)

This property is important for being able to construct a consistent quantum mechanics in a space with torsion.
See H. Kleinert, Path Integrals in Quantum Mechanics Statistics, Polymer Physics, and Financial Markets, World Scientific, Singapore 2008
(www.physik.fu-berlin.de/~kleinert/b5).

8.10 Symmetric Energy-Momentum Tensor

633

and for = i, = j
Z
Z
h
i
1
3
i
j0,k
0k,j
kj,0

d x x k (

+ ) (i j) = d3 x ij,0 .
(8.247)
2
The right-hand sides are the contributions of the spin to the total angular momentum.
For the electromagnetic field, the spin current (8.215) reads, explicitly
i
1h
(8.248)
, = F A ( ) .
c
From this we calculate the Belinfante correction
1
=
[ (F A F A ) (F A F A ) + (F A F A )]
2c
1
(F A ).
(8.249)
=
c
Adding this to the canonical energy-momentum tensor (8.154)
1
1
= (F A g F F ),
(8.250)
c
4
we find the symmetric energy-momentum tensor
1
1
1
(8.251)
T = (F F g F F ) + ( F )A
c
4
c
The last term vanishes due to the free Maxwell field equations, F = 0 and can
be dropped.
Note that the proof of the symmetry of T involves the field equations; via the
divergence equation (8.220).
It is useful to see what happens to Belinfantes energy momentum tensor in the
presence of an external current, i.e., if T is calculated from the Lagrangian
1
1
L = F F 2 j A .
(8.252)
4c
c
with an external current. The energy-momentum tensor is


1
1
1

F A g F F + 2 g j A
=
(8.253)
c
4
c
generalizing (8.25).
The spin current is the same as before, and we find Belinfantes energymomentum tensor [3]
1
T = + (F A )
(8.254)
c
1
1
1
1
= (F F g F F ) + 2 g j A + ( F )A .
c
4
c
c

Using Maxwells equations F = j , the last term can also be rewritten as


1
j A .
(8.255)
c
This term prevents T from being symmetric, unless the current vanishes.

634

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

8.10.1

Gravitational Field

The derivation of the canonical energy-momentum tensor is similar to the above


case of the electromagnetic field in Subsec. (8.9.1). We start from the quadratic
action of the gravitational field (4.375),
f

1
8

d4 x( h h 2 h h + 2 h h h h), (8.256)

and identify the canonically conjugate field ,


f

(8.257)

as
=

1
[( h h ) ( h h) h + h ]+( ).(8.258)
8

It is antisymmetric under the exchange , and symmetric in . From the


integrand in (8.256) we calculate, according to the general expression (8.137),
f

ab

L a cd
h ab L = bcd a hcd ab L
b hcd
1
(b hcd d hhc + bd ch bc d h bd e hec + bd e hcb ) a hcd
=
2
ab

( hhc a hbc 2c hab a hbc + 2 hab b h h a h).


8
=

(8.259)

In order to find the symmetric energy-momentum tensor T , we follow Belinfantes


construction rule. The spin current density is calculated as in Subsec. 8.9.1, starting
from the substantial derivative of the tensor field
s h = h + h .

(8.260)

Following the Noether procedure we find, as in (8.216),


, = 2

( ) = 2 [ h ( )] .
h

(8.261)

Combining the two results according to Belinfantes formula (4.57), we find the
symmetric energy-momentum tensor
f

= hc ( h h hc d + h h h ) L.
(8.262)

Using the field equation = 0 and the Hilbert gauge (4.406) with = 0,
this takes the simple form in h as well as
f

1
1
2
=
8
2




. (8.263)

635

8.11 Internal Symmetries

8.11

Internal Symmetries

In quantum field theory, an important role by classifying various actions is played by


internal symmetries. They do not involve any change in the spacetime coordinate
of the fields, i.e., they have the form
(x) = eiG (x)

(8.264)

where G are the generators of some Lie group and the associated transformation
parameters. The field may have several indices on which the generator G acts as
a matrix. The symmetry variation associated with (8.264) is obviously
s (x) = iG(x)

(8.265)

The most important example is that of a complex field and a generator G = 1,


where (8.264) is simply a multiplication by a constant phase factor. One also speaks
of U(1)-symmetry. Other important examples are those of a triplet or an octet of
fields i with G being the generators of an SU(2) vector representation or an SU(3)
octet representation (the adjoint representations of these groups). The first case is
associated with charge conservation in electromagnetic interactions, the other two
with isospin and SU(3) invariance in strong interactions. The latter symmetries are,
however, not exact.

8.11.1

U(1)-Symmetry and Charge Conservation

Given a Lagrangian density L(x) = L((x), (x), x) depending only on the absolute
squares ||2 , ||2 , ||. Then L(x) is invariant under U(1)-transformations
s (x) = i(x)

(8.266)

i.e., s L = 0. By the chain rule of differentiation we find, using the Euler-Lagrange


equation
!
"
#
L
d L
L
s L =
s +

s = 0
(8.267)
dt

Inserting (8.266), we find that
j =

(8.268)

is a conserved current.
For a free relativistic complex scalar field with a Lagrangian density
L(x) = m2

(8.269)

we have to add the contributions of real and imaginary parts of the field in formula
(8.268), and obtain the conserved current

j = i

(8.270)

636

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

where denotes the left-minus-right derivative:

( ).

(8.271)

For a free Dirac field, we find from (8.268) the conserved current

j (x) = (x)
(x).

8.11.2

(8.272)

SU(N)-Symmetry

For more general internal symmetry groups, the symmetry variations have the form
s = ii Gi ,

(8.273)

and the conserved currents are


ji = i

8.11.3

L
Gi

(8.274)

Broken Internal Symmetries

The physically important symmetries SU(2) of isospin and SU(3) are not exact. The
Lagrange density is not strictly zero. In this case we remember the alternative derivation of the conservation law from (8.116). We introduce the spacetime-dependent
parameters (x) and conclude from the extremality property of the action that

L
L
=
i (x)
i (x)

(8.275)

This implies the divergence law for the above derived current
ji (x) =

8.12

L
.
i

(8.276)

Generating the Symmetry Transformations


on Quantum Fields

As in quantum mechanical systems, the charges associated with the conserved currents obtained in the previous section can be used to generate the transformations
of the fields from which they were derived. One merely has to invoke the canonical
field commutation rules.
As an important example, consider the currents (8.274) of an internal U(N)symmetry. Their charges
Z
L
i
Q = i d3 x
Gi
(8.277)

can be written as
Qi = i

d3 xGi ,

(8.278)

8.12 Generating the Symmetry Transformations on Quantum Fields

637

where (x) L(x)/ (x) is the canonical momentum of the field (x). After
quantization, these fields satisfy the canonical commutation rules:
[(x, t), (x , t)] = i (3) (x x ),
[(x, t), (x , t)] = 0,
[(x, t), (x , t)] = 0.

(8.279)

From this we derive directly the commutation rule between the quantized charges
(8.278) and the field (x):
[Qi , (x)]

= i Gi (x)

(8.280)

We also find that the commutation rules among the quantized charges
i, Q
j ] = [Gi , Gj ].
[Q

(8.281)

Since these coincide with those of the matrices Gi this proves that the operators Qi
form a representation of the generators of symmetry group in the Fock space.
It is important to realize that the commutation relations (8.280) and (8.281)
remain valid also in the presence of symmetry braking terms as long as these do
not contribute to the canonical momentum of the theory. Such terms are called soft
symmetry breaking terms. The charges are no longer conserved, so that we must
attach a time argument to the commutation relations (8.280) and (8.281). All times
in these relations must be the same, in order to invoke the equal-time canonical
commutation rules.
The most important example is the commutation relation (8.95) which holds also
in the presence of a potential V (q) in the Hamiltonian. This brakes translational
symmetry, but does not contribute to the canonical momentum p = L/ q.
In this
case, the relation generalizes to
j = ii [
pi (t), xj (t)],

(8.282)

which is correct thanks to the validity of the canonical commutation relations (8.93)
at arbitrary equal times, also in the presence of a potential.
Another important example are the commutation rules of the conserved charges
associated with the Lorentz generators (8.238):
J

d3 xJ ,0 (x),

(8.283)

which are the same as those of the 4 4-matrices (8.206), and those of the quantum
mechanical generators (8.85):
[J , J ] = ig J .

(8.284)

The generators J d3 xJ ,0 (x), are sums J = L (t) + (t) of charges


(8.221) associated with orbital and spin rotations. According to (8.222), these individual charges are time dependent, only their sum being conserved. Nevertheless,
R

638

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

they both generate Lorentz transformations: L (t) on the spacetime argument of


the fields, and (t) on the spin indices. As a consequence, they both satisfy the
commutation relations (8.284):
, L
] = ig L
,
[L
,
] = ig
.
[

(8.285)

The commutators (8.281) have played an important role in developing a theory


of strong interactions, where they first appeared in the form of a charge algebra of
the broken symmetry SU(3) SU(3) of weak and electromagnetic charges. This
symmetry will be discussed in more detail in Chapter 10.

8.13

Energy Momentum Tensor of a


Relativistic Massive Point Particle

If we want to study energy and momentum of charged relativistic point particles in


an electromagnetic field it is useful to consider the action (8.68) with (8.72) as an
integral over a Lagrangian density:
A=

d xL(x),

with L(x) =

L(x ( )) (4) (x x( )).

(8.286)

Then we can derive for point particles similar local conservation laws as for fields.
Instead of doing this, however, we shall simply take the already known global conservation laws and convert them into the local ones by inserting appropriate -functions
with the help of the trivial identity
Z

d4 x (4) (x x( )) = 1.

(8.287)

Consider for example the conservation law (8.74) for the momentum (8.75). With
the help of (8.287) this becomes
0=

"

d
d
d x
p ( ) (4) (x x( )).
d

(8.288)

Note that in this expression the boundaries of the four volume contain the information on initial and final times. We then perform a partial integration in , and
rewrite (8.288) as
0=

d4 x

Z
i Z
d h
p ( ) (4) (x x( )) + d4 x
d p ( ) (4) (x x( )).
d

(8.289)

The first term vanishes if the orbits come from and disappear into infinity. The
second term can be rewritten as
0=

d x

Z

(4)

d p ( )x ( ) (x x( )) .

(8.290)

8.14 Energy Momentum Tensor of Massive Charged Particles in an Electromagnetic Field639

This shows that

(x)

d p ( )x ( ) (4) (x x( ))

(8.291)

satisfies the local conservation law


(x) = 0.

(8.292)

This is the conservation law for the energy-momentum tensor of a massive point
particle.
The total momenta are obtained from the spatial integrals over 0 :
P (t)

d3 x 0 (x).

(8.293)

For point particles, they coincide with the canonical momenta p (t). If the Lagrangian depends only on the velocity x and not on the position x (t), the momenta
p (t) are constants of motion: p (t) p .
The Lorentz invariant quantity
M 2 = P 2 = g P P

(8.294)

is called the total mass of the system. For a single particle it coincides with the
mass of the particle.
Subjecting the orbits x ( ) to Lorentz transformations according to the rules of
the last section we find the currents of total angular momentum
L, x x ,

(8.295)

to satisfy the conservation law:


L, = 0.

(8.296)

A spatial integral over the zeroth component of the current L, yields the conserved
charges:
Z

L (t) d3 x L,0 (x) = x p (t) x p (t).


(8.297)

8.14

Energy Momentum Tensor of Massive Charged


Particles in an Electromagnetic Field

Let us also consider an important combination of a charged point particle and an


electromagnetic field Lagrangian
A = mc

d g x ( )x ( )

1
4c

d4 xF F

e
c2

d x ( )A (x( )).
(8.298)

640

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

By varying the action in the particle orbits, we obtain the Lorentz equation of motion
dp
e
= F x ( ).
(8.299)
d
c
By varying the action in the vector potential, we find the Maxwell-Lorentz equation
e
(8.300)
F = x ( ).
c
The action (8.298) is invariant under translations of the particle orbits and the
electromagnetic fields. The first term is obviously invariant, since it depends only
on the derivatives of the orbital variables x ( ). The second term changes under
translations by a pure divergence [recall (8.135)]. Also the interaction term changes
by a pure divergence, which is seen as follows: Since the symmetry variation changes
x ( ) x ( ) , under which x ( ) is invariant,
x ( ) x ( ),

(8.301)

and A (x ) changes as follows:


A (x ) A (x ) = A (x + ) = A (x ) + A (x ),

(8.302)

Altogether we obtain
m

s L = L .

(8.303)

We now we calculate the same variation once more invoking the equations of
motion. This gives
s A =

d Lm
d
s x +
d x

em

L
dx
s A .
A
4

(8.304)

The first term can be treated as in (8.289)(8.290) after which it acquires the form

e
d
p + A
d
c


e
p + A (4) (x x( ) (8.305)
c



Z
Z
d (4)
e
(x x( ))
+
d4 x
d p + A
c
d

d4 x

d
d



and thus, after dropping boundary terms,

d
e
(p + A ) =
d
c

d4 x

dx (4)
e
(x x( )). (8.306)
d p + A
c
d

The electromagnetic part is the same as before, since the interaction contains no
derivative of the gauge field. In this way we find the canonical energy-momentum
tensor


Z
e

(x) =
d p + A x ( ) (4) (x x( ))
c


1
1

F A g F F .
(8.307)
c
4

8.14 Energy Momentum Tensor of Massive Charged Particles in an Electromagnetic Field641

Let us check its conservation by calculating the divergence:


e
(x) =
d p + A x ( ) (4) (x x( ))
c


1
1
1
F A (F F ) .
F A
c
c
4


(8.308)

The first term is, up to a boundary term, equal to


d (4)
e
(x x( )) =
d p + A

d
Z

"

e
d
p + A
d
d
c


#

(4) (x x( )).(8.309)

Using the Lorentz equation of motion (8.299), this becomes


e
c

d
d F x ( ) + A (4) (x x( )).
d

(8.310)

Inserting the Maxwell equation


F

= e

d (dx /d ) (4) (x x( )),

(8.311)

the second term in Eq. (8.308) can be rewritten as


eZ
dx (4)

d
A (x x( )),
c
d

(8.312)

which is the same as


e

dx dx (4)
F +
A (x x( )),
d
d

(8.313)

thus canceling (8.310). The third term in (8.308), finally is equal to

1 1
F F (F F )
c
4


(8.314)

due to the antisymmetry of F . With the help of the homogeneous Maxwell equation, the Bianchi identity
F + F + F = 0,

(8.315)

we see that this term vanishes identically.


It is easy to construct from (8.307) Belinfantes symmetric energy momentum
tensor. We merely observe that the spin density is entirely due to the vector potential, and hence the same as before [see (8.248)]
, =

i
1 h
F A ( ) .
c

(8.316)

642

8 Continuous Symmetries and Conservation Laws; Noethers Theorem

Hence the additional piece to be added to the canonical energy momentum tensor
is again [see (8.249)]
1
(F A )
c
1
=
( F A + F A ).
(8.317)
2
The second term in this expression serves to symmetrize the electromagnetic part
of the canonical energy-momentum tensor and brings it to the Belinfante form:
=

em

1
1
= F F g F F .
c
4


(8.318)

The first term in (8.317), which in the absence of charges vanished, is now just what
is needed to symmetrize the matter part of . Indeed, using once more Maxwells
equation, it becomes
Z
e

d x ( )A (4) (x x( ))
(8.319)
c
thus canceling the corresponding term in (8.307). In this way we find that the total
energy-momentum tensor of charged particles plus electromagnetic fields is simply
the term of the two symmetric energy-momentum tensor.
m

em

T = T + T
(8.320)


Z
1
1
1
d u u (4) (x x( ))
F F g F F
=
m
c
4
For completeness, let us cross check also its conservation. Forming the divergence
T , the first term gives now only
Z
e
d x ( )F (x( )),
(8.321)
c
in contrast to (8.310), which is canceled by the divergence in the second term
1
e
F F =
c
c
in contrast to (8.313).

d x ( )F (x( )),

(8.322)

Notes and References


For more details on the canonical formalism see
S. Schweber, Relativistic Quantum Fields, Harper and Row, N.Y., N.Y., 1961
A.O. Barut, Electrodynamics and Classical Theory of Fields and Particles, MacMillan, New York,
N.Y. 1964
J.D. Jackson, Classical Electrodynamics, John Wiley + Sons, New York, N.Y., 1975
L.D. Landau, E.M. Lifshitz, The Classical Theory of Fields, Addison-Wesley, Reading, Mess., 1951

The individual citations refer to:

Notes and References

643

[1] E. Noether, Nachr. d. vgl. Ges. d. Wiss. Gottingen, Math-Phys. Klasse, 2, 235 (1918);
See also
E. Bessel-Hagen, Math. Ann. 84, 258 (1926);
L. Rosenfeld, Me. Acad. Roy. Belg. 18, 2 (1938);
F. Belinfante, Physica 6, 887 (1939).
[2] S. Coleman and J.H. VanVleck, Phys. Rev. 171, 1370 (1968).
[3] The Belinfante energy-momentum tensor is discussed further in
H. Kleinert, Gauge Fields in Condensed Matter , Vol. II Stresses and Defects, World
Scientific Publishing, Singapore 1989, pp. 7441443 (kl/b2), where kl is short for
http://www.physik.fu-berlin.de/~kleinert;
H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation,
World Scientific, Singapore 2009, pp. 1497 (kl/b11).

All human things are subject to decay,


and when fate summons, monarchs must obey
John Dryden (1631-1700)

9
Scattering and Decay of Particles
So far we have discussed only free particles. If we want to detect the presence and
detailed properties of any of them it is necessary to perform scattering experiments
or to observe their decay products. In this chapter, we shall develop an appropriate
quantum mechanical description of such processes.

9.1

Quantum-Mechanical Description

We begin by developing the appropriate quantum-mechanical tools for describing


the scattering process.

9.1.1

Schr
odinger Picture

Consider a quantum mechanical system whose Schrodinger equation


S (t)i = it |S (t)i
H|

(9.1)

cannot be solved analytically (here we use natural units with h


=1). The standard methods of gaining information on such a system come from perturbation
0 whose
theory. The Hamilton operator is separated into a time-independent part H
Schrodinger equation
0 |S (t)i = it |S (t)i
H

(9.2)

is solvable, and a remainder V called interaction. The total Hamilton operator is


=H
0 + V .
H

(9.3)

0 is called the unperturbed Hamiltonian, the interaction V is often


The operator H
called the perturbation. For the sake of a simple notation we have omitted the hats on
top of all operators without danger of confusion. The time-dependent state |S (t)i
carries a subscript S to indicate the fact that we are dealing with the Schrodinger
S (p, x, t) whose matrix elements between states
picture. Observables are operators O
yield transition amplitudes. The interaction will be assumed to be of finite range.
644

645

9.1 Quantum-Mechanical Description

This ecludes, for the moment, the most important interaction potential, the Coulomb
potential. As an example, we may assume the unperturbed Hamiltonian to describe
a particle in a Rosen-Morse potential well
hx|V |xi = V (x) =

const
.
cosh2 |x|

(9.4)

If the constant is sufficiently negative, the system has one or more discrete bound
states. The important Coulomb potential does not fall directly into the class of
potential considered. It must first be modified by multiplying it with a very small
exponential screening factor. The consequences of the infinite range which arises
when removing this factor will have to be discussed separately. The interaction V
may in general depend explicitly on time.
0 will be
The results will later be applied to quantum field theory. There, H
a sum of the second-quantized Hamilton operator of all particles involved, and V
some as yet unspecified short-range interaction between them. If the particles are all
massive, all forces will be of short range. Moreover, the vacuum state is a discrete
state which is well separated from all other states by an energy gap. The lowest
excited state contains a single particle with the smallest mass at rest. This mass is
the energy gap.

9.1.2

Heisenberg Picture

The other, equivalent, description of the system due to Heisenberg works with timeindependent states which are equal to the Schrodinger states |S (t)i at a certain
fixed time t = t0 , which we shall take to be t0 = 0, for simplicity:
|H i |S (0)i = U (t, 0)1 |S (t)i.

(9.5)

When using time-independent states, the time dependence of the system is carried
by time-dependent Heisenberg operators
H (t) = U (t, 0)1 O
S U (t, 0).
O

(9.6)

0) is the unitary time evolution operator introduced in Section 1.6 which


Here U(t,
governs the motion of the Schrodinger state. It has the explicitly form [recall (1.268)]


0) = T ei
U(t,

Rt
0

)
dt H(t

(9.7)

has no explicit time dependence, which we shall assume from now on, then
If H

U (t, 0) has the explicit form U (t, 0) = eiHt/h and satisfies the unitarity relations
1 (t, 0) = U (t, 0) = U (0, t).
U

(9.8)

Then the relation (9.6) becomes


h iHt/
h
H (t) = eiHt/
O
OS e
.

(9.9)

646

9.1.3

9 Scattering and Decay of Particles

Interaction Picture

As far as perturbation theory is concerned it is useful to introduce yet a third


description called the Dirac or interaction picture. Its states are related to the
previous ones by

0)|H i.
|I (t)i = eiH0 t |S (t)i = eiH0 t U(t,

(9.10)

They have the property that in the absence of interactions they become time independent and coincide, in this case, with the Heisenberg states (9.5). The interactions,
however, drive I (t) away from H (t).
When using such states, the observables in the interaction picture are

I (t) = eiH 0 t O
S eiH 0 t = eiH 0 t eiHt
O
OH eiHt eiH0 t .

(9.11)

These operators coincide, of course, with those of the Heisenberg picture in the ab 0 |(t)i
sence of interactions. By definition, the unperturbed Schrodinger equation H
I (t)
can be solved explicitly implying that the time dependence of the operators O
d
0, O
I (t)]
OI (t) = [H
dt

(9.12)

S is explicitly time dependent,


is completely known. If the Schrodinger operator O
this becomes
d
0, O
I (t)] + [O
S (t)]I ,
OI (t) = [H
(9.13)
dt
where
S (t)]I eiH 0 t O
S (t)eiH 0 t .
[O

9.1.4

(9.14)

Neumann-Liouville expansion

A state vector in the interaction picture moves according to the following equation
of motion

it |I (t)i = it eiH0 t |S (t)i


0 |I (t)i + eiH 0 t He
iH 0 t |I (t)i
= H
= VI (t)|I (t)i

(9.15)

where

VI (t) eiH0 t VS eiH0 t

(9.16)

is the interaction picture of the potential VS . It is useful to introduce a unitary time


evolution operator also for the interaction picture. It determines the evolution of
the states by
|I (t)i = UI (t, t0 )|I (t0 )i.

(9.17)

647

9.1 Quantum-Mechanical Description

Obviously, UI (t, t0 ) satisfies the same composition law as the previously introduced
t0 ) in the Schrodinger picture:
time translation operator U(t,
I (t , t0 ).
UI (t, t0 ) = UI (t, t )U

(9.18)

Using the differential equation (9.15) for the states we see that the operator

UI (t, t0 ) satisfies the equation of motion


I (t, t0 ).
it UI (t, t0 ) = VI (t)U

(9.19)

The equation of motion (9.19) can be integrated resulting in the NeumannLiouville expansion [recall (1.200), (1.201)]
UI (t, t0 ) = 1 i

(i)2
T
dt1 VI (t1 ) +
2!
t0

T exp i

t
t0

t0

dt1 dt2 VI (t1 )VI (t2 ) + . . .

dt VI (t ) .

(9.20)

The expansion holds for t > t0 and respects the initial condition UI (t0 , t0 ) = 1.
Note that the differential equation (9.19) for UI (t, t0 ) implies that UI (t, t0 ) satisfies an integral equation
UI (t, t0 ) = 1 i

t0

dt VI (t )UI (t , t0 ).

(9.21)

Indeed, we may solve this equation by iteration, starting with UI (t , t0 ) = 1 which


yields the lowest approximation
UI (t, t0 ) = 1 i

t0

dt1 VI (t1 ).

(9.22)

Reinserting this into (9.2), we obtain the second approximation


UI (t, t0 ) = 1 i

(i)2
T
dt1 VI (t1 ) +
2!
t0

t
t0

dt1 dt2 VI (t1 )VI (t2 ).

(9.23)

Continuing this iteration we recover the full Neumann-Liouville expansion (9.20).


t0 ) = U 1 (t0 , t)
For t < t0 , we may use the relation (9.8) according to which U(t,
and the expansion is again applicable. The time-ordering operator T makes sure
that the operators VI (ti ) in the expansion appear in chronological order with all
later VI (ti ) standing to the left of earlier ones.
has no explicit time dependence, the Schrodinger state
If V and therefore H
evolves via the simple exponential time evolution operator

|S (t)i = U (t, t0 )|S (t0 )i = eiH(tt0 ) |S (t0 )i.

(9.24)

Combining this with (9.10) we see that

UI (t, t0 ) = eiH0 t eiH(tt0 ) eiH0 t0 .

(9.25)

648

9 Scattering and Decay of Particles

If the potential V depends explicitly on time, this has to be replaced by [recall (9.7)]


UI (t, t0 ) = eiH0 t T e

Rt

t0

)
dt H(t

= eiH0 t U (t, t0 )eiH0 t0 .

eiH0 t0
(9.26)

this relation shows that UI (t, t0 ) satisfies the same uniFor a time-independent H,

tarity relations (9.8) as U (t, t0 ):


I (0, t).
UI1 (t, 0) = UI (t, 0) = U

(9.27)

The Heisenberg representation (9.6) of an arbitrary operator is obtained from


the interaction representation (9.11) via
I (t)eiH 0 t U (t, 0)
(t) = U 1 (t, 0)eiH 0 t O
O
H
I (t)UI (t, 0)
= U 1 (t, 0)O
I

(9.28)

I (t) has a
We have observed above that the operator in the interaction picture O
particularly simple time dependence. Its movement is determined by the unperturbed equation of motion (9.12). Thus, (9.28) establishes a relation between the
H (t) in the presence of
complicated time dependence of the Heisenberg operator O

interaction and the simple time dependence of OI (t) which reduces to the Heisenberg
operator in the absence of an interaction.

9.1.5

Mller Operators

Consider now a scattering process in which two initial particles approach each other
from a large distance outside the range of their interactions. This implies that their
states at a very early time t obey the unperturbed Schrodinger theory, i.e.,
that their wave function I (t) is time independent. The same thing will be true a
very long time after the scattering has taken place. Let us study the behavior of the
operator UI (t, t0 ) in the two limits of large positive and negative time arguments.
In order to make all expressions well defined mathematically it is convenient to
introduce a simple modification of the potential by multiplying it with a switching
factor e|t| :
V e|t| V V (t),

VI (t) e|t| VI (t) VI (t),

(9.29)

with an infinitesimal parameter . As far as physical observations are concerned,


such a factor must have little relevance since can be chosen so that V remains
unchanged over any finite interval of time. An immediate consequence of the prescription (9.29), however, is that for t , the state vector in the interaction
picture becomes time independent, due to (9.15): Let us denote the limiting states
as | in i, i.e.,
out

|I (t)i |i.
t

(9.30)

649

9.1 Quantum-Mechanical Description

Using the time evolution operator, the limit (9.30) may be written as
UI (t, 0)|I (0)i |
t

in
out

(9.31)

I becomes a constant in the limit t . Inverting this relation


i.e., the operator U
we may write
|I (0)i = UI (0, )|

in
out

() |
i

in
out

i.

(9.32)

The operators
() UI (0, )

(9.33)

were first studied extensively by Mller1 and are named after him. Their most
important property is the following:

() =
() H
0,
H

() H
=H
0
() ,

(9.34)

To derive this, we consider, for simplicity, only the most common situation that V
has no explicit time dependence except for the very slow switching factor (9.29).
Then we multiply the explicit representation (9.25) of the time-evolution operator

by a factor eitH and find

ta t)eitH 0 .
eitH U (0, ta ) = eitH eita H eita H0 = ei(tta )H ei(ta t)H0 eitH0 = U(0,
(9.35)
In the limit ta , this yields

eitH U(0,
) = U (0, )eitH0 .

(9.36)

The time derivative of this at t = 0 yields precisely the first equation in (9.34). The
second follows by hermitian conjugation.
Let | in i be a steady-state solution of the time-independent unperturbed
out
Schrodinger equation
0 |
H

in
out

i = E|

in
out

i.

(9.37)

Due to (9.34) [or directly (9.36)], the interacting state |I (0)i solves the full
Schrodinger equation with the same energy
I (0)i = E|I (0)i.
H|

(9.38)

In the laboratory, one does not really observe steady states but wave packets, which
can be obtained from superpositions of steady state solutions with different momenta. The packets of free particles approach each other and enter into the range
1

C. Mller, Kgl. Danske Vidensk. Selsk. Mat.Fys. Medd 23, No. 1, (1945).

650

9 Scattering and Decay of Particles

of the interaction potential. Also there, the total energy remains the same, due to
energy conservation.
The amplitude for the scattering process is found as follows. Let |ain i be a
complete set of incoming eigenstates of the unperturbed Schrodinger equation
0 |ain i = Ea |aini
H

(9.39)

where a denotes the collection of all quantum numbers of these states. By applying
the interaction operator UI (t, ), the states |ain i, they are transformed into the
time-dependent |aI (t)i. After a long time, these develop into time-independent
states |aout i. The energies Ea remain, of course, unchanged.
Let us analyze the outgoing states |aout i
|aout i = UI (, ) |ain i

(9.40)

with respect to the incoming waves |bin i. The scalar products hbin |aout i are obviously
equal to the matrix elements of UI (, ):
I (, ) |ain i.
hbin |aout i = hbin |U

(9.41)

The right-hand side is defined as the scattering matrix or S-matrix:


Sba = hbin |UI (, ) |ain i.

(9.42)

The same name is also often used sloppily for the scattering operator
I (, )
S U

(9.43)

itself. This can be written as a product of the two Mller operators (9.33):
()
(+) .
S=

(9.44)

I (t, t0 ) being a unitary operator, the S-matrix is unitary and


Note that with U
satisfies:
SS = S S = 1.

(9.45)

The Mller operators, however, are not unitary. As we shall see in the next section,
() are inverse to
() only if H
0 has no bound states and then only
the operators
under multiplication from the left-had side:
()
() = 1,

(9.46)

whereas from the right-hand side they yield a projection operator onto the subspace

of continuous states of the Hamiltonian H:


() = Pcontinuum .
()

(9.47)

Only in an infinite-dimensional Hilbert space makes it possible that the left inverse
is not equal to the right inverse.

651

9.1 Quantum-Mechanical Description

9.1.6

Lippmann-Schwinger equation

The solution (9.32) of Eq. (9.38) can usually not be given analytically but only in
form of a perturbation series. This is found by inserting the Neumann-Liouville
expansion (9.20) into Eq. (9.32):
0

|I (0)i = |in ii

dt VI (t )|in i+(i)2

dt

dt VI (t )VI (t )|in i + . . . ,
(9.48)

With the help of (9.16) and inserting the damping factor as in Eq. (9.29) we see
that the second term is equal to
i

dtet eiH0 t V eiH0 t |in i.

(9.49)

0 with energy E, and V is independent of


By assumption, |in i is an eigenstate of H
time. Thus we can perform the integral and obtain
i

dtet ei(H0 E )t V |in i =

1
V |in i.
0 + i
EH

(9.50)

Treating the other expansion terms in (9.48) likewise we arrive at the expansion
|I (0)i = |in i +

1
1
1
V |in i +
V
V |in i + . . . .

E H0 + i
E H0 + i E H0 + i
(9.51)

In the second denominator we have replaced 2i by i since is infinitesimally small


and it actual size is irrelevant. The infinite sum is recognized as an iterative solution
of the so-called Lippmann-Schwinger equation 2
|I (0)i = |in i +

1
V |I (0)i.

E H0 + i

(9.52)

The fact that this state solves Eq. (9.38) is easily verified by multiplying both sides
0 , yielding
by E H
0 )|I (0)i = (E H
0 )|in i + V |I (0)i.
(E H

(9.53)

The first term on the right-hand side vanishes due to (9.37). The remainder coincides
with Eq. (9.38).
Comparing (9.52) with (9.32), we identify the Mller operator as
(+) = UI (0, ) = 1 +

1
V .
0 + i
EH

For a discussion see M. Gell-Mann and M. Goldberger, Phys. Rev. 91, 398 (1953).

(9.54)

652

9 Scattering and Decay of Particles

Because of this notation one often writes the scattering state |I (0)i in (9.52) as
| (+) i and the equation itself as
| (+) i = |in i +

1
V | (+) i.

E H0 + i

(9.55)

There also exists a Lippmann-Schwinger equation yielding a different set of interacting states | () i from the outgoing free state |out i:
| () i = |out i +

1
V | () i,

E H0 i

(9.56)

1
V .

E H0 i

(9.57)

where the Mller operator is


() = UI (0, ) = 1 +

The only difference with respect to (9.54) is the sign of the i-term.
The operators
0 (E)
G

ih
,
0 + i
EH

G(E)

ih
,
+ i
EH

(9.58)

are the resolvents of the free and interacting Hamiltonians, respectively [recall
(11.8)]. The second is related to the first by the equation

0 (E) i G
0 (E)V G(E).

G(E)
=G
h

(9.59)

h from the
This can easily be verified by multiplying it with [G(E)]
= (E H)/i
right:

i
i
1
0 (E)V G(E)

0 (E) i G
1= G
[G(E)]
= G
0 (E)(E H) G0 (E)V . (9.60)
h

The implicit equation (9.59) for G(E)


can be solved iteratively by the geometric
series
 2

0 (E)V G
0 (E) + i

0 (E) i G
G(E)
=G
h

0 (E)V G
0 (E)V G
0 (E) + . . . . (9.61)
G

In terms of the free resolvent, the Mller operators (9.54) and (9.57) can be
written as
0 (E)V ,
(+) = 1 i G

() = 1 i G
0 (E )V .

(9.62)

653

9.1 Quantum-Mechanical Description

9.1.7

Discrete States

The Mller operator has nontrivial properties with respect to the discrete spectrum
0 and the perturbed Hamiltonian H.
Let us first
of the unperturbed Hamiltonian H
understand the difference between the quasi-unitarity relations (9.46) and (9.47).
0 , similarly by
We denote by |a i the continuum and by | i the bound states of H
which solve the Lippmann|a() i the continuum and by | i the bound states of H
Schwinger equations (9.55) and (9.56), respectively, the latter being independent of
which of the two Lippmann-Schwinger equations is used. These states satisfy the
completeness relations
X
a

|a iha | +

| ih | = 1,

|a() iha() | +

| ih | = 1,

(9.63)

Only the continuum states carry the distinctive label () since only these depend
on the small quantity in the Lippmann-Schwinger equation (9.55) and (9.56). The
bound states are uniquely determined by the condition of quadratic integrability.
() transform the continuum states |a i into scattering
The Mller operators
()
states | i. Thus we can expand
() = UI (0, ) =

X
a

() = UI (0, ) =

|a() iha |,

X
a

|a iha() |.

(9.64)

From these expansions, we extract the following important properties of the Mller
operators:
() |a() i = |a i,

() | i = 0.

(9.65)

The second relation implies that


() Pbd = 0.

(9.66)

We further find
()
() =

X
a,b

()

|a iha() |b iha | = 1

0
| ih | = 1 Pbd
,

(9.67)

0 . In
the operator on the right-hand side it the projection on the bound states of H
the opposite order, the product yields
()
() =

X
a,b

()

|a() iha |b ihb | = 1

| ih | = 1 Pbd , (9.68)

The different orders of the products


where Pbd projects onto the bound states of H.

differ by the projection onto the bound states of H.


For the S-matrix, on the other hand, this dependence on the order disappears
and we find
()
(+) ] [
()
(+) ] =
(+)
()
()
(+) =
(+) 1 Pbd
(+)
S S = [
0
= 1 Pbd
,

(9.69)

654

9 Scattering and Decay of Particles

and
()
(+)
(+)
() =
() 1 Pbd
() =
()
() = 1.
SS =


(9.70)

The projection operator Pbd has dropped out because of (9.66).


By definition, the S-matrix is calculated only between states in the subspace of
0 . In this subspace, the
the Hilbert space formed by the continuous eigenstates of H
0
projection operator Pbd
vanishes and we find S S = 1 and SS = 1, in either order,
showing that the S-matrix is unitary.
From the property (9.34) of the Mller operator we derive immediately that S
0 , S] = 0:
does not change the energy of the incoming continuum states since [H
0S = H
0
()
(+) =
() H

(+) =
()
(+) H
0 = SH
0.
H

(9.71)

0 , the size of is irFor scattering states which all lie in the continuum of H
relevant at all times as long as it is very small and has a fixed sign. In contrast,
0 has a singular -dependence in the limit of small . In
a discrete eigenstate of H
order to understand this we observe that for any short-range potential, the two scattering particles lie practically all the time outside of each others range. Only for
a small time interval do they interact. The potential is not felt for large positive
and negative times where the wave packets are separated by a large distance. The
switching factor is therefore physically irrelevant for sufficiently small . It cannot
influence the physics of the system. For bound states, however, the particles interact
with each other all the time with the same strength. In order to understand this
consider a specific discrete state, assuming it to be well separated from the lower and
higher states by a finite energy gap . Then, for , the time dependence of
V (t) = e|t| V is so slow that the interaction is, over a long time interval, incapable
of causing transitions from the discrete state to a neighbor state. Such a slow time
dependence is called adiabatic. For an adiabatically time-dependent potential V (t),
the discrete state |(t)i is at any time t0 to a very good approximation the solution
of the time-independent Schrodinger equation containing the time-independent potential V (t0 ). The energy E(t) of the discrete state |(t)i is at any finite time equal
to E. Only for very large times |t| > 1/ does the interaction become so weak
0 , and the energy E(t) goes against
that the state |(t)i becomes an eigenstate of H
0.
the associated eigenvalue E0 of the unperturbed Hamilton operator H

9.1.8

Gell-Mann --Low Formulas

The Mller operators relate the wave functions |i of the interacting system which
satisfy the Schrodinger equation
0 + V )|i = E|i,
(H

(9.72)

to the corresponding unperturbed wave function |0 i satisfying


0 |0 i = E0 |0 i.
H

(9.73)

655

9.1 Quantum-Mechanical Description

Formally, this correspondence can be defined uniquely by multiplying V by a coupling constant g and varying g from g = 0 where the system is free to g = 1, where
it is fully interacting. The energy E(g) and the associated discrete state |(g)i are
continuous functions of g interpolating between E0 and E, and between |0 i and
|i, respectively.
() of Eq. (9.33) associated with the potential VI
The two Mller operators
carry a nondegenerate the discrete state |0 i into two solutions of the interacting
Schrodinger equation (9.72) as follows (first Gell-Mann --Low formula):3
1
| () i = lim UI (0, ) |0 i () .
0
z

(9.74)

To have a finite limit, the transformed state has been divided by a normalization
factor z() defined by the expectation value
I (0, ) |0 i.
z() h0 |U

(9.75)

The phase determines the energy shift via the formula (second Gell-Mann --Low
formula)
E = E E0 = lim g
0


.
g

(9.76)

In order to prove the statements (9.74) and (9.76), let us denote the states on the
right-hand side of (9.74) without the normalization factors by
I (0, ) |0 i.
| () i U

(9.77)

These states satisfy for small but fixed the Schrodinger equation
0 + V (0) | () i = E | () i,
H

(9.78)

where E is some -dependent energy. The proof is based on the observation that
in front of the state |0 i, the -dependent Mller operators
() U (0, )

(9.79)

satisfy the obvious commutation rule


0, U
I (0, ) |0 i = H
0 | () i E0 | () i.
H

(9.80)

We now expand UI (0, ) according to (9.20) as




ig
UI (0, ) = T e

=
3

R0

(ig)n
n!
n=0

dt et VI (t)

dt1

(9.81)
Z

dtn e

Pn

M. Gell-Mann and F. Low, Phys. Rev. 84 , 350 (1951).

t
i=1 i

T VI (t1 ) VI (tn ) .


656

9 Scattering and Decay of Particles

The operators VI (t) in the interaction picture have the time dependence (9.12) which
yields
h
i
d
0 , VI (t) .
VI (t) = i H
dt

(9.82)

Using (9.81) and (9.82) we find


h

0, U
(0, )
H
I

= i

Pn
0
(ig)n 0
dtn e i=1 ti
dt1
n!

n=0
n

X


T VI (t1 ) VI (ti ) VI (tn ) .


i=1 ti
Z

(9.83)

Since the integrand is symmetric in the time variables, we may replace the sum
Pn
i=1 /ti by n-times the time derivative /tn . Taking the derivative /tn to the
left of the switching factors eti , the integral becomes

i
0
0
0
h Pn ti 
(ig)n
dtn
dtn1
e i=1 T VI (t1 ) VI (tn )
dt1
n
n!
tn

n=0
Z
Z

n
P


0
0
X (ig)
n
+i
dtn e i=1 ti T VI (t1 ) V I (tn ) .
dt1
(9.84)
n
n!

n=0

The second sum is equal to


ig


U (0, ) .
g I

(9.85)

The first sum contains integrals over a pure derivatives which can be replaced by
surface terms. Moreover, since the integrand vanishes at tn = , only the tn = 0
-term contributes. The time tn = 0 is the latest of all time variables. Thus it can
be moved to the left of the time-ordering symbol, and we arrive at the equation
h

i
0, U
I (0, ) = g VI (0)U
I (0, ) + ig UI (0, ) .
H
g

(9.86)

In a similar way we derive


i
(0, +) ig U (0, +) .
0 , U (0, +) = g VI (0)U
H
I
I
g I

(9.87)

An important property of these formulas is that in the presence of a level shift, the
second term on the right-hand side proportional to is nonzero in the limit 0.
To see this, take Eq. (9.86) and rewrite it in a slightly different way as
U (0, ) =
H
I

0 + g VI (0) U
(0, )
H
I

I (0, ) H
0 ig UI (0, ) .
= U
g

(9.88)

657

9.1 Quantum-Mechanical Description

If this equation is applied to the state |0 i, it shows that the state | () i of (9.77)
satisfies the modified Schrodinger equation
!

() i = E0 ig | () i.
H|
g

(9.89)

If we were to take the limit 0 carelessly, we would arrive at the equation


() i = E0 | () i and erroneously conclude that |i were solutions of the fully
H|
interacting Schrodinger equation with energy E0 . For the continuum states, this
conclusion is indeed true and we have
UI (0, ) = U
I (0, ) H
0,
H

(9.90)

rather than (9.88), as stated before in Eq. (9.34). For the discrete states, the energy
is in general shifted by the interaction. Hence the derivative g(/g)| () i must
diverges like 1/ such as to give a finite contribution in (9.89). The origin of the
divergence is a diverging normalization of the state | () i. Thus is divergence is
eliminated by forming the states on the right-hand side of (9.74):
|() i | () i

1
()

(9.91)

These have a definite limit for 0. For these states, Eq. (9.89) implies
() i
H|

= E0 ig log z() |() i.


g

(9.92)

The level shifts caused by the interaction are given by the formula
E = lim ig
0

log z() .
g

(9.93)

Note that in the limit 0, the two states |() i coincide. This is seen as
(, ) |0 i.
follows. The state | (+) i develop after an infinitely long time into U
I
0 due to the commutation relation (9.71).
There it becomes again an eigenstate of H
(+)
Thus it must coincide with |0 i up to some overall factor ei (recall that |0 i is
nondegenerate by assumption). Similarly, we may develop the state |0 i backwards
in time via | () i into UI (, ) |0 i which is equal to the initial state rather than

|0 i, we see directly that |i |i+ ei . When forming | i from |i , the pure


phase between |0 i and UI (, ) |0 i cancels. Thus the two states | i must
indeed be identical. The phase associated with the total time evolution operator
UI (, ), defined by

ei h0 |UI (, ) |0 i,

(9.94)

is obviously related to the previously introduced phases by the relations


= + = 2 = 2+ .

(9.95)

658

9 Scattering and Decay of Particles

This equation can also be written in terms of matrix elements as


(, ) |0 i = h0 |U
(, 0) |0 ih0 |U
(0, ) |0 i,
h0 |U
I
I
I

(9.96)

which may come somewhat as a surprise. In general, there is certainly an identity


containing a sum over a complete set of orthonormal eigenstates in the middle:
h0 |UI (, ) |0 i =

X
n

(0, ) |0 i.
h0 |UI (, 0) |n ihn |U
I

(9.97)

It is the adiabatic slowness of the switching-on process of the interaction which does
not permit a transition to an excited intermediate state making (9.96) valid.
The level shift can also be expressed in terms of the -dependent S-matrix
S = UI (, ) = UI (, 0) UI (0, ) .

(9.98)

Indeed, by multiplying Eq. (9.89) for | (+) i by h(+) | from the left we find
E0 | () i = Eh() | () i = igh() |g | () i,
h() |H

(9.99)

such that with the abbreviation g /g:


E = ig

h() |g | () i

(9.100)

()

h | () i
h0 |UI (, 0) g UI (0, ) |0 i
= ig
.
h0 |UI (, 0) UI (0, ) |0 i

(9.101)

Similarly, taking the equation for h () | and multiplying it by |() i from the right
we find
h

h0 | g UI (, 0) UI (0, ) |0 i
E = ig
.
I (, 0) UI (0, ) |0 i
h0 |U

(9.102)

Adding both results for the upper sign gives



E = i g logh0 |S (, ) |0 i.
2 g

(9.103)

Note that the matrix element h0 |S (, ) |0 i is a pure phase ei . This


is seen as follows. The state | (+) i develop after an infinitely long time into
0 due to the comUI (, ) |0 i. There it becomes again an eigenstate of H
mutation relation (9.71). Thus it must coincide with |0 i up to some overall factor
ei (recall that |0 i is nondegenerate by assumption). Hence we arrive at the level
shift formula

E = g .
2 g

(9.104)

659

9.2 Scattering in External Potential

One sometimes finds this result sated the opposite direction:


= 2

dg
E(g ).

(9.105)

To lowest order in g, where E g, the formula becomes


 
g
2 E + O g 2 .

(9.106)

This formula is sometimes used for a lowest-order calculations of the change of mass
of a particle due to interactions.
Actually, in applications the exponential switching factor is somewhat difficult
to handle. A much more direct way of cutting off the interaction consists in a step
function along the time axis, i.e., by assuming the total interaction time T to be
finite. The connection between T and is found from the correspondence between
the integrals
Z

dte|t| =

T /2

T /2

dt = T.

(9.107)

Then we may write simpler


= T

9.2

dg
E(g )
g

(9.108)

Scattering in External Potential

After this development it is straightforward to derive scattering amplitudes for a


particle in an external potential V (x) of a finite range.

9.2.1

The T - Matrix

Imagine a scattering center with a region of nonzero potential concentrated around


the origin. A wave packet of a given average momentum p and energy E approaches
this region along the x-axis from negative infinity. As long as it is far from the
scattering center it follows the free-particle Schrodinger equation
0 |(t)i = it |(t)i,
H

(9.109)

0 = p2 /2M. As the particle approaches the


with the unperturbed Hamiltonian H
= H
0 + V . In the
potential, the state will be driven by the full Hamiltonian H
interaction representation we study the time evolution of a state which diagonalizes
0 and see how it changes due to the interaction
the unperturbed Hamiltonian H

VI exp (|t|). The switching factor exp (|t|) enables us to remove the scattering
potential long before the scattering happens. This frees us from the necessity of
assuming the incoming wave to be localized in a packet long before the scattering

660

9 Scattering and Decay of Particles

process. Instead, we can take it to be a pure plane wave, a momentum eigenstate


|pi solving the time-independent Schrodinger equation
0 |pi = Ep |pi,
H

(9.110)

with an energy Ep = p2 /2M. The incoming wave function is


1
hx|pi = eipx/h .
V

(9.111)

The particle has the same probability everywhere in space, also in the scattering
region, but since VI is switched off at large |t|, this will presents no problem in the
description of the scattering process. The amplitude for the incoming state |pi to
be scattered into another eigenstate with momentum |p i is given by the matrix
elements

hp |S|pi
= hp |UI (, ) |pi.

(9.112)

Decomposing UI (, ) into the product of Mller operators,


I (, ) = UI (, 0) UI (0, ) =
()
(+) ,
U

(9.113)

we can also write

hp |S|pi
= hp () |p(+) i

(9.114)

() |pi.
|p() i

(9.115)

where |p() i are the states


The time evolution does not change the energy under the assumption of an infinitesimally small , such that these states satisfy the full Schrodinger equation with the
same energy E = Ep = p2 /2M:
() i = Ep |p() i
H|p

(9.116)

It is useful to take advantage of this property of the scattering process by rewriting


the matrix element (9.112) in another way using the explicit operator (9.25):

hp |S|pi
=
=
=

lim hp |UI (t2 , 0)|p(+) i

t2

lim hp |eiH0 t2 eiHt2 |p(+) i

t2

lim ei(Ep Ep )t2 hp |p(+) i.

(9.117)

t2

Then we use the Lippmann-Schwinger equation (9.55) for |p(+) i and have
"

hp |S|pi
= lim ei(Ep Ep )t2 hp |pi +

t2

1
hp |V |p(+) i .
Ep Ep + i

(9.118)

661

9.2 Scattering in External Potential

The first term in brackets is nonzero only if the momenta are equal, p = p, in which
case also the energies are equal, Ep = Ep and the right-hand side is unaffected by
the limit t2 in the prefactor.
We shall imagine the system to be enclosed in a box of a large volume V (not
to be confused with the potential V ). Then the momenta are discrete and the first
term is simply
p,p .
It describes the amplitude for the direct beam to appear behind the scattering region,
containing all the unscattered particles. The second term descibes scattering. Here
the limit t2 in the prefactor is nontrivial. If E 6= E, it oscillates so fast that it
can be set equal to zero, by the Riemann-Lebesgue lemma (see Ref. [3] on p. 383).
For E = E it is identically equal to unity so that

ei(E E)t2
lim
=
t2 E E + i

0,
i/,

E =
6 E,

E = E.

(9.119)

Such a property defines a -function in the energy. Only the normalization needs
adjustment, which is done as follows:

i ei(E E)t2
lim
= (E E).
t2 2 E E + i

(9.120)

Indeed, if we integrate the left-hand side over a smooth function f (E ) and set
E E + /t2 , then the E -integral can be rewritten as

i Z d
ei
f (E + /t2 ) .
2 t2 /t2 i

(9.121)

In the limit of large t2 , the function f (E) can be taken out of the integral, and the
contour of integration can then be closed in the upper half-plane giving et2 . Since
is an infinitesimal quantity, this is the same result as would be obtained from the
right-hand side of (9.120). The -function (9.29) ensures the conservation of energy
in the scattering process.
Another way to verify formula (9.120) is to rewrite the left-hand side as
1
2

t2

dtei(E Ei)t ,

(9.122)

which obviously tends to (E E) in th limit t2 .


Let us also express the Lippmann-Schwinger equation in terms of wave functions.
Multiplying (9.55) by the local states hx| from the left, and identifying the interacting
states | (+) i with the states |p(+) i in Eq. (9.115) arising from an incoming particle
wave of momentum p, the left-hand side is the full Schrodinger wave function
(+) (x) hx|p(+) i,

(9.123)

662

9 Scattering and Decay of Particles

which is an eigenstate of Eq. (9.38):


(+) i = E|p(+) i.
H|p

(9.124)

The first term on the right-hand side is the plane wave (9.111). The second term in
(9.55)
hx|pisc hx|

1
V |p(+) i

E H0 + i

(9.125)

is the scattering state. ItR can be rewritten in a more convenient form, after inserting
a completeness relation d3 x |x ihx | = 1, as an integral
hx|pisc =

d3 x hx|

1
|x iV (x )hx |p(+) i.
0 + i
EH

Another way of writing this result is


hx|pisc =

0 (E)|x iV (x )hx |p(+) i


d3 x hx|G

(9.126)

0 (E)|x i are the matrix elements of the free-particle resolvent in


where hx|G
Eq. (9.58):
0 (E)|x i = hx|
hx|G

1
|x i.

E H0 + i

(9.127)

Thus the result of the full Lippmann-Schwinger is the wave function


hx|p(+) i = hx|pi + hx|pisc .

(9.128)

It solves the interacting Schrodinger equation (9.124) and shows explicitly the free
incoming wave plus the scattered wave. Asymptotically, the scattered wave is usually
an outgoing asymptotically spherical wave emerging from the scattering center.
We are now prepared to introduce the so-called T -matrix as the quantity
(+) |pi
hp |T (E)|pi hp |V |p(+) i = hp |V

(9.129)

as the T -matrix, so that the S-matrix decomposes into a direct beam contribution
plus an energy conserving scattering contribution as follows

hp |S|pi
= hp |pi 2i(E E)hp |T (E)|pi.

(9.130)

In terms of T, the Lippmann-Schwinger equation (9.55) reads


|p(+) i = |pi +

1
1
V |p(+) i = |pi +
T(E)|pi,

E H0 + i
E H0 + i

(9.131)

663

9.2 Scattering in External Potential

Equation (9.62) for the Mller operator implies that the T matrix obeys an implicit
equation
T (E) = V + V

1
T (E).

E H0 i

(9.132)

The matrix elements of this form an integral equation for the T -matrix.
Note that T (E) can be expressed in terms of the Mller operator (9.54) as
(+) .
T = V

(9.133)

If we multiply the state vectors (9.131) by ket vectos hx| from the ledt, we obtain
the position representation of the Lippmann-Schwinger equation:
hx|p(+) i = hx|pi + hx|

1
1
V |p(+) i = hx|pi + hx|
T(E)|pi, (9.134)
0 +i
0+ i
E H
E H

The first term hx|pi is once more the incoming wave, the second term is the T -matrix
representation of the scattered wave (9.125):
hx|pisc hx|

9.2.2

1
1
V |p(+) i = hx|
T |pi.

E H0 + i
E H0 + i

(9.135)

Asymptotic Behavior

Let us calculate the asymptotic form of the scattered wave (9.135) in the form
(9.126) where it is expressed in terms of the free-particle resolvent in Eq. (9.58):
hx|pi

sc

i
=
h

0 (E)|x iV (x )hx |p(+) i


d3 x hx|G

(9.136)

The matrix elements of the resolvent have the Fourier decomposition


d3 p V
ih
hx|pi
hp|x i
3
(2h)
E Ep + i

Z
eip(xx )/h
d3 p
.
= ih
(2h)3 E p2 /2M + i

0 (E)|x i = ih
hx|G

Introducing pE

(9.137)

2M(E + i), the momentum integral yields

0 (E)|x i = 2M ih
hx|G

2Mi 1 eipE |xx |/h


d3 p eip(xx )/h
=

.
(2h)3 p2 p2E
h
4 |x x |

(9.138)

Since the time dependence of this state is given by the exponential factor eiEt/h ,
this is an outgoing spherical wave emerging from x . Had we chosen the opposite i

in the denominator of (9A.1), the exponential would habe been eipE |xx |/h which
together with the time-depedence eiEt/h would have implied an incoming spherical

664

9 Scattering and Decay of Particles

wave. With the exponent as in (9A.1), the scattered wave (9.136) is composed
of a superposition of outgoing spherical waves emerging from all points x in the
potential V (x ). This is an adaptation of the good-old Huygens principle of optics
to scattering theory in the Schrodinger representation.
The outgoing spherical waves are observed at the point x far away from all points

x in the scattering potential V (x ). There they appear as a common outgoing


spherical wave whose shape is found from (9A.1) by going to the asymptotic regime
of large x where we can approximate
x + . . . .
|x x | r x

(9.139)

so that (9.136) simplifies to


eipE r/h 2M
hx|pi

r
h
2
sc

d3 x eipE x x /h V (x )hx |p(+) i.

(9.140)

Since the scattering is elastic the energy of the incoming particles Ep is equal to the
energy of the outgoing particles Ep . Those arriving at the point x emergy with a
. Thus we
momentum p whose size is equal to that of p and whose direction is x
define the outgoing momentum in the sperical wave as
,
p = pE x

(9.141)

and rewrite the exponent in (9.140) as

eipE x x /h eip x /h .

(9.142)

With the scattered wave (9.140), the total wave function (9.134) may be written
as a sum of an incoming plane wave and an outgoing spherical wave
(+)

hx|p

1
i

r
V

ipx/
h

eipr/h
+
fp p
r

(9.143)

The prefactor of the spherical wave fp p is called the scattering amplitude, and is
given by the integral4 .
2M Z 3 ip x /h
fp p = V 2
dxe
V (x )hx |p(+) i.
h

(9.144)

From the asymptotic behavior (9.143) where the wave function is decomposed into
an incoming plane wave and an outgoing spherical wave. we deduce immediately
that the square of fp p determines the scattering cross section: This argument goes
as follows: The incoming wave has the particle current density j =(i/M) =
v/V . The outgoing sperical wave as a curent density fp p eipr /r V and describes
4

For the phase convention see, for example, the textbook by L.D. Landau and E.M. Lifshitz,
Quantum Mechanics, Pergamon Press, London, 1965.

665

9.2 Scattering in External Potential

x/V r 2 . The ratio between the


radial particle flow with urrent density j = |fp p |2 v
two d/r 2 d = |j |/|j| defines the differential cross section:
d
= |fp p |2 .
d

(9.145)

If the potential is weak, the full wave function hx|p(+) i on the right-hand side can
be approximated by the incoming plane wave hx|pi, and one obtains the scattering
ampltitude in the Born approximation (Born 1926)
f

p p

2M
2
h

d3 x ei(p p)x /h V (x ).

(9.146)

The momentum difference q p p is equal to the momentum transfer of the


scattering process. Thus the scattering ampliude in the Born approximation is,
up to a facto 2M/h2 , equal to the Fourier components of the potential at the
momentum transfer q:
V (q)

d3 x ei(p p)x /h V (x ).

(9.147)

For a potential V (x) = g(x) where V (q) = g, the scattering amplitude in the Born
approximation is
fp p =

M
g.
2h2

(9.148)

In D dimensions the relation (9.146) will be derived in (9.247).

9.2.3

Partial Waves

The scattering amplitudes are often analyzed separately for each angular momentum of the scattered particles. The T -matrix hp |T|pi in Eq. (9.129) has for equal
incoming and outgoing energies Ep = Ep , i.e., |p | = |p| = p an expansion
hp |T |pi =

l
X
(2)3 X
Tl (p)
Ylm (
p )Ylm (
p),
V l=0
m=l

(9.149)

where Ylm (
p) are the spherical harmonics (4D.1) associated with the momentum
:
direction p
Ylm (
p) = Ylm (, ) (1)

"

2l + 1 (l m)!
4 (l + m)!

#1/2

Plm (cos )eim .

(9.150)

They satisfy the orthogonality relation (4D.3), which we shall write in the form
Z

Yl m (
d2 p
p)Ylm (
p) = l l m m ,

(9.151)

666
where

9 Scattering and Decay of Particles


R

denotes the integral over the surface of a unit sphere:


d2 p
Z

d2 p

d cos

d.

(9.152)

It is useful to view the spherical harmonics as matrix elements of eigenstates |l mi


of angular momentum [recall (4.818)] with localized states |
pi on the unit sphere:
h
p|l mi Ylm (
p).

(9.153)

),
h
p |
pi = (2) (
p p

(9.154)

) (cos cos )( ).
(2) (
p p

(9.155)

The latter are orthogonal:


) is the -function on a unit sphere:
where (2) (
p p

With the help of the addition theorem for the spherical harmonics
l
X
2l + 1

Pl (cos ) =
Ylm (
p )Ylm
(
p ),
4
m=l

p
,
cos p

(9.156)

the partial-wave expansion (9.149) becomes


hp |T |pi =

(2)3 X
2l + 1
).
Pl (
p p
Tl (p)
V l=0
4

(9.157)

where Pl (cos ) Pl0 (cos ) are the Legendre polynomials (4D.7).


A similar expansion of the S-matrix (9.130) requires expanding the scalar product hp |pi = p ,p into partial waves. In a large volume V , we replace this by
(2)3 (3) (p p)/V . The -function is decomposed into a directional and a radial
part
(2)3 1
(2)3 (3)
)
(p p) = (2) (
p p
(p p).
(9.158)
p,p =
V
V p2
The directional -function on the right-hand side has a partial-wave expansion which
is, in fact, the completeness relation (4C.1) for the states |l mi,
X
l
X

l=0 m=l

|l mihl m| = 1,

(9.159)

evaluated between the localized states |


pi on the unit sphere. Using the matrix
elements (9.153) and the orthogonality relation (9.154), we find from (9.159):
(2)

) =
(
p p

X
l
X

l=0 m=l

Ylm (
p )Ylm
(
p).

(9.160)

667

9.2 Scattering in External Potential

The radial part of (9.158) be rewritten as


(2)3 1 dE
(2)3 1
(2)3 1

(p p) =
(E E) =
(E E),
2
2
V p
V p dp
V E

(9.161)

where E = Ep = p2 /2M, E = Ep = p2 /2M, p |p|, and


E

p2
= Mp
dE/dp

(9.162)

is proportional to the density of energy levels on the E-axis in a large volume V ,


the density of states being E V /(2)3 . With the help of E we can rewrite the
sum over all momentum states as follows:
X

d3 pV
=
(2)3

d2 p

V
(2)3

dE E .

(9.163)

Thus we find the expansion


hp |pi = p ,p =

X
l
1
(2)3 X
Ylm (
p )Ylm (
p) (E E).
V l=0 m=l
E

(9.164)

In a partial-wave expansion for the S-matrix (9.130), the overall -function for
the energy conservation is conventionally factored out and one writes:

l
X
X
(2)3 1

hp |S|pi
=
(E E)
Sl (p)
Ylm (
p )Ylm (
p),
V E
l=0
m=l

(9.165)

Recalling (9.130), we find the relation between the partial-wave scattering amplitudes
Sl (p) and those of the T -matrix:
Sl (p) = 1 2iE Tl (p).

(9.166)

The potential scattering under consideration is purely elastic. For such processes,
the unitarity property (9.45) of the S-matrix implies
Z

d3 p V
hp |S |p ihp |S|pi = hp |pi.
(2)3

(9.167)

Inserting the partial-wave decomposition (9.165) and using (9.163) as well as the
orthogonality relation (9.151), we derive from this the unitarity property of the
partial waves of elastic scattering
Sl (p)Sl (p) = 1,

(9.168)

implying that Sl (p) is a pure phase factor: unitarity of partial waves


Sl (p) = e2il (p) .

(9.169)

668

9 Scattering and Decay of Particles

For the amplitudes Tl (p) in (9.166) this implies


Tl (p) =

1 il (p)
e
sin l (p).
E

(9.170)

The quantities l (p) are the phase shifts observed in the scattered waves of angular
momentum l with respect to the incoming waves.
In order to show this we use the expansion formula
iz cos

X
il Jl+1/2 (z)(2l + 1)Pl (cos ),
2z l=0

(9.171)

where the Bessel functions Jl+1/2 (z) can be expressed in terms of spherical Bessel
functions jl (z) as
s
2z
jl (z),
(9.172)
Jl+1/2 (z) =

to expand the incoming free-particle wave

X
X
X
1
2l + 1
), (9.173)
hx|pi = eipx =
l (pr)Pl (
xp
l (pr)
Ylm (
x)Ylm (
p) =
4
V
l=0
m=l
l=0

where r |x|, with the partial waves


1
l (pr) = il 4jl (pr).
V

(9.174)

Far from the scattering region, the spherical Bessel functions have the asymptotic
behavior
sin(z l/2)
,
(9.175)
jl (z)
pr
implying for the partial waves
sin(pr l/2)
1
.
l (pr) = il 4
pr
V

(9.176)

We now turn to the partial-wave expansion of the scattered wave (9.135) in


the Lippmann-Schwinger equation. Inserting a complete set of free momentum
eigenstates into the right-hand expression on the right-hand side, we obtain
hx|pi

sc

d 3 p V
1
hx|p i
hp |T|pi.
3

(2)
E E + i

(9.177)

Inserting the partial-wave decompositions (9.157) and (9.173), with (9.174), we find
Z

d 3 p V
1
hx|p i
hp |T|pi
3
(2)
E E + i

Z
X
l
1 X
jl (p r)Tl (p)

l
,
Ylm (
x )Ylm (
p)i 4
=
dp p2
E E + i
0
V l=0 m=l

(9.178)

669

9.2 Scattering in External Potential

so that the total partial waves including the the scattered waves are
ltot (pr)

"

1
= 4il jl (pr) +
V

jl (p r)Tl (p)
dp p
E E + i

(9.179)

For large r, we use the asymptotic form (9.175) to write the second term in the
brackets as
Z

Tl (p )
ei(p rl/2) ei(p rl/2)
dp p
.
E E + i
2ip r

(9.180)

Being at large r, the integral is very fast oscillating in p and can receive a nonzero
contribution only from a neighborhood of the pole term in the energy.5 There we
can approximate
dE
(p p ),
(9.181)
E E
dp
and p p in all terms except for the fast oscillating exponential. The ensuing
integral

Z
1
Tl (p ) ei(p rl/2) ei(p rl/2)

2
dp
p
(9.182)
dE/dp
p p + i
2ip r

can be extended to the entire p -axis with a negligible error for large r. The part
containing the second exponential is evaluated by closing the contour of integration
in the lower half-plane. There the integrand is regular so that the contour can be
contracted to a point yielding zero. In the part with the first exponential we close
the contour in the upper half-plane and deform it to pick up the pole at p = p + i.
Then Cauchys theorem yields

ip r
Tl (p ) ei(p rl/2)ei(p r/2l)
2p2
1
il/2 e
p2 dp
=

T
(p)e
, (9.183)
l
dE/dp
p p + i
2ipr
dE/dp
2p r

so that the total wave function (9.179) becomes


1
eipr
ltot (pr)
4 il jl (pr) 2E Tl (p)
2pr
V
l
i
h
r
1 4i ipr


e (1)l eipr 2iE Tl (p) eipr .
V 2ipr
r

"

(9.184)

Using (9.166) this is equal to


i
r
1 4il h
Sl (p)eipr (1)l eipr .
ltot (pr)

V 2ipr

(9.185)

Inserting the unitary form (9.169) for elastic scattering, this becomes
r
1 4 l il (p) sin(pr l/2 + l (p))
ltot (pr)

ie
.
r
V p
5

This is a consequence of the Riemann-Lebesgue lemma cited in Ref. [3] on p. 383.

(9.186)

670

9 Scattering and Decay of Particles

This shows that l (p) is indeed the phase shift of the scattered waves with respect
to the incoming partial waves (9.176).
At small momenta, the l = 0 (s-wave) phase shift dominates the scattering
process since the particles do not have enough energy to overcome the centrifugal
barrier.
For short-range potentials, the s-wave phase shift diverges for p 0 and the
leading two orders in p may be parametrized as follows
p cot 0 =

1
1
+ reff p2 + . . . .
as 2

(9.187)

where as is called the s-wave scattering length, and reff is called the effective range
of the scattering process.
The direct small-k expansion of 0 is
0 (p) = as p

a2s reff

a3
s p3 + . . . .
3
!

(9.188)

In the limit p 0, only the s-wave scattering length survives which goes to zero
like and the s-wave amplitude is [recall (9.170)]


i
1
2
as h
0 (p) + i02 (p) 03 (p) + . . .
1ias p+(as reff a2s )k 2 +. . . ,
Mp
3
M
(9.189)
implying via (9.157) that

T0 (p)

hp |T |pi (2)3

as
2
4 V

= 2

as 1
,
MV

(9.190)

which by (9.260) is also equal to gR /M.


For angular momentum l, the effective-range formula (9.187)
p2l+1 cot l =

1
1
+ reff l p2 + . . . .
al 2

(9.191)

Although (9.187) and (9.191) are small-p approximations, the range of validity
my extend to quite large momenta if the potential is of short range. It may include
a resonance in the scattering amplitue, in which case one may use the approximatin
fl (p)

4
.
p cot l (p) ip

(9.192)

In general, one may parametrize the S-matrix element in the vicinity of an resonance
by the unitary ansatz
Sl (p) = e2il (p)

E ER iR /2
,
E ER + iR /2

(9.193)

671

9.2 Scattering in External Potential

where ER is the energy of the resonance and R its decay rate. At every resonance
or bound state, l (p) passes through /2 plus an integer multiple of . This is
guaranteed by Levinsons theorem [4] which states that the number of bound states
in a system is equal to the difference [l (0) l ()]/.
We may also write
4
R /2
fl (p)
,
(9.194)
p E ER + iR /2
and

Sl (p) = e2ip aeff l (p) ,


with
aeff l al +

9.2.4

R (E ER )
1
arctan
.
2p
(E ER )2 + 2R /4

(9.195)
(9.196)

Off Shell T - Matrix

In scattering experiments, only matrix elements are observables in which the energy
E is equal to the equal energies Ep p2 /2M = Ep p2 /2M of incoming and
outgoing particles. For mathematical purposes, however, one also defines a T -matrix
for energies E 6= Ep 6= Ep as the matrix elements of the operator
T (E) = V + V

1
T (E).

E H0 + i

(9.197)

This is called the off-shell T -matrix. This can be expressed in terms the resolvents
(9.59) as
0 (E)T (E).
T(E) = V + V G

(9.198)

This implicit equation for T (E) can be solved iteratively leading to the geometric
series
0 (E)V + V G
0 (E)V G
0 (E)V + . . . .
T (E) = V + V G
(9.199)
Comparison with (9.61) shows that

T(E) = V + V G(E)
V .

(9.200)

There are two further equivalent equations:


0 (E)T (E) = G(E)

G
V ,

(9.201)

and
0 (E)T(E)G
0 (E) = G(E)

0 (E).
G
G

(9.202)

The latter can also be written as


0 )G(E)(E

0 ) (E H
0 ).
T (E) = (E H
H

(9.203)

672

9 Scattering and Decay of Particles

This equation gives a useful relation between the matrix elements of the off-shell
T -matrix and those of the resolvents:

0 (E)|pi]
hp |T (E)|pi = (E Ep )2 [hp |G(E)|pi
hp |G

= (E Ep )2 hp |G(E)|pi
(E Ep )hp |pi.

(9.204)

Let us express the right-hand side in terms of the spectral representation of the
resolvents. We take the completeness relation for the states (9.110):
X
p

|pihp| = 1,

(9.205)

0 ). Using the eigenvalue equation (9.110),


and multiply this by the operator 1/(E H
we obtain the spectral representation of the free resolvent:

G(E)
=

X
p

|pihp|
.
E Ep + i

(9.206)

The interacting resolvent G(E)


has a similar representation. Let |ni be the complete
set of eigenstates of the full Hamiltonian (i.e., the set of scattering states |a(+) i and
the set of bound states | i on the right hand of Eq. (9.63)]:

H|ni
= En |ni.
Then the completeness relation X
n

(9.207)

|nihn| = 1

(9.208)

leads to the spectral representation:

G(E)
=

X
n

|nihn|
.
E En

(9.209)

Rewriting E Ep = (E En ) + (En Ep ), then Eq. (9.204) becomes


hp |T(E)|pi =

X
n

X
hp |nihn|pi
+(E Ep )
hp |nihn|pihp|pi .
(En Ep )(E Ep )
E En +i
n
(9.210)
"

The second term vanishes due to the completeness relation (9.208), and can be
dropped. This is true only if the potential has no hard core forbidding completely
the presence of a particle inside some core volume Vcore . In this case, the wave
functions hx|ni vanish identically inside Vcore , and the wave functions are complete
only outside Vcore . Then the second term yields a correction terms to the first term
hp |Tcorr (E)|pi = (E Ep )

Vcore

dD x ei(p p)x ,

(9.211)

673

9.2 Scattering in External Potential

H
0 |ni =
which is zero on shell. Using the identity (En Ep )hp |ni = hp |H

hp |V |ni, the first term in (9.210) takes the form
hp |T(E)|pi =

X
n

(E Ep )

hp |V |nihn|pi
.
E En + i

(9.212)

Replacing further E Ep by (E En ) + (En Ep ), this becomes


hp |T (E)|pi =

X
n

hp |V |nihn|pi+

X
n

hp |V |nihn|V |pi
.
E En + i

(9.213)

Let us now focus attention upon a potential without bound states. Then we
can use for the states |ni the solutions |p(+) i of the Lippmann-Schwinger equation
(9.131) with E = Ep , and using Eq. (9.133) we obtain
hp |T(E)|pi =

X
p

(+)
(+)
hp |V |p ihp |pi+

X
p

hp |V |p (+) ihp (+) |V |pi


.
E Ep + i

(9.214)

Inserting for the states hp (+) | in the first term the right-hand side of the LippmannScwinger equation (9.131), we obtain
hp |T (E)|pi = hp |V |p(+) i


(+)
(+)

(+)
(+)
X
hp
|
V
|p
ihp
|
V
|pi
hp
|
V
|p
ihp
|
V
|pi

. (9.215)
+
+
+ i
+ i
E

E
E

p
p
p
p

Now we observe that the matrix elements hp |V |p(+) i are the half-on-shell matrix
elements of the T -matrix, whose energy E is equal to the energy Ep of the incoming
state |pi:
hp |V |p(+) i = hp |T (Ep )|pi.
(9.216)
The left-hand momentum is arbitrary. Hence we can rewrite Eq. (9.215) as
hp |T(E)|pi = hp |T (Ep )|pi
+

X
p

hp |T(Ep )|p ihp|T(Ep )|p i

1
1
. (9.217)
+

Ep Ep + i E Ep + i

After solving the Lippmann-Schwinger equation (9.131) to find hp |T (E)|pi, this


equation supplies us with the full off-shell T -matrix. For a hard-core potential, we
must add the correction term (9.211). The off-shell T -matrix has a partial-wave
expansion

l
X
(2)3 X
hp |T (E)|pi =
Tl (p , p; E)
Ylm (
p )Ylm (
p),
V l=0
m=l

(9.218)

If we want to calculate the off-shell T -matrix, we usually must do this separately


for each angular momentum. This has been done for various potentials. The result
for a hard-core potential of radius a will be given in Eq. (9.285).

674

9.2.5

9 Scattering and Decay of Particles

Cross Section

The square of the matrix elements of the T -matrix is experimentally observable as


differential cross section. This quantity is defined by the probability rate dP (t) for
= d cos d around the
an incoming particle to be scattered into solid angle d
= (sin cos , sin sin , cos ), divided by the incoming particle current
direction p
density j, which is the rate of incoming particles per area:
dP (t) 1
d
=
.
d
d j

(9.219)

The total rate of scattered particles is given by


=

d
.
d

(9.220)

The two quantities, which have the dimension of an area, are determined by the
matrix elements of the T -matrix, as we shall now derive.
The time-dependent version of the incoming wave functions (9.173) is
1
hx|p, ti = p (x, t) = ei(E0 tpx) .
V

(9.221)

The associated incoming current density is


j=

1
1 p
1
(x, t) (x, t) =
= v,
2Mi
V M
V

(9.222)

the factor 1/V reflecting the normalization hp |pi = p,p of one particle per total
volume. In order to find the rate of outgoing particles, we form the time derivative
d
dP (t)
I (t, t0 )|pi|2 = d hp |UI (t, t0 )|pihp |U
I (t, t0 )|pi
= |hp |U
dt
dt
dt
i
h
I (t, t0 )|pi
= 2 Im hp |VI (t)UI (t, t0 )|pihp |U

I (0, t0 i|pihp|UI (t, 0)U


I (0, t0 )|pi , (9.223)
= 2 Im hp |VI (t)UI (t, 0)U
h

which yields in the limit t0 the expression


i
h
dP (t)
I (t, 0)|p(+) i .
= 2 Im hp |VI (t)UI (t, 0)|p(+) ihp |U
dt

(9.224)

Inserting the explicit time dependence of VI (t) from Eq. (9.16), and of UI (t, 0) from
0 acting on |pi and H
on |p(+) i produce the
Eq. (9.25), and using the fact that H
same energy eigenvalues, this becomes
i
h
dP (t)

= 2 Im hp |eiH0 t V eiH0 t eiH0 t eiHt |p(+) ihp |eiH0 t eiHt |p(+) i


dt
h
i
= i hp |V |p(+) ihp |p(+) i hp |V |p(+) i hp |p(+) i .
(9.225)

675

9.2 Scattering in External Potential

We now use the Lippmann-Schwinger equation (9.131) to rewrite this as


(

"

dP (t)
1
= i hp |V |p(+) i hp |pi +
hp |T(E)|pi

dt
E E i
#)
"
1

(+)
hp |T (E)|pi
, (9.226)
hp |V |p i hp |pi +
E E + i
which becomes, on behalf of Eq. (9.129) and Sochockis formula (7.189),
dP (t)
= 2hp |pi Im hp|T|pi + 2(E E )|hp |T(E)|pi|2 .
dt

(9.227)

The first term gives the rate of the transmitted particles in the direct beam. The
second term gives the rate of scattered
particles. If we collect the scattered particles
R 3
P
of all final momenta in a sum p = d p V /(2)3 we find the total probability rate
dP
=
dt

d 3 p V
2(E E)|hp |T (E)|pi|2 .
(2)3

(9.228)

There exists a simple mnemonic rule for deriving this formula. The probability
for a particle to go from the initial to the final state is given by the absolute square
of the scattering amplitude (9.130):
2

Pp p = |hp |S|pi|
= |hp |pi 2i(E E)hp |T (E)|pi|2 .

(9.229)

Excluding the direct beam, this is equal to


Pp p = [2(E E)]2 hp |T (E)|pi|2.

(9.230)

Imagining the world to exist only for a finite time t, the -function in energy at
zero argument is finite:
(E)|E=0 =
Thus we can write (9.230) as

t/2

t/2

dt iEt
t
=
e
.

2
2
E=0

Pp p = t 2(E E)|hp |T(E)|pi|2 .

(9.231)

(9.232)

Integrating (9.232) over the final phase space volume we find the probability rate
for going to all final states:
P
dP
=
=
dt
t

d 3 p V
2(E E)|hp |T (E)|pi|2,
(2)3

(9.233)

leading directly to Eq. (9.228). This formula is written in natural units with c =
h
= 1. In proper physical units, the right-hand side carries a factor 1/h, and is
known as Fermis golden rule.

676

9 Scattering and Decay of Particles

The alert reader will have noted that in going from (9.230) to (9.232) have done
an operation which is forbidden in mathematics: we have calculated the square of a
distribution with the formula
(0)(E).
(9.234)
In the theory of distributions, such operations are illegal. Distributions form a linear
space, implying that they can only be combined linearly with each other.
In the present context, there are two ways of going from from (9.230) to (9.232)
properly. One is to use wave packets for the incoming and outgoing particles. This
was done in the textbook [1]. Another way proceeds by a more careful evaluation
of the -functions on a large but finite interval t, where instead of (E E) one
uses
Z t/2
sin |t(E E)/2|

.
(9.235)
(E E)
dt ei(E E)t =
|E E|/2
t/2
Squaring this yields
sin2 [t(E E)/2]
.
(E E)2 /4

(9.236)

sin2 [t(E E)/2]


1
[(E)]2
.
t
t(E E)2 /4

(9.237)

To get the rate we must divide this by t so we obtain a factor

This function has a sharp peak around E E = 0. The area under this is 2.
Hence we obtain for t :
1
sin2 [t(E E)/2]
[(E)]2

2(E E),
(9.238)
t
t(E E)2 /4 t
and thus once more the previous result (9.232).
There exists also a more satisfactory solution to lift the restriction of operations
between distributions to linear combination. We may define consistently products
of distributions in a unique way based on the requirement that path integrals, that
provide us with an alternative equivalent way of formulating quantum mechanics,
should be independent of the coordinates used in their definition [2].
The momentum integral can be split into an integral over the final energy and
the final solid angle d. For a nonrelativistic particle in D dimensions, this goes as
follows
Z
Z
Z
Z
Z
1
1
dD p
D1
dp p
=
dE E
=
d
d
(2)D
(2)D
(2)D
0
0
Z
Z
M
=
dE pD2 .
(9.239)
d
D
(2)
0

The energy integral removes the -function in (9.227) and by comparison with
(9.219) we can identify the differential cross section as
MV pD2
d
2 1
p|
=
2|T
,
p
d
(2)D
j

(9.240)

677

9.2 Scattering in External Potential

where we have set

hp |T (E)|pi Tp p ,

(9.241)

M 2 V 2 pD3
d
=
|Tp p |2 .
d
(2)D1

(9.242)

for brevity. Inserting j = p/MV from (9.222) this becomes

If thescattered particle moves relativistically, we have to replace M in (9.239) by


E = p2 + M 2 inside the momentum integral, so that
Z

Z
Z
dD p
1 Z
1 Z
D1
d
dp
p
=
d
dE rel
=
E
(2)D
(2)D
(2)D
0
0
Z
Z
1
d
dEE pD2 .
=
(2)3
M

(9.243)

Also in the relativistic case, the initial current is not proportional to p/M but to
the relativistic velocity v rel = p/E so that
1 p
j=
.
(9.244)
V E
Hence the cross section becomes in the relativistic case
E 2 V 2 pD3
d
=
|Tp p |2 .
(9.245)
d
(2)D1
Let us compare this with the result in Eq. (9.145) where we introduced a scattering amplitude fp p whose square is equal to the differential cross. The differential
cross section is therefore simply equal to the square of the scattering amplitude:
d
= |fp p |2 .
(9.246)
d
Thus we can identify
MV
fp p
Tp p
(9.247)
(2)(D1)/2
where we have chosen the sign to agree with the convention in Landau and Lifshitz
textbook.
Using the partial-wave expansion (9.149) and (9.157), we obtain for fp p in the
three-dimensional nonrelativistic case the expansion
fp p =

fl (p)

l=0

l
X

Ylm (
p )Ylm
(
p) =

m=l

X
l=0

fl (p)

2l + 1
),
Pl (
p p
4

(9.248)

where from (9.170)


4 2
2
4 il (p)
E Tl (p) =
[Sl (p) 1] =
e
sin l (p).
(9.249)
p
ip
p
This guarantees that the radial particle current leaving the scattering zone differs, on
the unit sphere, by a factor |fp p |2 from the incoming current, so that the differential
cross section is has the obvious normalization (9.246).6
fl (p) =

In D dimensions, is given in (9B.41).

678

9 Scattering and Decay of Particles

9.2.6

Partial Wave Decomposition of Cross Section

The contribution of the different partial waves to the total cross section is obtained
by forming the absolute square of (9.248) and integrating over all directions of
the outgoing beam, and using orthogonality relation (9.151) and addition theorem
(23.12):
=
=

d
p |fp p |2 =

X
l=0

|fl (p)|2

fl (p)fl (p)

l=0

XXZ

Ylm (
p)Ylm
(
p) =

lm

p)
p )Yl m (
d
p Ylm (
p )Ylm (
p)Ylm (

lm l m

X
l=0

|fl (p)|2

2l + 1
Pl (1).
4

(9.250)

Since Pl (1) = 1, we obtain simply


=

X
l=0

l =

X
l=0

|fl (p)|

+1
4 X
= 2
(2l + 1) sin2 l (p).
4
p l=0

2 2l

(9.251)

As mentioned at the end of Subsection 9.2.3, there can be a resonance at small


p if the effective range is large enough and a is positive. In the s-wave we we can
approximate f0 (p) in (9.249) by the expression (9.192): by
f0 (p)

4
.
k cot 0 (p) ik

(9.252)

Inserting the approximation (9.187), this becomes


f0 (p)

4
1 1
+ reff k 2 ik
a 2

(9.253)

corresponding to the general resonance formula (9.193) for the S-matrix. It leads
to a contribution to the cross section in Eq. (9.251):
0

4a2
.
(1 areff k 2 /2)2 + a2 k 2

(9.254)

The behavior of a near a low-momentum resonance is illustrated in Fig. 9.2.

9.2.7

Dirac -Function Potential

As an example, consider a -function interaction


V (x) = g (3) (x).

(9.255)

Then we find for plane waves the matrix elements


hp |V |pi =

g
V

d3 xeip x (3) (x)eipx =

g
,
V

(9.256)

679

9.2 Scattering in External Potential

Figure 9.1 Behavior of wave function for different positions of a bound state near the
continuum in a typical potential between atoms.

so that in Born approximation


fp p =

M
g.
2h2

(9.257)

Since g has the dimension energy times length3 , or h


2 /M times length, fp p has
the dimension length, as it should to make the cross section an area seen by the
incoming beam.
For the -function potential, it is instructive to try and calculate also the next
correction to the Born approximation. According to Eq. (9.132), it is given by
d3 p V
1
Vp p
Vp p + . . .
3
(2)
E E(p ) + i
"
#
Z
d3 p
1
1
2
g+g
+ ... .
=
V
(2)3 E E(p ) + i

Tp p = Vp p +

(9.258)

Thus, for zero incoming momentum and energy E, the coupling constant g is effectively replaced by
gR = g g

d3 p
1
+ ... ,
3
(2) E(p) i

(9.259)

and the T -matrix has the isotropic value



Tp p

|p|=0,E=0

1
gR .
V

(9.260)

The integral on the right-hand side diverges. A finite result can be obtained by
assuming the initial g as being infinitesimal small such as to compensate for the
divergence. We are confronted here for the first time with a renormalization problem
typical for all quantum theories with local interactions. The systematic treatment
of this problem will be presented in Chapters 7 and 8.
Here this problem occurs for the first time within the Schrodinger context. This
is due to the fact that the Schrodinger equation for the -function potential does

680

9 Scattering and Decay of Particles

not have a proper solution. It is, however, possible to bypass the infinity by renormalization and to solve the Lippmann-Schwinger equation (9.55) after all, yielding
the full scattering amplitude (9.258). The zero energy expression (9.259) can be
continued to a geometric series
gR = g g 2

1
d3 p
+ g3
3
(2) E(p) i

"Z

d3 p
1
3
(2) E(p) i

#2

+ ... .

(9.261)

This can be summed up to


1
1
= +
gR
g

d3 p
1
.
3
(2) E(p) i

(9.262)

Comparing (9.260) with (9.190) we see that the renormalized coupling determines
the s-wave scattering length as
as =

M
gR .
2h2

(9.263)

This identification of the renormalized coupling constant gR with the scattering


length may be attributed to the use of a so-called pseudopotential for which the
Schrodinger equation can be properly solved. Instead of the potential g (3) (x) which
in a Schrodinger equation picks out the value of a wave function at the origin when
forming the product, one uses the condition V (x)(x) = g (3) (x)r r(x). This
ensures that the s-wave scattering length is as = Mg/2h2 as for a hard sphere of
radius as .7
Note that (9.263) holds only for a particle of mass M in an external potential g (3) (x). A two-body system involves the relative Schrodinger equation which
contains the reduced mass Mred = 1/(1/M1 + 1/M2 ) rather than M. Hence in a
many-body system, gR is related to the scattering length by
gR = 4h2 /M.

(9.264)

A remark is useful in connection with the divergence of Eq. (9.265). It can be


made finite by cutting off the momentum integral at some large value for find
1
1
1
4 Z
4
dp = + 2M
= + 2M
,
3
gR
g
(2) 0
g
(2)3

(9.265)

and since is very large one might be tempted to conclude that gR must be positive
for any g. However, as we shall see in Chapters 7 and 8, this conclusion is not allowed
in renormalizable quantum field theories. According to the rules of renormalization,
any divergence may be cancelled by a counterterm, and the sign of it is arbitrary.
Only the renormalized quantity gR is physical, and it must be fixed by experiment.
7

See K. Huang, Statistical Mechancis, Wiley, N.Y., 1963 (p.233). For higher partial waves see
Z. Idziaszek and T. Calarco, (quant-ph/0507186).

681

9.2 Scattering in External Potential

9.2.8

Spherical Square-Well Potential

For a square-well potential


V (x) =

V0 ,
0,

r < r0 ,
r 0,

(9.266)

the radial Schrodinger equation reads


(

d2
2 d
l(l + 1)
+
+ K2
2
dr
r dr
r2
"

where
K

#)

l (r) = 0,

2M(E + V0 )/h.

(9.267)

(9.268)

There is only one solution which is regular at the origin:


l (r) = Ajl (Kr),

(9.269)

with an aarbitrary normalization factor A.


For positive energy E, the solution in the outer region r a receives an admixture of the associated spherical Bessel function nl (kr):
l (r) = B [cos l (k)jl (kr) + sin l (k)nl (kr)] ,
where
k

2ME/h,

(9.270)
(9.271)

and B, l (k) must be determined from the boundary condition at r = a. The


associated Bessel functions nl (z) have the asymptotic behavior orthogonal to (9.175):
nl (z)

cos(z l/2)
,
pr

(9.272)

The continuity of the wave function fixes the ratio B/A. The continuity of the
logarithmic derivative is independent of this ratio and fixes the phase shifts by the
equation
cos l (k)jl (kr0 ) sin l (k)nl (kr0 )
jl (Kr0 )
=k
.
(9.273)
K
jl (Kr0 )
cos l (k)jl (kr0 ) sin l (k)nl (kr0 )
For an s-wave, this reduces to
K cot Kr0 = k cot(kr0 + 0 ).

(9.274)

For small incoming energy, this equation becomes

where K0

K0 cot K0 r0 k cot(kr0 + 0 ),

(9.275)

2MV0 /h, which is solved by


0 (k) = r0

tan K0 r0
k + ... .
1
K0 r 0

(9.276)

682

9 Scattering and Decay of Particles

This determines the phase shift to be


as = r0

tan K0 r0
1
.
K0 r 0

(9.277)

The next term in the small-k expansion of (9.276) is


k3 3
3as
r0 3a2s r0 2a3s 2 ,
6
K0
!

(9.278)

implying an effective radius defined in (9.188):


reff =

1
r3
r0
02 +
.
2
6as 2as K02

(9.279)

At negative energies E, there are bound states which decrease for large r. The
only solution of (9.267) with this property is the spherical Bessel function of the
third kind (Hankel function of the first kind):8
(2)

l (k) = Bhl (kr) B [jl (kr) inl (kr)] ,

(9.280)

which behaves like e(rl/2) /kr for E < 0 where k = i 2ME i. The energies

Figure 9.2 Behavior of binding energy (in units of |V0 |) and scattering length in attractive
square well potential of depth |V0 | and radius R0 for identical bosons of reduced mass Mred .
The abcissa shows the potential depth in units of 2 h2 /8Mred R02 .
8

M. Abramowitz and I. Stegun, Handbook of Mathematical Functions, Dover, New York, 1965,
Subsection 10.1.1.

683

9.3 Two-Particle Scattering

of the bound states are found from the continuity of the logarithmic derivative at
r = r0 :
(2)

hl (ir0 )
(2)

hl (ir0 )

=K

jl (Kr0 )
jl (Kr0 )

(9.281)

For the s-wave, this becomes


K cot Kr0 = .

(9.282)

For a repulsive potential, Eq. (9.276) can be continued analytically to


0 (k) = r0

tanh |K0 |r0


k + ... .
1
|K0 |r0

(9.283)

In the hard-core limit V0 , this yields an s-wave scattering length


as = r0 ,

(9.284)

and a vanishing effective range.


For the hard-core potential, equation (9.203) has been solved [3]. The lth partial
wave amplitude is
h2 a

l 2
Tl (k , k; E) = (1)
(2M)3

kE2 k 2
ka jl (k a)jl+1 (ka) + (k k )
k 2 k 2
(2)

kE a
where kE

9.3

hl+1 (ka)

jl (k
(2)
hl (ka)

2ME/h.

)jl (k) ,

(9.285)

Two-Particle Scattering

Usually, the final state has two or more particles leaving the interaction zone. The
interaction energy is an integral over some local interaction density
V =

int (x),
d3 x H

(9.286)

integrated over all space at a fixed time x0 = t = 0. In this case, the matrix elements
of the potential factorizes as
hp |V |pi =

int (0)eipx |pi = V p p hp |H


int (0)|pi,
d3 x hp |eipx H

(9.287)

where p and p are the sum over all initial and final momenta, respectively. Thus
the matrix elements of always carry a V p p factor ensuring total momentum conservation. The same thing is true for the T -matrix Tp p . This factor is customarily
removed by defining the so-called t-matrix
Tp p = V p p tp p .

(9.288)

684

9 Scattering and Decay of Particles

The Born approximation is now


int (0)|pi.
tp p hp |H

(9.289)

Replacing the Kronecker for momentum conservation by its continuum limit,


p p

(2)3 (3)
(p p),
V

we can write (9.288) as


Tp p = (2)3 (3) (p p) tp p ,

(9.290)

i.e., the matrix elements (9.130) of the S-matrix become

hp |S|pi
= hp |pi i(2)4 (4) (p p) tp p .

(9.291)

In a world of a finite volume V existing for a finite time T , the probability of


going from an initial state p to a final state p is given by the absolute square of
(9.291). In analogy to (9.231) we have

(4)



(p)

p=0

d4 p ipx
TV
=
e
,
4

(2)
(2)4
p=0

(9.292)

so that we obtain the transition rate [compare (9.232)] Thus we can write (9.230)
as
P = T V (2)4 (4) (p p)|tp p |2 .
(9.293)

This has to be summed over all final particle states, and we obtain the rate.
dP
=V
dt

final states

(2)4 (4) (p p)|t f i |2 .

(9.294)

We have changed the notation for the t-matrix tp p and written it as t f i to emphasize
the validity for a more general initial state |ii and a larger variety of possible final
states |f i.
If the particles 1 , 2 , . . . , n leave the interaction zone, the sum over all final states
in a volume V consists of an integral over the phase space
X

=V

p1 ,p2 ,...

Z
d3 p1 Z d3 p2
d3 pn

,
(2)3 (2)3
(2)3

(9.295)

and (9.294) becomes


dP
= V n+1
dt

d3 p1
(2)3

d3 p2

(2)3

d3 pn
(2)4 (4) (p p)|t f i |2 ,
(2)3

(9.296)

where p = p1 + p2 and p = p1 + p2 + . . . + pn . By dividing out the current density


of the incoming particles j, we find from this the cross section.
=

dP 1
.
dt j

(9.297)

By omitting some of the final-state integrations, we may extract from this any
desired differential cross section.

685

9.3 Two-Particle Scattering

9.3.1

Center-of-Mass Scattering Cross Section

Let us assume the collision to take place in the center-of-mass frame with zero total
momentum and p = p1 = p2 and a total energy ECM . Then the -function fixes
p = p1 = p2 and the equality of the total energies before and after the collision:
E1 + E2 = E1 + E2 = ECM .

(9.298)

In order to be more explicit we have to distinguish the case of nonrelativistic and


relativistic particles. It is useful to treat the relativistic case first and obtain the
nonrelativistic result by going to the limit of small velocities. In the relativistic
case with initial momenta pCM = p1 = p2 , final momenta pCM = p1 = p2 and
energies
E1,2 =

2
p2CM + M1,2
, E1,2
=

2
pCM 2 + M1,2
,

(9.299)

the -functions for the total energy reads, in the center-of-mass frame, with zero
total momentum and

q

pCM 2

M1 2

pCM 2

M2 2

ECM .

(9.300)

This vanishes unless p makes the argument zero which happens at


pCM =

1
I (M, M1 , M2 ) ,
2M

(9.301)

where
I(M, M1 , M2 )
and

M2

(M1 + M2 )

ECM = M

M 2 (M1 M2 )2

(9.302)

(9.303)

is the total mass


in the center-of-mass frame, with zero total momentum and often

denoted by s. The individual energies of the two particles are


i
1 h 2  2
M M1 M22 ,
2M
q
i
1 h 2  2

2
pCM 2 M1,2
E1,2
=
M M1 M22 .
(9.304)
=
2M
To evaluate the momentum integral over the -function (9.300), we use the formula

E1,2 =

2
p2CM M1,2
=

(f (p ) f (pCM )) = [f (p )]

(p pCM )

(9.305)

to write

q

p2 + M1 2 +

p2 + M2 2 ECM
1

= [(/p ) (E1 + E2 )] (p pCM )


E1 E2
E1 E2

(p pCM ).
(p

p
)
=
=
CM
pCM (E1 + E2 )
pCM ECM

(9.306)

686

9 Scattering and Decay of Particles

Then we obtain from (9.296) the total rate of scattered relativistic particles
dP (t)
V 3 pCM E1 E2
=
dt
(2)2 ECM

d|t f i |2 .

(9.307)

Since, the t-matrix elements depend, in general, on more than just the initial and
final total momenta p and p , we have denoted the matrix elements of t between
initial and final states by t f i .
The initial current density is given by the relative velocity between two incoming
particles in the center-of-mass frame with zero total momentum and [see (9.222) and
(9.244)]. Hence
j=

p1
p2

E1 E2

1
pCM ECM 1
=
.
V
E1 E2 V

(9.308)

Dividing dP (t)/dt by j as demanded by (9.335) and omitting the integration


over the solid angle in (9.307), we obtain the differential cross section in the center
of mass:
1 pCM 1
d
=
E1 E2 E1 E2 V 4 |t f i |2 .
(9.309)
2
dCM
(2)2 pCM ECM
It is useful to realize that the combination of incoming currents and energies j
2V E1 E2 is a Lorentz-invariant quantity coinciding with I(M, M1 , M2 ) of Eq. (9.302).
This follows immediately by inserting (9.301) into (9.308):
j=

1
I(M, M1 , M2 ).
2V E1 E2

(9.310)

Thus we can write down an explicit formula for the cross section (9.335) any frame
of reference:
2V E1 E2
=
Vn
I(M, M1 , M2 )

9.3.2

d3 p1
(2)3

d3 p2

(2)3

d3 pn (4)
(p p)|t f i |2 .
3
(2)

(9.311)

Laboratory Scattering Cross Section

In most experiments, the directly measure quantity is the laboratory cross section
where an initial particle of momenta |p1 | = pL and energy
EL (E1 )Lab =

p2L + M12

hits the particle 2 at rest. Then the incoming current is


pL
vL
j=
= ,
EL V
V

(9.312)

(9.313)

rather than (9.308). The cross section of two particles going into two particles in
the laboratory is from (9.296) and (9.335):
=

d3 p1 V d3 p2 V
4 (4)
2V
(2)

(p

p)|t
|
.
f
i
(2)3 (2)3
j

(9.314)

687

9.3 Two-Particle Scattering

The -function for the total energy depends on the final laboratory momenta p1L
|p1L |, p2L |p2L | as follows:

q

p1L 2 + M1 2 +

p2L 2 + M2 2 EL M2

(9.315)

In contrast to the center-of-mass frame, the final momenta p1L and p2L are no
longer equal, which complicates the further evaluation. We want to express them as
functions of the initial energy and the scattering angle L . For this some kinematic
considerations are necessary.
The total mass introduced in (9.303) is given by the Poincare-invariant quantity
M 2 = s (p1 + p2 )2 = M12 + M22 + 2M2 EL .

(9.316)

Solving this we obtain the invariant expressions


EL =

M 2 M12 M12 M22 /2M2

pL = I (M, M1 , M2 ) /2M2 .

(9.317)
(9.318)

Observe that by inserting the last equation into the frame-independent expression
(9.310), we find the correct current (9.313) in the laboratory frame.
The relation between the laboratory and center-of-mass energies is
M2 EL + M12
,
E1CM = q
M12 + M22 + 2EL M2

M2 EL + M22
E2CM = q
.
M12 + M22 + 2EL M2

(9.319)
(9.320)

The relation between the laboratory the center-of-mass momenta is


M1
.
pCM = pL q
M12 + M22 + 2EL M2

(9.321)

The relation between the scattering angles in the two frames of reference is more
involved. To find it, we consider the Lorentz invariant formed from the momentum
transfer: the so-called Mandelstam variable
t (p1 p1 )2 = M12 + M1 2 + 2(p1 p1 E1 E1 ).

(9.322)

In either frame, the scattering angles appear in the scalar product p1 p1 = p1 p1 cos .
A calculation in the center-of-mass frame, leads to the formula
cos CM =

1
2pCM pCM

(t M12 M1 2 + 2E1CM E1CM


)

1
(9.323)
I(M, M1 , M2 )I(M, M1 , M2 )
[M 4 + M 2 (2tM12 M22 M1 2 M2 2 ) + (M12 M22 )(M1 2 M2 2 )].
=

688

9 Scattering and Decay of Particles

In the laboratory frame, we take advantage of the fact that p2 = (M2 , 0) and
calculate t using energy-momentum conservation as

t = (p1 p1 )2 = (p2 p2 )2 = M2 2 + M22 2M2 E2L


.

(9.324)

It is convenient to introduce another invariant momentum transfer

u (p1 p2 )2 = (p1 p2 )2 = M22 M1 2 2M2 E1L


.

(9.325)

The three invariants s, t, u are called Mandelstam variables, who used them extensively in his work on the scattering matrix. They are related by
s + t + u = M12 + M22 + M1 2 + M2 .

(9.326)

This implies that all different kinematical configurations of a two-body scattering


process can be represented in an invariant way by a point in an equilateral triangle
of height M12 + M22 + M1 2 + M2 2 , the so-called Mandelstam triangle. The heights
over the three sides represent the quantities s, t, u, and their sum is constant.
Similar considerations can be made for a decay process of a particle into three
particles. Then a Poincare-invariant way of picturing the distribution of the final
products in a Mandelstam triangle is known as Dalitz plot. Dalitz plot.
The laboratory energy of all particles can be calculated from (9.317) and
1
(M 2 + M2 2 t),
M2 2
1
=
(M22 + M1 2 u).
M2

E2L =

(9.327)

E1L

(9.328)

By evaluating the invariant (9.322) in the laboratory frame we find the scattering
angle in this frame:
cos L =

(M 4 M12 M22 )(M22 + M1 2 u) + 2M22 (t M12 M1 2 )


.
I(M, M1 , M2 )I(M, M1 , M2 )

(9.329)

The argument of the total energy -function (9.315) requires the knowledge of
the momenta p1 , p2 of the final particles as a function of the initial energy EL and
the scattering angle L . These can now be calculated as follows. From (9.316) we
find the invariant M as a function of EL . This is inserted into (9.329) to obtain the
invariant t as a function. Together with (9.326), this fixes the invariant u. Inserted

into Eqs. (9.328), we find the final energy E1L


, and from energy conservation also

E2L . These determine the final momenta.


It is useful to generalize the definition of the scattering amplitude f of (9.247)
(for the scattering of a nonrelativistic particle in an external potential) to the present
case of two relativistic particles scattering each other in the center-of-mass frame.
with zero total momentum and We define again f so that
d
= |f |2
dCM

(9.330)

689

9.3 Two-Particle Scattering

and see that f has to be identified with


1
f =
2

pCM 1 q 4
V E1 E2 E1 E2 t f i .
pCM ECM

(9.331)

In the case of elastic scattering on a very heavy particle 2, this reduces back to the
original definition.
For a more general scattering process in which two particles collide and produce
n particles in the final state the production cross section is given by
"Z
#
n
V Y
d 3 pi V
f i =
(2)4 (4) (p p)|t f i |2
j i=1
(2)

(9.332)

with the prefactor V /j = V 2 E1 E2 /pCM ECM as given by (9.308),


Scattering processes in a center-of-mass frame with zero total momentum and
can be observed directly in colliding beam experiments. There one has two beams
impinging almost with opposite momenta (see Fig. 24.8). If Vint is the volume of

Figure 9.3 Geometry of particle beams in a collider. The angle of intersection is


greatly exaggerated. The intersection defines the interaction volume Vint .

the intersecting zone, the interaction volume and n1 , n2 are the particle densities in
the two beams (particles per unit volume), the relative velocity vrel integrated over
the interaction volume is the so-called luminosity of the colliding beam:
L = vrel Vint n1 n2 =

I(M, M1 , M2 )
Vint n1 n2 .
2E1 E2

(9.333)

The rate of particle scattering is obtained by multiplying the differential cross section
d/dCM by this quantity:
dN
d
=
L.
dtdCM
dCM

(9.334)

Note that the luminosity increases linearly if the beams are focused upon a smaller
area (since n1 n2 increases like 1/area2 while Vint increases like area).

690

9.4

9 Scattering and Decay of Particles

Decay

The decay rate of an unstable particle into two particles is directly given by
Eq. (9.296):
dP
=V2
dt

d3 p1
(2)3

d3 p2
(2)4 (4) (p p)|t f i |2 .
(2)3

(9.335)

Evaluating this in the center-of-mass frame of the decaying particle yields


d
V 3 pCM E1 E2
= P (t) =
dt
(2)2
M

d |t f i |2

(9.336)

where M is the mass of the initial particle. If the particle has spin s and decays into
final spins s1 , s2 , an unpolarized initial state and all possible final spin orientations
m1 = s1 , . . . , s1 and m2 = s2 , . . . , s2 leads to an isotropic distribution, and we
can simplify (9.336) to
=

X
d
V 3 pCM E1 E2 4
P (t) =
|t f i |2 .
dt
(2)2
M
2s + 1 m ,m
1

9.5

(9.337)

Optical Theorem

The unitarity of the S-matrix (9.45),


SS = S S = 1,

(9.338)

implies an important theorem for the imaginary part of the diagonal elements of the
T -matrix. Writing
S f i = f i (2)4 (4) (pf pi ) i t f i

(9.339)

we see that


i tii tii =

X
f

t if (2)4 (4) (pf pi ) t f i

(9.340)

where the sum runs over all possible many particle states. According to Eq. (9.331),
the left-hand side is related to the imaginary part of the forward scattering amplitude
fii as follows
n
XY
2ECM
Im
f
=

ii
V 2 E1 E2
n i=1

"Z

d 3 pi V
(2)4 (4) (pf p1 ) |t f i |2
(2)3
#

(9.341)

Now, the right-hand side, if divided by the relative current density j =


pCM ECM /E1 E2 V , is equal to 1/V times the sum over all cross sections f i , i.e.,

691

9.6 Initial- and Final-State Interactions

it is equal to j/V times the total cross section (j/V )tot [see (9.314)]. For the
forward scattering amplitude fii this implies the simple relation
2
fii = tot .
pCM

(9.342)

This relation is referred to as the optical theorem (in reference to a theorem relating the total absorption of light to the imaginary part of the forward-scattering
amplitude).

9.6

Initial- and Final-State Interactions

w
Some inelastic scattering and decay processes are caused a weak interaction H
which can be treated in Born approximation, whereas the initial and final states are
s . Suppose the total Hamiltonian contains these
distorted by a strong interaction H
interactions additively, i.e.
=H
0 + Vs + Vw ,
H
(9.343)
w may be the weak interaction of beta decay, say, and H
s
In specific applications, H
w may
a combination of electromagnetic and strong interactions. Alternatively, H

be an electromagnetic interaction, and Hs the strong interaction.


For time-independent interactions, the Mller operators are given by (9.33) and
read explicitly

I (0, ) = lim ei(H 0 +V w +Vs )t eiH 0 t ,


() = U
t

(9.344)

The right-hand sides can obviously be decomposed into a product


I (0, ) = UIw (0, ) UI,s (0, ) ,
U

(9.345)

where
UIw (0, ) =
UIs (0, ) =

lim ei(H0 +Vs +Vw )t ei(H0 +Hs )t ,

lim ei(H0 +Vs )t eiH0 t ,

(9.346)
(9.347)

Thus we can write the Mller operators as


()
() = ()
w s ,

(9.348)

with
()

Iw (0, ) .
()
=U
w = UIw (0, ) , s

(9.349)

Using this, we can express the scattering amplitude as follows:


Is (0, )|ii,
Is (, 0)Sw U
Sfi = hf|U

(9.350)

692

9 Scattering and Decay of Particles

where, in analogy to (9.43),

Using the notation

Iw (, ) = () (+) ,
Sw U
w
w

(9.351)

Is (0, )|ii = (+) |ii,


|i()i U
w

(9.352)

in a slight variation of (9.115), the symbols () referring now only to the initialand final-state interactions, the scattering matrix elements become
= hj()|Sw |i(+)i.
hj|S|ii

(9.353)

Under the assumption of a weak Vw , we replace Sw by its Born approximation


and write
Z

dtVIw (t),
(9.354)
Sw i

where

VIw (t) ei(H0 +Vs +Vw )t Vw ei(H0 +Vs )t

(9.355)

is the interaction representation of the weak interaction in the presence of the initialand final-state interactions. The initial and final states are eigenstates of the oper 0 + Vs ) [see (9.16)]. The integration over t in (9.354) can therefore be done
ator H
yielding a -function of energy conservation [recall the derivation of (9.130)], and we
obtain
= hj| ii 2i(E E)hj()|Vw |i(+)i.
hj|S|ii

(9.356)

For a local interaction which is an integral over an interaction density as in


Eq. (9.286), we find the modified expression
int (0)|p(+) i.
tp p hp() |H

(9.357)

and for the amplitude of multiparticle scattering (9.294), the initial- and final-state
interactions are included in the formula
int (0)|i(+)i.
tfi hf()|H

9.7

(9.358)

Tests of Time-Reversal Violations

For a time-reversal invariant Hamiltonian, the property (7.116) has the consequence
that the scattering operator S of Eq. (9.43) satisfies
= S .
T ST

(9.359)

To derive observable consequences from this, we first use the properties (7.111) and
(7.112) to write down the trivial equality for any matrix element between an initial

693

9.7 Tests of Time-Reversal Violations

state |ii obtained by applying one or more particle creation operators upon the
vacuum, and a similar final state |fi:
= hf|1A S|ii
.
hf|S|ii

(9.360)

Invoking now time-reversal invariance in the form (9.359), we obtain


= hf|T S T |ii = hfT |S |iT i = hiT |S|f
T i.
hf|S|ii

(9.361)

Here |iT i denotes the time-reversed of the state |ii, in which all creation operators
are replaced by the right-hand sides of (7.106), (7.312), etc. Thus momenta and
spin directions are reversed, and there is a phase factor T for each particle.

9.7.1

Strong and Electromagnetic Interactions

Strong and electromagnetic interactions are time-reversal invariant. The matrix


the scattering amplitudes, satisfies the identity
elements of the scattering operator S,
a , s3a ; pb , s3b i = hpa , s3a ; pb , s3b |S|
pc , s3c ; pd , s3d i,
hpc , s3c ; pd , s3d |S|p
(9.362)
with some phase factor of unit norm. When calculating the unpolarized cross
section we have to sum over the final and average over the initial states. In the
center-of-mass frame, the cross section carries phase space factor pCM /pCM , the
ratio of final and initial center-of-mass momenta. It has to be removed to compare
the squared amplitudes. Hence we obtain
d(ab cd)
d
=
d(cd ab)
d

pcd
CM
pab
CM

!2

(2sc + 1)(2sd + 1)
.
(2sa + 1)(2sb + 1)

(9.363)

This is called the principle of detailed balance. It is important for the thermal
equilibrium of chemical reactions. This equation was used in 1951 to determine the
spin of the pion in the reaction in which two protons produce a pion and a deuteron9
p+p
d + +.

(9.364)

The principle of detailed balance says that


d()
2
d = 3 q (2s + 1) .
d()
4
p2p
d

(9.365)

This relation was satisfied with s = 0.


9

R.E. Marshak, Phys. Rev.82, 313 (1951); W.B. Cheston, Phys. Rev. 83, 1118 (1951); R.
Durbin, H. Loar, and J. Steinberger, Phys. Rev. 83, 646 (1951); D.L. Clark, A. Roberts, and R.
Wilson, Phys. Rev. 83, 649 (1951).

694

9.7.2

9 Scattering and Decay of Particles

Selection Rules in Weak Interactions

Consider a one-particle state |ii prepared by strong and electromagnetic interactions,


be an
which decays into a several-particle state |bi via weak interactions. Let O
observable which changes sign under time reversal:
T = O.

(9.366)

a long time after the decay. It is given by


Consider the expectation value of O
U (, )|ii = hi|S O
= hi|U(,

S|ii.

hOi
)O

(9.367)

Using the antiunitarity of the time-reversal operator as expressed by Eq. (7.111)


and (7.112), the right-hand side can be rewritten as
S|ii
= hi|1A S O
S|ii
= hi|T T S O
S|ii
.
hi|SO

(9.368)

Time-reversal invariance (9.359) brings this to


T S T |ii = hiT |SO
T S |iT i .
hi|T SO

(9.369)

was assumed to be an observable, the expectation is real, and we can drop


Since O
the complex conjugation at the end. This leaves us with the relation
= hi|S O
S|ii
= hiT |SO
T S |iT i.
hOi

(9.370)

If the decay is very slow, the scattering operator can ne replaced by its Born approximation
Z
dtVIw (t),
(9.371)
S SB i

which is a hermitian operator, up to a factor i (the direct term in (9.339) does not
contribute to a decay process). But then (9.370) becomes
= hi|SB O
SB |ii = hiT |SB O
T SB |iT i.
hOi

(9.372)

If the initial particle is spinless, then


|iT i = T |ii,

|T | = 1.

(9.373)

Together with (9.366), this implies


= 0.
hOi

(9.374)

Thus in a weak decay process, the expectation value of an observable which changes
sign under time reversal vanishes. Note that this result is independent of the phase
T of the decaying particle.
Without the weakness assumption, the exchange of the hermitian conjugation
on the scattering operators in (9.370) would have prevented us from any conclusion.

695

9.7 Tests of Time-Reversal Violations

Historically, the invariance under time reversal of weak interactions was apparently assured by the weak decays
K+

0 +
+ + ,

0 + + .

(9.375)

The outcoming muon was found to have no average polarization orthogonal to the
= 
production plane formed by the momenta p and p , i.e., the operator O
) had no measurable expectation value. Since this operator is odd under
(
p p
seemed to confirm time-reversal invariance.10
time reversal, the vanishing of hOi
However, as it turned out later, the breakdown of time-reversal invariance in weak
interactions is much too small to be detected by a polarization experiment. There
exists a much more sensitive experiment which shows the violation without effort:
0 meson system. This will be discussed later.
the K 0 -K

9.7.3

Phase of Weak Amplitudes from Time-Reversal Invariance

If a particle decays weakly into two final particles, time-reversal invariance fixes the
phase of the decay amplitude, if the final-state interaction is purely elastic, i.e., if
there are no other decay channel caused by the final-state interactions. A weak
decay can be treated in lowest-order perturbation theory and has the amplitude
w |i(+)i,
tba = hf()|H

(9.376)

where the labels () indicate that the states contain the full effect of the initialand final-state interactions (see Section 9.6). From time-reversal invariance we find
w |i(+)i = hiT ()|H
w |fT (+)i = hfT (+)|H
w |iT ()i .
hf()|H

(9.377)

If we choose the phase factors T of initial and final particles to be unity [see (4.223)],
we obtain
w |i(+)i = hfT (+)|H
w |iT ()i .
hf()|H
(9.378)
The initial state is a stationary single particles state if the weak interactions are
ignored. Hence
|i(+)i |i()i.
(9.379)
We now rotate the matrix element on the right-hand side by 180 degrees in the
center of mass frame, we reverses momentum and spin directions and bring the
time-reversed state back to the original form. Thus we obtain
w |a(+)i = hb(+)|H
w |a(+)i .
hb()|H

(9.380)

The amplitudes with + and minus interchanged differ by a pure phase factor which
is the scattering amplitude for the two-particle scattering in final state:
w |a(+)i = e2i hb(+)|H
w |aT ()i,
hb()|H
10

(9.381)

This possibility of testing time-reversal invariance of weak interactions was pointed out by J.J.
Sakurai, Phys. Rev. 109, 980 (1958). See also M. Gell-Mann and A.H. Rosenfeld, Ann. Rev.
Nucl. Sci. 7, 407 (1957), Appendix A.

696

9 Scattering and Decay of Particles

The pure-phase nature is a consequence of the assumption of elasticity. Together


with (9.380), this fixes the phases
w |a(+)i = const . ei .
hb()|H

(9.382)

This equation has many applications in phenomenological analyses.11

Appendix 9A

Green Function in Arbitrary Dimension

In D dimensions, the Green function (9A.1) in the scattered wave (9.136) of the LippmannSchwinger equation has the following form
0 (E)|x i = 2M ih
hx|G

dD p eip(xx )/h
.
(2h)D p2 p2E

(9A.1)

The momentum integral has been calculated in Section 1.12 (also 9.1) of Ref. [2], for E > 0 yielding
(1)

0 (E)|x i =
hx|G
where kE pE /
h=
behavior is

D2 H
M kE
D/21 (kE |x x |)
,
h (2)D/2 (kE |x x |)D/21

(9A.2)

p
2M (E + i)/h, and H (1) (z) is the Hankel function, whose leading large-z
H(1)

2 i(z/2/4)
e
,
z

(9A.3)

thus reproducing the previous result (9A.1) for D = 3.


In D = 1 dimension, the expression (9A.2) reduces to
0 (E)|x i =
hx|G

M ikE |xx |
e
,
hkE

so that the Lippmann-Schwinger equation (9.134) in one dimension reads




Z

M
1
(+)
(+)
hx|k i = 1/2 eikx + i 2 eikE |xx |
dx V (x )hx |k i ,
L
h kE

p
where kE = 2M (E + i) is equal to the wave number k of the incoming particles.
As a simple application consider the case of a -function potential
2
h
(x)
Ml

V (x) =

(9A.4)

(9A.5)

(9A.6)

where (9A.5) becomes


(+)

hx|k i =

1
L1/2



Z
1 ik|x|
(+)
e
h0|k i .
eikx +
ikl

(9A.7)

Setting x = 0 on the left-hand side determines


(+)

h0|k i =
11

G. Takeda, Phys. Rev. 101, 1547 (1956).

1
ikl
,
L1/2 ikl 1

(9A.8)

Appendix 9A

Green Function in Arbitrary Dimensions

and Eq. (9A.7) yields the fully interacting wave function




1
1
(+)
eik|x| .
hx|k i = 1/2 eikx +
ikl 1
L

697

(9A.9)

()

The corresponding states |k i are


1

()

hx|k i =

L1/2


eikx +


1
eik|x| .
ikl 1

(9A.10)

In the negative-x regime, the second wave runs to the left and its prefactor is the reflection
amplitude
r=

1
.
ikl 1

(9A.11)

The positive x the two exponentials run both to the right and the common prefactor
t=

ikl
ikl 1

(9A.12)

is the transmission amplitude. The unitarity relation for this process is |r|2 + |t|2 = 1.
For an attractive -function, there is a bound-state pole at kB = 1/|l| hwere the energy is
B =

2
2 kB
h
h2
g2M
= 2
=
.
2M
2l M
2

(9A.13)

()

It is easy to verify that hx|k i satisfies the Schroedinger equation



 2 2
h2
h x

()
+
(x) E hx|k i = 0

2M
Ml

(9A.14)

with the jump property at the position of the -function


2 x
h
2
h
()
()
()
[h0+ |k i h0 |k i] =
h0|k i.
2M
Ml

(9A.15)

Let us also calculate the matrix elements of the resolvent operator (9.58), the Greeen function

hx|G(E)|x
i = hx|

ih
|x i.
+ i
EH

(9A.16)

It has the spectral representation

hx|G(E)|x
i=

ih
dk
()
()
hx|k ihk |x i.
2 E Ek + i

Performing the integral over k we find from the residue at kE the result



1
M

eikE |xx | +
hx|G(E)|x
i=
eikE (|x|+|x |) .
hkE

ikE l 1

(9A.17)

(9A.18)

This is related to the free Green function (9A.4) via Eq. (9.59).
The Green function (9A.18) can also be obtained with the help of the Wronski method12 by
combining the two independent solutions

(+)
f2 (x) = f1 (x)
(9A.19)
f1 (x) = Lhx|kE i,
12

See Subsection 3.2.1 in the textbook [2].

698

9 Scattering and Decay of Particles

E)f (x) = 0 in the form


of the homogeneous Schrodinger equation (H

hx|G(E)|xi
= hx|

i
h

EH

|x i =

2M i
1
[(xx )f1 (x)f2 (x ) + (x x)f1 (x )f2 (x)], (9A.20)
h W (x, x)

where W (x, x) is the Wronski determinant


W (x, x) f1 (x)f2 (x) f1 (x)f2 (x ).

(9A.21)

= E + hx2 /2M to (9A.20) we obtain ih(x x ) as


Indeed, applying the operator E H
For the free system where f1 (x) = eikE x , f2 (x) = eikE x and W (x, x) = 2ikE , this formula
reproduces the above rasult (9A.4):
0 (E)|xi =
hx|G

1
M ikE |xx |
2M i
e
.
f1 (x> )f2 (x< ) =
h W (x, x)

hkE

(9A.22)

In the presence of the -potential, where the Wronski determinant is calculated from the solutions
(9A.19) as
W (x, x) =

2
2ikE
l
kE l + i

(9A.23)

the formula (9A.20) yields

hx|G(E)|xi
= hx|

i
h

EH

|x i =




M
1
eikE |xx |
eikE (|x|+|x |) .
hkE
1 ikE l

(9A.24)

From this we find the bilocal density of states





M
1
1
ikE (|x|+|x|)
ikE (xx )

+c.c.
hx|[G(E +i) G(E i)]|xi =
e
hx|
(E)|x i =
e
+
2
h
ikE l 1
2h2 kE
(9A.25)
R
R

Integrating this over all energies in the continuum, i.e., by forming 0 dE = h2 kE dkE /M ,
we obtain
Z

1
dEhx|
(E)|x i = (x x ) (l) e(|x|+|x |)l .
(9A.26)
l
0

According to Eq. (9A.17) the same quantity is equal to


Z
Z
Z
dk
dk
(+)
(+)
(+)
(+)
2
dE
hx|k i(E hk /2M )hk |x i =
hx|k ihk |x i,
0
2
2

(9A.27)

Hence
Eq. (9A.26) together with the contribution of the single bound state wave function hx| i =
(1/ l)e|x|l yields
Z
dk
(+)
(+)
hx|k ihk |x i + (l)hx| ih |x i = (x x ),
(9A.28)
2

(+)
thus proving the completeness relation (9.63) of the states hx|k i and hx| i = (1/ l)e|x|l .

Appendix 9B

Partial Waves in Arbitrary Dimensions

In D dimensions, angular decomposition of the scattering amplitudes proceeds as follows. The


T -matrix hp |T |pi in Eq. (9.129) has for equal incoming and outgoing energies Ep = Ep , i.e.,
|p | = |p| = p an expansion
hp |T |pi =

X
(2)D X
Tl (p)
Ylm (
p )Ylm (
p),
V
m
l=0

(9B.1)

Appendix 9B

Partial Waves in Arbitrary Dimensions

699

where Ylm (
p) are the hyperspherical harmonics generalizing (4D.1) associated with the momentum
, whose precise form is tedious to write down but will not be needed here. The magnetic
direction p
quantum number m is replaced by a D 2 dimensional vector of eigenvalues running though
dl =

(2l + D 2)(l + D 3)!


l!(D 2)!

different values for each l. The orthogonality (4D.3) becomes now:


Z
Yl m (
d2 p
p)Ylm (
p) = l l m m ,
where

(9B.2)

(9B.3)

denotes the integral over the surface of a unit sphere in D dimensions:


dD1 p
Z

= SD
dD1 p

2 D/2
.
(D/2)

(9B.4)

The spherical harmonics are matrix elements of eigenstates |l mi of angular momentum [recall
(4.818)] with localized states |
pi on the unit sphere:
h
p|l mi Ylm (
p).

(9B.5)

),
h
p |
pi = (D1) (
p p

(9B.6)

The latter are orthogonal:


(D1)

) is the -function on a unit sphere.


where
(
p p
The addition theorem (23.12) for the spherical harmonics are replaced by
X
2l + D 2 1 (D/21)
) =
Cl
(
p p
Ylm (
p )Ylm (
p),
D 2 SD
m

(9B.7)

()

where the functions Cl (cos ) are the hyperspherical Gegenbauer polynomials, the D-dimensional
versions of the Legendre polynomials Pl (cos ) in Eq. (4D.7).
The Gegenbauer polynomials are related to Jacobi polynomials, which are defined in terms of
hypergeometric functions by13
(,)

Pl

(z)

1 (l + 1 + )
F2,1 (l, l + 1 + + ; 1 + ; (1 + z)/2).
l! (1 + )

(9B.8)

The relation is14


()

Cl (z) =

(2 + l)( + 1/2) (1/2,1/2)


P
(z).
(2)( + l + 1/2) l

(9B.9)

Thus we can write


()

Cl (z) =

1 (l + 2)
F2,1 (l, l + 2; 1/2 + ; (1 + z)/2).
l! (2)

For D = 2 and 3, one has15


lim

0
(1/2)

Cl
13

1 ()
1
C (cos ) = cos l,
l
2l
(0,0)

(cos ) = Pl

(cos ) = Pl (cos ).

M. Abramowitz and I. Stegun, op. cit., Formula 15.4.6.


Ibid., Formula 15.4.5.
15
I.S. Gradshteyn and I.M. Ryzhik, op. cit.,Formula 8.934.4.
14

(9B.10)

(9B.11)
(9B.12)

700

9 Scattering and Decay of Particles

The partial-wave expansion (9B.1) becomes in D dimensions


hp |T |pi =

Tl (p)

l=0

2l + D 2 1 (D/21)
).
C
(
p p
D 2 SD l

(9B.13)

A similar expansion of the S-matrix (9.130) requires expanding the scalar product hp |pi = p ,p
into partial waves. In a large volume V , we replace this by (2)D (D) (p p)/V . The -function
is decomposed into a directional and a radial part
p,p =

(2)D 1
(2)D (D)
)
(p p).
(p p) = (D1) (
p p
V
V pD1

(9B.14)

The directional -function on the right-hand side has a partial-wave expansion which is the completeness relation [compare (9.159] for the states |l mi,
X
X
l=0 m

|l mihl m| = 1,

(9B.15)

evaluated between the localized states |


pi on the unit sphere. Using the matrix elements (9B.5)
and the orthogonality relation (9B.6), we find from (9B.15):
) =
(D1) (
p p

X
l
X

Ylm (
p )Ylm (
p).

(9B.16)

l=0 m

The radial part of (9B.14) can be rewritten as


(2)D 1 dE
(2)D 1
(2)D 1
(p p) =
(E E),
(E E) =
D1
D1
V p
V p
dp
V E
where
E

pD1
.
dE/dp

It allows us to rewrite the sum over all momentum states as follows:


Z
Z
X Z dD pV
V
D1

=
=
d
p
dE E .
(2)D
(2)D
p

(9B.17)

(9B.18)

(9B.19)

Thus we find the expansion


hp |pi = p ,p =

1
(2)D X X
Ylm (
p )Ylm (
p) (E E).
V

E
m

(9B.20)

l=0

In a partial-wave expansion for the S-matrix (9.130), the overall -function for the energy
conservation is conventionally factored out and one writes:

hp |S|pi

X
X
(2)D 1
(E E)
Sl (p)
Ylm (
p )Ylm (
p),
V E
m

(9B.21)

l=0

The relation between the partial-wave scattering amplitudes


Sl (p) and those of the T -matrix is again given by (9.166), the unitarity property (9.45) of the
S-matrix becomes Sl (p)Sl (p) = 1, so that Sl (p) = e2il (p) , as in (9.169), and Tl (p) is equal to
(1/E )eil (p) sin l (p), as in (9.170). Only E is the D-dimensional function of p in Eq. (9B.18).

Appendix 9B

701

Partial Waves in Arbitrary Dimensions

In order to show this we use the expansion formula corresponding to (9.171)


eih cos =

al (h)

l=0

l + D/2 1 1 (D/21)
C
(cos ),
D/2 1 SD l

(9B.22)

(2)D/2 h1D/2 il Jl+D/21 (h),

(9B.23)

where
al (h)

()

The expansion (9B.22) follows from the completeness of the polynomials Cl (cos ) at fixed ,
using the integration formulas16
Z

/2

/2

/2

/2

()

21 (2 + l)
h I+l (h),
l!()

(9B.24)

()

212 (2 + l)
ll .
l!(l + )()2

(9B.25)

d sin eh cos Cl (cos ) =


()

d sin Cl (cos )Cl (cos ) =

We now use (9B.22) to expand the incoming free-particle wave

hx|pi

=
=

X
X
1
eipx =
l (pr)
Ylm (
x)Ylm (
p)
V
m
l=0

X
l=0

where r |x|, and

l (pr)

l + D/2 1 1 (D/21)
),
C
(
xp
D/2 1 SD l

1
l (pr) = al (pr).
V

(9B.26)

(9B.27)

Far from the scattering region, the spherical Bessel functions have the asymptotic behavior
r
2
J (z)
sin(z /2 + /4),
(9B.28)
z
implying for the partial waves [compare (9.176)
1
sin{pr [l/2 (D 3)/4]}
l (pr) = il 2 (2)(D1)/2
.
(pr)(D1)/2
V

(9B.29)

We shall now calculate the asymptotic form of the outgoing waves. The second term of the
Lippmann-Schwinger equation (9.131) yields the scattered wave, which reads in configuration space

sc

hx|pi

1
dD p V
hx|p i
hp |T |pi.
(2)D
E E + i

Inserting the partial-wave decompositions (9B.1) and (9B.26), with (9B.27), we find
Z D
d pV
1
hx|p i
hp |T |pi
(2)D
E E + i
Z

al (p r)Tl (p)
1 XX
,
Ylm (
x )Ylm (
p)
dp pD1
=
E E + i
V
0
m
l=0

16

I.S. Gradshteyn and I.M. Ryzhik, op. cit., Formulas 7.321 and 7.313.

(9B.30)

(9B.31)

702

9 Scattering and Decay of Particles

so that the total partial waves including the the scattered waves are


Z

tot
D1 l (p r)Tl (p)
.
l (pr) = l (pr) +
dp p
E E + i
0

(9B.32)

In terms of the spherical Bessel functions of Eq. (9.172), this reads


1
1
ltot (pr) = 2(2)(D1)/2
il
(D3)/2
(pr)
V

Z
jl+(D3)/2 (pr) + p(D3)/2
dp p(D+1)/2 jl+(D3)/2 (p r)
0

(9B.33)
Tl (p )
.
E E + i


For large r, we use the asymptotic formula (9.175) to rewrite the integral as [compare (9.180)]
Z

ei{p r[l/2(D3)/4]}ei{p r[l/2(D3)/4]}


Tl (p )
(D+1)/2
.
(9B.34)
dp p
E E + i
2ip r
0
The integral can now be approximated as in (9.182):
Z

1
1  i{pr[l/2(D3)/4]}
Tl (p )
i{pr[l/2(D3)/4]}
,
e

e
dp p(D+1)/2
dE/dp
p p + i 2ip r

(9B.35)

and yields, together with the prefactor p(D3)/2 in (9B.33),

2pD1 il/2 i(D3)/4 eipr


e
e
,
dE/dp
2pr

(9B.36)

In this way we find for the wave function (9B.33):


r
1 2(2)(D2)/2
ltot (pr)

(9B.37)
V 2i(pr)(D1)/2
i
h

ei{pr[l/2(D3)/4]} ei{pr[l/2(D3)/4]} 2iE Tl (p)ei(D3)/4 eipr .

Expressing Tl (p) in terms of Sl (p) using Eq. (9.166), we have

i
r 1 2(2)(D1)/2 h
i(pr(D3)/4)
l ip(pr(D3)/4)
.
S
(p)e
(1)
e
ltot (pr)

l
V 2i(pr)(D1)/2

(9B.38)

Inserting the unitary form (9.169) for elastic scattering, this becomes
r 1 2(2)(D1)/2
sin{pr + l (p)[l/2 (D3)/4]}
ltot (pr)

il eil (p)
.
r(D1)/2
V p(D1)/2

(9B.39)

This shows again that l (p) is the phase shift of the scattered waves with respect to the incoming
partial waves (9B.29).
If we insert (9B.33) on the right-hand side of the partial wave expansion (9B.26), the first term
gives again the incoming plane wave and the Tl (p)-terms add to this via Eq. (9B.33) a scattered
wave, resulting in a total wave function for large r similar to (9.143):


1
ei(pr(D3)/4)
(+)
ipx
hx|p i

e
+
(9B.40)
fp p ,
r
r(D1)/2
V
where fp p has the partial wave expansion generalizing (9.248):
fp p =

X
l=0

fl (p)

X
m

Ylm (
p )Ylm (
p) =

X
l=0

fl (p)

l + D/2 1 1 (D/21)
),
C
(
p p
D/2 1 SD l

(9B.41)

Spherical Square-Well Potential in D Dimensions

Appendix 9C

703

with
fl (p) =

(2)(D+1)/2
(2)(D1)/2
2(2)(D1)/2 il (p)
E Tl (p) =
[Sl (p) 1] =
e
sin l (p). (9B.42)
(D1)/2
(D1)/2
p
ip
p(D1)/2

Taking the aboslute square of this and integrating over all directions of the final momentum p we
find the partial wave decomposition of the total cross section:
=
=

d
p |fp p |2 =

X
l=0

|fl (p)|2

fl (p)fl (p)

XX

Ylm (
p )Ylm (
p)Ylm (
p )Ylm (
p)

lm l m

l=0

X
l+D/21 1 (l+D3)!
l+D/21 1 (D/21)
|fl (p)|2
Cl
(1) =
,
D/2 1 SD
D/2 1 SD l!(D 2)!
l=0

where we have used the special value of the Gegenbauer function17




l + 2 1
()
Cl (1) =
.
l

(9B.43)

(9B.44)

Inserting (9B.42), the cross section decomposes as


=

X
l=0

l =

4(2)D1 X l+D/21 1 (l+D3)! 2


sin l (p).
pD1
D/2 1 SD l!(D 2)!

(9B.45)

l=0

In three dimensions, this reduces to (9.251).

Appendix 9C

Spherical Square-Well Potential


in D Dimensions

For a square-well potential (9.266) in D dimensions, the radial Schrodinger equation (9.267) becomes
 2


d
D1 d
l(l + D 2)
2
l (r) = 0,
(9C.1)
+
+ kE
dr2
r dr
r2
where
K

p
2M (E + V0 )/h.

(9C.2)

There is only one solution which is regular at the origin:


l (r) =

A
JD/21+l (Kr),
(Kr)D/21

(9C.3)

with an arbitrary normalization factor A.


For positive energy E, the solution in the outer region r a receives an admixture of the
associated spherical Bessel function nl (kr):
l (r) =
where



B
cos l (k)JD/21+l (Kr) + sin l (k)ND/21+l (Kr) ,
(kr)D/21
k

17

2M E/h,

M. Abramowitz and I. Stegun, op. cit., Formula 22.2.3.

(9C.4)

(9C.5)

704

9 Scattering and Decay of Particles

and B, l (k) must be determined from the boundary condition at r = a. The associated Bessel
functions ND/21+l (z) have the asymptotic behavior orthogonal to (9B.28):
r
2
N (z)
cos(z /2 + /4),
(9C.6)
z
The continuity of the wave function fixes the ratio B/A. The continuity of the logarithmic derivative
is independent of this ratio and fixes the phase shifts by the equation
r0 [(Kr0 )1D/2 JD/21+l (Kr0 )]
(Kr)1D/2 JD/21+l (Kr0 ))
=

(9C.7)

r0 [cos l (k)(Kr0 )1D/2 JD/21+l (Kr0 ) sin l (k)(Kr0 )1D/2 ND/21+l (Kr0 )]
.
cos l (k)(Kr0 )1D/2 JD/21+l (Kr0 ) sin l (k)(Kr0 )1D/2 ND/21+l (Kr0 )

This determines the phase shifts as follows:


tan l (k) =

kJD/2+l (kr0 )JD/21+l (K0 r0 ) K0 JD/2 (K0 r0 )JD/21 (kr0 )


.
kYD/2+l (kr0 )JD/21+l (K0 r0 ) K0 JD/2+l (K0 r0 )YD/21+l (kr0 )

(9C.8)

For the s-wave in D = 3 dimensions, this reduces to


tan 0 (k) =

k cos kr0 sin K0 r0 K0 cos K0 r0 sin kr0


,
k sin kr0 sin K0 r0 K0 cos K0 r0 cos kr0

(9C.9)

which is, of course, the same as (9.274). In D = 2 dimensions, we obtain


tan 0 (k) =
which is for small k equal to
0

kJ1 (kr0 )J0 (K0 r0 ) K0 J1 (K0 r0 )J0 (kr0 )


,
kY1 (kr0 )J0 (K0 r0 ) K0 J1 (K0 r0 )Y0 (kr0 )

(9C.10)

K0 r0 J1 (K0 r0 )

,
2 J0 (K0 r0 ) + K0 r0 J1 (K0 r0 )log(kr0 e /2)

(9C.11)

where 0.577 216 is the Euler-Mascheroni number.


p
The phase shifts for a repulsive potential are obtained by setting K0 = i0 i 2M |V0 |/h
and replacing


2
J (K0 r0 ) i I (0 r0 ), Y (K0 r0 ) i iI (0 r0 ) (1) K (0 r0 ) ,
(9C.12)

In the limit
p Bessel functions I (z) diverge
of an infinitely high repulsive potential, the modified
like ez / 2z, while K (0 r0 ) go to zero exponentially fast like /2zez . Hence
tan 0 (k)

V0

JD/21 (kr0 )
.
YD/21 (kr0 )

(9C.13)

In D = 3 dimensons, the right-hand side becomes tan kr0 , implying that 0 (k) kr0 . In D = 2
dimensions, the right-hand side is J0 (kr0 )/Y0 (kr0 ) which has the small-k expansion
tan 0 (k) =

.
2 log(kr0 e /2)

(9C.14)

For D > 2, the small-k expansion of (9C.13) is


tan 0 (k)

V0

(kr0 /2)D2
.
(D/2)(D/2 1)

(9C.15)

which reduces correctly to kr0 for D = 3, but not to (9C.14) for D = 2, since the small z
behavior of Y0 (z) (2/) log z cannot be obtained as 0 -limit of the > 0 behavior Y (z)
(1/)()(z/2) .18
18

M. Abramowitz and I. Stegun, op. cit., Formulas 9.1.8 and 9.1.9.

Notes and References

705

Notes and References


For more details on scattering theory see
S. Schweber, Relativistic Quantum Fields, Harper and Row, N.Y., 1961.
The individual citations refer to:
[1] See Chapter 3 in
M.L. Goldberger and K.N. Watson, Collision Theory, John Wiley and sons, New York 1964.
[2] The unique extension of the linear space of distributions to include products is developed
in the textbook
H. Kleinert, Path Integrals in Quantum Mechanics Statistics, Polymer Physics, and Financial Markets, World Scientific, Singapore 2008 (www.physik.fu-berlin.de/~kleinert/b5).
[3] J.M.J. Van Leeuwen and A.S. Reiner, Physica 27, 99 (1961));
S.A. Morgan, M.D. Lee, and K. Burnett, Phys. Rev. A 65, 022706 (2002).
[4] N. Levinson, Dansk. Videnskab. Selskab. Mat.-fys. Medd. 25, Np. 9 (1949).

The only place outside of Heaven where you can be perfectly safe
from all the dangers and perturbations of love is Hell
C. S. Lewis (18981963)

10
Quantum Field Theoretic Perturbation Theory
The only place outside of Heaven where you can be perfectly safe from all the dangers
and perturbations of love is Hell. C. S. Lewis ...
In this chapter we would like to develop a method for calculating the physical
consequences of a small interaction in a nearly free quantum field theory. The
results wil all be expressed as power series in the coupling strength. This powers
series will have many unpleasant mathematical properties. These will be discussed
in later chapters. In this chapter, we shall ignore such problems and show only how
these power series can be derived in principle. More details can be found in the
textbook [1].

10.1

The Interacting n-Point Function

We consider an interacting quantum field theory with a time-independent Hamiltonian. All physical informations on the theory are carried by the n-point functions
G(n) (x1 , . . . , xn ) =

H h0|T H (x1 )H (x2 ) H (xn )|0iH .

(10.1)

Here |0iH is the Heisenberg ground state of the interaction system, i.e., the lowest
steady eigenstate of the full Schrodinger Hamiltonian:
H|0iH = E|0iH .

(10.2)

The fields H (x) are the fully interacting time-dependent Heisenberg fields, i.e., they
satisfy
H (x, t) = eiHt S (x)eiHt

(10.3)

H = H0 + V

(10.4)

where

706

10.1 The Interacting n-Point Function

707

is the full Hamiltonian of the scalar field. Note that the field on the right-hand side
of Eq. (10.3) has the time argument t = 0 and is therefore the same in any picture
I (x, 0) = H (x, 0) = (x, 0) so that we shall also write
H (x, t) = eiHt (x, 0)eiHt .

(10.5)

We now express H (x) in terms of the field I (x) of the interaction picture and
rewrite
G(n) (x1 , . . . , xn ) =

H h0|T

[UI (0, t1 )I (x1 )UI (t1 , t2 )I (x2 )


UI (tn1 , tn )I (xn )UI (tn , 0)] |0iH

(10.6)

where we have used the properties of the time displacement operator


UI1 (t, 0) = UI (t, 0) = UI (0, t),
UI (t1 , t2 ) = UI (t1 , 0)UI1 (0, t2 ).

(10.7)

We shall now assume that the state |0iH is a non-degenerate eigenstate of the full
Hamiltonian. Then we can make use of the switching-on procedure of the interaction,
and the vacuum state must develop for t towards the vacuum of the free
field Hamiltonian H0 . According to the Gell-Mann-Low formula we may write,
UI (0, ) |0i
,
h0|UI (0, ) |0i
h0|UI (, 0)
,
H h0| =
h0|UI (, 0) |0i
|0iH =

(10.8)

where |0i is the free-particle vacuum and the presence of the switching parameter
and its limit 0 at the end are tacitly assumed. Then formula (10.6) becomes


G(n) (x1 , . . . , xn ) = h0|UI (, 0) T UI (0, t1 ) I (x1 )UI (t1 , t2 ) I (x2 )

UI (tn1 , tn )I (xn )UI (tn , 0)) UI (0, ) |0i


1/h0|UI (, 0) |0ih0|UI (0, ) |0i.
(10.9)

The product in the denominator can be combined to a single expression using the
relation (9.96):
h0|UI (, 0) |0ih0|UI (0, ) |0i = h0|UI (, ) |0i = h0|S|0i.

(10.10)

The numerator consists of the S matrix operator UI (, ) time-sliced into n + 1


pieces at t1 , . . . , tn , with (xi ), i = 1, . . . , n, inserted successively. It is gratifying to
observe that due to the definition of the time-ordering operator, the expression can
be written in the much more compact fashion


T S I (x1 )I (x2 ) I (xn ) ,

(10.11)

708

10 Quantum Field Theoretic Perturbation Theory

so that we arrive at the simple formula




G(n) (x1 , . . . , xn ) = h0|T S I (x1 ) I (xn ) |0i/h0|S|0i


=

h0|T e

dt VI (t)
i

h0|e

(x1 ) I (xn )|0i

R I

dtVI (t)

(10.12)
.

|0i

The fields I (x) are now expressed as follows: by


I (x, t) = eiH0 t S (x)eiH0 t
= eiH0 t (x, 0) eiH0 t .

(10.13)

This implies that the field (x, t) changes in time in the same way as the Heisenberg field H (x, t) of the free Hamiltonian H0 would do with the initial condition
H (x, 0) = (x, 0). This observation will be the key to the evaluation of the n-point
functions.
What is the interaction picture of the interaction VI itself? We assumed V to
be an arbitrary spatial functional of S ,
V = V [S (x)] .

(10.14)

But then we may use (10.13) to calculate


VI (t) = V [I (x, t)] .

(10.15)

Thus the potential VI (t) in the interaction picture is simply the Schrodinger interaction V with the fields S (x) replaced by I (x, t) which develop from (x, 0) via
the free field equations of motion.
The state |0i is the ground state of the free Hamiltonian H0 , i.e., the vacuum
state arising in the free-field quantization of Chapters 3 and 5. If we drop the indices
I, we can state the interaction n-particle Green function as
i

G(n) (x1 , . . . , xn ) =

h0|T e

dtV [(x,t)]
i

h0|T e

(x1 ) (xn )|0i

dtV [(x,t)]

(10.16)

|0i

where (x, t) is the free field and |0i the vacuum associated with it.
Note that the functional brackets only hold for the spatial variable x. All fields
in V [(x, t)] have the same time argument. In a local quantum filed theory, the
functional is a psatial integral over a density
i

dt V [(x,t)] i

dt

d3 x v((x,t))

(10.17)

709

10.2 Perturbation Expansion for Green Functions

10.2

Perturbation Expansion for Green Functions

In general, it is very hard to evaluate expressions like (10.16). If the interaction


term VI is very small, however, it is suggestive to perform a power series expansion
and write
i

dtV [(x,t)]

= 1i
+

dtV [(x, t)]

2 Z
(i)

2!

dt1 dt2 T (V [(x1 , t1 )] V [(x2 , t2 )]) + . . . (.10.18)

In this way, we are confronted in (10.16) with the vacuum expectation value of many
free fields (x) of which n are from the original product (x1 ) (xn ), the others
from the interaction terms. From Wicks theorem we know that we may reduce the
expression to a sum over products of free-particle Green functions G0 (x x ) with
all possible pair contractions. The simplest formulation of this theorem was given in
terms of the generating functional of all free-particle Green functions [recall (7.818)],
Z0 [j] = h0|T ei

d4 x (x)j(x)

|0i,

(10.19)

where the subscript 0 emphasizes now the absence of interactions. This functional
can also be used to compactly specify the perturbation expansion. Let us introduce
the generating functional also for the interacting case, where the Green functions
are expectation values of products of Heisenberg field H (x) in the full Heisenberg
vacuum state,
ZH [j] H h0|T ei

d4 x H (x)j(x)

|0iH ,

(10.20)

The functional derivatives of this yield the full n-point functions (10.1). The perturbation expansion derived above can now be phrased compactly in the formula
ZH [j] ZD [j]/Z[0],

(10.21)

where Z[j] is the generating functional in the interaction or Dirac picture:


i

Z[j] h0|T e

dtV [(x)]+i

d4 x (x)j(x)

|0i.

(10.22)

The fields and the vacuum state in Z[j] are those of the free field theory. Note
that ZH [j] and Z[j] differ only by an irrelevant constant Z[0] which appears in the
denominators of all Green functions (10.16), and which has an important physical
meaning to be understood in Section 10.3.1. The main difference bestween them is
the prescription of how they are to be evaluated.
As functionals of the sources j, That this expression for the generating functional
yields the perturbation expansion for all n-point functions can be verified by functional differentiations with respect to j and comparison of the results with (10.16)
and (10.1).

710

10 Quantum Field Theoretic Perturbation Theory

Note that while the generating functional ZH [j] is normalized to unity for j 0,
this is not the case for the auxiliary functional Z[j]. Nevertheless, for generating
n-point functions, Z[j] is just as useful as the properly normalized ZH [j] if one only
modifies the differentiation rule by the overall factor Z[0]1 :
G

(n)

"

Z[j]
(x1 , . . . , xn ) =
Z[j] ij(x1 )
ij(xn )

(10.23)

j=0

This is what we shall do from on do so that we can refer to Z[j] as a generating


functional. This will also be of advantage when we come to enumerating the different
perturbative contributions to each Green function. Indeed, formula (10.23) allows for
an immediate even though formal and implicit solution for the interacting generating
functional: Since differentiation /j(x) produces a field (x) we may rewrite the
interaction V [(x)] as V [i/j(x)]. Then it is no longer a field operator and may
be removed from the vacuum expectation value so that Z[j] becomes
i

Z[j] = e

dtV [i/j(x)]

Z0 [j].

(10.24)

Now Z0 [j] was calculated explicitly in (5.438) from Wicks theorem




Z0 [j] = exp

1
2

d4 y1 d4 y2 j(y1 )G0 (y1 , y2 )j(y2 ) .

(10.25)

The perturbation series of all n-point functions are now found by expanding the
exponential in (10.24) in powers of V and performing the derivatives with respect
to /j(x). These produce precisely all Wick contractions involving the fields in the
interaction.
The explicit evaluation is quite difficult for an arbitrary interaction. It is therefore advisable to learn dealing with such expressions by considering simple examples.

10.3

Feynman Rules for 4 -theory

In order to understand the systematics of the perturbation expansion let us focus


attention on a very simple field theory with the Lagrangian
m2 2 g 4
1
+ .
L = ()2
2
2
4!

(10.26)

This is usually referred to as 4 -theory. Her m is the mass of the free particles
and g the coupling strength. We shall assume g to be small enough to be able to
expand all interacting Green function in a power series in g. It is well-known that
the resulting series will be divergent since the coefficients of g k at large order k will
grow like k!. Fortunately, however, the limiting behavior of the coefficients is exactly
known which has made it possible to develop powerful resummation techniques for
extracting reliable results from this series.

10.3 Feynman Rules for 4 -theory

711

The interaction in the Schrodinger picture is


V [S (x)] =

g
4!

d3 x4S (x).

(10.27)

In the interaction picture, after substituting S (x) by the free field (x), the exponent in formulas (10.16), (10.22) becomes
i

dtV [S (x,t)]

= ei 4!

d4 x 4 (x)

(10.28)

In the functional formulation of the perturbation expansion we have to calculate the


series
g

Z[j] = ei 4!
=

"

d4 x(i/j(x))4

g
1i
4!

(i)2
+
2!
R
1

e 2

Z0 [j]

d x i
j(x)

g
4!

2 Z

!4

i
j(x1 )

d x1 d x2

d4 y1 d4 y2 j(y1 )G0 (y1 ,y2 )j(y2 )

!2

i
j(x2 )

!2

+ ...

(10.29)

The n-point functions are obtained according to (10.23) by expanding the exponential on the right-hand side in a power series, forming the nth functional derivatives
with respect to j, and setting j to zero. The result has to be divided by Z[0] which
is also a power series in g. Certainly, n has to be even, otherwise the result vanishes.
It we want to calculate G(n) up to a given power in g, say g k , there are many different contributions. The denominator Z[0] has to be expanded in powers of g and its


2k

/(2k)! term in the expansion


kth- order contributions come from the 21 jG0 j
of the exponential. Here and in most of the following structural formulas we shall
omit the integration variables, for brevity. The k th -order term has the form
g
Zk [0] = i
4!


k Z

i
j1

!4

i
jk

!4

1
1

(2k)!
2


jG0 j

2k

(10.30)

In the numerator of (10.23), there are contributions of zeroth order in g to G(n)




from the (n/2)th term which has the form 12 jG0 j




the (n/2 + 2)th term 12 jG0 j




n/2+4

n/2+2

n/2

, of first order in g form

, of second order in g from the (n/2 + 4)th

term 21 jG0 j
, etc. Forming the 4 derivatives (/j)4 associated with every
order in g, as well as the n derivatives for the Green function G(n) , the expression
to k th order have the structure
R

g
i
4!

k

i
i
j1
jn

i
j

!4

i
j

!4

1
1
jG0 j
(n/2 + 2k)!
2


 n +2k
2

(10.31)

712

10 Quantum Field Theoretic Perturbation Theory

The Green functions accurate to order g k are then obtained by dividing the two
power series through each other and expanding the result again up to order g k .
This all seems to be a horrendous task. It is, however, possible to devise a
diagrammatic procedure for keeping track of the different contributions which will
show that many simplifications arise. In particular, the division process is really
quite trivial due to the fact that Z[0] appears automatically as a factor in the
calculation of the numerator of each n-point function.
Actually formula (10.29), even though giving the most explicit answer to the
problem, is quite cumbersome when it comes to actual calculations. The derivatives
are an efficient analytic way of accounting for the set of all Wick contractions between
the field operations. In the calculation of a specific n-point function, however, it
is much more advantageous to insert the expansion (10.18) into formula (10.16),
separate numerator and denominator by writing
G(n) (x1 , . . . , xn )

1 (n)
G (x1 , . . . , xn ) ,
Z[0]

(10.32)

(n) (x1 , . . . , xn ) having the expansion


with the unnormalized Green function G
(n) (x1 , . . . , xn )
G

 Z
Z

n
1 ig 2
ig
d4 z1 d4 z2 4 (z1 )4 (z2 ) + . . .
d4 z(z) +
= h0|T 1
4!
2! 4!
o
(x1 ) (xn ) |0i,
(10.33)

and the denominator Z[0] of (10.32)


ig
Z[0] = h0|T 1
4!

1 ig
d z (z) +
2! 4!
4

2 Z

d z1 d z2 (z1 ) (z2 )+ . . . |0i,

(10.34)
(n)
and to perform the Wick contractions in the two expansion explicitly yielding G
p
and Zp [0], respectively.

10.3.1

The Vacuum Graphs

Because of its formal simplicity let us start a more explicit perturbation expansion
with the calculation of Z[0]. To first-order in the coupling constant g we have to
find
g Z 4
Z1 [0] = i
d zh0|T ((z)(z)(z)(z)) |0i,
(10.35)
4!
where we have written down the four powers of (z) separately see better how to
perform all pair contractions. The first field can be contracted with the three others.
After this the second field has only one choice. Thus there are 3 1 contractions, all
of the form
Z
g
d4 zG0 (z, z)G0 (z, z),
(10.36)
i
4!

10.3 Feynman Rules for 4 -theory

713

so that
g
Z1 [0] = i3
4!

d4 zG0 (z, z)G0 (z, z).

(10.37)

To order g 2 we have to evaluate


1
g 2
d4 z1 d4 z2
i
2!
4!
h0|T ((z1 )(z1 )(z1 )(z1 )(z2 )(z2 )(z2 )(z2 )) |0i.


 Z

(10.38)

Now there are 7 5 3 1 = 105 contraction, 32 of them with (z1 )s and (z2 )s
contracting among each other, for example,

(z1 )(z1 )(z1 )(z1 )|(z2 )(z2 )(z2 )(z2 )

(10.39)

where we have explicitly separated the two interactions by a vertical line. There
are further 4 3 2 = 24 contractions, where each (z1 ) connects with a (z2 ), for
example,

(z1 )(z1 )(z1 )(z1 )|(z2 )(z2 )(z2 )(z2 )

(10.40)

and 6 6 2 = 72 of the mixed type, for example,


(z1 )(z1 )(z1 )(z1 )|(z2 )(z2 )(z2 )(z2 )

(10.41)

The factors six counts the six choices of one contraction within each factor 4 after
which there are only two possible interconnections.
The 105 terms obtained in this way correspond to the following integrals
1
g
i
2!
4!


2 " Z

d4 z1 G0 (z1 , z1 )2

+ 72

2

+ 24

d4 z1 d4 z1 G0 (z1 , z2 )4
2

d zd zG0 (z1 , z1 )G0 (z1 , z2 ) G0 (z1 , z2 ) G0 (z2 , z2 ) . (10.42)

It is useful to picture the different contributions by means of so-called Feynman


diagrams: A line with x1 , x2 at the ends
q
x1

q
x2

= G0 (x1 , x2 )

(10.43)

represents a free-particle propagator. A vertex with four emerging lines

= i

g
4!

(10.44)

714

10 Quantum Field Theoretic Perturbation Theory

stands for the 4 (z) interaction at the point z with the convention to carry a coupling
constant ig/4!. The spacetime variables of each vertex have to be integrated over.
Then the only diagram to first order is
3

q
.

(10.45)

z1

To second order there are three diagrams


9

q
z1

q
z2

........q
q q
.
........
+ 72
.
+ 24 ...q..................

z1 z2

(10.46)

z1 z2

To third order we finds 11 9 7 5 3 1 = 10395 terms. The total number rapidly


proliferates.
Diagrams of this type consisting only of lines which close back with each other
are called vacuum diagrams.
When discussing Eq. (7.132) we have noted an important statstical interpretation
of the relativistic Euclidean propagator G(x, x ). It describes the probability for a
random walk of any length to go from x to x , provoded that its lengths are
distributed with a Boltzmann-like factor e . The loop expansion of the partition
function in terms of vacuum diagrams may therefore be interpreted as a direct
picture of the various topologies of random lines in a grand-canonical ensemble of
lines of any length. For this reason, relativistic quantum field theories may be used
to study such line ensembles, in which case they are called disorder field theories.
As mentioned in Chapter 7, such lines appear in many physical systems in the form
of vortex lines and defect lines.
This line interpretation of quantum fields has led to an entire quantum field theory of physical systems in which the statistical mechanics of line like excitations play
an important role in understanding the observed behavior. Consider, for example,
the superfluid phys of He. Conventionally, its behavior is understood by describing
it as an ensemble of a large number of atoms interacting by a strong van der Waals
type of interaction. At low temperatures, below the so-called -point T 2.17 K,
the atoms enter the superfluid phase in which all atoms behave in coherent fashion. At zero temperature the entire system lies in a ground state. As temperature
rises, thermal fluctuations create small loops of vortex lines. Their average length
grows, and at T it diverges. The vortex lines proliferate and fill the entire sample.
Since the inside of each vortex line contains a normal liquid, the superfluid becomes
normal. This picture gives rise to a completely alternative quantum field theoretic
decription of superfluid He. At zero temperature, the superfluid is a vacuum for
vortex lines, i.e., for the disorder field describing them. As the temperature rises,
more and more disorder excitations are generated, and the field acquires a finite
expectation value.

715

10.4 The Two-Point Function

The reader who wants to understand this interesting development is referred to


the original literature1 . Some details will also be discussed in Chapter 20.2.
At this place is is also worth mentioning that the opposite direction of research
has been pursued by a number of people. The understanding of field theory from it
description as an ensemble of lines. The formalism arising in this way is refereed to as
string-inspired approach to quantum field theory. It abandons the marvellous power
of the field theoretic description of ensembles of lines in favor of some calculational
advantages [5]. By construction, this approach conserved the number of lines, i.e.,
the number of particles. Thus it should not be used when particle condensation is
important.

10.4

The Two-Point Function

Let us now turn to the numerator in the perturbation expansion for the n-point
function (10.29). We shall first study the two-point function. Clearly, to zeroth
order there is only the free-particle expression
(2) (x1 , x2 ) = h0|T (x1)(x2 )|0i = G0 (x1 , x2 )
G

(10.47)

corresponding to the Feynman diagram (10.43). To first order we have to find all
contractions of the expression
ig
4!

d4 zh0|T (z)(z)(z)(z)(z1 )(z2 )|0i.

(10.48)

There are 5 3 1 = 15 of them. They fall into two classes: 3 diagrams contain
contractions only among (z)s multiplied by a contraction of (x1 ) with (x2 ).
Analytically, they correspond to
ig
3
4!


Z

d4 zG20 (z, z)G0 (x1 , x2 ) ,

(10.49)

i.e., this expression carries the same factor 3 that was found in the calculation (10.36)
of the vacuum diagrams by themselves. The diagrammatic representation consists
of a line and the vacuum diagram side by side
3 xq1

q
x2

q
.

(10.50)

Such a diagram is called disconnected. In general, the analytic expression corresponding to disconnected diagrams is the product of the expressions corresponding
to the individual pieces.
The second class of first-order diagrams collects the contractions between (z)s
and (x1 ) or (x2 ). There are 12 of them with the analytic expression
12
1

See Refs. [2, 3, 4].

ig
4!

Z

d4 zG0 (x1 , z) G0 (z, z) G0 (z, x2 ) .

(10.51)

716

10 Quantum Field Theoretic Perturbation Theory

They are pictured by the connected Feynman diagram

q q q .
x1 x2

(10.52)

Thus the expansion to first order has the diagrammatic expansion


(2) (x1 , x2 ) = q
G
x1

q
x2

+ 3

q
x1

q
x2

q q q
x1 x2

+ 12

(10.53)

Remembering the expansion of the denominator to this order

q
.

Z[0] = 1 + 3

(10.54)

we see that the two-particle Green function is to order g given by the free diagram
and the diagram in Fig. 10.52.
G(2) (x1 , x2 ) = xq1

q
x2

+ 12 q
q q .
x1

(10.55)

x2

The disconnected pieces involving the vacuum diagram have canceled.


we have to form all
Consider now the second-order contributions. To obtain G
contractions of

 Z
1 ig 2
d4 z1 d4 z2 h0|T (z1 ) (z1 )(z1 )(z1 )(z2 )(z2 )(z2 )(z2 )
2! 4!
(x1 )(x2 )|0i.
(10.56)
There are 9 7 5 3 1 = 945 of them. These decompose into three classes. The first
is disconnected and contains the 105 vacuum diagrams multiplied by G0 (x1 , x2 ),
1 q
2! x1

q
x2

+ 24

q
q

....
................ ...............
...

+ 72

q q

(10.57)

The second consists of mixed contributions in which the first order correction to G
combined with a first-order vacuum diagram
!

q
.
36 2 q qq
2!
x1 x2
The third contains only the connected diagrams
1
144 2 q
2!
x1

q q
x2

+ 144 2

q
q qq
x1
x2

(10.58)

qq
+ 96 2 xq1q
x2 .

(10.59)

In order to calculate the two-point function up to order g 2 we have to divide


1 , x2 ) =
G(x

q
x1

1
+ xq1
2!

q
x2

x1

q
x2

+36 q
q q
x1

+ 12 q qq + 3 xq1

x2

q
x2

x2

+ 24

q
+ 144 q
q q
x1 x2

q
q

....
................ ...............
...

+ 144

+ 72

q q q q
x1
x2

q q

(10.60)

qq
+ 96 xq1q
x2

717

10.5 The Four-Point Function

by Z[0] calculated up to the same order. This consists of the diagrams


1
q
9
Z[0] = 1 + 3
+
2!


........q
q q
........
+ 72
.
+ 24 ...q...................

(10.61)

1 , x2 ) by Z[0] gives the two-point function


Dividing G(x
G(2) (x1 , x2 ) = xq1

q
x2

+ 12 q qq
x1

1
+
144 2 q
2!
x1

10.5

q q
x2

(10.62)

x2

+ 144 2

q
q q
q
x1
x2

qq
+ 96 2 xq1q
x2

+ ... .

The Four-Point Function

Let us now study the four-point function. To zeroth order, the numerator has the
following trivial contributions in which all particles propagate freely
x3

(4) (x1 , x2 , x3 , x4 ) =
G
x4

x1 x3
+
x2 x4

x1 x3
+
x2 x4

x1

.
x2

To first order we must form all contractions in


ig
4!

d4 z1 d4 z2 h0|T (z1 ) (z1 )(z1 )(z1 ) (x1 )(x2 )(x3 )(x4 )|0i.

which yield the diagrams

q
.

(10.63)

718

10 Quantum Field Theoretic Perturbation Theory

q
We observe again the appearance of a factor 1 + 3
containing the first-order

vacuum diagram, which is canceled when forming the quotient (10.32). Thus G(4)
can be written to this order as
G(4) (x1 , . . . , xn ) =

We now turn to the second-order diagrams for which we must form all contractions
in (10.56) after exchanging its second line by (x1 )(x2 )(x3 )(x4 )|0i. Their total
(4) (x1 , x2 , x3 , x4 ). These can be
number is 11 9 7 5 3 1 = 10395 diagrams in G
grouped into 105 vacuum diagrams of second order multiplied with the previously
(4) (x1 , x2 , x3 , x4 ):
calculated zeroth-order diagram in G

.
Then there are those in which the vacuum diagrams appear in first order

Finally, there are 9504 terms without vacuum contributions

If the vacuum diagrams are divided out we remain with

719

10.6 Connected Green Function

G(4) (x1 , . . . , xn ) =

(10.64)

10.6

Connected Green Function

Faced with the rapid proliferation of diagrams for increasing order in the coupling
constant, there is need to economize the calculation procedure. The cancellation of
all disconnected pieces involving vacuum diagrams was a great simplification. But
the remaining diagrams are still many, even at low order in perturbation theory.
Fortunately, not all of these diagrams really require a separate calculation. First
of all, there are many diagrams which consist of disconnected pieces, each of which
already occurs in the expansion (10.62) of the two-point function. since the corresponding amplitude factorizes into the product of the individual expressions, each
of these diagrams being really known from the calculation of G(2) . Thus, we can
save a great deal of labor if we separate the connected diagrams in G(4) and consider
them separately. They are called the connected four-point function and denoted by
G(4)
c (x1 , x2 , x3 , x4 ). Their low-order expansion is simply
G(4) (x1 , . . . , x4 ) =

.
We only have to learn how to recover the full Green function from the connected
one and the omitted diagrams which are all known from G(2) . In the example it can
easily be verified that these omitted parts are simply the product of two propagators

720

10 Quantum Field Theoretic Perturbation Theory

G(2) together with the three perturbative corrections on of the external legs
(2)
G(2)
c Gc

+ 2 perm =

1
+
144 2
2!

q
x1

q q q
x1 x2

q
x2

+ 12 q qq
x1

+ 144 2

x2

q q q q
x1
x2

+ 96 2

qq
qq
x1
x2

)2

+ . . . + 2perm

.
We shall see later that this is a completely general law if the system is in the socalled normal phase as opposed to the condensed phase to be studied separately in
Chapter 12. The connectedness structure in the normal phase is
i

(2)
(2)
. 10.65)
G(4) (x1 , . . . , xn ) = G(4)
c (x1 , . . . , xn ) + Gc (x1 , x2 ) Gc (x3 , xn ) + 2 perm (

Note that the expansion (10.62) of G(2) is connected. This is a general feature of
the normal phase of the theory.
For the higher Green functions we expect more elaborate connectedness relations
than (10.65) and an even more drastic reduction of labor using these relations when
calculating all diagrams. The question arises as to the general composition law of
n-point functions from connected subunits.
To gain a first idea what this law could be, consider the free theory. Its generating
functional is [see (5.438)]


Z0 [j] = exp

1
2

d4 x d4 y j(x)G0 (x, y)j(y) .

(10.66)

If expanded in powers of j, it gives the sum of all n-point functions of the free theory.
In accordance with Wicks theorem all these free n-point functions are disconnected
and consist of sums of products of free Green function G0 which itself is the only
connected diagram of the theory. The important point is that the exponential tells
us in which way the connected diagrams G0 can be combined such as to form all
diagrams. This may best be visualized diagramically by expanding (10.66) in the
form

721

10.6 Connected Green Function

The numbers behind 1/n! in the second line show how many combinations of n/2
powers of G0 occur in Wicks expansion. To obtain these numbers we have rewritten
1
the (n/2)th coefficient (n/2)!
(1/2)n/2 as n!1 (n 1)!!. This establishes contact with the
previous counting rules: The denominator n! is factorized out since it is canceled
when going to the n-point function (which involves n differentiations /j). This
leaves (n 1)!! diagrams in agreement with the result found earlier when counting
the diagrams directly. Thus we have verified, in the free-field case the simple rule
for the reconstruction and proper counting of all n-point functions given only the
connected ones (of which there is only one in this case). By expanding diagramically
the exponential of the connected diagrams, which is here exp{ 12 }, we can
read off all connected plus disconnected diagrams behind the factors 1/n! . In this
way, the exponential of the connected diagram yields all diagrams.
Does this simple statement hold also in the interacting case? Here the generating
functional is given by
( Z

Z0 [j] = exp i

d x Lint

ij

!)

exp

1
2

d4 x d4 y j(x)G0 (x, y)j(y) . (10.67)

The interactions also enter exponentially. It is therefore suggestive that also here
the sum of all Green functions can be obtained by exponentiating all connected
ones. The proof will be given later after having developed more powerful formal
techniques. Let us state here only the result which may be written as a relation

1+

(n)

1 (n)
1 (n)
Gk = exp
Gc k ,

n=0,k=0 n!k!
n=0,k=1 n!k!

(10.68)

(n)

where Gk are all and Gc k all connected diagrams in kth-order perturbation theory
and the equality holds separately for each number of external lines.
Let us illustrate the statement (10.68) for the previously calculated diagrams for
n = 2 and n = 4. The left-hand side of relation (10.68) looks as follows

722

10 Quantum Field Theoretic Perturbation Theory

The right-hand side has the form

Indeed, by multiplying out the square in the last line we recover the correct sum of
disconnected diagrams of the four-point function.
Also the vacuum diagrams satisfy the law of exponentiation: Up to the second
order we have for all disconnected pieces
Z[0] = 1 + 3

1
9
+
2!


+ 24

q
q

....
.................................
.

+ 72

q q

(10.69)

and this can be obtained as an exponential of the connected vacuum diagrams


(

Z[0] = exp 3

exp{W [0]}.

1
........q
q q
.......
+
+ 72
24 ...q....................
2!


)

(10.70)

723

10.6 Connected Green Function

This is equation (10.68) for n = 0:

X
1 (0)
1 (0)
1+
Gk = exp
Gc k ,
k=0 k!
k=1 k!

"

(0)

(10.71)

(0)

where Gk collect all vacuum diagrams and Gc k all connected ones in kth-order
perturbation theory. This exponential relation will be of great help when it comes
to calculating physical amplitudes and cross sections.

10.6.1

One-Particle Irreducible Graphs

The decomposition into connected diagrams does not yet exhaust the possibilities
of reducing calculational labor. If we inspect the connected diagrams for two and
four-point functions
G(2) (x1 , x2 ) = xq1

q
x2

+ 12 q
q q +
x1

1
+
144 2
2!

x2

q q q
x1 x2

+ 144 2

q
q
q q
x1
x2

qq
+ 96 2 xq1q
x2

+ . . . (10.72)

and

G(4) (x1 , . . . , x4 ) =

,(10.73)
.

we discover that some of the diagrams contain a portion of others calculated at


a lower order of perturbation theory in the same connected two-point functions G(2)
c
(2)
or G(4)
c . An example is the fourth diagram in Gc . which is a simple repetition of
the second one. Similarly, the second diagram in G(4)
c is the composition of the first
(4)
(2)
in Gc with the second in Gc . It would be useful to find the rule according to
(4)
(2)
(4)
which lower subdiagrams of G(2)
c , Gc reappear in higher ones of Gc and Gc .
As far as G(2)
is concerned, this rule turns out to be really simple: Let us
c
characterize the repetition of a former subdiagram of G(2)
topologically by noting
c
that the diagram falls into two pieces by cutting one internal line. Such diagrams
are called one-particle reducible (OPR); otherwise irreducible (OPI). Then the full
connected two-point function G(2)
may be composed from all OPI subdiagrams as
c
follows: Consider the set of all OPI diagrams to the two-point function. They all
(2)
carry a free Green function G0 at the end of each leg which describes propagation of
the particle up to the first interaction vertex. Cutting off these last Green functions

724

10 Quantum Field Theoretic Perturbation Theory

amounts diagramically to amputating the two legs of the diagram. The lowest order
correction to the two-point function is amputated as follows:

The two short little trunks indicate the places of amputation. Let i be the sum
of all these amputated OPI two-point functions. Then the geometric series
G(2)
c =

1
= G0 + G0 (i)G0 + G0 (i)G0 + . . .
+

G1
0

(10.74)

gives precisely the connected two-point function G(2)


c . Thus the one-particle reducibility in the two-point function exhausts itself in a simple geometric series type
of repetition of the irreducible pieces, each term in the string having the same factor.
Also this result will be proved later in Chapter 14 when studying the general formal
properties of perturbation theory.
The sum of all OPI connected two-point functions i is usually referred to as
self-energy.
Consider now the four-point function G(4)
c . Here we recognize that any ornamentation of external legs can be taken care of by replacing this leg by the interacting
two-point function. Thus we decide to introduce the concept of an arbitrary oneparticle irreducible amputated Green function shortly called the vertex function
(n) (x1 , . . . , xn ). For any connected n-point function, cut all simple lines such that
the diagrams decompose. What remains are parts with two, four, or more trunks
sticking out. The first set consists of the OPI self-energy diagrams discussed before.
The others are called three-, four- etc. point vertex parts (n) , n = 3, 4, . . . . For
example,

can be cut into four proper self-energy diagrams and one four-point vertex part.
The sum of all composite diagrams obtained in this way is the n-point vertex function denoted by . The important reconstruction principle for all diagrams can now
be states as follows: The set of all connected diagrams in a four-point function is
obtained by connecting all vertex functions in the four-point vertex function G(4)
c
with the full connected Green function G(2)
c at each truncated leg. Analytically, this
amounts to the formula (valid in normal systems)

725

10.6 Connected Green Function

G(4)
c (x1 , . . . , x4 )

(2)

(4)

d4 x1 d4 x2 d4 x3 d4 x4 G(2)
c (x1 , x1 ) Gc (x4 , x4 ) (x1 , . . . , x4 ).

(10.75)
For the Green function G(4)
c as far as it has been calculated in (10.73) we see that
it can indeed be decomposed in this way with the vertex function, up to order g 2,
i(4)
c (x1 , . . . , xn ) =
(10.76)
and
G(2) (x1 , x2 ) = xq1

q
x2

+ 12 q
q q .
x1

(10.77)

x2

This decomposition of Green functions in terms of vertex functions shows its


particular strength when going to higher orders in perturbation theory. Then the
number of diagrams to be calculated is greatly reduced. For example, the thirdorder contribution to the vertex function i(4)
c (x1 , x2 , x3 , x4 ) are
i(4)
c (x1 , x2 , x3 , x4 )
.(10.78)
We leave it up to the reader to compare this with the diagrams in the connected
3
four-point function G(4)
c up to g .
For theories more general than 4 the composition law is more involved and will
be discussed in Chapter 6.

10.6.2

Momentum Space Version of Diagrams

The spacetime formulation of Feynman rules will be inconvenient when it comes


to an explicit evaluation of diagrams. It will be of great advantage to exploit the
translational invariance of the theory by going to momentum space. The free Green
function G0 has the very simple Fourier transform
G0 (q1 , q2 )
=

d4 x1 d4 x2 ei(q1 x1 +q2 x2 ) G0 (x1 , x2 )


d4 x1 d4 x2 ei(q1 x1 +q2 x2 )

= (2)4 (4) (q1 + q2 ) G0 (q1 )

i
d4 iq(x1 x2 )
e
4
2
(2)
q m2

(10.79)

is the propagator in momentum space. There is an overall (2)4 (4) -function which
ensures the conservation of four momenta. This is a consequence of the translational
invariance of G0 (x1 , x2 ) = G0 (x1 x2 ). The same factor appears in the Fourier

726

10 Quantum Field Theoretic Perturbation Theory

transform of all interacting n-point functions since G(n) (x1 , . . . , x1 ) depends only on
the differences between the coordinates
G(n) (x1 , . . . , xn ) = G(n) (x1 xn , x2 xn , . . . , xn1 xn , 0)

(10.80)

such that we can write


Z

d4 x1 . . . d4 xn eii=1 qi xi G(n) (x1 , . . . xn )

d4 (x1 xn ) d4 (xn1 xn )ei

Pn1
i=1

qi (xi xn )

Z

d4 xn eii=1 qixn

G(n) (x1 xn , x2 xn , . . . , xn1 xn , 0) .

(10.81)

Thus we may define the Fourier transform of an n-point function directly with the
factor of momentum conservation removed
(2)4 (4) (q1 + . . . + qn ) G(n) (q1 , . . . , qn )

d4 x1 d4 xn eii=1 qi xi G(n) (x1 , . . . , xn ) .


(10.82)

Consider now the vacuum diagrams evaluated via the Fourier transforms. To first
order we have
3

g
= 3i
4!

g
d4 z G2 (z, z) = 3i
4!

d4 z

"Z

d4 q
i
4
2
(2) q m2

#2

. (10.83)

The integral over z can be defined meaningfully only if the system is enclosed in a
finite box of volume V and studied in a finite time interval T . Then the integral
R 4
d z gives a factor V T . This would become infinite for large V T which is called
the thermodynamic limit.
If V T is finite, there is still a divergence coming from the integral over the
momenta p at large p. This is called ultraviolet divergence. It reflects the singularity
of G0 (x1 , x2 ) for x1 x2 (a so-called short-distance singularity). It will be the
subject of the next chapter to show how to deal with this type of divergence. For
(2) (x1 , x2 ) the diagram of first order in g is
G
12 q qq
x1

x2

= 12i

g
4!

dzG(x1 , z)G(z, z)G(z, x2 ) .

(10.84)

Going to the Fourier transform this gives


Z

i
d4 q iq(x1 x2 )
e
(2)4
q 2 m2

d4 k
i
(2)4 k 2 m2

i
,
q 2 m2

(10.85)

such that we obtain for the Fourier transformed Green function a contribution
i
(2) (q) = i g 12
G
4! q 2 m2

d4 k
i
i
.
(2)4 k 2 m2 q 2 m2

(10.86)

727

10.7 Green Functions and Scattering Amplitudes

As another example take


k+q1 +q2
q2
q.4.
..
..
q
...q....
...

.....
.
....
....

q3

k2

q1

ig
=
4!


2 Z

d-4 k

k2

i
i
2
m + i (k + q1 + q2 )2 m2 + i

(10.87)

It is easy to see that the following rules hold for the translation of the spacetime
diagrams to the analytic expression for the Fourier transformed Green function:
1. With each line associate a momentum label and specify its direction of flow.
Then the line corresponds to the Fourier transformed two-point function G0 (q),
the free-particle propagator:
=

G0 (q) =

q2

i
.
m2

(10.88)

Here the direction of the arrow does not matter since G0 is symmetric in q.
2. Each vertex

is associated with an amplitude


g
i
i
i
i
i (2)4 (4) (q1 + q2 + q3 + q4 ) 2
2
2
2
2
2
2
4!
q4 m q3 m q2 m q1 m2
containing a four momentum conservation (2)4 (4) function for the incoming
momenta. This is a consequence of the integration over z in the interaction.
3. We now distinguish external lines with at least one open and internal lines
which connect two vertices with each other. We also call the corresponding
momenta qi external and internal. Then for every internal, lines integrate over
R
all the internal four momentum qi with the measure d4 qi /(2)4 . At the end
remove an overall -function of energy-momentum conservation.
Obviously, all (2)4 (4) -functions associated with the vertices cancels one internal
R
momentum integral d4 p/(2)4 . Thus in a diagram of nth order with I internal
lines, only I n internal integrations remain. These are referred to as loop integrals.
In general, loop integrations diverge at large loop momenta, in the so-called
called the ultraviolet regime. We shall learn in the next chapter how to deal with
such divergent momentum space integrals.

10.7

Green Functions and Scattering Amplitudes

The Green functions carry all informations contained in the theory. In particular
they can be used to extract scattering amplitudes. For definiteness, let us discuss

728

10 Quantum Field Theoretic Perturbation Theory

here the simplest and most important case of the elastic scattering among two
particles. The free initial state long before the interaction takes place is
|in i = aq2 aq1 |0i.

(10.89)

Long after the interaction the state is given by


UI (, ) aq2 aq1 |0i

(10.90)

If we analyze this state according to its free-particle content we find the amplitudes
h0|aq4 aq3 UI (, ) aq2 aq1 |0i = h0|aq4 aq3 S aq2 aq1 |0i

(10.91)

We shall soon observe that this amplitude has a divergent phase as the switching
parameter tends to zero. This is caused by the same vacuum diagrams as before
in the corresponding Green function. In order to obtain a well-defined 0 -limit
we define the 2 2 scattering amplitude as the ratio
S (q4 , q3 |q1 , q2 )

SN (q4 , q3 |q1 , q2 )
,
Z[0]

(10.92)

(10.93)

with the numerator


i

SN (q4 , q3 |q1 , q2 ) h0|aq4 aq3 T e

dt VI (t)
aq2 aq1 |0i

and the denominator


i

Z[0] h0|e

dt VI (t)

|0i.

(10.94)

We shall often use the four-momentum notation S (q4 , q3 |q1 , q2 ) for S (q4 , q3 |q1 , q2 )
with the tacit understanding
that for the S-matrix, the energies are always on the

mass shells q 0 = q2 + m2 .
It is now easy to see how these amplitudes can be extracted from the Green
functions calculated in the last section. There exists a mathematical framwork to
do this known a the Lehmann-Symanzik-Zimmermann formalism (LSZ-reduction
formulas) [6]. Rather than presenting this we present here a simple pedestrian
approach with the same results.
For this we observe that aq , aq can be written as
aq =
aq

lim
0

lim

x0

2q 0 Z 3 i(q0 x0 qx)
d xe
(x),
V

(10.95)

2q 0
V

(10.96)

d3 x ei(q

0 x0 qx)

(x),

if q 0 is placed on the mass shell of the particles, i.e., q 0 = q = p2 + M 2 . The


infinite-time limits have the important effect of eliminating the undesired frequency

729

10.7 Green Functions and Scattering Amplitudes

contents in (x). Indeed, if we expand the field into creation and annihilation
operators we see that the right-hand side of Eq. (10.95) becomes
lim
0



2q0 Z 3 i(q0 x0 qx) X
1
0
d xe
eipx ap + c.c.
V
2p V
q

= 0lim

x p

p,q ei(q

0 p0 )x0

ap + ei(q

0 +p0 )x0

ap .

(10.97)

The spatial -function enforces q = p and thus q 0 = p0 , so that the right-hand side
becomes
aq + 0lim ei2q
x

0 x0

aq .

(10.98)

In the limit x0 , the second exponential function oscillates infinitely rapidly.


Such an oscillating expression can be set equal to zero. The argument why this
can be done uses the fact that no physical state is completely sharp in momentum
space but contains some, possibly very narrow, distribution function f (q q ) in
the momenta. Thus, instead of aq , we really deal with a packet state
Z

d3 q
f (q q )aq
(2)3

with f (q q ) sharply peaked around q. Then Eq. (10.103) has to be smeared out
with such a function, and the second term in (10.98) becomes
lim
0

d3 q i2q0 x0
e
f (q q )aq 0 .
(2)3

(10.99)

The vanishing of this in the limit x0 is a well-known consequence of the


Riemann-Lebesgue Lemma. The other equation (10.96) is proved similarly.
We can now make use of formula (10.96), replace the operators aq , aq by timeordered fields (x), and obtain for the numerator part of the S matrix elements in
Eq. (10.93) the following expression:
SN (q4 q3 |q2 q1 ) =

24 q10 q20 q30 q40


0 0
0 0
0 0
0 0
ei[q4 x4 q4 x4 +q3 x3 q3 x3 q2 x2 +q2 x2 q1 x1 +q1 x1 ]
lim
2
0
0
x >x
V
1
2
x0 <x0
4
3

h0|T (x4)(x3 )S(x2 )(x1 )|0i.

(10.100)

The last factor is precisely the four-point function G(4) (x4 , x3 , x2 , x1 ).


This formula looks somewhat cumbersome to implement in an actual calculation of the scattering amplitudes and it is useful to simplify it by evaluating the
infinite-time limits more explicitly. We observe that the perturbation series for
G(n) (xn , . . . , x1 ) consists of sums of products of free two-point functions G0 which
contain for each spacetime argument x1 , x2 , . . . , xn in G(n) a two-point function G0

730

10 Quantum Field Theoretic Perturbation Theory

whose line ends at that point. Consider, for example, the point x1 , and an associated
G0 (z1 , x1 ). The operation (10.97) at this point corresponds to taking the limit
s

2q10
lim
V x01

i(q10 x01 q1 x1 )

d x1 e

d4 q iq(zx1 )
i
e
.
4
2
(2)
q m2 + i

(10.101)

The spatial integral over x1 enforces q = q1 . The remaining integral over dq 0 can
be done via Cauchys residue theorem. Since x01 , the contour of integration
0
may be closed in the
q lower half of the complex q -plane, where it contains only a
pole at q 0 = q1 = q21 + m2 i. Thus the integral over q 0 can be done trivially,
and we find, that the limit (10.100) has the effect of replacing in the Green functions
G(z, x1 ) of the incoming external leg as follows:
1
G(z, x1 )

q
eiq1 z ,
0
2q1 V

with q10 on the mass shell q10 =


q

(10.102)

q21 + m2 . The right-hand side is simply the wave

function eiq1 z / 2q10 V of the incoming particle with the argument z of the nearest
vertex in the Feynman diagram.
Similarly, we obtain for an outgoing particle of momentum q3 the replacement
1
G(x3 , z)

q
eiq3 z .
0
2q3 V
As a specific example, consider the simple vertex diagram in (10.73):
= ig

(10.103)

d4 z G(x4 , z)G(x3 , z)G(z, x2 )G(z, x1 ).

(10.104)

Taking the limits (10.100) we obtain


4
Y

i=1

2V

qi0

(ig)

= ig
q

4
Y

i=1

d4 z ei(q4 z+q3 zq1zq2 z)


1

2V

qi0

(2)4 (4) (q4

+ q3 q2 q1 ) .

(10.105)

Thus, up to a factor 1/ 2V qi0 for each particle, we remain precisely with the vertex
contribution in momentum space including the total four-momentum conservation
factor.
In momentum space the Feynman diagram (10.104) reads

731

10.7 Green Functions and Scattering Amplitudes

Thus, by replacing the four external lines by external physical particles of momenta qi corresponds to dropping the factors
i
(10.106)
2
qi m2
and replacing them by

1
q

2V qi0

(10.107)

This corresponds to the amputation of the Feynman diagram for the four-point
function introduced earlier when defining the vertex functions.
How about the disconnected diagrams in (10.64)? Consider first the three diagrams containing two disconnected lines
.
In each of the diagrams we have to do the operations of the type (10.89). Take
the first diagram and consider G0 (x3 , x1 ). It contributes a factor
s

2q30
V

2q10
lim
V x03
x0
1

0 0

0 0

d3 x3 d3 x1 ei(q3 x3 q3 x3 q1 x1 +q1 x1 ) G (x3 , x1 ) .

(10.108)

Performing the first limit x1 we get


s

2q30
lim
V x03

1
0
0 0
d3 x3 eiq3 x3 eiq1 x3 q
= 0lim ei(q3 q1 )x3 q3 q1 = q3 q1 , (10.109)
x3
2q10 V

which is just the scalar product

hq3 |q1 i = h0|aq3 aq1 |0i = q3 ,q1

(10.110)

of the particle 1 running from the initial to the final state 2 without interaction.
The same factor appears for the lower line of the diagram
hq4 |q2 i.

(10.111)

Thus the diagram turns into the S matrix element


q3 q1 q2 q4 .

(10.112)

The second diagram results in the same product of -functions except with q3 and
q4 interchanged. The third diagram is different. Here the Fourier limit to be done
is
s

2q10
V

2q20
V

lim

t1 <t2

0 0

0 0

d3 x1 d4 x32 ei(q2 x2 q2 x2 q1 x1 +q1 x1 ) G(x2 , x1 ).


(10.113)

732

10 Quantum Field Theoretic Perturbation Theory

The first limit on x1 leads to


s

2q20
lim
V t2

1
,
d3 x2 eiq2 x2 iq1 x2 q
2q10 V

and the integration gives

(10.114)

lim ei(q2 +q1 )x2 q2 ,q1 .

(10.115)

x2

The limit x0 of the exponential oscillates infinitely rapid and the result
vanishes for the same reasons as before.
Let us now look at the diagrams

of (10.64). From the lower line this obviously gives a factor


hq4 |q2 i = q4 q2

(10.116)

just as in (10.109). The upper line corresponds to the expression




12

ig
4!

Z

d4 z G(x3 , z)G(z, z)G(z, x1 ) .

(10.117)

Adding this to the line without the loop correction leads to the one-loop corrected
Green function


GR (x3 , x1 ) = 12

ig
4!

Z

d4 z G(x3 , z)G(z, z)G(z, x1 ) .

(10.118)

The Fourier representation of the free Green function i/(q 2 m2 ) is replaced by


i
i
g
2
i
2
2
2
q m
q m
2

d4 q
i
i
4
2
2
2
(2) q m q m2
"

#2

(10.119)

This may be viewed as the first-order perturbation expansion of the the Fourier
transform of the renormalized propagator
GR (x2 , x1 ) =

i
d4 q iq(x2 x1 )
e
4
2
2
(2)
q m m2

with a mass shift


m2 =

g
2

d4 q
i
.
(2)4 q 2 m2

(10.120)

(10.121)

733

10.7 Green Functions and Scattering Amplitudes

The scattering amplitude is extracted from the aplitude with the renomalized Green
function as before, the only difference being that the factor (10.107) contains now
qi0 with the renormalized masses.
A simiar mass shift occurs in the diagrams

(10.122)
of (10.64). The remaining diagrams of (10.64) contain the second-order mass shifts.
To study such mass shifts in general we make use of the Gell-Mann Low formula
for the energy shift. For the vacuum, the energy shifts by E0 can be taken from
the matrix element
eiE0 (t2 t1 ) = h0|UI (t2 , t1 ) |0i

(10.123)

by going to the limit t2 , t1 . Consider now the single-particle state


aq |0i. If the interaction is applied, the state
UI (t2 , t1 ) aq |0i

(10.124)

will for t1 and t2 again be a solution of the free Hamiltonian. Because


of energy and momentum conservation, it must be equal to aq |0i up to a phase
which, due to (10.123), contains the information on the energy shift of this state. It
consists of E0 for the vacuum plus Eq for the particle. Collecting both together
in E we may write
eE(t1 t0 )

= h0|aq UI (t2 t1 ) aq |0i/h0|UI (t2 , t1 ) |0i.

t1
t2

(10.125)

Expanding UI in powers of g, the lowest diagrams on the right-hand side are precisely
the diagrams that appeared before:
G(2) (x1 , x2 ) = xq1

q
x2

+ 12 q
q q + .
x1

x2

Indeed, we find from (10.125) with t1 t0 = T :

ig
1 Z 4
d z G(z, z)
4! 2q 0 V
Z
i T
d4 q
i
= 1 g 0
.
2 2q
(2)4 q 2 m2

1 iE T = 1 + 12

(10.126)

Thus, if we use in formulas (10.95) and (10.96) the shifted energy q 0 q 0 +E = qR0 ,
and write them as
aq =
aq =

lim
0

2qR0
V

d3 x ei(qR x

lim

2qR0
V

d3 x ei(qR x

x0

0 qx)

(x),

0 qx)

(x),

(10.127)
(10.128)

734

10 Quantum Field Theoretic Perturbation Theory

then every particle line automatically receives a factor eiEq T . This removes precisely
the phase factor (10.123).
To higher orders in g, it is somewhat hard to proceed in this fashion. The
problem will be solved in the next section with more elegance.
Note that formula(10.125) allows us to evaluate the shift in the particle mass in
another way. If q 0 = q2 + m2 is the energy before turning on the interaction, and
q 0 + E =

q2 + m2 + m2 = q 0 +

m2
2q 0

(10.129)

is the energy afterwards, we find once more the mass shift (10.121).

10.8

Wick Rules for Scattering Amplitudes

The possibility of obtaining external particle states from fields by temporal limiting
procedures of the type (10.96) allows us to incorporate the cration and annihilation
ooperators of these particles into the general framework of Wicks contraction rules.
Confronted with the numerator of the perturbative scattering amplitude in (10.92),
i

SN (q4 q3 |q1 q2 ) h0|aq4 aq3 T e

dtVI (t)
aq2 aq1 |0i,

(10.130)

we may imagine the incoming particle operators to carry negative infinite time arguments, the outgoing ones positive infinite ones. Then they can be brought inside
the parentheses of the time-ordering operator. This, in turn, can be evaluated as
usual via Wick contractions.
The results of the last section show that the Wick contractions leading to an
external particles are
1
(x)ap = [(x),ap ] = q
eiqx ,
2V p
1
eiqx ,
ap (x) = [ap ,(x)] = q
2V p

(10.131)

(10.132)

where p = p + M 2 .
For Dirac particles, there are the corresponding contraction rules
1
u(p, s3 )eipx ,
(x)ap,s3 = {(x),ap } = q
V Ep
1
eipx u(p, s3 ),
ap,s3 (x) = {ap,s3 ,(x)} = q
V Ep

(10.133)

(10.134)

10.9 Thermal Perturbation Theory

735

and for antiparticles:


1

(x)b
eipx v(p, s3 ),
p,s3 = {(x),bp } = q
V Ep

where Ep

10.9

1
v(p, s3 )eipx .
bp,s3 (x) = {ap,s3 ,(x)} = q
V Ep

(10.135)

(10.136)

p + M 2.

Thermal Perturbation Theory

Since Wicks theorem was valid for thermal Green functions, we expect also the
perturbation expansion to have a simple generalization to the thermal case. Let us
define a thermal Heisenberg picture for operators by
OH ( ) = eH /h OS eH /h

(10.137)

and an interaction picture which moves according to the free equations of motion,
OI ( ) = eH0G /h OS eH0G /h .

(10.138)

Thus, a Heisenberg operator can be transformed to the free operator via


OH ( ) = UI (0, )OI ( )UI (, 0)

(10.139)

UI (2 , 1 ) = eH0G 2 /h eHG (2 1 ) eH0G 1 /h

(10.140)

where

is the time displacement operator along the Euclidean time axis . Therefore it has
the same factorization property as the quantum mechanical operator UI (t2 , t1 ) in
Eq. (9.18):
UI (3 , 2 )UI (2 , 1 ) = UI (3 , 1 ).

(10.141)

UI (1 , 1 ) = 1.

(10.142)

Certainly

However, in contrast to UI (t2 , t1 ), this operator does not satisfy the unitarity relation
(9.27), but instead:
UI (2 , 1 ) = UI (1 , 2 ).

(10.143)

736

10 Quantum Field Theoretic Perturbation Theory

It obeys the following equation of motion


h
UI (, ) = eH0 /h (H0G HG ) eHG ( )/h eH0G /h
= VI ( ) UI (, )

(10.144)

where VI ( ) is the time-dependent thermal interaction picture of the time independent Schrodinger perturbation V = VS
VI ( ) = eH0G /h V eH0G /h .

(10.145)

Therefore we can solve for UI (, ) using the standard exponential NeumannLiouville expansion (9.20) now time-ordered with respect to the Euclidean time
:

UI (, ) =


X

n=0

1
h

= T e

n

1
n!

d VI ( )

d1

d2 . . .

dn TT [VI (1 ) VI (n )]

(10.146)

By construction, UI (, 0) satisfies the relation


eHG /h = eH0 /h UI (, 0).

(10.147)

Thus, using the thermal value for the imaginary time /h = 1/kB T , we obtain
directly a perturbation expansion of the partition function


Z = Tr eHG = Tr eH0G UI (, 0)
=


X

n=0

1
h

n

1
n!

d1

(10.148)

dn Tr eH0G T (VI (1 ) VI (n )) .

Consider now a thermal two-body Green function in the presence of interaction,


defined for > as follows:
h

G(x, ; x , ) Tr eHG (x, ) (x , ) /Tr eHG


h

= Tr eH0G UI (, 0) UI (0, )I (x, )UI (, 0)


i

(10.149)

UI (0, )I (x , )UI ( , 0) /Tr eH0G UI (, 0) .


For < , on the other hand, we have
h

G(x, ; x , ) Tr eHG (x , )(x, ) /Tr eHG


h

= Tr eH0G UI (0, ) I (x , )UI ( , 0)


h

(10.150)
i

UI (0, )I (x, )UI (, 0)] /Tr eH0G UI (, 0) .

Both
equations
may
be
combined
insGreen+function,thermal,fermions
h

in

the
i

single
h

formula
i

G(x, ; x , ) = Tr eH0G T UI (, 0) (x, ) (x , ) /Tr eH0G UI (, 0) .


(10.151)

737

10.9 Thermal Perturbation Theory

As compared with the field theoretic formulas (10.9) and (10.12), the vacuum expectation values are replaced by the Botzmann-weighted thermal traces, and the vacuum expectation value of the S matrix operator UI (, ) in the denominators is
replaced by the Boltzmann-weighted trace of the interaction operator UI (h/kB T , 0)
along the Euclidean time axis .
Note that the denominator ensures the existence of the zero temperature limit
in just the same way as the phase factor did in the switching on limit 0. In
fact, the grand canonical partition function in the denominator may be written as
ZG = e =

e(En Nn ) ,

(10.152)

such that in the limit T 0 it becomes a pure exponential for the ground state
energy
ZG e(E0 N0 ) e .

(10.153)

T 0

This is the analogue of the infinitive phase factor for the field theoretic denominator
in Eq. (10.9).
Let us now expand the interaction operator UI (h/kB T , 0) in powers of the interaction, just as before in real time. This leads to a series of thermally averaged products of many fields I (x, ) which move according to the free field equations. Therefore Wicks theorem can be applied and we obtain an expansion of
G(x, ; x , ) completely analogous to the field theoretic one. The only difference is
the finite time interaction. When going to Fourier transformed space, the finite Euclidean time interval corresponds to, and is fully taken into account by, the discrete
set of imaginary Matsubara frequencies.
We have seen before that the evaluation fo field theoretic Green function proceeds best via a Wick rotation of the energy interactions to imaginary axis. This is
precisely the axis along which the Matsubara frequencies lie. Thus as far as perturbation theory in Fourier transformed space is concerned, the diagrammatic rules are
exactly the same except that the Wick rotated integrals over the imaginary energy
are to be replaced by the Matsubara frequency sums [recall Eq. (2.412)]
Z

Z
dp0
dpE
kB T X 1 X
i

=
,
2
h
m
m
2

2
m =
h

m
m+

1
2

for bosons,
(10.154)
for fermions.

In the limit of small temperature, the Matsubara frequencies move closer and closer
to each other, and the sum tends to the integral.
The two descriptions coincide apart from the trivial presence of the chemical potential in the grand canonical energy. We conclude that the Wick-rotated calculation
of field theoretic Green function really amounts to thermal equilibrium physics in
the limit of zero temperature.
Let us point out that the result (10.151) implies the same periodicity in h
/kB T =

of the full Green function G(x, ; x , ) as in the free case (2.410). First, we observe
that due to the time independence of the Hamiltonian, there is now translational

738

10 Quantum Field Theoretic Perturbation Theory

invariance in Euclidean time T such that G(x, ; x , ) depends only on the difference
:


G(x, ; x , ) = Tr eHG ( h) (x, 0)eHG h (x , 0)eHG /h / Tr(eHG )

= Tr eHG ( h) (x, 0)eHG /h (x , 0) / Tr(eHG ).

(10.155)

Similarly, we conclude the dependence only on x x by translational invariance in


space, using (x, ) = eiPx/h (0, )eiPx/h with the momentum operator P. Thus
we may write
G(x, ; x , ) = G(x x , ),

(10.156)

just as for the field theoretic Green functions in the vacuum. Now it is easy to see
the periodicity. For the interval (h/kB T , 0) we calculate
h

G (x x , ) = Tr eHG (x , )(x, ) / Tr(eHG )


h

= Tr (x, )eHG (x , ) / Tr(eHG )


h

= Tr eHG (x, + ) (x , ) / Tr(eHG )


= G (x x , + )

(10.157)

thus showing that interacting Boson and Fermi thermal Green functions are periodic
and antiperiodic under the replacement + , just as in the free case in
Eq. (2.410).

Notes and References


The particular citations in this chapter refer to:
[1] H. Kleinert and V. Schulte-Frohlinde, Critical Properties of 4 -Theories, World Scientific,
Singapore 2001, pp. 1489 (kl/b8),
where kl is short for http://www.physik.fu-berlin.de/~kleinert.
[2] H. Kleinert, Gauge Fields in Condensed Matter , World Scientific, 1989, pp. 1744 (kl/b1).
[3] H. Kleinert, Proceedings of a NATO Advanced Study Institute on Formation and Interactions of Topological Defects at the University of Cambridge, England, A-C. Davis and R.
Brandenberger, eds., Kluwer, London, 1995 (kl/227).
[4] H. Kleinert, Multivalued Fields, World Scientific, Singapore 2008, pp. 1497 (kl/b11).
[5] C. Schubert, Perturbative quantum field theory in the string-inspired formalism, Phys. Rept.
355, 73, (2001) (arXiv:hep-th/0101036).
[6] See for instance Chapter 7 in the textbook
S. Gasiorowicz, Elementary Particle Physics, John Wiley, Ney York, 1955.

But how can finite grasp infinity?


John Dryden (1631-1700)

11
Extracting Finite Results from Perturbation
Series. Regularization and Renormalization
The Feynman integrals introduced in the last chapter suffer from a serious defect:
They are almost always divergent. Fortunately, there exists a certain class of field
theories, called renormalizable field theories, for which it is nevertheless possible to
extract finite predictions for physically observable quantities. All infinities can be
absorbed into a few divergent prefactors in relations between input parameters of
the theory and their observed physical values. The price to pay is the impossibility
of predicting these parameters, but all other observable properties of the theory
can be expressed in terms of their physical values. Typically, these parameters are
masses and coupling constants, and the normalization of field operators.
The action of a scalar field with a 4 -interaction in four dimensions defines such
a renormalizable field theory and presents a very good demonstration object for this
procedure. It will now be discussed in detail.

11.1

Vacuum Diagrams

The Lagrangian of the theory is


L=

1
m2 2 g 4
()2
,
2
2
4!

(11.1)

so that there are lines


=
and vertices

i
,
q 2 m2

(11.2)

g
.
4!
The perturbation series for the Green functions
q

..... ....
........
..... .....

G(n) (x1 , , xn ) = h0|T ei 4!


(n)

dx4 (x)

(11.3)

(x1 ) . . . (xn )|0i/h0|T ei 4!

GN (x1 , . . . , xn )/Z[0]
739

dx4 (x)

|0i

(11.4)

740 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


is obtained by expanding numerator and denominator on the right-hand side in powers of the coupling constant g and performing all Wick contractions in the resulting
time-ordered
products of free-field operators. The dominator is a pure phase factor

e2i , as we know from the Gell-Mann-Low formula (9.74). Its diagrammatic expansion has been given in Section 10.3.1, where it was shown to consist of a sum of
all vacuum graphs. Here we are interested in the size of the individual terms. Using
translational invariance in space and time, we may rewrite (10.34) as
g

Z[0] = h0|T ei 4!
= 1 iV T

dx4 (x)

|0i

g
1 g
h0|4(0)|0i
4!
2! 4!


2 Z

dxh0| (x) (0)|0i + . . . . (11.5)

This expansion runs into an immediate difficulty: the first-order term does not exist
in an infinite spacetime volume.1 This difficulty can be avoided by continuing the
discussion for a finite system in a finite time interval.
This, however, is not yet enough to produce finite results on the right-hand
side of (11.5). Consider the first-order Feynman integral (10.36) in more detail. In
momentum space it reads
g
3i V T
4!

d4 g
i
4
2
(2) q m2 + i

!2

(11.6)

where 3 is the multiplicity factor of the diagram. Performing the Wick rotation
of Fig. 7.2, the mass shell singularity in the energy integral lies no longer on th
eintegration contour, so that the i-term may be omitted, and (11.6) is expressed
in terms of a euclidean integral:
g
i V T
4!

d4 qE
1
(2)4 qE2 + m2

!2

(11.7)

The integral has a quadratic divergence at large momenta. In order to obtain a


finite result we have to modify the system in such a way that in addition to the
finite spacetime volume V T , its Wick rotated energy-momentum space is finite as
well. Thus we restrict all euclidean momenta qE to lie within a sphere of rasius :
qE2 2 .

(11.8)

The integration radius is called an ultraviolet cutoff, or shorter UV cutoffsince it


eliminates the very short wavelengths of the field. Such an UV-cutoff would have
arisen naturally if the physical phenomena would take place on a 4-dimensional
lattice with spacing of the order l = / between the lattice points rather than a
1

Recall that a similar overall time factor T was found before in the general discussion in Chapter
9. There it appeared in the limit that the switching-on parameter tended to zero [see (9.107)].
In that general derivation we did not find a spatial volume factor, since in that chapter dealt with
quantum mechanics, and not with a local quantum field theory.

741

11.1 Vacuum Diagrams

continuous spacetime. On such a lattice, each term in the perturbation expansion


would have a finite value. If the lattice spacing can be chosen much finer than
the regions explored by presently available experiments, this would not present any
practical handicap. Any perturbation expansion containing an UV-cutoff is of course
unable to describe the physics going on between the lattice points.
Hopefully, this is a purely technical problem. If we could resum infinitely many
diagrams, the series might define a function which can be used also beyond the cutoff
(11.8). The simplest model function to illustrates such a hope is given by, say,
q2
2

!g

(11.9)

where is some characteristic mass parameter of the problem. In a perturbation


expansion, this function would appear as a series

g n log(q 2 /2 )

n=1

in

(11.10)

This expansion makes sense only if the momenta of the system are contained within
a sphere (11.8) whose size is smaller than
2 = 2 e1/g .

(11.11)

The unexpanded expression (11.9), on the other hand, is defined in the entire complex q 2 -plane, except for a cut from q 2 = 0 to . This characteristic behaviour will
indeed emerge in some partial resummation procedures of many perturbation series.
In general, however, the divergences may be of a much more violent nature. The
coefficients of the higher-order terms grow in general with n!, and the series have a
zero radius of convergence. The physical significance of this violent divergence will
be discussed later.
To second order in g, the Wick expansion of the second expectation value on the
right-hand side of (11.5) yields the Feynman integrals (10.42). In momentum space
they read
1
1
g
q

q
9
= 9
2!
2! 4!


2 "Z

1
dq 2
q m2 + i
-4

1
1
g
q q
72
= 72
2!
2!
4!

2 "Z

g
1
1 ...
........
24 q...........................................q = 24
2! 2!
4!

d-4 q1 d-4 q2 d-4 q3

2 Z

i
d q1 2
q1 m2 + i
-4

q12

#4

#2 Z

-4

d q2

!2

i
,(11.12)
2
q2 m2 + i

i
i
i
.
2
2
2
2
m + i q2 m + i q3 m2 + i

After performing again Wick rotations and introducing the euclidean momentum
cutoff (11.8), we see that these expressions diverge as 8 , 4 log , 4 respectively.
Analogous statements hold for higher orders.

742 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


Thus the infinite phase caused by the vacuum diagrams diverges quite catastrophically for two reasons, both deriving from fundamental properties of spacetime:
one is homogeneity (i. e., translational invariance) which causes the overall factors
V T . The other is continuity of spacetime, which causes the Fourier representations
fields to contain euclidean momentum integrals over an infinite momentum volume.
It is a pleasant feature of the Gell-Mann-Low formula, that the divergence of the
phase due to the vacuum diagrams is irrelevant. In Section 9.1.3 it was demonstrated
that the phase in the denominator cancels against an identical phase arising in the
numerator when expanding (11.4) in powers of g [recall (10.12)]. In terms of graphs
this amounts to a cancellation of all Feynman diagrams in the numerator which
contain a disconnected part occuring in the vacuum diagrams of the denominator.
This cancelation removes all vacuum diagrams from the numerator, together with
their unpleasant V T - and -divergences. In quantum mechanics, this was sufficient
to obtain finite results in the limit of a vanishing switching-on parameter .
We are thus confronted with only those Feynman graphs which are free of vacuum
contributions. Consider the corresponding two- and four-point functions.

11.2

The Two- and Four-Point Functions

In Eq. (10.62) we have derived the first terms of the perturbation series for the
two-point function:
G(2) (x1 , x2 ) = xq1

q
x2

+ 12

q
qq ,
+ 144 + 144 xq1 q q xq 2 + 96 xq1 q
x2
q
q q

x1

q q
x1

x2

q
x2

(11.13)
which amounts to the geometric series (10.74) in momentum space:
G(2) (q) =

(11.14)

(q)
= G0 (q)+G0(q) [i(q)] G0 (q)+G0(q) [i(q)] G0 (q) [i(q)] G0 (q) + . . . ,
q2

m2

with (q) being the sum of proper self-energy diagrams (= one-particle irreducible
diagrams with two amputated legs)
i(q) = 12

q .

+ 96 q
+ 144

(11.15)

The Feynman integrals associated with these diagrams can be written in momentum
space as follows:

ig
4!

d-4 k

i
,
k 2 m2 + i

743

11.2 The Two- and Four-Point Functions

ig

4!

2 Z

i
2
k1 m2 + i

-4

d k1

q
 ig 2 Z
q
q =

4!

d-4 k2 d-4 k1

k12

!2 Z

d-4 k2

k22

i
,
m2 + i

(11.16)

i
i
i
.
2
2
2
m + i k2 m + i (q k1 k2 )2 m2 + i

(11.17)

The four-point Green function is reduced to the vertex function (4) by factorizing
out the four Green functions:
G(4) (p1 , . . . , p4 ) = G(2) (p1 )G(2) (p2 )G(2) (p3 )G(2) (p4 )i(4) (p1 , . . . , p4 ).

(11.18)

The vertex function i(4) collects all one-particle irreducible graphs with four amputated legs. [see (10.6.1) and (10.78)]. Up to third order:
i

(4)

= 4!

.... ....
......
...
.... ....

1 .......
..
...q....
.q
+ (4!)
.. + 2perm
....
2
 

1 .....
..
q

...q...... + 2perm
+ (4!)3
....q
.
.
.
.
.
4 .




1
q
....
+
.. + 2perm
..q
...q...
...
....
.
.
2

 ........q....
1
q
q + 5perm
.
+
2

(11.19)

In momentum space, these graphs correspond to the Feynman integrals:


k+q1 +q2
q2
q4 ..
..
q3

.....
q
...q.....
....
...

q1
.

....
....
....
....
.....

k2

ig
=
4!


k1+q1+q2 k+q1+q2

q2
q4 ....
....
q3

2 Z

d-4 k

ig
q
q
q
=

4!
q1

.....
....
..
....
....

....
.......
....
...

k1

k2

3 "Z

i
i
,
2
m + i (k + q1 + q2 )2 m2 + i

d-4 k

(11.20)

i
i
k 2 m2 + i (k + q1 + q2 )2 m2 + i

#2

(11.21)
k

q4

q

 Z
q .... 2

ig 3
....
q
...q....
....
=

....

4!
k2 q1
Z

.....
....
.....
.....
....

q3

q4 .. .. q2
.....q...
k1 +q2 +q4
k1
q q1 =
q3 q
k
2

"

i
i
dk 2
k m2 + i (k + q1 + q2 )2 m2 + i
i
,
d-4 k 2
k m2 + i
-4

3 Z

#2

(11.22)

i
m2 + i
k1 -k2 -q1
i
i
i

.
2
2
2
2
(k1 + q2 + q4 ) m + i k2 m + i (k1 k2 q1 )2 m2 + i
(11.23)


ig
4!

d-4 k1 d-4 k2

k12

744 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


We may perform the Wick rotation in the loop integrals so that now d-4 q goes over
into i times the euclidean volume elements d4 qE /(2)4, and q 2 becomes qE2 =
(q2 + q42 ). Then we find the following analytic expressions for self-energy and
four-point vertex function:
g
g2
g2
(q) = D1 D2 D1 D3 (q),
2
4
6

(11.24)

i
g3 h 2
g2
I (q1 + q2 ) + 2 perm
[I (q1 + q2 ) + 2 perm ] +
2
4
3
g3
g
+ [I3 (q1 + q2 ) D1 + 2 perm ] + [I4 (qi ) + 5 perm ] .(11.25)
2
2

(4) (qi ) = g

Here, D1,2,3 and I are the following Feynman integrals


1
,
+ m2
Z
1
=
d-4 k 2
,
(k + m2 )2

D1 =
D2

D3 (q) =

d-4 k

(11.26)

k2

d-4 k1 d-4 k2

(11.27)
1
h

(k12 + m2 ) (k22 + m2 ) (q k1 k2 )2 + m2
Z
1
1
,
I(q) =
d-4 k 2
2
(k + m ) [(k + q)2 + m2 ]
Z
1
1
I3 (q) =
d-4 k 2
,
2
2
(k + m ) [(k + q)2 + m2 ]
Z
1
I4 (qi ) =
d-4 k1 d-4 k2 2
2
(k1 + m ) [(k1 + q2 + q4 )2 + m2 ]
1
i.
h

2
(k2 + m2 ) (+k1 k2 q1 )2 + m2

11.3

i,

(11.28)
(11.29)
(11.30)

(11.31)

Divergences, Cutoff, and Counterterms

The trouble with these expressions is that, except for I3 , all integrals diverge. In
order to extract physical information from them, we introduce an ultraviolet cutoff
(11.8) and integrate all momenta only over a spherical range q 2 2 . Then for
large , the different expressions diverge like D1,3 2 , D2 , I log 2 /m2 , I4
log2 2 /m2 .
In the presence of such a cutoff we must watch out that the size of the coupling
constant remains small enough to justify an expansions in powers of g. In the present
case we have to require at first that the quantities
g 2 /m2 ,

g log 2 /m2 ,

g 2 log2 2 /m2

(11.32)

745

11.3 Divergences, Cutoff, and Counterterms

are all very much smaller than unity. As far as the first term is concerned, this
severely restricts the size of g for narrow lattices with 2 m2 . The other contributions involving logarithms are not so restrictive and permit somewhat more
sizable couplings, since the logarithm grows very slowly with /m.
Before proceeding it is useful to separate each momentum integral into a term
which diverges for , and a remainder which remains finite in this limit,
and thus independent of . A standard procedure do this consists in expanding the
diverging integrals in a power series of the external momenta. Then we see that after
a few diverging expansion terms all coefficients are finite. We collect the remaining
finite powers and call this part the regular part of the diagram. For the diagrams
under consideration the expansions read
D3 (q) = D3 (0) + D3 (0)q 2 + D3R (q),

(11.33)

I(q) = I(0) + I R (q),

(11.34)

where a prime denotes the derivative with respect to q 2 . Only the initial expansion
terms D3 (0), D3 (0), I(0) are divergent for . The remainders D R (q), I R (q)
are cutoff-independent functions of the momenta.
The Feynman amplitude I4 (qi ) involves two integrations. Since the expansion
(11.25) contains all 6 permutations of the momenta q1 , q2 , q3 , the fourth being fixed
by energy-momentum conservation, q 4 = q1 q2 q3 , we may freely permute the
momenta in the integral (11.23) and calculate
I4 (qi ) =

d-4 k

1
(k 2

m2 )

(q1 + q2 k)2 + m2

iI

(q3 + k)

(11.35)

where I(q) is the integral (11.29), which is expanded as in (11.34). Thus we can
write
I4 (qi ) = I(q1 + q2 )I(0) +

d-4 k

1
h

(k 2 + m2 ) (q1 + q2 k)2 + m2

i I R (q3

+ k) (. 11.36)

Performing the subtraction in I (q1 + q2 ) we have


I4 (qi ) = I 2 (0) + I R (q1 + q2 )I(0)
+

d-4 k

(11.37)
1

(k 2

m2 )

(q1 + q2 k) +

m2

i I R (q3

+ k) .

The last term can now be subtracted at qi = 0, leaving a finite remainder. Since
I R (0) = 0, we may also write
I4 (qi ) = I4 (0) + I R (q1 + q2 ) I(0) + I4R (qi ).

(11.38)

746 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


There are two types of cutoff dependences: One which is momentum-independent
and one which is accompanied by the regularized amplitude I R (q1 + q2 ). This structure will be important in achieving a finite vertex function. With this separation we
can rewrite and (4)
(q) = (0) + (0)q 2 + R (q),
(11.39)
2 n
h
io
g
3I(0) + I R (q1 + q2 ) + 2 perm
(4) (qi ) = g
2
h
i
h
io
g3 n 2
+
3I (0)+2 I R (q1 +q2 )+2 perm I(0) + (I R (q1 + q2 ))2 + 2 perm
4
g3
[I3 (q1 + q2 ) + 2 perm ] D1
+
2
h
io
g3 n
+
6I4 (0) + I4R (qi ) + 5 perm ,
(11.40)
2
with the -dependent quantities
g2
g2
g
D1 D2 D1 D3 (0),
(0) =
2
4
6
2
g
(0) = D3 (0).
6

(11.41)
(11.42)

Let us interprete this separation physically. For this we consider the full singleparticle propagator (11.14). It has the form
G(2) (q) =

q2

m2

1
.
+ (0) + (0)q 2 + R (q)

(11.43)

The pole of this propagator corresponds to the mass of the particle after the perturbative corrections. If the cutoff is sufficiently large we may, for a moment, neglect
R (q). Then G(2) can be rewritten as
G(2) (q) =

1
q 2 [1 + (0)] + [m2 + (0)]
[1 + (0)]1
= 2
.
q + m2 [1 + (0)/m2 ]/[1 + (0)]

(11.44)

This looks like the propagator of a free field h0|T R(x)R (y)|0i, with
R
and a mass

m2R = m2

1 + (0) ,

(11.45)

1 + (0)/m2
.
1 + (0)

(11.46)

Experimentally, only the renormalized mass mR is observable; there is no way of


detecting the original mass parameter m in the Lagrangian.

747

11.3 Divergences, Cutoff, and Counterterms

Thus, in all amplitudes calculated from the Lagrangian, the mass mR has to be
used in the final expression rather than m. Technically, this turns out to be slightly
inconvenient. There is a more economic method due to Bogoljubov which is based
on adding to the Lagrangian a set of counterterms which have the same form as the
terms of the original Lagrangian. First the free parts:
2
1
2m
2 .
Lc = c ()2 cm
2
2

(11.47)

The coefficient functions are expanded in a power series in g starting with g:

c =

cn g n ,

cm =

n
cm
n g .

(11.48)

n=1

n=1

They are supposed to be small enough to be treated as a perturbation. If we now


calculate the self-energy (q) up to order g 2 we find the additional OPI Feynman
graphs
=

m 2
m2 ,
cm
1 g + c2 g

(11.49)

c1 g + c2 g 2 p2 ,

and loops of the form

1
2

2
2
cm
1 g m D2 ,

1
2

c1 g 2

d-4 k

(11.50)


1 2
k2
2
=
c
D

m
D
g
1
2 .
2 1
(k 2 + m2 )2

The crosses and little circles are called mass- and wave function insertions, respectively. respectively. When adding these terms to the cutoff-dependent parts of the
self-energy, they become
sg (q) =

g
g2
g2
g2
D1 D2 D1 D3 (0) D3 (0) q 2
2
4
6
6



m2
m2 2
2
+ c1 g + c2 g m + c1 g + c2 g 2 q 2

1 2 2 2
1 2
2
+ cm
D

m
D
.
g
m
D
+
c
g
1
2
2
2 1
2 1

(11.51)

The point is now that we can always choose the counterterms in such a way that the
physical mass arrives at the value m we started out with. This can be done order
by order in perturbation theory:

748 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


To order g we determine
1
2
c1m m2 = D1 ,
2

c1 = 0.

(11.52)

Inserting this into (11.51), we see that the second and 9th term cancel each other.
To second order we determine
2

1
D3 (0),
6
1
D (0).
=
6 3

2
cm
=
2 m

(11.53)

c2

(11.54)

With these choices, no cutoff-dependent parts are left in (q) and


(q) R (q) =

i
g2 h
D3 (q) D3 (0) D3 (0)q 2 .
6

(11.55)

As far as the physical results are concerned, this procedure is completely equivalent
to the first. The only difference is that with the counterterms the original mass
parameter in the Lagrangian winds up directly in the final expressions whereas in
the first method all expressions must be brought to the renormalized form at the
end of all calculations.
The counterterm method really means that instead of the original Lagrangian
we have worked with the modified Lagrangian
2

L + Lc =

1 + c
1 + cm 2 2
()2
m
2
2

(11.56)

where and m are the renormalized field and masses, respectively. This is equivalent
to starting out with
L0 L + Lc =

1
m2
(0 )2 0 20
2
2

(11.57)

where
0

1 + c ,

m20 m2 (1 + cm )(1 + c )1 .

(11.58)

Identifying this 0 with the previous and the new with the previous R we see
2
that to lowest order in gc = (0), cm = (0). To higher orders, an additional
change from m to mR is necessary in the normalization factors (11.45), (11.46) for
a comparison with the result (11.58) [the integrals (0), (0) in (11.41), (11.42)
2
are expressed in terms of the unrenormalized mass, while c , cm in (11.54),(11.53)
contain the renormalized mass].
The counterterm method has the obvious advantage that at every stage only
cutoff-independent quantities appear in the calculated amplitudes.

749

11.3 Divergences, Cutoff, and Counterterms

Note that the physical mass is not quite equal to the Lagrangian parameter m2 .
The propagator is
G(2) (q) =

1
,
q 2 + m2 + R (q)

(11.59)

and the renormalized quantity R (q) will cause a shift of the pole position to order
g in the coupling constant. This finite shift may, however, be easily calculated.
At this point we realize that the finite shift can also be avoided by expanding
the loop integrals not around zero but around q 2 = m2 . Then
R (q).
D3 (q) = D3 (q)|q2 =m2 + (q 2 + m2 )D3 (m2 ) + D
3

(11.60)

Now the counterterms would be (11.52), together with


2

i
1h
D3 (m2 ) m2 D3 (m2 ) ,
6
1
= D3 |q2 =m2 ,
6

2
cm
=
2 m

c2

(11.61)
(11.62)

leading to a propagator
G(2) (q) =

q2

1
,
R (q)
+

m2

(11.63)

where the regular part starts out with the second power in q 2 + m2 and therefore
leaves the pole at its initial place q 2 = m2 .
Consider now the four-point vertex (4) . To make it finite we have to include
the additional graphs caused by the counterterms (11.47). The lowest contribution
enters to order g 3 :

2
= g 3 cm
1 m [I3 (q1 + q2 ) + 2 perm ] ,

(11.64)
= g 3 c1 {[I (q1 + q2 ) m2 I3 (q1 + q2 )] + 2 perm } .
Using (11.52), this gives

g3
D1 [I3 (q1 + q2 ) + 2 perm ] ,
2

(11.65)

thus canceling the second-last line in (11.40).


To regularize the remaining -dependent quantities, we introduce a counterterm
of the interaction type
g
L c = cg 4
4!

(11.66)

750 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


where cg has an expansion:
cg =

cgn g n .

(11.67)

n=1

This adds to the four-point vertex (4) the diagrams


,

(11.68)

where the small circles denote the coupling constant counterterms. These insertions
amount to replacing g g (1 + cg1 g + cg2 g 2 ) in the first and second term of the
expansion (11.40) (the others receive corrections O(g 4 ).
It remains to combine mass and coupling constant insertions. To order g 3, there
are none. The corrected expression for (4) is therefore


(4) = g 1 + cg1 g + cg2 g 2


n
h
io
1 2
g + 2 cg1 g 3 3I(0) + I R (q1 + q2 ) + 2 perm

2
h
i
h
io
g3 n 2
+
3I (0)+2 I R (q1 +q2 )+2 perm I(0) + (I R (q1 + q2 ))2 + 2 perm
4
h
io
g3 n
(11.69)
6I4 (0) + I4R (qi ) + 5 perm .
+
2
We now determine cg order by order to cancel the remaining iinfinities:
3
I(0),
2
15 2
=
I (0) 3I4 (0).
4

cg1 =

(11.70)

cg2

(11.71)

The result is the renormalized (4) :


(4)

i
g2 h R
I (q1 + q2 ) + 2 perm
2
i
i
g3 h R 2
g3 h R
+
I (q1 + q2 ) + 2 perm +
I4 (qi ) + 5 perm .
4
2

R (qi ) = g

(11.72)

Finally, we have to consider the effects of the coupling constant counterterms upon
the self energy. There are the additional graphs

(11.73)
+...

751

11.4 Bare Theory and Multiplicative Renormalization

of which only the first contributes to the second-order expression (11.51), with an
amplitude
3
g2 g
c1 D1 = g 2I(0)D1 .
2
4
Adding this to (11.51), the quadratic mass counterterm is modified to
1
3
2
cm
2 = D3 (0) I(0)D1 .
6
4

(11.74)

In this way we have arrived at cutoff-independent expressions for and (4) . Note,
in particular, that the unpleasant terms in the vertex function (11.69) which depend
on both, cutoff and momenta, have dropped out.
The total counterterms to second order in g are
g2
3
1
2
cm m2 = gD1 + D3 (0) g 2 I(0)D1 ,
2
6
4
1

2
c =
D (0)g ,
6 3


15 2
g2
3
I (0) 3I4 (0) .
cg = g I(0) + g 2
2
4
3

(11.75)
(11.76)
(11.77)

This procedure can be extended to any order in perturbation theory. Moreover, if


more general vertex functions are calculated with an arbitrary number of legs the
counterterms regularizing the two and four point vertices are sufficient to guarantee
also their cutoff independence. The general argument will be formulated in the next
chapter.

11.4

Bare Theory and Multiplicative Renormalization

There is one important aspect of the counterterms which has to be stressed: All
counterterms have the same form as the terms in the original Lagrangian. Thus one
may rewrite L + Lc as
2

1 + c
1 + cm 2
g
L + Lc =
()2
m ()2 (1 + cg ) 4 .
2
2
4!

(11.78)

Therefore we can return to the original form by a multiplicative renormalization.


We define the so-called bare fields, mass, and coupling constant
1/2

B Z

= (1 + c )1/2 ,

m2B = m2 Zm Z1 = m2 (1 + c

(11.79)
m2

)/(1 + c ),

gB = gZg Z2 = gB (1 + cg )/(1 + c )2 .

(11.80)
(11.81)

and the total Lagrangian becomes


L + Lc LB =

m2
gB 4
1
(B )2 B 2B
.
2
2
4! B

(11.82)

752 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


In this bare Lagrangian, all quantities depend on the physical parameters m, g and
the cutoff . However, due to the renormalizability of the theory, all -dependence
drops out and only functions of m and g remain.
It is instructive to see the renormalization work if one uses the bare Lagrangian
(11.82) as a basis for a perturbation expansion wihtout constructing counterterms
order by order in g. Thus we calculate from the original graphs (11.15) and (11.19)
the self-energy (11.24), (11.25), now expressed in terms of the bare parameters gB
and mB instead of g, m. We find renormalized Green functions by multiplying the
1/2
bare Green functions with a factor Z
for each field,i.e.,
(2)

= Z1 GB ,

GR

(4)
GR

(11.83)

(4)
Z2 GB .

(11.84)

The bare vertex functions


(4)

(4)

1 1 1
B G1
B GB GB GB GB

become finite via the inverse factors


(2)

= B Z ,

(4)

= B Z2 .

R
R

(2)

(11.85)

(4)

(11.86)

In the previous scheme, the renormalized two-point vertex took the form
(2)

R (q) = q 2 + m2 + R (q)

(11.87)

where m is the renormalized mass defined before. Thus we may identify m from
(2)

(2)

R (0) = B (q)|q2=0 Z = m2B + B (0) Z = m2

(11.88)
(2)

The wave function renormalization is fixed by making the slope of R equal to unity
at the origin [note that R (q) starts out with q 4 because of the double substraction
in (11.55)]. Hence

"

(2) 2
(2)

=
Z = 1 + 2 B (0) Z = 1.

(q
)
B (q)
R
2
2
2
2
q
q
q
q =0
q =0

(11.89)

Finally, the renormalized coupling is given by


(4)

(4)

R (0) = gB B (0)Z2 = g.

(11.90)

This leads to the three equations:


gB B gB2 B B gB2 B
=
+ D1
D D
D (0) Z ,
2
4 2 1
6 3


3
3
3
2
gR = gB gB2 I B (0) + gB3 I B (0) + gB3 I3B (0)D1B + 3gB3 I4B (0) Z2 , (11.91)
2
4
2
"
#
gB2 B
1 = 1
D (0) Z .
(11.92)
6 3

m2R

"

m2B

753

11.4 Bare Theory and Multiplicative Renormalization

We now solve these three equations for gB , mB in terms of g, m by inserting on the


right-hand sides the power series expansions
gB =

an g n ,

n=1

m2B = m2R
Z = 1 +

bn g n ,

(11.93)

cn g n ,

n=1

and determining the coefficients an , bn , cn order by order in g. To first order in g


there are only few terms, and we can proceed less systematically in two steps. First,
we forget for a moment (11.92), set Z = 1, and call the masses and coupling constants emerging from (11.91) and (11.91) m
2 and g, respectively. Inverting (11.91)
we have
gB
m2B = m
2
D1 + O(gB2 ),
(11.94)
2
1 is the same expression as D1 , except with m2 replaced by m
where D
2 . Inserting
this into (11.91), we may replace mB by mR in all functions I, D. The only exception
is I B (0) where a first-order correction (11.94) must be taken into account, leading
to
I B (0) = I R (0) + gB I3R (0)D1R + O(gB2 ).

(11.95)

In this way we have accounted for all terms of order gB2 , and Eq. (11.91) yields
3 2
3
+ gB3 I(0)
+ 3gB3 I4 (0).
g = gB gB2 I(0)
2
4

(11.96)

Inverting this gives


15 2
3
+ g3 I(0)
3
g 3I4 (0).
gB = g + g2I(0)
2
4

(11.97)

Now we observe that with (11.94), the term D1B in (11.91) can be reexpanded as
1 + gB D
1D
2,
D1B = D
2

(11.98)

so that (11.91) takes the form


g2
gB
D1 (0) B D
3 (0).
2
6

(11.99)

3
1
g1
D1 g2 I(0)
D1 + g2 D
3 (0).
2
4
6

(11.100)

m
2 = m2B +
Together with (23.32), this yields
m2B = m
2

754 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


On the right-hand sides of this and (23.32) we recognize precisely the counterterms
(11.75) calculated with m,
g instead of m, g. This is in agreement with the multiplicative renormalization laws (11.80), since we assumed for the moment Z = 1 or
c = 0, in which case Eqs. (11.79)(11.81) take the form
B = 1 ,
m2B = m
2 Zm ,
gB = gZg .

(11.101)

Let us finally include the effect of Z and go from m,


g to m, g. Since
g2
Z = 1 D3R (0)
6
"

#1

(11.102)

is of second order in gB , we have to this order


Z = 1 +

g2
D3 (0) + . . . = 1 + cR + . . . .
6

(11.103)

We now determine m, g from


2

1
2
m
m
2 (1 + cm
1 ) = mR (1 + cR )(1 + cR ) ,
g(1 + cg1 ) = gR (1 + cgR )(1 + cR )2 .

(11.104)
(11.105)

The first of these equations shows that m2 and m21 differ by a term quadratic in g12:
m21

m2R

g2
1 1 D31 (0) + . . . ,
6

(11.106)

g2

1 R D3R (0) + . . . ,
6

(11.107)

"

which can be written as


m21 = m2R

"

without changing terms of order g 2 . Inserting this into (11.100) yields


m2B

m2R

2
gR R 3 2 R
1 2 R
R
2 gR R
D1 gR I (0)D1 + gR D3 (0) mR D3 (0).
2
4
6
6

(11.108)

Inserting (11.107) into (23.32), we can replace m1 by m without mistake up to order


g13 . But then we see from (11.105) that
g1 = gR

"

g2
1 R D3R (0) + . . . ,
3
#

(11.109)

i.e., g1 differs from gR only by a correction term of second order. Combined with
(11.96), this implies
15
g3
3
gB = gR + gR2 I R (0) + gR3 I R (0)2 3gR3 I4R (0) R D3R (0).
2
4
3

(11.110)

755

11.5 Dimensional Regularization of Integrals

Again we find precisely the same counterterms as in (11.77). This completes the
renormalization program with the bare quantities as a starting point and applying
the normalization conditions (11.91), (11.91), (11.92).
As we mentioned before, the finite values gR , mR do not correspond to what we
would experimentally identify as the physical coupling constants of the particles.
The mass would be defined by the pole of G(2) , i. e. by the zero of (2)

(2) (q)

q 2 =m2ph

and the physical coupling by

(4) (qi )

MS

=0

= gph

(11.111)

(11.112)

where the subscript MS denotes the point where all legs are on the mass shell, i. e.
correspond to physical particles
qi2 = m2
m2
qi qj =
(4ij 1)
3

(11.113)

such that (qi + qj )2 = 34 m2 .


The wave function normalization at the physical pole would be set equal to unity
by

R (q)
= 1.

q 2 q2 =m2

(11.114)

ph

Then all vertex functions would be directly expressed in terms of these physical
values. The difference between these and the previous m, g are calculable in terms
of a multiplicative renormalization whose factors are power series in g with only
finite coefficients.

11.5

Dimensional Regularization of Integrals

There is another method for making Feynman integrals finite with a specific control
over the divergences. It has the advantage of not needing the specification of a
subtraction point in momentum space. Moreover, in field theories other than 4
which contain massless particles and are gauge invariant it preserves most symmetry
relations between the Green functions which are easily be destroyed by the earlier
regularization methods (in particular the so-called Ward identities). Some classical
symmetries, however, cannot be salvaged into the quantum field theory. If this
happens, one speaks of an anomaly of this symmetry. Famous examples are the
chiral anomaly of massless quark theories, or the conformal anomaly of the twodimensional quantum field theory of strings. Instead of employing a cutoff it is

756 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


possible to analytically continue the Feynman integrals to a continuous spacetime
dimension in the neighborhood of the physical dimension D = 4.
Starting point of this method is the observation that a Feynman propagator can
be written as
Z
1
2
2
d e (k +2kq+m ) ,
=
2
2
k + 2kq + m
0

k 2 + 2kq + m2 > 0.

(11.115)

This can be generalized to


1
1
=
2
2
(k + 2kq + m )
()

d (k2 +2kq+m2 )
,
e

k 2 + 2kq + m2 > 0,
(11.116)

which is valid for all real > 0.


Consider now the typical Feynman integral
Z

d-D k

1
1
=
2
2
(k + 2kq + m )
()

2
2
d-D ke (k +2kq+m ) .

(11.117)

where D = 1, 2, 3, . . . , and
dD k
d-D k
.
(2)D

(11.118)

The D-dimensional momentum integral on the right-hand side is trivial:


Z

2
2
d-D k e (k +2kq+m ) =

1
(2)D

 D/2

e (m

2 q 2

),

(11.119)

since after a quadratic completion of the exponential it factorizes into D products


of Gaussian integrals. This equation forms the basis for the definition of Feynman
diagrams in a continuous number of spacetime dimensions. We simply define the
left-hand side of (11.117), (11.119) for all complex D via the analytic continuation
of the right-hand side of (11.119). Then we can perform in (11.117) the integral over
using the formula
Z

d D/21 e a = ( D/2)

1
aD/2

a > 0.

(11.120)

Note that although the integral on the left-hand side exists only for D < 2, the
right-hand side provides an analytic continuation to larger D > 2 if one goes around
the singularities at D = 2, 2 + 2, 2 + 4 . . . . The extension can be made explicit
by rewriting the left-hand side as follows:
Z

d D/21 e a =

d D/21 e a +

d D/21 (e a 1) +

cD/2
,
D/2
(11.121)

757

11.5 Dimensional Regularization of Integrals

with some parameter c > 0 on which the right-hand side does not depend. In
contrast to the original integral, the right-hand side is now defined for 2 < D <
2+2. In the limit c 0 we fall back onto the original integral. For 2 < D < 2+2,
however, we can take the limit c and we obtain the continuation rule
Z

d D/21 e a

d D/21 (e a 1).

(11.122)

If D approaches 2 from above, the decomposition (11.121) for c > 0 shows that
the integral is singular like 1/( D/2). The precise result is
Z

d-D k

1
1
D/2
( D/2)
=
.

2
2
D
2
2
D/2
(k + 2kq + m )
(2) (m q )
()

(11.123)

This equation is valid by analytic continuation in the entire complex D-plane except
at the singularities D = 2, 2 + 2, 2 + 4, . . . of the Gamma function.
The analytic continuation has a peculiar feature which is worth emphasizing since
it has important applications in massless theories (in particular gauge theories): For
all D which do not lie at the above singular points the limit 0 is well-defined
and gives
Z
d-D k = 0.
(11.124)

If one defines the integral at the singular points by analytic continuation from the
neighborhood it vanishes identically for all D. Let us split (11.124) into an angular
and radial integral, and rewrite it as
SD

dk k D1 = 0,

(11.125)

where SD is the surface of a sphere in D dimensions (S1 = 2, S2 = 2, S3 =


4, S4 = 2 2 , . . . ).
SD =

2 D/2
.
(D/2)

(11.126)

Since D is arbitrary, we see the vanishing of the general homogeneous integral


SD
(2)D

dk k D1 k 2 = 0,

(11.127)

which holds for all . Written differently, we have


Z

d-D k k 2 = 0.

(11.128)

This is known as the Veltman integral rule.


If we take q = 0 and the limit m 0 directly in the integral formula (11.123)
for 6= 0, the same result is obviously true only for D > 2
where the integral is
R
proportional to mD2 . Dor D > 2, the limiting integral dD k k 2 is regular at

758 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


k = 0. How that integral can be continued In 11.127 we show how that integral can
be continued to the singular regime D 2.
The integral (11.127) contains a prefactor which will occur in all Feynman integrals. It will be useful call it, by analogy with the measure of integration (11.118),
the reduced surface of a unit sphere ( S
1 = 1/, S
2 = 1/2, S
3 = 1/2 2 , S
4 = 1/8 2,
. . . ).

S
D

1
1
2 D/2
2
SD
=
=
.
D
D
D/2
(2)
(2) (D/2)
(4)
(D/2)

(11.129)

In the perturbation series to be derived, the factor S


D will accompany each power
of the coupling constant and may be absorbed into it [see (11.214)].
With the reduced surface of a sphere, the Feynman integral (11.123) may be
rewritten as
Z

d-D k

(k 2

1
(D/2) ( D/2)
1
= S
D
() 2
.(11.130)
2

+ 2kq + m )
2
(m q 2 )D/2

This integral may be used to derive several other Feynman integrals, by forming
successive derivatives with respect to the momentum four-vector k, and find the
formulas
(D/2) ( D/2)
1
k
=S
D
(q ), (11.131)
2

2
+ 2kq + m )
2()
(m q 2 )D/2
Z
k k
(D/2)
1
d-D k 2
=S
D
(11.132)
2

2
(k + 2kq + m )
2() (m q 2 )D/2


1
2
2
( D/2) q q + ( 1 D/2) (m q ) ,
2
Z
(D/2)
1
k
k
k

=S
D
(11.133)
d-D k 2
(k + 2kq + m2 )
2() (m2 q 2 )D/2


1
2
2
( D/2) k k k ( 1 D/2) ( k + k + k )(m q ) .
2
Z

d-D k

(k 2

By expanding (11.123 ) up to order O() and comparing the coefficients of we


find a further important integral
Z

1
d-D k log(k 2 + 2kq + m2 ) = S
D (D/2) (1 D/2) (m2 q 2 )D/2 , (11.134)
D

Strictly speaking, the logarithm on the left-hand side does not make sense since its
argument has the dimension of a square mass. It should therefore always be written
as log[(k 2 + 2kq + m2 )/2 ] with some auxiliary mass . If m if m 6= 0, the auxiliary
mass can, of course, be set equal to m. In dimensional regularization, however,
this proper way of writing the logartihm does not change the integral at all, since
it merely amounts to adding an integral over a constant, and this vanishes by the
Veltman rule (11.128).

759

11.5 Dimensional Regularization of Integrals

Note that when forming the trace of (11.132) in the indices , we obtain the
same result as in a direct calculation with k 2 in the numerator, the trivial trace
formula
D,

(11.135)

which is initially valid only for integer D, is comtinued to arbitrary continuous


dimensions D.
Because of this and for symmetry reasons we can replace in a rotationally invariant integral the tensors
k k

k2
,
D

k k k k

(k 2 )2
( + + ). (11.136)
D(D + 2)

In a general dimension D, all the above Feynman integrals are finite without a
cutoff , except for certain integer values of D. A typical example is the physical
spacetime dimension D = 4. To control the divergences in four spacetime dimensions
we set D = 4 with some small number . The divergences for D close to D = 4
appears now as a pole term in 1/n with some finite power n > 0. This way of
controlling the divergences of Feynman integrals is called the method of dimensional
regularization, or of analytic regularization.
Let us study the small- singularities of the Feynman integrals in Eqs. (11.26)
(11.31) for real D near the physical dimension D = 4. The first integral is (11.26):
1)

D1 =

d-D k

k2

1
1
=S
D (D/2)(1 D/2)(m2 )1+D/2 .
2
+m
2

(11.137)

The divergence of this integral manifests itself in the pole of the Gamma function
at the physical dimension D = 4. In the neighborhood of this dimension we set
D 4

(11.138)

and express the integral in terms of . The singularity at small may be isolated
by using the expansions
1
(1 + x) ,
x
2
x
= 1 + x2 + . . . .
(1 + x) (1 x) =
sin x
6
(x) =

(11.139)

Note that the second relation has only quadratic correction terms 2 . There is a
useful generalization of this,
 
(1 + an )
=
1
+
O
2
)

(1
+
a
m
m

if

X
n

an

X
m

am = 0.

(11.140)

(11.141)

760 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


This relation ensures the cancelation of all terms proportional to when expanding
(1 + a) = 1 + O (2 ). With (11.139) we have

1 +
= 1
1+
(D/2) (1 D/2) = 2
2
2

2
2
2
= + O()
(11.142)

and therefore
2

2 /2

D1 = S
D m (m )

1
+ O() .

(11.143)

This expression has a mass dimension (m2 )1/2 and a singularity proportional to
1/. It may be separated into a singular and a regular expression
D1 D1sg + D1reg ,

(11.144)

1
D1sg
SD (m2 )1/2 .

(11.145)

with

The singular is not uniquely defined. It may be modified by an arbitrary regular


term. For instance, we may introduce an arbitrary fixed mass scale and choose as
a singular part


D1sg
SD m2 2

/2

1
.

(11.146)

The two choices differ by a regular term 12 S


D log(m2 /2 ). The latter choice has
the form m2 multiplied function which is independent of the mass m. We shall see
below that this property persists to higher orders in perturbation. This property
will later be of great advantage in studying the theory in the limit m 0. The
scale parameter is unavoidable when taking this limit.
This parameter will play an important role in studying the scale properties of the
perturbation expansion via the so-called renormalization group in Chapter 21. That
group governs all critical phenomena observed in second-order phase transitions [1].
By comparing the singular term (11.146) with the previous cutoff divergence
which was
D1 =

1
2
(4)2

(11.147)

we realize that the quantity 1/ plays the same role as before 2 /m2 .
A similar subtraction procedure is applied to the other Fenyman integrals. Consider (11.27):
2)

D2 =

d-D k

1
1 (D/2) (2 D/2) 2 D/22
=
S

(m ) . (11.148)
D
(k 2 + m2 )2
2
(2)

761

11.5 Dimensional Regularization of Integrals

Now
2
(D/2) (2 D/2) = (1 D/2) + O()

2
1 + O()
=

(11.149)
(11.150)

so that
D2 = S
D (m2 )/2

1 1
+ O() D2sg + D2reg .
2


(11.151)

Here we choose as a singular part the expression


1
D2sg S
(11.152)
D .

The mass parameter ensures the correct dimension of both singular and regular
terms, and makes the singular term independent of the mass m. This is again useful
for a study of the theory in the massless limit. Another advantage of introducing
the mass parameter is that the singular parts (11.146) and (11.152) is that D2sg is
related to the singular part D1sg by derivative /m2 :
D2sg =

D1sg .
2
m

(11.153)

The pole term 1/ replaces a logarithmic divergence that woul appear in a cutoff
regularization. Since also the quadratic divergence in D1 had a 1/ pole, one may
wonder how the singularities in can reflect the difference between quadratic and
logarithmic divergence in the cutoff method: Comparison with D1 shows that while
D2 has its lowest singularity at D = 4, i.e., its highest singularity at = 0, D1 has
it at D = 2, i.e., the highest singularity lies at = 2.
Consider now the Feynman integral (11.29):
3)

I(q) =

d-D k

1
h

(k 2 + m2 ) (k q)2 + m2

Here we use Feynmans formula


1
=
AB

dx

1
=
[Ax + B(1 x)]2

dx

i.

1
[A(1 x) + Bx]2

which follows from the trivial relation




Z B
1
1
1
1
1
1
=
=

dz 2 .
AB
BA A B
BA A
z

(11.154)

(11.155)

(11.156)

For n denominators, Feynmans formula reads2


1
= (n 1)!
Ai An

( ni=1 xi 1)
dx1 . . . dxn
,
(A1 x1 + . . . + An xn )n
P

(11.157)

For a proof of this formula see Vol. I of E. Goursat and E.R. Hedrick, A Course in Mathematical
Analysis, Ginn and Co., Boston, 1904.

762 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


as can be proved by induction. By differentiating both sides i times with respect
to Ai , one finds more generally
n
1
( ni=1 i ) Y
=
Q
n
A1 1 Ann
i=1 (i ) i=1

dxi

Qn

i=1

xi i 1 (

Pn

Pn
i=1 xi
P
n

i=1 xi Ai ]

i=1

1)

(11.158)

By analytic continuation, this formula remains valid for complex values of i .


Using (11.155) and the substitution z = Ax + B(1 x), we rewrite (11.154) as
I(q) =

dx

dx

dD k n
dD k

1
i o2

(k 2 + m2 ) (1 x) + (k + q)2 + m2 x

1
.
[k 2 + 2kqx + q 2 x + m2 ]2

(11.159)

The momentum integration gives


Z 1
1
1
dx
I(q) = S
D (D/2) (2 D/2)
.
2
0
[q 2 x(1 x) + m2 ]2D/2

(11.160)

Going over to the variable , we rewrite this as


1
I(q) = S
D (2 /2) (/2)
2

dx

(2 )/2
[q 2 x(1 x) + m2 ]/2

(11.161)

We have used once more the mass parameter to make the integral dimensionless.
The integral is regular at = 1. A singularity comes only from the prefactor, which
possesses a 1/ -behavior to be extracted with the help of formula (11.139):


(2 /2) (/2) =

1+
1
2
2
2


2
1 + O().

(11.162)

Hence we obtain
1
2
I(q) = S
D
1

2
q2


!/2

Jm (q),

(11.163)

with the integral


Jm (q)

m2
dx x (1 x) + 2
q
"

#/2

(11.164)

For m = 0, the integral can immediately be done:


Jm (q) J0 =

xa (1 x)b =

(a + 1)(b + 1)
(a + b + 2)

(11.165)

763

11.5 Dimensional Regularization of Integrals

with
J0 =

2 (1 /2)
= 1 + + O(2 ).
(2 )

(11.166)

We may therefore expand




I(q) = S
D

1 1 2

+
2
q2


!/2

+ O().

(11.167)

The residue at the pole is independent of q, as a manifestation of the renormalizability of the theory. For m 6= 0, we can expand (11.164) in powers of m and find
the same pole term in . This property is a special feature of this regularization
procedure.
We now choose the singular part
I sg S
D

(11.168)

to form the subtracted Feynman integral


I reg (q) = I(q) I sg ,

(11.169)

which is smooth in the limit 0. As before, we have included the factor in


(11.168) for dimensional reasons.
For m 6= 0, we expand Jm (q) in powers of and find an additional factor
1

i
h m
L (q) L0 ,
2

(11.170)

where
m2
dx log x(1 x) + 2
L (q) =
q
0
2

q 2 + 4m2
q + 4m2 + q 2
m2
2
.
log 2
= 2 + log 2 +
q
q
q + 4m2 q 2
m

"

(11.171)

This can also be written as


Lm (q) = 2 + log

m2
+ 2 coth .
q2

(11.172)

with
sinh2 =
or

q2
4m2

2
1
q2
q + 4m2 + q 2
.
= atanh 2
= log 2
2
q + 4m2
q + 4m2 q 2

(11.173)

(11.174)

764 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


For m = 0, L0 (q) becomes a constant: L0 (q) L0 = 2.
Altogether, we obtain for any mass m:
I(q) = S
D

"

1 1 1 m
1
q2
L (q) log 2 + O().
2 2
2

(11.175)

The singularity is the same as in the m = 0 -expression (11.167), so that the subtraction (11.169) regularizes the integral for all m. Note the subtraction constant
of D2 (q) is the same as that of I(q):
I sg = D2sg ,

(11.176)

which is obvious from the definitions (11.27) and (11.29) and the mass and momentum independence of the singular parts.
The result (11.175) can now be applied to the calculation of the Feynman integral
(11.28):
4)

D3 (q) =

d-D k1 d-D k2

1
(k12

m2 )(k22

m2 )

1
.
[(q + k1 + k2 )2 + m2 ]

(11.177)

In D = 4 dimensions, this integral contains 8 momentum integrations and only 6


powers of momenta in the denominator, so that it diverges formally like 2 . Before
applying Feynmans formula (11.158), we decrease the degree of divergence by a
partial integration trick due to t Hooft and Veltman.3
Partial differentiation with respect to the momentum variables satisfies the trivial
identity:
k
= .
(11.178)
k
Contracting the indices in D dimensions as in (11.135) gives
k
= D.
k

(11.179)

Thus we may insert into the integral (11.177) the trivial identity
1
1=
2D

k1 k2
+
.
k1 k2
!

(11.180)

After an integration by parts neglecting surface terms, we may write


1
D3 (q) =
2D


d k1 d k2
+ k2
k1
k2
1
1
.
2
2
2
2
(k1 + m )(k2 + m ) [(q + k1 + k2 )2 + m2 ]

-D

-D

k1

G. tHooft and M.T. Veltman, Nucl. Phys. B 44 , 189 (1972). See also T. Curtright and G.
Ghandour, Ann. Phys. 106 , 209 (1977).

765

11.5 Dimensional Regularization of Integrals

Performing the differentiations leads to the following decomposition of D3 (q):


3m2 + q(q + k1 + k2 )
1 Z -D -D
d k1 d k2 2
D3 (q) =
D3
(k1 + m2 )(k22 + m2 ) [(q + k1 + k2 )2 + m2 ]3
i
1 h 2
=
3m A(q) + B(q) .
(11.181)
D3
where
A(q) =

d-D k1 d-D k2

B(q) =

d-D k1 d-D k2

1
(k12

(11.182)

q(q + k1 + k2 )
.
+ m2 ) [(q + k1 + k2 )2 + m2 ]2

(11.183)

m2 )2 (k22

m2 )(k22

m2 ) [(q

+ k1 + k2 )2 + m2 ]

and
(k12

The first integral is only logarithmically divergent. the second linearly. These integrals are now evaluated with the help of Feynmans formula (11.155). We begin
with A(q), which can obviously be rewritten as
A(q) =

d-D k1

(k12

1
I(k1 + q)
+ m2 )2

1
= S
D (D/2)(2 D/2)
2
Z
Z 1
1
1
. (11.184)
dx d-D k 2

2
2
2
(k + m ) [(q + k) x(1 x) + m2 ]2D/2
0
We remove a factor [x(1x)]2D/2 from the denominator and apply to the remaining
integrand the generalized Feynman formula (11.158) in the form
1 1
( + )
=

A B
()()

dy

y 1 (1 y)1
.
[Ay + B(1 y)]+

(11.185)

This leads to
1
(4 D/2)
A(q) = S
D (D/2)(2 D/2)
2
(2) (2 D/2)
Z 1
Z
Z 1
y(1 y)1D/2
dy d-D k 4D/2

dx [x(1 x)]D/22
f
(q; k, x, y)
0
0

(11.186)
(11.187)

with
m2
f (q; k, x, y) = y(k + m ) + (1 y) (q + k) +
x(1 x)
"
#
1y
2
2
2
= k + (2kq + q )(1 y) + m y +
.
x(1 x)
2

"

(11.188)

766 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


By integrating this over d-D k, we obtain
(4 D/2)
2 1
(D/2) (2 D/2)
A(q) = S
D
2
(2)(2 D/2)
Z 1
1 (D/2)(4 D)

dx [x(1 x)]D/22
2 (4 D/2)
0

1D/2

dy y (1 y)

q y(1 y) + m

"

(11.189)

1y
y+
x(1 x)

#)D4

which reads in terms of :


1
2 1 2
(11.190)
(2 /2) ()(m2 ) dx [x(1 x)]/2
A(q) = S
D
4
0
#)
"
(
Z 1
1y
q2
/21
y(1 y) + y +
.

dy y (1 y)
m2
x(1 x)
0

Defining the dimensionless expression in curly brackets as


1y
q2
,
F (q; x, y) 2 y(1 y) + y +
m
x(1 x)
#

"

(11.191)

we expand
F (q; x, y) = 1 log F (q; x, y) + O(2 ).

(11.192)

The first term gives [via (11.165)]


2 (1 /2) (2) (/2)
2 1 2
S
D
(2 /2) ()(m2 )
.
4
(2 ) (2 + /2)

(11.193)

This can be rewritten as


1
2 1
1
4 (1 /2)(1 + )(1 + /2)
2
S
D
(m2 ) (1 /2)2 2
. (11.194)
4
1 1 + /2
(1 ) (1 + /2)
The ratio of Gamma-functions has the general form (11.140) with (11.141) and can
be replaced by 1 + O(2 ). Thus we obtain the first contribution to A(q):
A(q) =

2
2 (m )
S
D

2
1 + O () .

(11.195)

We now turn to the second contribution in (11.190) due to the logarithm in the
expansion (11.192). A singular term at = 0 can only come from the end-point
singulaity in the y-integral of (11.190) at y = 1. Near that point, the logarithm goes
to zero like
log F (q; x, y) = O(1 y),
(11.196)

767

11.5 Dimensional Regularization of Integrals

so that the integral has no end-point singularity after all. Thus there is no further
singular contribution to A(q) in Eq. (11.231).
To complete the calculation of the pole terms in D3 (q) we have to study the
second integral B(q) in Eq. (11.181). After applying again the Feynman formula
twice we arrive at an expression of the type (11.190):
Z 1
2 1 2
B(q) = S
D
(2 /2) () q 2(m2 ) dx [x(1 x)]/2
4
0

(11.197)

dy y (1 y)/2 [F (q; x, y)] .

Here both the x- and the y-integrals are free of endpoint singularites and converge.
At = 0, the integrals are equal to 1/2. Thus we find the pole term of B(q):
2
B(q) = S
D

2 2
q2
q
(m2 ) 2
2 (m )
(2 /2)(1 + ) = S
D
[1 + O()] . (11.198)
4
2
4 2

Inserting this and (11.195) into (11.181), we obtain for the singular part of D3 (q):
D3sg (q)

SD
m2 2

6 1
m 1
q2
.
2 log + +
4

2 12m2
"

(11.199)

where we have introduced the mass parameter to make the singular term proportional to m2 as in D1sg and D2sg in (11.144) and (11.151). Apart from the overall
factor m2 , the singular term contains a logarithmic dependence on m, in contrast
to D1sg and D2sg , which had only the overall factor m2 . It will turn out that in the
present subtraction scheme all terms m2 log(m/) will eventually cancel in the divergent parts of self-energy and vertex functions. Observe that the singular term
this has only a linear q 2 -dependence. This is necessary for the renormalizability
of the theory, since a q 2 -term can be compensated by a counterterm of the type
(x)(x) in the Lagrangian densityno higher power in the momentum q can.
In some textbooks, the factor S
D is also expanded in powers of . Recalling 11.216,
we see that (11.199) may be written as
D3sg (q)

1 (m2 )1 6
q2
=
+
9

6
+
6
log
4
+
+ O () .
(4)4

2m2
"

(11.200)

Consider finally the Feynman integral (11.31):


5)

I4 (qi ) =

d-D k1 d-D k2

Thus can be written as


I4 (qi ) =

d-D k

k2

1
(k12

m2 )

(q1 + q12 k1 )2 + m2
1

(k22 + m2 ) (k1 k2 + q3 )2 + m2

1
1
I (k + q3 )
2
+ m (q1 + q2 k)2 + m2

i.

(11.201)

768 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


which becomes using (11.161)
1
1

I4 (qi ) = S
D 2
dx
(11.202)
2
2
2 0
Z
1
1
1
d-D k 2
i .
h
2
2
2
k + m (q1 + q2 k) + m (q + k)2 x(1 x) + m2 /2
3

 Z

The denominators are combined via (11.185), and we get

1
1

dy(1 y)/21 y
(2 + /2)
dx [x(1 x)]/2
2
2
2
0
0
(
1 (2 /2) () Z 1

dz yz(1 yz)(q1 + q2 )2 + y(1 y)q32


2 (2 + /2)
0

2 1

S
D

"

1y
y+
x(1 x)

2yz(1 y)q3 (q1 + q2 ) + m

# )

. (11.203)

The only pole term comes from the end-point singularity in the integral at y = 1.
For y = 1, the curly bracket can be expanded as
{. . .

i
2

} = (q1 + q2 )

m2
1 log z(1 z) +
(q1 + q2 )2
)
 
[...]
2
log
.
+O
z(1 z) (q1 + q2 )2 + m2
#

"

(11.204)

Using formula (11.140) with (11.141), the pre-factor of (11.203) is seen to have an
expansion
1
(1 ) + O () ,
4
so that
I4 (qi ) =

2
S
D

1
(1 + )
4

dx [x(1 x)]

/2

dy(1 y) 2 1 y

m2
dz 1 log z(1 z) +
(q1 + q2 )
0
(q1 + q2 )2

[...]
+ O().
(11.205)
log
z(1 z) (q1 + q2 )2 + m2
h

i Z

"

The first two terms in the integral are independent of x and y, and give
2 (1 /2) (/2) (2)
[1 Lm (q1 + q2 )] ,
(2 ) (2 + /2)

(11.206)

where Lm (q) is the integral (11.171) which occurred in the evaluation of I(q)of
Eq. (11.163)]. Thus, up to here the singular part of I4 (qi ) is
I4 (qi ) =

2
2
S
D
22

"

 

(q1 + q2 )2
1 Lm (q1 + q2 ) log
+
O
0 ,(11.207)
2
2
#

769

11.6 Renormalization of Amplitudes

where we have used the mass parameter to make the logarithm dimensionless.
We now look at the effect of the last logarithm in (11.204). It is accompanied by a
factor and it vanishes at y = 1 such that it fails to lead to an end-point singularity
at y = 1, = 0. Therefore it contributes to order O(0 ) and can be neglected as
far as the singular parts are concerned. Thus Eq. (11.206) remains unchanged.
Let us decompose this into singular and regular terms in accordance with the
decomposition (11.38) using the singular term I sg calculated in (11.168) of the Feynman integral (11.161):
I4 (qi ) I4sg + I4reg (qi )
= I4sg + I reg (q1 + q2 ) I sg + I4reg (qi )

(11.208)

with the momentum-independent term


I4sg

11.6

2
2
SD
22

.
1+
2


(11.209)

Renormalization of Amplitudes

Let us now apply these formulas. Consider at first the self-energy up to second
order where we have found it useful to recall grahically the origin of the different
contributions.
(q) = 12 gD1

14 g 2 D2 D1

m 2
+ cm
m2
1 g + c2 g

1 2
g D3
6

c1 g + c2 g 2 q 2
(11.210)

2
12 cm
1 g D2

21 c1 g 2 (D1 m2 D2 )

+ 12 cg1 g 2 D1

Obviously, we can regularize this expression in the same way as before by just
choosing the singular parts of D1 , D2 , D2 D1 , and D3 for the counterterms
1 sg
1
3
1
1
2
m2
cm
(D2 D1 )sg (D1sg D2 )sg (D2sg D1 )sg + D3sg (0),(11.211)
1 = D1 , c 2 =
2
4
4
4
6
1
sg
(11.212)
c1 = 0,
c2 = D3 .
6

770 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


Care has to be taken to distinguish between the singular parts of D2sg D1 , D1sg D2 ,
and (D2 D1 )sg , and we find from (11.144), (11.146), (11.151), (11.152), and (11.199)
sg
1 1
1
2
(D1sg D2 )sg = S
+ O()
D
m2 (m2 )/2
,

2
"
#
1
1
m
1
2
2 2
2 + log +
= S
D m
,

2
(

)sg
1

sg
sg
2
2 1/2
(D2 D1 )
+ O()
= S
D
(m )

#
"
1
m
1
2
2 2
,
2 + log
= S
D m





sg
1
1
sg
2
2 /2 1
2 1/2
(D2 D1 )
= S
D (m )
+ O()
+ O() (m )
2

#
"
2
m
1
1
2
,
= S
D
m2 2 2 + log +

2
"
#
1
2
m
1
sg
2
2 2 6
D3 (0) = S
D m
2 + log
,
(11.213)
4



In writing (11.212) we have used I sg = D2sg according (11.176). Due to the noninteger dimension all Feynman integrals carry a factor for each order in g. It
is convenient to absorb these factors, together with the surface factor S
D , into the
coupling constant and introduce the dimensionless reduced coupling constant
gS
D .

(11.214)

This considerably simplifies all expressions. Some people do not include the factor
S
D into the definition of , but only a factor S
4 = 2/(4)2 . The relation between the
two is obtained from the expansion
(D/2) = (2 /2) = 1 + ( 1)/2 + O(2 ),

(11.215)

S
D = (2/(4)2 )[1 + (log 4 + 1)/2 + O(2 )].

(11.216)

implying that

Collecting the different terms in (11.211) and (11.212), we obtain


c

m2

1
1 2
1

= + 2
2
2
4
2
1
.
= 2
48


(11.217)

Note the pleasant cancellation of all mass factors! This happens at all higher-orders
in the coupling constant.

771

11.6 Renormalization of Amplitudes

Consider now the coupling constant conterterms. Recalling the full vertex function
(4) (qi ) = g 12 g 2 [I (q1 + q2 ) + 2 perm ] + 41 g 3 [I 2 (q1 + q2 ) + 2 perm ]

+ 21 g 3 [I3 (q1 + q2 ) D1 + 2 perm ]

+(cg1 g + cg2 g 2)g

+ 21 g 3 [I4 (qi ) + 5 perm ]

cg1 g 3 [I (q1 + q2 ) + 2 perm ]

3
+ cm
1 g [I3 (q1 + q2 ) + 2 perm ]

+ c1 g 3 [I3 (q1 + q2 ) + 2 perm ]

(11.218)
we may identify
3 sg
I
(11.219)
2
i
1
1
1 h 2 sg
I
+ 2 perm [(I3 D1 )sg + 2 perm ] [I4sg + 5 perm ]
cg2 =
4
2
2
3 sg sg
1
sg
+ [(I I) + 2 perm ] + [(D1sg I3 ) + 2 perm ] .
(11.220)
2
2
Inserting (11.208) into (11.220) we separate I = I sg + I reg , D1 = D1sg + D1reg and use
the fact that I3 (qi ) is regular to cancel the term (I3 D1 )sg against the term (D1sg I3 )sg ,
so that
i
3h
cg2 = (I sg )2 + 2I sg I reg 3 (I4sg + I sg I reg )
4
i
9 h sg 2
+ (I ) + I sg I reg .
(11.221)
2
The terms containing I reg are seen to cancel each other, and we arrive at
cg1 =

15 sg 2
(I ) 3I4sg .
4
We now insert (11.176), (11.209), and obtain
cg2 =

(11.222)

15 1
1
1
3
3
+
= + 2
2
2
2
4
2
4


3
9
1
3
= + 2
.

2
4 2




(11.223)

772 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


Here we have to pause for a very important observation. Nowhere in the formulas
2
for cm , c , cg does the physical mass appear. There is only the arbitrary mass
scale which has been used up to make the coupling constant dimensionless. If this
situation prevails to all orders in perturbation theory we can conclude: The counterterms calculated according to the rules of this section can be used to renormalize
massive as well as massless theories. The multiplicative renormalization constants
depend only on the dimensionless coupling constant g.
The finite regularized vertex functions depend on the choice of g, m and the
parameter . For massive theories can be set equal to m and only the two parameters g, m are necessary to specify the manifold of all perturbative solutions to
the 4 field theory. For zero physical mass, the parameter can no longer be eliminated. Now the vertex functions will depend on g and . Since the parameter was
introduced in a completely arbitrary manner, we have the suspicion that it cannot
represent a true degree of freedom in the calculation of both massive and massless
theories. In fact we shall see in the next section that there is only a two parameter
manifold in g, m, space which corresponds to different sets of vertex functions.
As far as the renormalization is concerned, however, is absolutely necessary to
obtain finite expressions if the physical mass m is taken to zero.

11.7

Additive Renormalization of Vacuum Energy

The renormalization of the two- and four-point functions discussed above does not
remove all infinities of the theory. An infinity that is important in man-body theories
and cosmology is the infinity associated with the sum over all vacuum diagrams. The
emerge in the calculation of Z[0] of Eq. (11.5) which may be used to define the total
vaccum energy due to the interaction
vac

Z[0] = eT Eint .

(11.224)

vac
The energy Eint
has to be added to the total energy of all zero-point oscillations od
the field (x) which is

E0vac =

X
p

h
p
.
2

(11.225)

Recall Eqs. (7.651) and (7.424), and the discussion of the Casimir effect in Section ??.
This and the contribtion from all vacum diagrams are divergent,, and this infinity
must be subtracted from the total energy to ensure that the physical vacuum has
zero total energy. This is necessary for the final theory to be Lorentz invariant.

11.8

Generalization to O(N)-Symmetric Model

In many physical systems, the field has several degrees of freedom: In the superconductor it is a complex field which may be considered as a two-component object
with 2 , 4 replaced by
2 ||2 = 21 + 22

11.8 Generalization to O(N )-Symmetric Model

773

4 ||4 21 + 22

2

(11.226)

A magnetic system, on the other hand, is characterized by a vector field (1 , 2 , 3 )


and the Lagrangian contains the invariants

2 21 + 22 + 23
4

21 + 22 + 23

2

(11.227)

It is therefore useful to generalize all results to the case of an O(N)-vector a such


that instead of 2 , 4 we have:

N
X

2a

a=1
N
X

a=1

2a

!2

(11.228)

The propagator is the same as before except it is accompanied by a Kronecker ab .


The vertex function, on the other hand, carries a symmetric tensor

Tabcd =

1
(ab cd + ac bd + ad bc ) .
3

(11.229)

When calculating the Feynman diagrams, all integrations remain the same except
that each integral is accompanied by a product of T -tensors with a contraction of

774 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


indices for every internal line. For the previous diagrams these may be given as
follows:

D1 D1 Tabcc ,
D2 D1 D2 D1 Tabcd Tcdee ,
D3 D3 Tacde Tbcde

,
(11.230)

I ITabef Tcdef ,
I 2 I 2 Tabef Tef gh Tghcd ,
I3 D1 I3 D1 Tabef Tcdeg Tf ghh ,
I4 I4 Tabef Tecgh Tf dgh

For greater transparency we have underlined the external indices.


It is convenient to introduce the reduced matrix elements of the contracted tensors as follows
Tabcc = ab TD1 ,
Tabcd Tcdee = ab TD2 D1 = ab TD2 1 ,
Tacde Tbcde = ab TD3 ,
Tabef Tcdef |symm = Tabcd TI ,
Tabef Tef gh Tcdgh |symm = Tabcd TI 2 ,
Tabef Tcdeg Tf ghh |symm = Tabcd TI3 D1 Tabcd TI TD1 ,
Tabef Tecgh Tf dgh |symm = Tabcd TI4 .

(11.231)

The symbol symm denotes symmetrization in the indices abcd which replaces each
ab cd by Tabcd . In order to calculate the irreducible matrix elements we observe that
N +2
ab ,
3
N +2
ab ,
=
3

Tabcc =
Taef g Tbef g

(11.232)
(11.233)

11.8 Generalization to O(N )-Symmetric Model

775

so that
N +2
,
3
N +2
.
=
3

TD1 =

(11.234)

TD3

(11.235)

In order to calculate TI we first note that


Tabef Tcdef =

1
[(N + 4) ab cd + 2ac bd + 2ad bc ] .
9

(11.236)

It is useful to abbreviate the three tensors in the brackets as A1 , A2 , A3 , i.e., we


write
Tabef Tcdef

N +4
2
2
=
A1 + A2 + A3
9
9
9


(11.237)

abcd

The symmetrization of the tensors A1 , A2 , A3 gives each time Tabcd so that


Tabcf Tcdef |symm =

N +8
Tabcd
9

(11.238)

and hence
TI =

N +8
.
9

(11.239)

For the other products (11.231) we observe the multiplication rules


(p1 A1 + p2 A2 + p3 A3 )abef (q1 A1 + q2 A2 + q3 A3 )efcd
(11.240)
= {[p1 (q1 N + q2 + q3 ) + p2 q1 + p3 q1 ] A1 + (p2 q2 + p3 q3 ) A2 + (p2 q3 + p3 q2 ) A3 }abcd .
If we want to form Tabef Tef gh Tcdgh , we have to multiply A1 + A2 + A3 by
2
A + 92 A3 and symmetrize the result which gives
9 2
2 N +4 2 2
4
1 N +4
+
Tabcd
N+
+ +
3
9
9
3
9
9 9


N +4
A1
9

(11.241)

and hence
TI 2 =


1  2
N + 6N + 20 .
27

(11.242)

The last product in (11.231) is found by using the general rule


(p1 A1 + p2 A2 + p3 A3 )abcd (q1 A1 + q2 A2 + q3 A3 )cedf
= {[p1 (q1 + Nq2 + q3 ) + (p2 + p3 ) q2 ] A1
(p2 q1 + p3 q3 ) A2 + (p3 q1 + p2 q3 ) A3 }abef .

(11.243)

776 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


Applying this to 13 (A1 + A2 + A3 ) and N 9+4 A1 + 29 A2 + 92 A3 we find, after symmetrization






1 2 2
1 N +4 2
2 2
1 N +4
+
+2
+N +
+
+
TI4 =
3
9
9 9
3 9 9
3
9
9
1
=
(5N + 22) .
(11.244)
27
Diagrams D2 D1 and I3 D1 in Eq. (11.230) have only a single loop attached to a line.
Any such attachment inserts a tensor (11.232) into one of the index contractions,
producing merely a factor (N + 2)/3. Thus we obtain
N +2 2
TD2 D1 =
3



N +2
N +8
.
=
9
3


TI3 D1

(11.245)
(11.246)

Let us now consider the counterterms (11.211) and (11.212) in the presence of
T -tensors. They become
1
1
1 sg
2
sg
m2
(D2 D1 )sg TD2 D1 (D1sg D2 ) TD1 T
(cm
1 ) = D1 TD1 , (c2 ) =
2
4
4
1
3
(D2sg D1 )sg TI T + D3sg (0) TD3 (, 11.247)
4
6
1
sg
(c1 ) = 0,
(c2 ) = D3 TD3 .
(11.248)
6
Inserting into (11.248) the relations TD1 T = TD2 1 , TI T = TD2 TD1 , and
going over to the dimensionless coupling constant = g S
D of Eq. (11.214), we
obtain the generalization of Eqs. (11.217):
2

2 1
1
1
1
TD3 ,
TD1 +
T
+

D
D
2
1
2
4 2
2 2
1
TD ,
= 2
48 3 



3
1
1
2 3 1
=
TI4 .
TI +
TI 2 + 3
+
2
4 2
22 4

cm =
c
cg

(11.249)

For the counterterms (11.219) and (11.222) we find similarly


3 sg
I TI T
2

3
3  2 sg
I
TI 2 (I3 D1 )sg TI TD1 3(I4 )sg TI4
=

4
2

9 sg sg
3
+ (I I) TI sg I + (D1sg I3 )sg TD1sg I3 T ,
2
2

(cg1 ) =
(cg2 )

(11.250)

(11.251)

where
TI sg I T = TI T T = TI2 T ,

TD1sg I3 = TD1 T T = TD1 TI T .


(11.252)

11.8 Generalization to O(N )-Symmetric Model

777

The resulting generalization of (11.219) is


i
3 h sg 2
(I ) + 2I sg I reg TI 2 3 (I4sg + I sg I reg ) TI4
4
i
9 h sg 2
+ (I ) + I sg I reg TI2 .
2

cg2 =

(11.253)

With the reduced matrrix elements (11.239) and (11.244), the regular terms I reg
cancel as before (here somewhat more miraculously), and cg2 simplifies to
cg2

sg 2

= (I )

9
3
TI4 TI2 3I4sg TI4 .
2
4


(11.254)

Inserting the reduced matrix elements (11.238)(11.244), and the reduced coupling
constant = gS
D of Eq. (11.214), the counterterms become, finally,
c

m2

c
cg

1 N +2
1 N +2 2 1 1
1 N +2
,
+ 2

=
2
2
2 3
4
3
4
2
3
1 N +2
= 2
,
(11.255)
48 3
)
(


2
1 5N + 22
1
3 N +8
2 3 1 N + 6N + 20
.
+
+3
+
=
2 9
4 2
27
22 4
27
"

A further generalization which is sometimes encountered in the literature and


which is useful for applications is the case of n fields i (i = 1, . . . , n) each of which
occur in an O(q)-symmetric combination
L =

q
n,q h
q
n
i
X
X
1X
1 X
(i )2 + m2 (i )2 +
i j
k l . (11.256)
gijkl
2 i,
4! i,j,k,l=1
=1
=1

Among these, the case of a cubic symmetry in the indices is of special interest when
there are only two independent couplings
n
X

i,i =1

gii

q
X

=1

(i )2

q
X

(i )2
=1

= g1

q
n X
X

i=1 =1

(i )2

!2

+ g2

n
X
i=1

q
X

(i )2

=1

!2

. (11.257)

The special case of n complex fields (i. e. i with = 1, 2 governs the statistical
mechanics of ensembles of dislocation lines in a crystal. They also are used to
describe phase transforms in cubic crystals. The interaction (11.263) corresponds
to the tensor (11.229) being replaced by
Ti,i,k,l = g1 Si,j,k,l + g2 Fi,j,k,l
g1
=
(ij kl + ik jl + il jk )
3
g2
( + + ) ij ik il .
+
3

(11.258)

778 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


It is then straightforward to generalize the previous results. Take, for example, TI
arising from
Tabef Tcdef |symm = Tabcd TI .

(11.259)

Multiplying the tensors S, F with each other we find


nq + 8
S,
9
q+8
=
F,
9
q+2
2
F+
S.
=
3
9

SS|symm =
F F |symm
F S|symm

(11.260)

Hence, the lowest-order counterterms cg = 32 TI are replaced by


q+2
3 pq + 8
1 + 2
2 ,
2
9
9


4
3 q+8
1 + 2 .
=
2
9
3

cg 1 =
cg 2

11.9

(11.261)

Finite S-Matrix Elements

We are now in a position to calculate also finite S matrix elements. Again for
simplicity we shall consider here only elastic scattering between two -particles, the
generalization being straight-forward.
Recalling Eq. (10.75), the four-point function G(4) (x1 , . . . , xn ) has the following
structure: It consists of disconnected and connected parts with the disconnected
ones being the product of two interacting single-particle Green functions: We may
write
G(4) (x4 , x3 , x2 , x1 ) =

G(2) (x4 , x3 )G(2) (x2 , x1 ) + 2 perm + G(4)


c (x4 , x3 , x2 , x1 ) .
(11.262)

The two-particle Green functions were obtained by summing the one-particle irreducible graphs of the self energy (see (11.26)) and forming the geometric series in
momentum space
G(2) (q) =

q2

m2

i
.
(q) + i

(11.263)

The connected four-point function is then obtained


G(4) (q4 , q3 , q2 , q1 ) = G(2) (qn ) G(2) (q3 ) (4) (q4 , q3 , q2 , q1 ) G(2) (q2 ) G(2) (q1 ) , (11.264)
where the vertex function (4) (q4 , q3 , q2 , q1 ) is the sum of all one-particle irreducible
vertex graphs. We now prove that this vertex function gives directly the scattering

11.9 Finite S -Matrix Elements

779

amplitude if the momenta q4 , q3 , q2 , q1 , are evaluated on the physical mass


2shells 2of4
the incoming and outgoing particles where the energies are given by q0 = q + m .
The argument proceeds very similar to the previous discussion in Section (10.7). We
form again the limit (10.100) but with the energies qi0 on the physical mass shells
(rather than the bare mass shells used there). We are then confronted with studying
the limit (10.101) with the fully interacting two-point function G(2) (q),
lim

x01

2q10 i(q10 x01 q1 x1 ) Z d4 q iq(zx)


i
e
e
,
4
2
2
V
(2)
q m (q) + i

(11.265)

0
rather
2 than2 the free one. The integrand has poles at the physical energy value q =
q + m i. In addition,
q there are cuts from the higher Feynman amplitudes,
the lowest starting at q0 = q2 + (3m)2 from the second-order graph q q . The
cuts are slightly displaced in the imaginary direction due to the i attached to each
mass. Thus the complex q0 plane looks as shown in Fig. 11.1. If we now distort the

Figure 11.1 Singularities in complex q0 -plane of Feynman propagator. There are poles
and cuts due to three- and more-particle intermediate states in the diagram.

contour of the q 0 -integration we pick up contribution of the pole at q 0 = q2 + m2


as well as the cut. But in the limit x01 only the pole contributes since at
the cut q 0 and q10 do not cancel and an infinitely rapid oscillation remains which
are equivalent to zero by the standard Lebesgue lemma. Thus we obtain the same
(2)
rules as before but now for the interacting Greem functions G(2) rather than G0 ,
and with the particles on the physical mass shell. As a consequence, the four-point
function (11.264) gives the following S-matrix element
S (q4 , q3 |q2 , q1 ) = q3 q1 q4 q2 + q3 q2 q4 q1
+ (2)4 (4) (q4 + q3 q2 q1 )

4
Y

i=1

1
q

2V

qi0

(4) (q1 , q3 , q2 , q1 ) , (11.266)

with (4) (q1 , q3 , q2 , q1 ) being the amputated OPI


function evaluated for
2 four-point
0
2
the on-shell momenta q1 , q3 , q2 , q1 with q = q + m for the incoming and q 0 =
4

Remember that the n-point vertex functions (n) (qn .qn1 , . . . , q4 , q3 , q2 , q1 ) are defined for any
off-shell values of the energies.

780 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization

q2 + m2 for the outgoing particles. The first two terms in (11.266) correspond
to the particles running past each other without scattering (the so-called direct
beam right behind the sample in the experiment). The symmetrization in q3 , q4 is
a reflection of particle identity. The second term contains the amplitude for true
scattering. The factor after the total momentum conservation (2)4 (4) (pf pi )
function is i times the t-matrix defined in Section 10.7 [recall Eq. (9.291)]
In a scattering process where two particles enter and four emerge (a so-called
production process), the disconnected parts are absent and we have directly
4 (4)

S (q6 , q5 , q4 , q3 |q2 , q1 ) = (2)

6
X
i=3

qi

2
X

qi i(6) (q6 , q5 , q4 , q3 , q2 , q1 )

i=1

6
Y

i=1

1
q

2V qi0

(11.267)

Appendix 11A

Alternative Proof of Veltman Integral Rule

Here we shall convince ourselves in another way of the consistency of the Veltman
integral rule (11.128):
Z
dD kk 2 = 0,

D < 2,

(11A.1)

In Eq. (11.223), this was derived by taking the limit m 0 in the dimensionally
regularized integral (11.123) for D > 2. Instead of the integral (11A.1), we shall
discuss the more general integral
Z

dD k f (k),

(11A.2)

with a rotationally invariant f (k) which is bounded for large k by


|f (k)| < A k 2 .

(11A.3)

Then the integral exists for 0 < D < 2 whereas it is not immediately defined for
2 < D < 0 since, after integrating out the angular variables, the integrand has an
IR-singularity at k = 0:
Z

dD k f (k) SD

dk k D1 f (k) .

This singularity can be circumvented by the same trick as the singularity at = 0


in the integral (11.121). With some c > 0 we rewrite
Z

d kf (k) = SD
+

(Z

dk k D1 f (k)

dk k

D1

[f (k) f (0)] + f (0)

cD
.
D

(11A.4)

Appendix 11A

781

Alternative Proof of Veltman Integral Rule

This is independent of c and defined for 2 < D < 2. In the limit c 0 it gives
back the previous integral defined only for 0 < D < 2. For 2 < D < 0 we can
take the opposite limit c and only the second term survives.
Z

d kf (k) =

dD k[f (k) f (0)].

(11A.5)

For D 0, the third term gives


Z

d0 kf (k) = f (0).

(11A.6)

Repeated application of this procedure permits a continuation of the integral to


arbitrarily low dimensions 2l 2 < D < 2l (l = 0, 1, 2, . . .) via the formula
Z

dD kf (k) =

dD k [f (k) f (0)] k 2 f (0) . . . k 2l f (l) (0) .

with

d2l kf (k) = ()l f (l) (0).

(11A.7)

(11A.8)

Note that if an integral has initially no range of D where it exists, such as with the
function f (k) = k 2 (k 2 + a)/(k 2 + b), the above procedure serves to define it for
2l + 2 < D < 2 (l = 1, 2, 3, . . .). With l = 1, for example, we define
Z

d kk

2 k

k2 + a a
d kk

k2 + b
b
Z
2
ba k
=
dD kk 2
,
b k2 + b

+a

2
k +b

(11A.9)

with the right-hand side existing for 2 + 2 < D < 2.


We now observe that this continuation procedure can make sense only if the
difference between the two sides is formally equal to zero, i.e., if
Z

dD kk 2 = 0,

D < 2.

(11A.10)

The basic reason for this identity is, of course, the fundamental requirement of
dimensional regularization that the Gaussian integral (11.119) was assumed to hold
for all D, positive and negative values. Setting q and m equal to zero this implies
Z

k 2

d ke

 D/2

(11A.11)

For D < 0 the limit 0 gives once more (11.124) while for D = 0 we obtain
(11A.6). Inserting a factor k 2 we have
Z

dD kk 2 e k =

(D/2 ) D/2 D/2



.
(D/2)

(11A.12)

782 11 Extracting Finite Results from Perturbation Series. Regularization, Renormalization


For D < 2 the limit 0 gives
Z

dD kk 2 = 0,

(11A.13)

in agreement with (11A.10).


Note that this relation implies a specific choice af a small-k cutoff. Suppose we
had regularized the small-k regime of the Gaussian integral by an exponential cutoff
2
e/k . Then we would have obtained
Z

dD kk 2 e k

2 /k 2

(D/2 ) D/2
2

K
(2
).
D/2
(D/2) D/2

(11A.14)

where K (z) is the modified Bessel function. For small z this behaves like
 

z
1
K (z) = K (z) ()
2
2

(11A.15)

so that for small and the integral would have been


Z

dD kk 2 e k

2 /k 2

(D/2 ) D/2

(D/2)

D/2

,
D/2
D/2
,
D/2

D > 2,

(11A.16)

D < 2,

with the limit 0 no longer giving zero for D < 2 but a very large number. In
a renormalizable massless theory this infinity must cancel order by order in parturbation theory. It is this prpperty which makes it possible to use the infrared cutoff
R
implied by the integral formula dD kk 2 = 0 also for D > 2.

Notes and References


A detailed account of the techniques for extracting finite results from higher-order perturbation
exansions up to five loops can be found in the book:
H. Kleinert and V. Schulte-Frohlinde, Critical Properties of 4 -Theories, World Scientific, Singapore 2001, pp. 1489 (klnrt.de/b8).
The individual citations refer to:
[1] H. Kleinert, J. Neu, V. Schulte-Frohlinde, K.G. Chetyrkin, and S.A. Larin, Five-Loop Renormalization Group Functions of O( n)-symmetric 4 -Theory and -Expansions of Critical
Exponents up to 5 , Phys. Lett. B272, 39 (1991);
H. Kleinert and V. Schulte-Frohlinde, Exact Five-Loop Renormalization Group Functions
of 4 -Theory with O(N )-Symmetric and Cubic Interactions. Critical Exponents up to 5 ,
Phys. Lett. B 342, 284 (1995).

Nothing is mightier that an idea


whose time has come
V. M. Hugo (1802-1885)

12
Quantum Electrodynamics
In Chapter 4 we have learned how to quantize relativistic free fields and in Chapters
5, 6 how to deal with interactions perturbatively if the coupling is small. Fortunately,
there is a large set of physical phenomena for which the techniques developed so far
are powerful enough to supply theoretical results to be compared with observable
quantities. These, moreover, turn out to be in extremely good agreement with the
experimental measurements. It is the quantized theory of interacting electrons and
photons called quantum electrodynamics or shortly QED.

12.1

Gauge Invariant Interacting Theory

The free Lagrangians of electrons and the photons are known from Chapter 5 as

L(x) = L(x) + L(x)

(12.1)

with
e
(i/
m) (x)
L(x) = (x)

1
1 2
2
L(x) = F (x) = [ A (x) A (x)]
4
4

(12.2)
(12.3)

When quantizing the photon theory, there were subtleties due to the gauge freedom
in the choice of the gauge fields A :
A (x) A (x) + (x).

(12.4)

For this reason, there are different ways of constructing a Hilbert space of free
particles:
The first [see Subsec. 7.1.2] was based on a quantization of only the two physical
transverse degrees of freedom. The time component of the gauge field A0 and the
spatial divergence A(x) had no canonically conjugate field and were therefore
classical fields which had no manifestation at all in the Hilbert space. The two were
related by Coulombs law which in the absence of charges reads
2 A0 (x) = 0 A(x).
783

(12.5)

784

12 Quantum Electrodynamics

Only the transverse components, defined by




Ai (x) ij i j /2 Aj (x, t)

(12.6)

became operators. These components represent the proper dynamical variables of


the system. After quantization, the positive- and negative-frequency parts of these
fields defined creation and annihilation operators for the electromagnetic quanta,
the photons of right and left circular polarization.
This method had an esthetical disadvantage that two of the four components of
the vector field A (x) required a different treatment. The components which become
operators change with the frame of reference in which the canonical quantization
procedure is performed.
To circumvent this, a second quantization procedure was presented which, at first
sight, looked somewhat contrived. It started out by modifying the initial photon
Lagrangian by a gauge-fixing term
LGF (x) = D A (x) + D 2 (x)/2.

(12.7)

In the Gupta-Bleuler quantization scheme, the propagator took a pleasant covariant


form. But this was at the expense of another disadvantage was that this Lagrangian
describes the propagation of four particles of which only two correspond to physical
states. Accordingly, the Hilbert space contained two kinds of unphysical particle
states, states with negative and with zero norm. Still, a physical interpretation of
the formalism was found by means of a subsidiary condition selecting the physical
subspace in the Hilbert space of free particles.

12.1.1

Reminder of Classical Electrodynamics of Point Particles

In this chapter we want to couple electrons and photons with each other by an
appropriate interaction term and study the resulting interacting field theory which
is the famous quantum electrodynamics. Since the coupling should not change the
two physical degrees of freedom described by the four-component photon field A , it
is important to preserve the gauge invariance, which was so essential in assuring the
correct Hilbert space. The prescription how this can be done has been known for a
long time in the classical electrodynamics of point particles: There, a free relativistic
particle moving along an arbitrary parametrized path x ( ) in four-space is described
by an action
A = mc

Z
dx dx
v2 (t)
= mc2 dt 1 2
d d
c
"

#1/2

(12.8)

where x0 ( ) = t is the time and dx/dt = v(t) the velocity along the path. If the
particle has a charge e and lies at rest at position x, its potential energy is
V (t) = e(x, t)

(12.9)

785

12.1 Gauge Invariant Interacting Theory

where
(x, t) = A0 (x, t).

(12.10)

In our convention, the charge of the electron e has a negative value to have agreement
with the sign in the historic form of the Maxwell equations
E(x) = 2 (x) = (x),

B(x) E(x)
= A(x) E(x)

h
i
1

= 2 A(x) A(x) E(x)


= j(x). (12.11)
c

If the electron moves along a trajectory x(t), the potential energy becomes
V (t) = e (x(t), t) .

(12.12)

In the Lagrangian L = T V this contributes with the opposite sign


Lint (t) = eA0 (x(t), t)

(12.13)

giving a potential part of the interaction


int

A |pot = e

dtA0 (x(t), t) .

(12.14)

Since the time t coincides with x0 (t)/c of the trajectory, this can be expressed as
Aint |pot =

eZ
dx0 A0 (x).
c

(12.15)

In this form it is now quite simple to write down the complete electromagnetic
interaction purely on the basis of relativistic invariance. The minimal invariant
extension of (12.15) is obviously
Aint =

eZ
dx A (x).
c

(12.16)

Thus, the full action of a point particle can be written more explicitly as
A =

dtL(t) = mc

= mc

v2
dt 1 2
c
"

ds
#1/2

e
c

dx A (x)
1
dt A v A .
c


(12.17)

The canonical formalism supplies the canonically conjugate momenta


P=

v
e
e
L
= mq
+ A p + A.
v
c
1 v2 /c2 c

(12.18)

786

12 Quantum Electrodynamics

Thus the velocity is related to the canonical momenta and external field via
P ec A
v
.
= r
2
c
e
2
2
P cA + m c

(12.19)

This can be used to calculate the Hamiltonian via the Legendre transform
H=

L
vL= PvL
v

(12.20)

giving
H=c

s


e
P A
c

2

+ m2 c2 + eA0 .

(12.21)

In the non-relativistic limit, this has the expansion


1
e
H = mc +
P A
2m
c


2

+ eA0 + . . . .

(12.22)

The rest energy mc2 is usually omitted in this limit.

12.1.2

Electrodynamics and Quantum Mechanics

When going over from quantum mechanics to second quantized field theories we
found the rule that a non-relativistic Hamiltonian
H=

p2
+ V (x)
2m

(12.23)

became an operator
H=

2
d x (x, t)
+ V (x) (x, t),
2m

"

(12.24)

where we have omitted the hat on top of H, for brevity. With the same rules we see
that the second quantized form of the interacting nonrelativistic Hamiltonian in a
static A(x) field with the Hamiltonian (12.22) (minus mc2 ),
(P eA)2
+ eA0 ,
H=
2m

(12.25)

is given by
H=

"

1
e
d x (x, t)
i A
2m
c
3

2

+ eA

(x, t).

(12.26)

787

12.1 Gauge Invariant Interacting Theory

When going to the action of this theory we find


A=

dtL =

dt

d x (x, t) it + eA0 (x, t)


e
1
(x, t) i A
+
2m
c


2

(x, t) .

(12.27)

It is easy to verify that (12.26) reemerges from the Legendre transform


H=

L
(x, t) L.

(x,
t)

(12.28)

The action (12.27) holds also for time-dependent A (x) fields.


The equations show that electromagnetism is introduced into a free quantum
theory of charged particles following the minimal substitution rule
e
i A(x, t),
c
t t + ieA0 (x, t),
(12.29)
or covariantly:

e
i A (x).
(12.30)
c
The substituted action has the important property that the gauge invariance of
the free photon action is preserved by the interacting theory: If we perform the
gauge transformation
A (x) A (x) + (x),

(12.31)

A0 (x, t) A0 (x, t) + t (x, t)


A(x, t) A(x, t) (x, t),

(12.32)

i.e.,

the action remains invariant if we simultaneously change the fields (x, t) of the
charged particles by a spacetime-dependent phase
(x, t) ei(e/c)(x,t) (x, t).

(12.33)

Under this transformation, the derivatives of the field change like


e
i (x, t) ,
(x, t) e
c
i(e/c)(x,t)
t e
(t iet ) (x, t).
i(e/c)(x,t)

(12.34)

The modified derivatives appearing in the action have therefore the following simple
transformation law:




e
e
i(e/c)(x,t)
i A (x, t) e
i A (x, t),
c 
c



e 0
t + i A (x, t) ei(e/c)(x,t) t + ieA0 (x, t).
(12.35)
c

788

12 Quantum Electrodynamics

These combinations of derivatives and gauge fields are called covariant derivatives
and written as
e
D(x, t) i A (x, t)
c 

Dt (x, t) t + ieA0 (x, t)


(12.36)

or, in four-vector notation,

D (x) =

e
+ i A (x).
c

(12.37)

Here the adjective covariant does not refer to the Lorentz group but to the gauge
group. It records the fact that D transforms under local gauge changes (12.29)
of in the same way as itself in (12.33):
D (x) ei(e/c)(x) D (x).

(12.38)

With the help of this covariant derivative any action which is invariant under global
phase changes [i.e., U(1)-invariant in the sense discussed in Chapter 4] by a constant
phase angle
(x) ei (x)

(12.39)

can easily be made invariant under local gauge transformations (12.31). We merely
have to replace all derivatives by covariant derivatives (12.37) and adding the gauge
invariant photon Lagrangian (12.3).

12.1.3

Principle of Nonholonomic Gauge Invariance

The minimal substitution rule can be viewed as a consequence of a more general


principle of nonholonomic gauge invariance. The physics of the initial action (12.17)
is trivially invariant under the addition of a term
e
A =
c

dt x (t) (x).

(12.40)

The integral runs over the particle path and contributes only a pure surface term
from the endpoints:
e
A = [(xb ) (xa )] .
c

(12.41)

This it does not change the particle trajectories. If we now postulate that the
dynamical laws of physics remain valid if we admit multivalued gauge functions
(x) for which the Schwarz integrability criterion is violated, i.e., which possess
noncommuting partial derivatives:
( )(x) 6= 0,

(12.42)

789

12.1 Gauge Invariant Interacting Theory

then the derivatives


A (x) = (x)

(12.43)

have a nonzero curl F = ( A A ) = ( )(x) 6= 0, and the action


(12.41) coincides with the interaction (12.16).
Similarly we can derive the equations of motion of a wave function in an electromagnetic field from that in field-free space by noting the trivial invariance of
quantum mechanics without fields under gauge transformations (12.33) and extending the permissible gauge functions (x) to multivalued functions for which the
partial derivatives do not commute as in (12.42).
In either case, the nonholonomic gauge transformations convert the physical laws
obeyed by a particle in Euclidean spacetime without electromagnetism into those
with electromagnetic fields.
This principle is dicussed in detail in the literature [7]. It can be generalized to
derive the equations of motion in a curved spacetime from those in flat spacetime
by nonholonomic coordinate transformations which introduce defects in spacetime.

12.1.4

Electrodynamics and Relativistic Quantum Mechanics

Let us follow this rule for relativistic electrons and replace, in the Lagrangian (12.2),

/ =

e
e
+ i A = / + i A
/
c
c

D
/.

(12.44)

In this way we arrive at the Lagrangian of quantum electrodynamics (QED)


1 2
(i/
.
L(x) = (x)
D m) (x) F
4

(12.45)

The classical field equations can easily be found by extremizing the action and
variation with respect to all fields which gives
A
= (i/
D m) (x) = 0

(x)
1
A
= F (x) j (x) = 0.
A (x)
c
with the current density

(12.46)
(12.47)

j (x) ec(x)
(x).

(12.48)

1
F (x) = j (x).
c

(12.49)

Equation (12.47) coincides Maxwell equation for the electromagnetic field around a
classical four-dimensional vector current j (x):

In the Lorentz gauge A (x) = 0, this equation reads simply


1
2 A (x) = j (x).
c

(12.50)

790

12 Quantum Electrodynamics

The current j combines the charge density (x) and the current density j of
particles of charge e in a four-vector
j = (c, j) .

(12.51)

In terms of electric and magnetic fields E i = F i0 , B i = F jk , the field equations


(12.49) turn into the Maxwell equations
0 = e
E = = e
= 1 j = e .
BE
c
c

(12.52)

The first is Coulombs law, the second Amperes law in the presence of charges and
currents.
Note that the physical units employed here differ from those used in many books
of classical electrodynamics [8] by the absence of a factor 1/4 on the right-hand
side. The Lagrangian used in those books is
1 2
1
F (x) j (x)A (x)
8
c

i 
1
1 h 2
E (x) B2 (x) (x)(x) j(x) A(x) ,
=
4
c

L(x) =

(12.53)

which leads to Maxwells field equations


E = 4,
4
B =
j.
c

(12.54)

The form employed


conventionally
in quantum field theory arises from this by re
placing A 4A and e 4e2 . The charge of the electron in our units has
therefore the numerical value
q

e = 4 4/137
(12.55)

rather than e = .

12.2

Noethers Theorem and Gauge Fields

In electrodynamics, the conserved charge resulting from the U(1)-symmetry of the


matter Lagrangian is the source of a massless particle, the photon. This is described
by a gauge field which is minimally coupled to the conserved current. A similar
structure will be seen to exists for many internal symmetries giving rise to nonabelian
versions of the photon, such as gluons which give rise to strong interactions, and W and Z-vector mesons, which mediate the weak interactions. It is useful to reconsider
Noethers derivation of conservation laws in such theories.
The conserved matter current in a locally gauge invariant theory cannot be found
any more by the rule (8.118), which was so useful in the globally invariant theory. For

791

12.2 Noethers Theorem and Gauge Fields

the gauge tranformation of quantum electrodynamics, the derivative with respect


to the local field transformation (x) would be simply given by
j =

L
.

(12.56)

and yield identically zero, due to local gauge invariance. We may, however, subject
just the matter field to a local gauge transformation at fixed gauge fields. Then we
obtain the correct current

L
j
(12.57)
.

Since the complete change under local gauge transformations sx L vanishes identically, we can alternatively vary only the gauge fields and keep the electron field
fixed

L
(12.58)
.
j =
e

This is done most simply by forming the functional derivative with respect to the

gauge field and omitting the contribution of L:


e

L
j =
.

(12.59)

An interesting consequence of local gauge invariance can be found for the gauge
field itself. If we form the variation of the pure gauge field action

s A =

A
d4 x tr sx A
A

(12.60)

and insert for sx A an infinitesimal pure gauge field configuration


sx A = i (x)

(12.61)

the right-hand side must vanish for all (x). After a partial integration this implies
the local conservation law j (x) = 0 for the current

A
j (x) = i
.
A

(12.62)

In contrast to the earlier conservation laws derived for matter fields which were valid
only if the matter fields obey the Euler-Lagrange equations, the current conservation
law for gauge fields is valid for all field configurations. It is an identity, often
called Bianchi identity due to its close analogy with certain identities in Riemannian
geometry.
To verify this, we insert the Lagrangian (12.3) into (12.62) and find j =
F /2. This current is trivially conserved for any field configuration due to the
antisymmetry of F .

792

12.3

12 Quantum Electrodynamics

Quantization

The canonical formalism can be used to identify canonical momenta


(x) =

L
= (x)
(x)

(12.63)

and
Ai (x) i (x) = F 0i (x) = E i (x)

(12.64)

and to find the Hamiltonian density


L(x)
L
L(x)
(12.65)
(x) +

A k (x)
(x)
+ m) + 1 (E2 + B2 ) + A0 E + e
A .
= (i
2

H(x) =

Here and in all the subsequent discussion we use natural units in which the light
velocity is equal to unity.
The quantization procedure in the presence of interactions now goes as follows:
The Dirac field of the electron has the same equal-time anticommutation rules as in
the free case:
{ a (x, t), b (x , t)} = (3) (x x)ab ,
{(x, t), (x, t)} = 0,
{ (x, t) (x t)} = 0.

(12.66)

For the photon field we first write down the naive commutation rules of the spatial
components:
[Ei (x, t), Aj (x , t)] = i (3) (x x )ij
[Ai (x, t), Aj (x , t)] = 0,
[Ei (x, t), Ej (x , t)] = 0.

(12.67)
(12.68)

As in Subsec. 5.3.1, the first commutator cannot be true here, since by Coulombs
law (12.52),
2 A0
E = A
= e ,

(12.69)

and we want the canonical fields Ai to be independent of and commute with them.
The contradiction can be removed just as in the free case by postulating (12.67) only
for the transverse parts of Ei and using ijT (x x ) as in (7.335) while letting the

793

12.3 Quantization

longitudinal part A(x, t) be a c-number field, since it commutes with all Ai (x, t).
The correct commutation rules are the following:
A j (x, t), Aj (x , t)

= iijT (x x ),

[Aj (x, t), Aj (x , t)] = 0,


i
A i (x, t), A j (x , t) = 0.

(12.70)

To calculate the temporal behavior of an arbitrary observable composed of , and


Ai , A i fields, only one more set of commutation rules has to be specified which are
those with A0 . This occurs in the Hamiltonian density (12.65) and is not a canonical
variable. In contrast to the free-field case in Section 5.3 it is no longer a c-number.
To see this we have to express A0 in terms of the dynamical fields using Coulombs
law (12.69):
1
A (x, t) =
4
0

d3 x



1

(x , t).
e

A
|x x|

(12.71)

This replaces Eq. (4.265) in the presence of charges. In an infinite volume with
asymptotically vanishing fields there is no freedom of adding a solution of the homogenous Poisson equation (4.266). While A is a c-number field, the time component of the gauge field A0 is now a non-local operator involving the fermion field.
With these fields being independent of the electromagnetic field, A0 still commutes
with the canonical Ai , A j fields:
A0 (x, t), Ai (x ) = A0 (x, t), A i (x , t) = 0.

(12.72)

The commutator with the Fermi fields, on the other hand, is nonzero:
h

A0 (x, t), (x , t) =

e
(x, t).
4|x x |

(12.73)

Note the peculiar property of A0 : It does not commute with the electron field no
matter how large the distance between the space points is. This property is called
nonlocality. It is a typical property of the present transverse covariant quantization
procedure.
Certainly, the arbitrary c-number function A(x, t) can be made zero by an
appropriate gauge transformation as in (4.254).
In the Hamiltonian, the field A0 can be completely removed by a partial integration:
Z

d3 xA0 E =

d3 xA0 E,

(12.74)

setting the surface term equal to zero, and using the field equation
E = e .

(12.75)

794

12 Quantum Electrodynamics

Then
Z

H =

d3 x H

(12.76)

e
1 2
d3 x i ( i A) + m +
E + B2
c
2




When looking at this expression, one may wonder where the electrostatic interaction
has gone. The answer is found by decomposing the electric field
E i = 0 Ai + i A0

(12.77)

into longitudinal and transverse parts ELi and ETi with EL ET = 0:


ii
= A 2 Aj ,

!
ij
ij
0
Aj .
=
2

ELi
ETi

(12.78)

Then the field energy becomes


1
2

d x E +B

1
=
2

dx

E2T

+B

1
+
2

d3 xE2L .

(12.79)

Using (12.71) we see that the longitudinal field is simply given by


1
1
EL (x) = d3 x
e (x , t)(x , t)
4
|x x |


1

=
e (x)(x) .
2
Z

(12.80)

This is the Coulomb field caused by the charge density of the electron e (x)(x).
The field energy carried by ELi (x) is
1
2

2
e2
1

(x)(x)
d
=
d3 x
2
2
2 Z
1
e
d3 x (x)(x) 2 (x)(x)
=
2

2 Z
1
e
d3 xd3 x (x, t)(x, t)
(x , t)(x , t),
=
8
|x x |
3

x E2L (x)



(12.81)

and coincides precisely with the Coulomb energy associated with the charge density
(12.52).
R
The term 21 d3 x (E2T + B2 ) in Eq. (12.79) is an operator and contains the energy
of the field quanta.
In order to develop a perturbation theory for QED in this quantization, we
must specify the free and interacting parts of the action. Since A0 and A are

795

12.4 Perturbation Theory

unquantized and appear only quadratically in the action, they may be eliminated
in the same way as in the energy and lead to the action
A=

i
1h

d4 x (x)(i/
M)(x) + E2T (x) B2 (x) + Aint ,
2


(12.82)

where the first two terms are the actions of the quantized Dirac and transverse
electromagnetic fields and AT , and Aint is the interaction
Z
Z
e2 Z
1
1

dt d3 xd3 x (x, t)(x, t)


(x
,
t)(x
,
t)
+
d4 x j AT .

8
|x x |
c
(12.83)
The interaction contains two completely different terms: The first is an instantaneous Coulomb interaction at a distance which takes place without retardation
and involves four Dirac fields. It is nontrivial to show that this does not cause any
conflicts with relativity. This will be done at the end of Section 13.13.
The special role of the Coulomb interaction is avoided by the Gupta-Bleuler
quantization procedure. Here the free action is

Aint =

A=

1
d4 x (i/
m) F F D A + D 2 /2 ,
4


(12.84)

and the interaction has the manifestly covariant form


A

12.4

int

d4 x j A .

(12.85)

Perturbation Theory

It is easy to set up the Feynman diagrams. The free-particle propagators of the


electrons was given in (7.278). Since we want to calculate the effect of interactions,
we shall from now on attach to the free propagator a subscript 0. Thus
S0 (x x ) =

d4 p
i

eip(xx ) .
4
(2) p/ M + i

(12.86)

The propagator of the free photon depends on the gauge. It is most simple in the
Gupta-Bleuler quantization scheme, where [see (7.500)]

G
0 (x, x )

= g G0 (x, x ) = g

d4 q
i

eik(xx ) .
4
2
(2) q + i

(12.87)

In a Wick expansion, each contraction is represented by one of these propagators:


(x)(x ) = S0 (x x ),

(12.88)

A (x)A (x ) = G
0 (x x ).

(12.89)

They are pictured by the lines

796

12 Quantum Electrodynamics

i
p/ m

= g

q2

i
+ i

The interaction Lagrangian


A (x)
Lint = e

(12.90)

= e .

(12.91)

is pictured by the vertex

With these graphical elements we must form all Feynman diagrams which can contribute to a given physical process.
In the transverse quantization scheme, the Feynman diagrams are much more
complicated. Recalling the propagator (7.351), the photon line stands now for

= Pphys
(q)

i
q 2 + i

with the physical off-shell polarization sum [recall (12.7)]

Pphys
(q) = g

q q
q2
q + q
+
q

.

(q)2 q 2
(q)2 q 2
(q)2 q 2

(12.92)

The photon propagator is very complicated due to the appearance of the framedependent vector = (1, 0, 0, 0). As a further complication, there are diagrams
from the four-fermion Coulomb interactions in (12.83). These can be pictured by

i 0
0.
2
q

We now observe that these diagrams may be obtained from an auxiliary interaction
Aint =

d4 xj 0 A0

(12.93)

797

12.4 Perturbation Theory

assuming the A0 -field to have the propagator


0

i
q2

If this propagator is added to the physical one, it cancels precisely the last term in
the off-shell polarization sum (12.92), which becomes effectively

Pphys,eff
(k) = g

q + q
q q
+ q
.
2
2
(q) q
(q)2 q 2

(12.94)

Of course, the final physical results cannot depend on the frame in which the
theory is quantized. Thus is must be possible to drop all -dependent terms. We
prove this in three steps:
First, a photon may be absorbed (or emitted) by an electron which is on the
mass shell before and after the process. The photon propagator is contracted with
an electron current as follows

u(p , s3 ) u(p, s3 )Pphys,eff


(q).

q = p p.

(12.95)

Since the spinors on the right and left-hand side satisfy the Dirac equation, the
current is conserved and satisfies
u(p , s3 ) u(p, s3)q = 0.

(12.96)

This condition eliminates the terms containing the vector q in the polarization sum
(12.94). Only the reduced polarization sum

Pred
(q) = g

(12.97)

survives, which is the polarization tensor of the Gupta-Bleuler propagator (12.87).


The same elimination happens if a photon is absorbed by an internal line, although due to a slightly more involved mechanism. An internal line may arise in
two ways. An electron may enter a Feynman diagram on the mass shell, absorb a
number of photons, say n, and leave again on mass shell as shown in Fig. 12.1. The

Figure 12.1 An electron on the mass shell absorbing several photons and leaving again
on the mass shell.

798

12 Quantum Electrodynamics

associated off-shell amplitude is


a(p , p, qi ) =

p/

1
1
1
1
q/ n
q/ n1 q/ n1 q/ 3
q/ 2
q/ 1 . (12.98)
M p/ n1 M
p/ 2 M p/ 1 M

It has to be amputated and evaluated between the initial and final spinors, which
amounts to multiplying it from the left and right with u(p , s3 )(/
p M) and (/
p
M)u(p, s3 ), respectively. If an additional photon is absorbed, it must be inserted
n + 1 times as shown in Fig. 12.2. At each vertex, there is no current conservation

Figure 12.2 An electron on the mass shell absorbing several photons, plus an additional
photon, and leaving again on the mass shell.

since the photon lines are not on mass shell. Nevertheless, the sum of all n + 2
diagrams does have a conserved current.
To prove this we observe the following Ward-Takahashi identity for free particles
[1, 2]:
1
1
1
1
q/
=

.
(12.99)
p/ r + q/ M p/ r M
p/ r M
p/ r + q/ M

More details on this important identity will be given in the next section.
The sum of all off-shell absorption diagrams can be written as

1
1
q/ n
q/ n1
+ q/ M p/ n1 + q/ M
1
1
1
q/
q/ 2
q/ 1 u(p, s3). (12.100)

p/ r + q/ M p/ r M
p/ 1 M

a(p , p, qi ; q) =

p/

to be evaluated between u(p + q, s3 )(/


p + q/ M) and (/
p M)u(p, s3 ). With
the help of the Ward-Takahashi identity we can now remove the q/ recursively from
a(p , p, qi ; q) and remain with the difference
a(p , p, qi ; q) = a(p , p, qi ) a(p + q, p + q, qi ).

(12.101)

When evaluating the right-hand side between the above spinors, we see that the
first term in the difference vanishes since the left-hand spinor satisfies the Dirac

799

12.4 Perturbation Theory

equation. The same thing holds for the second term and the right-hand spinor.
Thus the polarization sum in the photon propagator can again be replaced by the
reduced expression (12.97).
Finally, the electron line can be a closed to a loop as shown in Fig. 12.3. Here

Figure 12.3 An internal electron loop absorbing several photons, plus an additional
photon, and leaving again on the mass shell.

the amplitude (12.101) appears in a loop integral with an additional photon vertex:
Z

Z
d4 p
d4 p

tr[/
q
a(p
,
p,
q
;
q)]
=
tr{/
q [a(p , p, qi ) a(p + q, p + q, qi )]}. (12.102)
i
(2)4
(2)4

If the divergence of the integral is made finite by a dimensional regularization, the


loop integral is translationally invariant in momentum space and the amplitude
difference vanishes. Thus, also in this case, the polarization sum can be replaced by
the reduced expression (12.97).
Thus we have shown that due to current conservation and the Ward-Takahashi
identity, the photon propagator in all Feynman diagrams can be replaced by
= G (q) =

ig
q 2 + i

(12.103)

As a matter of fact, for the same reason, any propagator


"

q q
i
= G (q) = 2 g 2 (1 )
q
q

(12.104)

can just as well be used, for any parameter . This is the propagators arising when
adding to the free Lagrangian the fixing term

LGF = D 2 D A .
(12.105)
2
The Feynman value = 1 reduces to (12.103).

800

12.5

12 Quantum Electrodynamics

Ward-Takahashi identity

As must be apparent from the application of Eq. (12.99), the Ward-Takahashi identity plays an important role in ensuring the gauge invariance of loop diagrams. In
fact, the renormalizability of quantum electrodynamics was completed only due to
the discovery of the diagonal part of it, the so-called Ward identity. For free particles, we observe that the Ward-Takahashi identity (12.99) can be written in terms
of the electron propagator (12.88) and the free vertex function 0 (p , p) = as
S01 (p ) S01 (p) = i(p q) 0 (p , p).

(12.106)

The Ward identity is obtained by forming the limit p p of this:


1
S (p) = i0 (p, p).
p 0

(12.107)

The important contribution of Ward and Takahshi was to prove that these identities are valid for the interacting propagators and vertex functions, order by order
in perturbation theory. Thus we may drop the subscripts zero in Eqs. (12.106) [and
(12.107)]:
S 1 (p ) S 1 (p) = i(p p) (p , p).
(12.108)
This identity is a general consequence of gauge invariance as was first conjectured
by Rohrlich [3].
The key observation is that the operator version of the fully interacting elec

tromagnetic current j (x) = e(x)


(x) satisfies at equal times the commutation
rules with the interacting electron and photon fields
[j 0 (x), (y)](x0 y 0 ) = e 0 (x) (x0 y 0 ),
0
0

[j 0 (x), (y)](x
y 0 ) = e(x)
(x0 y 0),
[j 0 (x), A (y)](x0 y 0 ) = 0.

(12.109)
(12.110)
(12.111)

This follows directly from the canonical equal-time anticommutation rules of the
electrons written in the form (x), (y)(x0 y 0 ) = (4) (x y), As a consequence
of (12.111), we find for any local operator O(x):
T j (x)O(y) = j 0 (x)O(yi ) (x0 y 0 ) + T ( j (x)O(y)) .


(12.112)

The first term on the right-hand side arises when the derivative is applied to the
Heaviside functions in the definition (2.232) of the time-ordered product. The generalization to many local operators reads:
n
 X

T j (x)O(y1) O(yi) O(yn ) =




i=1

T O(y1) j 0 (x), O(yi ) O(yn )(x0yi0)


h

+ T ( j (x)O(y1 ) O(yi ) O(yn )). (12.113)

Since the electromagnetic current is conserved, the last term vanishes.

801

12.6 Magnetic Moment of Electron

A particular case of (12.113) for a conserved current is the relation


1 )(y2 ) = eT (y
1 )(y2 ) [(x y1 ) (x y2 )] . (12.114)
Tj (x)(y


Taking this between single-particle states and going to momentum space yields to
all orders in pertrubation theory:
i(p p) S(p ) (p , p)S(p) = S(p ) S(p).

(12.115)

This is precisely the Ward-Takahashi identity (12.106).

12.6

Magnetic Moment of Electron

For dimensional reasons, the magnetic moment of the electron is proportional to the
Bohr magnetic moment
eh
.
(12.116)
B =
2Mc
Since it is caused by the spin of the particle, it is proportional to it and can be
written as
s
(12.117)
 = gB .
h

The proportionality factor g is called the gyromagnetic ratio. If the spin is polarized
in the z-direction, the z-component of  is
= gB

eh 1
1
=g
.
2
2Mc 2

(12.118)

We have discussed in Subsec. 4.16, that as a result of the Thomas precession an


explanation of the experimental fine structure requires the g-factor of the electron
magnetic moment to have the value 2. This is twice as large as that of a charged
rotating sphere of angular momentum L, whose magnetic moment is

=

eh L
,
2Mc h

(12.119)

i.e., whose g-value is unity. The result g = 2 has been found also in Eq. (6.120) by
bringing the Dirac equation in an electromagnetic field to the second-order Pauli
form (6.111).
Let us convince ourselves that a Dirac particle possesses the correct gyromagnetic
ratio g = 2. Consider an electron of momentum p in a electromagnetic field which
changes the momentum to p (see Fig. 12.91). The interaction Hamiltonian is given
by the matrix element
H int =

d3 xA (x)hp |j (x)|pi.

(12.120)

where in Diracs theory

hp , s3 |j (x)|p, s3 i = eh0|a(p , s3 ) (x)


(x) a (p, s3 )|0i

(12.121)

802

12 Quantum Electrodynamics

Inserting the field expansion (7.213) in terms of creation and annihilation operators
(x) =

p,s3

1
q

V Ep /M

eipx u(p, s3 )ap,s3 + eipx v(p, s3 )b (p, s3 ) ,

(12.122)

and using anticommutators (7.217) and (7.218), we obtain

hp

, s3 |j (x)|p, s3 i

ei(p p)x

q
= e
u(p , s3 ) u(p, s3 ) q
.
V Ep /M V Ep /M

(12.123)

The difference between the four-momenta


q p p

(12.124)

is the momentum transfer from the electromagnetic field to the electron.


In order to find the size of the magnetic moment we set up a constant magnetic
field in the 3-direction by letting the second component of the vector potential be
A2 (x) = x1 B3 . Then we put the final electron to rest, i.e., p = (M, 0), and let
the initial electron move slowly in the 1-direction. We create a spinor u(p, s3 ) by
1
0 1
applying a small Lorentz-boost ei (i )/2 to the rest spinors (4.670), and expand
the matrix element (12.123) up to the first order in p. To zeroth order, this yields
u(0, s3 ) u(0, s3 ) = (s3 )(s3 ) 0

(12.125)

showing that the charge is unity. The linear term in q 1 gives rise to a 2-component:
1 1

hp , s3 |j 2 (x)|p, s3 i = e
u(0, s3 ) 2 ei

1 (i 0 1 )/2

u(0, s3 )

eiq x
.
V

(12.126)

The two normalization factors on the right-hand side of (12.123) differ only by
second-order terms in q 1 , Now, since u(0, s3 ) 2 u(0, s3 ) = 0 and i 1 2 /2 = S3 , the
spinors on the right-hand side reduce to
ie 1 u (0, s3 )S3 u(0, s3) = ie 1 s3 .

(12.127)

Momentum conservation enforces 1 = q 1 /M, so that we find


1 1

hp

, s3 |j 2 (x)|p, s3 i

q 1 eiq x
= ie s3
.
M
V

(12.128)

Inserting this into the interaction Hamiltonian (12.120), we obtain


E

int

lim
1

q 0

1 iq 1 x1

d xA2 (x)iq e

d3 x 1 A2 (x)

e
1
s3
M V

1
e
s3 = lim
q 1 0
M V

d3 x A2 (x)1 eiq

1 x1

1
e
s3
M V
(12.129)

803

12.6 Magnetic Moment of Electron

Inserting now the vector potential A2 (x) = x1 B3 yields the magnetic interaction
Hamiltonian
H int = B3

e
s3 .
M

(12.130)

Since a magnetic moment interacts with a magnetic field with the energy B,
we read off the magnetic moment (12.118), implying a gyromagnetic ratio g = 2.
Note that the magnetic field caused by the orbital motion of an electron leads
to a coupling of the orbital angular momentum L = x p with a g-factor g = 1. In
order to see this relative factor 2 most clearly, consider the intraction Hamiltonian
H

int

d3 xA(x)hp , s3 |j(x)|p, s3 i,

(12.131)

and insert the Dirac current (12.123). For slow electrons we may neglect quantities
of second order in the momenta, so that the normalization factors E/M are unity,
and obtain
Z

(12.132)
H int = e d3 x A(x)
u(p , s3 ) u(p, s3 )ei(p p)x
At this place we make use of the so-called Gordon decomposition formula

u(p

, s3 ) u(p, s3 )

= u(p

, s3 )

1
i
(p + p ) +
q u(p, s3 ),
2M
2M


(12.133)

where q p p is the momentum transfer. This formula follows directly from the
anticommutation rules of the -matrices and the Dirac equation.
An alternative decomposition is
i
1
(p + p )F1 (q 2 ) +
q F2 (q 2 ) u(p ),
hp |j |pi = e
u(p )
2M
2M

(12.134)

with the form factors F1 (q 2 ), F2 (q 2 ) related to F (q 2 ), G(q 2 ) via (12.133) by


F (q 2 ) = F1 (q 2 ),

G(q 2 ) = F1 (q 2 ) + F2 (q 2 ).

(12.135)

Then we rewrite the interaction Hamiltonian as


H int =

e Z 3
d xA(x) u(p , s3 ) (p + q iq S) u(p, s3 )eiqx ,
M

(12.136)

where we have used the relations (4.515) and (4.512). We now replace q by the
derivatives ix in front of the exponential eiqx , and perform an integration by parts
to make the derivatives act on the vector potential A(x), with the opposite sign. In
the transverse gauge, the term A(x) q gives zero while iA(x) (q S) becomes
B S. For equal incoming and outgoing momenta, this leads to the interaction
Hamiltonian
Z
e
d3 x [A(x) p + B(x) S]
(12.137)
H int =
M

804

12 Quantum Electrodynamics

We now express the vector potential in terms of the magnetic field as


1
B x,
2

(12.138)

e
B (L + 2S) ,
2M

(12.139)

A(x) =
and rewrite (12.137) in the final form
H int =

where L = x p is the with orbital angular momentum. The relative factor 2


discovered by Alfred Land`e in 1921 between orbital and spin angular momentum
gives rise to a characteristic splitting of atomic energy levels in an external magnetic
field. If the field is weak, both orbital and spin angular momenta will precess around
the direction of the total angular momentum. Their averages will be, for example,
= J J L,
L
J2

= J J S.
S
J2

(12.140)

By rewriting
JL=


1 2
J + L2 S2 ,
2

JS=


1 2
J L2 + S2 ,
2

(12.141)

we see that
= fLJ J,
L

= fSJ J
S

(12.142)

with the factors


fLJ = [J(J + 1) + L(L + 1) S(S + 1)] /2J(J + 1),
fSJ = [J(J + 1) L(L + 1) + S(S + 1)] /2J(J + 1).

(12.143)
(12.144)

Inserting this into (12.139), we obtain the interaction energies of an atomic state
|JMi:
e
H int = gLS
BM,
(12.145)
2M
where
gLS = fLJ + 2fSJ = 1 + [J(J + 1) L(L + 1) + S(S + 1)] /2J(J + 1). (12.146)
is the gyromagnetic ratio, of the coupled system. This has been measured in many
experiments if the external field is not too large (Zeeman effect). If the external field
strength exceeds the typical field strength caused by the electron orbit, the orbital
and spin angular momenta decouple and precess independently around the direction
of the external magnetic field (Paschen-Back or anomalous Zeeman effect).

805

12.7 Decay of Atomic State

Figure 12.4 Transition of an atomic state from a state n with energy En to a lower state
n with energy En , thereby emitting a photon with a frequency = (En En )/h

12.7

Decay of Atomic State

The first important result of quantum electrodynamics is the explanation of the


decay of an atom. In quantum mechanics, this decay can only be studied by means
of the correspondence principle.
Consider an electron in an atomic state undergoing a transition from a state n
with energy En to a lower state n with energy En , whereby a photon is emitted
with a frequency = (En En )/h (see Fig. 12.4). Following the correspondence
principle one imagines the center of charge of the electronic cloud to oscillate with
this frequency with an amplitude into the direction :
x0 = hn | x|ni.

(12.147)

The oscillating charge emits antenna radiation. The classical theory of this has been
recapitulated in Section 5.2, where we have given in Eq. (5.25) the radiated power
per solid angle whose integral led to the Larmor formula (5.26) and reduced to (5.26)
for a harmonic oscillator.
According to the correspondence principle, the same result hold for an atom
decaying from level n to n , if we replace |x0 | by the absolute square of the quantum
mechanical matrix element (12.147):
|x0 |2 |hn | x|ni| ,
2

(12.148)

where  is the direction of oscillations of the charge cloud of the wave function. This
yields the radiated power per unit solid angle
e2 4
dEn n
2
= 2 3 |hn | x|ni| sin2 .
d
8 c

(12.149)

Integrating over all d yields the total radiated power, and if we divide this by the
energy per photon h
, we obtain the decay rate
4
4 e2 3
2
2
|hn | x|ni| = |k|2 |hn | x|ni| .
n n =
3
3 4 h
c
3

(12.150)

806

12 Quantum Electrodynamics

Let us now confirm this result by a proper calculation within quantum electrodynamics, We shall consider only a single electron in a hydrogen-like atom with central
charge Ze. For a nonrelativistic electron of mass M moving in a Coulomb potential
VC (x)

Z
,
|x|

(12.151)

the Hamiltonian reads in the transverse gauge with A(x) = 0,


H=

p2
1
e2 2
A(x, t) p +
A (x, t) + VC (x) + A0 (x, t) = HC + H int . (12.152)
2M
M
2M

where

p2
+ VC (x)
(12.153)
HC =
2M
is the Hamiltonian of the hydrogen-like atom by itself, and H int contains the interaction with the zeroth time-like component of the vector potential A0 (x, t) = 0, which
is determined by the electronic charge distribution via the Coulomb law Eq. (12.71),
and with the radiation field A(x, t) = 0, which has an expansion in terms of photon
creation and annihilation operators as in Eq. (7.338):
A(x, t) =

X
k,h

i
h
1
eikx (k, h)ak,h + h.c. .
2V k

(12.154)

Let |ni be an excited initial state of an atom with quantum number n, which decays
into a lower state a (k, h)|n i with quantum number n . The final state contains,
in addition, a photon of wave vector k, energy = ck, and helicity h. According
to Eq. (9.233), the decay probability of the initial state per unit time is given by
Fermis golden rule (in the remainder of this section we use physical units):
dPnn
=
dt

1
2
d3 kV
,
)
2
h
(E
+
h

E
hn
|a(k,
h)R|ni
n
n


(2)3
h

(12.155)

where R is the Rmatrix which coincides in lowest order perturbation theory with
H int [see (??)]. The matrix element is obviously
hn |a(k, h)H int |ni =

e
c
hn |eikx  (k, h) p|ni.
2V Mc

(12.156)

Performing the integral over the photon momentum (neglecting recoil) we find from
(12.155) the differential decay rate [compare (9.336)]
e2

dnn
= 2
|hn |eikx  (k, h) p|ni|2 .
2
3
d
8 h
M c

(12.157)

The further calculation is simplified by the observation that the wavelength of the
emitted photons is the inverse of their energy (in massless units), and thus of the
order of h
/Z2Mc, about 100 times larger than the atomic diameter which is of

807

12.7 Decay of Atomic State

the order of the Bohr radius aB = h


/ZMc for an atom of charge Z. The expoikx
nential e
is therefore almost unity and can be dropped. This yields the dipole
approximation to the atomic decay rate:
e2

dnn
2
| (k, h) hn |p|ni|2 .
d
8 h
M 2 c3

(12.158)

Another way of writing this result is


dn n
2

| (k, h) hn |x|ni|2 .
d
2h c2

(12.159)

x]/h, and thus, in the


The momentum operator p can be replaced by M x = i[H,

2
matrix element hn |p|ni| , by iM(En En )x/h = iMx. Multiplying the decay
rate by the energy of the photon h
to get the rate of radiated energy, the result
(12.159) coincides with the classical result (12.150).
It is customary to introduce the so-called oscillator strength for an oscillator in
the direction :
fn n

2M

2


X

hn | x|ni


|k| X
= 2 hn | x|ni .

e

(12.160)

This quantity fulfills the Thomas-Reiche-Kuhn sum rule:


X

fn n = 1.

(12.161)

For an atom with Z electrons, the right-hand side is equal to Z.


To derive this sum rule (and a bit more) we define the operator
 1 ei|k|x
E
|k|

(12.162)


i 
h
 i|k|x

e
E ]=
E  = [H,
 +  ei|k|x .
h

2M

(12.163)

has a time derivative

On behalf of the canonical commutator [


pi , xj ] = ihij , the Hermitian-conjugate

of E commutes with E  like

 ] = i h
[E  , E
.
M

(12.164)

Taking this commutator between states hn| and |ni, and inserting in the middle a
P
completeness relation n |n ihn | = 1 , going further to small k, we find indeed the
sum rule (12.161).
Let us calculate the angular properties of the radiation in more detail. The
decomposition of the hydrogen wave functions into radial and angular parts is
hx|nlmi = Rnl (r)Ylm (, ).

(12.165)

808

12 Quantum Electrodynamics

Then the matrix elements of x factorize:


hn l m |x|nlmi = rn l ;nl hl m |
x|lmi,

(12.166)

where the matrix elements of r,


rn l ;nl

drr 2Rn l (r)rRnl (r),

(12.167)

have been calculated by Gordon [22]:


n l

rn l ;nl

(1)
=
4(2l 1)!

v
u
u (n + l)!(n
t

n+n 2l2
+ l 1)!
l+1 (n n )
(4nn
)
(n l)!(n l 1)!
(n + n )n+n

4nn
F nr , 2l,

(n n )2

n n

n + n

!2

(12.168)

4nn

,
F nr 2, nr , 2l,
(n n )2

with F (a, b, c; z) being hypergeometric functions. The angular matrix elements of


the unit vector in (12.166),
hl m |
x|lmi

xYlm (, ).
d
x Yl m (, )

(12.169)

are easily found, since x3 = cos and the spherical harmonics satisfy the recursion
relation
cos Ylm (, ) =

v
u
u
t

v
u

u
(l + 1)2 m2
l2 m2
Yl+1 m (, ) + t
Yl1 m (, ).
(2l + 2)(2l + 1)
(2l + 1)(2l 1)

(12.170)

Using the orthonormality relation


Z

d
x Yl m (, )Ylm(, ) = l l m m ,

(12.171)

we obtain immediately the angular matrix elements of x3 :


hl+1 m|
x3 |lmi =

v
u
u
t

v
u

u
(l+1)2 m2
l2 m2
, hl1 m |
x3 |lmi = t
, (12.172)
(2l+2)(2l+1)
(2l+1)(2l1)

with all others vanishing. The matrix elements of x1 and x2 are found with the help
of the commutation rule
i , xj ] = iijk xk ,
[L

(12.173)

which states that xi is a vector operator. As a consequence, the matrix elements


satisfy the Wigner-Eckart theorem,
hl m 1|
xM |lmi = hl m |1M; lmi
xl l ,

(12.174)

809

12.8 Rutherford Scattering

where hl m |1m ; lmi are Clebsch-Gordan coefficients (see Appendix 4C) and xm
[recall the definition in Eq. (4.865)]:
are the spherical components of x

(12.175)
x3 = cos , x = (
x2 x2 )/ 2 = sin ei / 2.
Explicitly:
hl + 1 m 1|
x |lmi =
hl 1 m 1|
x |lmi =

12.8

v
u
u (l m + 2)(l m + 1)
t
,

2(2l + 3)(2l + 1)

v
u
u (l m)(l m 1)
t
.

2(2l + 1)(2l 1)

(12.176)

Rutherford Scattering

The scattering of electrons on the Coulomb potential of nuclei of charge Ze


VC (r) =

Z
ZE 2
=
4r
r

(12.177)

was the first atomic collision observed experimentally by Rutherford.


The associated scattering cross section can easily be calculated in an estimated
classical approximation.

12.8.1

Classical Cross Section

In a Coulomb potential the electronic orbits are hyperbola. If an incoming electron runs along the z-direction and is deflected by a scattering angle towards the
x direction (see Fig. 12.5) the nucleus has the coordinates
(xF , zF ) = (b, a)

(12.178)

where

= tan ,
= cos ,
(12.179)
b
2
d
2
The parameter d is equal to a, where > 1 is the excentricity of the hyperbola.
The distance of closest approach to the nucleus is
rc = d a.

(12.180)

It is determined by the conservation of the nonrelativistic energy:


E nr =

p2
p2
Z
p2
,
=
= c
2M
2M
2M
rc

(12.181)

810

12 Quantum Electrodynamics

Figure 12.5 Kinematics of Rutherford cross section.

where pc is the momentum at the point of closest approach. This momentum is


determined by the conservation of orbital angular momentum:
l = pb = p b = pc rc .

(12.182)

From these equations we find


b=

Z
.
2E tan(/2)

(12.183)

The quantity b is called the impact parameter of the scattering process. It is the
closest distance which an particle would have from the nucleus if it were not deflected
at all. Particles which would pass through a thin annular ring with the radii b and
b + db are, in fact, scattered into a solid angle d given by
db =

1 Z
1
d.
3
4 4E sin (/2) cos(/2)

(12.184)

The current density of a single randomly incoming electron is j = v/V . It would


pass through the annular ring, with a probability per unit time
dP = j 2bdb.

(12.185)

With this probability it winds up in the solid angle d. Inserting Eq. (12.184), we
find the differential cross section [recall the definition Eq. (9.240)]


dP
db
1
Z 2
d
=
= 2b
=
.
d
d
d
4 sin4 (/2) 2E

(12.186)

811

12.8 Rutherford Scattering

12.8.2

Quantum-Mechanical Born Approximation

Somewhat surprisingly, the same result is obtained in quantum mechanics within


the Born approximation. According to Eqs. (9.242) and (??), the differential cross
section is
M 2V 2
d

|Vp p |2 ,
(12.187)
d
(2)2
where
Vp p

1
=
V

d3 x ei(p p)x

Z
4 Z
=
.
r
V |q2 |2

(12.188)

The quantity
q2 |p p|2 = 2p2 (1 cos ) = 4p2 sin2 (/2) = 8EM sin2 (/2)

(12.189)

is the momentum transfer of the process. Inserting this into (12.188), the differential
cross section (12.187) coincides indeed with the classical expression (12.186).

12.8.3

Relativistic Born Approximation

Let us now see how the above cross section formula is modified in a relativistic
calculation involving Dirac electrons. The scattering amplitude is, according to
Eq. (10.130),
Z

Sfi = iehp , s3 | d4 x (x)


(x)|p, s3 iA (x),
(12.190)
where A (x) has only the time-like component
A0 (x) =

Ze
=
4r

d3 q iqx Ze
e
.
(2)3
|q|2

(12.191)

The time-ordering operator has been dropped in (12.190) since there are no operators
at different times to be ordered in first-order perturbation theory. By evaluating the
matrix element of the current and performing the spacetime integral we obtain
Sfi = i2(E E)

M2
Ze2

0
u

(p
,
s
)
u(p,
s
)
,
3
3
V 2E E
|q|2

(12.192)

where E and E are the initial and final energies of the electron, which are in fact
equal in this elastic scattering process.
Comparing this with (9.130) we identify the R-matrix elements
Rp p =

M
Ze2
u(p , s3 ) 0 u(p, s3) 2 ,
VE
|q|

(12.193)

This is to be compared with the nonrelativistic Barn approximation where (12.188)


corresponds via (??) to
1 Ze2
.
(12.194)
Rp p =
V |q|2

812

12 Quantum Electrodynamics

The relativistic amplitude receives therefore a correction factor


C=

M
u(p , s3 ) 0 u(p, s3 ).
E

(12.195)

The absolute square of this multiplies the differential cross section. Apart from this,
the relativistic cross section contains an extra kinematic factor E 2 /M 2 accounting for
the different phase space of a relativistic electron with respect to the nonrelativistic
one [the ratio between (9.245) and (9.242)]. The differential cross section receives
therefore a total relativistic correction factor
E2 2
C = |
u(p , s3 ) 0 u(p, s3 )|2 .
2
M

(12.196)

If we consider the scattering of unpolarized electrons and do not observe the final
spin polarizations, this factor has to be summed over s3 and averaged over s3 , and
the correction factor is
E2 2 1 X
C =
u(p , s3 ) 0 u(p, s3 )
u(p, s3 ) 0 u(p , s3 ).
M2
2 s ,s3

(12.197)

To write the absolute square in this form we have used the general identity in the
spinor space, valid for any 4 4 spinor matrix M:
u ,
(
u Mu) = uM

(12.198)

where the operation bar is defined for a spinor matrix in complete analogy to the
corresponding operation for a spinor:
0 M 01 .
M

(12.199)

The Dirac matrices themselves satisfy


= .

(12.200)

The bar operation has typical properties of adjoining operations. If it is applied


to a product of matrices, the order is reversed:
M1 Mn = M n M 1 .

(12.201)

We use now the semi-completeness relation (4.696) for the u-spinors and rewrite
(12.197) as
!
E2 2 1
/ + M 0 p/ + M
0p
.
(12.202)
C = tr

M2
2
2M
2M
The trace over product of gamma matrices occurring in this expression is typical
for quantum electrodynamic calculations. Its evaluation is somewhat tedious, but
follows a few quite simple algebraic rules.

813

12.8 Rutherford Scattering

The first states that a trace containing an odd number of gamma matrices vanishes. This is a simple consequence of the fact that , and thus a product of any
odd number of gamma matrices, changes sign under the similarity transformation
5 5 = , whereas the trace is invariant under any similarity transformation.
The second rule governs the evaluation of a trace containing an even number of
gamma matrices. It is a recursive rule which makes essential use of the invariance
of the trace under cyclic permutations
tr( 1 2 3 n1 n ) = tr( 2 3 n1 n 1 ),

(12.203)

and leads to an explicit formula which is a close analog of Wicks expansion formula
for a time ordered product of fermion field operators. For this purpose, we define a
pair contraction between a/ and b/ as
1
a b/ ) = ab.
a
/b/ tr(/
4

(12.204)

Then we may move the first gamma matrix in a trace such as


1
tr(/
a 1 a/ n )
4

(12.205)

step by step to the end, using the anticommutation rules between gamma matrices
(4.560), which imply
a/1 a/ i = /
ai a/1 + 2a1 ai = /
ai a/1 + 2 a
/1 a
/i

(12.206)

Having arrived at the end, it can be taken back to the front using the cyclic invariance
of the trace. This produces once more the initial trace, except for a minus sign, thus
doubling the initial trace on the left-hand side of the equation. In this way, we find
the recursion relation
1
1
1
tr(/
a1 a/2 a/3 a/n1 a/n ) = tr(/a1 a
/2 + tr(/a1 a
/2 a
/3 a/n1 a/n )
4
4
4
1
1
. . . + tr(/a1a
/2 a
/3 /an1 a/n ) + . . . + tr(/a1 a
/2 a
/3 a
/n1a
/n ), (12.207)
4
4
whose explict form is
1
1
1
tr(/
a1 a/2 a/3 a/n1 a/n )=(a1 a2 ) tr(/
a3 a/4 a/n1 a/n ) + (a1 a3 ) tr(/
a2 a/4 a/n1 a/n )
4
4
4
1
1
. . . + (a1 an1 ) tr(/
a2 a/3 a/n2 a/n ) + (a1 an ) tr(/
a2 a/3 a/n1 a/n ).(12.208)
4
4
The contractions within the traces are defined as in (12.204), but with a minus sign
for each permutation necessary to bring the Dirac matrices to adjacent positions.
By repeating this process, we arrive at the expansion rule
X
1
tr(/
a1 a/n ) =
()P (ap(1) ap(2) )(ap(3) ap(4) ) (ap(n1) ap(n) ). (12.209)
4
pair contractions

814

12 Quantum Electrodynamics

where P is the number of permutations to adjacent positions.


The derivation of the rule is completely parallel to that of the thermodynamic
version of Wicks rule in Section 4.14, whose Eqs. (7.835) and (7.836) can directly
be translated into anticommutation rules between gamma matrices and the cyclic
invariance of their traces, respectively.
Another set of useful rules following from (12.206) which will be needed later is
a/

a/
a/ b/
a/ b/ b/

=
=
=
=
=

a/ + 2a ,
4,
2/
a,
2/
a,
2/
c b/ a/ .

(12.210)
(12.211)
(12.212)
(12.213)
(12.214)

With the Wick rule (12.245) for -matrices we now calculate the expression
(12.202) as f
E2 2
1
1
C =
tr( 0 p/ 0 p/ ) + M 2 tr( 02 )
2
2
M
2M 4

1  0 0

2
.
2p
p

pp
+
M
=
2M 2
Inserting p0 = p0 = E and, in the center-of-mass frame,

pp = E 2 |p|2 cos = M 2 + 2|p|2 sin2 ,


2
and the total cross section in the center-of-mass frame
Z 2 2 M 2
E2 2
d
=
C .

dCM
4|p|4 sin4 (/2) M 2


(12.215)

(12.216)

(12.217)

This relativistic version of Rutherfords formula is known as Motts formula.


In terms of the incident electron velocity, the total modification factor reads
E2 2
1
v
1
C =
2
2
M
1 (v/c)
c
"

 2

.
sin2
2

(12.218)

It is easy to verify that the same differential cross section is valid for positrons. In
the nonrelativistic case, this follows directly from the invariance of (12.187) under
e e. But also the relativistic correction factor remains the same under the
interchange of electrons and positrons, where (12.197) becomes
E2 2 1 X
C =
v(p , s3 ) 0 v(p, s3 )
v (p, s3 ) 0 v(p , s3 ).
M2
2 s ,s3

(12.219)

Inserting the semi-completeness relation (4.697) for the spinors v(p, s3 ), this becomes
/ M 0 p/ M
E2 2 1
0p
C = tr

M2
2
2M
2M

(12.220)

which is the same as (12.197) since only traces of an even number of gamma matrices
contribute.

815

12.9 Compton Scattering

12.9

Compton Scattering

A simple scattering process whose cross section can be calculated to a good accuracy
by means of the above diagrammatic rules is photon-electron scattering, also referred
to as Compton scattering. It gives an important contribution to the blue color of
the sky.
Consider now a beam of photons with four-momentum ki and polarization i
impinging upon an electron target of four-momentum pi and spin orientation i .
The two particles leave the scattering regime with four-momenta kf , pf , and spin
indices f , f , respectively. Adapting formula (10.92) for the scattering amplitude
to this situation we have
Sfi = S (p , s3 ; k , h |k, h; p, s3)
R
i
dtV (t)
h0|a(p , s3 )a(k , h )T e I a (k, h)a (p, s3 )|0i
R

i
dtV (t)
h0|e I |0i
SN (p , s3 ; k , h |k, h; p, s3 ) /Z[0].
with
i

dt VI (t)

= eie

(x)A (x)

d4 x (x)

(12.221)

(12.222)

Expanding the exponential in powers of e, we see that the lowest-order contribution


to the scattering amplitude comes from the second-order term which gives rise to the
two Feynman diagrams shown in Fig. 12.6. In the first, the electron s1 of momentum
p absorbs a photon of momentum k, and emits a second photon of momentum k ,

Figure 12.6 Lowest-order Feynman Diagrams contributing to Compton Scattering and


giving rise to the Klein-Nishina formula.

to arrive in the final state of momentum p . In the second diagram, the acts of
emission and absorption have the reversed order. Before we calculate the scattering
cross section associated with these Feynman diagrams, let us first recall the classical
result.

12.9.1

Classical Result

Classically, the above process is described as follows. A target electron is shaken


by an incoming electromagnetic field. The acceleration of the electron causes an

816

12 Quantum Electrodynamics

emission of antenna radiation. For a weak and slowly oscillating electromagnetic


field of amplitude, the electron is shaken nonrelativistically and moves with an
instantaneous acceleration
e
e
=
x
E=
 E0 eit+ikx,
(12.223)
M
M
where is the frequency and E0 the amplitude of the incoming field.
The acceleration of the charge gives rise to antenna radiation following Larmors
formula. Inserting (12.223) into (5.25), and averaging over the temporal oscillations,
the radiated power per unit solid angle is
1
dE
=
d
2

e2
4M

!2

E02 sin2 ,

(12.224)

where is the angle between the direction of polarization of the incident light and
the direction of the emitted light.
For a later comparison with quantum electrodynamic calculations we associate
this emitted power with a differential cross section of the electron with respect to
light. According to the definition in Chapter 6, a cross section is obtained by dividing
the radiated power per unit solid angle by the incident power flux density cE20 /2.
This yields
!2
d
e2
=
sin2 = re2 sin2 ,
(12.225)
2
d
4Mc
where

e2
h

=
2.82 1013 cm
(12.226)
2
4Mc
Mc
is the classical electron radius. Formula (12.225) describes the so-called Thomson
scattering cross section. Formula (12.225) holds for incident light which is linearly
polarized. For unpolarized light, we have take the average between the cross section
(12.225) and another one in which the plane of polarization is rotated by 90%.
Suppose the incident light runs along the z-axis, and the emitted light along the
= (sin cos , sin sin , cos ). For a polarization direction  = x
along
direction k
the x-axis, the angle is found from
re =

)2 = cos2 + sin2 sin2 .


sin2 = (k

(12.227)

)2 = cos2 + sin2 cos2 .


sin2 = (k

(12.228)

along the y-axis, it is


For a polarization direction  = x

The average is

1
1 2
sin = (1 + cos2 ).
(12.229)
2
2
Integrating this over all solid angles yields the Thomson cross section for unpolarized
light
8 2
tot =
r .
(12.230)
3 e
sin2 = cos2 +

817

12.9 Compton Scattering

12.9.2

Klein-Nishina Formula

The scattering amplitude corresponding to the two Feynman diagrams in Fig. 12.6
is obtained by expanding formula (10.92) up to second order in e:
2

Sfi = e

(y)(x)/
(x)a |0i,
d4 xd4 yh0|ap ,s3 T (y)/
p,s3

(12.231)

and performing the Wick contractions of Section 7.8:


(y)(x)/
(x)a |0i
h0|ap ,s3 T (y)/
p,s3

(12.232)

/ (y)(x)
(y)(x)/
(x)a |0i|0i.
/ (x)a |0i+h0|T ap ,s (y)/
= h0|T ap ,s3 (y)
p,s3
p,s3
3
After Fourier-expanding the intermediate electron propagator,
G0 (y, x) =

d4 pi ipi (yx)
i
e
,
4
i
(2)
p/ M

(12.233)

we find
d 4 pi
M

4
(2) V 2 EE 22
"
i

i
i
/u(p, s3 )
ei(p p )y(pp )x u(p , s3 )/
i
p/ M
#
i

i(p pi )x(ppi )y

/ u(p, s3 ) .
+e
u(p , s3 )/
i
p/ M

Sfi = e2

d4 xd4 yek ykx

(12.234)

One of the spatial integrals fixes the intermediate momentum in accordance with
energy-momentum conservation, the other yields a (4) -function for overall energymomentum conservation. The result is
Sfi = i(2)4 (4) (p + k p k)e2

M
V

2 EE 22

u(p , s3 )Hu(p, s3), (12.235)

where H is the 4 4-matrix in spinor space


(/
p k/ ) + M
(/
p + k/ ) + M
/ + /
/ .
H /
(p + k)2 M 2
(p k )2 M 2

(12.236)

We have written (k, h), k as , , and (k , h ), k as , , respectively, with a


similar simplification for E and E . The second term of the matrix H arises from
the first by the crossing symmetry
,

k k .

(12.237)

Simplifications arise from the properties (12.238) It can, moreover, be simplified by


recalling that external electrons and photons are on their mass shell, so that
p2 = p2 = M 2 ,

k 2 = k 2 = 0,

(12.238)

818

12 Quantum Electrodynamics

k = k = 0.

(12.239)

A further simplification arises by working in the laboratory frame in which the initial
electron is at rest, p = (M, 0, 0, 0). Also, we choose a gauge in which the polarization
vectors have only spatial components. Then
p = p = 0,

(12.240)

since p has only a temporal component and only space components. We also
use the fact that H stands between spinors which satisfy the Dirac equation (/
p
M)u(p, s3 ) = 0, u(p, s3 )(/
p M) = 0. Further we use the commutation rules (4.560)
for the gamma matrices to write, as in (12.206),
p/ / = /
p/ + 2p.

(12.241)

The second term vanishes by virtue of Eq. (12.240). Similarly, we see that p/ anticommutes with / . Using these results, we may eliminate the terms p/ + M occurring
in M. Finally, using Eq. (12.239), we obtain

u(p

, s3 )Hu(p, s3)

= u(p

, s3 )

k/
k/
/ + /
/ u(p, s3 ). (12.242)

2pk
2pk
)

To obtain the transition probability, we must take the absolute square of this. If
we do not observe initial and final spins, we may average over the initial and sum
over the final spin components which produces a factor 1/2 times the double sum:
so that we and form
F =

s3 ,s3

|
u(p , s3 )Hu(p, s3)|2 =

u(p , s3 )Hu(p, s3)


u(p, s3 )Hu(p, s3 ). (12.243)

s3 ,s3

Now we use the semi-completeness relation (4.696) for the spinors to rewrite (12.243)
as
F =

1
tr [(/
p + M)H(/
p + M)H] .
4M 2

(12.244)

The trace over product of gamma matrices can be evaluated according to the Wicktype of rules explained on p. 812.
X
1
tr(/
a 1 a/ n ) =
()P a/ 1 a/ n .
4
pair contractions

(12.245)

After some lengthy algebra (see Appendix 9A), we find


1

F =
+ 2 + 4| |2 ,
2
2M

"

(12.246)

This holds for specific polarizations of the incoming and outgoing photons. If we sum
over all final polarizations and average over all initial ones, we find (see Fig. 12.7)

819

12.9 Compton Scattering

Figure 12.7 Illustration of the photon polarization sum h,h | |2 = 1 + cos2 in
Compton scattering in the laboratory frame. Incoming and outgoing photon momenta
with scattering angle are shown in the scattering plane, together with their transverse
polarization vectors.
P

1 X 2 1
| | = (1 + cos2 ),
2 h,h
2

(12.247)

This can also be found more formally using the transverse completeness relation
(4.324) of the polariztion vectors:
1 X 2 1 X i i
1 ij k i k j
2
| | =
(k , h ) (k, h)j (k, h)i (k , h ) =
2 h,h
2 h,h
2
k

k j k i
2
k

(k k )2
1
1
= 1 + 2 2 = (1 + cos2 ),
2
k k
2
"

ji

(12.248)

The average value of F is therefore


1X 1

F =
+ 2 + 4| | = 2
+ sin2 .
2
2 h,h 2M

"

"

(12.249)

We are now ready to calculate the transition rate, for which we obtain from
Eq. (9.296)
dP
= V (2)4
dt

d 3 p V
(2)3

d3 k V (4)
(p p)|tfi |2 ,
(2)3

(12.250)

with the squared t-matrix elements


1
M2
1
F.
(12.251)
4

V EE 22 2
The spatial part of the -function removes the momentum integral over p . The
temporal part of the -function enforces energy conservation. This is incorporated
into the momentum integral over k as follows. We set Ef = p0 + k0 , and write
|tfi |2 = e4

d3 k = k 2 dk d = 2

d
ddEf ,
dEf

(12.252)

where is the solid angle into which the photon has been scattered. Then (12.250)
becomes
dP
V2
= V e4 (2)4
dtd
(2)6

d
1 M2 1

F.
dEf Ef =Ei 4V 4 EE 2

(12.253)

820

12 Quantum Electrodynamics

For an explicit derivative d /dEf , we go to the laboratory frame and express the
final energy as
q
q

Ef = + p2 + M 2 = + (k p )2 + M 2 = + 2 2 cos + 2 + M 2 ,
(12.254)
where is the scattering angle in the laboratory. Hence:
cos
dEf
=
1
+
d
E

(12.255)

By equating Ei = M + with Ef , we derive the Compton relation


(1 cos ) = M( ),
or
=
and therefore

,
1 + (1 cos )/M

(12.256)
(12.257)

dEf
cos
E + cos
M + cos
M
= 1+
=
=
= . (12.258)

d
E
E
E
E
Since E = M in the laboratory frame, Eq. (12.253) yields the differential probability
rate
V2
dP
= V e4 (2)4
V
dtd
(2)6

!2

1 1
F.
4V 4 2

(12.259)

To find the differential cross section, this has to be divided by the incoming
particle current density. According to Eq. (9.313), this is given by
j=

v
,
V

(12.260)

where v is the velocity of the incoming particles. The incoming photons move with
light velocity, so that (in natural units with c = 1)
j=

1
.
V

(12.261)

This leaves us with the Klein-Nishina formula for the differential cross section

d
2
=
d

!2

1
F.
2

(12.262)

In the nonrelativistic limit where << M, the Compton relation (12.257) shows
that , and Eq. (12.249) reduces to
F

1
(1 + cos2 ).
M2

(12.263)

821

12.10 Electron-Positron Annihilation

Expressed in terms of the classical electron radius r0 = /M, the differential scattering cross section becomes, expressed in terms of the classical electron radius where
r0 = /M,
1
d
r0 2 (1 + cos2 ),
(12.264)
d
2
which is the Thomson formula for the scattering of low energy radiation by a static
charge. To find the total Thomson cross section, we must integrate (12.264) over all
solid angles:
Z
Z 1
d
1
= d
2
(12.265)
d cos r02 (1 + cos2 ),
d
2
1
which yields
8
Thomson r0 2 .
(12.266)
3
Let us also calculate the total cross section for relativistic scattering, integrating
(12.262) over all solid angles:
=

Z 1

d
2 1
= 2
+ sin2 .
d cos
d
2
d
2M

1
"

Inserting
=

(12.267)

,
1 + (1 cos )/M

(12.268)

which follows from (12.257), the angular integral yields


= Thomson f (),

(12.269)

where f () contains the relativistic corrections to Thomsons cross section:


f () =

"

log(1 + 2)
1 + 3
1 + 2(1 + )
. (12.270)
log(1 + 2) +

1 + 2
2
(1 + 2)2

For small :
f () = 1 2 + O( 2 ).

(12.271)

For large , f () decreases like


f () =

12.10

3
[log 2 + O(log /)] .
8

(12.272)

Electron-Positron Annihilation

The Feynman diagram in Fig. 12.6 can also be read from bottom to top in which
case it describes the annihilation of process corresponding to the two Feynman
diagrams in Fig. 12.6. In order use as much as possible the previous results, we
redraw the Feynman diagrams as shown in Fig. 12.9. Instead of an outgoing electron

822

12 Quantum Electrodynamics

Figure 12.8 Ratio between total relativistic Compton cross section and nonrelativistic
Thomson cross section.

Figure 12.9 Lowest-Order Feynman Diagrams contributing to Electron-Positron Annihilation. It arises from the Compton diagram by the crossing operation p p , k k,
.

with momentum p we let a negative-energy electron go out with momentum p .


Instead of an incoming photon with momentum k we let a negative-energy photon
go out with momentum k. The former are represented by spinors v(p , s3 ), which
are negative energy solutions of the Dirac equation with inverted momenta and spin
directions. The S-matrix element to lowest order is from a slightly modified (12.231)
as
Sfi = e2

(y)(x)/
(x)b ap,s |0i,
d4 xd4 yh0|T(y)/
3
p ,s
3

(12.273)

and performing the Wick contractions as in (12.233) we obtain


d 4 pi
M

(2)4 V 2 EE 22
"
i

i
i
/ u(p, s3 )
ei(p p )y(pp )x v(p , s3 )/
i
p/ M
#
i
i(p pi )x(ppi )y

+e
v(p , s3 )/
i
/ u(p, s3) .
p/ M

Sfi = e2

d4 xd4 yek y+kx

(12.274)

823

12.10 Electron-Positron Annihilation

Note that this arises from the Compton expression (12.234) by the crossing operation
p p , u(p, s3 ) v(p, s3 ),

k k, (k, h) (k, h).

(12.275)

As before, one of the spatial integrals fixes the intermediate momentum in accordance with energy-momentum conservation, the other yields a (4) -function for overall energy-momentum conservation. The result is
Sfi = i(2)4 (4) (k + k p p )e2

M
V

2 EE 22

v(p , s3 )Hu(p, s3 ), (12.276)

where H is the 4 4-matrix in spinor space


H /

p k/ ) + M
(/
p k/ ) + M
(/

/
+

/
/ .
(p k)2 M 2
(p k )2 M 2

(12.277)

As before, we have written (k, h), k as , , and (k , h ), k as , , respectively,


with a similar simplification for E and E . The second term of the matrix H arises
from the first by the Bose symmetry
,

k k.

(12.278)

Simplifications arise from the mass shell properties (12.238), the gauge conditions
(12.239), and the other relations (12.240), (12.241). We also work again in the
laboratory frame in which the initial electron is at rest,
p = (M, 0, 0, 0) and the
positron comes in with momentum p and energy E = p2 + M 2 . This leads to

v(p

, s3 )Hu(p, s3)

= v(p

, s3 )

k/
k/

/ + /
/ u(p, s3 ).
2pk
2pk
)

(12.279)

To obtain the transition probability, we must take the absolute square of this.
If we do not observe initial spins, we may average over the initial spin components
which gives a factor 1/4 times the spin sum
F =

s3 ,s3

|
v (p , s3 )Hu(p, s3 )|2 =

v(p , s3 )Hu(p, s3)


u(p, s3 )Hu(p, s3 ). (12.280)

s3 ,s3

Now we use the semi-completeness relations (4.696) and (4.697) for the spinors to
rewrite (12.280) as
F =

1
tr [(/
p + M)H(/
p + M)H] .
4M 2

(12.281)

The trace over product of gamma matrices is now evaluated as before and we obtain
for specific polarizations of the two outgoing photons. almost the same expression
as before in Eq. (12.249) (see Appendix 9A):
1

F =
+ + 2 4|  |2 ,
2
2M

"

(12.282)

824

12 Quantum Electrodynamics

This result can be deduced directly from the previous polarization sum (12.246) by
the crossing operation (12.275), apart from an overall minus sign, whose origin is
the negative sign in front of (12.281).
We are now ready to calculate the transition rate, for which we obtain from
Eq. (9.296)
dP
= V (2)4
dt

d3 k V
(2)3

d3 k V (4)
(p p)|tfi |2 ,
(2)3

(12.283)

with the squared t-matrix elements


|tfi |2 = e4

1
M2
1
F.
V 4 EE 22 4

(12.284)

The spatial part of the -function removes the momentum integral over k . The
temporal part of the -function enforces energy conservation. This is incorporated
into the momentum integral over k as follows. We set Ef = + , and write
d3 k = k 2 dkd = 2

d
ddEf ,
dEf

(12.285)

where is the solid angle into which the photon with momentum k emerges. Then
(12.283) becomes
dP
V2
= V e4 (2)4
dtd
(2)6

1 M2 1
d

F.
dEf Ef =Ei 4V 4 EE 4

(12.286)

To calculate d/dEf explicitly, we express the final energy in the laboratory frame
as
q
q

(12.287)
Ef = + k2 = + (p k)2 = + p2 2|p| cos + 2 ,
where is the scattering angle in the laboratory. Hence:


1 
dEf
1 d  2

2
p 2|p| cos + 2 =
=1 +

2|p
|
cos

d
2 d

 


q
1
k(p + p ) M(M + E )
2 +M 2 2|p | cos =
=
p

M
+
=
. (12.288)


We must now divide (12.286) by the incoming positron current density j = |p |/E V
[recall (9.313)], and find the differential cross section
d
2

d
|p |(M + E ) M


where

2

1 2
M F,
4

M + E |p | cos
M
=
,

M + E

(12.289)

(12.290)

825

12.10 Electron-Positron Annihilation

2
Figure 12.10 Illustration of the photon polarization sum
h,h | | in electronpositron annihilation in the laboratory frame. Incoming positron and outgoing photon
momenta with scattering angles and are shown in the scattering plane, together with
their transverse polarization vectors.

as follows from equating the right-hand sides of


1
1
kk = (k + k )2 = (p + p )2 = M(M + E ),
2
2

(12.291)

and
kk = k(p + p k) = (M + E |p | cos ).

(12.292)

If the incoming positron is very slow, then M, and the two photons
share equally the rest energy of electron and positron. We can now substitute in F
(12.282)

E |p |cos
=
,
(12.293)

M
and sum over all photon polarizations we obtain [see Figs. 12.10 and 12.248]
X

h,h

| |2 = 1 +

M
M
(k k )2
= 1 + cos2 ( + ) = 1

2
2
k k

2

(12.294)

The last expression is found by observing that


k k = [1 cos( + )],
2

(12.295)

0.9
0.8
0.7
0.6

/ lowenergy

0.5

Figure 12.11 Electron-positron annihilation cross section divided by its low-energy limit
p
(12.300) as a function of = 1/ 1 v 2 /c2 of the incoming positron in the laboratory
frame.

826

12 Quantum Electrodynamics

such that we can express


1 (k + k )2
1 (p + p )2
M(M + E )
k k
=
1

=
1

=
1

.

2
2

Energy conservation leads to

cos( + ) = 1

k k
M( + )
=
1

,
(12.296)


thus proving the right-hand side of (12.294). Note that from (12.295) and (12.296)
we find a relation
k k = M( + ).
(12.297)
cos( + ) = 1

We now introduce the relativistic factor of the incoming positron E /M and


write the relation (12.290) as
M
=1

1
cos ,
+1

1
cos
M
+1

=
.

2 1 cos
1

(12.298)

The result is now integrated over all solid angles, and divided by 2 to accound for
Bose statistics of the two final photons. This yields the cross section expressed in
terms of the classical electron radius r0 = /M:
1
=
2

q
r02 2 + 4 + 1
d
+3
2 1
=
log

+
d

d
1+
2 1
2 1
"

!#

(12.299)

For small incoming positron energy, the cross section diverges like
r 2 c
r 2
0 ,

20
v
1
|p|0

(12.300)

whereas in the high energy limit

|p|

r02
[log(2) 1] .

(12.301)

The detailed behavior is shown in Fig. 12.11.


The above result can be used to estimate the lifetime of a positron moving
through matter. We simply have to multiply the cross section by the incident
current density j = |p|/EV of a single positron and by the number N of target
electrons, which is Z per atom. For slow positrons we may use Eq. (12.300) to find
the decay rate
N
P = r02 v Z .
(12.302)
V
For lead, this yields a lifetime = 1/P 1010 s.
The formula (12.305) is actually not very precise, since the incoming positron
wave is strongly distorted towards the electon by the Coulomb attraction. This
is the so-called initial-state interaction. It is closely related to the problem to be
discussed in the next section.

827

12.11 Positronium Decay

12.11

Positronium Decay

The previous result can be used to calculate the lifetime of positronium. Since the
momenta in positronium are nonrelativistic, the annihilation cross section (12.300)
is relevant. The wave function of positronium at rest in an s-wave is approximately
given by [recall (7.301)]
|

S,S3

i=

d3 p
a3
8

|0ihS,
S
|s,
s
;
s,
s
i,
(p)
=
(p)a
b
, (12.303)
3
3
3
p,s3 p,s3
(2)3
(1 + a2 |p|2 )2

where S = 1 for ortho- and S = 0 for para-positronium, and a is the Bohr radius
of positronium, which is twice the Bohr radius of hydrogen: a = 2 aH = 2/M, the
factor 2 being due to the reduced mass being half the electron mass [recall the
general formula (6.133)]. This amplitude by be pictured by Feynman diagrams as
shown in in Fig. 12.12.

Figure 12.12 Lowest-Order Feynman diagrams contributing to decay of the spin singlet
parapositronium, the ground state.

In Eq. (7.300) we have calculated the charge conjugation parity C of these


states to be 1, respectively. Since a photon is odd under charge conjugation [recall
(7.540)], ortho-positronium can only decay into three photons (one is forbidden
by energy-momentum conservation). Thus only para-positronium decays into two
photons and the cross section calculated in the last section arises only from the spin
singlet contribution of the initial state. Since it was obtained from the average of a
total of four states, three of which do not decay at all, the decay rate of the singlet
state is four times as big as calculated prevously from (12.300). The integral over all
momenta in the amplitude (12.273) can be factored out since the low-energy decay
rate is approximately indedendent of v and hence of p. Thus we obtain directly the
decay rate [10]
para2 P para =
=


4r02 c

Mc2 5
.
h
2

d3 p
1
2
2
2
=
4r
c
|(0)|
=
4r
c
(p)

0
0

(2)3
a3

(12.304)

Using the electron energy Mc2 0.510 MeV and the Planck constant h
6.682
16
2
22
10 eV s, the ratio h
/Mc is equal to 1.288 10 s. Together with the factor 2/5
this leads to a lifetime = 1/ 0.13 109 s.

828

12 Quantum Electrodynamics

The decay of the spin triplet orthopositronium state proceeds at a roughly thousand times slower rate
Mc2 6 4( 2 9)
,
orth3 P orth =
h
2
9

(12.305)

with a lifetime of 140 109 s.

Figure 12.13 Lowest-Order Feynman diagrams contributing to decay of the spin triplet
orthopositronium, the fist excited state which lies 203.5 GHz above the ground state.

Decays into 4 and 5 photons have also been calculated and measured experimentally. The theoretical rates are [11]

Mc2 7
114.5 +O(2 ) 1.43 106 para2 . (12.306)
h

2
0.0189 orth3 0.959 106 orth3 .
(12.307)


para4 0.0138957
orth5

Experimentally, the branching ratios are [12] 1.14(33) 106 and 1.67(99) 106,
in reasonable agreement with the theoretical numbers. The validity of C-invariance
has also been tested by looking for the forbidden decays of parapositronium into
an odd and orthopositronium into an even number of photons. So far, there is no
indication of C-violation.

12.12

Bremsstrahlung

If a charged particle is de- or accelerated, it emits an electromagnetic radiation called


Bremsstrahlung. This is a well-known process in classical electrodynamics and we
would like to find the quantum field theoretic generalization of it. First, however,
we shall recapitulate the classical case.

12.12.1

Classical Bremsstrahlung

Consider a trajectory in which a particle changes its momentum abruptly from p to


p [see Fig. 12.14].
The trajectory may be parametrized as:

x ( ) =

p/M
p /M

for

< 0,
> 0,

(12.308)

829

12.12 Bremsstrahlung

Figure 12.14 Trajectories in the simplest classical Bremsstrahlung process: An electron


changing abruptly its momentum

where is the proper time. The electromagnetic current associated with this trajectory is

dx ( ) (4)
(x x( ))
d
Z
h
i
e
=
d ( )p (4) (x p/M) + ( )p (4) (x p /M) (. 12.309)
M

j (x) = e

After a Fourier decomposition of the -functions, this can be written as

j (x) =

d4 k ikx
e
j (k),
(2)4

(12.310)

with the Fourier components


p
p
j (k) = ie
.

pk p k
!

(12.311)

The vector potential associated with this current is found by solving the Maxwell
equation Eq. (12.50). Under the initial condition that at large negative time, the
vector potential describes the retarded Coulomb potential of the incident particle,
we obtain
Z
A (x) = i dx GR (x x )j (x )
(12.312)
where GR (x x ) is the retarded Green function defined in Eq. (7.154). At very
large times the particle has again a Coulomb field associated with it which can be
found by using the advanced Green function Eq. (7.160):
Aout (x)

=i

dx GA (x x )j (x ).

(12.313)

As the acceleration takes place, the particle emits radiation which is found from the
difference between the two fields,
Arad =

1
i

dx [GR (x, x ) GA (x, x )]j (x ).

(12.314)

830

12 Quantum Electrodynamics

Remembering the list (7.204) of Fourier transforms of the various Green functions,
we see that the Fourier components of the radiation field are given by
Arad (k) = i2(k 0 )(k 2 )j (k)

(12.315)

The energy of the electromagnetic field is by Eq. (7.417)


H=

1
d3 x (A A A A ).
2

(12.316)

A classical field which solves the field equations can be Fourier decomposed into
positive- and negative-frequency components as in (7.380),
A (x) =

X
k



1
eikx ak + eikx a
,
k
2V k

where
ak

3
X

(k, )ak,

(12.317)

(12.318)

=0

are classical Fourier components. Then the combination of time and space derivatives in (12.316) eliminates all terms of the form ak ak and ak ak , and we find
H=

X
k

(a
k ak ak ak ) g ,
2

(12.319)

just as in the calculation of the energy in (7.418). The radiation field (12.315)
corresponds to
ak = ij (k)|k0 =|k| .

(12.320)

Inserting everything into (12.319) we derive for large t an emitted energy


E =

X
d3 k
k0
j (k)j (k).
0
3
2k (2)
h=1,2

(12.321)

Inserting (12.311) we obtain the energy emitted into a momentum space element
d3 k:
"
#
2pp
M2
M2
1 d3 k 2
e

.
(12.322)
dE =
2 (2)3
(pk)(p k) (pk)2 (p k)2
where polarization vectors have vanishing zeroth components. Dividing out the
energy per photon k 0 , this can be interpreted as the probability of omitting a photon
into d3 k:
"
#
d3 k
M2
M2
2pp
2
dP = 0
.
(12.323)
e

2k (2)3
(pk)(p k) (pk)2 (p k)2
If we are interested in the polarization of the radiated electromagnetic field, we
we make use of the local current conservation law j which reads, in momentum
space,
k j = k0 j 0.
(12.324)

831

12.12 Bremsstrahlung

This allows us to rewrite


j (k)j (k) = |j 0 (k)|2 |j(k)|2 =
where
jTi (k)

1
|k j(k)|2 |j(k)|2 = |jT (k)|2
k2

ki kj
2
k
ij

j j (k)

(12.325)

(12.326)

is the transverse part of the current. In the second-quantized description of the


energy, the transverse projection is associated with a sum over the two outgoing
photon polarization vectors [recall (4.328)]:
ij

X
ki kj
=
i (k, h)i (k, h).
k2
h=1,2

(12.327)

The emitted energy (12.321) can therefore be resolved with respect to the polarization vectors as
Z
X
d3 k
0
k
E=
|(k, h) j(k)|2 .
(12.328)
0
3
2k (2)
h=1,2
This leads to an energy emitted into a momentum space element d3 k:
2

d3 k 2 X p
p
dE =
e


,
2(2)3 h=1,2 pk p k

(12.329)

and a corresponding probability of omitting a photon into d3 k analogous to (12.323).


Let us calculate the angular distribution of the emitted energy in Eq. (12.322).
Denote the direction of k by n:
k
.
(12.330)
n
|k|
Then we can write pk as

pk = E|k|(1 v n)

(12.331)

and find
d3 k e2
M2
M2
2(1 v v )
dE =
.

(2)3 2 |k|2 (1 v n)(1 v n) E 2 (1 v n)2 E 2 (1 v n)2


(12.332)
"

The radiation is peaked around the directions of the incoming and outgoing particle.

12.12.2

Bremsstrahlung in Mott Scattering

We now turn to the more realistic problem of an electron scattering on a nucleus.


Then the electron changes its momentum within a finite period of time rather than
abruptly. Still the Bremsstrahlung will be very similar to the previous case. We shall
consider immediately a Dirac electron, i.e., we study the Bremsstrahlung emitted by

832

12 Quantum Electrodynamics

Figure 12.15 Lowest-Order Feynman diagrams contributing to Bremsstrahlung. The


vertical photon line indicates the nuclear Coulomb potential.

Mott scattering. The lowest-order Feynman diagram by which this process occurs
is shown in Fig. 12.15. The vertical photon line indicates the nuclear Coulomb
potential,
Z
(12.333)
VC (x) =
4r
and the scattering amplitude is found from the Compton amplitude by simply interchanging the incoming photon field
q

eA (x) = k, eikx 2V k0
by the static vector potential
0 VC (r) = 0 Z
The scattering amplitude is therefore
Sfi = i

1
d4 q
2(q 0 )eiqx 2 .
4
(2)
|q|

(12.334)

1
4Ze
M

2(p0 + k p0 )
2
|q|
V 2 E E 2V k 0
"
#
1
1

0
0

u(p , s3 ) /

/ u(p, s3), (12.335)
p/ + k/ M
p/ k/ M

where
q = p + k p

(12.336)

is the spatial momentum transfer. The amplitude conserves only energy, not spatial
momentum. The latter can be transferred from the nucleus to the electron with any
amount. The unpolarized cross section following from Sfi is
1
1
d = M Z (4)
0

2k E E v
2

F
d 3 p d 3 k
2(p0 + k 0 p0 ) 2
6
(2)
|q|

(12.337)

where we have used the incoming particle current density v/V = p/EMV and set
1X
F
tr
2 h

"

/
p

/ k/ + M
+ k/ + M 0
0p

/
2p k
2pk

/
0p

+ k/ + M
p/ k/ + M 0
/ /

2p k
2pk

p/ + M
2M

p/ + M
2M

(12.338)

833

12.12 Bremsstrahlung

Introducing the solid angles of outgoing electrons and photons de and d , respectively, and dropping the primes on the emitted photon, so that and k become
and k. Setting the energy k 0 of the outgoing photon, equal to , the differential
cross section becomes
Z 2 3 M 2 |p |
d
=
F.
(12.339)
de d d
2 |q|4 |p|
The calculation of F is as tedious as that of the trace (12.281) in the Klein-Nishina
cross section. It can be done again with the help of the formulas in Appendix 9A.
Let us introduce the angles and between the outgoing photon momentum and
the initial and final electron momenta p and p , respectively, in addition the angle
Then
k
p

Figure 12.16 The angles , , in the Bethe-Heitler cross section formula


X
h

X
h

X
h

and

( p )2 = |p2 sin2 ,
( p)2 = |p|2 sin2 ,

( p )( p) = |p ||p| sin sin cos ,

p2 sin2
p2 sin2
1
2
2
(4E

q
)
+
(4E 2 q 2 )
F =
4 2 (E p cos )2
(E p cos )2
p2 sin2 + p2 sin2
+ 2 2
(E p cos )(E p cos )

p p sin sin cos

2
2
2
(4E
E

q
+
2
)
.
(E p cos )(E p cos )

(12.340)
(12.341)

(12.342)

With this function F , Eq. (12.339) is known as the Bethe-Heitler cross section
formula. For soft photon emission, 0, and the cross section has the limiting
form

2

d
d3 k p
d
p
2
e

(12.343)


de
de elastic
2(2)3 kp kp
the elastic cross section being simply multiplied by the cross section of the classical
Bremsstrahlung.

834

12 Quantum Electrodynamics

12.13

Electron-Electron Scattering

The leading Feynman diagrams are shown in Fig. 12.17. The associated scattering

Figure 12.17 Lowest-Order Feynman Diagrams contributing to Electron-Electron Scattering

amplitude is given by
Sf i = (2)4 4 (p1 + p2 p1 p2 )(ie)2
"
ig

u(p1 , 1 ) u(p1 , 1)
u(p2 , 2 ) u(p2 , 2 )
(p1 p1 )2
#
ig

+ u(p2 , 2 ) u(p1 , 1 )
u(p1 , 1 ) u(p2 , 2 ) .
(p1 p2 )2

(12.344)

For the sacattering amplitude tf i defined by


Sf i ie2 (2)4 4 (p1 + p2 p1 p2 )tf i ,

(12.345)

we find
tf i =

u(p1 , 1 ) u(p2, 2 )
u(p1 , 1 )u(p1, 1 )
u(p2 , 2 ) u(p2, 2 ) u(p2 , 2) u(p1, 1 )

.
(p1 p1 )2
(p1 p2 )2
(12.346)

There is a manifest antisymmetry of the initial or final states accounting for the
Pauli principle. Due to the identity of the electrons, the total cross section will be
obtained by integrating over only half of the final phase space.
Let us compute the differential cross section for unpolarized initial beams, when
the final polarizations are not observed. The kinematics of the reaction in the center
of mass frame is represented in Fig. 12.18, where is the scattering
angle in this
frame. The energy E is conserved, and we denote |p| = |p | = p = E 2 m2 . Using
the general formula (9.309) with the covariant fermion normalization V 1/E, we
obtain
M 2 e4
d
=
|tf i |2 ,
dCM
4E 2 (2)2

(12.347)

The bar on the right-hand side indicates an average over initial and a sum over final
polarizations. More explicitly, we must evaluate the traces
|tf i |2 =

1 X
|tf i |2
4 1 2
1 2

835

12.13 Electron-Electron Scattering

Figure 12.18 Kinematics of electron-electron scattering in the center of mass frame.

1
=
4

6 p1 + M 6 p1 + M
6 p + M 6 p2 + M
1
tr
tr
p

2M
2M
2M
2M
[(p1 p1 )2 ]2
!
6 p + M 6 p2 + M 6 p + M 6 p1 + M
1
tr

2M
2M
2M
2M
(p1 p1 )2 (p2 p1 )2
+(p1 p2 )} .
(12.348)

This is done using the formulas


tr [ (6 p1 + M) (6 p1 + M)] = 4(p1 p1 g p1 p1 + p1 p1 + M 2 g ),
tr [ (6 p1 + M) (6 p1 + M)] tr [ (6 p2 + M) (6 p2 + M)]
= 32[(p1 p2 )2 + (p1 p2 )2 + 2M 2 (p1 p2 p1 p2 )],
(12.349)
and further
(6 p1 + M) (6 p2 + M) = 2 6 p2 6 p1 + 4M(p2 + p1 ) 2M 2 , (12.350)
leading to
tr[ (6 p1 + M) (6 p2 + M) (6 p2 + M) (6 p1 + M)] = 32(p1 p2 )2 2M 2 p1 p2 ],
and thus to
|tf i

|2

1
=
2M 4

(p1 p2 )2 + (p1 p2 )2 + 2M(p1 p2 p1 p2 )


[(p1 p1 )2 ]2
(p1 p2 )2 + (p1 p1 )2 + 2M 2 (p1 p1 p1 p2 )
+
[(p2 p1 )2 ]2
)
(p1 p2 )2 2M 2 p1 p2
+2
.
(p1 p1 )2 (p2 p1 )2

(12.351)

This can be exressed in terms of the Mandelstam variables s, t, u whose properties


were discussed in Eqs. (9.316)(9.329), yielding
|tf i

|2

1
=
2M 4

1 s2 + t2
1 s2 + u 2
2
2
+
2m
(t

m
)
+
+ 2m2 (u m2 )
t2
4
u2
4



1
s
s
2
2
+
.
(12.352)
m
3m
tu 2
2
"

"

836

12 Quantum Electrodynamics

We may now easily express all invariants in terms of center-of-mass energy ECM and
scattering angle :
2
p1 p2 = 2ECM
M 2,
2
p1 p1 = ECM
(1 cos ) + M 2 cos ,
2
p1 p2 = ECM
(1 cos ) M 2 cos .

(12.353)

This leads to the Mller formula (1932):


2
2
d
2 (2ECM
M 2 )2
4
4
3
(ECM
M 2 )2
=
1+

+
4
2
2
2
2
2
2
2
2
dCM
4ECM (ECM M ) sin sin (2ECM M )
sin2

"

#

(12.354)

In the ultrarelativistic limit of high incident energies M/ECM 0,


2
d
2
dur
ECM

4
2
2
1
=

+
2
4ECM
sin4 sin2 4

1
1
+ 1 . (12.355)
+
4
sin /2 cos4 /2

2
2
For small energies where ECM
M 2 , v 2 = (p2 M 2 )/ECM
, we obtain

4
2 1
3

4
4
M 4v sin sin2
!


1
1
1
2 1

. (12.356)
+
=
M 16v 4 sin4 /2 cos4 /2 sin2 /2 cos2 /2

d
=
d nr

This result was first derived by Mott (1930).


Comparing (12.356) with the classical Rutherford formula for Coulomb scattering
in Eq. (12.186), we see that the forward peak is the same if we set Z = 1 and replace
M by the reduced mass M/2. The particle identity yields, in addition, the backward
peak.

12.14

Electron-Positron Scattering

Let us now consider electron-positron scattering. The kinematics and lowest-order


diagrams are depicted in Figs. 12.19 12.20. Polarization indices are omitted and in

Figure 12.19 General form of the diagrams contributing to electron-positron scattering.

837

12.14 Electron-Positron Scattering

Figure 12.20 Lowest-order contributions to electron-positron scattering.

Fig. 12.20 four-momenta are oriented according to the charge flow. The scattering
amplitude may then be obtained from (12.346) by substituting
p2 q1 ,
p2 q1 ,

u(p2 ) v(q1 ),
u(p2 ) v(q1 ),

and by changing the sign of the amplitude. The center of mass cross section is then
given by the formula
d
M 4 e4
|tf i |2
=
2
d
4ECM
(2)2

(12.357)

with
|tf i

|2

1
=
2M 4

(p1 q1 )2 + (p1 q1 )2 2M 2 (p1 q1 p1 q1 )


[(p1 + p1 )2 ]2
(p1 q1 )2 + (p1 p1 )2 + 2M 2 (p1 p1 + p1 q1 )
+
[(p1 + q1 )2 ]2
)
(p1 q1 )2 + 2M 2 p1 q1
+2
(p1 p1 )2 (p1 + q1 )2

(12.358)

This can be exressed in terms of the Mandelstam variables s, t, u whose properties


were discussed in Eqs. (9.316)(9.329) as follows:
|tf i |2

1
=
2M 4

1 u2 + t2
1 u 2 + s2
2
2
+ 2m (t m ) + 2
+ 2m2 (s m2 )
2
t
4
s
4



u
u
1
.
(12.359)
m2
3m2
+
st 2
2
"

"

It is then straightforward to derive the cross section formula first obtained by Bhabha
(1936):
+

4
2
d e e
5

8ECM
M 4
(2ECM
M 2 )2
=

+
(12.360)
2
2
2
2
d
2ECM
4 ECM
(ECM
M 2 )(1cos ) 2(ECM
M 2 )2 (1cos )2
#
4
2
2ECM
(1 + 2 cos + cos2 ) + 4ECM
M 2 (1 cos )(2 + cos ) + 2M 4 cos2
+
.
4
16ECM

"

838

12 Quantum Electrodynamics

In the ultrarelativistic limit, this becomes


+

d e e
d

1 + cos4 /2 1
cos4 /2
2
2
. (12.361)
+
=
(1
+
cos
)

2
ur 8E 2
2
sin4 /2
sin2 /2
CM
#

"

The nonrelativistic limit is simply


+

d e e

=
d nr M


2

1
16v 4 sin4

/2

(12.362)

This agrees again with the classical Rutherford cross section (12.186). The annihilation diagram does not contribute in this limit,

Figure 12.21 Experimental data for electron-electron and electron-positron scattering


at = 900 as a function of the incident electron energy in the laboratory frame. (a)
Electron-electron scattering. The solid line represents the Mller formula, the broken one
the Mller formula when the spin terms are omitted. (b) Electron-positron scattering.
The solid line is the Bhabha formula, the broken one the prediction when annihilation
terms are deleted. (From A. Ashkin, L.A. Page, and W.M. Woodward, Phys. Rev. 94,
p. 357, 1974.)

The results of Eqs. (12.354) and (12.360) may be compared with experimental
data. At low energies we show in Fig. 12.21 some data of Ashkin, Page, and Woodward (1954) for electron-electron scattering at 90 degrees. Mllers formula (12.354)
gives a good agreement. The agreement being a signal for the spin-1/2 property of
the electronsspinless particles would give a bad agreement (see the dashed curve
in Fig. 12.21).
Electron-positron scattering data are fitted well by Bhabhas cross section, and
the annihilation term is essential for the agreement. The energy of the incident
particle in the laboratory frame plotted on the abscissa is chosen in the intermediate range where neither the nonrelativistic nor the ultrarelativistic approximation
is valid. The numerical values show a significant departure from the ratio 2:1 between e e and e e+ cross sections, expected on the basis of a naive argument of
indistinguishability of the two electrons.

12.15 Anomalous Magnetic Moment of Electron and Muon

12.15

839

Anomalous Magnetic Moment of Electron and Muon

The most directly observable effect of loop diagrams is a change in the magnetic
moment of the electron. Recall the precession equation (6.74).
As a consequence of loop diagrams in quantum electrodynamics, the gyromagnetic ratio g in the relation (12.117) receives a corrections and becomes g = 2(1 + a).
The number
a = (g 2)/2 > 0
(12.363)
is called the anomalous magnetic moment of the electron. It has been measured
experimentally with great accuracy [13]:
a = 1 159 652 188.4(4.3) 1012 .

(12.364)

The numbers in parentheses indicate the error estimate in the last two digits. For
the positron:
a = 1 159 652 187.9(4.3) 1012 .
(12.365)
Quantum electrodynamics has been able to explain these numbers up to the last
digits a triumph of quantum field theory.
To lowest order in , the anomalous magnetic moment can easily be calculated.
Interestingly enough, it is found to be a finite quantity; no divergent integrals occur
appear in its calculation. The Feynman diagram responsible for it is the vortex
correction shown in Fig. 12.23. This diagram changes the electromagnetic current
of an electron from
hp , s3 |j |p, s3 i = e
u(p ) u(p)
(12.366)
to
hp , s3 |j |p, s3 i = e
u(p ) [ + (p , p)] u(p)

(12.367)

Figure 12.22 Cross section for Bhabha scattering at high energy, for scattering angle 450 < < 1350 as a function of total energy-momentum square s. The solid line
is calculated from quantum electrodynamics with first-order radiative corrections (From
K. Strauch in Proceedings of the Sixth Symposium on Electron and Photon Interactions at
High Energies, edited by H. Rollnik and W. Pfeil, North-Holland, Amsterdam, 1974.

840

12 Quantum Electrodynamics

Figure 12.23 The vertex correction responsible for the anomalous magnetic moment

where the vertex correction (p , p) is given by the Feynman integral

(p , p) = i4

d4 k
1
1
1

2 .
4
(2)
p/ + k/ M p/ + k/ M k

(12.368)

This integral is logarithmically divergent at large momenta k . It can be regularized


by cutting the integration off at some large but finite momentum which is later
removed by a renormalization of the charge in the Lagrangian. There is also an
infrared divergence which is kept finite by cutting off the k-integration at a small
mass value k 2 = 2 , much smaller than the electron mass, i.e., 2 M 2 .
To do the integral, rewrite the integrand as

p/ + k/ + M p/ + k/ + M
1

2
,

2
2
2
(p + k) M
(p + k) M
k 2

(12.369)

where we have introduced a small photon mass to avoid infrared divergencies at


small momenta. We now collect the product of denominators into a single denominator with the help of Feynmans formula (11.157) for three denominators
Z 1
Z x
1
=2
dx
dy[Ay + B(x y) + C(1 x)]3 ,
ABC
0
0

(12.370)

so that
(p

1
1
1
2
2
2
+ k) M (p + k) M k 2

=2

dx

dy

nh

(12.371)
i

(p + k)2 M 2 y + (p+k)2 M 2 (xy) + (k 2 2 )(1 x)

o3

This can be simplified to


2

dx

dy [k p y p(x y)] 2 (1 x) M 2 x2 + q 2 y(x y)

After performing a shift of the integration variable


k k + p y + p(x y),

o3

(12.372)

841

12.15 Anomalous Magnetic Moment of Electron and Muon

the vertex correction (12.368) takes the form


d4 k
(12.373)
(2)4
0
0
[/
p (1 y) k/ p/ (x y) + M] [/
p (1 x y) k/ p/ y + M]
.

[k 2 2 (1 x) M 2 x2 + q 2 y(x y)]3

(p , p) = i4 2

dx

dy

Instead of calculating this general expression, we shall restrict ourselves to matrix elements of the current between electron states and evaluate u(p , s3 ) (p , p)u(p, s3).
Then we can use the mass shell conditions p2 = p2 = M 2 and the Dirac equations p/ u(p, s3 ) = Mu(p, s3 ) and u(p , s3 )/
p = M u(p , s3 ). We now employ appropriately the anticommutation rules of the gamma matrices using the formulas
(12.212)(12.214), perform a Wick rotation k 0 ik 4 , and integrate over the euclidean four-momenta d4 kE = 2 2 dk k 3 , setting p2 = p2E , p2 = pE 2 . Cutting off
the k-integral at kE = , we arrive at the triple integral

u(p

, s3 ) (p , p)u(p, s3 ) =
h

dx

dy

dkE kE3

kE3
3
[kE2 + 2 (1x)+M 2 x2 +qE2 y(xy)]

u(p , s3 ) kE2 2M 2 (x2 4x + 2) + 2q 2 (y(x y) + 1 x)


o

(12.374)

4Mp (y x + xy) 4Mp (x2 xy y) u(p, s3 ).

The denominator is symmetric under the exchange y xy. Under this operation,
the coefficients of p and p are interchanged, showing that the vertex function is
symmetric in p and p . We can therefore replace each of these coefficients by the
common average
(

y x + xy
x2 xy y

1
1
[(y x + xy) + (x2 xy y)] =
,
2
2x(1 x)

and rewrite (12.374) as


u(p , s3 ) (p , p)u(p, s3) =

, s3 )

u(p

kE2

dx

dy

dkE kE3

kE3
3
[kE2 + 2 (1x)+M 2 x2 q 2 y(xy)]

2M (x 4x + 2) + 2q 2 (y(x y) + 1 x)

(12.375)

2M(p + p )x(1 x)} u(p, s3 ).

This expression may be decomposed as follows:

u(p

, s3 ) (p , p)u(p, s3)

= u(p

, s3 )

1
(p + p )G(q 2 ) u(p, s3 ),
H(q )
2M
(12.376)

with the invariant functions

H(q ) =

dx

dy

dkE kE3

kE2 2M 2 (x2 4x + 2) + 2q 2 [y(x y) + 1 x]


3
[kE2 + 2 (1 x) + M 2 x2 q 2 y(x y)]
(12.377)

842

12 Quantum Electrodynamics

and
G(q 2 ) =

2
M

dx

dy

4M 2 x(1 x)
3.
[kE2 + 2 (1 x) + M 2 x2 q 2 y(x y)]
(12.378)

dkE kE3

The momentum integral in the second invariant function G(q 2 ) is convergent


at small and large momenta, such that we can set the photon mass to zero and
ultraviolet to infinity. Using the integral formula

we obtain

dkE2 kE2

2
M

G(q 2 ) =

1
0

(kE2

dx

1
1
=
,
2 3
+ M1 )
2M12

dy

G(q ) = M 2

1
0

x(1 x)
.
q 2 y(x y)

(12.380)

(12.381)

M 2 x2

The integral over y yields


2

(12.379)

dx 4(1 x) q
arctan
q 2 (4M 2 q 2 )

q2
.
4M 2 q 2

leading to
2M 2

arctan
G(q 2 ) = q
q 2 (4M 2 q 2 )

q2
.
4M 2 q 2

(12.382)

q2
.
4M 2

(12.383)

which can be rewritten as

G(q 2 ) =

2
,
2 sin 2

with

sin2

For small q 2 , it has the expansion

q2
G(q ) =
1+
+ ... .
2
6M 2
!

(12.384)

In the first invariant function H(q 2 ), both the cutoff and the photon mass are
necessary to obtain a finite result. The divergence can be isolated by a subtraction
of the integrand, separating
H(q 2 ) = H(0) + H(q 2 ),

(12.385)

with a divergent integral

H(0) =

dx

dy

dkE kE3

kE2 2M 2 (x2 4x+2)


3
[kE2 + 2 (1x) + M 2 x2 ]
(12.386)

843

12.15 Anomalous Magnetic Moment of Electron and Muon

and a convergent one at large momenta


Z 1 Zx Z
dx dy
dkE kE3
H(q ) =
0
0
0
2

kE2 2M 2 (x2 4x+2) + 2q 2 [y(xy) + 1 x]


3
[kE2 + 2 (1x) +M 2 x2 q 2 y(xy)]
)
kE2 2M 2 (x2 4x + 2)
2
(12.387)
3 .
[kE + 2 (1x) + M 2 x2 ]
(

The divergent momentum integral in (12.386) must be performed with the help of
some regularization scheme, for which we choose the Pauli-Villars regularization,
replacing the photon propagator as follows:
1
1
1
2
2
,
(12.388)
2
2
2
k
k
k 2

where is a large cutoff mass. Using the formula


Z

dkE2

kE2

kE2 + M22
[2 2 ]
2
2 3
2
(kE + (1 x) + M1 )
2 (1 x) + M12 1
M22
= log
+
.
2 (1 x)
2 2 (1 x) + M12
#

("

(12.389)

The convergent momentum integral (12.387) yields

x
M 2 (x2 4x + 2) q 2 [y(x y) + 1 x]
1
dx dy
H(q ) =
2 0
2 (1 x) + M 2 x2 q 2 y(x y)
0
#)
"
M 2 (x2 4x + 2)
2 (1 x) + M 2 x2 q 2 y(x y)
+ 2
. (12.390)
log
(1 x) + M 2 x2
2 (1 x) + M 2 x2

If the photon mass is set equal to zero, the integral is divergent at x = 0. The
physical meaning of this infrared divergence will become clear later.
Let us first understand H(q 2 ) for small q 2 by expanding H(q 2 ) = H (0)q 2 +
O(q 4 ). Then H (0) is given by the integral
(

x
1
y(x y) + 1 x
H (0) =
dx dy
2 0
2 (1 x) + M 2 x2
0
)
y(x y)
M 2 (x2 4x + 2)y(x y)
+ 2

[2 (1 x) + M 2 x2 ]2
(1 x) + M 2 x2
!
1
M2
1
=
,
log 2
2 3

12

where we have discarded all terms which go to zero for 0.


The full result is
"
!
!
M2
2

2
log 2 2 1
H(q ) = H(0) +
2

tan 2
#
Z
4
2
+ tan +
dx x tan x +
tan 2 0
sin 2
which has precisely the first Taylor coefficient (12.391).

(12.391)

(12.392)

844

12.15.1

12 Quantum Electrodynamics

Form Factors

We now introduce the customary Lorentz-invariant decompositions of the matrix


elements of the current (12.367) of a spin-1/2 particle as follows:
i
hp |j |pi = e
u(p ) F (q ) +
q G(q 2 ) u(p ).
2M

(12.393)

The invariant functions F (q 2 ) and G(q 2 ) are the standard form factors of the electron. The relation between this and (12.376) follows directly from Gordons decomposition formula (12.133), showing that G(q 2 ) in (12.378) coincides with G(q 2 ) in
(12.393), whereas
F (q 2 ) = 1 + H(q 2 ) G(q 2 ).

(12.394)

The charge form factor F (q 2 ) at q 2 = 0 specifies the charge of the electron.


Inserting (12.392), we obtain

F (q ) = F (0) +
2
2

where

M2
2
log 2 2 1

tan 2
#
Z
4
+ tan +
dx x tan x
tan 2 0
"

9
M2
F (0) 1 +
log
+ log 2 .
2
M
4

(12.395)

(12.396)

The value F (0) contains both the ultraviolet and the infrared cutoff. The subtracted
function F (q 2 ) F (q 2 ) F (0) has only an infrared divergence. In writing down
the expression (12.395) and (12.396), we have ignored all contributions which vanish
for 0 and .
Due to the loop integral, the charge is changed to the new value
"

9
M
log
e1 = eF (0) = e 1 +
+ 2 log
2
M
4

!#

(12.397)

The factor is commonly denoted as the renormalization constant Z11 . To order ,


it is
Z1 F

9
M
log
(0) = 1
+ 2 log
2
M
4

(12.398)

According to the theory of renormalization, this has to be equated with the experimentally observed charge. After this we can replace in Eq. (12.395) the number
F (0) by 1, and the bare fine-structure constant by the physical one (keeping the
notation, for simplicity). The latter substitution is also done in G(q).

845

12.15 Anomalous Magnetic Moment of Electron and Muon

12.15.2

Charge Radius

For small momentum transfers q 2 , the renormalized charge form factor has the Taylor
series expansion
!

M
3
FR (q 2 ) = 1 + q 2
+ O(q 4 ).
log

2
3M

(12.399)

The form factor of the anomalous magnetic moment is for small q


GR (q 2 ) = q 2

1
+ O(q 4 ).
2
3M 4

(12.400)

Due to the emission and absorption of virtual photons, the electron is shaken
over a finite range. It is customary to define here a charge radius Re of the electron
by the first term in the expansion of the charge form factor1
F (q 2) = 1 q 2

Re2
+ ... .
6

(12.401)

The factor 1/6 is due to the fact that for a uniformly charged shell of radius Re , the
Fourier transform of the charge density has the expansion
1
F (q ) =
d x e (x) = d x 1 + iqx (qx)2 + . . . (x)
2
Z
2
R
1
(12.402)
= 1 q2 d3 x r 2 (x) = 1 q2 e + . . . .
6
6
2

iqx

Setting q 0 = 0 in Eq. (12.399), we find the charge radius Revc of the electron due to
the vertex correction to be given by
Re2 vc

2
M
3
=
.
log

2
M

(12.403)

If the zeroth component of the current (12.393) couples to a static electric potential A0 (x), the q 2 -term in (12.399) yields a correction factor
R2
1 + e 2 A0 (x)
6

"

(12.404)

The fluctuations of the electron position leads to an extra term proportional to


the harmonic average of the potential. Since the potential obeys Gauss law, this
is nonzero only where there are charges. In an atom of nuclear charge Ze at the
origin, the potential is
eA0 (x) = VC (x) =
1

Z
r

(12.405)

Note the difference of this quantity with respect to the classical electron radius (12.226).

846

12 Quantum Electrodynamics

and the harmonic average produces an additional -function at the origin:


e2 A0 (x) = 4Z (3) (x).

(12.406)

Thus the correction factor changes the Coulomb potential into an effective potential
VCeff =

Z Re2 vc
+
Z 4 (3) (x).
r
6

(12.407)

The prediction of this additional term is the origin of one of the early triumphs of
QED. The -function leads to an energy shift of s-wave orbits with respect those of
nonzero orbital angular momentum, called the Lamb shift. It removes the degeneracy between the 2S1/2 and the 2P1/2 predicted by Diracs equation in an external
Coulomb field. It will be discussed in detail below.

12.15.3

Anomalous Magnetic Moment

On the right-hand side of the current matrix element (12.393) we replace the term by a combination of vectors p + p and q as before in (12.133), and see
that the magnetic moment of the free electon acquires a correction factor 1 + G(0).
Thus the gyromagnetic ratio g in Eq. (12.118) is changes from g = 2 to
g 2[1 + a] = 2[1 + G(0) + . . . ]

(12.408)

[recall (12.363)]. The number G(0) yields therefore directly the anomalous magnetic
moment:
a = G(0).
(12.409)
In contrast to the charge, this quantity is finite and has the value [14]
a = G(0) =

1 161 409 74292 1012 .


2

(12.410)

This result was first calculated by Schwinger [15]. It is about 1.5% larger than
the experimental value (12.365). (12.410). The difference can be explained by
higher-order electrodynamic and strong-interaction corrections [16]. By including
all diagrams of sixth order in perturbation theory one finds, after a considerable
effort (there are 72 Feynman diagrams to sixth order), the expansion
 2

a=
+ c2
2

+ c3

 3

(12.411)

with the coefficients [17]


c2
c3

3
1
197
+
3 log 2 (2) + (4) = 0.328 478 965 . . . ,
=
144
2
4
= 1.176 11(42).
(12.412)


847

12.15 Anomalous Magnetic Moment of Electron and Muon

Up to c2 , the theory is lower than the experimental number by only 1 part in 105 .
Adding the c3 -term, the theoretical value becomes
a = 1 159 652 140.4(27.1) 1012 ,

(12.413)

reducing the discrepancy to 1 part in 108 . The error is mainly due to the uncertainty
in the fine-structure constant quoted in the footnote of p. 878.
The above calculation can be modified only slightly to show that the anomalous
magnetic moment of an antiparticles positron is the same as for a particle. The
matrix element of the initial current between positron states is

hp , s3 |j (x)|p, s3 i = eh0|b(p , s3 )(x)


(x)b (p, s3 )|0i = e
v (p , s3 ) u(p, s3)eiqx .
(12.414)
rather than (12.121). The minus sign is due to the odd number of operator exchanges
necessary to evaluate the anticommutators. If both momenta are zero, only the
zeroth component survives showing that the charge of the positron is e. For small
momentum transfers we calculate the second spatial component of (12.414) more
explicitly as

hp , s3 |j 2 (x)|p, s3 i = e
v (0, s3 ) 2 ei

1 (i 0 1 )/2

v(0, s3),

(12.415)

and compare this with (12.126). The term linear in 1 contains the contribution
of the charge form factor to the magnetic moment. We see that (12.415) has an
opposite overall sign of (12.126) which is compensated by an opposite sign in the
exponent of the Lorentz transformation. Thus we obtain the matrix element of j 2
between positrons:
ie 1 v(0, s3 )S3 v(0, s3 ) = ie 1 s3 .
(12.416)

with an opposite sign in comparison with the electron in (12.127). The sign change at
the end is caused by the opposite spin orientation of the two-spinors (s3) contained
in the spinors v(0, s3 ) [recall (4.678),(4.679) and (4.670)].
An opposite sign is also found for the contribution from the second form factor
where the matrix element of the electron is
i 21
e
u(0, s3 )
q1 G(0)u(0, s3) = ie
u(0, s3 )S3 1G(0)u(0, s3 ) = ie 1 s3 G(0).
2M
(12.417)
This produces the correction to the g-factor
g = 2[1 + G(0)].

(12.418)

For a positron, the matrix element of the current is


i 2
q G(q 2 )]v(p, s3 ),
2M
and the second term becomes for p = 0 and small p in the x-direction
e
v (p , s3 )[ 2 F (q 2 ) +

e
v (0, s3 )

(12.419)

i 21
q1 G(0)v(0, s3) = ie
v (0, s3 )S3 1G(0)v(0, s3 ) = ie 1 s3 G(0),
2M
(12.420)

848

12 Quantum Electrodynamics

exactly the opposite of (12.417). Thus a positron has the same correction (12.418)
to the g-factor as an electron. Note that the sign change found in (12.416) is present
also here, but it is compensated by a minus sign from an extra 0 -matrix acting
upon v .
For a muon, the coefficients are [18]
c2 = 0.765 857 376(27),
c4 = 24.050 508 98(44),

c3 = 24.050 508 98(44),


c5 = +930(170),

(12.421)

leading to
a = 1 165 847 057(29) 1012 .

(12.422)

The strong interactions contribute via the composite Feynman diagram shown in
Fig. 12.24. Their contribution is estimated from [19]

hadrons

Figure 12.24 Leading hadronic vacuum polarization corrections to a .


11
astr
= 6 924(62) 10

to 6 988(111) 1011 ,

(12.423)

The amplitude is calculated from the formula


astr
(vac. pol.) =

1 Z
ds K (s) 0 (s)e+ e hadrons ,
4 3 4m2

(12.424)

where 0 (s)e+ e hadrons is the cross section for the indicated process with some
radiative corrections such as initial state radiation subtracted from the measured
cross sections, and
2

K(s) = x
x =

1
1+

x2
1
2
q

+ (1 + x)

1 4m2 /s

1 4m2 /s

1
1+ 2
x

"

1+x 2
x2
+
ln(1 + x) x +
x ln x
2
1x
#

(12.425)

comes from the remaining diagram. The different parts of the cross sections are
separately listed in Table 12.15.3. The weak interactions in the standard model add

849

12.15 Anomalous Magnetic Moment of Electron and Muon

11
astr
(vac. pol.) 10
6343 60
338.7 4.6
143.1 5.4
68.7 1.1
12.1 0.5
18.0 0.1
6924 62

s (GeV)
2m 1.8
1.8 3.7
3.7 5 + (1S, 2S)
5 9.3
9.3 12
12
Total

to this [20]
aEW
(1 loop)

m2
1
5 G m2
2
2

1 + (1 4 sin W ) + O
=
3 8 2 2
5
M2
"

!#

195 1011. (12.426)

2
where G = 1.16637(1)105 GeV2 , sin2 W 1MW
/MZ2 0.223. and M = MW
or MHiggs . See Fig. 12.25 for the Feynman diagrams.

(a)

(b)

(c)

Figure 12.25 One-loop electroweak radiative corrections to a . The wiggly lines are
gluons.

Two-loop corrections change this slightly by found for a Higgs particle mass
mH 150 GeV (with little sensitivity to the exact value)
aweak
(2 loop) = 43(4) 1011 ,

(12.427)

Altogether, we obtain
atheory
= 1 165 915 97(67) 1011 ,

(12.428)

in agreement with the experimental numbers [21]


9
9
9
aexp
= 11 659 204(7) 10 , 11 659 202(22) 10 , 11 659 204(7) 10 . (12.429)

See Fig. 12.26 for a comparison of theory and experiment.

850

12 Quantum Electrodynamics

Figure 12.26 Measured values of a and prediction of the standard model (SM). The
small arror bars of the theoretical value come from the left-hand estimate for the hadron
contribution in Eq. (12.423). For sources see Footnote 21.

12.16

Vacuum Polarization

Let us now turn to the vacuum polarization. The free propagator of the photon has

Figure 12.27 Lowest-order Feynman diagram for the vacuum polarization

the form
G
0 (q) = i

P (q)
,
q2

(12.430)

where the polarization tensor P (q) depends on the gauge [see (12.104)]. From
the lowest-order Feynman diagram shown in Fig. 12.27, the propagator receives a
lowest-order correction

G
(12.431)
0 (q) [i (q)] G0 (q),

851

12.16 Vacuum Polarization

where (k) is given by the Feynman integral.


i (q) = e2

p + q/ + M) (/
p + M)]
d4 p tr[ (/
.
(2)4 [(p + q)2 M 2 ](p2 M 2 )

(12.432)

Using Feynmans formula (11.155)


1
=
AB

dz

1
,
[Az + B(1 z)]2

(12.433)

we rewrite (12.432) as
Z

i (q) = e2

dz

d4 p tr[ (/
p + q/ + M) (/
p + M)]
.
4
2
2
2
(2) [(p + qz) + q (z z ) M 2 ]2

(12.434)

This expression is symmetric in and . A shift in the p-integration by an amount


qz brings it to the from
2

i (q) = e

dz

d4 p tr{ [/
p + q/ (1 z) + M] (/
p q/ z + M)}
(12.435)
4
2
2
2
(2)
[p + q (z z ) M 2 ]2

After evaluating the trace, dropping odd terms in p, and using the symmetry in
and , we obtain
d4 p (g q 2 + 2q q )(z z 2 ) g (p2 M 2 )+2pp
.
(2)4
[p2 + q 2 (z z 2 ) M 2 ]2
0
(12.436)
Because of the rotational symmetry of the intgrand we can use the first replacement
rule in Eq. (11.136) to replace g (p2 M 2 ) + 2p p by g (p2 /2 M 2 ).
The momentum integral is quadratically divergent, since there are two more powers of integration variables in the numerator than in the denominator. The situation
is improved by imposing the requirement of gauge invariance, q (q) = 0. This
makes the final results independent of the gauge choice in the photon propagator.
Thus we postulate for the moment the vanishing of the divergent integral
2

i (q) = 4e

dz

iq (q) = e2 q 4

dz

d4 p q 2 (z z 2 ) (p2 /2 M 2 )
.
(2)4 [p2 + q 2 (z z 2 ) M 2 ]2

(12.437)

We shall verify in the next Subsection that this is guarenteed if we calculate the
Feynman digrams by analytic regularization in D dimensions rather than D = 4.
Thus we are left with the logarithmically divergent integral
1
d4 p
.
i (q) = (g q + q q ) 8e
dz(z z )
4
2
2
(2) [p + q (z z 2 ) M 2 ]2
0
(12.438)
Let us define the invariant function that accompanies the tensor (g q 2 + q q ) =
q 2 P (q) as i(q 2 ), i.e., we write
2

i (q) iP (q)q 2 (q 2 ).

(12.439)

852

12 Quantum Electrodynamics

If we expand
(q 2 ) = (0) + (0)q 2 + . . . ,

(12.440)

we see that only (0) is logarithmically divergent:


i(0) = 8e2

dz(z z 2 )

d4 p
1
,
(2)4 (p2 M 2 )2

(12.441)

whereas all remaining terms in the expansion (12.440) are finite, for example the
first term:

i (0) = 16e

2 2

dz(z z )

d4 p
1
4
2
(2) (p M 2 )3

(12.442)

Since the mass M carries a small negative imaginary part, we now perform a Wickrotation of the integration contour setting p0 = ip4 and letting p4 run from to
. Thus we substitute p2 = p2E and d4 p/(2)4 = id4 pE /(2)4, and calculate from
(12.442):
1
e2 4 2
=
.
(12.443)
(0) =
4
2
(2) 15 M
6M 2 5
In the divergent quantity (0), we perform a Wick rotation so that the integral d4 p/(2)4 turns into an integral over Euclidean momentum d4 pE /(2)4 =
S4 dpE p3E = dp2E p2E /16 2 , and introduce an ultraviolet momentum cutoff at 2 :
d 4 pE
1
21 1
=
8e
2
(2)4 (pE + M 2 )2
6 16 2
!
1
2
=
log 2 1 .
3
M
1
6

(0) = 8e2

dp2E

p2E
(p2E + M 2 )2
(12.444)

The complete invariant function of the polarization tensor is therefore


2
1
1
2
log 2 1 +
(q ) =
+
3
M
3 3 sin2
"

1
tan

(12.445)

For small q 2 , we expand


2
1
(q 2 ) =
+
3 3 sin2


1
tan

1
4
+ 2 + . . . ,
9 15

(12.446)

and find the bracket to behave like


2
1 q2
1
log 2 1 +
+ O(q 4 ) .
(0) + (0)q + . . . =
2
3
M
15 M

"

(12.447)

The last term agrees with the result (12.443) of a direct calculation of (0).
It is gratifying to find out that the condition (12.437) is fulfilled by the dimensional regularization in d = 4 dimensions.

853

12.16 Vacuum Polarization

Note that the expression (12.445) is real only for q 2 < M 2 . When q 2 > (2M)2 the
external field can produce electron-positron pairs and (q 2 ) acquires an imaginary
part. The imaginary part causes a decrease of the probability amplitude for the
occurrence of a pure scattering process below the threshold of pair production.
A string of vacuum polarization diagram, produces the geometric series

G = G
0 + G0 [i ] G0 + . . . + G0 [i ]G0 [i ]G0 + . . . , (12.448)

which can be summed up to


1 1
G (q) = {[G1
(q) + i (q)] } .

(12.449)

Inserting the tensor decomposition (12.439), this can be written as


G (q) = i

P (q)
P (q)

i
q 2 [1 (q 2 )]
q 2 [1 + (0) + (0)q 2 . . .]

(12.450)

For small q 2 , the photon propagator is, therefore,


G (q) i

P (q)
1
+
(0)
+
q2
q2
15M 2


1

(12.451)

The divergent number 1 + (0) can be absorbed into the field renormalization factor
defining
1/2
AR (x) = Z3 A(x)
(12.452)
with

2
Z3 = [1 + (0)] = 1
log 2 + . . . .
3
m
This corresponds to renormalizing the charge to
1

(12.453)

R = Z3 .

(12.454)

The propagator of the renormalized field AR is then


G
R (q)

P (q)
R q 2
=i
1
+
q2
15 M 2

!1

(12.455)

For a given static source, of charge Ze at the origin, the Coulomb potential
VC (x) =

Z
r

(12.456)

(3)
is obtained by applying the free propagator G
0 to the current j (x) = 0 Ze (x).
To lowest order in , the vacuum polarization changes G0 (q) by a factor

R q 2
1
15 M 2

854

12 Quantum Electrodynamics

which is equivalent to multiplying the potential VC (x) by a factor


R 2
1
.
15 M 2
!

The potential is therefore modified to


VCR (x)

1
+
4 (3) (x) .
= Z
r 15M 2


(12.457)

Comparison with (12.399)(12.407) shows that the vacuum polarization decreases


the effective radius of the electron (12.403) derived from the vertex correction by
Revp =

1
,
3M 2 5

(12.458)

thus giving rise to a total effective radius to lowest order in :


Revp

1
=
3M 2 5

"

M
3 1
log

8 5

!#1/2

(12.459)

Since the finite radius of the electron gives rise to the Lamb shift to be derived
below, the vacuum polarization decreases the Lamb shift.

12.17

Dimensional Regularization

We still owe the reader the proof that the divergent integral (12.437) is really zero to
ensure the gauge invariance of the self-energy (12.460). In D dimensions, the Dirac
matrices have the dimension 2D/2 and (12.460) reads [23]
dD p (g q 2 +2q q )(zz 2 )g (p2 M 2 )+2pp
.
(2)D
[p2 + q 2 (z z 2 ) M 2 ]2
0
(12.460)
and the previous replacement under the integral p p g /4 becomes p p
g/D [recall (11.136)]. The integral dD pE over Euclidean momentum spacetime
may now be Wick-rotated to
i (q) = 2

D/2 2

dz

dD pE /(2)D = iSD /2(2)D

dp2E (p2E )D/21 ,

(12.461)

where SD is the surface of the unit sphere (11.126). Hence we can rewrite the integral
(12.437) in Euclidean spacetime as
iq (q) = e2 q 2D/2 i

dz

dD pE q 2 (zz 2 )+ (1 2/D)p2E + M 2
, (12.462)
(2)D
[p2E + m2 ]2

where
m2 = m2 (q 2 , z) M 2 q 2 z(1 z),

(12.463)

855

12.18 Two-Dimensional QED

and further as
iq (q) = e2 q 2D/2 i

dz

dD pE 2m2 /D + (1 2/D)(p2E + m2 )
.
(2)D
[p2E + m2 ]2

(12.464)

Now we use Formula (11.130) to replace


Z

(D/2) ( D/2)
1
d D pE
= S
D
()[m2 ]D/2 ,
2
D
2

(2) (pE + m )
2

(12.465)

so that
(D/2)
2


Z 1

2
2
dz[m2 ]D/21 = 0, (12.466)

(2D/2) (1)
(1D/2) (2) + 1
D
D
0

iq (q) = e2 q 2D/2 i
SD

thus guaranteeing the gauge invariance of the vacuum polarization tensor in any
dimension D [6].

12.18

Two-Dimensional QED

An interesting special situation arises in two dimensions if the initial electron is massless. Then the self-energy makes the photon massive in spite of gauge invariance.
This was observed by Schwinger [4] and for this reason two-dimensional massless
QED is called the Schwinger model [24]. Consider the self-energy (12.438) in D
spacetime dimensions where the prefactor 4 in Eq. (12.460) is 2D/2 so that (12.438)
reads
dD p
1
.
D
2
2
(2) [p + q (zz 2 )M 2 ]2
0
(12.467)
The integral over dD p in momentum spacetime may be Wick rotated as in (12.461)
to the Euclidean momentum integral
i (q) = (g q 2 + q q ) 2D/2+1 e2

SD
i
2(2)D

dp2E

(p2E

dz(zz 2 )

SD
(D/2)(2D/2)
1
=
i[m2 ]D/22
,
2
2
D
+m )
2(2)
(2)

(12.468)

with m2 from Eq. (12.463). Inserting this into (12.467) yields for i (q) of massless electrons
(g q 2 +q q ) 2D/2+1 e2 i

dz [z(1z)]D/21 (q 2 )D/22

SD (D/2)(2D/2)
2(2)D
(2)
(12.469)

which is equal to2


(g q 2 + q q )2D/2+1 e2 i
2

ibid., Formula 3.191.3.

SD (D/2)(2D/2)
2 (D/2)
(q 2 )D/22
(12.470)
(D)
2(2)D
(2)

856

12 Quantum Electrodynamics

so that we obtain in D = 2 dimensions


i (q) = i(g q 2 + q q )

4e2 S2
e2
e2
2
=
i(g
q
+
q
q
)
=
iP
.

q 2 2(2)2
q 2
q 2
(12.471)

or, recalling (12.439),

! (q) =

e2
.
q 2

(12.472)

Inserting this into (12.450), we obtain the renormalize photon propagator


G (q) = i

P (q)
P (q)
=
i
,
q 2 + q 2 (q)
q 2 e2 /

(12.473)

which shows that the photon has acquired a nonzero mass m2 = e2 /. The effective
Lagrangian of the photon to this order is
Leff = 41 F (1 + m2 /

)F .

(12.474)

The Schwinger model illustrates that a mass term can be generated by a loop diagram in spite of gauge invariance.

12.19

Self-Energy of Electron

The lowest-order Feynman diagram for the self-energy of the electron is shown in
Fig. 12.27.

Figure 12.28 Lowest-order Feynman diagram for the vacuum polarization

It adds to the electron propagator a term


G0 (p)[i(p)]G0 (p)
where
G0 (p) =

i
p/ M

(12.475)

(12.476)

is the free Dirac propagator and the self-energy i(p) of the electron is given by
the diagram in Fig. 12.28. Explicitly:
i(p) = e2

d4 k p/ k/ + M
1

(2)4 (p k)2 M 2 k 2

(12.477)

857

12.19 Self-Energy of Electron

Using the anticommutation rule (12.206), we can simplify in the numerator


(/
p k/ + M) = 2(/
p k/ ).

(12.478)

The integral is logarithmically divergent at large k. It also has an infrared diergence. To enforce convergence, we employ the Pauli-Villars regularization method
and modify the photon propagator as follows:
1
1
1

.
k2
k 2 2 k 2 2

(12.479)

where is a large cutoff mass and is a small photon mass. The self energy becomes
i(p) = e2

p k/ ) + 4M
d4 k 2(/
4
(2) (p k)2 M 2

1
1
2
.
2
2
k
k 2

(12.480)

The mass parameter cuts off the contribution of short-wave photons with k 2 .
At the end we shall take the cutoff to infinity.
By adding the same Feynman diagram repeatedly to an electron line, we obtain
the geometric series
G0 (p) + G0 (p)[i(p)]G0 (p) + G0 (p)[i(p)]G0 (p)[i(p)]G0 (p) + . . . , (12.481)
which can be summed up to
G(p) =

i
.
p/ M (p)

(12.482)

Using Feynmans formula (12.433), this can be rewritten as


d4 k
(2)4
0
(
)
2(/
p k/ ) + 4M

( = ) . (12.483)
[(k px)2 p2 x(1 x) + M 2 x + 2 (1 x)]2
Z

(p) = ie2

dx

A simplification occurs by shifting the integration variable from k to k + px. Then


the terms proportional to k are off in k and can be dropped. After performing a
Wick rotation of the integration contour we obtain
2 2 e2
(p) =
(2)4

2/
p (1 x) + 4M
dkE kE3
ds
2 ( = 0) .
2
2
0
0
[kE p x(1 x) + M 2 x + 2 (1 x)]
(12.484)
The kE -integral is easily done and yields
Z

2 e2
(p) =
(2)4

dx [2/
p (1 x) + 4M]

p2 x(1 x) + M 2 x + 2 (1 x)
( = ) .
log
p2 x(1 x) + M 2 x
(

"

(12.485)

858

12 Quantum Electrodynamics

The right-hand side is a 4 4-matrix in spinor space which may be decomposed into
invariant functions
(p) = (/
p M)A(p2 ) + B(p2 )
(12.486)
where
n
o
2 e2 1
2
A(p ) =
dx[2(1

x)]
log
f
(p
,
x,
)

(
=
)
(2)4 0
n
o
2 e2 Z 1
2
2
dx
2M(1
+
x)
log
f
(p
,
x,
)

(
=
)
,
B(p ) =
(2)4 0
Z

(12.487)

with
f (p2 , x, ) =

p2 x(1 x) + M 2 x + 2 (1 x)
.
p2 x(1 x) + M 2 x

(12.488)

The invariant functions are logarithmically divergent for large .


We may expand them around the mass shell p2 = M 2 in powers of p2 M 2 .
Then only the lowest expansion terms carry the logarithmic divergence:
2 e2 1
M 2 x2 + 2 (1 x)
,
dx
[2(1

x)]
log
A(M ) =
(2)4 0
M 2 x2
"
#
2 2
2
2 2 Z 1
M
x
+

(1

x)

e
dx 2M(1 + x) log
,
B(M 2 ) =
(2)4 0
M 2 x2
Z
2 e2 1
dx 2M(1 + x)(1 x)
B (M 2 ) =
(2)4 0
#
"
2 (1 x)
2 (1 x)
.

M 2 x[M 2 x2 + 2 (1 x)] M 2 x[M 2 x2 + 2 (1 x)]


"

(12.489)

All higher expansion terms are finite and can be evaluated with an infinite cutoff
. We have also dropped all terms which vanish in the limit of zero photon mass .
Omitting the regular parts of the self-energy, the propagator becomes
i
(/
p M)[1 +
+ B(M 2 ) + B (M 2 )(p2 M 2 )
i
1
=
2

2
1 + A(M ) + 2MB (M ) p/ M M

G(p) =

A(M 2 )]

where
M

B(M 2 )
.
1 + A(M 2 ) + 2MB (M 2 )

(12.490)

(12.491)

The prefactor in the denominator is commonly denoted by


Z2 1 + A(M 2 ) + 2MB (M 2 ).

(12.492)

859

12.20 Ward-Takahashi identity


1/2

It can be removed by renormalizing the field (x) to R (x) = Z2 (x). The


renormalized field has then a propagator with a pole term i/(/
p M). For large 2 ,
we can simplify the results for Z2 and M 2 to order :
1
2 (1x)
1x
1x2
2
Z2 1 =
1 2 2
,
dx (1x) log 2 +log 2 2
2 0
M
x
x
M x + 2 (1x)
"
#
Z
2
1x
1
2
dx(1+x) log 2 + log 2 .
(12.493)
M =
2 0
M
x
Z

"

"

#)

Performing the integrals over x yields

5
M2
Z2 1 =
log
+ + 1 log 2 ,
2
M
4



Z 1
3

.
+
dx 3 log
M 2 =
2 0
M
4
"

12.20

(12.494)

Ward-Takahashi identity

It is important to realize that Z2 coincides with the renormalization constant Z1 =


F 1 (0) defined in Eq. (12.398) by the charge form factor to make the current matrix
element finite. This equality is a consequence of the Ward identity fulfilled by the
vertex correction (p , p) defined in Eq. (12.367):
(p, p) =

(p).
p

(12.495)

For the total Dirac matrix (p , p) + (p , p) in the current (12.367) this


implies
(p, p) =

[/
p M (p)] .
p

(12.496)

This holds to all orders in . There exists an even more general relation for offdiagonal matrix elements of the current:
(p p) (p , p) = [/
p M (p)] [/
p M (p)],

(12.497)

from which (12.496) can be derived in the limit p p. This is the famous WardTakahashi identity. Its validity for the bare quantities is obvious. For the interacting
quantities it will be proved below.
To derive Z1 = Z2 we use (12.496) to rewrite

(p) + (p , p) (p, p)
p

= +
[/
p M (p)] + (p , p) (p, p). (12.498)
p

(p , p) =

860

12 Quantum Electrodynamics

From the renormalization equation of the electron propagator (12.490) we see that
p/ M (p) = Z21 [/
p MR R (p)]

(12.499)

where MR = M + M. This leads to


(p , p) = (Z21 1) Z22

R (p) + (p , p) (p, p). (12.500)


p

This is to be compared with the definition of the renormalized vertex function


(p , p) = (Z11 1) + Z11 R (p , p).

(12.501)

Thus we find
Z1 = Z2 ,
R (p , p) = Z1 [ (p , p) (p, p)]

R (p).
p

(12.502)

The proof of the Ward-Takahashi identity [25] follows from the canonical commutator of the current with the field which holds in the presence of interactions:
[j 0 (x, x0 ), (x , x0 )] = e (3) (x x )(x, x0 ),
, x0 )] = e (3) (x x )(x,
x0 )
[j 0 (x, x0 ), (x

(12.503)

This follows from Noethers theorem which makes j 0 the generator of phase transformations [recall Eq. (8.280)]. We now form the derivative of the time-ordered
expectation

z h0|T (x )j (z)(x)|0i
= eh0|T(x )(x)|0i
(4) (x z) (4) (z x) .
(12.504)
Expressed in terms of the full propagators, this becomes
h

iS(x z) z (z) iS(z x) = iS(z x) iS(x z)

(12.505)

After a Fourier transformation, this becomes


S(p )(p p) (p , p)S(p) = S(p) S(p ),

(12.506)

(p p) (p , p) = S 1 (p) S 1 (p ),

(12.507)

or

which is precisely the Ward-Takahashi identity (12.497).


As a consequence of the equality Z1 = Z2 , we find for the matrix elements of the
current between one one-loop corected electron states a unit charge rather than the
divergent charge from the charge form factor F (0) = Z11 . The charge is therefore
unrenormalized.

861

12.21 Lamb Shift

12.21

Lamb Shift

One of the most important early confirmations of the correctness of quantum electrodynamics in describing the theory of electrons and photon was the atomic Lamb
shift. According to Diracs theory, the energy spectrum of an electron in an external
Coulomb field is

Enl = Mc2 1 +

Z
n +

1/2

(12.508)

(j + 1/2)2 2 Z 2

where n = n j 1/2 = 0, 1, 2, . . . is the radial quantum number, and


j =
1/2, 3/2, . . . the total angular momentum. Up to lowest order in the fine-structure
constant = e2 /hc, this is approximately equal to
1
Z2
2 Z 2
Enl = Mc 2 Mc2 2 1 +
2
n
n
2

"

1
3
+ ... .

j + 1/2 4n

(12.509)

These formulas show that the Schrodinger degeneracy of all levels with the same
principal quantum number and different values of the orbital angular l is modified
in Diracs theory, where levels with the same quantum numbers n and j are degenerate for different ls. The lowest states for this degeneracy can be compared with
experiment are the n = 2 - states 2S1,2 and 2P1/2 , and they are found to have a
slightly different energy.
The energy difference is explained by quantum electrodynamics, due to three
physical effects. First, the electron encircling the nucleus is shaken by the vacuum
fluctuations of the electromagnetic field over a range of the order of the Compton
wavelength. Thus it sees a harmonic average of the Coulomb potential over this
length scale. This lifts the the level 2S1/2 agauinst the level 2P1/2 by roughtly
27 MHz. Second, the anomalous magnetic moment of the electron changes slightly
the Coulomb attraction. Third, the photon running through the vacuum can excite
an electron-positron pair. These three effects together cause ultimately an upwards
shift of the level 2S1/2 with respect to the level 2P1/2 equal to [26] E = 1 010 MHz,
68 MHz, and 27 MHz, respectively. The sum is roughly 1 052 MHz, a number
which was confirmed experimentally [27]. The calculation of these effects will now
be reviewed.

12.21.1

Rough Estimate of Effect of Vacuum Fluctuations

The order to estimate the first effect, consider a free nonrelativistic electron of mass
M in the vacuum. It is shaken by the zero-point oscillation of the electromagnetic
field, which causes an acceleration
M x = eE.

(12.510)

For a given frequency , the electron is shaken around its average position by an
amount
e
x = 2 E().
(12.511)
M

862

12 Quantum Electrodynamics

Its square-average is

e2 Z d 2
h(x) i = 2
hE ()i.
(12.512)
M 0 4
The right-hand side can be estimated from the energy (7.331) of the free electromagnetic field in the vacuum, where it has the value [recall (7.424)]
1X
k .
(12.513)
E =
2 k,
2

The polarization sum runs over the two helicities. Hence, with the usual limiting
phase space integral (7.21) for the momentum sum, we have
E=

d3 k
1
k =
3
(2)
2

d 3

(12.514)

Since the vacuum energy is equally distributed between electric and magnetic fields,
we find
Z
1 d 3
2
,
(12.515)
hE i =
2 0
and hence
Z
2 d 1
2
(12.516)
h(x) i = 2
M 0
The integral is divergent at small and large frequencies . A priori, it is unclear
which are the relevant frequencies will contribute in a proper calculation. If we
consider only electromagnetic waves with wavelength shorter than the Bohr radius
aB = 1/M, the integral starts at min = M. Alternatively, we may expect the
energy of the atomic electrons to supply the relevant cutoff. Then min = M2 . On
the high-frequency end, we omit wavelengths shorter than the Compton wavelength
of the electron, were classical considerations become invalid. Thus we cut off the
integral at max = M. In this way we obtain a mean square somewhere in the range
2
hxi2 =
C
(12.517)
M
with a constant C somewhere in the range


1
1
C log
4.92 (1, 2).
(12.518)
, 2 log
min
min
The electric interaction energy of an electron shaken over this region is modified as
follows. The Hamiltonian at the position x + x(t) is
H =e
Averaging over x gives

d3 xA0 (x + x(t)).

1
H = e d x A0 (x) + hxi xj ii j A0 (x)
2


Z
1
2
3
= e d x 1 + hx xi A0 (x)
6


Z
1
2
3
= e d x 1+
C A0 (x).
3 M 2
Z

(12.519)

(12.520)

863

12.21 Lamb Shift

In an atom of nuclear charge Ze with a Coulomb potential (12.405), the Laplace


operator yields (12.406) and the potential is changed into an effective one
VCeff =

Z
1
+
CZ4 (3) (x).
r
3 M 2

(12.521)

For an atomic s-state with wave function n , we treat the extra potential in (12.521)
perturbatively, i.e., we merely have to calculate its expectation value in an s-state
of principal quantum number n (x) and find the positive energy shift
En =

4
ZC|n (0)|2 .
3M 2

(12.522)

Thus, the present rough estimate of the effect of vacuum fluctuations produces the
same term as before in (12.407), except for a different logarithmic factor.
For a hydrogen atom


1
1 3/2

(12.523)
n (0) =
n3 aB
where aB = 1/M is the Bohr radius. If the nuclear charge is Z, then aB , is
diminished by this factor. Thus, we obtain the energy shift
En =

42 Z
1
(MZ)3 3 C.
2
3M
n

(12.524)

For a hydrogen atom with n = 2, this becomes


E2 =

3 2
MC.
6

(12.525)

The quantity M2 is the unit energy of atomic physics determining the hydrogen
spectrum to be En = M2 /2n2 . Thus
M2 = 4.36 1011 erg = 27.21eV = 2 Ry = 2 3.288 1015 Hz (12.526)
Inserting this together with 1/137.036 into (12.532) yields3
E2 135.6MHz C,

(12.527)

667.15 MHz < E2 < 1334.3 MHz.

(12.528)

and thus the estimate

The experimental Lamb shift4 and


ELamb shift 1057 MHz,
is indeed contained in this range.
3
4

The precise value of the Lamb constant 4 M/6 is 135.641 0.004 MHz.
See Notes and References.

(12.529)

864

12.21.2

12 Quantum Electrodynamics

Relativistic Estimate

The above simple estimate of the effect of vacuum fluctuations produces the same
type of correction to the Coulomb potential as the vertex correction in Eq. (12.407).
and the vacuum polarization in (12.457). Those two corrections yield an energy
shift in an s-state n (x) of principal quantum number n:
En =

4
43Z 4 2
2
Z|
(0)|
C
=
MCrel ,
n
rel
3M 2
3n3

(12.530)

with the constant Crel :

M
3 1
.
(12.531)

8 5
For a hydrogen atom in an s-state of principal quantum number n = 2 with n = 2,
this implies
3
E2 =
M2 Crel 135.641 MHz Crel .
(12.532)
6
The result is undetermined since it contains the infrared cutoff in the constant Crel
of Eq. (12.531). In a first approximation, we may imagine the atomic energy M2
to provide an infrared cutoff for the photon energies, and find an estimate for the
energy shift of the 2S1/2 levels with respect to the 2P1/2 levels of the hydrogen atom
which is about 6% smaller than the larger of the non-relativistic estimates (12.528):
Crel log

E2 (1 334.3 51 27.3) MHz. 1 256 MHz

(12.533)

The intermediate expression exhibits the contribution of the terms 3/8 and 1/5 in
(12.531) (thus showing that vacuum polarization gives a negative shift 27.31 MHz).
This shift was first calculated by Uehling citeUHLC11, who initially thought that
vacuum polarization was the main cause of the Lamb shift, and was disappointed to
see it contribute only about 3% to it. In muonic atoms, however, vacuum polarization does produce the dominant contribution to the Lamb shift for a simple reason:
While the above-calculated energy shifts contain in a factor 1/M 2 in formulas such as
(12.583), where M is the mass of the muon, the leading vacuum polarization graph
still involves an electron loop containing the electron mass, thus being enhanced by
a factor (M /Me )2 2102 .

12.21.3

Effect of Wave Functions

In the above calculations the finite size of the electron was derived from a oneloop Feynman-diagram in which the electron lines described free particles. In an
atom, however, the electrons move in a Coulomb potential. The electron is bound
to the nucleus A more accurate calculation should take into account the atomic
wave functions of the electron. This is most simply done in an approximation which
treats the electrons as non-relativistic particles. Such an approximation carries an
intrinsic error caused by the fact that if a non-relativistic electron emits a photon
with energy of the order Mc2 and larger, the recoil will necessarily make the electron

865

12.21 Lamb Shift

relativistic. This error can, however, be avoided by separating the relativistic from
the nonrelativistic contributions. In the first, the effect of the atomic binding of
the electrons is negligible so that the electrons can be treated as free relativistic
particles. In the second, the electrons remain approximately nonrelativistic. There
exists a natural energy scale K = M which is much larger than the atomic energy
M2 , but much smaller than the rest energy of the electron M. The energy scale K
serves to make the separation quantitative. For photons in the upper energy regime,
to be referred to as the hard-photon regime, we may equip the photon with a mass
2 M M, and deduce the Lamb shift from Eq. (12.530) to be
En =

43 Z 4 2
4
2
Z|
(0)|
C
=
MChard ,
n
hard
3M 2
3n3

with the constant


Chard log

M
3 1
.

8 5

(12.534)

(12.535)

The renormalization procedure performed when calculating the vertex correction


and vacuum polarization has removed the ultraviolet divergence.
This leaves us with the task of calculating the contribution from the photons in
the lower energy regime, the soft-photon regime, where the electrons stay nonrelativistic. In the transverse gauge with A(x) = 0, the Hamiltonian has the form
(12.152), with the Coulomb Hamiltonian (12.153). The radiation field (12.154) contains, in the soft-photon regime under consideration, only photon energies within
the limited interval (0, K) where K satisfies M2 K M.
In the transverse gauge with A(x) = 0, the radiation field is expanded in
terms of photon creation and annihilation operators as in Eq. (7.338)
A (x) =

X
k,

h
i
1
eikx (k, )ak, + h.c. .
2V k

(12.536)

The hats on the field operators are displayed, for clarity.


To see what effect we can expect we consider first the influence of the radiation
field upon a free electron.
Effect of Vacuum Oscillations upon Free Electron
To gain experience we first we calculate the size of the energy shift for a free electron.
This is the quantum-mechanical version of the calculation in Subsection 9.10.1.
0 = p
2 /2M, and
The Hamilton operator of a free electron of momentum p is H
the wave functions is a plane wave
1
hx|pi = eipx
V

(12.537)

of energy Ep = p2 /2M. If the electromagnetic field is quantized, there is a ground


state for the fluctuating vector potential, the vacuum state |0i of the photon field.

866

12 Quantum Electrodynamics

The combined state will be denoted by |p; 0i. The calculation is done perturbatively.
Thus we expand the energy shift in powers of the coupling constant:
Ep = Ep(1) + Ep(2) + . . . .

(12.538)

assuming the charge e to be sufficiently small. The first-order shift Ep(1) is simply
the expectation value of the interaction operator
e2 2

int = e p
A(x)
(x).
+ 2 2A
H
Mc
M c

(12.539)

Since this is odd in the field A(x) which has no expectation value, the first-order
energy shift vanishes. Thus we turn to the second-order shift Ep(2) , in which we
may ignore the second term in the interaction (called the seagull term), since it
contributes equally to all atomic levels. Then we have
Ep =

e2
1

A(x)|p;
p
0i.
hp; 0|
p A(x)
2
0
(Mc)
Ep H

(12.540)

Inserting a complete set of intermediate electron-plus-single photon states |p; ki, we


find, in natural units with c = 1, h
= 1,
2

Ep =

e
M2

d3 x

2V |k|

k,k
, =1,2

1
q

2V |k |

p (k, )eikx p  (k , )eik x

|p (k, )|2
1
e2 X
.
k2
M 2 k,=1,2 2V |k| p k

|k|
M
2M
Performing the polarization sum

ih

p2
1

(p k )2 |k |
2M
2M

(12.541)

=1,2

|p (k, )|2 = p2

(p k)2
,
k2

and replacing the sum over momenta by a phase space integral, V


obtain
p2
Ep =
JM (p2 )
2M
with
"
#
e2 Z d3 k 1
(p k)2
1
.
JM (p ) =
1 2 2
3
pk
k2
M (2) |k|
pk

|k|
M
2M
Writing the last factor as
2

pk
k2
1 1 2M M
,
+
pk
k k k2

+k
2M
M

(12.542)
R

d3 k/(2)3 , we
(12.543)

(12.544)

(12.545)

867

12.21 Lamb Shift

we obtain

JM (p2 ) = JM (0) + JM
(p2 )

with

e2
JM (0) =
M

(12.546)

(p k)2 1
d3 k 1
1 2 2
(2)3 k
pk
k
"

(12.547)

and

JM
(p2 ) =

e2
M

(p k)2
d3 k 1
1

(2)3 k
p2 k 2
"

pk
k2

2M
M
.
k2
pk

+k
2M
M

(12.548)

The first term can easily be calculated:


e2 Z Z
d cos (1 cos2 )
dk
4 2 M 0
0
Z
4
dk.
(12.549)
=
3M 0
Being in the soft-photon regime, a quadratic divergence at large k is avoided. The
integral is cut off at k = K << M. With the resulting finite JM (0), the kinetic
energy of the electron is changed from E(p) = p2 /2M to the renormalized energy
JM (0) =

p2
[1 + JM (0)].
(12.550)
2M
Such a factor may be absorbed in the mass of the electron, by defining a renormalized
mass
M
M MJM (0).
(12.551)
MR =
1 + JM (0)
ER (p) =

In terms of this and the subtracted function JM


(p2 ) of Eq. (12.548), we find the full
momentum dependence of the energy to order

ER (p) =

p2

[1 + JM
(p2 )].
R
2MR

(12.552)

Effect of Vacuum Oscillations upon Bound Electron


We now perform the same calculation once more in the presence of the Coulomb potential VC (x) and study an electron in an orbit of principal quantum number n with
a wave function n (x), moving through a photon vacuum |0i. Then Eq. (12.540)
becomes
e2
1

A(x)|n;
En = 2 hn; 0|
p A(x)
p
0i.
(12.553)
C
M
En H
Inserting a complete set of states |n; 0i between the operators in Eq. (12.553) leads
to
En =

Z
peikx )n n  (k, )
peikx )nn (k, )(
1 (
d3 k
e2 X X
M 2 n , (2)3 2V |k|
En En k

(12.554)

868

12 Quantum Electrodynamics

where (
peikx )nn denotes the matrix elements
(
peikx )nn

d3 xn (x)(
peikx )nn n (x).

(12.555)

A Schrodinger wave function correspond graphically to an infinite set of static photon


exchanges. The energy shift to be calculated form Eq. (12.553) has therefore the
following graphical representation: The additional photon provides for a radiative
nucleus

X
n

electron

Figure 12.29 Diagrammatic content in the calculation of the energy shift with the help
of Schr
odinger wave functions. A hydrogen atom is represented by the fat line on the left
which results from an infinite sum of photon exchanges.

correction to the static photons which create the bound state. The sum over n
must, of course, include also an integral over the continuous wave functions. By
rewriting as in (12.545)
1
En En
1
= +
,
En En k
k k(En En + k)

(12.556)

we obtain
En =



e2 X
ikx
ikx

(
p
e
)
p

e
i
nn
j
n n
M 2 n

ki kj
d3 k 1
ij 2
3
2
(2) 2k
k

!

En En
.
1 +
En En + k


(12.557)

The |k|-integration is again restricted to the soft-photon regime |k| < K << M.
In the integral involving only the first term in the brackets of (12.557), the energies
En are absent and we can replace the sum over wave functions |n i by an integral
over plane waves. Then we obtain the expectation value of the energy calculated for
a free electron in Eq. (12.550):

1
hn|p2 |niJM (0).
2M

(12.558)

This produces the same mass renormalization as before. Subtracting this from En
and we are left with
En =

e2 X
pj eikx )n n
(
pi eikx )nn (
2M 2 n

d3 k 1
ki kj
ij 2
3
2
(2) k
k

En En
.
En En + k

(12.559)

869

12.21 Lamb Shift

To understand the behavior of the integral it is useful to split the k-integral in the
soft-photon regime further. Thus we introduce an energy M2 K K = M,
and distinguish an upper regime with wave number K < k < K and a lower regime
with k < K . The corresponding energy shifts are denoted by low En and up En ,
respectively. In the upper regime we may approximate (12.559) by
up En

e2 X
pj eikx )n n (En En ) + (i j)
(
pi eikx )nn (
=
4M 2 n
"

d3 k 1
ki kj
ij 2
3
2
(2) k
k

(12.560)

C
The sum over n can now be expressed with the help of the Hamilton operator H
of the Coulomb system as
C , pj eikx ]]|ni.
pi eikx [H
(En En ) + (i j) = hn|[
n
(12.561)
Working out the commutators, this reduces to
X

pi eikx

nn

pj eikx

n n

C , pj eikx ]] = i j VC (x) + . . . ,
[
pi eikx , [H

(12.562)

where the omitted terms contain a factor ki , kj or both. Due to their longitudinal nature, the do not contribute to (12.560), where they are contracted with the
transverse projection tensor ij ki kj /k 2 . Thus we find
up En

Z
ki kj
d3 k 1
e2
ij 2
hn|

V
(x)|ni
=
i
j
C
2
3
3
4M
k
K>|k|>K (2) k

(12.563)

Doing the integral over all k-directions yields


up En

e2
1 Z K dk
2
=
hn| VC (x)|ni 2
.
4M 2
3 K k

(12.564)

Inserting
2 VC (x) = 4Z (3) (x),

(12.565)

this can be rewritten as


up En =

K
4Z
|n (0)|2 log .
2
3M
K

(12.566)

Note that this expression matches smoothly to the energy shift (12.534) caused by
the hard photons. By adding the two results together, the intermediate energy scale
K cancels producing an energy shift depending only on the separation parameter
K :


M
3 1
4
2
log
.
(12.567)
|
(0)|

(hard + up )En =
n
3M 2
K 8 5
Consider now the lower soft-photon regime part of the integral (12.559). Here
we take advantage of the fact that due to the presence of the atomic probability

870

12 Quantum Electrodynamics

< a = 1/M.
distribution |n (x)|2 , the integration over x is limited to a range |x|
B

Since |k| < K M we see that

|kx| 1,

(12.568)

so that we can neglect the exponential eikx in the matrix elements (


peikx )nn . Thus
we may evaluate the simpler expression
low En

e2 X
(pi )nn (pj )n n
2M 2 n

ki kj
d3 k 1
ij 2
3
2
(2) k
k

En En
.
En En + k

(12.569)

This approximation amounts to neglecting the recoil of the atom. Performing the
integral over all k-directions yields
low En

X
n

|pnn |2
JM (n, n )
2M

(12.570)

with

2 2 Z K
En En
JM (n, n ) =
dk
M 3 0
En En + k
After doing the k-integral, this becomes

(12.571)

K
4
+ i(En En ) .
(En En ) log
JM (n, n )
3M
|En En |
#

"

(12.572)

The imaginary part contributes to the decay rate of an atom from the state n
into a lower state n . It can be dropped in the final formula for the energy shift.
Let us decompose
K
K
En En
log
= log
log
,
En En
2E1S
2E1S

(12.573)

(1)

thus separating (12.570) into two sums low En and low En . The first of these can
be treated the same way as in (12.561) and yields a contribution
(1)

low En =

K
4Z
2
|
(0)|
log
.
n
3M 2
2E1S

(12.574)

Together with the energy shift (12.567), this becomes


(1)

(hard + up + low )En =

4Z
43 Z 4 2
2
|
(0)|
C
=
MC
n
3M 2
3n3

(12.575)

with the constant


3 1
1
3 1
M
= log 2 9.265.
C = log
2E1S
8 5

8 5

(12.576)

871

12.21 Lamb Shift

The separation parameter K has disappeared, and the result is unambiguous.


For the n = 2 -state of the hydrogen atom, the numerical value is
(1)
(hard + up + low )En 135.6 MHz C
(1 334 51) MHz 1 283 MHz,

(12.577)

still larger than the experimental value 1057 MHz. The relativistic treatment of the
hard regime together with the upper soft-photon regime have produced a number
which lies about 4% below the upper rough estimate (12.528).
It remains to calculate the second sum low En in the lower soft-photon regime
which contains the effect of the wave functions in an essential way. This sum is
slightly more involved and requires evaluating a detailed spectral sum. We shall
write it as
2
low En = M 2 Sn
(12.578)
3
where Sn denotes the sum
Sn =

X
n

|pnn |2 (En En ) log

|En En |
.
2E1S

(12.579)

It is convenient to define an average energy E a by the relation


Sn =

"
X
n

|p

nn

E av
| (E En ) log
.
2E1S
2

(12.580)

The bracket is, by virtue of (12.561), equal to


X
n

|pnn |2 (En En ) =
=

1
1
C, p
eikx ]]|ni = hn|2 VC |ni
hn|[
peikx [H
2
2
1
2Z 4
4Z|n (0)|2 =
.
2
n3

(12.581)

This provides a useful check for the convergence of the calculation. An explicit
evaluation of the sum gives for the 2S level [29]
E av = 8.320 2 M.

(12.582)

By writing Sn in the form (12.580), the effect of the correction is to subtract from
C in (12.575) a term log(2E1S /E av ), thus producing the result
En =

43 Z 4 2
4
2
|
(0)|
C
=
MC2 ,
n
2
3M 2
3n3

(12.583)

with
2E1S
1
3 1
2E1S
tot

C2S
= log 2 log av 7.146.
1/2 = C log
av
E

8 5
E

(12.584)

872

12 Quantum Electrodynamics

In combining the hard-photon with the two soft-photon results we have been
a bit careless since the first result (12.534) was derived with a finite photon mass
2 M M as an infrared cutoff parameter. The calculation of the finite
correction (12.585) should be done in the same way, i.e., we should integrate
low En

e2 X
(pi )nn (pj )n n
2M 2 n
Z

d3 k
ki kj
1

ij
(2)3 k 2 + 2
k 2 + 2

En En

. (12.585)
En En + k 2 + 2

The photon mass changes (12.571) into

JM
(n, n )

2
=
M

1 k2
dk 1
3 k 2 + 2

En En

.
En En + k 2 + 2

(12.586)

The difference between the two integrals is


"

5
4
+ + i(En En ) .
(En En ) log
JM (n, n ) =
3M
2|En En | 6
(12.587)
tot
Comparison with (12.572) shows that the constant C2S
in
(12.584)
receives
an
1/2
additional correction of 5/6 log 2 0.140, bringing it up to

JM
(n, n )

tot
C2S
1/2 = log

3 1
2E1S 5
M
log av + log 2 7.286,
2E1S
8 5
E
6

(12.588)

corresponding to an energy shift


E2S 1/2 (1 334.3 51 27.3 287 + 20) MHz 989 MHz,

(12.589)

which is smaller than the experimental value 1 057 MHz.


An important correction is missing in this calculation: that caused by the anomalous magnetic moment of the electron. This produces an energy shift of about 68.5
MHz, which brings the Lamb shift from the value (12.589) up to 1 057.5 MHz, in
good agreement with the experimental number 1 057 MHz.
This contribution will be calculated in the next subsection. Before we come to
that, however, we want to observe that while the relativistic Lamb shift applied only
to s-waves, the effect of the wave functions changes also the energy of states with
orbital angular momenta l 0. For such wave functions, we may define an average
energy analog to (12.580) as
Sn =

"
X
n

|pnn | (En En )

log
l=0

Eav
.
2E1S

(12.590)

The bracket must be taken for l = 0 since it vanishes for l > 0 by virtue of the same
commutator calculation as in (12.581), to be evaluated between l 6= 0-states. By
doing the spectral sum one finds the average energy [30]
av
E2P
= 0.9704 2 M.

(12.591)

873

12.21 Lamb Shift

This raises the p-wave slightly by


av E2P =

3 2
av
MC2P
,
6

(12.592)

av
C2P
= log

2E1S
0.03
av
2E2P

(12.593)

with

i.e., by
av
av E2P 135.6 MHz C2P
4 MHz.

12.21.4

(12.594)

Effect of the Anomalous Magnetic Moment

The relativistic current of the electron was found in Eq. (12.393) to have the form
i
hp |j |pi = e
u(p ) F (q ) +
q G(q 2 ) u(p ),
2M

(12.595)

with the form factors F (q 2 ) and G(q 2 ) given by Eqs. (12.395) and (12.383). For small
momentum transfers, these can be approximated by [recall (12.399) and (12.410)]
F (q 2 ) 1 + q 2 Revc 2 ,

Revc 2

3
,
log
=

2
3M

(12.596)

.
2

G(q 2 )

In configuration space, this amounts to an effective extra electromagnetic interaction


of the Dirac field of the electron which can be written as a Lagrangian density
Leff

= e(x)
1 Revc 2
h

i
i A (x) (x). (12.597)
A (x) +
2M
2


In a static electric field, the radiative corrections provide a solution of the Dirac
equation njm (x) with an additional energy


E = e(x)
0 Revc 2 2 A0 (x) +

i
E(x) (x).
4M


(12.598)

To lowest order in , we approximate the solutions to the Dirac equation by combinations of the nonrelativistic Schrodinger wave functions nlm with rest spinors
u(0, s3 ), combining them to state of total angular momentum j with the help of
Clebsch-Gordan coefficients, as shown in Eq. (6.185). The first term in (12.598)
leads precisely to the relativistic energy shift hard E calculated in (12.534). That
calculation lacked, however, the energy shift due to the second term, which arises
from the anomalous magnetic moment:
E a = i

e
4M

(x)E(x).
d3 x(x)

(12.599)

874

12 Quantum Electrodynamics

In order to calculate this we have to approximate the solutions of the Dirac


equations a little better than up to now. We go to the Dirac representation of
the gamma matrices and decompose the bispinor into two simple spinors, as in
Eq. (6.185),
!
(x)
(x) =
,
(12.600)
(x)
These satisfy the Dirac equation (4.497), and we observed in (4.574), that the lower,
small components are related to the upper, large ones by [see also (6.114)]
(x) = i


(x) + O().
M

(12.601)

Neglecting the corrections of the order of and inserting (12.601) into (12.603),
using the explicit form of the electric field
E(x) = Z

x
,
|x|3

(12.602)

the integrand is found to contain an expression


x (x) = (x)  x (x) (x)  x (x)
(x)
|x|3
|x|3
|x|3
"
#
1
x

(x) 3  (x) + h.c. .


2Mi
|x|

(12.603)

After an integration by parts this leads to


Z

x (x) (x)  ,  x (x).


d x(x)
|x|3
|x|3
"

(12.604)

We now observe that

1
x
= 2
= 4 (3) (x),
3
|x|
|x|

(12.605)

and calculate the commutator using the chain rule


"

 ,

x
|x|3

= 2

1
1
1
2i (x ) 3 = 4 (3) (x) + 4 3 L S. (12.606)
|x|
|x|
|x|

The first term gives rise to another energy shift for s-waves, which may be written
in the same general form as the previous ones:
En =
with the constant being now

4
Z|n (0)|2 C a ,
3M 2
3
Ca .
8

(12.607)

(12.608)

Appendix 12A

Calculation of Dirac Trace for Klein-Nishina formula

875

The second term depends on the angular momentum of the Dirac wave function.
Writing 2L S = J2 L2 S2 , the eigenvalues of L S are j(j + 1) l(l + 1) 3/4.
The expectation value of 1/|x|3 in l 6= 0 -states is
hnlm|

2
1
|nlmi =
(ZM)3 .
3
3
|x|
l(l + 1)(2l + 1)n

(12.609)

Thus we obtain for l 6=-states an energy shift


En =

4
43 Z 4 2
2 a
a
Z|
(0)|
C
=
MC2S
1/2
1/2 ,
n
2S
3M 2
3n3

with

(12.610)

1
1
j
=
l
+
,
3 1
l
+
1
2
a
for
C2S 1/2
(12.611)
1
8 2l + 1

1
j = l ; (l 1).
l
2
Remarkably, the anomalous magnetic moment causes a small energy shift also in
states of higher angular momenta. The 2P 1/2 -state of hydrogen is shifted by

a E2P 1/2 =

3 2
a
MC2P
1/2 ,
6

(12.612)

with

1
a
C2P
(12.613)
1/2 = .
8
Together with the s-wave splitting (12.611) this increases the Lamb shift by
a E =

with

3 2
MC a ,
6

(12.614)

1
C a = ,
2

(12.615)

a E 135.6 MHz C a 68.9 MHz.

(12.616)

i.e., by
Recently, Lamb shifts have been measured also in heavy atoms where the atomic
levels are accessible to X-ray spectoscopy. There the theoretical analysis is much
more difficult due to the narrowness of the atomic wave functions and the large size
of the nucleus [32].

Appendix 12A

Calculation of Dirac Trace for


Klein-Nishina formula

The trace (12.281) can be expanded as




1 t81 + M 2 t61
t82 + M 2 t62
t83 + M 2 t63
t84 + M 2 t64
F = 2
,
+
+
+
M
(2pk)2
(2pk)(2p k ) (2pk)(2pk )
(2pk )2

(12A.1)

876

12 Quantum Electrodynamics

where tij denotes the following Dirac traces involving i gamma matrices:
t81
t83
t61
t63

=
=
=
=

1
k/ / p/ / k/ / p/ ),
4 tr(/
1
k/ / p/ / k/ / p/ ),
4 tr(/
1
k/ / / k/ / ),
4 tr(/
1
k/ / / k/ / ),
4 tr(/

t82
t84
t62
t64

=
=
=
=

1
k/ / p/ / k/ / p/ ),
4 tr(/
1
k/ / p/ / k/ / p/ ),
4 tr(/
1
k/ / / k/ /),
4 tr(/
1
k/ / / k/ /),
4 tr(/

(12A.2)

For brevity of notation, we have omitted the symbols of complex conjugation on the outgoing
polarization vectors, which may be taken to be real and purely spatial, corresponding to linear
polarizations. Traces involving an odd number of gamma matrices have been omitted, since they
vanish. We now use the crossing symmetry (12.278) to rewrite F as


1 t81 + M 2 t61
t82 + M 2 t62

F = 2
+
+ ( , k k ) ,
(12A.3)
M
(2pk)2
(2pk)(2p k )
The traces are now evaluated by the analog of Wicks expansion theorem, Eq. (12.245), together
with the properties
p2 = p2 = M 2 , k 2 = k 2 = 0, 2 = 2 = 1.
(12A.4)
Using further the equalities in the laboratory frame with p = (M, 0, 0, 0):
p
p

= p = 0,
= (p + k k ) = k,

we calculate t81 as follows:


First we use p = 0, 2 = 1, and the Dirac relation (12.241) to reduce
/ p/ / = p/ / / = p/ .

(12A.5)

k/ p/ k/ = k/ k/ p/ + 2(kp)/
k = 2(kp)/
k.

(12A.6)

Similarly, k 2 = 0 allows us to rewrite

Then t81 becomes

1
k/ / p/ ).
t81 = 2(pk) tr(/
4
Now there are only three Wick contractions in the expansion a la (12.245):

(12A.7)

t81 = 2(pk) [( k)( p ) ( )(kp ) + ( p )(k )] ,

(12A.8)

t81 = 2(pk) [2( k)( p ) + (kp )] .

(12A.9)

which yield
With the help of the substitution
kp = 21 [(p k)2 M 2 ] = 21 [(p k )2 M 2 ] = pk ,

(12A.10)

this becomes
t81 = 4(pk)( k)2 + 2(pk )(pk).

(12A.11)

Note that the similar relation


kp = 21 [(p + k)2 M 2 ] = 12 [(p + k )2 M 2 ] = p k ,

(12A.12)

leads to
t84 = 4(pk )(k )2 + 2(pk )(pk),

(12A.13)

Appendix 12A

Calculation of Dirac Trace for Klein-Nishina formula

877

arising from t81 via the crossing operation (12.278). The other traces are
t61 = t64 = 0,

(12A.14)

since they contain in the middle the products k/ / / k/ = k/ k/ = k 2 = 0 and k/ / / k/ = k/ k/ =


k 2 = 0. We further find


t62 = t63 = 41 tr(/
/ / / k/ k/ ) = M 2 2(kk )( )2 2()(k )(k )(kk ) ,
(12A.15)
and finally,

t82 = t83 = 4( )2 (kp)(k p) + 2( )2 (kk )M 2 2( )(k )( k)M 2 2(k )2 (kp)


+ 2(k)2 (k p) (kk )M 2 + 2(kp)(k p).

(12A.16)

Hence

 8

1
1
t1 + M 2 t61
t82 + M 2 t62

( , k k ) =
2( )2
+
2M 2
(2pk)2
(2pk)(2p k )
2M 2



1
2
2

(k ) (kp) + ( k) (k p) (kk )(k p) + ( , k k ) .


+
2(pk)(pk )

The bracket is, explicitly,

1
{(kk )[(kp) (k p)]} ,
2(pk)(pk )

(12A.17)

1
(kk )2 ,
2(pk)(pk )

(12A.18)

kk = k(p + k p ) = kp kp = kp k p.

(12A.19)



(k k)2
1
2
.
+
4(
)
2M 2 (pk)(pk )

(12A.20)

which can be simplified to

using the equation


Thus we obtain [31]

F =

In the laboratory frame, where according to Comptons relation (12.256)




1
1

= M ( ) ,

k k = (1 cos ) = M

and
pk = M ,

pk = M ,

expression (12A.20) reduces to (12.246).


In the electron-positron annihilation process, the trace in Eq. (12.281) becomes


1 t81 M 2 t61
t82 M 2 t62

F = 2
+
+ ( , k k ) ,
M
(2pk)2
(2pk)(2p k )
and we find, after the replacement p = k + k p,


(k k) [(kp) + (k p)]
1
2
.

4(
)
F =
2M 2
(kp)(k p)
Inserting (12.297) and using k p = M and k p = M , this becomes


(k k)2
1
2
4( ) ,
F =
2M 2 4(kp)(k p)
instead of (12A.20), thus yielding Eq. (12.282) which we wanted to derive.

(12A.21)

(12A.22)

(12A.23)

(12A.24)

878

12 Quantum Electrodynamics

Notes and References


More on this subject canbe found in the textbooks
S. Schweber, Relativistic Quantum Fields, Harper and Row, N.Y., N.Y., 1961,
J.D. Bjorken, and S.D. Drell, Vol. I: Relativistic Quantum Mechanics, Vol. II: Relativistic Quantum Fields (McGraw-Hill, New York, 1965). and
C. Itzykson and J.B. Zuber, Quantum Field Theory, McGraw-Hill (1985).
The individual citations refer to:
[1] J.C. Ward, Phys. Rev. 78, 182 (1959).
[2] Y. Takahashi, Nuovo Cimento 6, 371 (1957).
[3] F. Rohrlich, Phys. Rev. 80, 666 (1950).
[4] J. Schwinger, Phys. Rev. 125, 397 (1962); ibid. 128, 2425 (1962).
[5] S. Coleman, R. Jackiw, and L. Susskind, Ann. Phys. 93, 267 (1975);
S. Coleman, ibid. 101, 239 (1976);
C. Adam, Nucl. Phys. B 54, 198 (1997).
[6] G. t Hooft and M.T. Veltman, Nucl. Phys. B 44, 189 (1972).
For earlier related work see:
C. Bollini, J. Giambagi, and A.G. Dominguez, Nuovo Cimento 31, 550 (1964);
C. Bollini and J. Giambagi, Nuovo Cimento B 12 (1972);
P. Breitenlohner and H. Mitter, Nucl. Phys. B 7, 443 (1968);
J.F. Ashmore, Lett. Nuovo Cim. 4, 289 (1972).
[7] See the textbook
H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation,
World Scientific, Singapore 2008, pp. 1-497, (klnrt.de/b11)
[8] See for example
J.D. Jackson, Classical Electrodynamics, Wiley and Sons, New York, 1967.
[9] W. Gordon, Ann. d. Phys. 2, 1031 (1929).
They can be found very simply by group theory, as shown in
H. Kleinert, Fortschr. Phys. 6, 1, (1968), and Group Dynamics of the Hydrogen Atom,
Lectures presented at the 1967 Boulder Summer School, published in Lectures in Theoretical
Physics, Vol. X B, pp. 427482, ed. by A.O. Barut and W.E. Brittin, Gordon and Breach,
New York, 1968.
[10] For a review of the experimental situation see the internet page http://int.phys.
washington.edu/~int talk/WorkShops/int 02 3/People/Vetter P/pstalk.pdf.
[11] G.S. Atkins and E.D. Pfahl, Phys. Rev. A 59, R915 (1999);
G.S. Atkins and F.R. Brown, Phys. Rev. A 28, 1164 (1983).
[12] For a review of the experimental situation see the internet page http://int.phys.
washington.edu/~int talk/WorkShops/int 02 3/People/Vetter P/pstalk.pdf.
[13] See
R.S. Van Dyck, Jr., in Quantum Electrodynamics, ed. by T. Kinoshita, World Scientific,
Singapore, 1990, p. 322.
[14] The numerical value uses the fine-structure constant 1 = 137.035 997 9(3 2) found with
the help of the quantum Hall effect. See M.E. Cage et al., IEEE Trans. Instrum. Meas. 38 ,
284 (1989).
[15] J. Schwinger, Phys. Rev. 73, 416 (1948).

Notes and References

879

[16] For a discussion see the review


J. Gailey and E. Picasso, Progr. Nucl. Phys. 12 , Part 1, eds. D.M. Brink and J.H. Mulvey,
Pergamon, Oxford, 1970, pp. 43-.
[17] The calculation of c2 was performed in the Feynman gauge by
A. Petermann, Helv. Phys. Acta 30, 407 (1957);
C.M. Sommerfield, Phys. Rev. 107, 328 (1957);
and in another gauge by
G.S. Adkins, Phys. Rev. D 39, 3798 (1989).
The number for c3 was published in T. Kinoshita, IEEE Trans. Instrum. Meas. 38, 172
(1989).
See also the review articles by T. Kinoshita as well as by
R.Z. Roskies, E. Remiddi, and M.J. Levine, in Quantum Electrodynamics, ed. by T. Kinoshita, World Scientific, Singapore, 1990, p. 218 and p. 162, respectively.
[18] T. Kinoshita in Quantum Electrodynamics, ed. by T. Kinoshita, World Scientific, Singapore,
1990, p. 419; A. Czarnecki and W.J. Marciano (hep-ph/0102122).
[19] The small errors are estimated by
M. Davier, Nucl. Phys. B (Proc. Suppl.) 76, 327 (1999) (hep-ex/9912044).
The second, more conservative, estimate is due to
F. Jegerlehner, in Radiative Corrections, edited by J. Sol`
a (World Scientific, Singapore,
1999), pp. 7589, and in seminar at New York University in honor of A. Sirlins 70th
Birthday.
[20] R. Jackiw and S. Weinberg, Phys. Rev. D 5, 2473 (1972);
W.A. Bardeen, R. Gastmans, and B.E. Lautrup, Nucl. Phys. B 46, 315 (1972).
[21] J. Bailey et al., Phys. Lett. B 68, 191 (1977);
F.J.M. Farley and E. Picasso, in Quantum Electrodynamics, ed. by T. Kinoshita, World
Scientific, Singapore, 1990, p. 479.
H.N. Brown at al. Phys. Rev. Lett. 86, 2227 (2001) (hep-ex/0102017);
G.W. Bennett at al., Phys. Rev. Lett. 89, 101804 (2002); Erratum-ibid. 89, 129903 (2002)
(hep-ex/0208001).
See also http://phyppro1.phy.bnl.gov/g2muon.
[22] W. Gordon, Ann. d. Phys. 2 , 1031 (1929). They can be found very simply by group theory,
as shown by H. Kleinert, Fortschr. Phys. 6 , 1, (1968), and Group Dynamics of the Hydrogen
Atom, Lectures presented at the 1967 Boulder Summer School, published in Lectures in
Theoretical Physics, Vol. X B, pp. 427482, ed. by A.O. Barut and W.E. Brittin, Gordon
and Breach, New York, 1968.
[23] For details the reader is referred to Bernard de Wits lecture on this subject in
http://www.phys.uu.nl/~bdewit/ftip/AppendixE.pdf.
[24] The Schwinger model with a nonzero initial electron mass is discussed in Ref. [5]
[25] Y. Takahashi, Nuovo Cimento 6, 370 (1957).
[26] H.A. Bethe, Phys. Rev. 72 , 339 (1947); J. Schwinger and V. Weisskopf, Phys. Rev. 73 ,
1272A (1948); R.P. Feynman, Phys. Rev. 76 , 939 (1948); N.M. Kroll and W.E. Lamb, Phys.
Rev. , 388 (1949); J.B. French and V. Weisskopf, Phys. Rev. , 1240 (1949); H.A. Bethe,
L.M. Brown, and J.R. Stehn, Phys. Rev. 77 , 370 (1950); A.J. Layzer, Phys. Rev. Lett. 4 ,
580 (1960); G.W. Erickson and D.R. Yennie, Annals of Phys. 35 , 271, 447 (1965); G.W.
Erickson, Phys. Rev. Lett. 27 , 780 (1971); P.J. Mohr, Phys. Rev. Lett. 34 , 1050 (19).
[27] The first experiment was done by
W.E. Lamb and R.C. Retherford, Phys. Rev. 72, 241 (1947); 79, 549 (1950); 81, 222 (1951);

880

12 Quantum Electrodynamics
86, 1014 (1951).
See also the review
W.E. Lamb, Rep. Progr. Phys. 14, 19 (1951).
A more accurate value E = 1 057.8 0.1 MHz was found in
S. Triebwasser, E.E. Dayhoff, and W.E. Lamb, Phys. Rev. 89, 98 (1953).

[28] E.A. Uehling, Phys. Rev. 49, 55 (1935).


[29] See
H.A. Bethe and E.E. Salpeter in Encyclopedia of Physics (Handbuch der Physik), Springer,
Berlin, 1957, p. 405.
[30] For more average energy values see
J.M. Harriman, Phys. Rev. 101, 594 (1956).
[31] The calculation can actually be performed with a simple reduce program which can be
downloaded from http://www.physik.fuberlin.de/~kleinert/b6/programs/reduce.
[32] For a review see
P.J. Mohr, G. Plunien, and G. Seff, QED-Corrections in Heavy Atoms, Phys. Rep 293, 227
(1998).

Perfection is achieved, not when there is nothing more to add,


but when there is nothing left to take away
Antoine de Saint-Exupery (19001944)

13
Functional Integral Representation
of Quantum Field Theory
In theoretical physics, Fourier transformations have always played an important role
in yielding complementary insights into mathematical structures. This is also true
for generating functionals.

13.1

Functional Fourier Transform

Our goal is to study the functional Fourier transform of the generating functional
Z[j] of a scalar field theory. This will be written as

Z[]

Dj(x)Z[j]ei

dD x j(x)(x)

(13.1)

where the symbol Dj(x) is called a functional integral.


The associated mathematics is an own discipline and discussed in many
textbooks.1 Functional integrals can be used to represent physical amplitudes without the use of operators in terms of fluctuating time dependencies of dynamical
variables q(t). They are defined by grating space-time into a fine lattice. For every
coordinate x , we introduce discrete lattice points
R

x xn n,

n = 0, 1, 2, 3,

(13.2)

with a very small lattice spacing . Then we may write


Z

dD x j(x)(x) = D

X
n

j(xn )(xn ) D

jn n

(13.3)

where n is to be read as a RD-dimensional index (n0 , n1 , . . . , nD ), one for every


coordinate. Now we define Dj(x) as the infinite product of integrals over jn at
every point xn :
Z
YZ
djn
q
.
(13.4)
Dj(x) =
n
2i/D
1

See, for example, H. Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer
Physics, and FinancialMarkets, World Scientific Publishing Co., Singapore 2004, Third extended
edition, pp. 11468, and references therein.

881

882

13 Functional Integral Representation of Quantum Field Theory

Operations with functional integrals are very similar to those with ordinary integral.
For example, the inverse Fourier transform reads
Z

Z[j] =

D(x)Z[]e

dD x j(x)(x)

(13.5)

There are analogs of the Dirac -function:


Z

Dj(x) ei
D(x) ei

dD x j(x)(x)

= [],

(13.6)

dD x j(x)(x)

= [j],

(13.7)

called -functionals. In the lattice approximation corresponding to (13.4), they are


defined as infinite product of ordinary -functions
[j] =

Yq

2i/D (jn ).

(13.8)

They have the obvious property


Z

D [] = 1,

Dj [j] = 1.

(13.9)

A commonly-used notation for the measure (13.4) of functional integrals employs


a continuously infinite product of integrals which must be imagined
as the continuum
R
limit of the lattice product (13.4). In this notation one writes Dj(x) simply as
Z

Dj(x) =

YZ
x

dj(x)

,
2i

(13.10)

and the associated -functionals as


[j] =

2i (j(x)).

(13.11)

13.2

Gaussian Functional Integral

Only very few functional integrals can be solved explicitly. The simplest nontrivial
example is the Gaussian integral2
Z

Dj(x)e 2

dD x dD x j(x)M (x,x )j(x )

(13.12)

In the discretized form, this can be written as

YZ
n

djn
q

2i/D

i 2D

e 2

j M
j
n,m n nm m

(13.13)

Mathematically speaking, integrals with an imaginary quadratic exponent are more accurately
called Fresnel integrals, but field theorists do not make this distinction.

883

13.2 Gaussian Functional Integral

We may assume M to be a real symmetric functional matrix, since its antisymmetric


part would not contribute to (13.12). Such a matrix may be diagonalized by a
rotation
jn jn = Rnm jm

(13.14)

which leaves the measure of integration invariant


(j1 , . . . , jn )
= det R1 = 1.
(j1 , . . . , jn )

(13.15)

In the diagonal form, the multiple integral (13.13) factorizes into a product of Gaussian integrals, which are easily calculated:
YZ
n

djn

i 2D

2i/D

e 2

j M j
n n n n

Y
n

1
= det 1/2 (D M).
D Mn

(13.16)

On the right-hand side we have used the fact that the product of diagonal values
Mn is equal to the determinant of M. The final result

YZ

djn

2i/D

i 2D

e 2

j M j
n,m n nm m

= det 1/2 (D M)

(13.17)

is invariant under rotations, so that it holds also without diagonalizing the matrix.
This formula can be taken to the continuum limit of infinitely fine gratings 0.
Recall the well-known matrix formula
det A = elog det A = etr log A ,

(13.18)

and expand tr log A into power series as follows


tr log A = tr log [1 + (A 1)] = tr

()k
(A 1)k
k
k=1

(13.19)

The advantage of this expansion is that when inserting for A the matrix D M, the
traces of powers of D M remain well-defined objects in the continuum limit 0:
D

tr M =

X
n

tr D M

2

= 2D

Mnn

n,m

..
.

dD x M(x, x) TrM

Mnm Mmm

(13.20)

dD x dD x M(x, x )M(x , x) TrM 2 , (13.21)

We therefore rewrite the right-hand side of (13.17) as exp[(1/2)tr log(D M)], and
expand
h

tr log(D M) = tr log 1 + D M 1

()k
k=1 k

i
h

dD x1 dD xk M(x1 , x2 ) (D) (x1 x2 )


ih

M(x2 , x3 ) (D) (x2 x3 ) M(xk , x1 ) (D) (xk x1 ) . (13.22)

884

13 Functional Integral Representation of Quantum Field Theory

The expansion on the right-hand side defines the trace of logarithm of a the functional matrix M(x, x ), and will be denoted by Tr log(M). This, in turn, serves
to define the functional determinant of M(x, x ) by generalizing formula (13.18)
to functional matrices:
Det (M) = elog det (M ) = eTr log(M ) .

(13.23)

Thus we obtain for the functional integral (13.12) the result as


Z

Dj(x)e 2

dD x dD x j(x)M (x,x )j(x )

= Det 1/2 (M) = e 2 Tr log(M ) .

(13.24)

This formula can be generalized to complex integration variables


by separating
the currents into real and imaginary parts, j(x) = [j1 (x) + ij2 (x)]/ 2. Each integral
gives the same functional determinant so that
Z

Dj (x)Dj(x)ei

dD x dD x j (x)M (x,x )j(x )

= eTr log M ,

(13.25)

where M(x, x ) is now an arbitrary hermitian matrix, and the measure of integration
for complex j(x) is defined as the product of the measures for real and imaginary
parts: Dj (x)Dj(x) Dj1 (x)Dj2 (x).

13.3

Functional Formulation for Free Quantum Fields

The functional integrals for free fields can all be done with the help of the Gaussian
integral formulas (13.24) and (13.25), which arise for real fields in the form
Z

D(x)e 2

dD x dD x (x)M (x,x )(x )

= (Det M) 2 ,

(13.26)

where and a real symmetric functional matrix


M(x, x ) is a real symmetric functional
matrix. For complex fields = (1 + i2 )/ 2, they arise in the form

D (x)D(x)ei

dD x dD x (x)M (x,x )(x )

= Det 1 M,

(13.27)

where M(x, x ) is a hermitian functional matrix. The complex measure of integration


is defined by D (x)D(x) D1 (x)D2 (x).
Expression (13.39) with the normalization factor (13.41) may be used to define
a free quantum field theory by functional integration,
Z0 [j] = Det

1/2

(iG0 )

D(x)ei

dD x [L0 (,)+j(x)(x)]

(13.28)

and the integrand in Eq. (13.12) with M(x, y) = iG0 (x, x ) is the result of this
functional integral:
1

Z0 [j] = e 2

dD x dD x j(x)G0 (x,x )j(x )

(13.29)

885

13.3 Functional Formulation for Free Quantum Fields

The Gaussian functional integrals in Eqs. (13.24) and (13.25) can be used to
calculate the Fourier transform of the generating functional Z0 [j] of the free theory,
which was defined in (7.818) and (7.821) and calculated explicitly for real scalar
fields in Eq. (7.829),
1

Z0 [j] = e 2

dD x dD x j(x)G0 (x,x )j(x )

(13.30)

(13.31)

and for complex scalar fields in Eq. (7.829):


Z0 [j] = e

dD x dD x j (x)G0 (x,x )j(x )

If we omit the spacetime indices, and use an obvious functional vector notation,
then Z0 [j] can be written in the case of real fields as
A0 [] =

i T 1
G j+ij T
i 0

Dje 2 j

(13.32)

The exponent may be completed quadratically as

T 1


i
i T
1
j + iG1

G
j
+
iG

+ T iG1
0
0
0
0 .
2
i
2

(13.33)

We now replace the variable j + iG1


0 by j which in each of the infinite integrals
amounts only to a trivial shift of the center of integration. Thus

A0 [] =

Z

2i j 1i G0 j

Dj e

e2

T iG1
0

(13.34)

We can now apply formula (13.25) and find


A0 [] = Det (D)

1/2

i
2

dD x dD x (x)D(x,x )(x )

(13.35)

where the functional matrix in the exponent is the inverse free-particle Green function:

2
2 (D)
D(x, x ) = iG1
(x x ).
(13.36)
0 (x, x ) = ( m )
The determinant is a constant prefactor which does not depend on the field . We
shall abbreviate it as a normalization factor
N = Det 1/2 (D) = Det 1/2 (iG0 ).

(13.37)

Using the explicit form (13.36), we can rewrite (13.35) after a partial integration of
the derivative term as
i

w0 [] = N e

dD x

2
1
()2 m2 2
2

= N ei

dD x L0 (,)

(13.38)

Thus the Fourier-transformed generating functional is, up to the normalization factor N , just the exponential of the classical action under consideration.

886

13 Functional Integral Representation of Quantum Field Theory

By Fourier-transforming (13.38), we undo the first transformation and return to


the generating functional Z0 [j], thereby obtaining a functional integral representation for the generating functional
Z0 [j] = N

D(x)ei

dD x [L0 (,)+j(x)(x)]

(13.39)

The functional integral over the -field is defined as in (13.4) by the product of
integrals
Z

D(x) =

dn

YZ

2i/D

(13.40)

Let us check the normalization of (13.39). At zero source, Z0 [j] has to be equal
to unity [recall (13.30)]. This is ensured if
N

= Det

1/2

(iG0 ) =

Dei

dD x L0 (,)

(13.41)

Indeed, by rewriting the right-hand side as


Z

D(x)e

dD x

2
1
()2 m2 2
2

Dei

dD x (x)D(x,x )(x )

(13.42)

we can apply the Gaussian formula (13.24) and see that (13.18) is true due to (13.37).
With (13.41) we can rewrite the expression (13.38) for the Fourier transform w0 []
as
w0 [] = R

ei

dD x L0 (,)

D(x)ei

This ratio is obviously normalized:


Z

dD x L0 (,)

Dw0[] = 1,

(13.43)

(13.44)

and has precisely the form of the quantum mechanical version (1.490) of the thermodynamical Gibbs distribution (1.488).
1
wn ZQM
(tb ta )eiEn (tb ta )/h .

(13.45)

There exists useful formula for harmonically fluctuating fields which is encountered in many physical observables and which can be derived immediately from this.
Consider the correlation function of two exponentials of a free field (x). Inserting
into Eq. (13.58) a special current
j12 (x) = a

dD x[a(x x1 ) b(x x2 )],

(13.46)

887

13.4 Interactions

the partition function (13.28) reads


Z0 [j12 ] = Det

1/2

(iG0 )

D(x) eia(x1 ) eib(x2 ) ei

As such it coincides with the harmonic expectation value

dD x [L0 (,)]

heia(x1 ) eib(x2 ) i.

. (13.47)

(13.48)

Inserting the special current (13.46) into Eq. (13.29), we obtain the equation
heia(x1 ) eib(x2 ) i = e 2 [a
1

2 G(x

2
1 ,x1 )2abG(x1 ,x2 )+b G(x2 ,x2 )

(13.49)

In the brackest of the exponent we recognize the expectation values of pairs of fields
a2 h(x1 )(x1 )i 2abh(x1 )(x2 )i + b2 h(x2 )(x2 )i,

(13.50)

such that we may write (13.49) also as


1

heia(x1 ) eib(x2 ) i = heia(x1 )ib(x2 ) i = e 2 h[a(x1 )b(x2 )] i .

(13.51)

This result may also be derived for field operators using Wicks theorem of Subsection 7.14.1.

13.4

Interactions

Let us now include interactions. We have seen in Eq. (10.24) that the generating functional in the interaction picture [more precisely the functional ZD [j] of
Eq. (10.22)] may simply be written as
Z[j] = ei

dD x Lint (i/j(x))

Z0 [j].

(13.52)

This can immediately be Fourier-tranformed to


A[] =

dD x j(x) (x)

dD x Lint ()

Dj ei

h R

ei

dD x Lint ()(i/j)

Z0 [j] .

(13.53)

Removing the second exponential by a partial functional integration, we obtain


A[] = ei
= ei
= R

dD x Lint (,)

ei

Djei
N ei

dD x L(,)

D(x)ei

dD x j(x)(x)

dD x

dD x L0 (,)

1
2

Z0 [j]

[()2 m2 2 ]

(13.54)

The Fourier transform of this expression renders the generalization of (13.39), the
functional integral representation for the generating function of the interacting theory:
Z[j] = N
=

D(x)ei

Dei

dD x [L(,)+j(x)(x)]

dD x [L(,)+j(x)(x)]

De

dD x L0 (,)()

(13.55)

888

13 Functional Integral Representation of Quantum Field Theory

This representation is to be compared with the perturbation theoretic formula


of operator quantum field theory
Z[j] = h0|T ei

dD x [Lint ()+j(x)(x)]

|0i,

(13.56)

where the symbol (x) denotes free-field operators, and the vacuum expectation
value of products of these fields proceeds following Wicks theorem. In the new
integral representations (13.53), (13.55), on the other hand, is a classical c-number
object. All quantum properties of Z[j] arise from the infinitely many integrals over
, one at every space time point, rather than from field operators.
Note that in formula (13.55), the full action appears in the exponent, whereas
in (13.56) only the interacting part appears.
In contrast to w0 [] of Eqs. (13.43) and (13.44), the amplitude for the interacting
theory is no longer properly normalized. In fact, we know from the perturbative
evaluation of (13.56) that it represents the sum of all vacuum diagrams displayed in
(10.70). Since the denominator does not normalize A[] anyhow, it is convenient y
to drop it and work with the numerator only, using the unnormalized
Z[j] =

D(x)ei

dD x [L(,)+j(x)(x)]

(13.57)

as the generating functional.


The normalization in (13.57) has an important advantage over the previous one
in (13.55). In the euclidean formulation of the theory to be discussed in Section 13.5,
it makes Z[0] equal to the thermodynamic partition function of the system.
For free fields, Z[0] is equal to the partition function of a set of harmonic oscillators of frequencies (k) for all momenta k. This statement can be proved only in
the lattice version of the theory. In the continuum limit the statement is nontrivial,
since the determinants on the right-hand sides are infinite. However, we shall see
in Section 13.7, Eqs. (13.122)(13.132). that correct finite partition functions are
obtained if the infinities are removed by the method of dimensional regularization
that was used before in Section 11.5 to remove divergences from Feynman integrals.
Even though the operator formula (13.56) and the functional integral formula
(13.57) are completely equivalent, there are important advantages of the latter. In
some theories it may be difficult to find a canonically quantized set of free fields
on which to construct an interaction representation for Z[j] as in Eq. (13.56). The
photon field is an important example where it was quite hard to interprete the
Hilbert space. In particular, we encountered the problem that in the Gupta-Bleuler
quantization scheme the vacuum energy contained the quanta of two unphysical
polarization states of the photon. Within the functional approach, this will easily
be explained in Chapter 18.
A second and very important advantage is the possibility of deriving from the
representation (13.57) directly the Feynman rules without any knowledge of the
Hilbert space. For this we simply take the non-quadratic piece of the action, which

889

13.4 Interactions

defines the vertices of the perturbation expansion, outside the functional integral as
in (13.52), i.e., we rewrite (13.57) as:
Z[j] = ei

dD x Lint (i/j(x))

D(x)ei

dD x [L0 (,)+j(x)(x)]

(13.58)

Using (13.35), this reads more explicitly


i

Z[j] = e

dD x Lint (i/j(x))

i
D(x)e[ 2

dD x dD x (x)iG1
0 (x,x )(x )+j(x)(x)]

. (13.59)

A shift in the field variables to


(x) = (x) +

dy G0 (x, x )j(x ),

(13.60)

and a quadratic completion lead to


Z[j] = ei

dD x Lint (i/j(x)) 21

D(x)e 2

dD x dD x j(x)G0 (x,x )j(x )


dD x dD x (x)iG1
0 (x,x ) (x )

(13.61)

The functional integral over the shifted field (x) can now be performed with the
help of formula (13.24), inserting there M(x, x ) = iG0 (x, x ). The result is, recalling (13.37),
Z

D (x)e 2


dD x dD x (x)iG1
0 (x,x ) (x )

= Det 1/2 (iG0 ),

(13.62)

such that we find


Z[j] = Det

1/2

(iG0 ) e

dD x Lint (i/j(x)) 12

dD x dD x j(x)G0 (x,x )j(x )

. (13.63)

Expanding the prefactor in (13.59) in a power series yields all terms of the
perturbation expansion (10.29) which correspond to the Wick contractions in Section 10.3.1, and to the associated Fenyman diagrams. The free-field propagators
is the functional inverse operator between the fields in the quadratic part of the
Lagrangian. If this is written as
i
2

dD x dD x (x)D(x, x )(x )

(13.64)

G0 (x, x ) = iD 1 (x, x ).

(13.65)

then

This formal advantage of obtaining perturbation expansions from the functional


integral representations has far-reaching consequences. We have seen in Chapter 11
that the evaluation of the perturbation series proceeds most conveniently by Wickrotating all energy integrations to run along the imaginary axes in the complex
energy plane. In this way one avoids the singularities in the propagators at the

890

13 Functional Integral Representation of Quantum Field Theory

physical particle energies. In Section 10.7, on the other hand, these singularities
were shown to be responsible for the fact that particles leave a scattering region
and form asymptotic states. Thus Wick-rotated perturbation expansions describe a
theory which does not possess any particles states. In fact, they cannot be described
by field operators and a Hilbert space.
These expansions can, however, be obtained from a simple modification of the
above functional integral representation of the generating functional Z[j]. We simply
perform the x-space version of the Wick-rotation in Fig. 7.2.

13.5

Euclidean Quantum Field Theory

In Eq. (7.129), we replaced in D = 4 spacetime dimensions the coordinates x


( = 0, 1, 2, 3) by the euclidean coordinates xE = (x1 , . . . , x3 , x4 = ix0 ). The same
operation is done to the time x0 in any dimensions. Under this, the action
A

d x L(, E )

d x

("

1
m2 2
()2
+ Lint ()
2
2
#

(13.66)

goes over into i times the euclidean action


AE =

d xE

("

1
m2 2
(E )2 +
+ Lint () .
2
2
#

(13.67)

The euclidean versions of the Gaussian integral formulas (13.26) and (13.88) are
as
Z

12

D(x)e

dD x dD x (x)M (x,x )(x )

= (Det M) 2 ,

(13.68)

and
Z

D (x)D(x)e

dD x dD x (x)M (x,x )(x )

= (Det M)1 .

(13.69)

for complex fields = (1 + i2 )/ 2, where D (x)D(x) D1 (x)D2 (x).


The amplitude (13.54) becomes
wE [] = R

dD xE LE (,)

dD xE LE (,)

D(x)e

The normalized version of this

w[] = R

D(x)e

dD xE LE 0 (,)

dD xE LE (,)

(13.70)

(13.71)

represents the functional version of the proper quantum statistical Gibbs distribution
corresponding to (13.45) [recall (1.488)]:
eEn /kB T
wn = P En /k T .
B
ne

(13.72)

891

13.6 Functional Integral Representation for Fermions

The functional integral representation for the unnormalized generating functional


of all Wick-rotated Green functions corresponding to (13.57) is then
Z

ZE [j] =

D(x)e

dD xE [LE (,E )j(x)(x)]

(13.73)

The euclidean action corresponds to an energy of a field configuration. The


integrand plays the role of a Boltzmann factor and gives the relative probability for
this configuration to occur in a thermodynamic ensemble.
We now understand the advantage of working with the unnormalized functional
integral: At zero external source, ZE [j] corresponds precisely to the thermodynamic
partition function of the system. This will be seen explicitly in the examples in
Section 13.7.

13.6

Functional Integral Representation for Fermions

If we want to use the functional technique to describe also the statistical properties
of fermions, some modifications are necessary. Then the fields must be taken to be
anticommuting c-numbers In mathematics, such objects form a Grassmann algebra
G. If , are real elements of G, then
= .

(13.74)

A trivial consequence of this condition is that the square of each Grassmann element
vanishes, i.e., 2 = 0. If = 1 + i2 is a complex element of G, then 2 = =
2i1 2 is nonzero, but ( )2 = ()2 = 0.
All properties of operator quantum field theory for fermions can be derived from
functional integrals if we find an appropriate extension of the integral formulas
in the previous sections to Grassmann variables. Integrals are linear functional.
For Grassmann variables, these are completely determined from the following basic
integration rules:
Z

d
0,
2

d
1,
2

d
n 0, n > 1,
2

(13.75)

for real and


d
d
d d
1,
1,
( )n n1 . (13.76)
2
2
2 2

for complex = (1 + i2 )/ 2, where dd = id1 d2 .


Note that these integration rules make the linear operation of integration in
(13.75) coincide with the linear operation of differentiation. A function F () of a
real Grassmann variable , is determined by only two parameters: the zeroth- and
the first-order Taylor coefficients. Indeed, due to the property 2 = 0, the Taylor
Z

d
0,
2

892

13 Functional Integral Representation of Quantum Field Theory

series has only two terms F () = F0 + F , where F0 = F (0) and F dF ()/d.


But according to (13.75), also the integral gives F :
Z

d
F () = F .
2

(13.77)

The coincidence of integration and differentiation has the important consequence


that changes in integration variable do not transform with the Jacobian but with
the inverse Jacobian:
Z
Z
d
d(a)
=a ,
(13.78)
2
2
We shall use this transformation property below in Eq. (13.84).
As far as perturbation theory is concerned, it is sufficient to define only Gaussian
functional integrals such as (13.26) and (13.88). In the discretized form, we may
derive the formula
"
Y Z
n

d
i 2D X
n

m Mmn n = det 1/2 (D M).


exp
2
2iD
m,n

(13.79)

The right-hand side is the inverse of the bosonic result (13.26). In addition, there
is an important difference: only the antisymmetric part of the functional matrix
contributes.
If the matrix Mmn is hermitian, complex Grassmann variables are necessary to
produce a nonzero Gaussian integral. For complex variables we have
"
Y Z
n

X
dn dn

exp i2D
m
Mmn n = det (D M),
D
D
2i 2i
m,n
#

(13.80)

the right-hand side being again the inverse of the corresponding bosonic result
(13.88):
We first prove the latter formula. After bringing the matrix Mmn to diagonal
form via a unitary transformation, we obtain the product of integrals
"
Y Z
n

X
dn dn

exp i2D
n Mn n .
2iD 2iD
n
#

(13.81)

Expanding the exponentials into a power series leaves only the first two terms since
(n n )2 = 0, so that the integral reduces to
YZ
m

dn dn

(1 + i2D n Mn n ).
D
D
2i 2i

(13.82)

Each of these integrals is performed via formulas (13.76), and we obtain the product
of eigenvalues Mn , which is the determinant:
Y
m

Mn = det M.

(13.83)

893

13.6 Functional Integral Representation for Fermions

For real fermion fields, we observe that an arbitrary real antisymmetric matrix
Mmn can always be brought by a real orthogonal transformation T to a canonical
form C which is zero except for 2 2 matrices c = i 2 along the diagonal. Thus
M = T T CT . The matrix C has a unit determinant so that det T = det 1/2 (M).

Let m
Tmn n , then the measure of integration in (13.79) changes according to
(13.78) as follows:
Y
Y
dn = det T
dn .
(13.84)
n

Applying now formulas (13.75), the Grassmannian functional integral (13.79) can
be evaluated as follows:
"
Y Z
m

#
#
!
"

X
X
Y Z
dn
d
n
2D

exp i
exp
m Cmn n
m Mmn n = det T
2iD
2iD
m,n
m
k,l

= det 1/2 (iD M).

(13.85)

The integrals over n decompose into a product of two-dimensional Grassmannian


integrals involving the antisymmetric unit marix c = i2 . The have the generic form
"
Y Z
n



d
d

2n
2n+1 1 + i2D 2n
2n+1
= 2D .
D
D
2i
2i
#

(13.86)

There is one such factor for every second lattice site, which changes det T = det 1/2 M
into det 1/2 (D M), thus proving (13.79).
In the continuum limit, the result of this discussion can be summarized in an
extension of the Gaussian functional integral formulas (13.26) and (13.25) to
Z

D(x)e 2

dD x dD x (x)M (x,x )(x )

= Det 2 M,

(13.87)

where M(x, x ) is real symmetric or antisymmetric for bosons or fermions, respectively, and
Z

D (x)D(x)ei

dD x dD x (x)M (x,x )(x )

= Det 1 M.

(13.88)

where the matrix is hermitian.


The functional integral formulation of fermions follows now closely that of bosons.
For N relativistic real fermion fields a obtain an amplitude A[a ] from the Fourier
franform
A[] =

Dj(x)Z[j]ei

dD x ja (x)a (x)

(13.89)

and find
A[] = hR

ei

dD x L(,)

Dei

dD x L0 (,)

i.

(13.90)

894

13 Functional Integral Representation of Quantum Field Theory

The functional
Y Z

Z[j] = N

Da ei

dD x L(,)

dD x L0 (,)

(13.91)

with a normalization factor


N

Y Z
a

Da e

(13.92)

provides us with a functional integral representation of the generating functional of


all Green functions.
An obvious extension of this holds for complex fermion fields In the case of a
Dirac field, for example, where the sources j are commonly denoted by , we obtain
Z[, ] = N

DD e

/ m)+Lint ()+
]
+
dD x [(i

(13.93)

with
N

DD ei(i / m) = Det iG1


m).
0 (x, x ) = Det (i/

(13.94)

As in the boson case, we shall from now on work with the unnormalized functional
without the factor N ,
Z[, ] =

De

/ m)+Lint ()+
]
+
dD x [(i

(13.95)

which has also her the advantage that the euclidean version of Z[0, 0] becomes
directly the thermodynamic partition function of the system.
The functional representations of the generating functionals can of course be
continued to a euclidean form as in Section 13.5, thereby replacing operator quantum
physics by statisticla physics. The corresponding Gaussian formulas for boson and
fermion fields are the obvious generalization of Eqs. (13.87) and (13.88):
Z

D(x)e 2

dD x dD x (x)M (x,x )(x )

= Det 2 M,

(13.96)

and
Z

D(x)D(x)e

dD x dD x (x)M (x,x )(x )

= Det 1 M,

(13.97)

where we have defined the measure of the euclidean functional integrals in the same
way as before in (13.40) and (13.79), except without the factors i under the square
roots. The euclidean version of the generating functional (13.93) renders all Wickrotated Green functions.

13.7 Relation Between Z [j ] and Partition Function

895

As a side result of the above development we can note the following a functional
formulas known as the so-called Hubbard-Stratonovich transformation
Z

D(x, t)e 2

d3 xdtd3 x dt [(x,t)A(x,t;x ,t )(x ,t )+2j(x,t)(x,t)3 (xx ,t)(tt )]

=e
Z

n o

i( 2i Trlog 1i A) 2i

D (x, t)D(x, t)ei

d3 xdtd3 x dt

= ei(iTrlogA)i

d3 xdtd3 x dt j(x,t)A1 (x,t;x ,t )j(x ,t )

(13.98)
}

(x,t)A(x,t;x ,t )(x,t )+[ (x,t)(x)3 (xx )(tt )+c.c.]

d3 xdtd3 x dt (x,t)A1 (x,t;x ,t )(x ,t )

(13.99)

These integration formulas will be needed repeatedly in the remainder of this text.
They are the basis for the treatment of any interacting quantum field theory in
terms of collective quantum fields.

13.7

Relation Between Z[j] and Partition Function

The introduction of the unnormalized functional integral representation (13.57) for


Z[j] was motivated by the fact that in the euclidean version (13.73), Z[0] is equal
to the thermodynamic partition function of the system, up to a trivial factor. Let
us verify this for a free field theory in D = 1 dimension. Then ZE [0] of Eq. (13.73)
becomes
Z =

Dx exp

1
2
d x 2 ( ) + x2 ( )
2
2
"

#)

(13.100)

For D = 1. the fields ( ) may be interpreted as paths x( ) in imaginary time


= it, and we have changed the notation accordingly. In the exponent, we
recognize the euclidean version
Z

AE =

1
2
d x 2 ( ) + x2 ( )
2
2
"

(13.101)

of the action of the harmonic oscillator:


A=

tb

ta

2 2
1 2
dt x (t) x (t) ,
2
2
#

"

(13.102)

for tb ta = ih = ih/kB T . Thus Z is a quantum-statistical path integral for a


harmonic oscillator. The measure of path integration is defined on lattice of points
on the -axis n = n with n = 0, . . . , N + 1 as (13.40):
Z

Dx( ) =

N Z
Y

n=0

dx
n .
2

(13.103)

where xn x(n ) and N + 1 = h


/. For a finite -interval h
, the paths have to
satisfy periodic boundary conditions
x(h) = x(0),

(13.104)

896

13 Functional Integral Representation of Quantum Field Theory

as a reflection of the quantum-mechanical trace. On the -lattice, this implies


xN +1 = x0 , and the action becomes
+1
(xn xn1 )2
1 NX
AN
=

+ 2 x2n .
E
2 n=1
2

"

(13.105)

+1
1 NX
xn (2 + 2 2 )xn .
2 n=1

(13.106)

This can be rewritten as


AN
E =

where x denotes the lattice version of x. It may be represented as an (N + 1)


(N + 1)-matrix

2 =

2 1 0
1 2 1
..
.
0
1

0
0

...
...

0
0

0
0

1
0
..
.

0 . . . 1 2 1
. . . 0 1 2

(13.107)

The Gaussian functional integral can now be evaluated using formula (13.26), and
we obtain
Z = detN +1 [2 + 2 ]1/2

(13.108)

The determinant is calculated recursively,3 with the result


Z =

1
,
2 sinh(h
/2)

(13.109)

with the auxiliar frequency

2
(13.110)

arsinh .

2
In the continnum limit 0, the frequency
goes against , and Z becomes
Z =

1
.
2 sinh(h/2)

(13.111)

Expanding this as
Z = eh/2kB T + e3h/2kB T + e5h/2kB T + . . . ,

(13.112)

this is the quantum statistical partition function of the harmonic oscillator, as we


wanted to prove. The ground state has a nonzero energy, as observed in the operator
discussion in Chapter 7.32.
3

See Section 2.12 of the textbook in Footnote 1.

13.7 Relation Between Z [j ] and Partition Function

897

For the quantum-mechanical version of the functional integral (13.113)


Z =

( Z

Dx exp i

1
2
dt x 2 (t) x2 (t)
2
2
"

tb

ta

#)

(13.113)

we obtain with the measure of functional integration (13.40):


Z

Dx(t) =

N Z
Y

n=0

dx
n ,
2i

(13.114)

via the Gaussian integral formula (13.26) the result


Z = detN +1 [2 2 ]1/2 .

(13.115)

Thus we only have to replace i in (13.110)(13.111). In the continnum limit,


we therefore obtain
1
.
2 sin(h/2)

Z =

(13.116)

We end this section by mentioning that the path integral representation of the
partition function (13.113) with the integration measure (13.103) can obtained from
a euclidean phase space path integral 4
Z =

Z Z

Dp( )
Dx( )
exp
2h

(Z

1
2
d ipx p2 x2
2
2
"

#)

(13.117)

by going to a -lattice andintegrating out the momentum variables. The momentum


integrals give the factors 2 in the denominators of the measure (13.103). In the
quantum-mechanical version of (13.117)
Z =

Z Z

Z tb
Dp(t)
2
1
Dx(t)
exp i
dt px p2 x2
2h
2
2
ta
(

"

they p-integrals produce the denominators containing the factors


(13.103).
The measure of funtional integration on the -lattice
N Z
Y

n=0

dxn

"
 NY
+1 Z
n=1

dpn
2h

#)

(13.118)

i in the measure

(13.119)

is the obvious generalization of the classical statistical weight in phase space


Z

dx

dp
2h

(13.120)

For a detailed discussion of the measure see Chapters 2 and 7 in the textbook on path integrals
quoted in Footnote 1.

898

13 Functional Integral Representation of Quantum Field Theory

to fluctuating paths with many variables xn = x(n ).


For completeness, we write down the free energy F = log Z associated with
the partition function (13.111):
F =

1
h

1
log[2 sinh(h/2)] =
+ log(1 eh ).

(13.121)

At zero temperatures, only the ground-state oscillations contribute.


The detailed time-sliced calculation of the partition functions has to be compared
with the formal evaluation of the partition function (13.113) according to formula
(13.68), which would yield
Z =

Dx( ) exp

= Det

1/2

(2

2
1
d x 2 ( ) + x2 ( )
2
2
"

#)

+ )..

(13.122)

In this continnum formulation, the right-hand side is at first meaningless. It differs


from the results obtained by proper time-sliced calculation (13.108)(13.111) by a
temperature-dependent infinite overall factor. To see what this factor is we note that
for periodic boundary condition (13.104), the eigenvectors of the matrix 2 + 2
2
are eim with eigenvalues m
+ 2. Since m grows with m like m2 , the functional
1/2
2
2
determinant Det
( + ) is strongly divergent. Indeed, the bosonic lattice
result (13.111) can only be obtained dividing out of this determinant an equally
divergent -independent functional determinant and calculating the ratio
Det 1/2 (2 + 2 )
.
h
Det 1/2 (2 )

(13.123)

The prime in the denominator indicates that the zero-frequency 0 must be omitted
to obtain a finite result. The factor 1/h) is the regularized contribution of the zero
frequency. This follows from a simple integral consideration. An integral
Z

b/2

b/2

dx
2 2
e x /2
2

(13.124)

gives 1/ only for finite ( 1/b). If 0, it gives b/ 2. In the path integral,


the zero mode is associated with the fluctuations of the average the path x( ). To
indicate this origin of the factor h
in the ratio (13.123), we may write (13.123) as
Det 1/2 (2 + 2 )
1/2

Det R

(2 )

(13.125)

where the subscript R indicates the above regularization.


The ratio (13.123) is calculated via the product of eigenvalues:
2
m
1 Y
Det 1/2 (2 + 2 )
.
=
2 + 2
h
Det 1/2 (2 )
h
m>0 m

(13.126)

13.7 Relation Between Z [j ] and Partition Function

899

The product can be found in tables:5


2
m
h
/2
=
,
2
2
sinh(h/2)
m6=0 m +

(13.127)

so that (13.126) is equal to 1/2 sinh(h/2), and the properly renormalized partition
function (13.122) yields the lattice result (13.111).
1/2
In a lattice calculation, the determinant Det R (2 ) for is equal to
unity.
It is curious to see that a formal evaluation of the functional determinant via
the analytic regularization procedure of Sections ?? and 11.5 is indifferent to this
denominator, and produces precisely the same result from the formal expression
(13.122) as from the proper lattice calculation. Recalling formula (11.134) and
setting D = 1, we find


1
1/2
2
2
2
2
Det
( + ) = exp Tr log( + )
2
)
(
Z
h

d
1
2
2
log( + ) = e 2 .(13.128)
= exp h
2
2
The exponent gives is precisely the free energies in at zero temperatures in
Eq. (13.121).
We now admit a finite temperatures. Then we have to replace the integral over

by a sum over Matsubara frequencies (2.412), and evaluate


Z
Z
1 X
1 X
d
dm
2
2
2
2
2
log(m
+ 2 ).
log( + ) +
log(m + ) =

h
m
2
h
m
2
(13.129)
!

In contrast to the first term whose evaluation required the analytic regularization,
the second term is finite. It can be rewritten as
!
1 X Z dm Z 2
1

d 2
.
(13.130)

2
h
m
2
m + 2

Recalling the summation formula (2.421), this becomes

d 2
2

"(

coth(h /2)
tanh(h /2)

2
d 2

2 e 1

= kB T 2 log(1 eh ).

(13.131)

For later applications, we have treated also the case of fermionic Matsubara frequencies (the lower case). Together with the zero-temperature result (13.128), we find
at any temperature


1
Z = eF = Det 1/2 (2 + 2 ) = exp Tr log(2 + 2)
2
( "
#)
h

= exp
+ log(1 eh ) ,
(13.132)
2
5

I.S. Gradshteyn and I. M. Ryzhik, op. cit., Formula 1.431.2: sinh x =

m=1 (1

+ x2 /m2 2 ).

900

13 Functional Integral Representation of Quantum Field Theory

in agreement with (13.121).

13.8

Bosons and Fermions in a Single State

The discussion in the last section cannot be taken over to fermionic variables x( )
since in this case the action (13.113) vanishes identically, due to the symmetry of the
functional matrix D(, ) = (2 + 2)( ). A fermionic version of the above
path integral can only be introduced within the canonical formulation (13.118) of
the harmonic path integral. With the help of a canonical transformation
a =

(13.133)

d (a it a a a).

(13.134)

1/2h(x ip), a =

1/2h(x + ip),

this may be transformed into the action


A QM =

tb

ta

A canonical quantization with commutation and anticommutation rules


[a(t), a
(t)] = 1,
[a (t), a
(t)] = 0,
[a(t), a
(t)] = 0.

(13.135)

produces a second-quantized Hilbert space of the type discussed in Chapter 2. Since


a (t), a(t) carry no space variable, they desribe Bose and Fermi particles at a single
point.
The path integral representation of the quantum-mechanical partition function
of this system is
Z
Da(t)Da (t) iA QM
Z QM
e
.
(13.136)
2
The measure of integration is
Z Z

Da Da

D a1 D a2
,
2

(13.137)

where a = (a1 + ia2 )/ 2, is directly obtained from the canonical measure of path
integration in (13.117).
At a euclidean time = it, the action of the free nonrelativistic field becomes
A =

d (a a + a a),

(13.138)

and the thermodynamic partition function has the path integral representaion
Z

Da( )Da ( ) A
e
.
2

(13.139)

901

13.8 Bosons and Fermions in a Single State

In this formulation, there is no problem in trating simultaneously the case of bosons


and fermions. We simply have to assume the fields a( ), a ( ) to be periodic and
antiperiodic in the imaginary-time interval
a(h) = a(0),

(13.140)

aN +1 = a0 .

(13.141)

or in the sliced form


Using formula (13.97), the continuum partition function partition function can be
written as
Z =

Da Da
exp
2
"

d (a a + a a)

= Det 1/2 ( + ).

(13.142)

Contact with the previous oscillator calculation is established by observing that


in the determinant, the operator + can be replaced by the conjugate operator
+, since all eigenvalues come in complex-conjugate pairs, except for the m = 0 value, which is real. Hence the determinant of + can be substituted everywhere
by
q
(13.143)
det ( + ) = det ( + ) = det (2 + 2 ),

In the boson case, we thus reobtain the result (13.122). In both cases, we may
therefore write
Z = Det 1/2 (2 + 2 ).
(13.144)

The right-hand side was evaluate for bosons in Eqs. (13.128)(13.132). Anticipating
the present problem, we calculated the determinant up to Eq. (13.131) for bosons
and fermions. The result for the two cases is then the obvious extension of (13.132):
F

Z = e

= Det
"

= exp

1/2

(2

1
+ ) = exp Tr log(2 + 2)
2
#)


+ log(1 eh )
2

(13.145)

This can be written as


Z =

[sinh(h/2)]1
cosh(h/2)

for

bosons,
fermions.

(13.146)

For bosons, the physical interpretation of this expression was given after
Eq. (13.109). In analogy we rewrite for fermions
Z = eh/2kB T + eh/2kB T ,

(13.147)

and see that the system has two states, one with no particle and one with one
particle, where the no-particle state has a negative vacuum energy, as observed in
the operator discussion in Chapter 2..

902

13 Functional Integral Representation of Quantum Field Theory

13.9

Free Energy for Free Scalar Fields

The results of the last section are easily applied to fluctuating scalar fields. Consider
the free-field partition function
Z0 =

D(x)e

dD x

[()2 +m2 2 ] ,

1
2

(13.148)

which is of the general form (13.73) for j = 0. If the fields are decomposed into its
spatial Fourier components in a finite box of volume V ,
Z
i
1 X h ikx

e k ( ) + c.c. ,
(x) =
2V k

(13.149)

the partition function becomes


YZ

Z0 =

Dk ( )e

R h
0

dt

1
2

{| k ( )|2 +k2 |k ( )|2 +m2 |k ( )|2 } .

(13.150)

For each k, the functional integral is obviously the same as in (13.113), so that
Z0 =

Z(k) ,

(13.151)

where (k)

k2 + m2 , and by (13.111)
Z(k) =

1
.
2 sinh[h(k)/2]

(13.152)

For a real field (x), the anticommuting alternative cannot be accommodated


into the functional integral (13.148). We must first go to the field-theoretic analog
of the path integral in phase space
Z Z

Z0 =

exp

D(x)
Z

D(x)
2h


1
1
d x i(x) (x) 2 (x) [((x))2 + m2 2 (x)]
2
2
4



, (13.153)

where (x) are the canonical field momenta (7.1). After a Fourier decomposition
(13.149) and a similar one for (x), we perform again a canonical tranformation
corresponding to (13.133),
ak ( ) =

1/2h[(k)k ( ) ik ( )],

ak ( ) =

1/2h[(k)k ( ) + ik ( )],
(13.154)

and arrive at the analog of (13.142):


Z =

YZ
k

Y
k

Dak Dak
exp
2
"

Det 1/2 ( + (k)).

d (ak ak + ak ak )

(13.155)

903

13.10 Interacting Nonrelativistic Fields

which can be evaluated as in (13.146) to


Z =

Y
k

{2 sinh[h(k)/2]}1
2 cosh[h(k)/2]

for

bosons,
fermions.

(13.156)

The associated free energy is


F0 = log Z0 =

X
k

i
h
(k)
+ log 1 eh(k) .
2

(13.157)

For a complex field, the canonical transformation is superfluous. The field in the
partition function
Z0 =

D(x)D (x)e

dD x

1
2

( +m2 ) ,

(13.158)

can directly be assumed to be of the bosonic or fermionic type. A direct application


of the Gaussian formula (13.97) yields
Z =

Det [2 + 2 (k)]1

{2 sinh[h(k)/2]}2
{2 cosh[h(k)/2]}2

for

bosons,
fermions.

(13.159)

with a free energy twice as large as (13.157).

13.10

Interacting Nonrelativistic Fields

The quantization of nonrelativistic particles was amply discussed in Chapter 2 and


applied to many-body Bose and Fermi systems in Chapter 3. Her we shall demonstrate that a completely equivalent formulation of the second-quantized nonrelativistic field theory is possible with the help of functional integrals.
Consider a many-fermion system described by an action
A A0 + Aint =

d3 xdt (x, t) [it (i)] (x, t)

1
2

(13.160)

d3 xdtd3 x dt (x , t ) (x, t)V (x, t; x t )(x, t)(x , t )

with a translationally invariant two-body potential


V (x, t; x , t ) = V (x x , t t ).

(13.161)

In the systems to be treated in this text we shall be concerned with the potential
is, in addition, instantaneous in time
V (x, t; x , t ) = (t t )V (x x ).

(13.162)

904

13 Functional Integral Representation of Quantum Field Theory

This property will greatly simplify the discussion.


The fundamental field (x) may describe bosons or fermions. The complete
information on the the physical properties of the system resides in the Green functions. In the operator Heisenberg picture, these are given by the expectation values
of the time-ordered products of the field operators
(13.163)
G (x1 , t1 , . . . , xn , tn ; xn , tn , . . . , x1 , t1 )



= h0|T H (x1 , t1 ) H (xn , tn )H


(x1 , t1 ) |0i
(xn , tn ) H
The time-ordering operator T changes the position of the operators behind it in
such a way that earlier times stand to the right of later times. To achieve the final
ordering, a number of field transmutations are necessary. If F denotes the number
of transmutations of Fermi fields, the final product receibes a sign factor (1)F .
It is convenient to view all Green functions (13.163) as derivatives of the generating functional
 Z

Z[ , ] = h0|T exp i

d3 xdt H
(x, t)(x, t) + (x, t)H (x, t)

i

|0i (13.164)

namely
(13.165)
G (x1 , t1 , . . . , xn , tn ; xn , tn , . . . , x1 , t1 )

n+n

Z[ , ]


.

= (i)n+n

(x1 , t1 ) (xn , tn )(xn , tn ) (x1 , tn ) = 0

Physically, the generating functional describes the amplitude that the vacuum remains a vacuum in spite of the presence of external perturbations.
The calculation of these Green functional is usually performed in the interaction
picture which can be summarized by the operator expression for Z:

Z[ , ] = Nh0|T exp iAint [ , ] + i

d xdt (x, t)(x, t) + h.c.

i

|0i.(13.166)

In the interaction picture, the fields (x, t) possess free-field propagators and the
normalization constant N is determined by the condition [which is trivially true for
(13.164)]:
Z[0, 0] = 1.

(13.167)

The standard perturbation theory is obtained by expanding exp{iAint } in (13.166)


in a power series and bringing the resulting expression to normal order via Wicks
expansion technique. The perturbation expansion of (13.166) often serves conveniently to define an interacting theory. Every term can be pictured graphically and
has a physical interpretation as a virtual process.
Unfortunately, the perturbation series up to a certain order in the coupling constant is unable to describe many important physical phenomena, for example bound

905

13.10 Interacting Nonrelativistic Fields

states in the vacuum and collective excitations in many-body systems. Those require
the summation of infinite subsets of diagrams to all orders. In many situations it
is well-known which subsets have to be taken in order to account approximately for
specific effects. What is not so clear is how such lowest approximations can be improved in a systematic manner. The point is that as soon as a selective summation is
performed, the original coupling constant has lost its meaning as an organizer of the
expansion and there is need for a new systematics of diagrams. Such a systematics
will be presented in what follows.
As soon as bound states or collective excitations are formed, it is very suggestive
to use them as new quantum fields rather than the original fundamental particles .
The goal would then to be rewrite the expression (13.166) for Z[ , ] in terms of
new fields whose unperturbed propagator has the free energy spectrum of the bound
states or collective excitations and whose Aint describes their mutual interactions.
In the operator form (13.166), however, such changes of fields are hard to conceive.

13.10.1

Functional Formulation

In the functional integral approach, the generating functional (13.164) is given by

Z[ , ] = N

D (x, t)D(x, t)


exp iA[ , ] + i

d xdt [ (x, t)(x, t) + c.c.] . (13.168)

It is worth emphasizing that the field (x, t) in the path integral formulation is a
complex number and not an operator. All quantum effects are accounted for by
fluctuations; the path integral includes not only the classical field configurations but
also all classically forbidden ones, i.e., all those which do not run through the valley
of extremal action in the exponent.
By analogy with the development in Section 13.4 we take interaction outside the
integral and write or the vacuum expectation value in either formula as
(

Z[ , ] = exp iAint

"

1 1
,
i i

#)

Z0 [ , ],

(13.169)

where Z0 is the generating functional for the free fields. Thus in Eq. (13.168) there
is only A0 of (13.160) in the exponent. Since
A0 [ , ] =

dxdt (x, t) [it (i)] (x, t)

(13.170)

the functional integral is of the Gaussain type (13.25) with a matrix


A(x, t; x , t ) = [it (i)] (3) (x x )(t t ).

(13.171)

This matrix is the inverse of the free propagator



A(x, t; x , t ) = iG1
0 (x, t; x , t )

(13.172)

906

13 Functional Integral Representation of Quantum Field Theory

where

dE
2

G0 (x, t; x , t ) =

i
d3 p i[E(tt )p(xx )]
e
.
4
(2)
E (p) + i

(13.173)

Inserting this into (13.99), we see that


 

Z0 [ , ] = N exp i

iTr log iG1


0

d xdtd x dt (x, t)G0 (x , t )(x , t ) .

We now fix N in accordance with the normalization (13.167) to


N = exp [i (iTr log iG0 )]

(13.174)

and arrive at


Z0 [ , ] = exp

d xdtd x dt (x, t)G0 (x, t; x , t )(x , t ) . (13.175)

This coincides exactly with what would have been obtained from the operator expression (13.166) for Z0 [ , ] (i.e., without Aint ).
Indeed, according to Wicks theorem [2, 3, 4], any time ordered product can
be expanded as a sum of normal products with all possible contractions taken via
Feynman propagators. The formula for an arbitrary functional of free fields , is
R

T F [ , ] = e

G0 (x,t;x ,t ) (x,t
d3 xdtd3 x dt (x,t)
)

: F [ , ] : .

(13.176)

Applying this to
 Z

h0|T F [ , ]|0i = h0|T exp i

dxdt( + ) |0i

(13.177)

one finds:


Z0 [ , ] = exp

dxdtdx dt (x, t)G0 (x, t; x , t )(x , t )


 Z

h0| : exp i

dxdt( + ) : |0i.

(13.178)

The second factor is equal to unity thus proving the equality of this Z0 with the
path integral result (13.175) [which holds for the full Z[ , ] because of (13.169)].

13.10.2

Grand-Canonical Ensembles at Zero Temperature

All these results are easily generalized from vacuum expectation values to thermodynamic averages at fixed temperatures T and chemical potential . The change
at T = 0 is trivial: The single particle energies in the action (13.160) have to be
replaced by
(i) = (i)

(13.179)

907

13.10 Interacting Nonrelativistic Fields

and new boundary conditions have to be imposed upon all Green functions via an
appropriate i prescription in G0 (x, t; x , t ) of (13.173) [see [2, 5]]:
T =0

G0 (x, t; x , t ) =

i
dEd3 p iE(tt )+ip(xx )
e
. (13.180)
4
(2)
E (p) + i sgn (p)

Note that, as a consequence of the chemical potential, fermions with < 0 inside
the Fermi sea propagate backwards in time. Bosons, on the other hand, have in
general > 0 and, hence, always propagate forward in time.
In order to simplify the notation we shall often use four-vectors p = (p0 , p) and
write the measure of integration in (13.180) as
dEd3 p
=
(2)4

d4 p
.
(2)4

(13.181)

Note that in a solid, the momentum integration is really restricted to a Brioullin


zone. If the solid has a finite volume V , the integral over spacial momenta becomes
a sum over momentum vectors,
Z

d3 p
1 X
=
,
(2)3
V p

(13.182)

and the Green function (13.180) reads


T =0

G0 (x, t; x , t )

dE 1 X ip(xx )
i
.
e
0
2 V p
p (p) + i sgn (p)

(13.183)

The resulting formulas for T =0 Z[ , ] can be brought to conventional form by performing a Wick rotation in the complex energy plane in all energy integrals (13.180)
implied by formulas (13.178) and (13.166). For this, one sets E = p0 i and
replaces
Z

dE
i
2

d
.
2

(13.184)

Then the Green function (13.180) becomes


T =0

G0 (x, t; x , t ) =

d d3 p (tt )+ip(xx )
1
e
.
2 (2)3
i (p)

(13.185)

Note that with formulas (13.175) and (13.169), the generating functional T =0 Z[ , ]
is the grand-canonical partition function in the presence of sources [5].
Finally, we have to introduce arbitrary temperatures T . According to the standard rules of quantum field theory (for an elementary introduction see Chapter 2 in
Ref. [2]), we must continue all times to imaginary values t = i , restrict the imaginary time interval to the inverse temperature6 1/T , and impose periodic or
6

Throughout these lectures we shall use natural units so that kB = 1, h = 1.

908

13 Functional Integral Representation of Quantum Field Theory

antiperiodic boundary conditions upon the fields (x, i ) of bosons and fermions,
respectively [2, 5]:
(x, i ) = (x, i( + 1/T )).
(13.186)

When there is no danger of confusion, we shall usually drop the factor i in front
of the imaginary times in the field arguments, for brevity. The same thing will be
done in the Green functions.
By virtue of (13.169) and (13.175), also the Green functions satisfy these boundary conditions. With the above notation:
T

G0 (x, + 1/T ; x , ) G0 (x, i ; x , i ).

(13.187)

This property is enforced automatically by replacing the energy integrations


d/2 in (13.185) by a summation over the discrete Matsubara frequencies [in
analogy to the momentum sum (13.182), the temporal volume being = 1/T ]
R

X
d
T
2
n

(13.188)

which are even or odd multiples of T


n =

2n
2n + 1

T for

bosons
fermions

(13.189)

The prefactor T of the sum over the discrete Matsubara frequencies accounts for the
density of these frequencies yielding the correct T 0-limit.
Thus we obtain for the imaginary-time Green function of a free nonrelativistic
field at finite temperature (the so-called free thermal Green function) the following
expression:
T

G0 (x, , x , ) = T

XZ
n

d3 p in ( )+ip(xx )
1
e
.
3
(2)
in (p)

(13.190)

Incorporating the Wick rotation in the sum notation we may write


T

X
p0

= iT

X
n

= iT

(13.191)

p4

where p4 = ip0 = . If both temperatures, and volume are finite, the Green
function will be written as
T

G0 (x, , x , ) =

1
T X X in ( )+ip(xx )
.
e
V p0 p
in (p)

(13.192)

At equal space points and equal imaginary times, the sum can easily be evaluated.
One must, however, specify the order in which . Let denote an infinitesimal
positive number and consider the case = + , i.e., the Green function
T

G0 (x, , x, + ) = T

XZ
n

1
d3 p in
e
.
3
(2)
in (p)

909

13.10 Interacting Nonrelativistic Fields

The sum is now found by changing it into a contour integral


T

ein

1
T
=
in (p)
2i

dz

1
ez
.
ez/T 1 z

(13.193)

The upper sign holds for bosons, the lower for fermions. The contour of integration
C encircles the imaginary z axis in the positive sense, thereby enclosing all integer
or half-integer valued poles of the integrand at the Matsubara frequencies z = im
(see Fig. 2.8).
The factor ez ensures that the contour in the left half-plane does not contribute.
By deforming the contour C into C and contracting C to zero we pick up the
pole at z = and find
T

ein

1
1
1
= (p)/T
= (p)/T
= n((p)).
in (p)
e
1
e
1

(13.194)

The phase ez ensures that the contour in the left half-plane does not contribute.
The function on the right is known as the Bose or Fermi distribution function.
By subtracting from (13.194) the sum with replaced by , we obtain the
important sum formula
T

X
n

n2

1
(p)
1
=
coth1
2
+ (p)
2(p)
T

(13.195)

In the opposite limit = , the phase factor in the sum would be eim
and would be converted into a contour integral
kB T

X
m

eim

1
kB T
=
im (p)
2i

dz

ez
ez/kB T

1
,
1z

(13.196)

and we would find 1 n(p) .


In the operator language, these limits correspond to the expectation values
T

G (x, ; x, + ) = h0|T H (x, )H


(x, + ) |0i = h0|H
(x, )H (x, )|0i

G (x, ; x, ) = h0|T H (x, )H


(x, ) |0i = h0|H (x, )H
(x, )|0i

= 1 h0|H
(x, )H (x, )|0i

The function n((p)) is the thermal expectation value of the number operator
= (x, )H (x, ).
N
H

(13.197)

Also in the case of T 6= 0 ensembles, it is useful to employ a four-vector notation.


The four-vector
pE (p4 , p) = (, p)

(13.198)

910

13 Functional Integral Representation of Quantum Field Theory

is called the euclidean four-momentum. Correspondingly, we define the euclidean


spacetime coordinate
xE (, x).

(13.199)

The the exponential in (13.190) can be written as


pE xE = + px.

(13.200)

Collecting integral and sum in a single four-summation symbol, we shall write


(13.190) as
T

G0 (xE x )

T X
1
.
exp [ipE (xE xE )]
V pE
ip4 (p)

(13.201)

It is quite straightforward to derive the general T 6= 0 Green function from a


path integral formulation analogous to (13.168). For this we consider classical fields
(x, ) with the periodicity or anti-periodicity
(x, ) = (x, + 1/T ) .

(13.202)

They can be Fourier-decomposed as


(x, ) =

T X X in +ipx
T X ipE xE
e
a(pE )
e
a(n , p)
V n p
V pE

(13.203)

with a sum over even or odd Matsubara frequencies n . If now a free action is
defined as

A0 [ , ] = i

1/2T
1/2T

d3 x (x, ) [ (i)] (x, )


(13.204)

formula (13.99) renders [1, 6]


T

Tr log A+

Z0 [ , ] = e

with

R R 1/2T

1/2T

d d

d3 xd3 x (x, )A1 (x,,x , )(x , )

A(x, ; x , ) = [ + (i)] (3) (x x )( ),

(13.205)

(13.206)

and henceforth A1 equal to the propagator (13.190), the Matsubara frequencies


arising due to the finite interval of Euclidean space together with the periodic
boundary condition (13.202).
Again, interactions are taken care of by multiplying TZ0 [ ] with the factor
(13.169). In terms of the fields (x, ), the exponent has the form:
Aint

1
=
2

Z Z
Z

1/2T
1/2T

d d

d3 xd3 x (x, ) (x , )(x , )(x, )V (x, i ; x , i (13.207)


).

911

13.10 Interacting Nonrelativistic Fields

In the case of an instantaneous potential (13.161), the potential becomes instantaneous in :


V (x, i ; x , i ) = V (x x ) i( ).

(13.208)

In this case Aint can be written in terms of the interaction Hamiltonian as


Aint = i

1/2T

1/2T

d Hint ( ).

(13.209)

Thus the grand canoncial partition function in the presence of external sources may
be calculated from the path integral [6]:
T

Z[ , ] =

i A+

D (x, )D(x, )e

R 1/2T

1/2T

d3 x[ (x, )(x, )+c.c.]

(13.210)

where the grand-canonical action is


T

i A[ , ] =

1/2T

1/2T

d3 x (x, ) [ + (i)] (x, )

(13.211)

Z
i Z 1/2T
+
d d d3 xd3 x (x, ) (x , )(x, )(x, )V (x, i ; x, i ).
2 1/2T

(13.212)
G (x1 , 1 , . . . , xn , n ; xn , n , . . . , x1 , 1 )

n+n

Z[ , ]


= (i)n+n
.

(x1 , 1 ) (xn , n )(xn , n ) (x1 , n ) = 0

The right-hand side consists of the functional integrals


Z

,]
1 , 1 ) (x
n , n ) (xn , n ) (x1 , 1 )eiA[(13.213)
.
N D (x, t)D(x, t) (x

In the sequel, we shall always assume the normalization factor to be chosen in such
a way that Z[0, 0] is normalized to unity. Then the functional integrals (13.213) are
obviously the correlation function of the the fields commonly written in the form
1 , 1 ) (x
n , n ) (xn , n ) (x1 , 1 )i
h(x
In contrast to Section 1.2, the bra and ket symbols denote now a thermal average
of the classical fields.
The functional integral expression (13.210) for the generating functional offers
the advantageous flexibility with respect to changes in the field variables.
Summarizing we have seen that the functional (13.210) defines the most general
type of theory involving two-body forces. It contains all information on the physical
system in the vacuum as well as in thermodynamic ensembles. The vacuum theory is
obtained by setting T = 0, = 0, and continuing the result back from T to physical

912

13 Functional Integral Representation of Quantum Field Theory

times. Conversely, the functional (13.168) in the vacuum can be generalized to


ensembles in the straight-forward manner by first continuing the times t to imaginary
values i via a Wick rotation in all energy integrals and then going to periodic
functions in .
There is a complete correspondence between the real-time generating functional
(13.168) and the thermodynamic imaginary-time expression (13.210). For this reason it will be sufficient to exhibit all techniques only in one version for which we
shall choose (13.168). Note, however, that due to the singular nature of the propagators (13.173) in real energy-momentum, the thermodynamic formulation specifies
the way to specifies how to avoid singularities.

13.11

Interacting Relativistic Fields

Let us see how this formalism looks for relativistic boson and fermion systems.
Conseider a Lagrangian of Klein-Gordon and Dirac particles consisting of a sum
= L0 + Lint .
L , ,


(13.214)

As in the case of nonrelativistic felds, all time ordered Greens functions can be
obtained from the derivatives with respect to the external sources of the generating
functional
Z [, , j] = const h0|T ei

dx(Lint +
++j
)

|0i.

(13.215)

The fields in the exponent follow free equations of motion and |0i is the free-field
vacuum. The constant is conventionally chosen to make Z [0, 0, 0] = 1, i. e.


const = h0|T e

)
dxLint (,,

|0i

1

(13.216)

This normalization may always be enforced at the very end of any calculation such
that Z [, , j] is only interesting as far as its functional dependence is concerned,
modulo the irrelevant constant in front.
It is then straight-forward to show that Z [, , j] can alternatively be computed
via the Feynman path integral formula
Z [, , j] = const

DD De

)+Lint +

dx[L0 (,,
++j
]

(13.217)

Here the fields are no more operators but classical functions (with the mental reservation that classical Fermi fields are anticommuting objects). Notice that contrary
to the operator formula (13.215) the full action appears in the exponent.
For simplicity, we demonstrate the equivalence only for one real scalar field (x).
The extension to other fields is immediate [?], [?]. First note that it is sufficient to
give the proof for the free field case, i. e.
Z0 [j] = h0|T ei

dxj(x)(x)

= const

Dei

|0i
R

dx[ 21 (x)(x 2 )(x)+j(x)(x)]

(13.218)

913

13.11 Interacting Relativistic Fields

For if it holds there, a simple multiplication on both sides of (13.218) by the differential operator
ei

dxLint ( 1i j(x)
)

(13.219)

would extend it to the interacting functionals (13.215) or (13.217). But (13.218)


follows directly from Wicks theorem according to which any time ordered product
of a free field can be expanded into a sum of normal products with all possible time
ordered contractions. This statement can be summarized in operator form valid for
any functional F [] of a free field (x):
T F [] = e

1
2

dxdy (x)
D(xy) (y)

: F [] :

(13.220)

where D(x y) is the free-field propagator

Z
i
i
d4 q iq(xy)
(x

y)
=
e
. (13.221)
D(x y) =
2
4
2
x + i
(2)
q 2 + i

Applying this to (13.220) gives


1

Z0 = e 2

dxdy (x)
D(xy) (y)

R
1

= e 2

R
1

= e 2

dxdyj(x)D(xy)j(y)
dxdyj(x)D(xy)j(y)

h0| : ei

h0| : ei

dxj(x)(x)

dxj(x)(x)

: |0i

: |0i

(13.222)

The last part of the equation follows from the vanishing of all normal products of
(x) between vacuum states.
Exactly the same result is obtained by performing the functional integral in
(13.218) and using the functional integral formula (13.98). The matrix A is equal
to A(x, y) = (x 2 ) (x y), and its inverse yields the propagator D(x y):
A1 (x, y) =

1
(x y) = iD(x y)
x 2 + i

(13.223)

yielding again (13.222).


For the generating functional of a free Dirac field theory
Z0 [, ] = h0|T ei

)dx
(
+

= const
with the free-field Lagrangian

|0i
R
]
i dx[L0 (,)++
.
DD e

(13.224)

(i M) (x),
L0 (x) = (x)

(13.225)

we obtain, similarly,
1

Z0 [
, ] = e 2

21

= e

dxdy (x)
G0 (xy) (y)

R
1

= e 2

dxdy (x)G0 (xy)(y)


dxdy (x)G0 (xy)(y)

h0| : ei

h0| : ei
.

)
dx(
+

)
dx(
+

: |0i
: |0i

(13.226)

914

13 Functional Integral Representation of Quantum Field Theory

Now,
A(x, y) = (i M) (x y),

(13.227)

and its inverse yields the fermion propagator G0 (x y):


A1 (x, y) =

1
(x y) = iG0 (x y).
M + i

(13.228)

Note that it is Wicks expansion which supplies the free part of the Lagrangian
when going from the operator form (13.220) to the functional version (13.217).

13.12

Plasma Oscillation

The functional formulation of second-quantized many-body sytems allows us to treat


efficiently various collective phenomena such systems. As a first example we shall
consider a many-electron system that interact only via long-range Coulomb forces.
The Coulomb forces give rise to collective modes called plasmons.
The other extremely important example caused by attractive short-range interactions will be treated in the next chapter.

13.12.1

General Formalism

Let us give a first application of the functional method by transforming the grand
partition function (13.210) to plasmon coordinates.
For this, we make use of the Hubbard-Stratonovic transformation (13.98) and
observe that a two-body interaction (13.160) in the generating functional can always
be generated by an auxiliary field (x) as follows.
i
exp
2


dxdx (x) (x )(x)(x )V (x, x )

= const

(13.229)

h
i
iZ
dxdx (x)V 1 (x, x )(x )2(x) (x)(x)(x x )
2

To abbreviate the notation, we have used four-vector notation with


x (x, t),

dx d3 xdt,

(x) 3 (x)(t).

The symbol V 1 (x, x ) denotes the functional inverse of the matrix V (x, x ), i.e., the
solution of the equation
Z

dx V 1 (x, x )V (x , x ) = (x x ).

(13.230)

The constant prefactor in (13.229) is [det V ]1/2 . Absorbing this in the always
omitted normalization factor N of the functional integral, the grand-canonical partition function = Z becomes

Z[ , ] =

D DD exp iA + i

dx ( (x)(x) + (x)(x))

(13.231)

915

13.12 Plasma Oscillations

where the new action is


A[ , , ] =

dxdx (x) [it (i) (x)] (x x )(x ) (13.232)


1
+ (x)V 1 (x, x )(x ) .
2


Note that the effect of using formula (13.98) in the generating functional amounts
to the addition of the complete square in in the exponent:
1
2

dxdx (x)

dyV (x, y) (y)(y) V 1 (x, x )

(x )

dy V (x , y ) (y )(y )

(13.233)

together with the additional integration over D. This procedure of going from
(13.160) to (13.232) is probably simpler mnemonically than formula (13.98). The
fact that the functional Z remains unchanged by this addition follows, as before,
since the integral D produces only the irrelevant constant [det V ]1/2 .
The physical significance of the new field (x) is easy to understand: (x) is
directly related to the particle density. At the classical level this is seen immediately
by extremizing the action (13.232) with respect to variations (x):
Z
A
= (x) dyV (x, y) (y)(y) = 0.
(x)

(13.234)

Quantum mechanically, there will be fluctuations around the field configuration


(x) determined byR Eq. (13.234), making the Green functions of (x) and of the
composite operator dyV (x, y) (y)(y) different. But due to the Gaussian nature
of the D integration, the fluctuations are quite simple. One can easily show that,
for example, the propagators of either field differ only by the direct interaction, i.e.,
hT ((x)(x ))i

(13.235)


= V (x x ) + T

Z

dyV (x, y)(y)

 Z

dy V (x , y ) (y )(y )



For the proof, the reader is referred to Appendix 13A. Note, that for a potential
V which is dominantly caused by a single fundamental-particle exchange, the field
(x) coincides with the field of this particle: If, for example, V (x, y) represents the
Coulomb interaction
e2
(t t )
|x x |

(13.236)

4e2
(x, t)(x, t)
2

(13.237)

V (x, x ) =
Eq. (13.234) amounts to
(x, t) =

916

13 Functional Integral Representation of Quantum Field Theory

revealing the auxiliary field as the electric potential.


If the particles (x) have spin indices, the potential will, in this example, be thought
of a spinRconserving at every vertex and Eq. (13.234) must be read as spin contracted:
(x) d4 yV (x, y) (y) (y). This restricion is just for convenience and can easily be lifted later. Nothing in our procedure depends on this particular property of
V and . In fact, V could arise from the exchange of many different fundamental particles and their multiparticle configurations (for example , , , , etc. in
nuclei) so that the spin dependence is the rule rather than the exception.
The important point is now that the auxiliary field (x) can be made the only
field of the theory by integrating out , in Eq. (13.231), using formula (13.99).
Thus one obtains
Z[ , ] [ , ] = NeiA

(13.238)

where the new action is




A[] = Tr log iG1


+

1
2

dxdx (x)G (x, x )(x )

(13.239)

with G being the Green function of the fundamental particles in an external classical field (x):
[it (i) (x)] G (x, x ) = i(x x ).

(13.240)

The field (x) is called a plasmon field. The new plasmon action can easily be
interpreted graphically. For this, one expands G in powers
G (x, x ) = G0 (x x ) i

dx1 G0 (x x1 )(x1 x ) + . . .

(13.241)

Hence the couplings to the external currents , in (13.239) amount to radiating


one, two, etc. fields from every external line of fundamental particles (see Fig.
13.1). An expansion of the Tr log expression in gives

i
E(p)

Figure 13.1 This diagram displays the last, pure current, piece of the collective action
(13.239). The original fundamental particle (fat line) can enter and leave the diagrams
only via external currents, emitting an arbitrary number of plasmons (wiggly lines) on its
way
1
iTr log(iG1
) = iTr log(iG0 ) iTr log(1 + iG0 )

X
1
(iG0 )n .
= iTr log(iG1
)

iTr
0
n
n=1

(13.242)

917

13.12 Plasma Oscillations

Figure 13.2 The non-polynomial self-interaction terms of the plasmons arising from the
Tr log in (13.239) are equal to the single loop diagrams emitting n plasmons

The first term leads to an irrelevant multiplicative factor in (13.238). The nth term
corresponds to a loop of the original fundamental particle emitting n lines (see Fig.
13.2).
Let us now use the action (13.239) to construct a quantum field theory of plasmons. For this we may include the quadratic term
iTr(G0 )2

1
2

(13.243)

into the free part of in (13.239) and treat the remainder perturbatively. The free
propagator of the plasmon becomes
{0|T (x)(x )|0} (2s + 1)G0 (x , x).

(13.244)

This corresponds to an inclusion into the V propagator of all ring graphs (see Fig.
13.3). It is worth pointing out that the propagator in momentum space Gpl (k)

Figure 13.3 Free plasmon propagator containing an infinite sequence of single loop
coorections (bubblewise summation)

contains actually two important physical informations. From the derivation at fixed
temperature it appears in the transformed action (13.239) as a function of discrete
Euclidean frequencies n = 2nT only. In this way it serves for the time independet
fixed T description of the system. The calculation (13.244), however, renders it as
a function in the whole complex energy plane. It is this function which determines
the time dependent collective phenomena for real times7
With the propagator (13.244) and the interactions given by (13.242), the original theory of fundamental fields , has been transformed into a theory of
fields whose bare propagator accounts for the original potential which has absorbed
ringwise an infinite sequence of fundamental loops.
7

See the discussion in Ch. 9 of the last of [53] and G. Baym and N.D. Mermin, J. Math. Phys. 2,
232 (1961).

918

13 Functional Integral Representation of Quantum Field Theory

This transformation is exact. Nothing in our procedure depends on the statistics


of the fundamental particles nor on the shape of the potential. Such properties
are important when it comes to solving the theory perturbatively. Only under
appropriate physical circumstances will the field represent important collective
excitations with weak residual interactions. Then the new formulation is of great
use in understanding the dynamcis of the system. As an illustration consider a dilute
fermion gas of very low temperature. Then the function (i) is (i) with
(i) = 2 /2m.

13.12.2

Physical Consequences

Let the potential be translationally invariant and instantaneous


V (x, x ) = (t t )V (x x ).

(13.245)

Then plasmon propagator (13.244) reads in momentum space


Gpl (, k) = V (k)

1
1 V (k)(, k)

(13.246)

where the single electron loop symbolizes the analytic expression8


(, k) = 2

1
1
T X
.
2
V p i p /2m + i( + ) (p + k)2 /2m +

(13.247)

The frequencies and are odd and even multiples of T . In order to calculate
the sum we introducing a convergence factor ei , and rewrite (13.247) as
d3 p
1
3
(2) (p + k) (p) i
"
#
X
1
1
in
e
T
,

i(n + ) (p + k) in (p)
n

(, k) = 2

(13.248)

and using the summation formula (13.194), this becomes


(, k) = 2

d3 p n(p + k) n(p)
,
(2)3 (p + k) (p) i

(13.249)

or, after some rearrangment,


(, k) = 2

1
1
d3 p
(13.250)
.
n(p)
+
(2)3
(p + k) (p) i (p k) (p) + i
#

"

Let us study this function for real physical frequencies = i where we rewrite it
as
d3 p
1
1
(, k) = 2
, (13.251)
n(p)
+
3
(2)
(p+k)(p) (pk)(p) +
Z

"

The factor 2 stems from the trace over the electron spin.

919

13.12 Plasma Oscillations

which can be brought to the form


(, k) = 2

k 2 Z d3 p
1
n(p)
.
2
3
m
(2)
( p k/m + i)2 (k 2 /2M)2

(13.252)

For || > pF k/m + k 2 /2m, the integrand is real and we can expand
k2
(, k) = 2
m 2

pk
d3 p
2p k
n(p) 1 +
+3
3
(2)
m
m
pk
+
m

!3

!2

80(p k)4 + m2 2 k 4
+ . . .(13.253)
.
+
16m2 4

Zero Temperature
For zero temperature, the chemical potential = pF and n(p) = (p pF ). Then
the integral in (13.253) can be performed trivially using the integral
2

N
p3F
d3 p
nT =0 (p) =
= n = 2,
(2)3
V
3

(13.254)

and we obtain

3
k2 n
(, k) = 2 1 +
m
5

pF k
m

!2

1
+
5

pF k
m

!4

1 k4
+ . . . .
+
16 m2 2

(13.255)

Inserting this into (13.246) we find, for long wavelength,s the Green function
#1

"

V (k) n
G (, k) V (k) 1
+ ...
2 m
pl

(13.256)

Thus the original propagator is modified by a factor


(, k) = 1

4e2 n
+ ....
2 m

(13.257)

The dielectric constant vanishes at the frequency


= pl =

4e2
,
m

(13.258)

which is the famous plasma frequency of the electron gas. At this frequency, the
plasma propagator (13.246) has a pole on the real- axis, implying the existence of
an undamped excitation of the system.
For an electron gas we insert the Coulomb interaction (13.237), and obtain
#1

4e2
4e2
G (, k) 2 1
n + ...
k
m 2
pl

"

(13.259)

920

13 Functional Integral Representation of Quantum Field Theory

Thus the original Coulomb propagator is modified by factor


(, k) = 1

4e2
n + ...,
m 2

(13.260)

which is simply the dielectric constant.


The zero temperature limit can al so be calculated exactly starting from the
expression (13.253), written in the form
1
d3 p
(p pF )
+ ( ) , (13.261)
(, k) = 2
3
(2)
p k + k 2 /2m
"

Performing the integral yields

!2
1 2
mpF
k m
k 2 + 2m 2kpF
(, k) = 2 1
pF
+
+ p2F log 2
2
2kpF
2
k
k + 2m + 2kpF

+ ( ).

(13.262)

The lowest terms af a Taylor expansion innpowers of k agree with (13.255).


Short-Range Potential
Let us also find the real poles of Gpl (, k) for a short-rang potential where the
singulatity at k = 0 is absent. Then a rotaionally invariant [V (k)]1 has the longwavelength expansion
[V (k)]1 = [V (0)]1 + ak2 + . . .

(13.263)

as long as [V (0)]1 is finite and


positive, i.e., for a well behaved overall repulsive
R 3
potential satisfying V (0) = d xV (x) > 0. Then the Green function (13.246)
becomes

Gpl (, k) = 2 2 [V (0)]1 + a 2 k 2 k 2

n
3
1+
m
5

This has a pole at = c0 k where

c0 = V (0)

n
.
m

pF k
m

!2

+ . . . (13.264)

(13.265)

This is the spectrum of zero sound with the velocity c0 .


In the neighbourhood of the positive-energy pole, the propagator has the form
Gpl (k0 , k) V (0)

|k|
.
c0 |k|

(13.266)

921

13.13 Gauge Fields and Gauge Fixing

Nonzero Temperature
In order to discuss the case of nonzero temperature it is convenient to add and
subtract in the numeraotr of (13.249) a term n(p + k)n(k), and rewrite it as
(, k) = 2

d3 p n(p + k) [1 n(p)] n(p) [1 n(p + k)]


,
(2)3
(p + k) (p) i

(13.267)

which can be rearranged to


(, k) = 4

d3 p
(p + k) (p)
.
n(p) [1 n(p + k)]
3
(2)
[(p + k) (p]2 + 2

(13.268)

In the high-temperature limit the Fermi distribution becomes Boltzmannian,


2
n(p) e(p /2) , and we evaluate again most easily expression (13.253) as follows:
(, k) = 2

d3 p (p2 /2)[(p+k)(p)]
e
+ ( ) . (13.269)
(2)3

The right-hand side is equal to


Z

d cos d3 p (p2 /2)[(pk cos /m+k2 /2m)]


d
e
+ ( ). (13.270)
2
(2)3
1
Z

Performing the angular integral yields

m2
k2

13.13

d3 p (p2 /2)+ 1
d
e
(pk cosh pk sinh pk) + ( ).(13.271)
2 2 (2)3
p2
Z

Gauge Fields and Gauge Fixing

The functional integral formalism developed in the last sections does not immediately apply to electromagnetism and any other gauge fields. There are subtleties
which we are now going to discuss. These will lead to an explanation of the mistake
in the vacuum energy observed in Eq. (7.496) when quantizing the electromagnetic
field via the Gupta-Bleuler formalism. Consider a set of external electromagnetic
currents described by the current density j (x). Since charge is conserved, these
satisfy
j (x) = 0.
(13.272)
The currents are the sources of an electromagnetic fields F determined from the
field equation (12.49):
F (x) = j (x),

(13.273)

if we employ natural units with c = 1. The action reads


1 2
(x) j (x)A (x)
A =
d x F
4
 h

Z
i 
1
1
4
2
2
=
dx
E (x) B (x) (x) j A(x) ,
2
c
Z

(13.274)

922

13 Functional Integral Representation of Quantum Field Theory

The field strengths are the four-dimensional curls of the vector potential A (x):
F = A A ,

(13.275)

so that they satisfy the Bianchi identity


F = 0.

(13.276)

The decomposition (13.275) is not unique. If we add to A (x) the gradient of an


arbitrary function (x),
A (x) A (x) = A (x) + (x)

(13.277)

then does not appear in the field strengths if it satisfies Schwarzs integrability
condition
( )(x) = 0,
(13.278)

i.e., if the derivatives in from of (x) commute. In the theory of partial differential
equations, this is referred to as the Schwarz integrability condition for the function
(x). In general, a function (x) which satisfies (13.278) in a simply-connected
domain can be defined uniquely in this domain. This property makes (13.277) a
local gauge transformation.
If the domain is multiply connected, there is more than one path along which
to continue the function (x) from one spatial point to another and (x) becomes
multi-valued. This happens, for example, if (13.278) is nonzero on a closed line in
three-dimensional space in which case the set of paths between two given points
decomposes into equivalence classes, depending on how often the closed line is being
encircled. Each of these paths allows for another continuation of (x). By (13.278),
such functions are not allowed in gauge transformations (13.277).
Let us now see how we can construct the generating functional of fluctuating
free electromagnetic fields. in the presence of external currents j (x). As in (13.57),
we would like to calculate a functional integral, now in the unnormalized version
corresponding to (13.57):
Z0 [j] =

DA (x)ei

2 /4j A )
d3 x(F

(13.279)

2
The F
-terms in the exponents can be partially integrated and rewritten as

2
xF

1
=
2

d4 xA ( 2 g )A .

(13.280)

Hence we can identify the functional matrix D(x, x ) in (13.35) as


D (x, x ) = ( 2 g ) (4) (x x ).

(13.281)

Recalling (13.30) [but in the unnormalized version, with the normalization factor N
of Eq. (13.37) divided out] it appears, at first, as though the generating functional
(13.279) should simply be equal to
1

Z0 [j] = (Det G0 )1/2 e 2

d4 xd4 x j (x)G0 (x,x )j (x )

(13.282)

923

13.13 Gauge Fields and Gauge Fixing

where in analogy to (13.36)


1
G0 (x, x ) = iD
(x, x ),

(13.283)

and we have divided out a normalization factor N = (Det G0 )1/2 , since we are
dealing with unnormalized version of Z0 [j].
Unfortunately, the expression (13.282) is meaningless, since the inverse of the
functional matrix (13.281) does not exist. In order to see this explicitly we diagonalize the functional part (i.e., the x, y-part) of G0 (x, x ) by going to the Fourier
transform
D (q) = q 2 g + q q .
(13.284)
For every momentum q, this matrix has obviously an eigenvector with zero eigenvalue, namely q . This prevents us from inverting the matrix D (q). Correspondingly, when trying to form the inverse determinant of the functional matrix D in
(13.282), we encounter an infinite number of infinities, one for every momentum q.
The difficulty can be resolved using the fact that the action (13.274) is gauge
invariant. For F this is trivially true; for the source term j (x)A (x) this is a consequence of current conservation. Indeed, if we change A according to (13.277), the
R 4
source
term
is
changed
by
d x j (x) (x). By a partial integration, this is equal
R
to d4 x j (x)(x)(x), and this expression vanishes due to current conservation
(13.272).
Because of this invariance, not all degrees of freedom over which are integrated
over in the functional integral (13.279) are associated with a Gaussian integral. The
fluctuations corresponding to pure gauge transformations which leave the exponent
invariant. Since a path integral is a product of infinitely many integrals from minus
to plus infinity, the gauge invariance of the integral produces an infinite product
of infinite factors. This is precisely the infinity which occurs in the functional determinant origin of the infinities encountered when trying to calculate the inverse
determinant of the functional matrix D (x, x ).
This infinity must be avoided by restricting the functional integral to field fluctuations by specific gauge condition. The restriction is done by inserting into the
integral gauge-fixing functionals, for example
F1 [A] = [ A (x)],
F2 [A] = [ A(x)],
F3 [A] = [A0 (x)],
F3 [A] = [A3n(x)],
o
i R 4
F4 [A] = exp 2
d x [ A (x)]2 .

(Lorentz gauge)
(Coulomb gauge)
(Hamiltonian gauge)
(axial gauge)

(13.285)

The first is a -functional enforcing the Lorentz gauge at each spacetime point:
A (x) = 0. The second enforces the Coulomb gauge, the third corresponds to
the axial gauge, and the fourth can be used to derive the Feynman diagrams of
the Gupta-Bleuler quantization formalism, thereby correcting the mistake in the
vacuum energy.

924

13 Functional Integral Representation of Quantum Field Theory

If we insert into any of these gauge-fixing functionals F [A] the gauge-transformed


vector potential (13.277), and integrate functionally over all (x), the integral receives a finite contribution for that gauge function (x) which enforces the desired
gauge. The result is a gauge-invariant functionals of A (x):
(13.286)
[A] =

DF [A ] =

DF [A + ].

(13.287)

Explicitly, we find for the above cases (13.285):


1 [A] =
2 [A] =
3 [A] =
4 [A] =

DF1 [A ] =
DF2 [A ] =
DF2 [A ] =
DF3 [A ] =

D[ A (x) + 2 ],

(13.288)

D[ A(x) + 2 ],

(13.289)

D[A0 (x) + 0 ],


D exp

i
2

(13.290)


d4 x [ A + 2 ]2 . (13.291)

If we form the ratios Fi [A]/i [A], we obtain gauge fixing functionals which all yield
unity if integrated over all gauge transformations. If any of these are inserted into
the functional integral (13.279), they will all remove the gauge degree of freedom,
and lead to a finite functional integral which is the same for each choice of Fi [A].
Let us calculate the functionals i [A] explicitly. For 1,2,3 [A] we simply observe
that the trivial identity for -functions
(ax) = a1 (x),

(13.292)

which is proved by multiplying both sides with a smooth function f (x) and integrating over x, has a functional generalization
[OA] = Det 1 O [A],

(13.293)

where O is an arbitrary differential operator acting on the field A (x). Hence we


find immediately
1 [A] = Det 1 ( 2 ),
2 [A] = Det 1 (2 ),
3 [A] = Det 1 ( 0 ),

(13.294)
(13.295)
(13.296)

The last functional 4 [A] is simply a Gaussian functional integral. The additive term
A (x) can be removed by a trivial shift of the integration variable (x) (x) =
(x) A (x)/ 2 , under which the measure of integration remains invariant, D =
D . Using formula (13.24) we obtain
4 [A] = Det 1/2 ( 4 /).

(13.297)

925

13.13 Gauge Fields and Gauge Fixing

The functional determinants 1


are called Fadeev-Popov,determinants. .
i
The Fadeev-Popov determinants in the four examples happen to be independent
of A (x), so that we shall write them as i without arguments. This independence
is a very useful property. Complications arising for A-dependent functionals [A]
will be illustrated below in an example.
We now study the consequences of inserting the gauge-fixing factors Fi [A]/i .
into the functional integrands (13.279). For F4 [A]/4 , the generating functional
becomes
Z0 [j] = (Det G0 )1/2 Det 1/2 ( 4 /)

DA (x)ei

2 1 ( A )2 j A
d4 x[ 41 F

. (13.298)

The free-field action in the exponent can be written in the form


A=

1
1
( A )2 .
d x ( A )2 + 1
2

(13.299)

The associated Euler-Lagrange equation is


1
(A) = 0,

2 A 1

(13.300)

which is precisely the field equation (7.371) of the covariant quantization scheme.
With the additional term in the action, the matrix (13.284) becomes
D (q) = q

"

1 q q
.
1
q2


(13.301)

This can be decomposed into projection matrices with respect to the subspaces
transverse and longitudinal to the four-vector q ,
q q
q q
l
l
P
(q) = 2 , P
(q) = g 2 ,
(13.302)
q
q
as


1 l
t
(q) .
(13.303)
(q) + P
D (q) = q 2 P

It is easy to verify that the matrices are really projections, since they satisfy
t
t
l
l
t
P
P t = P
, P
P l = P
, P
P l = 0.

(13.304)

A similar transverse projection appeared before in the three-dimensional subspace


space [see (4.328), (12.6), and (12.78), (12.326) where the projection was indicated
by a capital subscript T ].
Due to the relations (13.304), there is no problem in inverting D (q), and we
find the free photon propagator in momentum space
i t
l
[P (q) + P
(q)]
q 2
#
q q
(1 ) 2 .
q

1
G0 (q) = iD
(q) =

"

i
= 2 g
q

(13.305)
(13.306)

926

13 Functional Integral Representation of Quantum Field Theory

For = 0, this is the propagator G derived in the Gupta-Bleuler canonical field


quantization in Eq. (7.500), there obtained from canonical quantization rules with
a certain filling of the vacuum with unphysical states.
Taking the Fadeev-Popov determinant 1
4 of (13.297) into account, we obtain
for the generating functional (13.298) the result:
1

Z[j] = [Det (iG0 )]1/2 Det 1/2 ( 4 /)e 2

d4 xd4 x j (x)G0 (x,x )j (x )

(13.307)

Let us calculate the functional determinant Det G0 . For this we take in the momentum representation (13.305) q along the 0-direction and see that there are three
spacelike eigenvectors of eigenvalues iq 2 , and one timelike eigenvector of eigenvalue
i/q 2 , The total determinant is therefore
Det (iG0 ) =

1
Det ( 2 )2 Det ( 4 /)

(13.308)

The two prefactors together are therefore proportional to


[Det (iG0 )]1/2 Det 1/2 ( 4 /)

1
.
Det ( 2 )

(13.309)

Recalling the discussion from Eq. (13.128) to (13.132), we see that the associated
free energy is
F0 = log Z0 = 2

X
k

i
h
(k)
+ log 1 eh(k) .
2

(13.310)

It contains precisely the energy of the two physical transverse photons. The unphysical polarizations have been eliminated by the Fadeev-Popov determinant in
(13.307).
We now understand why the Gupta-Bleuler formalism failed to get the correct
vacuum energy in Eq. (7.496). It has no knowledge of the Fadeev-Popov determinant.
Taking current conservation into account, the exponent in (13.307) reduces to
j (x)G0 (x, x )j (x ) = j (x)

i
j (x),
2

(13.311)

so that the generating functional becomes


1

Z[j] = const Det 1 ( 2 )e 2

d4 xd4 x j (x)(i/ 2 )j (x )

(13.312)

where const is an infinite product of identical constant factors.


Let us see what happens in the other gauges (13.285). The results in the Lorentz
gauge are immediately obtained by going to the limit 0 in the previous calculation. In this limit, the functional 4 coincides with 1 , up to a trivial factor.

927

13.13 Gauge Fields and Gauge Fixing

The Hamiltonian and the axial gauges are quite similar, so we may only discuss
the one of them. In the Hamiltonian gauge where the Fadeev-Popov determinant is
given by (13.296) we have, instead of the generating functional (13.307),
1

Z[j] = (Det G0 )1/2 Det (0 )e 2

d4 xd4 x j i (x)G0 ij (x,x )j j (x )

(13.313)

where the matrix G0 ij (q) has only spatial entries, and is equal to
G0 ij (q) = iDij1 (q)

(13.314)

Dij (q) = q 2 ij + qi qj

(13.315)

with
this being the spatial part of the 4 4-matrix (13.284). With the help of 3 3
projection matrices
PijT (q) = ij qi qj /q2

(13.316)

PijL (q) = qi qj /q2 ,

(13.317)

this can be decomposed as follows:


Dij (q) = q 2 PijT (q) + q 0 2 PijL (q).

(13.318)

The inverse of this is


Dij1 (q)

1
qi qj
1
1
= 2 PijT (q) + 0 2 PijL (q) = 2 ij 0 2
q
q
q
q

(13.319)

In the exponent of (13.313) we have to evaluate iDij1 (q) between two conserved
currents, and find
"

1
i
j i (q)G0 ij (q)j j (q) = 2 j (q) j(q) 0 2 q j (q) q j(q) .
q
q

(13.320)

Inserting the local current conservation law j (x) = 0 in momentum space


q j(q) = q 0 j 0 (q),

(13.321)

we obtain
j i (q)G0 ij (q)j j (q) =

i
j (q)j (q).
q2

(13.322)

Let us calculate the functional determinant Det G0 in this gauge. From


(13.319) we see that G0 has two eigenvectors of eigenvalue i/q 2 , and one eigenvector of eigenvalue i/q 0 2 . Hence:
Det G0 =

1
.
Det (i 2 )2 Det (i02 )

(13.323)

928

13 Functional Integral Representation of Quantum Field Theory

The two prefactors in (13.313) together are therefore proportional to


(Det G0 )1/2 Det (0 )

1
,
Det (i 2 )

(13.324)

which is the same as (13.309). Thus the generating functional (13.313) agrees with
the previous one (13.312)
Let us also show that Coulomb gauge leads to the same result. We rewrite the
exponent in (13.312) in momentum space as
Z

Z
1
1
1
d4 q
d xj (x) 2 j (x) =
c2 (q) 2 (q) c2 j(q) 2 j(q) .
4

(2)
q
q
4

"

(13.325)

Here we keep explicitly the light velocity c in all formulas since we want to rederive
the interaction equivalent to Eq. (12.83) where c is not set equal to unity. Writing
the denominator as q 2 = q02 q2 and j 2 (q) = j2L (q) + j2T (q) with jL (q) = q j(q)/|q|
and jT (q) jL (q) = 0, we can bring (13.325) to the form
Z

d4 q
1
1
1
c2 (q) 2
(q) jL (q) 2
jL (q) jT (q) 2
jT (q) . (13.326)
4
2
2
(2)
q q
q0 q
q q2
"

Now we use the current conservation law cq0 (q) = q j(q) to rewrite (13.326) as
Z

2
1
1
1
d4 q
2
2 q0
2
c
(q)
(q)

c
(q)
(q)

c
j
(q)
jT (. 13.327)
T
(2)4
q 2 q2
q2
q02 q2
q 2 q2

"

The first two terms can be combined to

d4 q 2
1
c (q) 2 (q).
4
(2)
q

(13.328)

The inverse Fourier transform of this times e2 /c2 yields precisely the first (Coulomb)
term in the interaction (12.83). The third term in (13.327) is the result of integrating
out the transverse vector potential (12.6) in the interaction (12.83), which produces
the interaction (13.319).

13.14

Nontrivial Gauge and Fadeev-Popov Ghosts

The Fadeev-Popov determinants in the above examples were all independent of the
fields. As such they were irrelevant for the calculation of any Green function. This
is not always true.
Consider the following nontrivial gauge-fixing functional (see also [96])
h

F [A] = (A)2 + gA2 .


As in Eqs. (13.288)(13.291), we calculate
[A] =

DF [A ] =

(13.329)

D [A + gA2 + 2 + 2gA + g()2 ]. (13.330)

929

13.14 Nontrivial Gauge and Fadeev-Popov Ghosts

This path integral can trivially be performed in analogy with the ordinary integral
1
dx (ax + bx2 ) = ,
a

(13.331)

and the result is


[A] = Det ( 2 + 2gA )1 .

(13.332)

The generating functional is therefore


Z[j] =

 Z

DA Det ( + 2gA ) [ A + gA ] exp i

1 2
j A .
d x F
4
(13.333)
4



Contrary to the previous gauges, the determinant is no longer a trivial overall factor
but depends now functionally on the field A . It can therefore no longer be brought
outside the functional integral.
There is a simple way of including its effect within the usual field-theoretic
formalism. On e introduces an auxiliary Fadeev-Popov ghost field. We may consider
the determinant as the result of a fluctuating complex fermion field c with a complexconjugate c , and write
2

Det ( + 2gA ) =

Dc Dcei

d4 x(c c2gA c c)

(13.334)

Note that the Fermi fields are necessary to produce the determinant in the numerator; a Bose field would have put it into the denominator. A complex field is taken
to make the determinant appear directly rather than the square-root of it.
The ghost fields interact with the photon fields. This interaction is necessary in
order to compensate for the interactions induced by the constraint [A + gA2 ] in
the functional integral.
It is possible to exhibit the associated cancellations order by order in perturbation
theory. For this we have to bring the integrand to a form in which all fields appear
in the exponent. This can be achieved for the -functional by observing that the
same representation (13.333) would be true with any other choice of gauge, say
[ A (x) + gA2 (x) (x)]

(13.335)

since this would lead to the same Fadeev-Popov ghost term (13.334). Therefore
we can average over all possible functions (x) with a Gaussian weight and replace
(13.335) just as well by
Z

De 2

d4 x2 (x)

[ A (x) + gA2 (x) (x)].

(13.336)

Now the generating functional has the form


Z[j] =

DA

Dc Dcei

d4 x(Lj A )

(13.337)

930

13 Functional Integral Representation of Quantum Field Theory

with a Lagrangian
2
1 
1 2
A + gA2 c c + 2gA c c.

L = F
4
2

(13.338)

The photon and ghost propagators are


A (x)A (0) =

"
#
(13.339)
d4 k ikx i
k k
g

(1

)
,
e

(2)4
k2
k2

(13.340)

Contrary to the previous gauges, there are now photon-ghost and photon-photon
interaction terms
g
g 2  2 2
A
A A2
+ 2gA c c

(13.341)

with the corresponding vertices

It can be shown that the Fadeev-Popov ghost Lagrangian has the property of canceling all these unphysical contributions order by order in perturbation theory. We
leave it as an exercise to show, for example, that there is no contribution of the
ghosts to photon-photon scattering up to, say, second order in g, and that selfenergy corrections to the photon propagator due to photon and ghost loops cancel
exactly.
In the context of quantum electodynamics, there is little sense in using a gauge
fixing term (13.329). The present discussion is, however, a useful warm-up exercise
to gauge-fixing procedures in nonabelian gauge theories, where the Fadeev-Popov
determinant will always be field dependent.

931

13.15 Functional Formulation of Quantum Electrodynamics

Of course, also the previous field-independent Fadeev-Popov determinants can be


generated from fermionic ghost fields. The determinant 1
4 in (13.297), for example,
can be generated from a complex ghost field c(x) and its complex conjugate c by a
functional integral
1
4

Dc Dc ei

d4 x c c/

(13.342)

It should be pointed out that the sign of the kinetic term of these c-field Lagrangians are opposite to those of a normal field. If these fields were associated with
particles, their anticommutation rules would carry the wrong sign and the states
would have a negative norm. Such states are commonly referred to as ghosts, and
this is the reason for the name of the fields c, c .
Note that the determinant cannot be generated by a real fermion field via a
functional integral
Z
R
?
i d4 x 2 c 2 c/
1
=
Dc
e
.
(13.343)
4

The reason is that the differantial operator 4 is a symmetric functional matrix, so


that the exponent vanishes after diagonalization by an orthogonal tranformation.
In the language of ghost fields, the mistake in calculating the vacuum energy
in Eq. (7.496) when quantizing the electromagnetic field via the Gupta-Bleuler formalism can be phrased as follows. When fixing the gauge in the action (7.366) by
adding an Lagrangian density LGF (10.75), we must also add a ghost Lagrangian
density

Lghost = c c/ .
(13.344)

The ghost fields have to be quantized canonically, and the physical states must
satisfy, beside the Gupta-Bleuler subsidiary condition in Eq. (7.492), the condition
of being a vacuum to the ghost fields:
c|phys i = 0.

(13.345)

The Fadeev-Popov formalism is extremely useful offering many other possibilities


of fixing a gauge and performing the functional integral for the generating functional
over all A -components.

13.15

Functional Formulation of Quantum Electrodynamics

For quantum electrodynamics, the functional integral from which we can derive all
time-ordered vacuum expectation values reads
Z[j, , ] =

DD DA
DD eiAi

(x)(x)] i
d4 x [(x)(x)+

d4 x j (x)A (x)

(13.346)

where A is the sum of the free-field action (12.84) and the minimal interaction
(12.85), which we shall write here as
A=

e
1
d x i/
A
/ m F F D A + D 2 /2 ,
c
4
4

(13.347)

932

13 Functional Integral Representation of Quantum Field Theory

The Dirac field appears only quadratically in the action. It is therefore possible to
integrate it out using the Gaussian integral formula (13.94), and we obtain
Z[j, , ] = Det (i/
M)

DA DD eiA

d4 x d4 x (x)G0 (x,x )(x )(x)i

d4 x j (x)A (x)

(13.348)

where A is the action


A=

1
d4 x F F D A + D 2 /2 + Aeff [A].
4


(13.349)

The effects of the electron are collected in the effective action


eff

eiA

= exp [Tr log (i/


e/
A M) Tr log (i/
M)] ,

(13.350)

where TrO combines the functional trace of the operator O and the matrix trace
in the 4 4 space of Dirac matrices. We have performed a subtraction of the
infinite vacuum energy caused by the filled-negative energy states. The subtraction
is compensated by the determinant in the prefactor of Eq. (13.348).

13.15.1

Decay Rate of Dirac Vacuum in Constant Electric Field

An important application of the functional formulation (13.348) of quantum electrodynamics was made by Heisenberg and Euler [97, 98]. They observed that in a
constant external electric field, the vacuum becomes unstable. There exists a finite
probability of creating an electron-positron pair. This process has to overcome a
large energy barrier 2Mc2 , but is the pair is separated sufficiently far, the total
energy of the pair can be made arbitrarily low, so that the process will occur with
nonzero probability. The rate can be derived from the result (6.255). For a large
total time t, the time dependence of an unstable vacuum state will have the form
ei(E0 i/2)t , where E0 is the vacuum energy, the desired decay rate, and t the
total time over which the amplitude is calculated. If Leff is the effective Lagrangian
density causing the decay, and Aeff the associated action, we identify

Aeff
= 2 Im
= 2 ImLeff .
V
V t

(13.351)

We now make use of the fact that due to the invariance under charge conjugation,
the right-hand side of (13.350) can depend only on M 2 . Thus we also have
exp [Tr log (i/
e/
A M)] = exp [Tr log (i/
e/
A +M)]


1
= exp
[Tr log (i/
e/
A M) (i/
e/
A +M)] ,(13.352)
2
and may use the product relation (6.108) to calculate (13.350) from half the trace
log of the Pauli operator in Eq. (6.242) to find the effective action
o
n
1
e
1
iAeff = Tr log [i eA(x)]2 F M 2 Tr log 2 M 2 .
2
2
2
(13.353)


933

13.15 Functional Formulation of Quantum Electrodynamics

We now use the integral identity


log

Z
a
d ia
=
[e eib ]
b

(13.354)

and relation (6.207) to rewrite (13.353) as


1
2

iAeff =

d i (M 2 i)
2 e e
2
e
trhx|ei {[ieA(x)] + 2 F } ei |xi. (13.355)

Recalling (6.190) and (6.207), the first, unsubtracted, term can be re-expressed as
iAeff
1

1
=
2

1
d

trhx|ei H |xi =

d
trhx, |x 0i.

(13.356)

Inserting (6.255), and subtracting the field-free second term in (13.355), we obtain
the contribution to the effective Lagrangian density
L

eff



eE
1 Z d
2
1 ei (M i) .
=
2
3
2(2) 0
tanh eE

(13.357)

The integral over is logarithmically divergent at = 0. We can separate the


divergent term by a further subtraction, splitting
eff
Leff = Leff
div + LR

(13.358)

into a convergent intgral


Leff
R

1
=
2(2)2

d
3

e2 E 2 2 i (M 2 i)
eE
e
,
1
tanh eE
3
!

(13.359)

and a divergent one


Leff
div

e2 2 1
= E
2
3(2)2

d i (M 2 i)
e
,

(13.360)

The latter is proportional to the electric part of the original Maxwell Lagrangian
density in (4.234). It can therefore be removed by the following formal manipulation
called renormalization. We add (13.360) to the Maxwell Lagrangian density, and
define a renormalized charge eR by the equation
1
1
1
= 2+
2
eR
e
12 2

d i (M 2 i)
1
e

Z3 e2

(13.361)

to obtain the modified electric Lagrangian density


LE =

e2 2
E .
2e2R

(13.362)

934

13 Functional Integral Representation of Quantum Field Theory

Now we redefine the electric fields by introducing renormalized fields


ER

e
E,
eR

(13.363)

and identify these with the physical fields. In terms of these, (13.362) takes again
the usual Maxwell form
1
LE = ER2 .
2

(13.364)

The finite effective Lagrangian density (13.359) possesses an imaginary part


which by Eq. (13.351) determines the decay rate of the vacuum per unit volume

1
= Im
(2)2

d
3

e2 E 2 2
eE
1
tanh eE
3

ei (M

2 i)

(13.365)

For comparison we mention that for a charged boson field, the expression (13.350)
defining the effective action is replaced by
eff

eiA

= exp Tr log [i eA(x)]2 M 2 + Tr log 2 M 2

i

(13.366)

Hence the last factor 4 cosh eE in (6.255) is simply replaced by 2, and the the
unsubtracted effective action (13.365) becomes
iAeff
u =

i Z d eE
d
2
hx, |x 0i =
ei (M i) ,
2
3

4(2) 0 sinh eE

(13.367)

implying a twice subtracted effective Lagrangian density


Leff
R

1
=
4(2)2

d
3

e2 E 2 2 i (M 2 i)
eE
e
,
1+
sinh eE
6
!

(13.368)

and a charge renormalization


1
1
1
= 2 2
2
eR
e
6

d i (M 2 i)
1
e

Z3 e2

(13.369)

The decay rate per unit volume


KG
1
= Im
V
4(2)2

d
3

e2 E 2 2
eE
1+
sinh eE
6

ei (M

2 i)

(13.370)

The integrands in (13.365) and (13.370) are even in , so that the integrals for
the decay rate can be extended symmetrically to run over the entire -axis. After
this, the countour of integration can be closed in the lower half-plane and the the
integral can be evaluated by Cauchys residue theorem. To find the pole terms we
expand the integrand for fermions in Eq. (13.365) as

X
X
(e )2
2
e
,
1 =
=
2
2
2 2
2
2
tanh e
n= (e ) + n
n=1 + n

n
.
e

(13.371)

13.15 Functional Formulation of Quantum Electrodynamics

935

The relevant poles lie at = in and yield the result

e2 E 2 X

1 nM 2 /eE
=
e
.
3
V
4 n=1 n2

(13.372)

An technical remark is necessary at this place concerning the integral (13.370).


At first sight it may appear as if the second subtraction term in the integrand
e2 E 2 2 /3 can be omitted [99]. First, it is unnecessary to arrive at a finite integral in
the imaginary part, and second, it seems to contributes only to the real part, since
for all even powers > 2, the integral
Z

d i (M 2 i)
1
e
=
( 2).
3

(iM 2 )2

(13.373)

is real. The limit at hand 2, however, is an exception since for 2, the


integral possesses an imaginary part due to the divergence at small
1

( 2) log M 2 i + O( 1).
2
2
(iM )
2

(13.374)

The right-hand side of Eq. (13.372) is a polylogarithmic function (2.273), so that


we may write
e2 E 2

2
=
2 (eM /eE ).
(13.375)
3
V
4
For large fields, this has the so-called Robinson expansion 9
(e ) = (1 )1 + () +

1
()k ( k).
k!
k=1

(13.376)

This expansion plays an important role in the discussion of Bose-Einstein


condensation10 For 2, the Robinson expansion becomes
2 (e ) =

2
2 3
+ (1 + log )
+
+ O(5 ).
6
4
72

(13.377)

Hence we find the strong-field expansion

#
!
!2
!3
"

1 M 2
M 2
e2 E 2 2
M 2
1 M 2

=
+
1
+
log
+
+
.
.
.
.

V
4 3 6
eE
eE
4 eE
72 eE
(13.378)
For bosons, we expand

X
X
(eE )2
2
n
e
n
(1)
,

1 = 2 (1)n
=
2
.
n
sinh e
(eE )2 + n2 2
2 + n2
eE
n=1
n=1
9

(13.379)

J.E. Robinson, Phys. Rev. 83 , 678 (1951).


See Chapter 7 in the textbook H. Kleinert, Path Integrals in Quantum Mechanics, Statistics,
Polymer Physics, and Financial Markets, World Scientific, Singapore 2004,
Third extended edition, pp. 11450. (http://www.physik.fu-berlin.de/~kleinert/b5)
10

936

13 Functional Integral Representation of Quantum Field Theory

Comparison with (13.371) shows that the bosonic result for the decay rate will
differ from the fermionic (13.372) only by an alternating sign a factor 1/2, due to
the absence of spin, and a minus sign for bosons [recall (13.370)]. Thus we find the
decay rate per volume

KG
1
e2 E 2 1 X
2
(1)n1 2 enM /eE .
=
3
V
4 2 n=1
n

(13.380)

The sum can again be expressed in terms of the polylogarithmic function (13.376)
as follows:
k

X
X
zk
z2
(1)k1 z k
=

2
= (z) 21 (z 2 ).
(z)

k
k
(2k)
k=1
k=1
k=1

(13.381)

For z = e 1, the Robinson expansion (13.377) yields


2
2 3
5
2 (e ) =
log 2 +

+
+ O(8 ).
12
4
24 960

(13.382)

The expansion 2 (e ) replaces the curly bracket in (13.378), with = M 2 /eE.

13.15.2

Constant Electric and Magnetic Background Fields

In the presence of both E and B fields, Eq. (6.254) reads


e
exp i F
2


0
eie (BiE)
0
eie (B+iE)

(13.383)

The trace of this can be found by adding the traces of the 2 2 block matrices
ee (iBE) separately. These are equal to ee1 + ee2 and their complex conjugates, respectively, where 1 , 2 are the eigenvalues of the matrix  (iB E),
which are
q

(13.384)
1,2 = (E + iB)2 = E2 B2 + 2iE B
Thus we find

e
tr exp i F
2


= 2(cos e1 + cos e1 ).

(13.385)

The eigenvalues are, of course, Lorentz-invariant quantities. They depend only on


the two quadratic Lorentz invariants of the electromagnetic field: the scalar S and
the pseudoscalar P defined by

1
1 2
E B2 ,
S F F =
4
2

1
P F F = E B.
4

(13.386)

In terms of these, Eq. (13.384) reads



1,2 = 2 S + iP ,

(13.387)

13.15 Functional Formulation of Quantum Electrodynamics

937

which can be rewritten as


1/4

1,2 = 2 S 2 + P 2
(cos /2 + i sin /2) ,

(13.388)

where
tan =

P
,
S

(13.389)

so that
cos =
and cos /2 =

S
,
S2 + P 2

sin =

(1 + cos )/2, sin /2 =


(

cos /2
sin /2

P
,
S2 + P 2

(13.390)

(1 cos )/2 become

S2 + P 2 S

2 (S 2 + P 2 )1/4

(13.391)

We shall abbreviate the result (13.388) as


1,2 = ( + i) ,

(13.392)

where
( )

S2

P2

r
1 q 2
(E B2 )2 + 4(E B)2 (E2 B2 ). (13.393)
S =
2

In terms of and , the invariants (13.386) becomes




1 2
1 2
1
E B2 =
2 ,
S F F =
4
2
2

1
P F F = E B = ,
4
(13.394)

and the trace (13.385) becomes


e
tr exp i F
2


= 2 cosh ( + i) + 2 cosh ( i) = 4 cosh cos . (13.395)

Some special cases will simplify the upcoming formulas:


1. If B = 0, then reduces to |E|, whereas for E = 0, reduces to |B|.
2. If E 6= 0 and B 6= 0 are orthogonal to each other, then either = 0 for E > B
or = 0 for B > E. The formulas are then the same as for pure electric or
magnetic fields.
3. If E 6= 0 and B 6= 0 are parallel to each other, then = |E| =
6 0, = |B| =
6 0.

938

13 Functional Integral Representation of Quantum Field Theory

In all these cases, the calculation of the exponential (13.383) can be done very
simply directly. Take the third case. Due to rotational symmetry, we can assume the
fields to point in the z-direction, B = Bz, E = Ez. Then the exponential (13.383)
reads
eie 3 (BiE)
0
ie 3 (B+iE)
0
e

e
exp i F =
2


eie(BiE)
0
0
0
0
eie(BiE)
0
0
0
0
eie(B+iE)
0
0
0
0
eie(B+iE)

(13.396)

and has a trace


e
tr exp i F
2


= 2 cosh (E + iB) + 2 cosh (E iB) = 4 cosh E cos B ,

(13.397)
in agreement with (13.395). In fact, given an arbitrary constant field configuration
B and E, it is always possible to perform a Lorentz transformation to a coordinate
frame in which the tranformed fields, call them BCF and ECF , are parallel. This
frame is called center-of-fields frame. The transformation has the form (4.275) and
(4.276) with a velocity of the transformation determined by
EB
v/c
=
.
2
1 + (|v|/c)
|E|2 + |B|2

(13.398)

By Lorentz invariance,
E2 B2 = E2CF B2CF,

E B = ECF BCF ,

(13.399)

which shows that |ECF | and |BCF| in Eq. (13.397) coincide with and in
Eq. (13.394). This is the reason why Eq. (13.397) gives the general result for arbitrary constant fields if E and B are replaced by |ECF | = and |BCF | = .
Let us now calculate the determinantal prefactor in Eqs. (6.215) and (6.226) for
a general constant field strength F . The basic matrix to be diagonalized is eeF
in Eq. (6.246). For an electric field pointing in the z-direction, this has the form
eiM3 eE . In contrast to (6.254), this is a boost matrix with rapidity = eE in
the defining 4 4 -representation [compare (4.63)]. The fact that the rapidity in
the Dirac representation (6.253) was twice as large has its origin in the value of the
gyromagnetic ratio 2 of the Dirac particle in Eq. (6.120).
If the field points in any direction, the obvious generalization is eiMEe . In the
presence of a magnetic field, the generators of the rotation group (4.57) enter and
(6.246) can be written as eeF = eie(ME+LB) . This is the defining four-dimensional
representation of the complex Lorentz transformation, whose chiral Dirac representation was written down in (13.383), apart from the factor 2 in rapidity and rotation
vector.

939

13.15 Functional Formulation of Quantum Electrodynamics

As before in the Dirac representation, much labor is saved by working in the


center-of-fields frame where electric and magnetic fields are parallel, point in the
z-direction, and have the invariant values and , respectively. The associated
transformation eeF has then the simple form

eeF = ei(M3 +L3 )e =

cosh e
0
0
sinh e

0
0
sinh e
cos e sin e
0
sin e
cos e
0
0
cosh e

(13.400)
From this we find for sin eF = [ei(M3 +L3 )e ei(M3 +L3 )e ]/2 the matrix

sin eF =

0
0
0
sinh e

and from this

eF =

0
0
sinh e
0
sin e
0
sin e
0
0
0
1

0
0
e
0 e
0

.
e
0
0
0
1

0
0
0
e

(13.401)

(13.402)

This leads to the desired prefactor in the amplitude (6.226)


det

1/2

sinh eF
eF

= e 2

e 2
.
sinh sin e

(13.403)

Thus we can obtain the imaginary part of the vacuum energy in a constant
electromagnetic field if we simply replace the term eE coth eE in the integrand of
the rate formula (13.365) as follows:
eE
e
e

tanh eE
tanh e tan e

13.15.3

(13.404)

Decay Rate in Constant Electromagnetic Field

For a constant electromagnetic field the effective Lagrangian density (13.359) becomes, after the replacement (13.404)
Leff
R

1
=
2(2)2

e2 (2 2 ) i (M 2 i)
e
e
d
e
. (13.405)

3 tanh e tan e
3
#

"

The subtracted small- divergence lies in the integral [recall the equality 2 2 =
E2 B2 from Eq. (13.394)]
Leff
div =

1
e2 2
(E B2 )
2
3(2)2

d i (M 2 i)
e
.

(13.406)

940

13 Functional Integral Representation of Quantum Field Theory

which is proportional to the full original Maxwell Lagrangian density in (4.234), and
can therefore be absorbed into it by a renormalization of the charge as in (13.361)
and of the fields
e
e
B, ER
E.
(13.407)
BR
eR
eR
From twice the imaginary part of (13.405) we obtain the decay rate per unit volume extending Eq. (13.365) in an obvious manner. Inserting the expansion (13.371),
extending the integral over the entire -axis and rotating the contour of integration
as we did in evaluating (13.365), we find the generalization of Eq. (13.372) to constant electromagnetic fields

KG
e2 E 2 X
1
n/
2
=
enM /eE .
3
2
V
4 n=1 n tanh n/

(13.408)

For bosons obeying the Klein-Gordon equation, we obtain by analogy the extension of (13.380) to constant electromagnetic fields:

e2 E 2 1 X
(1)n1 n/

2
=
enM /e .
3
2
V
4 2 n=1 n
sinh n/

13.15.4

(13.409)

Effective Action in Purely Magnetic Field

If there is only a magnetic field, the integral representation (13.405) reduces to


[97, 98]
Leff
R

e2 B 2 2 i (M 2 i)
eB
e
.
1+
tan eB
3
!

1 Z d
=
2(2)2 0 3

(13.410)

The integral still contains a divergence at small, which is precisely the B-version of
Eq. (13.360). This is removed by the same subtraction as before. Going to renormalized quantities as in (13.361) and (13.407) and rotating the contour of integration
clockwise to move it away from the poles, as trivial change of the integration variable
leads to
Leff
R

e2 B 2
=
2(2)2

1 s sM 2 /eB
ds
e
.
coth s
2
s
s 3


(13.411)

This is a typical Borel tranformation of the expression in parentheses. This implies


that its power series expansion leads to coefficients of B 2n which grow like (2n)!.
The expansion has therefore a vanishing radius of convergence. It is an asymptotic
series, and we shall understand later the physical origin of this.
For Klein-Gordon particles, the result becomes
LR KG =

e2 B 2
(2)2

1 s sM 2 /eB
1
ds
e
.

+
s2 sin s s 6


(13.412)

941

13.15 Functional Formulation of Quantum Electrodynamics

13.15.5

Heisenberg-Euler Lagrangian

By expanding (13.404) in powers of e, we obtain





1 e
1
e2  2
e
4
2
4
2 2
4

+
5

=
+
3 tan e tanh e 3 3
45
3 


(13.413)
26 + 74 2 72 4 2 6 + . . . .
+ e6
945
Inserting this into (13.405) and performing the integral over leads to an expansion
in powers of the fields, whose lowest terms are, with e2 = 4:
o 163 n
o
22 n 2
2 2
2
2
2 3
2
2
2
+
.
(E
B
)
+
7(EB)
2(E
B
)
+
13(E
B
)(EB)
45M 4
315M 8
(13.414)
In each term we can express in terms of R = (1 + O()) and the fields in terms
of the renormalized fields via (13.407), and obtain the same series as in (13.414) but
for the physical renormalized quantities, plus higher-order corrections in for each
coefficients, which we ignore in this lowest-order calculation.
Each coefficient is exact to leading order in . To illustrate the form of the
higher-order corrections we include without derivation the leading correction into
the first term, which becomes (see Appendix 13C for details)

Leff
R =

Leff
R =

22
45m4e



1+

40
1315
(E2 B2 )2 + 7 1 +
(E B)2 + . . . .(13.415)
9
252


Electrons in Constant Magnetic Field


For Dirac particles in arbitrary constant fields we expand
e coth e =

X
2
(e )2
,
=
2
2 2
2
2
n= + n,
n= (e ) + n

X
(e )2
2
,
e cot e =
=
2
2
2 2
2
m= (e ) m
m= m,

n
,
e

n,

m,

(13.416)

m
.
e

(13.417)

The i accompanying the mass term in the -integral (13.365) is equivalent to re2
2
placing ei (M i) by ei (1i)M , implying that is to be integrated slightly below
the real axis. Equivalently we may shift the n, slightly upwards in the complex
plane to n, . This leads to an additional contribution to the action of a constant
eff
electromagnetic field Leff = Leff
div + LR with the logarithmically divergent part
Leff
div

1
=
2(2)2

X
1
1
d i (1i)M 2 X

e
2
2
2

m=1 m,
n=1 n,

(13.418)

and the finite part


Leff
R =

1
2(2)2

0
2

d
12 2
2
3
2
2
2
n,m= + n, m,

ei (1i)M .

2
n=0 n,

2
m=1 m,

(13.419)

942

13 Functional Integral Representation of Quantum Field Theory

Performing the sums

n=1

k
n,

e
=

k

(k),

m=1

1
k
m,

!k

(k),

(13.420)

we see that (13.418) coincides with (13.406), as it should. The remaining sum is
finite:
Leff
R

#
2
2
Z "

e M
e M
1
1 X

d 2
... ,
2
= 2
2
2 + 2
2 +i
8 n,m= n,
n,
+ m,
i
m, 0

(13.421)

where the dots indicate the subtractions. Now we decompose

=
2 + i
2 n,

P
,
( + n, ) i ( n, ) + 2
2
2
2
n,

(13.422)

where P indicates the principal value under the integral. The integrals over the
-functions contribute

X
e2 2 2 X
1
1
i
2
n, M 2
e
=i
enM /e .
L =
2
2
3
2
2
2
2
8 n>0,m6=0 n, + m,
8 n>0,m0 n + m
(13.423)
This leads to a decay rate per volume which agrees with (13.409) if we expand in
that expression [recall (13.371) and (13.351)]:
eff

X
X
(n/)2
n/
n2 2
= 1+
= 1+
.
2
2 2
2 2
2 2
tanh n/
m6=0 (n/) + m
m6=0 n + m

(13.424)

It remains to do the principal-value integrals. Here we use the formula11


J(z) P

where
Ei(z)

i
e
1 h z
z
e
Ei(z)
+
e
Ei(z)
,
=

2 z2
2
z

dt

X
P t
zk
e = log(z) +
t
k=1 k k!

(13.425)

(13.426)

is the exponential integral, being the Euler-Mascheroni constant


= 0.577216 . . . .

(13.427)

The function (13.425) has the small-z expansion


i
1h z
e log(z) + ez log(z) cosh z
2

X
X
z 2l+1
z 2l
cosh z +
sinh z,

l=1 (2l + 1)(2l + 1)!


l=1 (2l)(2l)!

J(z) =

11

(13.428)

I.S. Gradshteyn and I.M. Ryzhik, Table of Integrals, Series, and Products, Academic Press,
New York, 1980, Formulas 3.354.4, 8.211.

943

13.15 Functional Formulation of Quantum Electrodynamics

and behaves for large z like


J(z) =

1
6
120 5040
4 6 8 + ... ,
2
z
z
z
z

(13.429)

Using these formulas, we find from the principal-value integrals in (13.421) [100]:
P Leff =

1 2
e2 X
( an + 2 bn ),
4 4 n=1 n2

(13.430)

with
nm2
n/
Ci
an =
tanh n/
e


nm2
nm2
cos
+ si
e
e

1 n/
nm2
bn =
exp
2 tanh n/
e


nm2
Ei
e

nm2
sin
,
e


nm2
+ exp
e

(13.431)
nm2
Ei
e

!

where [103]
Ci(z)

dt
cos t = + log z +
t

dt
(cos t 1) , si(z)
t

dt
sin t.(13.432)
t

The prefactors come once more from sums of the type (13.424).
For large arguments of J(z), the expansion (13.429) becomes

1 X
1

P L = 2
2
2
8 n,m=n, + m,
eff

"

1
6
120
+ 4
+ 6
+ . . . . (13.433)
2
4
8
n, M
n, M
n, M 12

Applying the summation formulas (13.420), this is seen to agree with the small-field
expansion (13.414).
For E = 0, the sum (13.430) contains only the n = 0 -terms. If B is large, the
n-sum is dominated by the first term in the expansion (13.428), and yields
P L

eff

X
1
1
= 22
log(m, M 2 ).
2
8 m=1 m,

(13.434)

The leading part of this is


P Leff =

M2
1 2 2
eB
e2 B 2
1 X
log
+
.
.
.
=

e
B
log
,
4 2 m=1 2 m2
eB
24 2
M2

(13.435)

as derived first by Weisskopf [101].


Note that this logarithmic behavior can be understood as the result of a lowest
expansion in e2 of an anomalous power behavior of the effective acion
1
2
2
Leff = (E 2 B 2 )1+e /12 .
2

(13.436)

Due to th smallness of e2 , this power can be observed only at very large field
strengths. In self-focussing materials, however, it is visible in present-day laser
beams [113].

944

13 Functional Integral Representation of Quantum Field Theory

13.15.6

Alternative Derivation for Constant Magnetic Field

The case of a constant magnetic field can also be treated in a different way [102].
The eigenvalues of the euclidean Klein-Gordon operator of a free scalar field
OKG = h2 2 + M 2 c2

(13.437)

p = p 2 + M 2 c2 .

(13.438)

are
If a constant magnetic field is present which points in the z-direction, the kinetic
energy p21 + p22 in the xy-plane changes to the Landau energies, i.e.,
p21 + p22 2M

p21 + p22
1
, n = 0, 1, 2, . . . ,
= 2M h
L n +
2M
2


(13.439)

where L = eB/Mc is the Landau frequency. Thus the eigenvalues become


n,p = p2 + 2

h
e
B(n + 1/2) + M 2 c2 .
c

(13.440)

where p is the two-dimensional momentum in the x3 x4 -plane.


For a Dirac electron satisfying the Pauli equation (6.111), this changes to
n,p = p2 + 2

h
e
B(n + 1/2 + s3 ) + M 2 c2 ,
c

s3 = 1/2.

(13.441)

In this expression, the famous g-factor which accounts for the anomalous magnetic
moment and approximate by the Dirac value 2. Radiative corrections would insert a
factor g/2 in front of 3 , with g = 2 + / + . . . . The contribution of the electrons
to the effective action in Eq. (13.353) can therefore be written as
iA

eff

Z
Z
1 X
d 2 p
d4 p
=
log

log p .
n,p

2 n, (2h)2
(2h)4
!

(13.442)

These are divergent expressions. In order ot deal with them efficiently it is useful to
introduce the Hurwitz -function of number theory [106, 107]:
H (s, z) =

1
,
s
n=0 (n + z)

Re(s) > 1 ,

v 6= 0, 1, 2, . . . .

(13.443)

This Hurwitz -function can be analytically continued in the s plane to define an


analytic function with a single simple pole at s = 1. In quantum field theory, one
may extend this definition to an arbitrary eigenvalue spectrum of an operator O as
O (s) =

1
,
(/d )s

(13.444)

945

13.15 Functional Formulation of Quantum Electrodynamics

where is some mass parameter to make the eigenvalues of mass dimension d dimensionless. For large enough s, this is always convergent. By analytic continuation
in s, one can derive a finite value for the functional determinant of O:
Det O = eTr log O ,

Tr log O = O (0).

(13.445)

For the Dirac spectrum (13.441), the -function becomes, in natural units,
XZ
2
+ eB(2n + 1 1)
eB X
d 2 k M 2 + k
D (s) =
2
2 n=0
(2)
2

"

#s

(13.446)

where the prefactor of eB/2 is the degeneracy of the Landau levels which ensures
P
that theR sum (eB/2) n Rconverges in the limit B 0 against the momentum
integral dk 1 dk 2 /(2)2 = dk 2 /4. Performing the integral over the momenta p
yields
D (s) =

o
B 2s 1 n 2
1s
2
1s
. (13.447)
[M
+
2eBn]
+
[M
+
2eB(n
+
1)]

8 2
s1

This can be expressed in terms of the Hurwitz -function (13.443) as


M4
D (s) =
4 2

eB
M2

2 X

n=0

2
2eB

!s

From this we obtain

1
M2
2H s 1,
s1
2eB

(eB)2
M2

=
2H 1,
4 2
2eB
(

D (0)

1
M2
+ +
6
2eB

!2 "
1 + log

M2

2eB

M2
M2

log
2eB
2eB
2
2eB

!#

!1s
. (13.448)

(13.449)

Choosing = M and subtracting the zero-field contribution, which is simply


3M 4 /32 2 , we find
Leff

M2
(eB)2

1,
=
H
2 2
2eB
(

+ H

M2
M2
1,
log
2eB
2eB
!

1
1
+
12 4

M2
2eB

!2

(13.450)
where we have used the property [104, 106]
1
z z2
+ .
(13.451)
12 2
2
Contact with the previous result in (13.411) is established with the help of the
integral representation of the Hurwitz -function [104]:
H (1, z) =

1 Z ez t ts1
dt , Re(s) > 1 , Re(z) > 0
(s) 0 1 et


Z
1 t
z s sz 1s 2s1 dt 2z t
z 1s
coth
t

,(13.452)
+
+
+
e

=
s1
2
12
(s) 0 t1s
t 3

H (s, z) =

946

13 Functional Integral Representation of Quantum Field Theory

where the second expression is valid for Re(s) > 2 [111]. Thus we can evaluate
the derivative at s = 1 as follows:

H
(1, z)

1
1 t
z2
1 Z dt 2z t
=
coth t
.(13.453)

H (1, z) log z
e
2
12
4
4 0 t
t 3


Inserting this into (13.450) we recover exactly the previous Eq. (13.411) .
Let us now derive the strong-field limit. For this we use the following relation
between the Hurwitz -function and the -function [104, 106]:

H
(1, z)

z
z
= (1) log(2) (1 z) +
2
2

log (x)dx .

(13.454)

This identity follows from an integration of Binets integral representation [108] of


log (z). Thus we can write

M2
3
(eB)2 1

(1)

+
Leff =
2
2 12
4eB 4

1
M2
1
+ +

12 4eB 2

M2
2eB

M2
2eB

!2

M2
log
2eB

!2

M2
log(2)
4eB

M 2 /2eB

dx log (x) . (13.455)

In the strong-field limit, the range of integration in the last term vanishes, so we
can use the Taylor expansion [104, 105] of log (x):
log (x) = log x x +

(1)n
(n) xn
n
n=2

(13.456)

where (n) is the usual Riemann -function. This leads to :


L

eff

(eB)
1
M2
3

=
+ (1)
+
2

2
12
4eB 4

M
1
1

+ +
12 4eB 2

M2
M2
+
1 log
2eB
2eB
"

M
2eB

!#

M2
2eB

!2

M2
log
2eB

(1)n (n)
+
n=2 n(n + 1)

The leading behavior in the strong-field limit is


Leff =

!2

(eB)2
2eB
log 2 + . . . ,
2
24
M

in agreement with Weisskopfs result (13.435).

M2
log(2)

4eB
!2

M2
2eB

M2
2eB

!n+1

(13.457)

(13.458)

947

13.15 Functional Formulation of Quantum Electrodynamics

Charged Scalar Field in Constant Magnetic Field


For a charged scalar field obeying the Klein-Gordin equation there is no spin sum
and the -function (13.446) becomes
Z
2
eB X
d 2 k M 2 + k
+ eB(2n + 1)
KG (s) =
2 n=0 (2)2
2

"

(eB)2
=
4 2

"

2
2eB

!s

#s

1
1
M2
H s 1, +
(s 1)
2 2eB

!#

(13.459)

Setting again = m, and subtracting the zero field contribution 3m4 /64 2 , we
obtain
Leff
KG

1
M2
(eB)

1, +

=
4 2 H
2 2eB
2

M2
+ 1 + log
2eB
"

!#

3
+
4

M2
2eB

1
M2
1, +
2 2eB

!2

!)

(13.460)

In the Bose case, the equivalence with the previous result is shown with the help of
the integral representations of the Hurwitz -function [compare (13.452)]
1 Z z t s1 et/2
1
e t
dt , Re(s) > 1, Re(z) >
t
(s) 0
1e
2


1s
s1 Z
1s
dt 2z t
sz
2
1 t
1
z
,(13.461)

+
e
+
=
1s
s1
24
(s) 0 t
sinh t t 6

H (s, 1/2 + z) =

where the second expression is valid for Re(s) > 2, where the subtracted integral
converges. We can therefore use it to find the derivative at s = 1 required in
Eq. (13.460):
1
z2 z2
(1 + log z)
+ log z
24
4
2


Z
1 dt 2z t
1
1 t

.
e
+
4 0 t2
sinh t t 6

H
(1, 1/2 + z) =

(13.462)

Inserting this into (13.460), we recover exactly the previous result (13.412).
In order to find a strong-field expansion of Leff
KG we use the following relation
between the Hurwitz -function and the log of the -function [104, 108]:
1
z
log 2 z 2

H
(1, 1/2 + z) = (1) log 2
+
+
2
2
24
2

dz log (x+ 12 ), (13.463)

where we have used the formula [110, 111, 112]


Z

1
2

log (x) dx =

1
1 3
5
log 2 + log + (1) .
24
4
8 2

(13.464)

948

13 Functional Integral Representation of Quantum Field Theory

Then we expand


log x + 1/2 =

X
(1)n1 (1 2n )
1
log ( + 2 log 2) x +
(n) xn , (13.465)
2
n
n=2

and obtain the strong-field expansion

(eB)2 5
=
4 2 4

Leff
KG

M2
2eB

!2

1
1
+
24 2

M2
2eB

1
log 2
M2
1
(1)

log 2
2
24
4eB
2

M2
2eB

(1)n1 (1 2n )(n)
+
n(n + 1)
n=2

Appendix 13A

M2
1 + log
2eB

M2
2eB

!n+1

!2

!#

( + 2 log 2)

(13.466)

with the leading behavior


Leff
KG =

!2 "

(eB)2
2eB
log 2 + . . . .
2
96
M

(13.467)

Propagator of the Bilocal Pair Field

Consider the Bethe-Salpeter equation (4.32) with a potential V instead of V


= iV G0 G0 .

(13A.1)

Take this as an eigenvalue problem in at fixed energy-momentum q = (q 0 , q)=


(E, q) of the bound states. Let n (P |q) be all solutions, with eigenvalues n (q).
Then the convenient normalization of n is:
i

d4 P
q
q

(P
|q)
G
+
P
G
P n (P |q) = nn .
0
0
n
(2)4
2
2


(13A.2)

If all solutions are known, there is a corresponding completeness relation (the sum
may comprise an integral over a continuous part of the spectrum)
i

X
n

G0

q
q
+ P G0
P n (P |q)n (P |q) = (2)4 (4) (P P ). (13A.3)
2
2


This completeness relation makes the object given in (4.121) the correct propagator
of . In order to see this write the free action A2 [ ] as
1
1
+ iG0 G0
A2 =
2
V


(13A.4)

where we have used V instead of V . The propagator of would have to satisfy




= i.
+ iG0 G0
V


(13A.5)

Appendix 13A

Propagator of the Bilocal Pair Field

949

Performing this calculation on (4.48) one has, indeed, by virtue of (13A.1) for n , n :


X n n
1
+ iG0 G0 i
V
n n (q)
1
X
+ iG0 G0 n n
V n n
= i
n (q)
n

= i

X
n

= i i

n(q) + 1
(iG0 G0 n n )
n (q)

X
i

G0

G0 n n

= i.

(13A.6)

Note that the expansion of the propagator in powers of

=i

X
k

X
n

n (q)

!k

n n

(13A.7)

corresponds to the graphical sum over one, two, three, etc. exchanges of the potential
V . For n = 1 this is immediately obvious due to (13A.1):
i

X
n

n n =
n (q)V G0 G0 n n = iV.
n (q)
n (q)

(13A.8)

For n = 2 one can rewrite, using the orthogonality relation,


i

X
n

n (q)

!2

n n =

X
nn

= V G0 G0 V(13A.9)
n n G0 G0 n n
n (q)
n (q)

which displays the exchange of two V terms with particles propagating in between.
The same procedure applies at any order in . Thus the propagator has the expansion

= iV iV G0 G0 iV + . . . .

If the potential is instantaneous, the intermediate


placing
G0 G0 i

(13A.10)

dP0 /2 can be performed re-

1
E E0 (P|q)

(13A.11)

where
q
q
E0 (P|q) =
+P +
P
2
2


is the free particle energy which may be considered as the eigenvalue of an operator
H0 . In this case the expansion (13A.10) reads


= i V + V

E H0
1
V + . . . = iV
.
E H0
E H0 V


(13A.12)

950

13 Functional Integral Representation of Quantum Field Theory

We see it related to the resolvent of the complete Hamiltonian as

= iV (RV + 1)

(13A.13)

X n n
1
=
E H0 V
n E En

(13A.14)

where
R

with n being the Schrodinger amplitudes in standard normalization. We can now


easily determine the normalization factor N in the connection between n and the
Schrodinger amplitude n . Eq. (13A.2) gives in the instantaneous case
Z

d3 P
1
n (P|q)
n (P|q) = nn
3
(2)
E H0

(13A.15)

Inserting from (4.48) renders


1
N2

d3 P
(E H0 )n (P|q) = nn .
(2) n

(13A.16)

But since
(E H0 ) =

(13A.17)

this is also
1 Z d3 P
(P|q) V n (P|q) = nn .
N 2 (2)3 n

(13A.18)

For n wave functions in standard normalization the integral expresses the differential

dE
.
d

For a typical calculation of a resolvent, the reader is referred to Schwingers treatment [95] of the Coulomb problem. His result may directly be used for a propagator
of electron hole pairs bound to excitons.

Appendix 13B

Fluctuations around the Composite Field

Here we show that the quantum mechanical fluctuations around the classical equations of motion (13.234)
(x) =

dyV (x, y) (y)(y),

(13B.1)

or (4.5)
(x, y) = V (x y)(x)(y),

(13B.2)

Appendix 13B

951

Fluctuations around the Composite Field

are quite simple to calculate. For this let us compare the Green functions of (x) or
(x, y) with those of the composite operators on the right-hand side of Eqs. (13B.1)
or
(13B.2). The Green
functions or or are generated by adding external currents
R
R
dx(x)I(x) or 1/2 dxdy((y, x)I (x, y) + h.c.) to the final actions (13.239) or
(4.13), respectively, and by forming functional derivatives /I. The Green functions
of the composite operators, on the other hand, are obtained by adding
Z

dx

Z

dyV (x, y) (y)(y) K(x)

1
dxdyV (x y)(x)(y)K (x, y) + h.c.
2
to the original actions (13.232) or (4.4), respectively, and by forming functional
derivatives /K. It is obvious that the sources K can be included in the final
actions (13.239) and (4.13) by simply replacing
(x) (x) = (x)
or

dx K(x )V (x , x)

(x, y) (x, y) = (x, y) K(x, y).


If one now shifts the functional integrations to these new translated variables and
drops the irrelevant superscript prime, the actions can be rewritten as
A[] = iTr log(iG1
)+
+
or

1
2

dxdx (x)V 1 (x, x )(x ) + i

dx(x) [I(x) + K(x)] +

1
2

dxdx (x)G (x, x )(x)

dxdx K(x)V (x, x )K(x )

(13B.3)



i
1
1
A[] = Tr log iG1
dxdx |(x, x )|2
+

2
2
V (x, x )
Z
1
i
dxdx j (x)G (x, x )
+
2
V (x, x )
Z
n
h
i
o
1
dxdx (y, x) I (x, y) + K (x, y) + h.c.
+
2Z
1
+
dxdx |K(x, x )|2 V (x, x ).
(13B.4)
2
In this form the actions display clearly the fact that derivatives with respect to
the sources K or I coincide exactly, except for all possible insertions of the direct
interaction V . For example, the propagators of the plasmon field (x) and of the
R
composite operator dyV (x, y) (y)(y) are related by
Z

(x)
(x
) =

(2) Z
(2) Z
1

=
V
(x,
x
)

I(x)I(x )
K(x)K(x )
Z

(13B.5)

= V 1 (x, x ) + h0|( dyV (x, y) (y)())( dy V (x y ) (y ) (y )(y ))|0i

952

13 Functional Integral Representation of Quantum Field Theory

in agreement with (13.235). Similarly, one finds for the pair fields:

(x,
x )(y,
y ) = (x y)(x y )iV (x x )
+h0|(V (x , x)(x )(x))(V (y , y) (y) (y ))|0i.

(13B.6)

Note that the latter relation is manifestly displayed in the representation (13A.10)
of the propagator . Since

= iV G(4) V

one has from (13B.6)


h0|V ()( V )|0i = V G(4) V

(13B.7)

which is correct, remembering that G(4) is the full four-point Green function. In the
equal-time situation at instantaneous potential, G(4) is replaced by the resolvent R.

Appendix 13C

Two-Loop Heisenberg-Euler
Effective Action

The next correction to the Heisenberg-Euler Lagrangian density (13.405) is [102, 109]
(2) Leff =

ie2
128 4

e4 2 2
i(M 2 i)( + )
e
sin e sin e sinh e sinh e
n

4M 2 [S( )S( ) + P ( )P ( )] I0 i I , (13C.1)


where
S( ) cos e cosh e
1
b
I0
log
ba
a

a e (cot e + cot e )
2e2 2 cosh e( )
p
sin e sin(e )

P ( ) sin e sinh e.
(q p)
b
aq bp
I
log
,
2
(b a)
a ba(b a)
b e (coth e + coth e ) ,
2e2 2 cos e( )
q
.
(13C.2)
sinh e sinh e

,
,
,
,

This expression contains divergences which require renormalization. First, there


is a subtraction of an infinity to make (2) Leff vanish for zero fields. Then there
are charge and wavefunction renormalization, just as for the one-loop effective Lagrangian, which involves identifying a divergent term in (2) Leff of the form of the
zero-loop Maxwell Lagrangian. This is done simply by expanding the integrand to
quadratic order in the fields and . This divergence can be absorbed by redefining
the electric charge and the fields as
1/2

eR = e Z3

1/2

B R = B Z3

1/2

E R = E Z3

(13C.3)

953

Notes and References

where Z3 is some divergent normalization constant, which was given by (13.361)


in previous result (13.407). The invariants and are renormalized accordingly
1/2
1/2
(R = Z3 , R = Z3 ).
Then we re-express (2) Leff in terms of the renormalized charges and fields.
Finally, we have to renormalize the mass:
m2R = m20 + M 2 ,
(1)

(1)

2
Leff
R (mR )

(1)
LR (m20 )

LR (m20 )
+ M
.
m20
2

(13C.4)

The second term in (13C.4) is of order 2 , since M 2 and (1) Leff


R are both of order
. See the original papers [109] for details of removing the divergencies. The final
answer for the renormalized two loop effective Lagrangian is
(2) Leff
R

K0 ( )
ie2
d K(, )
d
=
4
64 0

0


ie2 Z
5
2

d K0 ( ) log(iM ) +
64 4 0
6

"

(13C.5)

where 0.577... is Eulers constant, and the functions K(, ) and K0 ( ) are
i
(e)2 (e)2 h
2

4M
(S(
)
S(
)
+
P
(
)
P
(
))I

iI
0
P ( ) P ( )
"
!#)

2
2
2
1
2i
5i

e
(

)

4M 2
2M 2 ( 2 2 2 2 )
+
( + )
+
3
+
!
#
"

1
e
e2 ( 2 2 ) 2
e
2
iM 2
4M + i
K0 ( ) = e
.
1+
2 tan e tanh e
3
(13C.6)

iM 2 ( + )

K(, ) = e

The fields in this expression can be replaced by the renormalized fields, and everything is finite. The lowest contribution is of the fourth power in the fields and
reads
(2)

eff

263
16 2
e6
( 2 )2 +
( )2 + . . . ,
=
4
4
64 m 81
162


(13C.7)

which has been added to the one-loop result in Eq. (13.415).


In the limit of strong magnetic field it yields
(2)

eff

e4 2
e
=
log
4
128
M 2

Notes and References

"

+ constant + . . . .

(13C.8)

954

13 Functional Integral Representation of Quantum Field Theory

[1] R.P. Feynman, Rev. Mod. Phys. 20, 367 (1948); R.P. Feynman and A.R. Hibbs, Path Integrals and Quantum Mechanics, McGraw-Hill, New York (1968).
[2] H. Kleinert, Path Integrals in Quantum Mechanics, Statistics and Polymer
Physics,
World Scientific Publishing Co., Singapore 1995, pp. 1890.
[3] J. Rzewuski, Quantum Field Theory II, Hefner, New York (1968).
[4] S. Coleman, Erice Lectures 1974, in Laws of Hadronic Matter, ed. by A.
Zichichi, p. 172.
[5] See for example: A.A. Abrikosov, L.P. Gorkov, I.E. Dzyaloskinski, Methods of
Quantum Field Theory in Statistical Physics, Dover, New York (1975); L.P.
Kadanoff, G. Baym, Quantum Statistical Mechanics, Benjamin, New York
(1962); A.L. Fetter, J.D. Walecka, Quantum Theory of Many-Paricle Systems,
McGraw-Hill, New York (1971).
[6] R.L. Stratonovich, Sov. Phys. Dokl. 2, 416 (1958), J. Hubbard, Phys. Rev.
Letters 3, 77 (1959); B. M
uhlschlegel, J. Math. Phys. 3, 522 (1962); J. Langer,
Phys. Rev. 134, A 553 (1964); T.M. Rice, Phys. Rev. 140 A 1889 (1965); J.
Math. Phys. 8, 1581 (1967); A.V. Svidzinskij, Teor. Mat. Fiz. 9, 273 (1971);
D. Sherrington, J. Phys. C4 401 (1971).
[7] The first authors to employ such identities were P.T. Mathews, A. Salam,
Nuovo Cimento 12, 563 (1954), 2, 120 (1955).
[8] H.E. Stanley, Phase Transitions and Critical Phenomena, Clarendon Press,
Osford, 1971; F.J. Wegner, Phase Transitions and Critical Phenomena, ed.
by C. Domb and M.S. Green, Academic Press 1976, p. 7; E. Brezin, J.C. Le
Guillou and J. Zinn-Justin, ibid. p. 125, see also L. Kadanoff, Rev. Mod. Phys.
49, 267 (1977).
[9] For the introduction of collective bilocal fields in particle physics and applications see
H. Kleinert, On the Hadronization of Quark Theories, Erice Lectures 1976
on Particle Physics, publ. in Understanding the Fundamental Constituents of
Matter , Plenum Press 1078, (ed. by A. Zichichi). See also H. Kleinert, Phys.
Letters B62, 429 (1976), B59, 163 (1975).
[10] An excellent review on this equation is given by N. Nakanishi, Progr. Theor.
Phys. Suppl. 43, 1 (1969).
[11] See Ref. [?]

Notes and References

955

[12] See the last of Ref. [?] or D. Saint-James, G. Sarma, E.J. Thomas, Type II
Superconductivity, Pergamon Press, New York (1969).
[13] J.G. Valatin, D. Butler, Nuovo Cimento X, 37 (1958); see also: R.W. Richardson, J. Math. Phys. 9, 1327 (1908).
[14] V.A. Adrianov, V.N. Popov, Theor. Math. Fiz. 28, 340 (1976).
[15] A.L. Leggett, Rev. Mod. Phys. 47, 331 (1975) and Lectures presented at
1st erice Summer School on Low Temperature Physics, 1977. P.W. Anderson, W.F. Brinkman, Lectures presented at XV Scottish Universities Summer
School, The Helium Liquids, Acad. Press, New York (1975), p. 315; D.M. Lee,
R.C. Richardson, The Physics of Liquid and Solid Helium (in preparation).
[16] J.C. Wheatley, Rev. Mod. Phys. 47, 415 (1975) and Lectures presented at the
XV Scottish Universities Summer School, The Helium Liqudis, Acad. Press,
New York (1975), p. 241.
[17] K. Maki, H. Ebisawa, Progr. Theor. 50, 1452 (1973); 51; 337 (1974).
[18] V. Ambegaokar, P.G. De Gennes, D. Rainer, Phys. Rev. A9, 2676 (1974); A
12, 245.
[19] V. Ambegaokar, Lectures presented at the 1974 Canadian Summer School.
[20] D.R.T. Jones, A. Love, M.A. Moore, J. Phys. C9, 743 (1976).
[21] N.D. Mermin, G. Stare, Phys. Rev. Letters 30, 1135 (1973); G. Stare, Ph. D.
Thesis, Cornell (1974); See also a related problem in N.D. Mermin, Phys. Rev.
B13, 112 (1976); G. Barton, M.A. Moore, J. Phys. C7, 2989, 4220 (1974).
[22] K. Maki, Phys. Rev. B11, 4264 (1975) and Paper presented at Sanibel Symposium USC preprint 1977; K. Maki, P. Kumar, Phys. Rev. Letters 38, 557
(1977); Phys. Rev. B16, 174, 182 (1977), B14, 118 (1976); P.G. De Gennes,
Phys. Letters 44A, 271 (1973); K. Maki, Phys. Rev. Letters 39, 46 (1977).
[23] N.D. Mermin, Remarks prepared for the Sanibel Symposium on Quantum
Fields and Solids, Cornell preprint 1977 and Erice Lecture Notes, June 1977;
N.D. Mermin, Phyisca 90B, 1 (1977).
[24] G. t Hooft, Nuclear Phys. B79, 276 (1974) and Phys. Rev. Letters 37, 81
(1976); A. Belavin, A. Polyakov, A. Schwartz, Y. Tyupkin, Phys. Letters B59,
85 (1975).
[25] A. Engelsberg, W.F. Brinkman, P.W. Anderson, Phys. Rev. A9, 2592 (1974).
[26] W.F. Brinkman, H. Smith, D.D. Osheroff, E.I. Blount, Phys. Rev. Letters 33,
624 (1974).

956

13 Functional Integral Representation of Quantum Field Theory

[27] J.M. Delrieu, J. de Physique Letters 35 L-189 (1974), Eratum ibid 36, L22(1975).
[28] C.M. Gould, D.M. Lee, Phys. Rev. Letters 37, 1223 (1976).
[29] D.N. Poulson, N. Krusius, J.C. Wheatley, Phys. Rev. Letters 36, 1322 (1976);
M. C. Temperature Physics, Ontaniemi, Finland 1975, edited by M. Kusius,
M. Vuorio (North Holland, Amsterdam (1975), Vol. 1, p. 29.
[30] W.F. Brinkman, J.W. Serene, P.W. Anderson, Phys. Rev. A 10, 2386 (1974);
M.T. Bealmonod, D.R. Fredkin, S. K. Ma, Phys. Rev. 174, 227 (1968); Y.
Kuroda, A.D.S. Nagi, J. Low. Temp. 23, 751 (1976).
[31] D. Rainer, J.W. Serene, Phys. Rev. B13, 4745 (1976); Y. Kuroda, A.D.S.
Nagi, J. Low Temp. Phys. 25, 569 (1976).
[32] For a review see: D.R. Bes, R.A. Broglia, Lectures delivered at E. Fermi
Varenna Summer School, Varenna, Como. Italy, 1976. For recent studies: D.R.
Bes, R.A. Broglia, R. Liotta, B.R. Mottelson, Phys. Letters 52B, 253 (1974);
56B, 109 (1975), Nuclear Phys. B A260, 127 (1976). See also: R.W. Richardson, J. Math. Phys. 9, 1329 (1968), Ann. Phys. (N.Y.) 65, 249 (1971) and
N.Y.U. Preprint 1977 as well as references therein.
[33] H. Kleinert, Phys. Letters B69, 9 (1977).
[34] B.S. De Witt, Rev. Mod. Phys. 29, 377 (1957); K.S. Cheng, J. Math. Phys.
13 1723 (1972).
[35] N.D. Mermin, T.L. Ho, Phys. Rev. Letters 36, 594 (1976); L.J. Buchholtz,
A.L. Fetter, Phys. Rev. B 15, 5225 (1977).
[36] G. Toulouse, M. Kleman, J. Phys. (Paris) 37, L-149 (1976); V. Poenaru, G.
Toulouse (preprint).
[37] P.W. Anderson, G. Toulouse, Phys. Rev. Letters 38, 508 (1977).
[38] G.E. Volovik, N.B. Kopnin, Pisma, Zh. eksper. teor. Fiz. 25, 26 (1977).
[39] G.E. Volovik, V.P. Mineev, Zh. eksper. teor. Fiz. 73, 167 (1977).
[40] For an extensive discussion see: K. Maki, Physica 90B, 84 (1977); Quantum
Fluids and Solids, Plenum Publishing Corporation, p. 65 (1977); Lecture presented at the 6th Hokone Symposium, Sept. 1977 available as USC preprint).
[41] The situation is quite similar to liquid crystals (nematics): See P.G. de Gennes,
The Physics of Liquid Crystals, Clarendon press, Oxford 1974.

Notes and References

957

[42] L.D. Landau, E.M. Lifshitz, Quantum Mechanics, Pergamon Press, New York
(1965), p. 71; I. Kay, H.E. Moses, J. Appl. Phys. 27, 1503 (1956); J. Rubinstein, J. Math. Phys. 11, 258 (1970).
[43] M.J. Ablowitz, D.J. Kaup, A.C. Nevell, H. Segur, Phys. Rev. Letters 30, 1262
(1973).
[44] K. Maki, H. Ebisawa, J. Low Temp. Phys. 23 351 (1976).
[45] K. Maki, P. Kumar, Phys. Rev. 16, 174 (1977).
[46] J. Schwinger, J. Math. Phys. 5, 1606 (1964).
[47] N.N. Bogoliubov, Sov. Phys. JETP 7, 41 (1958).
[48] J. Bardeen, L.N. Cooper and J.R. Schriefer: Phys. Rev. 108, 1175 (1957).
[49] J.G. Bednorz and K.G. Mueller, Z. Phys. B 64, 198 (1986).
[50] See the www addresses:
http://www.users.qwest.net/csconductor/Experiment Guide/History%20of
%20Superconductivity.htm
http://www.eere.energy.gov/superconductivity/about history.html
http://www.chemsoc.org/exemplarchem/entries/igrant/history noflash.html
[51] R.P. Feynman, Rev. Mod. Phys. 20, 367 (1948); R.P. Feynman and A.R. Hibbs, Path Integrals and Quantum Mechanics, McGraw-Hill, New York (1968).
[52] J. Rzewuski, Quantum Field Theory II, Hefner, New York (1968); S. Coleman,
Erice Lectures 1974, in Laws of Hadronic Matter, ed. by A. Zichichi, p. 172.
[53] See for example: A.A. Abrikosov, L.P. Gorkov, I.E. Dzyaloshinski, Methods of
Quantum Field Theory in Statistical Physics, Dover, New York (1975); L.P.
Kadanoff, G. Baym, Quantum Statistical Mechanics, Benjamin, New York
(1962); A.L. Fetter, J.D. Walecka, Quantum Theory of Many-Paricle Systems,
McGraw-Hill, New York (1971).
[54] J. Hubbard, Phys. Rev. Letters 3, 77 (1959); B. M
uhlschlegel,J. Math. Phys. ,
3, 522 (1962); J. Langer, Phys. Rev. 134, A 553 (1964); T.M. Rice,
Phys. Rev. 140 A 1889 (1965); J. Math. Phys. 8, 1581 (1967); A.V. Svidzinskij,
Teor. Mat. Fiz. 9, 273 (1971); D. Sherrington, J. Phys. C4 401 (1971).
[55] F.W. Wiegel, Phys. Reports C16 57 (1975); V.N. Popov, Kontinualnye Integraly v Kvantovoj Teorii Polja i Statisticheskoj Fizike, Atomizdat, Moscow
(1976).
[56] The first authors to employ such identities were
P.T. Mathews and A. Salam, Nuovo Cimento 12, 563 (1954), 2, 120 (1955).

958

13 Functional Integral Representation of Quantum Field Theory

[57] H.E. Stanley, Phase Transitions and Critical Phenomena, Clarendon Press,
Oxford, 1971; F.J. Wegner, Phase Transitions and Critical Phenomena, ed. by
C. Domb and M.S. Green, Academic Press 1976, p. 7; E. Brezin, J.C. Le Guillou and J. Zinn-Justin, ibid. p. 125, see also L. Kadanoff, Rev. Mod. Phys. 49,
267 (1977).
[58] For the introduction and use of such bilocal fields in particle physics see
H. Kleinert, Erice Lectures 1976 on Particle Physics (ed. by A. Zichichi); also
Phys. Letters B 62, 429 (1976), B 59, 163 (1975).
[59] An excellent review on this equation is given by N. Nakanishi,
Progr. Theor. Phys. Suppl. 43, 1 (1969).
[60] See Ref. [58]
[61] See the last of Ref. [53] or D. Saint-James, G. Sarma, E.J. Thomas, Type II
Superconductivity, Pergamon Press, New York (1969).
[62] J.G. Valatin, D. Butler, Nuovo Cimento X, 37 (1958); see also: R.W. Richardson, J. Math. Phys. 9, 1327 (1908).
[63] V.A. Adrianov, V.N. Popov, Theor. Math. Fiz. 28, 340 (1976).
[64] A.L. Leggett, Rev. Mod. Phys. 47, 331 (1975) and Lectures presented at 1st
Erice Summer School on Low Temperature Physics, 1977.
P.W. Anderson, W.F. Brinkman, Lectures presented at XV Scottish Universities Summer School, The Helium Liquids, Acad. Press, New York (1975), p.
315; D.M. Lee, R.C. Richardson, The Physics of Liquid and Solid Helium (in
preparation).
[65] J.C. Wheatley, Rev. Mod. Phys. 47, 415 (1975) and Lectures presented at the
XV Scottish Universities Summer School, The Helium Liqudis, Acad. Press,
New York (1975), p. 241.
[66] K. Maki, H. Ebisawa, Progr. Theor. Phys. 50, 1452 (1973); 51; 337 (1974).
[67] V. Ambegaokar, P.G. De Gennes, D. Rainer, Phys. Rev. A 9, 2676 (1974); A
12, 245.
[68] V. Ambegaokar, Lectures presented at the 1974 Canadian Summer School.
[69] D.R.T. Jones, A. Love, M.A. Moore, J. Phys. C9, 743 (1976); A. Fetter,
Stanford Preprint 1977.
[70] N.D. Mermin, G. Stare, Phys. Rev. Letters 30, 1135 (1973); G. Stare,
Ph. D. Thesis, Cornell (1974); See also a related problem in N.D. Mermin,
Phys. Rev. B 13, 112 (1976); G. Barton, M.A. Moore, J. Phys. C7, 2989,
4220 (1974).

Notes and References

959

[71] K. Maki, Phys. Rev. B 11, 4264 (1975) and Paper presented at Sanibel Symposium USC preprint 1977; K. Maki, P. Kumar, Phys. Rev. Letters 38, 557
(1977); Phys. Rev. B 16, 174, 182 (1977), B 14, 118 (1976); P.G. De Gennes,
Phys. Letters 44A, 271 (1973); K. Maki, Phys. Rev. Letters 39, 46 (1977).
[72] N.D. Mermin, Remarks prepared for the Sanibel Symposium on Quantum
Fields and Solids, Cornell preprint 1977 nell preprint 1977 and Erice Lecture
Notes, June 1977; N.D. Mermin, Phyisca 90B, 1 (1977).
[73] G. t Hooft, Nuclear Phys. B 79, 276 (1974) and Phys. Rev. Letters 37, 81
(1976); A. Belavin, A. Polyakov, A. Schwartz, Y. Tyupkin, Phys. Letters B
59, 85 (1975).
[74] A. Engelsberg, W.F. Brinkman, P.W. Anderson, Phys. Rev. A 9, 2592 (1974).
[75] W.F. Brinkman, H. Smith, D.D. Osheroff, E.I. Blount, Phys. Rev. Letters 33,
624 (1974).
[76] J.M. Delrieu, J. de Physique Letters 35 L-189 (1974), Eratum ibid 36, L22(1975).
[77] C.M. Gould, D.M. Lee, Phys. Rev. Letters 37, 1223 (1976).
[78] D.N. Poulson, N. Krusius, J.C. Wheatley, Phys. Rev. Letters 36, 1322 (1976);
M.C. Temperature Physics, Ontaniemi, Finland 1975, edited by M. Kusius,
M. Vuorio (North Holland, Amsterdam (1975), Vol. 1, p. 29.
[79] W.F. Brinkman, J.W. Serene, P.W. Anderson, Phys. Rev. A 10, 2386 (1974);
M.T. Bealmonod, D.R. Fredkin, S.K. Ma, Phys. Rev. 174, 227 (1968); Y.
Kuroda, A.D.S. Nagi, J. Low. Temp. 23, 751 (1976).
[80] D. Rainer, J.W. Serene, Phys. Rev. B 13, 4745 (1976); Y. Kuroda, A.D.S.
Nagi, J. Low Temp. Phys. 25, 569 (1976).
[81] For a review see: D.R. Bes, R.A. Broglia, Lectures delivered at E. Fermi
Varenna Summer School, Varenna, Como. Italy, 1976. For recent studies: D.R.
Bes, R.A. Broglia, R. Liotta, B.R. Mottelson, Phys. Letters 52B, 253 (1974);
56B, 109 (1975), Nuclear Phys. A 260, 127 (1976). See also: R.W. Richardson,
J. Math. Phys. 9, 1329 (1968), Ann. Phys. (N.Y.) 65, 249 (1971) and N.Y.U.
Preprint 1977 as well as references therein.
[82] H. Kleinert, Phys. Letters B 69, 9 (1977).
[83] B.S. De Witt, Rev. Mod. Phys. 29, 377 (1957); K.S. Cheng J. Math. Phys. 13
1723 (1972).
[84] N.D. Mermin, T.L. Ho, Phys. Rev. Letters 36, 594 (1976); L.J. Buchholtz,
A.L. Fetter, Phys. Rev. B 15, 5225 (1977).

960

13 Functional Integral Representation of Quantum Field Theory

[85] G. Toulouse, M. Kleman, J. Phys. (Paris) 37, L-149 (1976); V. Poenaru, G.


Toulouse (preprint).
[86] P.W. Anderson, G. Toulouse, Phys. Rev. Letters 38, 508 (1977).
[87] G.E. Volovik, N.B. Kopnin, Pisma, Zh. eksper. teor. Fiz. 25, 26 (1977).
[88] G.E. Volovik, V.P. Mineev, Zh. eksper. teor. Fiz. 73, 167 (1977).
[89] For an extensive discussion see: K. Maki, Physica B 90, 84 (1977); Quantum
Fluids and Solids, Plenum Publishing Corporation, p. 65 (1977); Lecture presented at the 6th Hokone Symposium, Sept. 1977 8available as USC preprint).
[90] The situation is quite similar to liquid crystals (nematics): See P.G. de Gennes,
The Physics of Liquid Crystals, Clarendon press, Oxford 1974.
[91] L.D. Landau, E.M. Lifshitz, Quantum Mechanics, Pergamon Press, New York
(1965), p. 71; I. Kay, H.E. Moses, J. Appl. Phys. 27, 1503 (1956); J. Rubinstein, J. Math. Phys. 11, 258 (1970).
[92] M.J. Ablowitz, D.J. Kaup, A.C. Nevell, H. Segur, Phys. Rev. Letters 30, 1262
(1973).
[93] K. Maki, H. Ebisawa, J. Low Temp. Phys. 23 351 (1976).
[94] K. Maki, P. Kumar, Phys. Rev. 16, 174 (1977).
[95] J. Schwinger, J. Math. Phys. 5, 1606 (1964).
[96] C. Itzykson and J.-B. Zuber, Quantum Field Theory, McGraw-Hill (1985). See
Section 12.2.
[97] W. Heisenberg and H. Euler, Z. Phys. 98, 714 (1936). English translation available at http://www.physik.fu-berlin.de/~kleinert/files/heisenbergeuler.pdf.
[98] See also
J. Schwinger, Phys. Rev. 84, 664 (1936); 93, 615; 94, 1362 (1954).
[99] Such an omission was done in Eq. (4.117) of the textbook [96].
[100] U.D. Jentschura, H. Gies, S.R. Valluri, D.R. Lamm, E.J. Weniger, Can. J.
Phys. 80, 267 (2002) (hep-th/0107135).
[101] V. Weisskopf, The electrodynamics of the vacuum based on the quantum theory
of the electron, Kong. Dans. Vid. Selsk.Math-fys. Medd. XIV No. 6 (1936);
English translation in: Early Quantum Electrodynamics: A Source Book, A.I.
Miller, Cambridge University Press, 1994.

Notes and References

961

[102] G.V. Dunne, Heisenberg-Euler Effective Lagrangians : Basics and Extensions,


Phys. Rep. 355, 73 (2002); to appear in Ian Kogan Memorial Collection,
From Fields to Strings: Circumnavigating Theoretical Physics, M. Shifman,
A. Vainshtein, and J. Wheater, eds., World Scientific, Singapore, 2006.
[103] See Section 5.2 in M. Abramowitz and I. Stegun, Handbook ofMathematical
Functions, (Dover, New York, 1972).
[104] A. Erdelyi (ed.), Higher Transcendental Functions, Vol. I, Kreiger, Florida,
1981.
[105] M. Abramowitz and I. Stegun, Handbook ofMathematical Functions, Dover,
New York, 1972.
[106] E. Whittaker and G. Watson, A Course in Modern Analysis, 4th ed., Cambridge, 1950, pp. 268269.
[107] http://mathworld.wolfram.com/HurwitzZetaFunction.html.
[108] E. Whittaker and G. Watson, A Course in Modern Analysis, 4th ed., Cambridge, 1950.
[109] V.I. Ritus, Lagrangian of an intense electromagneticfield and quantum electrodynamics at short distances, Zh. Eksp. Teor.Fiz 69, 1517 (1975) [Sov. Phys.
JETP 42, 774 (1975)]; Connection between strong-field quantum electrodynamics with short-distance quantum electrodynamics, Zh. Eksp. Teor. Fiz 73, 807
(1977) [Sov. Phys. JETP 46, 423 (1977)]. The Lagrangian Function of an
Intense Electromagnetic Field , in Proc. Lebedev Phys. Inst. Vol. 168, Issues
in Intense-field Quantum Electrodynamics, V.I. Ginzburg, ed., (NovaScience
Pub., NY 1987); Effective Lagrange function of intense electromagnetic field
in QED, (hep-th/9812124).
[110] I.S. Gradshteyn and I.M. Ryzhik, Table ofIntegrals, Series and Products, Academic Press, New York,1972), Formula 6.441.1.
[111] V. Adamchik, Symbolic and Numeric computation of the Barnes function,
Conference on Applications of Computer Algebra, Albuquerque, June 2001;
Contributions to the theory of the Barnes function, (math.CA/0308086).
[112] E.W. Barnes, The theory of the G-function, Quart. J. Pure Appl.Math XXXI
(1900) 264.
[113] G. Mourou, T. Tajima, and S.V. Bulanov, Reviews of Modern Physics, 78,
309, (2006); and references therein.

If you lose the path


you get to know it better
Tanzanian Proverb

14
Formal Properties of Perturbation Theory
In Chapter 10 we have observed simple composition rules which help to construct
all Feynman diagrams of a field theory from connectd diagrams which, in turn, can
be decomposed in a simple way into one-particle irreducible ones, These composition rules can be derived by simple functional operations on appropriate generating
functionals [1].

14.1

Connectedness Structure of Feynman Diagrams

In view of the applications to come, we write the action in a rather general form
A[] = A0 [] + Aint []

(14.1)

1Z D D

A [] =
d x d x (x)iG1
0 (x , x)(x)
2

(14.2)

with a free action


0

and an interaction
1
dD x1 dD x2 dD x3 V3 (x1 , x2 , x3 )(x1 )(x2 ) (x3 )
(14.3)
3! Z
1

dD x1 dD x2 dD x3 dD x4 V4 (x1 , x2 , x3 , x4 )(x1 )(x2 )(x3 )(x4 ).


4!

Aint[] =

In the relativistic scalar theory discussed in the previous two chapters the operator
iG1
0 between the -fields had the specific form

(D)
iG1
(x x )( 2 m2 ).
0 (x, x ) =

(14.4)

We shall find it convenient to introduce a short notation and write (14.2) as


1
A0 [] = iG1
0 .
2

(14.5)

Here the fields are considered as vectors with continuous indices x, whereas

G1
0 (x, x ) is a corresponding matrix. Such objects will be referred to as functional vectors and matrices, respectively. This notation will permit an immediate
962

963

14.2 Functional Differential Equations

adaptation of all formulas to be derived in this chapter to many different field theories. With x being considered as a functional index, V3 , and V4 may be viewed as
functional tensors which are symmetric in all indices. We can save many integral
symbols by rewriting (14.3) in analogy with (14.5) short as
1
1
Aint [] = V3 3 V4 4 ,
3!
4!

(14.6)

where the successive contraction of the functional indices x is implied.


For the above general action and in this notation, the Heisenberg field operator
satisfies an equation of motion
iG1
0

Aint []
= 0.
+

(14.7)

or explicitly
iG1
0 (x)

14.2

V3 2 V4 3
= 0.
2!
3!

(14.8)

Functional Differential Equations

The operator equation (14.8) can be used to derive a differential equation for the
generating functional Z[j] of the theory defined in Eq. (7.818). In the functional
vector notation introduced above we, shall write the generating functional as
Z[j] = hT eij i.

(14.9)

In order to apply (14.8), we form the derivative

Z[j] = hT eij i
ij

(14.10)

1
and multiply this by the functional matrix iG1
0 . When passing iG0 to the right
of the time-ordering operator we generate the 2 and 3 terms of (14.8). These
can also be obtained from higher derivatives (/j)2 , (/j)3 applied to Z[j]. The
passage of the time-ordering operator produces some additional terms in the same
way as in the derivation of the equation for the Green functions in (7.42)(7.44). Its
calculation requires the use of the canonical equal-time commutation rules. Since
these are the same for interacting and free fields, we can evaluate them for the free
fields and consider taking iG1
to the right of the time-ordering operator in the
0
free-field expression

iG1
0

Z0 [j].
ij

(14.11)

The free partition function was calculated in Section 7.15. In Eqs. (7.821) and
(7.822)] we found
1

Z 0 [j] = e 2 j G0 j .

(14.12)

964

14 Formal Properties of Perturbation Theory

satisfying the functional differential equation


iG1
0

Z 0 [j] + j(x)Z 0 [j] = 0.


ij(x)

(14.13)

The generalization to the interacting case follows directly by adding the interaction
terms of (14.8) expressed in terms of derivatives with respect to the current:

V3
iG1

0
ij
2!

ij

!2

V4

3!

ij

!3

+ j Z [j] = 0.

(14.14)

For the subsequent development it is useful to introduce the short-hand notation

Z[j] = Zjj...j [j],


ij ij
ij

(14.15)

with the arguments x1 , . . . , xn of the currents suppressed. Then we may write (14.14)
as
1
V3 1
V4 1
iG1
Zj
Zjj
Zjjj + jZ = 0 .
0
2
i
2! i
3! i3

(14.16)

This equation can be used to calculate Z[j] recursively. We have seen, however,
in the previously discussed example of 4 -theory that Z[j] is not a very economic
quantitiy to deal with. Its expansion coefficients, the n-point Green functions, contain many disconnected pieces. A substantial simplification arises by going over to
the functional iW [j] defined by
iW [j] = log Z[j].

(14.17)

It serves as a generating functional for the connected part of the Green functions
which are obtained from
G(n)
c (x1 , . . . , xn ) =

iW [j].
ij(x1 )
ij(xn )

(14.18)

Using the differential equation (14.16) we are able to prove quite generally that
these Green functions do deserve their name since their graphical representation
consists only of connected diagrams. Moreover, the connected Green functions collect precisely all the connected diagrams in G(n) (x1 , . . . , xn ) and the full Green
functions G(n) (x1 , . . . , xn ) can be recovered in a simple way from the connected
ones Gc(n) (x1 , . . . , xn ). Let us first prove this by deriving the general relationship
between the two types of Green functions. This relationship serves to transform the
functional differential equation (14.16) for Z[j] into a more useful one for W [j] which
will enable us to prove the connectedness of the Green functions G(n)
c (x1 , . . . , xn ).

14.3 Decomposition of Green Functions into Connected Green Functions

14.3

965

Decomposition of Green Functions


into Connected Green Functions

Using the exponential relation (14.17) between iW and Z we can now derive general
relations between the n-point functions and the connected ones. For the one-point
function we find

G(1) (x) = Z 1 [j]


Z[j] = W [j] = G(1)
(14.19)
c (x).
ij(x)
j
Thus the one-point function, which represents the ground state expectation value of
the field, is always connected:
h(x)i G(1) (x) = G(1)
c (x).

(14.20)

We may convince ourselves by drawing the Feynman diagrams corresponding to the


formula
[]

h0|T (x)eiAint +j |0i


G (x) =
,
(14.21)
h0|T eiAint [] |0i
that this is indeed true. All disconnected pieces cancel when forming the quotient.
Consider now the two-point function where

Z[j]
G(2) (x1 , x2 ) = Z 1 [j]
ij(x1 ) ij(x2 )
(
!
)

1
= Z [j]
iW [j] Z[j]
ij(x1 )
ij(x2 )


1
= Z 1 [j] Wj(x1 )j(x2 ) + Wj(x1 ) Wj(x2 ) Z[j]
i
(2)
(1)
= Gc (x1 , x2 ) + G(1)
(14.22)
c (x1 ) Gc (x2 ) .
(1)

In addition to the connected diagrams with two ends there are two connected diagrams ending in a single line. These are absent in a pure 4 -theory at j = 0 if
the ground state of the system is normal, a specification which we are going to
understand only later when discussing phase transitions.
For the three-point function we find

G(3) (x1 , x2 , x3 ) = Z 1 [j]


Z[j]
ij(x1 ) ij(x2 ) ij(x3 )
("
#
)

1
= Z [j]
iW [j] Z[j]
ij(x1 ) ij(x2 )
i(x3 )



1

Wj(x2 )k(x3 ) + Wj(x2 ) Wj(x3 ) Z[j]


(14.23)
= Z 1 [j]
i(x1 )
i


1
1
= Z [j] 2 Wj(x1 )j(x2 )j(x3 ) +i Wj(x1 ) Wj(x2 )j(x3 ) +Wj(x2 ) Wj(x1 )j(x3 )
i

o
+ Wj(x3 ) Wj(x1 )j(x2 ) + Wj(x1 ) Wj(x2 ) Wj(x3 ) Z[j]
h

(1)
(2)
(1)
(1)
(1)
= G(3)
c (x1 , x2 , x3 ) + Gc (x1 )Gc (x2 , x3 ) + 2 perm + Gc (x1 )Gc (x2 )Gc (x3 ).

966

14 Formal Properties of Perturbation Theory

and for the four-point function


h

(3)
(1)
G(4) (x1 , . . . , x4 ) = G(4)
c (x1 , . . . , x4 ) + Gc (x1 , x2 , x3 ) Gc (x4 ) + 3 perm

(2)
+ G(2)
c (x1 , x2 ) Gc (x3 , x4 ) + 2 perm

(1)
(1)
+ G(2)
c (x1 , x2 ) Gc (x3 ) Gc (x4 ) + 4 perm
(1)
+G(1)
c (x1 ) Gc (x4 ).

(14.24)

In the pure 4 -theory (and in normal systems) there are no odd Green functions and
we remain with (10.65), which was observed there diagrammatically up to second
order in the perturbation expansion.
For the general Green function G(n) , the total number of terms is most easily
retrieved by dropping all indices and factors i and differentiating with respect to j
(the prime stands for /j)


G(1) = eW eW


G(2) = eW eW


G(3) = eW eW


G(4) = eW eW

= Wj = G(1)
c
(1) 2
= Wjj + Wj2 = G(2)
c + Gc

jjj

(14.25)

jjjj

(2) (1)
(1)3
= Wjjj + 3Wjj Wj + Wj3 = G(3)
c + 3Gc Gc + Gc

= Wjjjj + 4Wjjj Wj + 3Wjj2 + 6Wjj Wj2 + Wj4

(3) (1)
(2)2
(1)2
= G(4)
+ 6G(2)
+ G(1)4
c + 4Gc Gc + 3Gc
c Gc
c .

All relations follow from the recursion relation


(n1)

G(n) = Gj

+ G(n1) G(1)
c ,

n > 2,

(14.26)

(n1)

if one uses Gc j
= Gc(n) and the initial relation G(1) = G(1)
c . By comparing the
first four relations with the explicit forms (14.22)(14.24) we see that the numerical
factors on the right-hand side of (14.25) refer to the permutations of the arguments
x1 , x2 , x3 , . . . of otherwise equal expressions. Since there is no problem in reconstructing the explicit permutations we shall, from now on, write all composition
laws in the short-hand notation (14.25).
The formula (14.25) and its generalization is often referred to as cluster decomposition or also as the cumulant expansion of the Green functions.
Let us now prove the statement that the connected Green functions collect precisely all connected diagrams in the n-point functions. For this we observe that
the decomposition rules can be inverted by simply writing iW [j] = log Z[j] and
differentiating such that one finds
G(1)
c
G(2)
c
G(3)
c
G(4)
c

=
=
=
=

G(1)
G(2) G(1) G(1)
G(3) 3G(2) G(1) + 2G(1)3
G(4) 4G(3) G(1) + 12G(2) G(1)2 3G(2)2 6G(1)4 .

(14.27)

14.4 Functional Differential Equation for W [j [

967

Each equation follows from the previous one by differentiation with respect to j,
and replacing the derivatives on the right-hand side according to the rule
(n)

Gj

= G(n+1) G(n) G(1) .

(14.28)

Again the numerical factors imply different permutations of the arguments, and the
primes denote functional differentiations with respect to j.
Now it is obvious that any connected diagram contained in G(n) must also be
contained in G(n)
c , since all the terms added or subtracted in (14.27) are products of
G(n) s and thus necessarily disconnected.
The proof that G(n)
c collects only connected diagrams, is somewhat more involved
and requires the use of the equations of motion (14.8). Since the information on
Gc is contained most directly in W [j], we shall first translate these equations into a
functional differential equation for W [j].

14.4

Functional Differential Equation for W[j]

Since we already know the consequences of the equation of motion for Z[j] we may
simply insert (14.17) into (14.16) and find directly the following differential equation
for W [j]:
iG1
0 Wj



1 1
1 1
V3 2 iWjj Wj2 V4 3 iWjjj 3Wjj Wj iWj3 + j = 0.
2! i
3! i
(14.29)

Multiplying this by iG0 gives


Wj = iG0



1 1
1 1
V3 2 iWjj Wj2 iG0 V4 3 iWjjj 3Wjj Wj iWj3 + iG0 j.
2! i
3! i
(14.30)

This may be expressed in terms of the connected j-Green function


G(1)
c = Wj (x)

(14.31)

as

V3  (1)
(1)
Gc j + iG(1)
G
c
c
2!

V4  (1)
(1)
(1)3
+G0
+ iG0 j.
iGc jj 3Gc j G(1)
c iGc
3!

G(1)
= G0
c

14.5

(14.32)

Iterative Solution

Equation (14.30) can best be solved graphically as shown in Fig. 14.1. The lowest,

968

14 Formal Properties of Perturbation Theory

Figure 14.1 Graphical solution of recursion relation (14.30) for the generating functional
W [j] of all connected Green functions.

zeroth order in V3 , V4 , we have1


G(1)
c = iG0 j.

(14.33)

From this we find the zeroth order generating functional W [j] by functional integration
W 0 [j] =

(0)

Dj Wj [j] =

i
Dj G(1)
c = jG0 j,
2

(14.34)

a result already known from previous discussions.


Reinserting (14.34) on the right-hand side of (14.32) we find the first-order equation
G(1)
= G0
c
1

i
V3
V4 h
3G0 G0 j (G0 j)3 + iG0 j. (14.35)
(iG0 iG0 j G0 j) + G0
2!
3!
(1)

(2)

The superscripts (1), (2), . . . on Green functions Gc , Gc , . . . refer to the number of points
in the connected Green function, while those on W 1 , W 2 , . . . record the order in the interactions
V3,4 of the perturbative expressions.

969

14.6 Vertex Functions

This can be integrated functionally to obtain W [j] up to first order in the potentials.
Graphically, this process amounts to multiplying the open line in each diagram by
a current j and dividing the resulting powers j n by n. Thus we have (see Fig. 14.1)
i
i
i
jG0 j jG0 V3 G0 + V3 (G0 j)3
2
2!
3!
1
1
2
V4 G0 (G0 j) V4 (G0 j)4 .
(14.36)
+
4
24
This procedure can be continued to any order in V3 , V4 .
We are now in a position to confirm the statement made above that the generating functional W [j] collects only connected diagrams in its Taylor coefficients
n W/j(x1 ) . . . j(xn ). Recall that we convinced ourselves by the argument after
Eq. (14.28) that these were all connected diagrams contained in G(n) (x1 , . . . , xn ).
For the lowest two orders we can verify the connectedness directly by inspecting the
third line in Fig. 14.1. After this, the diagrammatic form of the recursion relation
shows that this topological property remains true to all orders in V3,4 , by induction.
Since, if it was true for some n, then all G(1)
c to be inserted on the right-hand side are
connected and so are the diagrams constructed from these giving G(1)
c to (n + 1)st
order. Note that this calculation is unable to recover the value of W [j] at j = 0
since this is an unknown integration constant of the functional differential equation.
For the purpose of generating Green functions, this constant is irrelevant. We have
seen in Eq. (10.70) that Z[0] consists of the sum of all connected vacuum diagrams
contained in W [0].
W 0 [j] + W 1 [j] =

14.6

Vertex Functions

Having understood the connectedness structure of Feynman diagrams, we now turn


to the second important composition property: that in terms of one-particle irreducible graphs. A crucial role in the perturbation expansions of Chapter 10 was
played by the vertex functions [the one-particle irreducible amputated Green functions introduced on p. 724]. In his section we shall describe their properties by
functional differential equations. In particular, we shall present a simple algorithm
yielding the composition rules of connected diagrams into one-particle irreducible
ones.

14.7

The Generating Functional for Vertex Functions

First we introduce a new generating functional [] which is defined as the Legendre


transform of W [j]:
[] W [j] Wj j,

(14.37)

where Wj j is a short-hand notation for


Wj j =

dD x Wj (x) j(x),

(14.38)

970

14 Formal Properties of Perturbation Theory

and the new variable is the functional derivative of W [j]:


(x)

W [j]
Wj = hij ,
j(x)

(14.39)

thus giving the ground state expectation of the field operator in the presence of the
current j. The physical situation corresponds to j = 0 and we shall denote the
associated field expectation by
0 = hi|j=0,

(14.40)

which may be zero or non-zero and space-time dependent. The most common situation found in quantum field theoretic systems is 0 = 0. In many interesting
physical applications to many-body systems, however, the state 0 at j = 0 is nonzero and in fact, not even isotropic, but carries interesting structures: It may be
perforated by vortices, as in superconductors, exhibit a plane wave-like modulation,
as in magnetic superconductors and pion condensates, or occur in a lattice-like modulation, as in the blue phase of cholesteric liquid crystals. At present we shall allow
for any such possibility.
In this section we want to demonstrate an important property of the functional
derivatives of []
(n)

... (x1 , . . . , xn )

...
[] .
(x1 )
(xn )

(14.41)

They collect precisely all OPI vertex diagrams formed with vertices V3 + V4 and
V , and with propagators G0 A1
|. The expression V4 is short notation for
R4 D
d xV4 (x1 , x2 , x3 , x4 )(x4 ). With the above Thus in the previously discussed special case of a 4 -theory with 0 = hi = 0, the vertex functions (14.41) for = 0
coincide with those defined graphically in Eq. (10.6.1), for example.
In order to prove this, we take again recourse to a functional differential equation
for []. This can be derived quite simply from that for W [j]. First we observe that
G(1)
c is simply the field expectation . Second we see that
iG(2)
c

(1)
Gj

= Wjj =
=
j

!1

= 1
.

(14.42)

Here the inverse is to be understood in the functional sense, i.e.,


1

(x, y)

(x)(y)
"

#1

(14.43)

is the functional matrix which satisfies


Z

1
dy
(x, y) (y, z) = (x z).

(14.44)

Equation (14.42) says that the second derivative of the generating functional []
determines directly the propagator of the interacting theory in the presence of the

14.7 The Generating Functional for Vertex Functions

971

external source j. Since j is an auxiliary quantity which eventually has to be turned


off such that becomes 0 , the actual physical propagator is given by
1
G = G(2)
c = i ,

(14.45)

evaluated at = 0 . The third relation to be used follows from (14.42) by applying


the chain rule and deriving once more functionally with respect to j.
Wjjj = 2

3
= 3
= iG .
j

(14.46)

This equation has a simple physical meaning. The third derivative of W [j] on the
left-hand side is the full three-point function (up to the sign), so that
G(3) = Wjjj = iG3 .

(14.47)

This equation states that the full three-point function arises from a third derivative of
[] by attaching to each derivation a full propagator. This is exactly the structure
observed before in (10.75) for the four-point function.
We shall express Eq. (14.52) graphically as follows:

where

denotes the connected n-point function and

the n-point vertex function.


Thus seems to be a three-point vertex function in the same sense as defined
graphically in Subsection 10.6.1. For this it has to contain all OPI diagrams. This
will emerge from the discussion to follow.
To go on to higher orders, we observe that (14.46) states that the functional
derivative of the full propagator G with respect to the current j is
1
Gj = Wjjj = G(3) = iG3 .
i
This is pictured graphically as follows:

(14.48)

972

14 Formal Properties of Perturbation Theory

This equation may be differentiated further with respect to the current j by


graphical methods. From the definition (14.18) we deduce the trivial recursion
relation
G(n)
c (x1 , . . . , xn ) =

G(n1) (x2 , . . . , xn ) ,
ij(x1 ) c

(14.49)

which is represented graphically as

Thus, by applying /ij repeatedly to Eq. (14.48), we generate all higher connected
Green functions. On the right-hand side of (14.48), the differentiation of each line
is done using once more (14.48), which changes a line into a three-point function.
It remains to find the derivative of the vertex function. This is calculated via the
chain rule as

= ... iG,
(14.50)
...j = ...
j
which is represented graphically as

With these graphical rules, we can differentiate (14.46) any number of times, and
derive the graphical structure of the connected Green functions with an arbitrary
number of external legs. The result up to n = 5 is shown in Fig. 14.2.
The diagrams generated in this way have a tree-like structure which is why they
are called tree diagrams. The tree decomposition reduces all diagrams to their oneparticle irreducible contents.
The one-particle reducible structure of all graphs can also be illustrated in another way. Let us separate the effective action into a free and a nontrivial part
[] 0 [] + int [] =
=

1
m2 2
()2
+ int []
2
2

1
int
iG1
0 + [].
2

(14.51)

973

14.7 The Generating Functional for Vertex Functions

G(n)
c

1 ...n

Figure 14.2 Tree decompositon of connected Green functions into one-particle irreducible
perts.

The free part coincides with the free action A0 [] in Eq. (14.2), except that the
fluctuating field is replaced by its expectation value . The same decomposition
is done with the generating functional W [j] [recall (14.34)]:
i
W [j] = W 0 [j] + W int [j] = j G0 j + W int [j].
2

(14.52)

Then we see that


i
int [] = W int [j] G1
0
2
i
+ jG0 j j.
2

(14.53)

But using the equation for the current


1
j = = int
iG0

(14.54)

i
i
i int
1
G0 int
= jG0 j G0 j
2
2
2

(14.55)

we find

974

14 Formal Properties of Perturbation Theory

which is obviously a one-particle reducible contribution. Inserting this into (14.53),


we may write
i
W int [j] = int [] int
G0 int
.
2

(14.56)

This equation shows explicitly how the sum of all connected diagrams in W int [j]
is composed of all particle irreducible connected diagrams in int [] plus all oneparticle reducible diagrams.
The effective action [] can also be used to prove the fact that the full propagator can be expressed as a geometric series (10.74) of repetitions of the self-energy,
a fact that was observed graphically to low orders in Subsection 10.6.1. We use once
more the decomposition (14.51):
[] = 0 [] + int [].

(14.57)

Differentiating this twice with respect to , we obtain


int
= iG1
0 + ,
= int
,

(14.58)

and (14.45) can be rewritten as




G = G0 1 iG0 int

1

(14.59)

Expanding the denominator and recollecting terms this can also be written as an
integral equation
G = G0 + iG0 int
G.

(14.60)

In this equation we recognize the self-energy


= int
.

(14.61)

This was introduced graphically in Eq. (10.74), where it was defined by the relation
1
G i[iG1
0 ] ,

1
iG1
0 iG .

(14.62)

Comparison with (14.59) shows that (14.61) is the correct identification.


Thus the self-energy is given by the interacting part of the second derivative of the
effective action. This proves the statement that, in general, all Feynman diagrams in
the propagator can be obtained by a geometric series type of repetition of self-energy
diagrams. What we do not know as yet, but what will become clear immediately in
the next section is that the self-energy collects precisely all one-particle irreducible
amputated two-point functions.

14.8 Functional Differential Equation for []

14.8

975

Functional Differential Equation for []

We shall now prove the fact that [] contains only one-particle irreducible diagrams
by means of a functional differential equation for []. For a cubic and quartic
interaction, this is immediately obtained from the differential equation (14.29) for
W [j]. We simply substitute
Wj ,

j ,

Wjj iG,

Wjjj G(3) = iG3 ,

(14.63)

make use of the separation (14.57), and find for the interacting part the functional
differential equation
1
1
1
1
3
int
V4 4 (V3 + V4 ) G V4 iG3 int
= V3
.
2!
3!
2!
3!

(14.64)

This equation can be solved solved iteratively together with (14.59). Before starting
it is useful to observe that there exists a trivial relation between [] and int
which
n
merely consists in a factor 1/n in front of every power . Thus the diagrams in []
differ from those in [] only by numerical prefactors. The resulting equations are
displayed graphically in Fig. 14.3. First we neglect int on the right-hand side of the
equation and find the first-order correction 1 [] to the interacting part int []:
1
1
1
1
1 [] = V3 3 V4 4 V3 G0 V4 G0 2 .
3!
4!
2
4

(14.65)

The first two terms are simply the interacting parts of the original action Aint []
(14.3) with replaced by . The remaining two terms are new. They may be
obtained by partial contractions in the interacting part of the original interaction
(14.3),and setting = in the remaining uncontracted fields:
1
1
V3 3 V3 3 ,
3!
3!
1 43
1
.
V 4 4 V 4
4!
4!
2!

(14.66)

The iteration has been continued graphically up to the second order in the interactions in Fig. 14.3.
The lowest term 1 [] contains obviously only OPI diagrams. The graphical
iteration shows that this property persists to all orders. If all diagrams of nth order
are OPI, those composed to next order on the right-hand side of Eq. (14.64) always
contain at least two more lines and thus do not decompose by cutting a single line.
They are therefore are OPI (in normal field theories in which the expectation 0
vanishes at zero external source j). This proves the statement made on that behalf in
the beginning. It is thereby confirmed that the selfenergy collects only one-particle
irreducible vertex diagrams. Thus the full propagator is indeed obtained from the
geometric iteration of such diagrams implied by (14.59):
G = G0 + G0 (i)G0 + G0 (i)G0 (i)G0 + . . . .

(14.67)

976

14 Formal Properties of Perturbation Theory

Figure 14.3 Graphical solution of functional differential equation (14.64) for the interacting part int [] of the effective action.

As argued before, all OPI two-point vertex diagrams must be contained in =


iint
since if it occurs in G it must also occur in .
We have also proved the another important statement observed graphically in the
low-order diagrams of Chapter 10. All n-point functions can be obtained by composing all OPI vertex diagrams and full Green functions will be via tree diagrams.
The composition rules were derived above.
For completeness, let us treat the functional differential equation for [] once
more in a different way in a non-normal system (corresponding to a magnetized

14.8 Functional Differential Equation for []

977

phase in solid-state physics) with a nonvanishing field expectation 0 = hi =


6 0 for
e j = 0 , keeping explicit reference to 0 . Setting
= 0 +

(14.68)

[ ] [0 + ]

(14.69)

we define

which then satisfies the equation



1 
1
V3 (0 + 20 ) V4 30 + 320 + 30
2!
3!


1
1
1
2
V4 i 1

V
i
+

0
3

2
2!


1

3
2 2
= 0.

+
2

V4 3 + 3i1

4!
Let us now split

iG1
0 0

[ ] = 0 [ ] + int [ ],

(14.70)

(14.71)

with
1
0 [ ] = A[0 ] + A [0 ] + A [0 ] .
2

(14.72)

Inserting this into (14.70), we find for the functional differential equation

Here we insert



1
= (V3 + V4 0 ) i 1 + 2
2!


1

3
V4 3 + 3i1
.
+
3!

(14.73)

(14.74)

= int
.

(14.75)

= A [0 ] + int

and

Then Eq. (14.73) becomes structurally identical with (14.64), only that V3 , V4 are
replaced by (V3 + V4 0 ) and V4 , respectively, and the free Green function G0 by


1
0
G
0 = i iG0 V3 0

V4 2
.
2 0


(14.76)

Thus int [], when expanded in powers of , does indeed collect only one-particle
irreducible diagrams. This is what we wanted to prove in the general case 0 6= 0
[see (14.41)]:
(n)
...


n

=

.
(x) (xn ) =0

(14.77)

978

14.9

14 Formal Properties of Perturbation Theory

Effective Action as a Basis for Variational


Calculation

In the last section we have seen that the effective action contains all the information
necessary to construct all n point functions. Moreover, its calculation requires only
a certain restricted class of Feynman diagrams, namely those which are OPI. There
is, however, another important feature which makes the effective action a powerful
tool in the study of physical properties of a system and which, finally, justifies its
name: A physical configuration of the field expectation = hi|j extremizes []
at j = 0. This is a direct consequence of the definition (14.37). If we write [] in
the form
[] = W [j]

dD x (x)j(x),

(14.78)

we find

d x

W
j

j(x) = j(x),
j

(14.79)

implying that in the absence of the auxiliary external source j, [] is indeed extremal. The functional [] has thus exactly the same property in quantum theory
as the usual action A[] in classical theory.

14.10

Effective Potential

A special role in the extraction of observational consequences from a theory will


be played by the effective potential. Remember that in classical mechanics, the
potential is defined by the action evaluated for a time-independent orbit q(t) as
A=

tb

ta



dt L(q, q)

q=const

= T V (q),

T tb ta .

(14.80)

By analogy, we define the effective potential of a field theory, if it has the special
property 0 = const, by

[]

=const

= T L3 V (),

L3 = spatial volume.

(14.81)

Comparing this with (14.41) we see that the effective potential has the expansion
T L3 V () =

1 Z D
d x1 dD xn (n) (x1 , . . . , xn ) ( 0 ) . . . ( 0 ) .(14.82)
n!

Introducing the Fourier transforms


(n) (q1 , . . . , qn )

dD x1 dD xn ei(q1 x1 +...+qn xn ) (n) (x1 , . . . , xn ),

(14.83)

979

14.11 Higher Effective Actions

and removing the overall -function of energy-momentum conservation contained in


(n) (q1 , . . . , qn ) due to the translational invariance of (n) (x1 , . . . , xn ),
(n) (q1 , . . . , qn ),
(n) (q1 , . . . , qn ) (2)D (D) (q1 + . . . + qn )

(14.84)

we can write (14.86) as


V () =


1 (n)
( 0 ) . . . ( 0 ).
(q1 , . . . , qn )
n!
qi =0

(14.85)

The discussion has been given for a single Bose field. For a set of Bose fields a ,
the index a may be imagined as being included in the spacetime index x. Thus
Eq. (14.86) reads more explicitly:
V (a ) =

14.11


1 (n)

(q1 , . . . , qn )
( 0 )a1 . . . ( 0 )an .
qi =0
n!

(14.86)

Higher Effective Actions

Effective actions are an important theoretical tool to find extremal principles in


quantum field theory and statistical mechanics [?], and their history goes back to the
extremal principles of Lee and Yang [?]. They were generalized to bilocal variables
by de Dominicis, P.C. Martin, and others [2]. The functional [G] is specified in
terms of a graphical expansion according to the number of fermion loops. We shall
sketch this formalism in real time, since then the lowest contributions lead directly to
the famous time-dependent Hartree-Fock Bogoljubov equations. Higher corrections
are found from a graphical iteration procedure. Let us briefly recall that procedure
for a fermion system.
These provide us with an effective action functional [G], whose extrema yield
the bilocal density matrix G of the interacting quantum system.
Starting point is an action of the form (14.1) with an arbitrary two-body force.
It will be written in terms of the quasireal doubled field
!

=
as

with

c = ,

c = 1 =

0 1
1 0

(14.87)

1
1
A[] = G1
v .
0
2
4!

(14.88)

with G1
0 being now
G1
0

2 /2 + V1 (x) T
+ p
0
2
/2 V1 (x) +
0
p

=c

(14.89)

More explicitly, the matrix elements of G0 contain the free correlation functions

G(x, t; x , t )

hTt (x, t) (x , t )i0 hTt (x, t) (x , t )i0


hTt (x, t) (x , t )i0 hTt (x, t) (x , t )i0 ,

(14.90)

980

14 Formal Properties of Perturbation Theory

where , are spin indices, all being suppressed in (14.88) and the subsequent
formulas.
To the action (14.88) we add a bilocal source term and form the generating
functional

 
Z
1
(14.91)
Z[K] D exp i A[] + K ,
2
where
!

,
Kc
T
and K stands short for dtdt a (t)Kab (t, t )b (t ). The superscript T deT
notes functional transposition, i.e., if spin indices are displayed: Kab
(t, t )
Kba (t , t), T (t, t) (t , t). Note that the potential v with four doubled indices is antisymmetric.
Formally, Z[K] can be calculated by removing the interacting part from the
integral via functional derivatives, writing
R

i
Z[K] = exp v
Z0 [K],
6 iK iK

(14.92)

where Z0 [K] is given by a Gaussian functional integral, which can be performed to


give




Z
1
i
1
(14.93)

=
exp
Tr
log
iG
Z0 [K] = D exp iG1
K
K ,
2
2
1
with G1
K = G0 + K. Obviously, GK is the Green function of the fermions in the
presence of an external field K. Expanding Z in powers of v results in the standard
diagrammatic rules of perturbation theory. If we form the logarithm generating
functional
iW [K] = log Z[K],
(14.94)
this contains all connected vacuum graphs formed with four-vertices and propagators
GK . Because of the fermion nature of these are one-particle irreducible and look
as follows (we drop there all diagrams with three-vertices)
W [K] =
"

1 3
+
3! 16

1
8

1
+
8

1 1
2! 8


+
1
+
4

1
24

+
1
+
8

1
6
3
+
4

(14.95)
3
+
4

Given W [K], it is easy to find the exact density matrix of the system
W [K]

1
G=
=
W [K].
2
K
K

(14.96)

An effective action [G] of the system is now introduced as the Legendre transform

1
[G] W [K] Tr(KG)
2
K=K[G]

(14.97)

981

14.11 Higher Effective Actions

such that (14.96) amounts to


[G]
1
= K.
(14.98)
G
2
At the end, the auxiliary source K is set equal to 0. Equation (14.98) implies that
that the physical density matrix extremizes [G], thus providing us with an extremal
principle.
Let us evaluate the lowest contributions explicitly: At the one-loop level we have
i
W (1) [K] = Tr log iG1
K
2

(14.99)

and (14.96)yields G = GK . Reinserting this into (14.97), we obtain


i
i
1
[G] (0) [G] + (1) [G] = Tr(G1
0 G) Tr log iG .
2
2

(14.100)

The extremum of this is the density matrix of the free fermion system G = G0 .
We now add to W [K] the lowest two-loop correction in (14.96).
1
W (2) [K] = v GK GK .
8

(14.101)

Exhibiting upper and lower spin component, as well as other particle indices, this
contains again Hartree, Fock, and Bogoljubov contributions, whose explicit spin-up
and spin-down configurations are

1
g [(GK ) (GK ) (GK ) (GK ) (GK ) (GK ) ].
24

(14.102)

By differentiating W2 [K] = W (0) [K]+W (1) [K]+W (2) [K] with respect to K according
to (14.96), we obtain
i
G = GK GK (vGK )GK ,
2

(14.103)

1
K = iG1 iG1
0 + vG,
2

(14.104)

which is solved for

to be reinserted into W2 [K], yielding via (14.97):


1
(2) [G] = gGG.
8

(14.105)

Extremizing 2 [K] = (0) [K] + (1) [K] + (2) [K] yields


G=i

iG1
0

1
vG
2

1

(14.106)

982

14 Formal Properties of Perturbation Theory

This is precisely to the well-known time-dependent Hartree-Fock-Bogoljubov equations.


Let us indicate how to proceed in this fashion to higher-loop corrections. Most
efficiently, we may use the obvious functional identity
Z

exp iA[] + K = 0,

2


(14.107)

and work out the differentiations and derive the functional differential equation
i
i
2
(iG1
= 0,
0 + K)W,K + v(W,KK + iW,K )
3
2

(14.108)

where subscritp separated by a komma indicates functional differentiation with re1


spect to that variable. Using the relations (14.96) and iW,KK = G,K /2 = K,G
/2 =
1
,GG /4:
1
v(i1
(14.109)
(iG1
,GG ) i = 0.
0 2,G )G
3!
By separating out the trivial parts the interacting part satisfies the coupled equations
i
1
vG4 ,
int
,G [G]G = vGG +
4
12

(14.110)

int 1
int
= 4int
= 4int
,GG (1 2iGG,GG )
,GG + 2i,GG GG .

(14.111)

The quantity represents e the exact four-particle vertex function of the theory.
The equation is pictured in Fig. 14.4. This equation generates precisely those vacG

Figure 14.4 Recursion relation for two-paticle-irreducible graphs in the effective action
int [G].

uum graphs which do not fall into pieces by cutting two lines, i.e. the two-particle
irreducible diagrams. Up to two loops, we have
[G] = (0) [G] + (1) [G] + int [G],

(14.112)

with
int [G] =

1
8

1
24

1
48

+ ... .

(14.113)

983

14.11 Higher Effective Actions

In general, effective action has the expansion

X
i
1
1
1
[, G] = Tr log(iG0 ) Tr(G) vGG +
(n) [G].
2
2
8
n3

(14.114)

with (n) collecting all two-paticle-irreducible graphs with n loops. The extremality
condition G [G] = 0 produces a Green function
1
G = i{G1
0 [G]} ,

(14.115)

with a self-energy

X
1
i
1
1
(n)
3
,G [G].
[G] = int
[G]
=
vG

vG

=
vG

G
2
4
12
4
n3

(14.116)

Equation (14.115) was first used by Dyson [?] with the prescription that []
contains all those self-energy graphs which are one-particle irreducible, i.e. which
do not fall apart when cutting one line. Obviously, this agrees precisely with the
present procedure, since removing one line form all two-particle irreducible diagrams
produces all one-particle irreducible diagrams of the self-energy.
We remark that large amplitude collective excitations are studied by solving
the above equations semiclassically. For any periodic state one chooses some initial
self-energy , and solves the eigenvalue equation
{iG1
0 }k (t) = k k (t),

(14.117)

where k (t) are antiperiodic wave functions and k the corresponding Bloch-Floquet
indices. Then one determines
i
exp [in (t t )] k (t)k (t )

n
k
k n
X
1
=
exp[ik (t t )]k (t)k (t ).
exp(i
T
)
+
1
k
k

G(t, t ) =

XX

(14.118)

where the Fermi distribution contains the period of the oscillations T like an imaginary temperature, as a result of the sum over the frequencies n = 2(n + 1/2)/T .
This Green function (14.118) is inserted into (14.117) for a next iteration. The static
case follows by taking the limit T where the sum reduces to the states below
the Fermi surface, and k (t) become time-independent wave functions.
The resulting G can be inserted into the effective action [G] for a quantization
of the large-amplitude oscillating orbits (or for a determination of the energy in the
case of static G configurations, E = [G]/T ). For this purpose one only has to
form the semiclassical quantum propagator as the Fourier transform [3]
Z

dT
W (E)
exp(i[G] + ET )
exp[iW (E)],
2
E

(14.119)

984

14 Formal Properties of Perturbation Theory

where T is the period. Running through the same orbit many times and inserting a
phase 1 for each turning point gives a Green function

W/E

exp[inW (E)] = W/E exp[iW (E)]

n=1

1
,
1 + exp[iW (E)]

(14.120)

which has pole at W (En ) = 2(n + 1/2) [3]).


If an imaginary-time solutions passes through a potential barrier, exp(T E) =
exp{i[G]} gives the amplitude of barrier penetration [4].

14.12

High Orders in Simple Model

It is instructive to apply the treatment to an exactly solvable model where one


can calculate the higher effective action to arbitrary order. Consider the classical
partition function of the anharmonic oscillator in zero dimensions, where it is given
by the simple integral
Z=

dx

2/ 2

e (

2 x2 /2+gx4 /4!

It is easy to expand this in powers of g as Z =


coefficients are
z (k) =

k=0

).

(14.121)

z (k) g k where g g/ 4. The

(1)k (2k + 1/2)


6k k!
(1/2)

(14.122)

35 385 25025 1616615 260275015 929553625 835668708875


starting out at like 81 , 384
, 3072 , 98304 , 2359296 , 113246208 , 100663296 , 19327352832 , . . . . For large k,

k
they grow like (1) (2/3)k k!/k 2. The series is divergent, but Borel-summable.

It possesses an absolutely convergent strong-coupling expansion in powers of 1/ g .


This is found by rewriting (14.121) as

1 Z dy y/g y2
Z = g 1/4
e
e
0
y

(14.123)

and expanding the first exponential. The result is the absolutely convergent strongcoupling series:
!k

X
1
1/4
(14.124)
Z = (g )
k
g
k=0
with coefficients

1
(k/2 + 1/4)
.
k = (1)k
2
k!

(14.125)

The series is may be expressed in terms of well-known mathematical functions as


1/8g

Z=e

1
W0,1/4 (1/4g ) =
2

1
2g

!1/2

e1/8g K1/4 (1/8g )

(14.126)

985

14.12 High Orders in Simple Model

where W0,1/4 (z) is Whittakers function and K1/4 (z) the modified Bessel function.
For g the function approaches the limit
Z
0 (g /6)1/4

(14.127)

1 Z dy y2
(1/4)
0 =
= 1.0228 .
e
0
y
2

(14.128)

with

The free energy of the model is defined by Z eF . It has the power series
expansion

g
fn
F =
4
n=1

!n

(14.129)

with the expansion coefficients fn beig equal to


11
619
709 858437
54193 18639247
, 17
fn = ( 18 , 121 , 96
72 , 960 , 324 , 96768 , 1296 , 82944 , . . . ).

(14.130)

The function Z is plotted in Fig. 14.5.

Figure 14.5 The anharmonic model integral Z as a function of g = g/ 4 in comparison with the three variational approximations, Z1 (long-dashed) which coming from the
Hartree-Fock-Bogoljubov approximation to the energy, from the the new extended approximation Z3 , and from the much worse old extension Z3old derived from the higher effective
action. The new extension Z3 is not distinguishable from Z on this plot, lying less than
0.1% below Z at g /4! = 2.5. Observe that the old extension Z3old is worse than Z1 .

How well is this classical partition function reproduced by the Hartree-Fock


approximation and its successive improvements?
The function [G, [G]] of Eq. (14.114) becomes for = 1:
2 2 2 2
1
1
5 4 16 101 5 20
1
a +3ga4 g 2a8 + g 3 a12
g a +
g a ...
[G, [G]] = log 2 +
2

2
48
48
128
960
(14.131)

986

14 Formal Properties of Perturbation Theory

with a2 = 1/, where the higher terms can be calculated to all orders. They have
the form (1)n1 cn g n a4n with cn satisfying the recursion relation
n1
X
1
1
(n m)m(2m 1)cm cnm
cn1 (2n 2)(2n 3) + 8
cn =
2n
6
m=1

"

(14.132)

5
101
93
8143
starting out with c1 = 1/8, c2 = 1/48, and continuing with 128
, 960
, 256
, 5376
,
271217
36864 , . . . .
An important feature of the Hartree-Fock-Bogoljubov approximation is that it
explains quite well the strong-coupling behaviour g of the system. In this
limit, we see from the first-order terms in
(14.131) that the extremal
2 diverges

1/2
1/4
1/4
for large g like d(g/6)
and Z1 e / d(g/6)
with d = 3. Indeed, the
1/4
associated approxmation for Z is e / d 0.9756 compares reasonably with the
correct value 0 in (14.128).
However, when trying to calculate this behaviour in the presence of the higher
correction terms, we find a severe weakness of the present improvement scheme:
The extremlality equation to be solved for d becomes, with correction terms to
Hartree-Fock-Bogoljubov approximation

d +

3
810 16362
6
3+ 7 +
. . . = 0.
d
d
d
d9


(14.133)

The factorial growth of the coefficients makes it impossible to extract a solution


without resummation procedures. This is manifest in the plot in Fig. (14.6), where
the third-order approximation is worse than the first-order one.

Figure 14.6 Approximations to F obtained from the extrema of the higher effective
action [G, [G]] in Eq. (14.131) up to order g N , denoted by N . The errors increase
rapidly with increasing g and there is no uniform convergence.

The only scheme to converge at higher order is the one based on collective classical fields to be presented in Section ??.

Notes and References

987

Notes and References


[1] In this chapter we follow the review paper
H. Kleinert, Higher Effective Actions in Bose Systems,
Fortschr. Phys. 30, 187 (1982) (kl/82/82.pdf), where kl is short for
klnrt.de.
See also
C. Itzykson and J.-B. Zuber, Quantum Field Theory, McGraw-Hill (1985);
H. Kleinert and V. Schulte-Frohlinde, Critical Phenomena in 4 -Theory, World
Scientific, Singapore 2001, pp. 1487 (kl/b8).
[2] C. de Dominicis, J. Math. Phys. 3, 983 (1962);
C. de Dominicis, P.C. Martin, ibid 5, 16, 31, (1964);
H. Dahmen, G. Jona-Lasinio, Nuovo Cimento 52A, 807 (1967);
A.N. Vasilev, A.K. Kazanskii, Teor. Math. Phys. 12, 875 (1972);
J.M. Cornwall, R. Jackiw, E. Tomboulis, Phys. Rev. 10, 2428 (1974);
H. Kleinert, Fortschr. Phys. 30, 187 (1982).
[3] H. Kleinert and H. Reinhardt, Nucl. Phys. A 42, 331 (1981).
See also Chapter 4 in
H. Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets, World Scientific, Singapore 2009
(klnrt.de/b5).
[4] H. Kurasuji and T. Suzuki, J. Math. Phys. 21, 472 (1980);
Phys. Lett. B 92, 19 (1980).

It is best to do things systematically,


since we are only human, and disorder is our worst enemy
Hesiod (800 BC)

16
Systematic Graphical Construction of Feynman
Diagrams and Their Multiplicities
In Section 14.4 we have calculated the generating functional for the connected Green
functions W [j] by solving the functional differential equation (14.29) with respect
to j. In addition, W [j] may also be considered as a functional of the free propagator
G0 . As such it also obeys functional differential equation which can be turned into
a recursion relation. This is solved order by order in the coupling constant to find
all connected vacuum diagrams with their proper multiplicities. Let us apply this
procedure here to W [0] which contains all connected vaccum diagrams, first for a
pure 4 -theory, subsequently to a theory with two scalar fields and A and an
interaction 2 A.
After obtaining the connected vacuum diagrams, all Feynman diagrams with
external lines are obtained from functional derivatives of the connected vacuum
diagrams with respect to the free correlation function.
Finally, the recursive graphical construction will be automatized by computer
algebra with the help of a unique matrix notation for the Feynman diagrams. The
methods developed in this chaoter have been set up a long time ago,1 but for a long
time they have only been solved only for low orders in the coupling strength, and
to all orders only for a rather trivial zero-dimensional quantum field theories, the
anharmonic integral. Here we present an extension of this work2 consisting in an
efficient graphical algorithm for solving the functional differential equation for W [0]
as a functional of G0 . We shall do this first for a multicomponent scalar 4 -theory,
then also for a theory with two scalar fields and A with an interaction 2 A.
In a second step, the connected vacuum diagrams are used to derive all connected
diagrams with external lines with the help of functional derivatives of the connected
vacuum diagrams with respect to the free correlation function. This operation will
represented graphically by cutting lines. Finally, we automatize our construction
1

H. Kleinert, Fortschr. Phys. 30, 187 (1982); 30, 351 (1982).


H. Kleinert, A. Pelster, B. Kastening, and M. Bachmann, Phys. Rev. E 62, 1537 (2000)
(hep-th/9907168). The associated algebraic program and its output can be downloaded from
http://www.physik.fu-berlin.de/kleinert/294/programs .
2

1008

16.1 Generalized Scalar 4 -Theory

1009

method by computer algebra with the help of a unique matrix notation for Feynman
diagrams.

16.1

Generalized Scalar 4 -Theory

The folowing treatment will involve somewhat lengthy equations such that it is
useful to introduce a a shoter notation for tha symbols used. For the free proagator
G0 (x1 , x2 ) we shall employ a functional matrix notation G12 . The fully interacting
two-point
function G(x1 , x2 ) will Rbe denoted by G12 . The D-dimensional integral
R D
d x1 will simply be written as 1 . We shall further work in euclidean spacetime
to avoid factors of i. The scalar field is allowd to have N components, and the
thermal flcutuations are controlled by a euclidean action which plays the role of an
energy functional:
Z
Z
1
g
E[] =
G1

+
V1234 1 2 3 4
(16.1)
1 2
2 12 12
4! 1234
with some coupling constant g. The functional matrix G1
12 is the inverse of the free
propagator G12 . Both contain, in addition , a Kronecker for the N field components, which we suppress. Throughout the manipulations to follow, the functional
matrix G12 will be considered as a completely general symmetric nonsingular functional matrix, such that G1
12 exists. Only at the end shall we specify G12 , identifying
it with the free propagator, i.e., for the inverse


2
2
1
(x1 x2 ) ,
G1
12 G1 ,2 (x1 , x2 ) = 1 ,2 x1 + m

(16.2)

In applications to critical phenomena to be discussed in Chapter 21, the mass m2


will be proportional to the temperature distance T Tc from the critical point.
The interaction will likewise be considered as a completely general symmetric
function of four arguments, gV (x1 , x2 , x3 , x4 ), where g parametrizeds the coupling
strength and serves as an expansion parameter. In the short-hand notation, it is a
symmetric functional tensor of rank four, and will eventuzally be identified with the
true interaction
V1234 V1 ,2 ,3 ,4 (x1 , x2 , x3 , x4 )
(16.3)
1
= {1 ,2 3 ,4 +1 ,3 2 ,4 +1 ,4 2 ,3 } (x1 x2 )(x1 x3 )(x1 x4 ) .
3
By using natural units in which the Boltzmann constant kB times the temperature
T equals unity, the partition function is determined as a functional integral over the
Boltzmann weight eE[]
Z=

D eE[]

(16.4)

and may be evaluated perturbatively as a power series in the coupling constant g.


From this we obtain the negative free energy W = ln Z as an expansion
1 g
W =
4!
p=0 p!

p

W (p) .

(16.5)

1010

16 Systematic Graphical Construction of Feynman Diagrams . . .

The coefficients W (p) may be displayed as connected vacuum diagrams constructed


from lines and vertices. Each line represents a free correlation function
2

G12 ,

(16.6)

whose functional inverse G1 appears in the energy functional (16.1), defined by


Z

G12 G1
23 = 13 .

(16.7)

The vertices represent an integral over the interaction

1234

V1234 .

(16.8)

To construct all connected vacuum diagrams contributing to W (p) to each order p in


perturbation theory, one connects p vertices with 4p legs in all possible ways according to Feynmans rules which follow from Wicks expansion of correlation functions
into a sum of all pair contractions. This yields an increasing number of Feynman
diagrams, each with a certain multiplicity which follows from combinatorics. In
total there are 4!p p! ways of ordering the 4p legs of the p vertices. This number is
reduced by permutations of the legs and the vertices which leave a vacuum diagram
invariant. Denoting the number of self-, double, triple and fourfold connections
with S, D, T, F , there are 2!S , 2!D , 3!T , 4!F leg permutations. An additional reduction arises from the number N of identical vertex permutations where the vertices
remain attached to the lines emerging from them in the same way as before. The
resulting multiplicity of a connected vacuum diagram in the 4 -theory is therefore
given by the formula3
=
ME=0
4

4!p p!
.
2!S+D 3!T 4!F N

(16.9)

The superscript E records that the number of external legs of the connected vacuum
diagrams is zero. The diagrammatic representation of the coefficients W (p) in the
expansion (16.5) of the negative free energy W is displayed in Table up to five
loops4
For higher orders, the factorially increasing number of diagrams makes it more
and more difficult to construct all topologically different diagrams and to count
their multiplicities. In particular, it becomes quite hard to identify by inspection
the number N of identical vertex permutations. This identification problem is solved
by introducing a uniqued matrix notation for the graphs, to be explained in detail
3

H. Kleinert and V. Schulte-Frohlinde, Critical Properties of 4 -Theories, World Scientific,


Singapore, 2000.
4
B. Kastening, Phys. Rev. D 54, 3965 (1996); Phys. Rev. D 57, 3567 (1998); S.A. Larin, M.
M
onnigmann, M. Str
osser, V. Dohm, Phys. Rev. B 58, 3394 (1998).

1011

16.2 Basic Graphical Operations

in Section IV.
In the following, we shall generate iteratively all connected vacuum diagrams.
We start in Subsection II.A by identifying graphical operations associated with
functional derivatives with respect to the inverse propagator G1 , or the propagator
G. In Subsection II.B we show that these operations can be applied to the one-loop
contribution of the free partition function to generate all perturbative contributions
to the partition function (16.4). In Subsection II.C we derive a nonlinear functional
differential equation for the negative free energy W , whose graphical solution in
Subsection II.D. yields all connected vacuum diagrams order by order in the coupling
strength.

16.2

Basic Graphical Operations

Each Feynman diagram is composed of integrals over products of free correlation


functions G and may thus be considered as a functional of the inverse propagator G1 . The connected vacuum diagrams satisfy a certain functional differential
equation, from which they will be constructed recursively. This will be done by
a graphical procedure, for which we set up the necessary graphical rules in this
subsection. First we observe that functional derivatives with respect to the inverse
propagator G1 or to the free propagator G correspond to the graphical prescriptions
of cutting or of removing a single line of a diagram in all possible ways, respectively.

16.2.1

Cutting Lines

Since is a real scalar field, the inverse propagator G1 is a symmetric functional


matrix. This property has to be taken into account when performing functional
derivatives with respect to the inverse propagator G1 , whose basic rule is
G1
1
12
{13 42 + 14 32 } .
1 =
2
G34

(16.10)

From the identity (16.7) and the functional chain rule, we find the effect of this
derivative on the free propagator
2

G12
= G13 G42 + G14 G32 .
G1
34

(16.11)

This has the graphical representation


2

G1
34

. (16.12)

Thus differentiating a propagator with respect to the inverse propagator G1


amounts to cutting the associated line into two pieces. The differentiation rule
(16.10) ensures that the spatial indices of the inverse propagator are symmetrically

1012

16 Systematic Graphical Construction of Feynman Diagrams . . .

attached to the newly created line ends in the two possible ways. When differentiating a general Feynman integral with respect to G1 , the product rule of functional
differentiation leads to a sum of diagrams in which each line is cut once.
With this graphical operation, the product of two fields can be rewritten as a
derivative of the energy functional with respect to the inverse propagator
1 2 = 2

E[]
,
G1
12

(16.13)

as follows directly from (16.1) and (16.10). Applying the substitution rule (16.13)
to the functional integral for the fully interacting two-point function
G12

1
=
Z

D 12 eE[] ,

(16.14)

we obtain the fundamental identity


G12 = 2

W
.
G1
12

(16.15)

Thus, by cutting a line of the connected vacuum diagrams in all possible ways, we
obtain all diagrams of the fully interacting two-point function. Analytically this has
a Taylor series expansion in powers of the coupling constant g similar to (16.5)
G12

1 g
=
4!
p=0 p!

p

(p)

G12

(16.16)

with coefficients
(p)

G12 = 2

W (p)
.
G1
12

(16.17)

The cutting prescription (16.15) converts the vacuum diagrams of pth order in the
(p)
coefficients W (p) in Table to the corresponding ones in the coefficients G12 of the
two-point function. The results are shown in Table up to four loops. The numbering
of diagrams used in Table reveals from which connected vacuum diagrams they are
obtained by cutting a line.
For instance, the diagrams #15.1-#15.5 and their multiplicities in Table follow
from the connected vacuum diagram #15 in Table . We observe that the multiplicity
of a diagram of a two-point function obeys a formula similar to (16.9):
ME=2
=
4

4!p p! 2!
.
2!S+D 3!T N

(16.18)

In the numerator, the 4!p p! permutations of the 4p legs of the p vertices are multiplied by a factor 2! for the permutations of the E = 2 end points of the two-point
function. The number N in the denominator counts the identical permutations of

1013

16.2 Basic Graphical Operations

both the p vertices and the two end points.


Performing a differentiation of the two-point function (16.14) with respect to the
inverse propagator G1 yields
2

G12
= G1234 G12 G34 ,
G1
34

(16.19)

where G1234 denotes the fully interacting four-point function


G1234 =

1 Z
D 1 2 3 4 eE[] .
Z

(16.20)

The term G12 G34 in (16.19) subtracts a certain set of disconnected diagrams from
G1234 . By subtracting all disconnected diagrams from G1234 , we obtain the connected four-point function
Gc1234 G1234 G12 G34 G13 G24 G14 G23

(16.21)

in the form
Gc1234 = 2

G12
G13 G24 G14 G23 .
G1
34

(16.22)

The first term contains all diagrams obtained by cutting a line in the diagrams of
the two-point-function G12 . The second and third terms remove from these the
disconnected diagrams. In this way we obtain the perturbative expansion
Gc1234

1 g
=
4!
p=1 p!

p

c,(p)

(16.23)

G1234

with coefficients
(p)

c,(p)
G1234

p
X
G12
= 2

G1
34
q=0

p
q

(pq)

G13

(q)

(pq)

G24 + G14

(q)

G23 .

(16.24)

They are listed diagrammatically in Table up to three loops. As before in Table II,
the multiple numbering in Table indicates the origin of each diagram of the connected four-point function. For instance, the diagram #11.2.2, #11.4.3, #14.1.2,
#14.3.3 in Table stems together with its multiplicity from the diagrams #11.2,
#11.4, #14.1, #14.3 in Table .
The multiplicity of each diagram of a connected four-point function obeys a
formula similar to (16.18):
=
ME=4
4

4!p p! 4!
.
2!S+D 3!T N

(16.25)

1014

16 Systematic Graphical Construction of Feynman Diagrams . . .

This multiplicity decomposes into equal parts if the spatial indices 1, 2, 3, 4 are assigned to the E = 4 end points of the connected four-point function, for instance:
20736

62208

+ 20736

+ 20736

(16.26)
Generalizing the multiplicities (16.9), (16.18), and (16.25) for connected vacuum diagrams, two- and four-point functions to an arbitrary connected correlation function
with an even number E of end points, we see that
4!p p! E!
,
2!S+D 3!T 4!F N

ME4 =

(16.27)

where N counts the number of permutations of vertices and external lines which
leave the diagram unchanged.

16.2.2

Removing Lines

We now study the graphical effect of functional derivatives with respect to the free
propagator G, where the basic differentiation rule (16.10) becomes
G12
1
= {13 42 + 14 32 } .
G34
2

(16.28)

We represent this graphically by extending the elements of Feynman diagrams by


an open dot with two labeled line ends representing the delta function:
2

12 .

(16.29)

Thus we can write the differentiation (16.28) graphically as follows:

1
=
2

(16.30)

Differentiating a line with respect to the free correlation function removes the line,
leaving in a symmetrized way the spatial indices of the free correlation function on
the vertices to which the line was connected.
The effect of this derivative is illustrated by studying the diagrammatic effect of
the operator
=
L

12

G12

.
G12

(16.31)

to a connected vacuum diagram in W (p) , the functional derivative /G12


Applying L
removes successively each of its 2p lines. Subsequently, the removed lines are again

1015

16.3 Perturbation Theory

reinserted, so that the connected vacuum diagrams W (p) are eigenfunctions of L,


whose eigenvalues 2p count the lines of the diagrams:
W (p) = 2p W (p) .
L

(16.32)

As an example, take the explicit first-order expression for the vacuum diagrams, i.e.
W

(1)

=3

1234

V1234 G12 G34 ,

(16.33)

and apply the basic rule (16.28), leading to the desired eigenvalue 2.

16.3

Perturbation Theory

For the calculation of the vacuum energy W [0] we use functional derivatives with
respect to the inverse propagator G1 in the energy functional (16.1) rather than
with respect to the current j. This allows us to substitute the previous expression
(13.52) for the partition function by
(

g
Z = exp
6

1234

V1234

2
W (0)
,
1
1 e
G12 G34
)

(16.34)

where the zeroth order of the negative free energy has the diagrammatic representation
1
1
W (0) = Tr ln G1
2
2

(16.35)

Expanding again the exponential in a power series, we obtain


Z =

2
g
V1234
(16.36)
1+
1
6 1234
G1
12 G34
)


1 g 2 Z
4
(0)
+
V1234 V5678 1 1 1 1 + . . . eW .
2 6
G12 G34 G56 G78
12345678
Z

Thus we need only half as many functional derivatives as in the expansion derived
with the help of the currents j. Taking into account (16.10), (16.11), and (13.22),
we obtain
1
W (0)
G12 ,
1 =
2
G12

1
2 W (0)
{G13 G24 + G14 G23 } ,
1
1 =
4
G12 G34

(16.37)

such that the partition function Z becomes


Z =



g Z
1 g 2 Z
3
V1234 V5678
1+
V1234 G12 G34 +
4!
2 4!
12345678
1234
"

(16.38)
#

9 G12 G34 G56 G78 + 24 G15 G26 G37 G48 + 72 G12 G35 G46 G78 + . . .

eW

(0)

1016

16 Systematic Graphical Construction of Feynman Diagrams . . .

This has the diagrammatic representation

Z =

1+

1
2

g
3
4!

g
4!

2

(16.39)

+ 24

+ 72

+ ...

eW

(0)

All diagrams in this expansion follow directly by successively cutting lines of the
basic one-loop vacuum diagram (16.35) according to (16.36). By going to the logarithm of the partition function Z, we find a diagrammatic expansion for the negative
free energy W

W =

1
2

g
3
4!

1
2

g
4!

2

24

+ 72

+ ... ,
(16.40)

which turns out to contain precisely all connected diagrams in (16.39) with the same
multiplicities. In the next section we show that this diagrammatic expansion for the
negative free energy can be derived more efficiently by solving a differential equation.

16.4

Functional Differential Equation for W = ln Z

Regarding the partition function Z as a functional of the inverse propagator G1 ,


we derive a functional differential equation for Z. We start with the trivial identity
Z

o
n
2 eE[] = 0 ,
1

(16.41)

which follows via functional integration by parts from the vanishing of the exponential at infinite fields. Taking into account the explicit form of the energy functional
(16.1), we perform the functional derivative with respect to the field and obtain
Z

D 12

G1
13 2 3

g
6

345

V1345 2 3 4 5 eE[] = 0 .

(16.42)

Applying the substitution rule (16.13), this equation can be expressed in terms of the
partition function (16.4) and its derivatives with respect to the inverse propagator
G1 :
12 Z + 2

G1
13

2
Z
g
1 =
3
G23

345

V1345

2Z
1 .
G1
23 G45

(16.43)

16.4 Functional Differential Equation for W = ln Z

1017

Note that this linear functional differential equation for the partition function Z is,
indeed, solved by (16.34) due to the commutation relation
(

g
exp
6

g
2
2
1
1
G

G
exp

V1234
V
1234
56
56
1
1
6 1234
G1
G1
1234
12 G34
12 G34
(
)
Z
Z

2
g
g
V5678
V
=
exp

(16.44)
1234
1
3 78
6 1234
G1
G1
78
12 G34
)

which follows from the canonical one


1

1
1
{13 24 + 14 23 } .
1 G34 G34
1 =
2
G12
G12

(16.45)

Going over from Z to W = ln Z, the linear functional differential equation (16.43)


turns into a nonlinear one:
12 + 2

G1
13

W
2
g
1 =
3
G23

345

V1345

2W
W W
1
1
1 +
G23 G45
G1
23 G45

(16.46)

If the coupling constant g vanishes, this is immediately solved by (16.35). For a nonvanishing coupling constant g, the right-hand side in (16.46) produces corrections
to (16.35) which we shall denote with W (int) . Thus the negative free energy W
decomposes according to
W = W (0) + W (int) .

(16.47)

Inserting this into (16.46) and taking into account (16.37), we obtain the following
functional differential equation for the interaction negative free energy W (int) :
Z

12

G1
12

g
g
W (int)
W (int)
=
V
G
G

V
G
1234 12 34
1234 12
4 1234
3 1234
G1
G1
12
34
(
)
Z
2
(int)
(int)
(int)
g
W
W
W
+
V1234
.
1 +
3 1234
G1
G1
G1
12 G34
12
34
Z

(16.48)

With the help of the functional chain rule, the first and second derivatives with
respect to the inverse propagator G1 are rewritten as

=
G1
12

34

G13 G24

G34

(16.49)

and
2
1 =
G1
12 G34

5678

1
+
2

56

G15 G26 G37 G48

2
G56 G78

{G13 G25 G46 + G14 G25 G36 + G23 G15 G46 + G24 G15 G36 }

(16.50)

,
G56

1018

16 Systematic Graphical Construction of Feynman Diagrams . . .

respectively, so that the functional differential equation (16.48) for W (int) takes the
form5
Z

12

G12

16.5

W (int)
g
W (int)
=
(16.51)
V1234 G12 G35 G46
V1234 G12 G34 g
G12
4 1234
G56
123456
)
(
Z
W (int) W (int)
g
2 W (int)
+
.

V1234 G15 G26 G37 G48


3 12345678
G56 G78
G56 G78
Z

Recursion Relation and Graphical Solution

We now convert the functional differential equation (16.51) into a recursion relation
by expanding W (int) into a power series in G:
W (int) =

1 g
4!
p=1 p!

p

W (p) .

(16.52)

Using the property (16.32) that the coefficient W (p) satisfies the eigenvalue problem
of the line numbering operator (16.31), we obtain the recursion relation
W

(p+1)

= 12
+ 4

123456
p1
X
q=1

V1234 G12 G35 G46


p
q

!Z

12345678

W (p)
+4
G56

12345678

V1234 G15 G26 G37 G48

V1234 G15 G26 G37 G48

W (pq) W (q)
,
G56 G78

2 W (p)
G56 G78
(16.53)

and the initial condition (16.33). With the help of the graphical rules of Subsection
16.2, the recursion relation (16.53) can be written diagrammatically as follows

W (p+1) =

2 W (p)
1 2 3

p1
X
q=1

p
q

1
2
3
4

W (p)
1 2

+ 12

W (pq)
1 2

W (q)
,
3 4

1
2

p 1. (16.54)

This is iterated starting from


W (1) = 3 3

(16.55)

The right-hand side of (16.54) contains three different graphical operations. The
first two are linear and involve one or two line amputations of the previous perturbative order. The third operation is nonlinear and mixes two different one-line
5

Compare Eq. (51) in the second of the papers quoted in Footnote 1.

1019

16.5 Recursion Relation and Graphical Solution

amputations of lower orders.


An alternative way of formulating the above recursion relation may be based on
the graphical rules

W (p) =

W (p) 1
=
2
G12

2
2 W (p)
=
G12 G34 3

(16.56)

With these, the recursion relation (16.54) reads

p+1 =4

+ 12

+ 4

p1
X

p
q

q=1

pq

To demonstrate the working of (16.54), we calculate the connected vacuum diagrams


up to five loops. Applying the linear operations to (16.53), we obtain immediately
W (1)
= 6
1 2

2 W (1)
1 2 3

= 6

(16.58)

Inserted into (16.54), these lead to the three-loop vacuum diagrams


W (2) = 24

+ 72

(16.59)

Proceeding to the next order, we have to perform one- and two-line amputations on
the vacuum graphs in (16.59), leading to
W (2)
= 96
1 2

+ 144

+ 144

(16.60)

and subsequently to
2 W (2)
1 2 3

288

+ 144

144

+ 288

+ 144

4
3

4
3

2
2

1
2

+ 144

. (16.61)

1020

16 Systematic Graphical Construction of Feynman Diagrams . . .

Inserting (16.60) and (16.61) into (16.54) and taking into account (16.58), we find
the connected vacuum diagrams of order p = 3 with their multiplicities as shown in
Table . We observe that the nonlinear operation in (16.54) does not lead to topologically new diagrams. It only corrects the multiplicities of the diagrams generated
from the first two operations. This is true also in higher orders. The connected
vacuum diagrams of the subsequent order p = 4 and their multiplicities are listed in
Table .
As a cross-check we can also determine the total multiplicities M (p) of all connected vacuum diagrams contributing to W (p) . To this end we recall that each of
the M (p) diagrams in W (p) consists of 2p lines. The amputation of one or two lines
therefore leads to 2pM (p) and 2p(2p 1)M (p) diagrams with 2p 1 and 2p 2
lines, respectively. Considering only the total multiplicities, the graphical recursion
relations (16.54) reduce to the form derived before:6
M (p+1) = 16p(p + 1)M (p) + 16

p1
X

p!
M (q) M (pq) ;
(p

1)!(q

1)!
q=1

p (16.62)
1.

These are solved starting with the initial value


M (1) = 3 ,

(16.63)

leading to the total multiplicities


M (2) = 96 ,

M (3) = 9504 ,

M (4) = 1880064 ,

(16.64)

which agree with the results listed in Table . In addition we note that the next
orders would contain
M (5) = 616108032 , M (6) = 301093355520 , M (7) = 205062331760640

(16.65)

connected vacuum diagrams.

16.6

Scalar 2 A-Theory

For the sake of generality, let us also study the situation where the quartic interaction
of the 4 -theory is generated by a scalar field A from a cubic 2 A-interaction. The
associated energy functional
E[, A] = E (0) [, A] + E (int) [, A]

(16.66)

decomposes into the free part


E (0) [, A] =
6

1
2

12

G1
12 1 2 +

See the second of the papers quoted in Footnote 1.

1
2

12

1
H12
A1 A2

(16.67)

16.6 Scalar 2 A-Theory

1021

and the interaction


E

(int)

Z
g
[, A] =
V123 1 2 A3 .
2 123

(16.68)

Indeed, as the field A appears only quadratically in (16.66), the functional integral
for the partition function
Z=

D DA eE[,A]

(16.69)

can be exactly evaluated with respect to the field A, yielding


Z=

D eE

(eff) []

(16.70)

with the effective energy functional


1
1
E (eff ) [] = Tr ln H 1 +
2
2

12

G1
12 1 2

g
8

123456

V125 V346 H56 1 2 3 4 . (16.71)

Apart from a trivial shift due to the negative free energy of the field A, the effective
energy functional (16.71) coincides with one (16.1) of a 4 -theory with the quartic
interaction
V1234 = 3

56

V125 V346 H56 .

(16.72)

If we supplement the previous Feynman rules (16.6), (16.8) by the free correlation
function of the field A
1

H12

(16.73)

and the cubic interaction


Z

123

V123 ,

(16.74)

the intimate relation (16.72) between the 4 -theory and the 2 A- theory can be
graphically illustrated by
=

(16.75)

This corresponds to a photon exchange in the so-called s-, t- and u-channels of


Mandelstams theory of the scattering matrix. Their infinite repetitions yield the
relevant forces in the Hartree, Fock and Bogoliubov approximations of many-body
physics. In the following we analyze the 2 A-theory along similar lines as before the
4 -theory.

1022

16.7

16 Systematic Graphical Construction of Feynman Diagrams . . .

Perturbation Theory

Expanding the exponential in the partition function (16.69) in powers of the coupling
constant g, the resulting perturbation series reads
 p Z

1
g
Z=
p=0 (2p)! 4

D DA

Z

123456

V123 V456 1 2 4 5 A3 A6

p

eE

(0) [,A]

. (16.76)

Substituting the product of two fields or A by a functional derivative with respect


to the inverse propagators G1 or H 1, we conclude from (16.76)
(2g)p
Z=
p=0 (2p)!

123456

V123 V456

3
1
1
G1
12 G45 H36

!p

eW

(0)

(16.77)

where the zeroth order of the negative free energy reads


1
1
1
W (0) = Tr ln G1 Tr ln H 1
2
2
2

1
2

(16.78)

Inserting (16.78) in (16.77), the first-order contribution to the negative free energy
yields
W

(1)

=2

V123 V456 H36 G14 G25 +

123456

123456

V123 V456 H36 G12 G45 ,

(16.79)

which corresponds to the Feynman diagrams


W (1) = 2

16.8

(16.80)

Functional Differential Equation for W = ln Z

The derivation of a functional differential equation for the negative free energy W
requires the combination of two independent steps. Consider first the identity
Z

D DA

o
n
2 eE[,A] = 0 ,
1

(16.81)

which immediately yields with the energy functional (16.66)


12 Z + 2

G1
13

1 + 2 g
G23

34

V134

{hA4 iZ}
= 0,
G1
23

(16.82)

where hAi denotes the expectation value of the field A. In order to close the functional differential equation, we consider the second identity
Z

D DA

E[,A]
e
= 0,
A1

(16.83)

1023

16.9 Recursion Relation and Graphical Solution

which leads to
hA1 iZ =

234

V234 H14

Z
.
G1
23

(16.84)

Inserting (16.84) in (16.82), we result in the desired functional differential equation


for the negative free energy W = ln Z:
12 + 2

W
= 2g
G1
23

G1
13

34567

V134 V567 H47

W W
2W
1
1 . (16.85)
1 +
G23 G56
G1
23 G56
)

A subsequent separation (16.47) of the zeroth order (16.78) leads to a functional


differential equation for the interaction part of the free energy W (int) :
Z

12

G1
12

W (int)
g
=
1
4
G12

V123 V456 H36 {G12 G45 + 2G14 G25 }

V123 V456 G12 H36

123456

W (int)
(16.86)
G1
123456
45
)
(
Z
W (int) W (int)
2 W (int)
.
g
V123 V456 H36
1 +
G1
G1
G1
123456
12 G45
12
45

+g

Taking into account the functional chain rules (16.49), (16.50), the functional derivatives with respect to G1 in (16.86) can be rewritten in terms of G:
Z

12

G12

W (int)
g
=
G12
4
+ g
+g

16.9

123456

123456

123456789
1

V123 V456 H36 {G12 G45 + 2G14 G25 }

(16.87)
W (int)
G78
)
(int)
W
W (int)
+
.
G78 G91

V123 V456 H36 {G12 G47 G58 + 2G14 G27 G58 }

V123 V456 H36 G17 G28 G39 G41

2 W (int)
G78 G91

Recursion Relation and Graphical Solution

The functional differential equation (16.87) is now solved by the power series
W

(int)

1
=
p=1 (2p)!

 p

g
4

W (p) .

(16.88)

Using the property (16.32) that the coefficients W (p) satisfy the eigenvalue condition
of the operator (16.31), we obtain both the recursion relation
W

(p+1)

= 4(2p + 1)

(Z

12345678

V123 V456 H36 (G12 G47 G58 + 2G14 G27 G58 )

W (p)
G78

1024

16 Systematic Graphical Construction of Feynman Diagrams . . .

p1
X

(p)

W
2p
+
V123 V456 H36 G17 G28 G39 G41
G78 G91 q=1 2q

123456789
1

(pq)

(q)

W
G78

W
G91

(16.89)

and the initial value (16.79). Using the Feynman rules (16.6), (16.73) and (16.74),
the recursion relation (16.89) reads graphically

(p+1)

W (p)
= 4(2p + 1)
1 2
(

p1
X
q=1

2p
2q

W (p)
+2
1 2

1
2

W (pq)
1 2

2 W (p)
+
1 2 3

1
2

W (q)
, p 1,
3 4
)

1
2

3
4

(16.90)

which is iterated starting from (16.80). In analogy to (16.57), this recursion relation
can be cast in a closed diagrammatic way by using the alternative graphical rules
(16.56):

p+1

4(2p + 1)

p1
X

2p
2q

q=1

+ 2

(16.91)

pq

p 1,

We illustrate the procedure of solving the recursion relation (16.90) by construction


the three-loop vacuum diagrams. Applying one or two functional derivatives to
(16.80), we have
W (1)
= 2
1 2
2 W (p)
1 2 3

+ 4

,
2

+ 4

(16.92)

This is inserted into (16.90) to yield the three-loop diagrams shown in Table with
their multiplicities. The table also contains the subsequent 4-loop results which we
shall not derive here in detail. Observe that the multiplicity of a connected vacuum
diagram D in the 2 A-theory is given by a formula similar to (16.9) in the 4 -theory:
MD2 A =

(2p)!4p
.
2!S+D N

(16.93)

1025

16.10 Computer Generation of Diagrams

Here S and D denote the number of self- and double connections, whereas N represents the number of identical vertex permutations.
The connected vacuum diagrams of the 2 A-theory in Table can, of course, be
converted to corresponding ones of the 4 -theory in Table , by shrinking wiggly
lines to a point and dividing the resulting multiplicity by 3 in accordance with
(16.75). This relation between connected vacuum diagrams in 4 - and 2 A-theory is
emphasized by the numbering used in Table . For instance, the shrinking converts
the five diagrams #4.1-#4.5 in Table to the diagram #4 in Table . Taking into
account the different combinatorial factors in the expansion (16.5) and (16.88) as
well as the factor 3 in the shrinkage (16.75), the multiplicity ME=0
of a 4 -diagram
4
2
results from the corresponding one ME=0
2 A of the A-partner diagrams via the rule
=
ME=0
4

16.10

1
M E=0
.
2
(2p 1)!! A

(16.94)

Computer Generation of Diagrams

Continuing the solution of the graphical recursion relations (16.54) and (16.90) to
higher loops becomes an arduous task. We therefore automatize the procedure by
computer algebra. Here we restrict ourselves to the 4 -theory because of its relevance
for critical phenomena.

16.11

Matrix Representation of Diagrams

To implement the procedure on a computer we must represent Feynman diagrams


in the 4 -theory by algebraic symbols. For this we use matrices as defined in the
textbook in Footnote 3. Let p be the number of vertices of a given diagram and label
them by indices from 1 to p. Set up a matrix M whose elements Mij (0 i, j p)
specify the number of lines joining the vertices i and j. The diagonal elements Mii
(i > 0) count the number of self-connections of the ith vertex. External lines of a
diagram are labeled as if they were connected to a single additional dummy vertex
with number 0. As matrix element M00 is set to zero as a convention. The offdiagonal elements lie in the interval 0 Mij 4, while the diagonal elements for
i > 0 are restricted by 0 Mii 2. We observe that the sum of the matrix elements
Mij in each row or column equals 4, where the diagonal elements count twice,
p
X

j=0

Mij + Mii =

p
X

Mij + Mjj = 4 .

(16.95)

i=0

The matrix M is symmetric and is thus specified by (p+1)(p+2)/21 elements. Each


matrix characterizes a unique diagram and determines its multiplicity via formula
(16.27). From the matrix M we read offdirectly the number of self-, double, triple,

1026

16 Systematic Graphical Construction of Feynman Diagrams . . .

fourfold connections S, D, T, F and the number of external legs E = pi=1 M0i . It


also permits us to calculate the number N of identical vertex permutations. For
this we observe that the matrix M is not unique, since so far the vertex numbering
is arbitrary. In fact, N is the number of permutations of vertices and external lines
which leave the matrix M unchanged [compare to the statement after (16.27)]. If
nM denotes the number of different matrices representing the same diagram, the
number N is given by
P

N=

p
p! Y
M0i ! .
nM i=1

(16.96)

The matrix elements M0i count the number of external legs connected to the ith
vertex. Inserting Eq. (16.96) into the formula (16.27) for E = 0, we obtain the
multiplicity of the diagram represented by M. This may be used to cross-check the
multiplicities obtained before when solving the recursion relation (16.54).
As an example, consider the following diagram of the four-point function with
p = 3 vertices:
.

(16.97)

This diagram can be represented by altogether nM = 3 different matrices, depending


on the labeling of the top vertex with two external legs by 1, 2, or 3:
0
1
2
3

0
0
2
1
1

1
2
0
1
1

2
1
1
0
2

3
1
1
2
0

0
1
2
3

0
0
1
2
1

1
1
0
1
2

2
2
1
0
1

3
1
2
1
0

0
1
2
3

0
0
1
1
2

1
1
0
2
1

2
1
2
0
1

3
2
1 .
1
0

(16.98)

From the zeroth row or column of these matrices or by inspecting the diagram
(16.97), we read off that there exist two one-connections and one two-connection
between external legs and vertices. Thus Eq. (16.96) states that the number of
identical vertex permutations of the diagram (16.97) is 4 (compare the corresponding entry in Table ).
The matrix M contains of course all information on the topological properties
by removing the zeroth row and
of a diagram. For this we define the submatrix M
column from M. This allows us to recognize the connectedness of a diagram: A
is a block madiagram is disconnected if there is a vertex numbering for which M
trix. Furthermore a vertex is a cutvertex, e.g. a vertex which links two otherwise
has an almost block-form for an
disconnected parts of a diagram, if the matrix M
appropriate numbering of vertices in which the blocks overlap only on some diago ii , i.e. the matrix M
takes block form if the ith row and column are
nal element M

1027

16.12 Practical Generation

allows us to recognize a one-particle-reducible diremoved. Similarly, the matrix M


agram, which falls into two pieces by cutting a certain line. Removing a line amounts
by one unit. If the
to reducing the associated matrix elements in the submatrix M

resulting matrix M has block form, the diagram is one-particle-reducible.


So far, the vertex numbering has been arbitrary, making the matrix representation of a diagram non-unique. To achieve uniqueness, we proceed as follows. We
associate with each matrix a number whose digits are composed of the matrix elements Mij (0 j i p), i.e. we form the number with the (p + 1)(p + 2)/2 1
elements
M10 M11 M20 M21 M22 M30 M31 M32 M33 . . . Mpp .

(16.99)

The smallest of these numbers is chosen to represent the diagram unique. For
instance, the three matrices (16.98) carry the numbers
201101120 ,

102101210 ,

101202110 ,

(16.100)

the smallest one being the last. Thus we uniquely represent the diagram (16.97) by
this number.

16.12

Practical Generation

We are now prepared for the computer generation of Feynman diagrams. First the
vacuum diagrams are generated from the recursion relation (16.54). From these the
diagrams of the connected two- and four-point functions are obtained by cutting
or removing lines. A MATHEMATICA program performs this task. The resulting
unique matrix representations of the diagrams up to the order p = 4 are listed in
Table . They are the same as those derived before by hand in Table . Higherorder results up to p = 6, containing all diagrams which are relevant for the five
loop renormalization of the 4 -theory in d = 4 dimensions7 are made available
on internet8 , where also the program can be found.

16.12.1

Connected Vacuum Diagrams

The computer solution of the recursion relation (16.54) necessitates to keep an exact record of the labeling of external legs of intermediate diagrams which arise from
differentiating a vacuum diagram with respect to a line once or twice. To this end
we have to extend our previous matrix representation of diagrams where the external legs are labeled as if they were connected to a simple additional vertex with
number 0. For each matrix representing a diagram we define an associated vector
which contains the labels of the external legs connected to each vertex. This vector
has the length of the dimension of the matrix and will be prepended to the matrix.
7
8

See the textbook in Footnote 3.


http://www.physik.fu-berlin.de/kleinert/294/programs.

1028

16 Systematic Graphical Construction of Feynman Diagrams . . .

By doing so, it is understood that the rows and columns of the matrix are labeled
from 0 to the number of vertices as explained in Subsection IV.A, so that we may
omit these labels from now on. Consider, as an example, the diagram (16.97) of
the four-point function with p = 3 vertices, where the spatial indices 1, 2, 3, 4 are
assigned in a particular order:

(16.101)

In our extended matrix notation, such a diagram can be represented in total by six
matrices:

{}
{1, 2}
{3}
{4}

{}
{1, 2}
{4}
{3}

0
2
1
1

2
0
1
1

1
1
0
2

1
1
2
0

0
2
1
1

2
0
1
1

1
1
0
2

1
1
2
0

{}
{3}
{1, 2}
{4}

{}
{4}
{1, 2}
{3}

0
1
2
1

1
0
1
2

2
1
0
1

1
2
1
0

0
1
2
1

1
0
1
2

2
1
0
1

1
2
1
0

{}
{3}
{4}
{1, 2}

{}
{4}
{3}
{1, 2}

0
1
1
2

1
0
2
1

1
2
0
1

2
1
1
0

0
1
1
2

1
0
2
1

1
2
0
1

2
1

.
1
0
(16.102)

In the calculation of the vacuum diagrams from the recursion relation (16.54), starting from the two-loop diagram (16.55), we have to represent three different elementary operations in our extended matrix notation:
1. Taking one or two derivatives of a vacuum diagram with respect to a line. For
example, we apply this operation to the vacuum diagram #2 in Table
1

3
G12 G34

4 2
=

= 3

G12

This has the matrix representation


2

G12 G34

{}

{}
{}

.(16.103)

{} 0 1 1
{} 0 1 1
0 0 0

0 0 4 =2
{3} 1 0 3 + {4} 1 0 3


G12
{3} 1 3 0
{4} 1 3 0
0 4 0

{}

= 3 {1, 3}
{2, 4}
"

{}
0 2 2

{2,
3}
2 0 2
+

{1, 4}
2 2 0

0 2 2
2 0 2

2 2 0

1029

16.12 Practical Generation








{}

+ {1, 4}
{2, 3}

{}
0 2 2

2 0 2 + {2, 4}
{1, 3}
2 2 0

0 2 2 #
2 0 2
.
2 2 0

(16.104)

The first and fourth matrix as well as the second and third matrix represent
the same diagram in (16.103), as can be seen bu permutating rows and columns
of either matrix.
2. Combining two or three diagrams to one. We perform this operation by creating a block matrix of internal lines from the submatrices representing the
internal lines of the original diagrams. Then the zeroth row or column is
added to represent the respective original external spatial arguments. Let us
illustrate the combination of two diagrams by the example
1

{}

{1}
{2}

{}
{1, 2}
{1}
{2}

0 1 1
1 0 3

1 3 0








0
2
1
1

2
1
0
0

1
0
0
3

1
0
3
0

and the combination of three diagrams by


1
2

1
2

{}
{1, 2}

3
4

0 2
2 1

{}
{1, 2}
{1, 2, 3, 4}
{3, 4}

{}
{1, 2}

0 2
2 1

(16.105)

3
4

,
0
2
4
2

2
1
0
0

{}
{1, 2, 3, 4}
4
0
0
0

2
0
0
1

0 4
4 0

{}
{3, 4}

0 2
2 1

(16.106)

We observe that the ordering of the submatrices in the block matrix is arbitrary
at this point, we just have to make sure to distribute the spatial labels of the
external legs correctly.
3. Connecting external legs with the same label and creating an internal line.
This is achieved in our extended matrix notation by eliminating the spatial
labels of external legs which appear twice, and by performing an appropriate
entry in the matrix for the additional line. Thus we obtain, for instance, from
(16.105)

{}
{}
{}
{}

0
0
0
0

0
1
1
1

0
1
0
3

0
1
3
0

(16.107)

1030

16 Systematic Graphical Construction of Feynman Diagrams . . .

and similarily from (16.106)

{}
{}
{}
{}

0
0
0
0

0
1
2
0

0
2
0
2

0
0
2
1

(16.108)

As we reobtain at this stage connected vacuum diagrams where there are no


more external legs to be labeled, we may omit the prepended vector.
The selection of a unique matrix representation for the resulting vacuum diagrams
obtained at each stage of the recursion relation proceeds as explained in detail in
Subsection IV.A. By comparing we find out which of the vacuum diagrams are topologically identical and sum up their individual multiplicities. Along these lines, the
recursion relation (16.54) is solved by a MATHEMATICA program up to the order
p = 6. The results are shown in Table , available on the internet at the address
stated in Footnote 8. To each order p, the numbers n(0)
p of topologically different
connected vacuum diagrams are
p
n(0)
p

1
1

2
2

3
4

4
10

5
28

6
97

(16.109)

Two- and Four-Point Functions G12 and Gc1234 from


Cutting Lines

16.12.2

Having found all connected vacuum diagrams, we derive from these the diagrams
of the connected two- and four-point functions by using the relations (16.17) and
(16.24). In the matrix representation, cutting a line is essentially identical to removing a line as explained above, except that we now interpret the labels which
represent the external spatial labels as sitting on the end of lines. Since we are
not going to distinguish between trivially crossed graphs which are related by exchanging external labels in our computer implementation, we need no longer carry
around external spatial labels. Thus we omit the additional vector prepended to the
matrix representing a diagram when generating vacuum diagrams. As an example,
consider cutting a line in diagram #3 in Table

G1

= 2

(16.110)

which has the matrix representation

0 0 2
0 2 0
0 1 1
0 0 0

1 0 1 2 = 2 1 1 1 + 2 0 2 + 0 1 2 . (16.111)
G
2 2 0
0 2 1
1 1 1
0 2 1

1031

16.12 Practical Generation

Here the plus signs and multiplication by 2 have a set-theoretical meaning and are
not to be understood as matrix algebra operations. The last two matrices represent,
incidentally, the same graph in (16.110) as can be seen by exchanging the last two
rows and columns of either matrix.
To create the connected four-point function, we also have to consider second
derivatives of vacuum diagrams with respect to G1 . If an external line is cut, an
additional external line will be created which is not connected to any vertex. It can
be interpreted as a self-connection of the zeroth vertex which collects the external
lines. This may be accomodated in the matrix notation by letting the matrix element
M00 count the number of lines not connected to any vertex. For example, taking
the derivative of the first diagram in Eq. (16.110) gives

G1

+ 2

(16.112)

with the matrix notation

G1

1 1 1
0 2 2
0 1 3
0 3 1
0 1 1

1 1 1 = 3 0 1 + 1 1 1 + 2 1 0 + 2 1 1 1 .
1 1 1
2 0 1
3 1 0
1 1 1
1 1 1
(16.113)

The first two matrices represent the same diagram as can be seen from Eq. (16.112).
The last two matrices in Eq. (16.113) correspond to disconnected diagrams: the first
because of the absence of a connection between the two vertices, the second because
of the disconnected line represented by the entry M00 = 1. In the full expression
c,(2)
for the two loop contribution G1234 to the four-point function in Eq. (16.24) all
(2)
disconnected diagrams arising from cutting a line in G12 are canceled by diagrams
resulting from the sum. Therefore we may omit the sum, take only the first term
and discard all disconnected graphs it creates. This is particularly useful for treating
low orders by hand. If we include the sum, we use the prescription of combining
diagrams into one as described above in Subsection IV.B.1, except that we now omit
the extra vector with the labels of spatial arguments.

16.12.3

Two- and Four-Point Function G12 and Gc1234 from Removing Lines

Instead of cutting lines of connected vacuum graphs once or twice, the perturbative
coefficients of G12 and Gc1234 can also be obtained graphically by removing lines.
Indeed, from (16.15), (16.37), (16.47) and (16.49) we get for the two-point function
G12 = G12 + 2

34

G13 G24

W (int)
,
G34

(16.114)

1032

16 Systematic Graphical Construction of Feynman Diagrams . . .

so that we have for p > 0


(p)

G12 = 2

34

G13 G24

W (p)
G34

(16.115)

(p)

at our disposal to compute the coefficients G12 from removing one line in the connected vacuum diagrams W (p) in all possible ways. The corresponding matrix operations are identical to the ones for cutting a line so that in this respect there is no
difference between both procedures to obtain G12 .
Combining (16.114) with (16.11), (16.22) and (16.49), we get for the connected
four-point function
Gc1234

2 W (int)
G56 G78

= 4

G15 G26 G37 G48

G15 G27 (G36 G48 + G46 G38 )

5678

5678

W (int) W (int)
,
G56 G78

(16.116)

which is equivalent to
c,(p)
G1234

2 W (p)
= 4
G15 G26 G37 G48
G56 G78
5678
Z

p1
X

q=1

p
q

!Z

5678

G15 G27 (G36 G48 + G46 G38 )

W (q) W (pq)
. (16.117)
G56 G78

Again, the sum serves only to subtract disconnected diagrams which are created by
the first term, so we may choose to discard both in the first term.
Now the problem of generating diagrams is reduced to the generation of vacuum
diagrams and subsequently taking functional derivatives with respect to G12 . An
advantage of this approach is that external lines do not appear at intermediate steps.
So when one uses the cancellation of disconnected terms as a cross check, there are
less operations to be performed than with cutting. At the end one just interprets
external labels as sitting on external lines. Since all necessary operations on matrices have already been introduced, we omit examples here and just note that we can
again omit external labels if we are not distinguishing between trivially crossed
graphs.
The generation of diagrams of the connected two- and four-point functions has
been implemented in both possible ways. Cutting or removing one or two lines in
the connected vacuum diagrams up to the order p = 6 lead to the following numbers
(p)
c,(p)
(2)
n(2)
p and np of topologically different diagrams of G12 and G1234 :

p
n(2)
p
n(4)
p

1
1
1

2
3
2

3
8
8

4
30
37

5
118
181

6
548
1010

(16.118)

1033

16.12 Practical Generation

Table 16.1 Connected vacuum diagrams and their multiplicities of the 4 -theory up to
five loops. Each diagram is characterized by the vector (S, D, T, F ; N ) whose components
specify the number of self-, double, triple and fourfold connections, and of the identical
vertex permutations, respectively.

W (p)

p
#1
3

(2,1,0,0;1)

#2
24

#3
72

(0,0,0,1;2)

(2,1,0,0;2)

#4
1728

#5
3456

#6
1728

#7
2592

(0,3,0,0;6)

(1,0,1,0;2)

(3,0,0,0;6)

(2,2,0,0;2)

#8
62208

#9
248832

#10
55296

#11
497664

(0,4,0,0;8)

(0,2,0,0;8)

(0,0,2,0;4)

(1,2,0,0;2)

#12
165888

#13
248832

#14
165888

#15
248832

(2,0,1,0;2)

(2,1,0,0;4)

(1,1,1,0;2)

(3,1,0,0;2)

#16
62208

#17
124416

(4,0,0,0;8)

(2,3,0,0;2)

1034

16 Systematic Graphical Construction of Feynman Diagrams . . .

Table 16.2 Connected diagrams of the two-point function and their multiplicities of the
4 -theory up to four loops. Each diagram is characterized by the vector (S, D, T ; N ) whose
components specify the number of self-, double, triple connections, and of the identical
vertex permutations, respectively.
(p)

G12

#1.1
12
(1,0,0;2)

#2.1
192

#3.1
288

#3.2
288

(0,0,1;2)

(1,1,0;2)

(2,0,0;2)

#4.1
20736

#5.1
6912

#5.2
20736

#5.3
13824

(0,2,0;2)

(0,0,1;4)

(1,1,0;2)

(1,0,1;1)

#6.1
10368

#6.2
10368

#7.1
10368

#7.2
20736

(2,0,0;4)

(3,0,0;2)

(1,2,0;2)

(2,1,0;1)

#8.1
995328

#9.1
1990656

#9.2
1990656

#10.1
221184

(0,3,0;2)

(0,1,0;4)

(0,2,0;2)

(0,0,2;2)

#10.2
663552

#11.1
995328

#11.2
1990656

#11.3
995328

(0,1,1;2)

(0,2,0;4)

(1,2,0;1)

(1,2,0;2)

#11.4
3981312

#12.1
995328

#12.2
331776

#12.3
663552

(1,1,0;1)

(2,1,0;2)

(2,0,1;2)

(2,0,1;1)

#12.4
663552
(1,0,1;2)

#13.1
995328

#13.2
995328

(2,0,0;4)

#13.3
1990656

(1,1,0;4)

(2,1,0;1)

#14.1
995328

#14.2
663552

#14.3
663552

#14.4
331776

(1,2,0;2)

(1,1,1;1)

(1,0,1;2)

(0,1,1;4)

#15.1
995328

#15.2
497664

#15.3
497664

#15.4
995328

(3,1,0;1)

(3,1,0;2)

(2,1,0;4)

(2,1,0;2)

#15.5
995328
(3,0,0;2)

#16.1
497664
(3,0,0;4)

#16.2
497664

#17.1
497664

(4,0,0;2)

(1,3,0;2)

#17.2
995328

#17.3
497664

(2,2,0;1)

(2,2,0;2)

1035

16.12 Practical Generation

Table 16.3 Connected diagrams of the four-point function and their multiplicities of the
4 -theory up to three loops. Each diagram is characterized by the vector (S, D, T ; N )
whose components specify the number of self-, double, triple connections, and of the
identical vertex permutations, respectively.
c,(p)

G1234
#1.1.1

24
(0,0,0;24)

#2.1.1,#3.1.1

#3.1.2,#3.2.1

1152,576

1152,1152

1728

2304

(0,1,0;8)

(1,0,0;6)

#4.1.1,#7.1.1

#4.1.2,#5.1.1,#5.2.1

#5.1.2,#5.3.2

41472,20736

165888,41472,41472

27648,27648

62208

248832

55296

(0,2,0;8)

(0,1,0;4)

(0,0,1;6)

#5.2.2,#6.1.1

#5.2.3,#5.3.1,#7.1.2,#7.2.1

#6.1.2,#6.2.2,#7.2.2

82944,41472

82944,82944,41472,41472

20736,20736,82944

124416

248832

124416

(1,0,0;8)

(1,1,0;2)

(2,0,0;4)

#6.1.3,#6.2.1

#7.1.3,#7.2.3

41472,41472

41472,41472

82944

82944

(2,0,0;6)

(1,1,0;6)

#8.1.1,#17.1.1

#8.1.2,#9.2.1,#10.2.1

#8.1.3,#11.1.2,#11.3.1

1990656,995328

3981312,3981312,3981312

7962624,1990656,1990656

2985984

11943936

11943936

(0,3,0;8)

(0,2,0;4)

(0,2,0;4)

#9.1.1,#13.2.1
3981312,1990656
5971968
(0,1,0;16)

#9.1.2
7962624
(0,0,0;24)

#9.1.3,#9.2.3,#11.1.1,#11.4.1
15925248,15925248,7962624,7962624
47775744
(0,1,0;2)

1036

16 Systematic Graphical Construction of Feynman Diagrams . . .

#12.1.4,#12.3.3,#15.1.1,#15.3.2

#12.2.1,#12.4.2

#12.3.2,#12.4.3,#14.2.2,#14.3.2

3981312,3981312,1990656,1990656

1327104,1327104

1327104,1327104,2654208,2654208

11943936

2654208

7962624

(2,1,0;2)

(1,0,1;6)

(1,0,1;2)

#12.3.1,#12.4.4

#13.1.2,#13.3.4,#15.4.2,#15.5.1

#13.1.3,#16.1.1

1327104,1327104

7962624,7962624,3981312,3981312

1990656,995328

2654208

23887872

2985984

(1,0,1;6)

(2,0,0;2)

(2,0,0;16)

#13.2.4,#13.3.5

#13.3.2,#15.2.1,#15.3.3

#14.1.3,#14.2.3,#17.1.3,#17.3.1

3981312,3981312

3981312,995328,995328

3981312,3981312,1990656,1990656

7962624

5971968

11943936

(1,1,0;6)

(2,1,0;4)

(1,2,0;2)

#14.1.4,#15.4.4

#14.3.1,#14.4.1

#15.1.2,#15.5.3,#16.1.3,#16.2.2

3981312,1990656

1327104,1327104

3981312,3981312,1990656,1990656

5971968

2654208

11943936

(1,1,0;8)

(0,0,1;12)

(3,0,0;2)

1037

16.12 Practical Generation

Table 16.4 Connected vacuum diagrams and their multiplicities of the 2 A-theory up to
four loops. Each diagram is characterized by the vector (S, D; N ) whose components specify the number of self- and double connections as well as the identical vertex permutations,
respectively.
W (p)

#1.1
2

#1.2
1

(0,1;2)

(2,0;2)

#2.1
48

#2.2
24

#3.1
96

#3.2
96

#3.3
24

(0,0;8)

(0,2;4)

(0,0;4)

(1,0;2)

(2,1;2)

#4.1
3840

#4.2
11520

#4.3
3840

#4.4
960

#4.5
5760

(0,0;12)

(0,0;4)

(0,0;12)

(0,3;6)

(0,1;4)

#5.1
11520

#5.2
23040

#5.3
11520

#5.4
5760

#6.1
7680

(0,1;2)

(0,0;2)

(1,0;2)

(1,1;2)

(0,0;6)

#6.2
11520

#6.3
5760

#6.4
960

#7.1
11520

#7.2
5760

(1,0;2)

(2,0;2)

(3,0;6)

(0,0;4)

(0,0;8)

#7.3
11520

#7.4
5760

#7.5
1440

#7.6
2880

(1,0;2)

(1,1;2)

(2,2;2)

(2,0;4)

1038

16 Systematic Graphical Construction of Feynman Diagrams . . .

Table 16.5 Unique matrix representation of all connected vacuum diagrams of 4 -theory
up to the order p = 4. The number in the first column corresponds to their graphical representation in Table . The matrix elements Mij represent the numbers of lines connecting
two vertices i and j, with omitting Mi0 = 0 for simplicity. The running numbers of the
vertices are listed on top of each column in the first two rows. The further columns contain
the vector (S, D, T, F ; N ) characterizing the topology of the diagram, the multiplicity M
and the weight W = M/(4!)p p!.
W (1) : 1 diagram
i
j
#
1

i
j
#
2
3

i
j
#
5
4
7
6

i
j
#
10
8
12
14
17
11
9
13
15
16

1
1
Mij
2

(S,D,T,F ;N )

(2,1,0,0;1)

M
3

W 1
8

W (2) : 2 diagrams
1 22
1 12
Mij
(S,D,T,F ;N )
M
0 40 (0,0,0,1;2) 24
1 21 (2,1,0,0;2) 72

1
1
0
0
0
1

1
1
0
0
0
0
0
0
0
0
0
1

22
12
Mij
11
20
21
11

22
12
00
00
01
01
01
01
10
11
11
01

W 1
48
16

W (3) : 4 diagrams
333
123
(S,D,T,F ;N )

310
220
201
111

(1,0,1,0;2)
(0,3,0,0;6)
(2,2,0,0;2)
(3,0,0,0;6)

M
3456
1728
2592
1728

W (4) : 10 diagrams
333 4444
123 1234
Mij
(S,D,T,F ;N )
130 3100 (0,0,2,0;4)
220 2200 (0,4,0,0;8)
111 3100 (2,0,1,0;2)
120 3010 (1,1,1,0;2)
201 2200 (2,3,0,0;2)
210 2110 (1,2,0,0;2)
120 2110 (0,2,0,0;8)
101 2110 (2,1,0,0;4)
111 2001 (3,1,0,0;2)
111 1101 (4,0,0,0;8)

W 1
24
48
32
48

M
55296
62208
165888
165888
124416
497664
248832
248832
248832
62208

W 1
144
128
48
48
64
16
32
32
32
128

1039

16.12 Practical Generation

Table 16.6 Unique matrix representation of all connected two-point function of 4 theory up to the order p = 4. The numbers in the first column correspond to their
graphical representation in Table . The matrix elements Mij represent the numbers of
lines connecting two vertices i and j, with omitting Mi0 = 0 for simplicity. The running
numbers of the vertices are listed on top of each column in the first two rows. The further
columns contain the vector (S, D, T ; N ) characterizing the topology of the diagram, the
multiplicity M and the weight W = M/(4!)p p!.
(2)

(1)

i
j
#
1.1

11
01
Mij
21

i
j
#
7.1
5.1
5.3
4.1
6.1
7.2
5.2
6.2

11
01

G12 : 1 diagram

(S,D,T ;N)

(1,0,0;2)

M
12

i
j
#
3.1
2.1
3.2

W 1
4

(3)

00
00
00
00
01
01
01
01

222
012
Mij
021
030
111
120
011
101
110
111

G12 : 8 diagrams
3333
0123
(S,D,T ;N)

2200
2110
1300
1210
2110
1210
1120
1101

(1,2,0;2)
(0,0,1;4)
(1,0,1;1)
(0,2,0;2)
(2,0,0;4)
(2,1,0;1)
(1,1,0;2)
(3,0,0;2)

M
10368
6912
13824
20736
10368
20736
20736
10368

W 1
16
24
12
8
16
8
8
16

(4)

i
j
#
17.1
12.4
14.2
11.3
14.4
11.1
10.1
9.2
15.3
15.4
13.2
12.3
13.3
11.4
11.2
8.1
9.1
17.2
14.1
15.2
14.3
10.2
12.2
16.1
15.1
17.3
13.1
15.5
12.1
16.2

11
01

222
012

00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
01
01
01
01
01
01
01

001
001
001
001
010
010
010
010
011
011
011
011
011
011
020
020
020
021
021
021
030
030
030
001
001
001
001
011
011
011

G12 : 30 diagrams
3333
44444
0123
01234
Mij
(S,D,T ;N)
0220
22000
(1,3,0;2)
0310
21100
(1,0,1;2)
1120
13000
(1,1,1;1)
1210
12100
(1,2,0;2)
0130
22000
(0,1,1;4)
0220
21100
(0,2,0;4)
1030
13000
(0,0,2;2)
1120
12100
(0,2,0;2)
0111
22000
(2,1,0;4)
0201
21100
(2,1,0;2)
0210
21010
(1,1,0;4)
1011
13000
(2,0,1;1)
1101
12100
(2,1,0;1)
1110
12010
(1,1,0;1)
1011
12100
(1,2,0;1)
1020
12010
(0,3,0;2)
1110
11110
(0,1,0;4)
1001
12010
(2,2,0;1)
1100
11020
(1,2,0;2)
1101
11001
(3,1,0;2)
1001
11110
(1,0,1;2)
1010
11020
(0,1,1;2)
1011
11001
(2,0,1;2)
0111
21100
(3,0,0;4)
1011
12100
(3,1,0;1)
1020
12010
(2,2,0;2)
1110
11110
(2,0,0;4)
1001
11110
(3,0,0;2)
1010
11020
(2,1,0;2)
1011
11001
(4,0,0;2)

M
497664
663552
663552
995328
331776
995328
221184
1990656
497664
995328
995328
663552
1990656
3981312
1990656
995328
1990656
995328
995328
497664
663552
663552
331776
497664
995328
497664
995328
995328
995328
497664

W 1
32
24
24
16
48
16
72
8
32
16
16
24
8
4
8
16
8
16
16
32
24
24
48
32
16
32
16
16
16
32

11
01

G12 : 3 diagrams
222
012

Mij
01
220
10
130
11
111

(S,D,T ;N)

(1,1,0;2)
(0,0,1;2)
(2,0,0;2)

M
288
192
288

W 1
8
12
8

1040

16 Systematic Graphical Construction of Feynman Diagrams . . .

Table 16.7 Unique matrix representation of all connected two-point function of 4 theory up to the order p = 4. The numbers in the first column correspond to their
graphical representation in Table . The matrix elements Mij represent the numbers of
lines connecting two vertices i and j, with omitting Mi0 = 0 for simplicity. The running
numbers of the vertices are listed on top of each column in the first two rows. The further
columns contain the vector (S, D, T ; N ) characterizing the topology of the diagram, the
multiplicity M and the weight W = M/(4!)p p!.
c,(1)

i
j
#
1.1.1

G1234 : 1 diagram
11
01
Mij (S,D,T ;N ) M
40
(0,0,0;24) 24

W 1
24

c,(2)

i
j
#
3.1.2, 3.2.1
2.1.1, 3.1.1

G1234 : 2 diagrams
11 222
01 012
Mij
(S,D,T ;N )
11 310 (1,0,0;6)
20 220 (0,1,0;8)

M
2304
1728

W 1
12
16

c,(3)

i
j
#
5.1.2, 5.3.2
4.1.1, 7.1.1
6.1.3, 6.2.1
7.1.3, 7.2.3
5.2.2, 6.1.1
5.2.3, 5.3.1, 7.1.2, 7.2.1
4.1.2, 5.1.1, 5.2.1
6.1.2, 6.2.2, 7.2.2

G1234 : 8 diagrams
11 222 3333
01 012 0123
Mij
(S,D,T ;N )
00 130 3100 (0,0,1;6)
00 220 2200 (0,2,0;8)
01 111 3100 (2,0,0;6)
01 120 3010 (1,1,0;6)
01 210 2110 (1,0,0;8)
10 111 2200 (1,1,0;2)
10 120 2110 (0,1,0;4)
11 101 2110 (2,0,0;4)

c,(4)

i
j
#
13.2.4, 13.3.5
12.3.1, 12.4.4
11.4.5, 15.3.1, 15.4.1
11.1.3, 11.2.1
8.1.1, 17.1.1
9.1.1, 13.2.1
15.2.3, 15.4.6
17.1.4, 17.2.4
14.1.4, 15.4.4
12.2.1, 12.4.2

11
01
00
00
00
00
00
00
00
00
00
00

G1234 : 37
222 3333
012 0123
Mij
011 1210
011 1300
011 2110
020 1120
020 2020
020 2110
021 1101
021 1200
021 2100
030 1011

M
55296
62208
82944
82944
124416
248832
248832
124416

W 1
36
32
24
24
16
8
8
16

diagrams
44444
01234
(S,D,T ;N )

31000
30100
22000
31000
22000
21100
31000
30010
21010
31000

(1,1,0;6)
(1,0,1;6)
(1,1,0;4)
(0,2,0;6)
(0,3,0;8)
(0,1,0;16)
(2,1,0;6)
(1,2,0;6)
(1,1,0;8)
(1,0,1;6)

M
7962624
2654208
11943936
7962624
2985984
5971968
3981312
3981312
5971968
2654208

W 1
24
72
16
24
64
32
48
48
32
72

1041

16.12 Practical Generation

(Table continued.)
c,(4)

i
j
#
14.3.1, 14.4.1
10.2.2, 12.4.1
11.2.4, 11.3.2, 17.1.2, 17.2.1
12.3.2, 12.4.3, 14.2.2, 14.3.2
9.2.2, 14.1.1, 14.4.3
9.1.3, 9.2.3, 11.1.1, 11.4.1
10.1.1, 10.2.3, 14.2.1, 14.4.2
13.3.2, 15.2.1, 15.3.3
11.2.3, 11.4.2, 13.2.2, 13.3.1
8.1.3, 11.1.2, 11.3.1
15.1.4, 15.4.5
13.1.3, 16.1.1
16.1.4, 16.2.1
15.3.4, 15.5.4
12.1.3, 16.1.2
11.3.3, 11.4.4, 12.1.1, 12.4.5
14.1.3, 14.2.3, 17.1.3, 17.3.1
15.1.2, 15.5.3, 16.1.3, 16.2.2
13.1.2, 13.3.4, 15.4.2, 15.5.1
15.1.3, 15.4.3, 17.2.3, 17.3.2
12.1.4, 12.3.3, 15.1.1, 15.3.2
11.4.6, 13.1.1, 13.2.3
8.1.2, 9.2.1, 10.2.1
12.1.2, 12.2.2, 13.3.3, 17.2.2
11.2.2, 11.4.3, 14.1.2, 14.3.3
9.1.2
15.2.2, 15.5.2

G1234 : 37 diagrams
11 222 3333 44444
01 012 0123 01234
Mij
00 030 1110 30010
00 030 2010 21010
00 101 1210 22000
00 101 1300 21100
00 110 1120 22000
00 110 1210 21100
00 110 1300 20200
00 111 1101 22000
00 111 1200 21010
00 120 1200 20110
01 001 1120 31000
01 001 2110 21100
01 011 1011 31000
01 011 1110 30010
01 011 2010 21010
01 100 1120 21100
01 100 1210 20200
01 101 1101 21100
01 101 1110 21010
01 101 1200 20110
01 110 1101 20200
01 110 1110 20110
10 100 1120 12100
10 101 1101 12100
10 101 1110 12010
10 110 1110 11110
10 111 1101 11001

(S,D,T ;N )

(0,0,1;12)
(0,0,1;8)
(1,2,0;2)
(1,0,1;2)
(0,2,0;4)
(0,1,0;2)
(0,1,1;2)
(2,1,0;4)
(1,1,0;2)
(0,2,0;4)
(2,1,0;6)
(2,0,0;16)
(3,0,0;6)
(2,0,0;12)
(2,0,0;8)
(1,1,0;2)
(1,2,0;2)
(3,0,0;2)
(2,0,0;2)
(2,1,0;2)
(2,1,0;2)
(1,0,0;4)
(0,2,0;4)
(2,1,0;2)
(1,1,0;2)
(0,0,0;24)
(3,0,0;6)

M
2654208
3981312
11943936
7962624
11943936
47775744
7962624
5971968
23887872
11943936
3981312
2985984
3981312
3981312
5971968
23887872
11943936
11943936
23887872
11943936
11943936
23887872
11943936
11943936
23887872
7962624
3981312

W 1
72
48
16
24
16
4
24
32
8
16
48
64
48
48
32
8
16
16
8
16
16
8
16
16
8
24
48

It is simplicity that makes the uneducated


more effective than the educated
Aristoteles (384 BC322 BC)

15
Path Integral Calculation of Effective Action.
Loop Expansion
The path integral formula for Z[j] can be employed for a rather direct evaluation
of the effective action of a theory. Diagrammatically, the result will be the loop
expansion organized by powers of the Planck quantum h
[1].

15.1

General Formalism

Consider the generating functional of all Green functions


Z[j] = eiW [j]/h ,

(15.1)

where W [j] is the generating functional of all connected Green functions. The vacuum expectation of the field, the average
(x) h(x)i,

(15.2)

is given by the first functional derivative


(x) = W [j]/j(x).

(15.3)

This can be inverted to yield j(x) as an x-dependent functional of (x):


j(x) = j[](x).

(15.4)

This is used to form the Legendre transform of W [j]:


[] W [j]

d4 xj(x)(x).

(15.5)

where j(x) on the right-hand side is replaced by (15.4). This is the effective action
of the theory. Its first functional derivative gives back the current
[]
= j(x).
(x)
988

(15.6)

989

15.1 General Formalism

The generating functional of all connected Green functions can be recovered from
the effective action by the inverse Legendre transform
W [j] = [] +

d4 xj(x)(x).

(15.7)

Let us compare this with the the path integral formula for the generating functional Z[j]:
Z[j] =

D(x)e(i/h){A[]+ d xj(x)(x)}
R
.
D(x)e(i/h)A0 []
R

(15.8)

Using (15.1) and (15.5), this amounts to the path integral formula for []:
e {[]+
i
h

d4 xj(x)(x)}

=N

D(x)e(i/h){A[]+

d4 xj(x)(x)}

(15.9)

with the normalization factor


N =

Z

D(x)e(i/h)A0 []

1

(15.10)

In writing this we have explicitly displayed the fundamental action quantum h


,
which is a measure for the size of quantum fluctuations.
There are many physical systems, for which quantum fluctuations are rather
small except in the immediate vicinity of certain critical points. For these systems,
it is desirable to develop a method of evaluating (15.9) as an expansion in powers of h
,
which amounts to taking into account successively increasing quantum fluctuations.
In the limit h
0, the path integral over the field (x) in (15.8) is dominated
by the classical solution cl (x) which extremizes the exponent

A
= j(x).

=cl (x)

(15.11)

At this level we therefore identify


W [j] = [] +

d xj(x)(x) = A[cl ] +

d4 xj(x)cl (x).

(15.12)

Of course, cl (x) is a functional of j(x), so that we may write it more explicitly, as


cl [j](x). By differentiating W [j] with respect to j, we have from the general fist
part of Eq. (15.2):
(x) =

W
=
++j .
j
j
j

(15.13)

Inserting the classical field equation (15.11), this becomes


(x) =

A cl
cl
+ cl + j
= cl .
cl j
j

(15.14)

990

15 Path Integral Calculation of Effective Action. Loop Expansion

Thus, to this approximation, (x) coincides with the classical field cl (x). Replacing
cl (x) (x) on the right-hand side of Eq. (15.12), we therefore obtain the lowestorder result (zeroth order in h
):
[] = A[]

(15.15)

i.e., the effective action equals the fundamental action. We have shown in Chapter
14 that all vertex functions can be obtained from the derivatives of [] with respect
to at the equilibrium value of for j = 0.
For the 4 -theory with O(N)-symmetry the zeroth-order effective action is
[] =

1
m2 2
g  2 2
dx

.
(a )2
a
2
2
4! a
4

"

(15.16)

0, and there are only two


For massive particles,1 this has an extremum at
non-vanishing vertex functions (n) (x1 , . . . , xn ):
n=2




2A
2


(2)
=


(x1 , x2 )ab
a (x1 )b (x2 ) a =0
a (x1 )b (x2 ) a =a =0
= ( 2 m2 )ab (4) (x1 x2 ).

(15.17)

This determines the inverse of the propagator:


(2)

ab (x1 , x2 ) = [ihG1 ]ab (x1 , x2 ),

(15.18)

Thus we find to this zeroth-order approximation that Gab (x1 , x2 ) is equal to the free
propagator:
Gab (x1 , x2 ) = G0ab (x1 , x2 )
(15.19)
n=4
(4)

abcd (x1 , x2 , x3 , x4 )

4
= gTabcd ,
a (x1 )b (x2 )c (x3 )d (x4 )

(15.20)

with

1
Tabcd = (ab cd + ac bd + ad bc ),
(15.21)
3
which is just the fundamental vertex (11.229).
According to the definition of the effective action, all diagrams of the theory
can be composed from the propagator Gab (x1 , x2 ) and this vertex via tree diagrams.
Thus we see that in this lowest approximation, we recover precisely the subset of
all original Feynman diagrams with a tree-like topology. These are all diagrams
which do not involve any loop integration. Since the limit h
0 corresponds to the
classical equations of motion with no quantum fluctuations we conclude: Classical
field theory corresponds to tree diagrams.
The use of the initial action as an approximation to the effective action neglecting
fluctuations is often referred to as mean-field theory.
1

Also the case m2 < 0 corresponds to physical systems and will be discussed separately

991

15.2 Quadratic Fluctuations

15.2

Quadratic Fluctuations

In order to find the h


-correction to this approximation we expand the action in
powers of the fluctuations of the field around the classical solution
(x) (x) cl (x),

(15.22)

and perform a perturbation expansion. The quadratic term in (x) is taken to be


the free-field action, the higher powers in (x) are the interactions. Up to second
order in the fluctuations (x), the action is expanded as follows:
A[cl + ] +

= A[cl ] +
+

d4 xj(x)[cl (x) + (x)]


Z

d4 xj(x)cl (x) +

A
d4 x j(x) +

(x)
(x) =cl




2A

3
.

(y)
+
O
()
d xd y (x)
(x)(y) =cl
4

(15.23)

The curly bracket term that is linear in the variation vanishes due to the extremality property of the classical field cl expressed by the field equation (15.11).
Inserting this expansion into (15.9), we obtain the approximate expression

i Z

2A

cl }
D exp
Z[j] N e(i/h){A[cl ]+

d4 xd4 y (x)
(y)
h

(x)(y) =cl
(15.24)
due to
We now observe that the fluctuations in will be of average size h
the h
-denominator in the Fresnel integrals over in (15.24). Thus the fluctuations
()n are on the average of relative order h
n/2 . If we ignore corrections of order h
3/2 ,
the fluctuations remain quadratic in and we may calculate the right-hand side of
(15.24) as
R

(i/
h){A[cl ]+

Ne

d4 xj

d4 xj(x)cl (x)}

"

2A
det
(x)(y)

= (det iG0 )1/2 e(i/h){A[d ]+

(15.25)
=cl

d4 xj(x)cl (x)+i(
h/2)Tr log[2 A/(x)(y)|=cl }

Comparing this with the left-hand side of (15.9), we find that to first order in h
, the
effective action may be recovered by equating
[] +

d4 xj(x)(x) = A[cl ] +

d4 xjcl +

2A
ih
Tr log
(cl ) .
2
(x)(y)

(15.26)

In the limit h
0, the trace log term disappears and (15.26) reduces to the classical
action.
To include the h
-correction into [], we expand W [j] as
W [j] = W0 [j] + h
W1 [j] + O(h2 ).

(15.27)

992

15 Path Integral Calculation of Effective Action. Loop Expansion

Correspondingly, the field differs from cl by a correction of order h


2.
= cl + h
1 + O(h2 ).

(15.28)

Inserting this into (15.26), we find


[] +

d xj = A [ h
1 ] +

d xj h

+O h
2 .

2 A
i

Tr log
d xj1 + h
2
a b =h1

Expanding the action up to the same order in h


gives
(

A[]
[] = A[] + h

j 1 +

+O h
2 .

i
2 A
d4 x j + h
Tr log

2
a b =

(15.29)

But because of (15.11), the curly-bracket term is only of order O(h2 ), so that we
find the one-loop form of the effective action
1
m2 2
g  2 2
2
[] = 0 [] + 1 [] =
dx

()
a
2
2
4! a


i
g
2
2
2
+ h
ab c + 2a b , (15.30)
Tr log m
2
6
Z

"

Tr log [ 2 m2 ] . In the special


where we dropped the infinite additive constant 2i h
case of a one-component real field, this becomes
m2 2
g
1
[] = d x ()2
4
2
2
4!


i
g
+
h
Tr log 2 m2 2 .
2
2
Z

"

(15.31)

What is the graphical content of the set of all Green functions at this level? For
j = 0, we find that the minimum lies at = 0 j = 0, as in the mean-field
approximation. Around this minimum, we may expand the trace log in powers of
, and obtain
g
i
h
Trlog 2 m2 2
2
2
!


i
i
i
2
2
2
=
h
Trlog m + h
Trlog 1+ 2
ig
2
2
m2 2


The second term can be expanded in powers of 2 as follows:

h
X
g
i
i
2 n=1
2

n

1
i
Tr
2
2
n
m2


n

(15.32)

993

15.2 Quadratic Fluctuations

If we insert
G0 =

i
,
m2

(15.33)

then (15.32) can be written as



g
h
X
h

i
i Tr log 2 m2 i
2
2 n=1
2

n

n
1 
Tr G0 2 .
n

(15.34)

More explicitly, the terms with n = 1 and n = 2 read:


h

g d4 xd4 y (4) (x y)G0(x, y)2 (y)


2
Z
g2
d4 xd4 yd4z 4 (x z)G0 (x, y)2 (y)G0(y, z)2 (z) + . . . .
+ih
16
Z

(15.35)

The expansion terms of (15.34) for n 1 correspond obviously to the Feynman


diagrams

A[cl ] =

(15.36)

Thus the series (15.34), is a sum of all diagrams with one loop and any number of
fundamental 4 -vertices
To systematize the entire expansion (15.34), the trace log term may be pictured
by a single-loop diagram


h

i Tr log 2 m2 =
2

(15.37)

The first two diagrams in (15.36) contribute corrections to the vertices (2) and
(4) of (15.17), (15.20). The remaining ones produce higher vertex functions and
lead to more involved tree diagrams. Note that only the first two corrections are
formally divergent, all other Feynman integrals converge. In momentum space we
find from (15.35)
dk 4
i
g
(15.38)
4
2
2 (2) k m2 + i
#
"Z
g2
d4 k
i
i
(4)
(qi ) = g i
+ 2 perm .
2
(2)4 k 2 m2 + i (q1 + q2 k)2 m2 + i
(15.39)
(2) (q) = q 2 m2 h

The convergence of all higher diagrams the expansion (15.36) is ensured by the
renormalizability of the theory since only up to n = 4, a counter term may be formed

994

15 Path Integral Calculation of Effective Action. Loop Expansion

in the same form as those in the original Lagrangian. In euclidean form, we may
write (15.38) and (15.39) as
g
(2) (q) = q 2 + m2 + h
D1 ,
2
2
g
(4) (qi ) = g h
[I (q1 + q2 ) + 2 perm] ,
2


(15.40)
(15.41)

where D1 and I(q) are the Fennman integrals (11.26) and (11.29). These are renormalized with the help of counter terms calculated in Section 11.3, yielding the renormalized diagrams
(2)

R (q) = (q 2 + m2R
g2
(4)
R (qi ) = g h
{[I (q1 + q2 ) + 2 perm] 3I(0)} .
2

(15.42)
(15.43)

If we regularize the integrals dimensionally in 4 dimensions, the counterterms


take the forms calculated in Section 11.5, and (15.42) becomes
(2)
R (q)

g
= q 2 + m2 + h
S
D m2
2

m2
1
(2 /2) (1 + /2)
2
2

(4)

R (qi ) = g h

!/c

g2
[IR (q1 + q2 ) + 2 perm] .
2

1
+ ,

(15.44)
(15.45)

where [compare (11.175)]


q2
1
D 1 + Lm (q) + log 2 + O().
IR (q) = S
2

"

(15.46)

As far as the original effective action (15.31) to order h


is concerned, we may add
the counter terms as
[] =

m2 2
g
1
d x
()2
4 i Tr log 2 m2
2
2
4!
2
Z
Z
Z
Z
4
g
g2
dq
d4 q
1
1
4
2
+
d x (x)
4 (2)4 q 2 + m2
16 (2)4 (q 2 + m2 )

g 2

d4 x 4

The divergent integrals in the last two terms can be evaluated with any regularization
method. In dimensional regularization in 4 dimensions, only the singular 1/parts of the two integrals are selected, and (15.47) becomes
[] =

m2 2 g 4
g
1
()2
+ i Tr log 2 m2 2
d4 x
2
2
4!
2
2
2
2
g
2m
2 cg 4 ,
(15.47)
cm
2
4!
(

995

15.2 Quadratic Fluctuations


2

where m c and g c a the divergent expressions (11.51) and (11.70).


The expansions in powers of has an interesting feature which is worth pointing
out at this place and which has interesting applications in the description of critical
phenomena. For small g, Eq. (15.44) can be rewritten as
(2)
R

g
m2
= q 2 + m2 + h
S
D log 2
4

(15.48)

(15.49)

which is, to the same order in g, equal to

m2
(2)
R (q) = q 2 +
2

!1+h g SD
4

2 .

This means that the vertex function at q = 0 has a non-integer power in the mass
parameter which depends on the coupling strength.
(2)
R (0)

m2
=
2

(15.50)

where
g
=1+h
S
D
(15.51)
4
The important point is that this power is measureable as an experimental quantity
called the susceptibility. It is the so-called critical exponent of the susceptibility [see
(21.153).2 A similar power behaviour is found for (4) , but here only in the limit
m2 0 where Lm log m2 /2 , and thus

(4)

3
m2
g 1+h
gS
D log 2
4

m2
g
2

m0

!h 3 g SD

+ O(m2 )

(15.52)

Also this power behaviour is measurable and defines the critical index via the
so-called scaling relation
3
D ,
2 h
gS
4

(15.53)

so that
1
1
h
gS
D .
(15.54)
2
4
The higher powers of are accompanied by terms which are more and more singular
in the limit m 0. We see from the Feynman integrals in (15.34) that the diagrams
in (15.36) behave like m4n for m 0, and so do the associated effective action
terms n .
=

In Section 21.8 we shall see that the coupling strength


 governing the behavior of all observable
quantities in the limit m2 0 is given by hg S
D = 32 .

996

15.3

15 Path Integral Calculation of Effective Action. Loop Expansion

Massless Theory

The case m = 0 deserves special attention. It is referred to the critical theory. This
name reflects the relevance of this limit for the behaviour of physical systems at a
critical temperature where fluctuations are of infinite range. Let us study directly
the effective potential of a critical 4 -theory at a constant field at the one loop
level. It is defined by v() = []/V T . Then the argument in the trace log term
becomes diagonal in momentum space and reduces to a simple momentum integral.
After a Wick rotation the trace log term, it can be calculated, and the renormalized
potential corresponding to the effective action (15.47) can be written as
g 2
dD q
m2 2 g 4 1
+ +
log
1
+
v() =
2
4!
2 (2)D
2 q 2 + m2
Z
Z
g
dD q
g2
dD q
1
1
2

+
4 ,
D
2
2
D
4 (2) q + m
16 (2) (q 2 + m2 )2
!

(15.55)

where we dropped the infinite additive constant


1
2



dD q
2
2
.
log
q
+
m
(2)D

If we attempt to take this expression to the m = 0-limit, the counter terms are
no longer defined. We now see an important advantage of the -expansion: The
singular 1/-terms are independent of the mass, so that the poles at = 0
g 2
1

mS
D 2 ,
4

g 2 1 4
S
D
16

(15.56)

regularize the potential for any mass. Note, however, that an auxiliary mass had
to be introduced to define these expressions. In the case m 6= 0, can be chosen
equal to m. But for m = 0 we must use an arbitrary nonzero auxiliary mass.
t us now verify that the effective potential remains indeed finite for m = 0.
There, v() becomes
g
1 Z dD q
g 2 g 2 S
D 4
v() = 4+
+
log
1+
.
D
2
4!
2 (2)
2q
16
!

(15.57)

and displays an important feature: When expanding the logarithm in powers of ,


the expansion terms correspond to increasingly divergent Feynman integrals
Z

dD q 1
.
(2)D (q 2 )n

Contrary to the previously regularized divergencies coming from the large-q 2 regime,
these divergencies are due to the q = 0 -singularity of the massless propagators
G0 (q) = i/q 2 . This means that they are IR-singularities. It is a pleasant feature

997

15.3 Massless Theory

of the sum over all these terms that the final result has lost the IR singularities in
favor of a weak logarithmic singularity, thuse removing the IR problem. The only
singularities which remain lie in the ultraviolet, and these are taken care of by the
counter terms.
Let us calculate the renormalized potential explicitly. For this we have to extend
the formulas of dimensional regularization to the integral in (15.57). Using
Z

d-D k

1 (D/2)(1 D/2)
1
1
=
S

D
(k 2 + m2 )
2
(1)
(m2 )1D/2

(15.58)

we integrate in m2 and find


Z



1 (D/2)(1 D/2) 2  2 D/2
m
.
D
d-D k log k 2 + m2 = S
2
(1)
D

(15.59)

In 4 dimensions, this behaves like


S
D [1/ + O()]

  2
1
(1 + ) m2 2 ,
2

(15.60)

So that the potential becomes


1 1
g 4 1
+ S
D
v() =
4!
4
4




g 2

2

g2
1
S
D 4 .
16

(15.61)

To perform an -expansion of this, we divide the coupling constant and field by a


scale parameter involving . Then g and /1/2 are dimensionless quantities,
in terms of which

g
v() = 4
4!

1/2

!4

g 2 2

+
S
D
1/2
16

1 1
1
D
+ S
4
4


!4

g 2
2 2

!2
2

(15.62)

If we use the dimensionless coupling constant S


D g and the reduced field
/1/2 as new variables, then

1 1
4 4 1

+
v() =
S
D 4!
4
4


!2

1 log
2
2

2 4
+
. (15.63)
16

To zeroth order in , the prefactor is equal to 4 8 2 . Thus the massless limit of the
effective potential is well defined in D = 4 dimensions. There is, however, a special
feature: The finiteness is achieved at the expense introducing the extra parameter
.

998

15 Path Integral Calculation of Effective Action. Loop Expansion

The most important property of the critical potential is that it cannot be expanded in powers of . Instead, the expression (15.63) can be rewritten, correctly
up to order 2 , as as
"

3
4 1
2
1


v() =
S
D 6 2
16


4 1
3
2
1
=
S
D 6 2
16
"

#2

3
+
8

#2+ 3

!2

+ O(3 ).

2
log
2

(15.64)

is
This is the typical power behavior of crtical phenomena. The power of
3
referred to as critical index 1 + , so that here = 3 + 2 .
What is the coupling strength in the critical regime? The counter term proportional to 4 Eq. (15.57) implies that the use of the renoralized coupling constant in
the subsequent subtracted expressions that the Lagrangian contains as bare coupling
constant
3
S
D
gB = g + g 2
+ ... ,
2

(15.65)

Alternatively, this may be rewritten as


1
3S
D
1
=
+ ... ,

gB
g
2

(15.66)

In the strong-coupling limit gB , the renormalized coupling has a finite value


3S
D
1
=
+ ... .

g
2

(15.67)

2
S
D g + . . . ,
3

(15.68)

This implies that

with the omitted terms being of higher order in .


For the critical exponents , , in (15.52), (15.54), (15.64), this implies that
1
1 1
= 1 + , = , = 3 + .
6
2 6

(15.69)

It should be pointed out that the potential in the first line of (15.64) has another
minimum away from the origin at
2 1
2 3
3 2 3
+ log

+
= 0.
3!
8
2
2
16
!

(15.70)

4
2
=
log
2
3

(15.71)

This is solved by

999

15.3 Massless Theory

or
2 = 2 e4 3.

(15.72)

However such a solution found for small cannot be trusted. The higher loops
to be discussed later and neglected up to this point will produce more powers of
2 /2, and the series cannot be expected to converge at such a large . As
log
a matter of fact, the approximate exponentiation performed in the second line of
(15.64) does not show this minimum and will be seen, via the methods to be described later, to be the correct analytic form of the potential to all orders in for
small enough and .
If we want to apply the formalism to a gas of bosons, we must discuss the case
of a general O(N)-symmetric version of the effective potential based on the action
(15.30), Introducing the variable y g2 /m2 , this becomes
4 1

v() = m g

y y2
y
3N
g
y
1+ 2 log 1+
+

+ (N 1) 12 +
2 24 2(4N) 16
6
6

2 

 )
g
y
y
+
1+
1+2 log 1+
.
16
2
2




(15.73)

where g = 6/(N + 8).


It is instructive to determine from this all critical exponents of the phenomenological scaling relations. First we deduce from v() the equation of state a la Widom
by forming the derivative

H v()/
.

(15.74)

x m2 /y 1/2 ,

(15.75)

Introducing the variable

we find

"

3
N 1
(x + 16 ) +
(x + 61 )(log(x + 16 ) + 1)
N +8
6
#)
1
1
1
+ (x + 2 )(log(x + 2 ) + 1)
(15.76)
2

where the variable x is related to y by the scaling relation


x m2 /y 1/2 .

(15.77)

To first order in , the relation is


m2
1
x=
1
1 log y + . . . ,
y
2
"

(15.78)

1000

15 Path Integral Calculation of Effective Action. Loop Expansion

or
1
m2
1+
1 log x + . . . ,
y=
y
2

(15.79)

1
3
1 =
.
2
N +8

(15.80)

"

where

Inserting this into (15.76) we find


f (x),
H=

(15.81)

= 3 + ,

(15.82)

f (x) = 1 + x + f1 (x),

(15.83)

with

and

and
f1 (x) =

15.4

1
[(N 1)(x + 1) log(x + 1) + 3(x + 3) log(x + 3)
2(N + 8)
9(x + 1) log 3 + 6x log 2].

(15.84)

-2
Effective Action to Order h

Let us now find the next correction to the effective action.3 Instead of truncating
the expansion (15.23), we keep all terms, reorganizing only the linear and quadratic
terms as in the passage to (15.23). This yields the exact expression
i

e h {[]+j} = e h W [j] = e h (A[cl ]+jcl ) 2 Tr log A [cl ] e h h

W2 [cl ]

(15.85)

where W2 [cl ] is defined by


e

i 2
h W2 [cl ]

De h { 2 A [cl ]+I[cl ,]}


,
1
R
i
De h { 2 A [cl ]}
i

(15.86)

and R is the remainder of the fluctuating action

1
R [cl , ] = A [cl + ] A[cl ] A [cl ].
2

(15.87)

The subscript denotes functional differentiation and integration symbols. Adjacent


space-time variables have been omitted, for brevity. We have renamed (x) by (x).
3

R. Jackiw, Phys. Rev. D 9, 1687 (1976)

15.4 Effective Action to Second Order in h

1001

Note that by construction, R is at least cubic in . Thus the path integral


(15.87) may be considered as the generating functional Z of a theory of field with
a propagator
G[cl ] = ih{A [cl ]}1
and an interaction R which both depend in a complicated manner on j via cl . We
know from the previous sections, and will immediately see this explicitly, that h
2 W2
is indeed of order h
2 . Let us again set
= cl + h
1 .

(15.88)

We shall express W [j] in the form


W [j] = A[cl ] + cl j + h
1 [cl ]

(15.89)

and collect one-and two-loop corrections in the term


i
1 [cl ] = Tr log ihG 1 + h
W2 [cl ].
2

(15.90)

The functional W [j] depends explicitly on j only via the second term, in all others
the j dependence is due to cl [j]. We may use this fact by expressing j as a function
of cl and writing
W [cl ] = A[cl ] + cl j[cl ] + 1 .

(15.91)

If we now insert (15.28) and re-expand everything around X rather than xcl :
1 2
[] = A[] h
A []1 h
1 j[] + h
2 1 j []1 + h
1 iG 1 []1
2
+h
1 [] h
2 1 []1 + O(h3 )
(15.92)


1
= A[] + h
1 [] + h
2
1 iG 1 []1 + 1 j []1 1 1
2
From this we find the correction 1 to the vacuum expectation value of . Considering W [j] in (15.89) again as functional in j we see that
= W [j]/j = cl + h
1cl [cl ]cl /j,

(15.93)

1 = 1cl [cl ]cl /j.

(15.94)

such that

But the derivative cl /j is known from


j/cl = A [cl ] = ihG 1 [cl ]

(15.95)

1002

15 Path Integral Calculation of Effective Action. Loop Expansion

such that
1 = 1cl [cl ]iG[cl ].

(15.96)

Inserting these equations into (15.92), the h


2 -term becomes
h
2
1 iG 1 []1 + O(h3 ).
2

(15.97)

In the derivative 1 , only the trace log term in (15.90) has to be included, i.e.,
!

i
D ,
D[X] = Tr D 1
2
X

(15.98)

so that the effective action becomes, to this order in h


,
[] = A[] + h
1 [] + h
2 2 []
i
2 W2 []
= A[] + Tr log iG 1 [] + h
2
!
!
i
h
2 i
1
1
iG Tr G G
.

Tr G G
2 2

(15.99)

We now calculate W2 [] to lowest order in h


. The remainder R in (15.87) has the
expansion
R[; ] =

1
1 3
A [] + 4 A [],
3!
4!

where we have replaced cl by + O(h). In order to obtain W2 , we have to calculate


all connected vacuum diagrams for the interactions r with a -particle propagator
h
G[](x, x ). Since every contraction brings in a factor h
, we can truncate the
4
expansion (15.100) after . Thus the only contribution to W2 [] are the connected
vacuum diagrams

(15.100)

where a line stands now for G[], a four-vertex for


A [] = iG 1 [] ,

(15.101)

A [] = iG 1 [] .

(15.102)

and a three-vertex for

15.4 Effective Action to Second Order in h

1003

Only the first two graphs are one-particle irreducible. It is now a pleasant feature
that the third graph cancels with the last term in (15.99). In order to see this we
write the diagram more explicitly as
h
2
G1 2 A1 2 3 G3 3 A3 1 2 G1 2
8

(15.103)

which is now part of iW2 . Thus, only the one-particle irreducible vacuum graphs as
make up the h
2 -correction to []:
ih2 2 [] = ih2

3
1
G12 A1 2 3 4 G34 + ih2 2 A1 2 3 G1 1 G2 2 G3 3 A1 2 3
4!
4!
(15.104)

whose graphical representation is


ih2 2 [] =

(15.105)

This topological property of the diagram is true for arbitrary orders in h


. To calculate , we define fundamental vertices of nth order
A1 ...n [],
and sum up all connected one particle irreducible vacuum graphs

ihn n [] =

n2

(15.106)

Note the content of the lines G in terms of fundamental Feynman diagrams. The
lines in these diagrams contain infinite sums of diagrams in the original perturbation
4

Note that each line carries a factor h from the propagator, the n-vertex contributes a factor
.

1n

1004

15 Path Integral Calculation of Effective Action. Loop Expansion

expansion. If propagators of the original -theory are indicated by a dashed line,


an expansion of G in powers of reveals that every line corresponds to a sum

(15.107)
line emits 1, 2, 3 . . . lines via vertices n in arbitrary combinations.
The expansion to order h
n is an expansion according to the number of loops.
The general structure of the series (15.106) becomes rapidly involved functional
formalism. It is treated in the next section.

15.5

Effective Action to all Orders in h

In order to find the general loop expansion, let us expand not around the classical
colution cl but around some other functional 0 = 0 [j] whose properties will be
specified later. Then the generating functional is
i

Z[j] = e h W [j] = N e h {A[0 ]+0 j+h1 [0 [j]]}

(15.108)

where
i1 [0 ]

G e h {A[0 + ]A[0 ]+j[0] } .

(15.109)

On the right-hand side we have inverted the relation 0 [j] and expressed j as a
functional of 0 . The field may be calculated as
(

1 [0 ]
A[0]
W [j]
+j +h

= 0 +
=
j
0
0

0
.
j

(15.110)

Since 0 is not equal to cl , the first two terms in the bracket do not cancel. Instead,
we shall determine 0 by the requirement that the whole bracket vanishes for j =
j[0 ]
1 [0 ]
A[0]
+ j[0 ] + h

= 0.
0
0

(15.111)

15.5 Effective Action to all Orders in h

1005

This requirement makes 0 directly equal to and


[] = A[] + h
1 [].

(15.112)

Since 1 [0 ] depends on i also via j[0 ], equation (15.111) looks impossible to solve.
In fact, reinserting (15.111) into (15.109) we can find a pure equation for 0 :
i1 [0 ]

G e h {A[0 + ]A[0 ] A0 [0 ]h 10 [0 ] }
i

(15.113)

which is a functional integral-different equation. Still it is possible to extract from


(15.113) the desired information. Let us consider an auxiliary generating functional
eiWaux [0 ,K]

G e h {Aaux [0 , ]+
i

d4 xK}

(15.114)

with an action
Aaux [0 , ] = A[0 + ] A[0 ] A0 [0 ].

(15.115)

It defines, for an arbitrary fixed function 0 , an auxiliary field theory with an external
source K. If expanded Waux [K] in powers of K, it collects all connected diagrams
for the action (15.115). They are built up from propagators
= G[0 ]

(15.116)

and vertices

n
1

5
2

n A[0 ]
.
0 . . . 0

(15.117)

4
3

Thus, if we succeed in enforcing 0 = , we are dealing exactly with the field


theory whose one-particle irreducible vacuum graphs are supposed to make up the
corrections 1 []. How can we show this? By comparing (15.113) with (15.109) we
see that for the special choice of the current
K K0 h
10 [0 ]

(15.118)

1006

15 Path Integral Calculation of Effective Action. Loop Expansion

the auxiliary functional Waux [0 , K0 ] becomes identical to the generating functional


we are trying to calculate. The auxiliary functional also has an auxiliary Legendre
transformed field
aux =

Waux [0 , K]
.
K

(15.119)

Thus, if we succeed in enforcing = , we are dealing exactly with the field theory
whose one-particle irreducible vacuum graphs are supposed to make up the corrections 1 []. Indeed, by comparing (15.113) with (15.109) we see that for the special
choice of the current
K K0 h
10 [0 ],

(15.120)

the auxiliary functional Waux [0 , K0 ] becomes identical to the generating functional


we are trying to calculate.
With the help of
aux =

Waux [0 , K]
= aux [0 , K]
K

(15.121)

we now define an auxiliary effective action


aux [aux ] Waux [0 , K]

d4 xKaux ,

(15.122)

where aux [aux ] is also a functional of 0 . There is one important property of


aux [aux ]. If it is evaluated at aux = 0, it collects all one-particle irreducible
vacuum graphs of Aaux [see (11.45), (11.46)]. But these are exactly the graphs we
want. Hence the proof can be completed by showing that the condition K = K0 =
10 [0 ] leads to
aux = 0.

(15.123)

But this is trivial to see: The condition K = K0 was just made up to enforce = 0 .
The fluctuating fields are the difference between and 0 :
= 0 = .

(15.124)

Now hi implies hi = 0. But in the auxiliary theory,


aux hi.

(15.125)

Thus aux = 0, and Waux [0 , K] collects the one-particle irreducible vacuum graphs
with propagators and vertices (15.120), (15.121). This completes the proof.

Notes and References

Notes and References

1007

[1] In this chapter we follow the textbooks


C. Itzykson and J.-B. Zuber, Quantum Field Theory, McGraw-Hill (1985);
H. Kleinert, Path Integrals in Quantum Mechanics, Statistics, Polymer Physics, and Financial Markets, World Scientific, Singapore 2009
http://www.physik.fu-berlin.de/~kleinert/b5).
The technique goes back to
C. De Dominicis, Jour. Math. Phys. 3 983 (1962).
See also
J.M. Cornwall, R. Jackiw, E. Tomboulis, Phys. Rev. D 10, 2428 (1974).

If God lived on earth, people would break his windows


Yiddish Proverb

17
Spontaneous Symmetry Breakdown
In the perturbation expansion developed in Chapter 10 the classical ground state
carries no field. The associated vacuum state is called normal since it is invariant
under all symmetry transformations of the action. There are many physical systems
where this is not true. As an example, we shall discuss here the 4 -theory with N
components 1 , . . . , N coupled in an O(N)-symmetric way.

17.1

Scalar O(N)-Symmetric 4 -Theory

The Lagrangian
L=

m2 2 g  2 2
1
.
(a )2

2
2 a 4 a

(17.1)

This is invariant under all N-dimensional rotations of its components


a Rab b ,
which conserve the length of the field vector
|| =

(17.2)

 (a),

2a .

(17.3)

The perturbation expansion of the 4 -theory in Chapters 10 and 11, was based on
a Taylor expansion of the total generating functional
i(g/4)

Z[j] = e

dt[i/j(x)]4

Z0 [j]

(17.4)

in powers of the interaction g, where Z0 [j] is the generating functional of the freefield theory with a Lagrangian
L0 =

1
m2 2
(a )2
.
2
2 a

(17.5)

This expansion makes sense only for a positive square mass m2 . Assuming a positive
coupling constant g, the classical field potential
v(a ) =

m2 2 g  2 2

+
2 a 4 a
1042

(17.6)

17.1 Scalar O(N )-Symmetric 4 -Theory

1043

has a minimum at the O(N)-symmetric field configuration a = 0a = 0. Only in this


case can the perturbation expansion of the second-quantized field theory, proceed
around the free-field theory described by L0 ,
A non-trivial and physically very important situation is, however, described by
the case m2 < 0 (at g > 0). Then the classical field configuration minimizing L is
given by a nonzero field
0a 6= 0.
(17.7)

This happens in various many-body systems, where field theories of the type (17.6)
are used to describe orientational properties of materials, for instance the direction
of magnetization. In that case n = 3, and the vector (1 , 2 , 3) is identified with
the local size and orientation of the local magnetization. The generating functional
Z[j] =

Dei

dx[L(,)+ja a ]

(17.8)

can serve to understand the fluctuation phenomena in such systems. When comparing theory and experiment it turns out that there is usually a certain temperature
Tc , above which the system can be described via the perturbation expansion given in
the last chapter. As the temperature is lowered toward Tc , flucuations increase drastically. This can be accounted for in the field theory by letting the mass parameter
be temperature dependent such that m2 vanishes for some temperature T = TcMF,
say
2

m =

T
1 2
MF
Tc

(17.9)

where is some mass scale. In the bare mass, the temperature carries a superscript
MF to indicate that it is the places where the bare mass changes sign, and the
mean-field approximation has its phase transition. The fluctuations will shift this
temperature downwards, making the true critical temperature Tc smaller than Tc .
In the low-temperature phase, the directon of the spin vectors a is no more random
but the system settles at a specific non-zero expectation value of the field which we
denoted before as
a h0|a|0i.

(17.10)

For example, magnetic systems acquire a certain spontaneous magnetization. Eq.


(17.10) introduces a new qualitative feature into the theory. The ground state no
longer has the symmetry which was contained in the action. A rotation O(N)
changes the direction of the ground state just as a magnet can be rotated in space.
One speaks of a spontaneous breakdown of the original symmetry of the action.
The perturbative calculation order by order cannot account for this feature since,
as we have seen, the evenness of the 4 interaction prevents the existence of any oddpoint function. The effect must therefore be due to a non-perturbative mechanism.
It is gratifying to note that this does not prevent us from giving a theoretical
explanation of this phenomenon using perturbation-theoretic techniques. In fact,

1044

17 Spontaneous Symmetry Breakdow

the effective action calculated in the proceeding chapters from such techniques is
the ideal tool for this purpose. First of all we notice that whatever symmetry is
carried by the action A[a ], this will also wind up in the effective action [a ]. For
if A[a ] does not change under the transformation (17.2), then W [ja ] is invariant
under ja Rab jb . Thus a W/ja transforms according to the same law as a
and [a ] = W [ja ] ja a must also be invariant.
This statement is obvious at the lowest-order approximation to the effective
action which is equal to the classical action A [a ] taken at a :
[a ] A [a ] .

(17.11)

At a constant field expectation a , this determines the effective potential


[a ] = V v(a ),

(17.12)

where V = dD x is the spacetime volume. For the Lagange density (17.6), the
effective potential is
1
g  2 2
.
v (a ) = m2 2a +
2
4 a

(17.13)

Already at this lowest approximation, the effective action is capable of describing a


spontaneous symmetry breakdown. As long as m2 > 0, this has an O(N)-symmetric
minimum at a single field point, the origin a = 0. But if m2 is lowered and becomes
smaller than zero, the origin becomes unstable and the potential looks as shown in
(17.1). The O(N)-symmetry of the potential gives rise to an infinitely degenerate

Figure 17.1 Effective potential of the 4 -theory for N = 2 in mean-field approxmation.


It is plotted for m2 > 0 and m2 < 0, corresponding to temperatures above and below the
critical temperature, respectively.

17.1 Scalar O(N )-Symmetric 4 -Theory

1045

minimum characterized by a fixed length of the field vector a :


|0 |

0a 2 =

m2
.
g

(17.14)

The direction of this minimum can be changed arbitrarily by an O(N)-rotation.


without changing the energy.
The length scale over which the fluctuations take place is given by the mass term
as
= m1 = 1 1/2 .

(17.15)

Experimentally one finds the slightly different result


= 1 ,

(17.16)

and the power is the critical exponent or or critical index of the temperature
dependence of the coherence length. In mean-field approximation, it is equal to 1/2.
The system is unstable at the origin. Small fluctuations which are neglected within
the present classical approximation will drive the field down to a new minimum at

0 . The non-perturbative aspect of the solution manifests itself in the power 1/ g


appearing even at the level of no fluctuations.
Notice how the symmetry is spontaneously broken: The system has now many
directions to choose from since there is the whole surface of an O(N) sphere (O(N)
shell) of equal |0 | which has been pictured in (17.1) as a circle. One says that for
m2 < 0 the vacuum is infinitely degenerate. As soon as a has made its choice, the
ground state has lost its symmetry.
From (17.14) we can calculate how the strength of the magnetization varies with
temperature using (17.9):
s

T
|0 | = = 1 MF .
g
g
Tc
Experimentally, one finds a power law slightly different from this
|0 | ( ) ,

(17.17)

and the power is the critical exponent or critical index of the temperatur behavior of the magnetic field, that lies experimentally quite close to the mean-field
approximation shown is plotted in Fig. 17.2.
The deviation of and from 1/2 is due to fluctuations. As a matter of fact,
in Section 21.8 we shall derive that these corrections. It will be obtained from an
expansion of a general scaling relation to be derived in Eq. (21.177)]. Anticipating
the result for v(a ), we write
v() =


g 
1 + const 2/(D2+) + . . . ,
4

(17.18)

1046

17 Spontaneous Symmetry Breakdow

0
0 |T =0

T
Tc

Figure 17.2 Magnetization 0 in mean-field approximation as a function of the temperature ratio T /TcMF in mean-field approximation.

where for N = 2, = (D + 2 )/(D 2 + ) 4.7, 0.033, 2/3, 0.8.


This expression is valid for positive m2 . What can we say about m2 < 0? Looking at
the original one-loop formula we see that the whole effective action can be continued
analytically to negative m2 as long 2 is kept nonzero, so that
g 2
m2 + 0 > 0.
2

(17.19)

But this is too close to the new minimum which with the normalization of the
coupling used there (gold = 3!gnew ) was at
2

0 =

m2
3!
g

(17.20)

such that m2 + g2 /2 = 2m2 > 0. Moreover, the full loop expansion formula
to all orders can be calculated for close to this minimum since all propagators
D = i/ ( 2 m2 g2 /2) correspond to particles of proper positive mass term.
Also: The renormalization program needs no modification. The same counter terms
which made the results finite above TcMF keep the results finite below TcMF. We only
have to forbid to drop.
The only point to watch out is that the never drops below a certain critical
value
m2
1
2 < c2 = 02 = 2
3
g

(17.21)

at which place logarithms pick up negative arguments and give rise to imaginary
parts. Thus, as long as we avoid the critical value (17.21) we may use (17.18) also

17.1 Scalar O(N )-Symmetric 4 -Theory

1047

for m2 < 0, T < TcMF. Actually, that result was derived only for a single field.
But it can easily be shown that the same form remains also for the O(N) field a
except with slightly different numbers for the powers , .
We may now use (17.18) for the same discussion as carried through previously
and find now a vacuum expectation below TcMF of
m2
|0 | =
g

= ( ) .

(17.22)

By comparing this one-loop corrected theoretical result with experiment one does
find an improvement in the right direction.
The theory of critical indices has been the subject of many investigations in the
past years. Therefore they deserve a more detailed discussion which will be given in
a separate chapter. Let us end this section by pointing out that the appearance of
an imaginary part for 2 < 2c has a good physical reason: Imaginary parts in an
energy are always related to an instability of the system. The origin of the instability
for 2 < 2c can easily be found: Let us turn on an external field j which drives
into the region of instability. The effective potential becomes
v() =

m2 2 g 4
+ j
2
4!

(17.23)

and looks as plotted in (17.1).


The value of at the minimum increases with j. The other minimum at the
diametrally opposite place, however, begins to move closer to the axis. We can
display the two minima as a function of j by plotting the derivative of v(), (see
(17.3)
j = m2 +

g 3
.
3!

(17.24)

For very large j or j, there is only a single minimum on the right or left-hand
side, respectivly. In a region |j| jc there are three solutions, two minima and a
maximum. The value jc is determined from the maximum of the curve (17.24)
s

jc =

2m2 2 2
m
g 3

(17.25)

which lies at
c2 = 2m2 /g.

(17.26)

There the minimum fuses with the maximum and disappears. But this is precisely
the place at which the effective potential develops an imaginary part. Thus the
instability indicated by the appearance of an imaginary part in the effective potential
is simply a signal for the fact that the local minimum at the elevated place can no

1048

17 Spontaneous Symmetry Breakdow

Figure 17.3 Magnetization 0 as a function of external source j in mean-field approximation.

longer be a solution of the field equation. The point becomes locally unstable.
Notice that the same instability point appears to all orders in the loop expansion,
i. e. to all orders in the fluctuation expansion.
From our knowledge of simple quantum mechanical systems we may conclude
that this result cannot contain the whole truth about the instability. While the system certainly can be in the higher minimum, it alsways has a possibility of escaping
through the potential barrier via tunneling. Thus there must be an imaginary part
in the true effective potential as soon as || < |0 |. Obviously, the loop expansion
cannot account for this. Even the hope that a full summation of all terms might
produce a corresponding singularity at = 0 is futile. From quantum mechanics
it is known that tunneling amplitudes behave as ea/gh . But such an expression can
never be recovered from a loop expansion in a power series in g. (The power series
expansion of ea/gh would be 1 + 0 + 0 + . . .). Methods to describe physical effects
of this type will be developed in a later chapter.

17.2
17.2.1

Nambu-Goldstone Particles
The Mechanism

There is an important physical consequence of spontaneous symmetry breakdown.


The system has an infinite choice between different directions of . Consider a field
configuration which is not constant in space but changes smoothly from space-time

1049

17.2 Nambu-Goldstone Particles

point to space-time point along the degenerate O(N) shell. Obviously, the energy
of such a configuration can be made arbitrarily small. But this means that there is
a large region in field space over which fluctuations hardly change the action. Thus
the suppression of fluctuation cannot be very effective and the angular fluctuations
of the field are large. Let us see this explicitly for the case N = 2 for which may
be written as a single complex field = 1 + i2 . Going to a radial decomposition
= ei

(17.27)

1
m2 2 1 2
g
2
A[] = ()
+ ()2 4 .
2
2
2
4

(17.28)

we may write the action as

For m2 < 0, consider variations of around 0 =


we may expand = 0 + and write

m2 /g. If (x) is kept constant

2m2
1
2
()2 .
A[] = () +
2
2
2

(17.29)

Notice that the size fluctuations take place with a mass twice the opposite of the
T > TcMF case.
An O(2)-symmetric field theory with 4 -interaction can be used to describe the
phase transition from normal to superfluid in liquid helium near 2.7 degrees Kelvin.
The O(2) degeneracy of the ground state manifests itself in the independence of
the action on as long as it is constant. If has small space time variations we may
write = 0 + and get an additional piece in the action (to quadratic order)
1
2 A[] = 0 2 ()2 .
2

(17.30)

Consider now the eigenmodes of quadratic fluctuations. Previously, for m2 > 0, we


had two (in general N) particles both (all) of mass m. Now we have one field
with propagator
G =

i
,
+ 2m2

(17.31)

and another field with propagator


G =

1 i
.
0 2 2

(17.32)

This is a massless particle called Nambu-Goldstone boson whose presence in the


case of spontaneously broken symmetry was first discovered by Nambu, followed by
a more rigorous treatment by Goldstone.
Remember that the physical origin is the arbitrarily small change of action for
smooth continuous variations of the ground state directions.

1050

17 Spontaneous Symmetry Breakdow

At long distances, the only experimantal correlations come from the massless
modes. There exist many physical phenomena which are dominated by these. They
can be studied approximately by ignoring completely the variations of the size of
the field , freezing it at the equilibrium value 0 . This is called the hydrodynamic
limit of the field theory. It describes only the Nambu-Goldstone boson. In the
O(N)-symmetric theory, the freezing of the size of the field leaves the directional
vector na = a /|| free to fluctuate. Then the field theory model goes over into the
classical Heisenberg model.
This limit will be discussed further in Section 20.2.

17.2.2

General Considerations

The necessity of a Nambu-Goldstone particle as a consequence of spontaneous symmetry breakdown can easily be proved for any continous symmetry and to all orders
in perturbation theory by using the full effective action. Let us look only at the case
of O(N) symmetry. The infinitesimal symmetry transformation on the currents are
ja ja icd (Lcd )ab jb

(17.33)

where Lcd are the N(N 1)/2 generators of O(N) rotations. Under this, the generating functional is invariant
W [j] =

dx

W [j]
i (Lcd )ab jb cd = 0.
ja (x)

(17.34)

Let us insert the Legendre tranformation laws for and j. Then (17.34) amounts
to
0=

dxa (x)i (Lcd )ab

[]
cd
b (x)

(17.35)

which expresses the infinitesimal invariance of the effective action under


a a icd (Lcd )ab b .
The invariance law (17.35) is often called the Ward-Takahashi identity for the functional []. It can be used to find an infinite set of identities with this name for all
the vertex functions by forming higher functional derivatives and setting = 0 at
the end. For example the first derivative of (17.35) gives (dropping the infinitesimal
parameter cd )
0 = (Lcd )ab

Z
[]
2 []
+ dx a (x ) (Lcd )a b
.
b (x)
b (x )a (x)

Setting = 0 where ja = []/b (x) = 0 gives


Z

dx 0a (x) (Lcd )a b

2 [0 ]
0b (x )0a (x)

(17.36)

1051

17.2 Nambu-Goldstone Particles

At = 0 , the second derivative of [0 ] is simply the vertex function (2) at zero


momentum which in turn, by (14.45)), equals the inverse of the connected two-point
function in momentum space G(p). Hence we obtain
0a (Lcd )a b [G1
c ]ba (q = 0) = 0.

(17.37)

This statement implies that certain fully interacting propagators have to contain
a singularity at q = 0. In most cases this caused by a pole in the variable q 2 .
Then (17.37) implies the existence of a massless particle in the Hilbert space, the
Nambu-Goldstone boson. We shall see later what exceptions there can be.
We have argued before that the directions of massless fluctuations must be associated with the valley of degeneracy in the effective potential. Eq. (17.37) is
a mathematical expression of this observation. As an example consider the case
N = 2. Then (Lcd cd )ab is simply the 2 2-matrix
0 1
1
0

12

Let the vacuum expectation of the two-component field a (x) be




0 = 0 , 0 .
Then (17.37) states that
[G1
c ]22 (q = 0) = 0,

(17.38)

calling for a massless mode along the second field direction. Translating this into
the complex field language studied previously, we rewrite a = ( cos , sin ) as
= ei .
For close to 0 and small , this is precisely the massless mode in the azimuthal
variable. For N = 3, the rotation may be written as
(Lcd cd )ab = dab d ,
where is a small angle of rotation and 123 = 1. Then (17.37) becomes
abd 0a [G1
c ]bc (q = 0) = 0.
(2)

(17.39)
(2)

For 0a pointing in 3-direction, the modes [Gc ]11 and [Gc ]22 are massless.
Note that only statements on diagonal elements are non-trivial since [G(2)
c ]ij
vanishes unless i = j, as is obvious at the one-loop level and can be shown to hold
n
for any invariant power (2a ) appearing in the effective action.
The Goldstone mode has an exception which has turned out to provide the basis
of present theories of weak interactions. It therefore deserves being discussed in
detail in the next section.

1052

17.2.3

17 Spontaneous Symmetry Breakdow

Experimental Consequences

The Nambu-Goldstone phenomenon has important observational consequences. The


most simple experiment which can be done on a system is to study the change of
the energy as a function of an external field j. The object
W [j]
= h0|(x)|0ij=const =
j

(17.40)

is observable directly as an order parameter of the system (say magnetization). The


change of this order parameter with respect to changes in j
(x)
2 W [j]
=
= h0|T (x)(y)|0i = (x, y)
j(y)
j(x)j(y)
is measured as a susceptibility. Its Fourier transform
(2)4 (4) (q + q )(q) =

d4 x

d4 x ei(qx+q x ) (x, x )

(17.41)

at q0 = 0 is called the static susceptibility. Using our identities of Ch. 6, we may


write equivalently
(2)4 (4) (q q )1 (q) =

2 []
,
a (q)b (q )

(17.42)

or
1 (0) =

2v
.
a b

(17.43)

Let us calculate these quantities at the mean-field level. For > 0 we find
1
ab (q) =

m2 + q2 q02 ab = 2 + q2 q02 ab

2
2
1
ab (0) = m ab = ab .

(17.44)

Thus 1 increases linearly with and is a favorite quantity for experimental plots.
Very close to a critical point, the measurements show deviations from this law due
to fluctuations. Near T = TcMF, one measures a power law parametrized by
2
1
ab = ab .

The value of is called the the critical index of susceptibility above TcMF . The
derivation of this power behavior will be given in Subsec. 21.9.1 [see Eq. 21.155)].
As the temperature passes below TcMF, the spontaneous breakdown of O(N)
symmetry leads to a drastic change: The susceptibility ab (q) looses its isotropy in
filed space. The component longitudinally to the direction 0a behaves like
2
2
2
2
2
2
1
11 (q) = 2m + q q0 = 2 + q q0
2
2
< 0.
1
11 (0) = 2m = 2

<0
(17.45)

1053

17.3 Domain Walls in O(1)-Symmetric Theory

Then N transverse components of 1


ab , on the other hand, vanish at q = 0 (see
(17.37)) since
1
2
1
(q) = G (q) = q q0 .

(17.46)

Thus diverges for q 0 and this leads to a dramatic experimental effect: If


the external source describes incoming waves with weak space-time dependence, the
system scatters these very strongly. For example, if j corresponds to an electromagnetic field in the optical range (as is the case in nematic liquid crystals) the
system becomes completely opaque (it looks milky). When cooling through TcMF
this phenomenon is called critical opalescence.
Notice that a spontaneously broken continuous symmetry is necessary for the
opalescence below TcMF . In the case of a single real field (Ising like system)
there are no Goldstone modes and there is no 1
which stays identically zero for
T < TcMF, q = 0. There, one sees critical opalescence just in the immediate vicinity
of TcMF and the system becomes again clear below TcMF.

17.3

Domain Walls in O(1)-Symmetric Theory

The field equations following from the Lagrangian (17.6) are


h

2 + m2 + g2b (x) a (x) = 0.

(17.47)

For m2 < 0, where the potential looks like the bottom of a champaign bottle in
Fig. 17.1, these equations possess nontrivial static solutions solving the spatial part
of the differential equation (17.47):
h

2 + m2 + g2b (x) a (x) = 0.

(17.48)

The simplest nontrivial solution varies only along one axis, say the x-axis, and only
in one field component. It will turn out to be a stable solution only if the ather
components are absent altogether. There the Lagrangian (17.6) possesses merely a
repfection symmetry , which in the general context is called O(1)-symmetry.
Thus we shall study the simplified field equation
"

d2
2 + m2 + g2(x) (x) = 0.
dx
#

(17.49)

For a single field component, the potential has the form of a double-well pictured in
Fig. 17.4. The equation (17.49) has two trivial constant solutions
(x) 0

m2
g

(17.50)

which lie at the two minima of the potential


V () =

m2
m2 2 g 4
V (x) =
+ .
2
2
4

(17.51)

1054

17 Spontaneous Symmetry Breakdow


V ()

Figure 17.4 Plot of the symmetric double,-well potential V (x) = 21 2 + 41 4 . We have


added a trivial constant 1/4 to place the minima on the -axis.

At the minimum, the potential has the value


Vc =

m4
,
g

(17.52)

which is called the condensation energy of the field system. It will be convenient to
subtract this from V () and write the subtracted potential
V () =

m2
( 0 )2 ( + 0 )2 ,
2
80

(17.53)

exhibiting more directly the symmetric minima at = 0 (see Fig. 17.4). The
coupling strength is
g = m2 /220 .
(17.54)
Near the minima, the potential looks approximately like a harmonic oscillator potential V (0 ) = m2 (x 0 )2 /2:
m2
0
V () =
( 0 )2 1
+ . . . V () + V () + . . . .
2
0
!

(17.55)

The height of the potential barrier at the center is


Vmax

(m0 )2
.
=
8

(17.56)

In the limit 0 at a fixed frequency m, the barrier height becomes infinite and
the system decomposes into a sum of two independent harmonic-oscillator potentials
widely separated from each other.
The nontrivial solution of the field equation connects the two trivial solutions
(17.50) across the barrier. As we shall derive immediately, this reads
(x) =
cl (x) 0 tanh[m(x x0 )/2],

(17.57)

1055

17.3 Domain Walls in O(1)-Symmetric Theory

10 V ()
Figure 17.5 Classical kink solution (solid curve) in double-well potential (short-dashed
curve with units marked on the lower half of the vertical axis). The solution connects the
two degenerate maxima in the reversed potential. The long-dashed curve shows a solution
which starts out at a maximum and slides down into the adjacent abyss.

with an arbitrary parameter x0 specifying the point on the x-axis where the crossing
takes place. The crossing takes place over a distance of the order of 2/m. For large
positive and negative x, the solution approaches 0 exponentially (see Fig. 17.5).
The subscript cl on cl emphasized that these solutions solve the classical field
equations. In allon to their shape, the solutions
cl (x) are called kink and antikink
solutions, respectively.1
To derive these solutions, consider the equation of motion (17.49) in the pseudotime time t = ix,
= V ((t)),
(t)
(17.58)
where V () dV ()/d). Since the differential equation is of second order, there is
merely a sign change in front of the potential with respect to thr spatial differential
equation (17.49). The equation of motion (17.58) corresponds therefore to a usual
equation of motion of a point particle as a function of the pseudotime t, whose
potential is turned upside down with respect to Fig. 17.4. This is illustrated in
Fig. 17.6. The reversed potential allows obviously for a classical solution which
starts out at = 0 for x and arrives at = 0 for x +. The particle
needs an infinite time to leave the initial potential mountain and to climb up to the
top of the final one. The movement through the central valley proceeds within the
finite time 2/m. If the particle does not start its movement exactly at the top but
slightly displaced towards the valley, say at = a + , it will reach = a after
a finite time, then return to = a + , and oscillate back and forth forever. In
1

In field-theoretic literature, such solutions are also referred to as instanton or anti-instanton


solutions, since the valley is crossed within a short time interval. See the references quoted at the
end of the chapter.

1056

17 Spontaneous Symmetry Breakdow

V ()

Figure 17.6 Reversed double-well potential governing the motion of the position as a
function of the imaginary time x.

the limit 0, the period of oscillation goes to infinity and only a single crossing
of the valley remains.
To calculate this movement, the differential equation (17.47) is integrated once
after multiplying it by = dx/dx and rewriting it as
1 d 2
d
x =
V ((x)).
2 dx
dx

(17.59)

The integration gives

x2
+ [V ((x))] = const .
(17.60)
2
If x is reinterpreted as the physical time, this is the law of energy conservation
for the motion in the reversed potential V (). Thus we identify the integration
constant in (17.60) as the total energy E in the reversed potential:
const E.

(17.61)

Integrating (17.60) further gives


1
x x0 =
2

(x)

(x0 )

E + V ()

(17.62)

A look at the potential in Fig. 17.6 shows that an orbit starting out with the particle
at rest for x must have E = 0. Inserting the explicit potential (17.53) into
(17.62), we obtain for || < 0
x x0

d
1
0 +
2phi0 Z
= log
=

m 0 (0 )( + 0 )
m 0

2
= artanh .
m
0

Thus we find the kink and antikink solution (17.57) crossing the barrier.

(17.63)

1057

17.3 Domain Walls in O(1)-Symmetric Theory

The euclidean action of such a solution can be calculated as follows [using (17.60)
and (17.61)]:
Acl =

Z
2
dx cl + V (cl (x)) =
dx(2cl E)
2

"

= EL +

dx 2[E + V ()].

(17.64)

The kink has E = 0, so that


q

2(E + V ()) =

m 2
( 2 ),
20 0

and the classical action becomes


Z
m3
2
m 0
.
d(20 2 ) = 20 m =
Acl =
20 0
3
3g

(17.65)

(17.66)

Note that for E = 0, the classical action is also given by the integral
cl 2
Acl =
dx
.
(17.67)
2

There are also solutions starting out at the top of either mountain and sliding
down into the adjacent exterior abyss, for instance (see again Fig. 17.6)
Z

x x0

20 Z
d
1
+ 0
=
= log

m ( 0 )( + 0 )
m 0

2
= arcoth .
m
0

(17.68)

However, these solutions cannot connect the bottoms of the double well with each
other and will not be considered further.
The real 4 -theory can be used as a model for a spin system, in which the
spins can oint only in two directions. In statistical mechanics, such a system is
usually described by the Ising model on a lattice. In fact, it can be shown that the
Ising model can be transformed into a real 4 -field theory whose potential can be
approximated, near the phase transition, by a 4 -potential.2 The two spin directions
are represented in the 4 -theory by the two minima of the potential. The classical
solution to the field equation may be interpreted as a domain wall between up and
down directions. Domain walls can occur in many possible shapes which are tow
hard to calculate analytically. They possess a high configurational entropy, and the
demagnetization transition can be understood as a transition in the ensemble of
fluctuating domain walls of any shape, whose configurational entropy overcomes the
energy it takes to create the domain walls.
Let us finally relaize that domain walls in O(N)-symmetric theories cannot be
stable: It is always possible to rotate the domain with one spin direction into the
other lowering continuously the energy. This is impossible for a single field .
2

For details see the textbook H. Kleinert, Gauge Fields in Condensed Matter , World Scientific,
1989 (http://www.physik.fu-berlin.de/kleinert/kleiner re.html#b1).

1058

17.4

17 Spontaneous Symmetry Breakdow

Vortex Lines in O(2)-Symmetric Theory

The most important domain of application of the O(2)-symmetric theory is the


physics of the phase transition in superfluid helium at Tc 2.3 K. The twocomponent 4 theory is then referred to as Landau-Pitajevski
theory. Its euclidean
action is written in terms of a complex field = (1 +i2 )/ 2 in 3 space dimensions
as an energy density
h
2 2
H = (x)
(x) + V |(x)|4 ,
2M
!

(17.69)

where is the chemical potential which pases through zero at some critical meanfield temperature
!
T
= 0
1 .
(17.70)
TcMF
The quartic term approximates the repulsion between the helium atoms by a function potential [recall the second-quantized Hamiltonian (2.138)]
The field equation which minimizes this energy density is
h
2 2
+ V ||2 = 0.

2M
!

(17.71)

It possesses cylindrical solutions which are observable in the laboratory in the form
of straight vortex lines. To find these, we decompose into its polar components,
= ei , and find from the real and imaginary parts of (17.71) the two equations
(

i
h
2 h 2

()2 + V 2 = 0,
2M
)

(17.72)

and
(2 ) = 0.

(17.73)

The latter equation is simply the radial-azimuthal form of the current conservation
law for the particles


1
j =
= 0.
(17.74)
2i
Equation (17.73) can be solved by a purely circular flow of particles, in which depends on the spatial distance r from the cylindrical axis, and the phase of the complex field is an integer multiple of the azimuthal angle in space, tan1 (x2 /x1 ),
i.e.,
= n.
(17.75)
Then, (17.72) reduces to the radial differential equation
h
2

2M

r2

1 n2
+
+ V (2 20 ) = 0,
r
2
!

(17.76)

1059

17.4 Vortex Lines in O(2)-Symmetric Theory


q

where 0 = |0 | = 0 (T /TcMF 1)/V [compare (17.70)].


In order to solve Eq. (17.76), it is convenient to go to reduced quantities r,
which measure the distance r in units of the coherence length
=

2size =

v
u
u
t

h
2
1
2M (0 )

T
1
TcMF

(17.77)

and the size of the order parameter in units of 0 . Thus we introduce


x x/,

r r/,

(17.78)

in configuration space, and


= /0

(17.79)

in field space. Then (17.76) takes the form


"

r2

n2
1
+ r 2
r
r

+ (
1) (
r) = 0.

(17.80)

For small r 1, this is dominated by the terms in the first parentheses, leading
to a differential equation of the Bessel type for (
r ). Thus, close to the origin, the
solution is
(
r ) = An Jn (
r ) rn ,
(17.81)
where Jn (
r) is the standard Bessel function. Multiplying this with the phase factor
ein = ein tan

1 (x

2 /x1 )

(17.82)

we see that the complex field (x) has the following small |x| behavior
(x) rn ein tan

1 (x

2 /x1 )

= (x1 + ix2 )n ,

(17.83)

exhibiting a zero of n-th order in the complex plane x1 + ix2 .


For large r 1, (
r ) approaches the asymptotic value = 1. In fact, from (1.44)
we can extract the 1/
r expansion,
n (
r) = 1

n2
1
1
1
n
1
n2 + n4 4 8 + 2n2 + n4 6 O 8 . (17.84)
2
2
r
8
r
16
r
r


Integrating the differential equation numerically inward, we find the solution


displayed Fig. 17.7. What is the energy of these vortex lines? In order to find a
convenient expression for it we use the fact that if (x) is a solution of the differential
equation (17.71), the rescaled solution
(x) e (x)

(17.85)

must extremize the energy inn for = 0. In terms of the reduced quantities of
Eqs. (17.78) and (17.79), the energy is given by
E = Ec

2 ||
2 + ||
4 ,
d3 x ||

(17.86)

1060

17 Spontaneous Symmetry Breakdow

Figure 17.7 The order parameter = ||/|0 | around a vortex line of strength n =
1, 2, 3, . . . as a function of the reduced distance r = r/ where r is the distance from the
axis and the healing length.

where

2
(17.87)
4V
is the condensation energy. Inserting the reduced version of (17.85) into (17.86), we
obtain
Ec =

E = Ec

d3 x e2 ||
2 e2 ||
2 + e4 ||
4 .

(17.88)

Setting the derivative with respect to equal to zero gives at = 0


Ec

d3 x ||
2 ||
2 + 2||
4 = 0.

(17.89)

Subtracting this from (17.88) for = 0 we find that the energy of a solution of the
field equation is simply given by
E = Ec

d3 x||
4.

(17.90)

Most of this energy is due to the asymptotic value ||


1, where E is equal to the
negative condensation energy Ec . Subtracting this from E we find the additional
energy of the vortex line as
Ev = Ec

d3 x(1 ||
4 ).

(17.91)

Going over to cylindrical coordinates r, , z, the integral becomes, for a vortex


line of length L along the z-direction,

Ev = Ec 2 L

d
rr 1 4 (
r) .

(17.92)

Before inserting the numerical solutions for (


r) shown in Fig. (17.7), we note that
due to the factor r, the additional energy comes mainly from the large r regime, i.e.,

1061

17.4 Vortex Lines in O(2)-Symmetric Theory

from the far zone. In fact, if we insert the leading asymptotic behavior (17.84), we
obtain an integral

Ev = Ec 4n2 L

d
r
,
r

(17.93)

which diverges logarithmically for large r. An immediate conclusion is that a single


vortex line can have a finite energy only in a finite container. In a finite cylindrical
container of radius R, the integral no longer diverges at the upper end, and we
obtain
log(R/).
Ev = Ec 4n2 L

(17.94)

Consider now the energy in the small-


r regime, where Eq. (17.81) tells us that (
r)
behaves like rn . Hence, 1 4 1, and the energy of a thin cylindrical section of
radius r grows like r 2 . For increasing r, the rate of growth slows down rapidly, and
settles at the asymptotic rate 4n2 L log(
r/), where is the coherence length.
The proper inclusion of the non-asymptotic behavior gives simply a finite correction
to the asymptotic energy and the energy of a vortex line in a container of radius R
becomes

Ev = Ec 4n2 L[log(R/)
+ c].
(17.95)
This is the same result we would have obtained had we replaced in (17.92) the
integrand r[1 4 (
r )] by the asymptotic form 2n2 /
r, and integrated from the radius
c
r0 = e to R.
The precise numerical evaluation of the differential equation (17.80) and the
integral (17.92) shows that for the lowest vortex line, c has the value
c = 0.385.

(17.96)

Hence the energy of the vortex line becomes


4n2 [log(R/) + 0.385].
Ev = Ec L

(17.97)

The logarithmic divergence of the energy has a simple physical origin. In order to
see this let us calculate the energy once more in another way, using the original
expression (17.86), i.e. without invoking the virial theorem. It reads, for the radial
solutions,

Ev = Ec 4 L

n2
1
d
rr (r)2 + (1 2 )2 + 2 2 .
2
r
"

(17.98)

The first two terms are rapidly convergent. Thus the energy of the far zone
resides completely in the last term
n2 2 n2
2 .
r2
r

(17.99)

1062

17 Spontaneous Symmetry Breakdow

This energy is a consequence of the angular behavior of the condensate phase =


n tan1 (x2 /x1 ) around a vortex line. It comes entirely from the azimuthal part of
the gradient energy (h2 /2M 2 )||2()2 , i.e., the term which describes the NambuGoldstone modes. This is not surprising. We have discussed before that the long
range properties of the system are dominated by these modes. A phase of the field
which behaves like = n tan1 (x2 /x1 ) can be viewed as a coherent pile-up of these
modes determining the energy in the far zone.
The dominance of the energy carried by the phase gradient can also be described
in a different and more physical way. In the last section we identified the gradient
(h/M) with the superfiuid velocity. For the vortex line at hand, the superfiuid
velocity is found, far away from the line, to be
vs =

x2
n tan1
M
x1


h
n
h
1
(x2 , x1 , 0) =
n e,
2
Mr
M r

(17.100)

where e is the unit vector in azimuthal direction. Thus, around every vortex line,
there is a circular fiow of the superfiuid whose velocity decreases like the inverse
distance from the line. The hydrodynamic energy density of this flow is
s 2 s h
2 n2
(x) = vs =
.
2
2 M 2 r2

(17.101)

This is precisely the dominant third Nambu-Goldstone term in the energy integral
(17.98). Thus the energy of the vortex line is indeed mainly due to the hydrodynamic
energy of the superfiow around the line. For the major part of the volume, the
hydrodynamic limit (17.100), (17.101) gives an excellent approximation. Only in
the neighborhood of the line, i.e., for small radii r , the energy density differs
from (17.101) due to gradients in the size of the field ||. It is therefore suggestive to
idealize the superfiuid and assume the validity of the pure gradient energy density
(x) =

h
2
s 2
vs = ()2 , s 2 ,
2
2
M

(17.102)

everywhere in space, where s is the superfluid density. insdensity,superfluid


The deviations from this law, which become significant only very clnse to a
vortex line, i.e., at distances of the order of the coherence length , are treated
approximately by simply cutting off the energy integration at a radius away from
a vortex line. In other words, we pretend as though there were no superflow at all
within the thin tubes of radius , with a sudden onset of idealized flow outside ,
moving with the limiting velocity (17.100).
Although the internal part of the thin tube carries no superfiow, it nevertheless
carries rotational energy. Within the present approximation, this energy is associated with the number 0.385 in (17.97). This piece will be called the core energy.
The core energy has a physical interpretation. At distances smaller than the coherence length, the different parts of the liquid can no longer slip past each other freely.

1063

17.4 Vortex Lines in O(2)-Symmetric Theory

Hence the core of a vortex line is expected to rotate roughly like a solid rod, rather
than with the diverging velocity vs 1/r. Indeed, if we use the approximation
1
, r > ,
vs n rr

, r ,
2

(17.103)

for a line of vortex strength n, the energy density does behave like r 2 , for small r,
as observed before. Moreover, the energy integration gives
2

"Z

R/
1

d
r/
r+

d
rr

= n2 [log(R/) + 0.25],

(17.104)

and we see that the number for the core energy, 0.25, emerges with the right order
of magnitude. We complete our discussion of the hydrodynamic picture, let us
calculate the circulation of the superftuid velocity field around the vortex lines:
h
h

h
I
n dx =
2n = n
= n1 .

dx vs =
M B
M
M
B
I

(17.105)

This integral is the same for any size and shape of the circuit B around the vortex
line. Thus the circulation is quantized and always appears in multiplets of 1 =
h/M 103 cm2 /sec. The number n is called vortex strength .
The integral (17.105) can be transformed into a surface integral, via Stokes
theorem:
Z

SB

dS ( vs )=

2n,
M

(17.106)

where S B is some surface spanned by the circuit B in (17.105), and dS is the surface
element. This integral is the same for any size and shape of S B . From this result
we conclude that the third component of the curl of vs , must vanish everywhere
except at the origin. There it must have a singularity of such a strength that the
two-dimensional integral gives the correct vortex strength. Hence
vs =

2n (2) (xT )z,


M

(17.107)

where z is the unit vector along the z-axis and xT (x1 , x2 ) are the coordinates
orthogonal to the vortex line.
If the nonlinearities of the field equation are taken into account, the (2) -function
is really smeared out over a circle whose radius is of order . As an example, let us
replace 1/r 2 in (17.101) by 1/(r 2 + 2 ). Then the rotation of the superftuid velocity
becomes
vs =

22
n 2
z.
M (r + 2 )2

(17.108)

1064

17 Spontaneous Symmetry Breakdow

The right-hand side is non-zero only within a small radius r 6= , where it diverges
with the total strength
Z

d2 x

Z
22
2r
2
dr 2
=
2
= 2.
2
2
2
(r + )
(r + 2 )2
0

(17.109)

This shows that (17.108) is, indeed, a smeared out version of the -function relation
(17.107).
Because of their rotational properties, vortex lines can be generated experimentally by rotating a vessel with an angular velocity m. Initially, the lack of friction
will cause the superfluid part of the liquid to remain at rest. This situation cannot,
however, persist forever since it is not in a state of thermal equilibrium. After some
time, vortex lines form on the walls which migrate into the liquid and distribute
evenly. This goes on until their total number is such that the rotational Helmholtz
free energy
F = F L

s 2
v x s vs
dx
2 s
3

(17.110)

is minimal. This equilibration process has been observed in the laboratory and has
even been photographed directly using the property that vortex lines trap ions which
can be accelerated against a photographic plate. Integrating (17.110) with vs from
(17.100), we find that in a cylindrical vessel of radius R, the first vortex line n = 1
appears at a critical angular velocity
c =

1
R
log
2
R

(17.111)

and settles on the axis of rotation. It is useful to observe that the vortex lines
of higher n are all unstable. Since the energy increases quadratically with n, it is
favorable for a single line of higher n to decay into n lines of unit strength each.
When generating vortex lines by stirring a vessel, one may nevertheless be able to
create, for a short time, such an unstable line, and to observe its decay.

Notes and References


For more information on vortex lines see the textbook
H. Kleinert, Gauge Fields in Condensed Matter , Vol I, World Scientific, 1989
(http://klnrt.de/b1).
See also:
H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism, and Gravitation, World Scientific, Singapore 2009, pp. 1497 (http://klnrt.de/b11);
H. Kleinert and V. Schulte-Frohlinde, Critical Properties of 4 -Theories, World Scientific, Singapore 2001 (http://klnrt.de/b8).

A complex system that works is invariably found


to have evolved from a simple system that works
John Gall

18
Scalar Quantum Electrodynamics
In nature, there also exist charged scalar particles which are coupled to the electromagnetic field. In three spatial dimensions, a historic example which supplied
many important predictions of experimental data is the Ginzburg-Landau theory of
superconductivity. The generalization of this theory to four spacetime dimensions
is of great importance in elementary particle physics, where it is known as scalar
quantum electrodanymics (scalar QED).

18.1

Action and Generating Functional

The Ginzburg-landau theory is a three-dimensional euclidean quantum field theory


containing a a complex scalar field
(x) = 1 (x) + i2 (x)

(18.1)

coupled to a magnetic vector potential A. The scalar field describes bound states
of pairs of electrons, which arise in a superconductor at low temperatures due to an
attraction coming from elastic forces. The detailed mechanism will not be of interest
here; we only note that the pairs are bound in an S-wave and a spin singlet state
of charge q = 2e. Ignoring for a moment the magnetic interactions, the ensemble
of these bound states may be described in the neighborhood of the superconductive
transition temperature Tc by a complex scalar field theory of the 4 type, with a
euclidean action
AE =

1
m2
g
d x +
+ ( )2 .
2
2
4
3

"

(18.2)

The mass parameter depends on the temperature like m2 (T Tc )/Tc . Above


the transition temperature, the parameter m2 is positive, below it is negative. As
a consequance, the system exhibit a spontaneous symmetry breakdown discussed in
Chapter 27. The symmetry group which is broken is the U(1)-symmetry
(x) ei (x)
1065

(18.3)

1066

18 Scalar Quantum Electrodynamics

under which AE is obviously invariant. Alternatively we may write


1 cos 1 + sin 2
2 sin 1 + cos 2

(18.4)

so that we may equally well speak of a broken O(2)-symmetry.


The action (18.2) is the euclidean version of a relativistic field theory in D = 1+2
dimensions with an action
A=

1
m2
g
d x
( )2 .
2
2
4
D

"

(18.5)

For the sake of generality, we shall discuss this theory for arbitrary D. Most formulas
will be written down explicitly in D = 4 dimensions, to emphasize analogies with
proper QED in Chapter 12.
The phenomenon of spontaneous symmtry breaking in this system has been
studied in detail in Chapter 27. For T > Tc , where the mass term is positive, small
oscillations around = 0 consist of two degenerate modes carried by 1 , 2 , both
of mass m2 . For T < Tc , where the mass is negative, the energy is minimized by a
field with a real non-vanishing vacuum expectation value, say
h(x)i 0 .

(18.6)

Then the symmetry between 1 and 2 is broken and there are two different modes
of small oscillations: one orthogonal to hi which is the massless Nambu-Goldstone
mode, and one parallel to hi which has a positive mass 2m2 > 0.
In order for the euclidean version of (18.5) to describe the phenomena of superconductivity, we must include electromagnetism. According to the minimal substitution rules described in Chapter 12, we simply replace in (18.5) by the covariant
derivative
D ( iqA ) ,
(18.7)
and add the free electromagnetic action1
A =

1
dD x F 2 .
4

(18.8)

The combined action becomes


A=

m2 2 g 4 1
1
|D |2
|| || F 2 .
d x
2
2
4
4
D

(18.9)

This action is invariant under local gauge transformations


(x) eiq(x) (x),
1

A (x) A (x) + (x).

In this chapter we use natural units with h = c = 1.

(18.10)

1067

18.1 Action and Generating Functional

The generating functional is given by the functional integral


Z [, , j] =

DD

phys.

DA ei{A[,

]+iA [, ,A;, ,j]}


s

(18.11)

where
As [, , A; , , j] j A

(18.12)

collects the sources (x), (x) for the complex scalar fields (x) and (x), and
j is the usual electromagnetic current, the latter being a conserved quantity with
R
j (x) = 0. The path integral phys. DA infdicates that a Fadeev-Popov gauge fixing factor is inserted, together with a compensating determinant (recall the general
discussion in Section 13.13):
Z

phys.

DA

DA 1 [A]F [A].

(18.13)

For m2 > 0, we may choose [recall (13.285)]


1

F [A] = F4 [A] = e 2 ( A

)2

(18.14)

and derive from it Feynman rules for perturbation expansions. with a photon propagator (13.305):
G0 (x y) =

dD k ik(xy) i
k k
g

(1

)
. (18.15)
e

(2)D
k2
k2
"

We have seen in Chapter 12 that in QED all scattering amplitudes are independent
of of the gauge parameter .
In addition, the pertubation expansion associated with the action (18.12) involves propagators of scalar particles

ik(xy)

e
and vertices

G0 (x y) =

dD k ik(xy)
i
e
D
2
(2)
k m2

(18.16)

(18.17)
Since the scalar field is complex, the particle carries a charge and the line in the
Feynman diagram an orietation. The coupling of the scalar field to the vector
potential gives not only rise to a three-vertex analog to (12.91) in QED, but also
to a four vertex in which two photons are absorbed or emitted by a scalar particle.
This diagram is commonly referred to as seagull diagram.

1068

18.2

18 Scalar Quantum Electrodynamics

Meissner-Higgs Effect

The situation is drastically different for T < Tc where m2 < 0. At the q


classical level,
the lowest energy state of the pure theory is reached if || = 0 = m2 /g, say
s

= 0 ei0 =

m2 i0
e
.
g

(18.18)

Inserting this into the action (18.12), its photon part becomes
A =

q 2 m2 2
1 2

A .
d x F
4
2 g
!

(18.19)

Extremizing this gives the Euler-Lagrange equation


2 A A =

q 2 m2
A = 0.
g

(18.20)

Note that we could have chosen as well


i0 (x)

= 0 e

m2 i0 (x)
e
,
q

(18.21)

with an arbitrary spacetime-dependent phase 0 (x). Then the photon Lagrangian


would read
A =

1 2
q 2 m2
d x F

(A 0 )2 .
4
2 g
D

"

(18.22)

with Euler Lagrange equations


2 A (A) q 2

m2
(A 0 ) = 0
g

(18.23)

and the minimal energy would be reached for


A (x) = 0 (x).

(18.24)

Of course, the two results are equivalent by a gauge transformation A A +


0 (x) and their physical content is independent on the particular choice of 0 (x).
Consider now the fluctuation properties of this theory. The generating functional has the form (18.11) in which the measure sums over all and physical A
configurations orthogonal to the gauge degrees of freedom. If we decompose into
radial and azimuthal parts
(x) = (x)ei(x) ,

(18.25)

the path integral takes the form

Z [, , j] =

D D

phys

DAeiA[1 A ]+iAs [,

,A;, ,j]

(18.26)

1069

18.2 Meissner-Higgs Effect

But from the structure of the covariant derivative it is obvious that no matter what
the fluctuating (x) configuration is, it can be absorbed into A(x) as a longitudinal degree of freedom. Thus we have the option of considering complex fields (x)
together with physical, transverse, A fluctuations, or real together with transverse and longitudinal A (x)-fluctuations. This may be expressed by rewriting the
functional integral (18.26) as

Z [, , j] =

all

DA eiA[,A ]+iAs [,

,A;, ,j]

(18.27)

with the action


A[, A ] =

q2
m2 2 g 4 1
1
( )2 + 2 A 2
+ F 2 .
d x
2
2
2
4
4
D

"

(18.28)

The formerly irrelevant gauge degree of freedom in A (x) has now become a physical
one, accounting for the phase oscillations of the complex -field.
Let us describe the situation more precisely in terms of a gauge-fixing functional
F [A] of (18.14) and a compensating Fadeev-Popov determinant 1 [A]. expression
(13.287). Whereas all former gauge consitions involved only the A (x)-field itself,
this is no longer true here. Since the transformations of A and are coupled, the
gauge condition may involve both fields (or any other field) of the system. Here
F [A , ] = 0,

(18.29)

to be compensated by a functional [A , ], which also depends on both fields. The


latter is determined by the integral over the gauge volume as in (13.287):
[A , ] =

DF A , .

(18.30)

The above discussion corresponds to choosing a gauge condition which ensures


the reality of the complex scalar field via the functional:
F [A , ] = [2 (x)].

(18.31)

where 2 (x) = Im(x). Since the gauge transformations change


ei(x) = [cos (x) 1 + sin (x) 2 ] + i [ sin (x) 1 + cos (x)2 ] ,(18.32)
we see that
[A , ] =

D [ sin 1 + cos 2] .

(18.33)

The result must be independent of the choice of gauge and can therefore be obtained
at = (||, 0) where we find
[A , ] =

1
.
|(x)|

(18.34)

1070

18 Scalar Quantum Electrodynamics

Note that involves no derivative terms as in the eaarlier gauges. For this reason
it does not require Fadeev-Popov ghost fields. Setting = (x)ei(x) we arrive
precisely at the functional integral (18.28). The Faddev-Popov determinant 1 =
(x) turns out to be responsible for give the integral at each point a form expected
from the radial-azimutzal decomposition = ei and the associated integration
measure dd = d1 d4 = dd.
The generating functional (18.27) may now be used for perturbation expansions
below Tc . Inserting into (18.28)
(x) = 0 + (x) =

m2
+ (x),
g

(18.35)

we can rewrite
A=

d x

"

1
(2m2 ) 2 g 4 3
2
( )
g0 3
2
2
4
2
#
2
q 2 2 2
1
2
2 2

+ A q 0 A 2 + A .
F +
4
2
2

(18.36)

with 2 qm2 /g. From the quadratic part in we can deduce the massive
propagator

(x) (0) =

dq iqx
i
e
.
(2)4
q 2 (2m2 ) + i

(18.37)

the field has cubic 3 and quadratic 4 interactions. The originally massless
Godstone field has disappeared. It has been eaten up by the photon, making it
massive, and giving it one more polarization degree of freedom.
In the action (18.36) there are electromagnetic vertices in which one or two
particle couple to two photons
q 2 0

(18.38)

The most important new feature is found in the propagator of the photon field. If
we the quadratic photon part of the action in momentum space it becomes
A=

1
2

i

h
dk
2
2

A (k)
k

A
(k)
k
k

(2)4

(18.39)

we can immediately form the inverse of




M = k k g k 2 2
= g

k k
k2

(18.40)


k 2 2 + 2

k k
k2

(18.41)

as
1
M

g k k /k 2
1 k k
g k k /2
=
+ 2 2 =
k 2 2
k
k 2 2

(18.42)

1071

18.2 Meissner-Higgs Effect

such that the massive propagator becomes


G (x) =

dk ikx i
k k
e
g 2
4
2
2
(2)
k
k

(18.43)

The photon has acquired a mass and can no longer propagate over a long range. the
dramatic experimental consequence is known as the Meiner effect: If a superconductor is cooled below Tc , the magnetic field lines are expelled and can invade only
a thin surface layer of penetration depth 1/.
At k 2 = 2 there is a massive particle pole. Since at this place, the propagator
G (x) is purely transverse, it satisfies
k G (k) = 0,

(18.44)

and these particles have three internal degrees of freedom corresponding to a massive
particles of unit spin. This is to be contrasted to the massless gauge invariant
theory for T > Tc where gauge invariance permits to choose, say, A0 = 0 while still
satisfying (18.44) at the pole , thereby eliminating one more degree of freedom in the
asymptotic states. Only two transverse photon polarizations remain far away from
the interaction. There is no contradiction in the number of degrees of freedom since
for T > Tc there are two modes carried by the complex fields. The third photon
degree is a consequence of absorbing into the vector field the phase oscillations,
i.e., the former Nambu-Goldstone degrees of freedom. One sometimes describes
this process by saying: The photon has eaten up the would-be Nambu-Goldstone
boson and has grown fat and massive.
Certainly, the same theory can also be calculated with the earlier gauge fixing
Lagnranian
LGF =

1
( A )2 ,
2

(18.45)

in which case the quadratic piece of the action is


A

quad

1
(2m2 ) 2
2
d x
( )

2
2
D

"

1
2
1
A
F 2 +
4
2
q

(18.46)
!2

1
( A )2 .

Recalling the Faddev-Popov determinant 1


4 of (13.297), for this gauge we have to
add the Fadeev-Popov ghost action corresponding ton (13.345):
Aghost =

1
c c
2

to complete the list of fluctuating fields present in this gauge.

(18.47)

1072

18 Scalar Quantum Electrodynamics

The generating functional is obtained by integrating eiA over all A (x), all complex (x), and all ghost configurations c (x), c(x):

Z [, , j] =

DcDc

DD

DAeiA .

(18.48)

What is the particle content of this theory? Since the Fadeev-Popov ghosts do not
interact witgh the other particles, they can be ignored. As far as the photon field is
concerned, let us split the vecter potential a
A = At + Al ,

(18.49)

where At is the transverse part satisfying the Lorentz gauge condition At = 0,


and Al is the longitudinal part Al = 2 A , which is a pure gauge.2 Then the
Lorentz part is diagonal and has the propagator
A t (x)A t (0) =

k k
i
dk
g 2
4
2
2
(2) k
k

(18.50)

The gauge-like pieces, on the other hand, are coupled with the would-be NambuGoldstone mode. In terms of A = k and 1q , we find the quadratic action
k 2 1 k 4 k 2
k2
k2

2
(, )
2

1 k2 u
2 k 2
k2
t
1

k =
1
+

2
2 2
22

!2

(18.51)

which has the determinant 4 /4k 6 and eigenvalues

v
u

(18.52)

which both vanish at k 2 = 0 (and only there). Thus there are two more massless
asymptotic states in the Hilbert space. At first sight, this formulation describes
a physical situation different from the previous one. We do know, however, that
the generating functionals are the same. In the previous formulation, there were
only two types of particles. Therefore all the additional zero mass asymptotic states
found in the present gauge must be such that if a scattering process takes place with
the two physical particles coming in, the unphysical ones can never be produced.
Physical and unphysical states should have no mutual interaction. We may say
that the S-matrix is irreducible on the set of physical states. For the Fadeev-Popov
ghost fields this is trivial to see. With respect to the other states, however, it is not
and the path integral is the only efficient tool for convincing us of the correctness
of this statement as we demonstrated in Section 13.13. The irreducibility of the
S-matrix becomes most transparent by using a particular gauge due to tHooft in
2

As in (13.302), the letters l and t stand for longitudinal and transverse to the four -vector field
A (x).

1073

18.2 Meissner-Higgs Effect

which physical and unphysical parts of the Hilbert space receive a clear distinction
which depends on the gauge parameter. Consider the gauge condition

2
1
F [A, ] = exp
A
2
q

!2

(18.53)

For 0 this enforces the Lorentz gauge. The exponent has the pleasant advantage
that when added to the free part of the Lagrangian, the term which mixes with the
gauge-like component of A1 and which requires diagonalization in (18.51), cancels.
Therefore, barring the calculation of [A, ]1 , gauge-fixed free action reads
A

free

d x

2m2 2 1
2
1
2
( )
F 2 + A2
2
2
4
2
)
2 h
i

1
2
2 2
2
+ 2 ( )
( A ) .
2q
2

(18.54)

Thus the photon has a quadratic piece




1
1
A (k) k 2 2 q + k k k k A (k)
2

"
!
#


k
k
k 2 2 k k
1

2
2
q 2

A (k),
= A (k) k
2
k

k2


(18.55)

whose inversion leads to a propagator


G (x) =

"

dk ikx
i
k k

(2)4
k 2 2
k2

i
k k
2
.
k 2 k 2
#

(18.56)

Apart from the physical propagator, there is a ghost-like state in A (k) proportional
to k .
The new feature of this gauge is that the would-be Nambu-Goldstone mode (x)
has acquired a mass term 2 2 (x). The reason for this lies in the fact that the
gauge fixing term destroys explicitly also the global invariance of the Lagrangian
(x) ei (x),

(x) (x) + ,

A A ,

(18.57)

under which the gauge pieces transform as


1
2

A
2
q

!2

2
1

A ( + )2 .

2
q
"

(18.58)

Therefore, pure phase transformations change the energy and this results in a mass
term.

1074

18 Scalar Quantum Electrodynamics

Let us now take a look at the functional [A, ] in order to see that this does
not modify the Lagrangian (18.54). According to (13.287), we have to integrate
[A, ] =

2
1
A
D exp
2
q

!2

2
1
D exp
A + 2 ( + )
2
q

!2

2
2
1
A + 2
D exp
2
q
q
!

which after a shift of the integration variable gives the trivial constant

2
[A, ] = det + / .
q
2

(18.59)
!2

(18.60)

This can be expressed in terms of Fadeev-Popov ghost fields as


1

[A, ] =

"

i
DcDc exp

2
dx c c + c c
q

!#

(18.61)

The situation is similar to that of the last gauge condition: There are three additional
kinds of asymptotic states: Fadeev-Popov ghosts, ghost-like poles in the gauge part
of the photon field and a state carried by the would-be Nambu-Goldstone field. But
contrary to the previous case. all these states now have a mass which depends on
the gauge parameter . This proves that they must be artifacts of the gauge choice
and cannot contribute to physical processes. In particular, by letting the parameter
tend to infinity, all these particles become infinitely heavy and are thus eliminated
from the physical Hilbert space.

18.3

Spatially Varying Ground States

The theory of a complex scalar field interacting with a photon has another interesting
feature. The ground state must not necessarily be uniform. There are ranges of
mass, and external source where the system prefers to be filled with string like field
configurations and others where the fields have spatially periodic structure of the
hexagonal type. In order to see how this comes about and to be able to compare
results with experimental data we consoder the euclidean three-dimensiona scalar
QED, which is the Ginzburg-Landau theory of superconductivity. Superconductivity
arises if electrons in a crystal are attracted by the effects of the elastic forces so
strongly that the attraction overcomes the Coulomb repulsion. Then they form
Cooper pairs which can form a condensate very similar to the bosonic helium atoms.
In the absence of electric fields, the mean-field approximation to the effective action
is
1
[] =
T

1
m2 2 g 4
1
d x H2 + (D) D +
|| + || .
2
2
2
4
3

"

(18.62)

1075

18.4 Two Natural Length Scales

where Since the charge carriers are electron pairs, the charge of the field is twice the
electronic charge, and the covariant derivative (18.7) becomes
D = iqA = i

2e
A.
h
c

(18.63)

The first expression is in natural units with h


= c = 1, the second in physical cgs
units. The magnetic field is the curl of the vector potential:
H(x) = A(x).

(18.64)

The time-independent equations of motion are


(i qA)2 + m2 + g||2 , = 0

(18.65)

H = j,

(18.66)

and

where j is defined as the current operator of the matter field


j=

q
q 2 ||2 A.
2i

(18.67)

In order to establish contact with the case of a superconductor, we again insert a


temperature dependent mass term
2

m =

T
1 = 2
Tc


(18.68)

thereby choosing the minus sign since our interest will be focussed on the regime
below Tc .

18.4

Two Natural Length Scales

Before proceeding it is useful to introduce reduced field quantities and define


=

|0 |

(18.69)

with
|0 | =


2
=
g
g

(18.70)

being the nonzero vacuum expectation of the -field for T < Tc 0. Also we define a
length scale associated with the mass of the -particle
1
( ) = ,
2

(18.71)

1076

18 Scalar Quantum Electrodynamics

and with that of the photon, after eating up the Nambu-Goldstone boson

g
g
1
( ) =
= =
( ).
q|0 |
q
q

(18.72)

The first length scale is usually referred to as coherence length, the second as penetration depth for reasons which will become plausible soon. The ratio of the two
length scales is an important temperature independent material constant denoted
by :

g
( )
=
,
(18.73)

( )
q
which measures the ratio between the coupling strength g with respect to q 2 .
The coherence length is used to introduce a reduced dimensionless vector potential
a q( )A

(18.74)

h = a = q( ) A.

(18.75)

and a reduced magnetic field

In the absence of a magnetic field, the action may then be expressed as


1
[] = 4fc 3 F .
T

(18.76)

where fc is the so-called condensation energy density:


fc =

20 m2
m4
4
=
= ,
4
4g
4g

(18.77)

and F is the reduced free energy


1
F = 3

d3 x

1
1
1
1
|( )( ia)|2 ||2 + ||4 + 2 ( )h2 .
2
2
4
2


(18.78)

Note that h measures the magnetic field in units of


H0 =

q
q
1
2
2
= m 0 = 2 fc
2Hc
q

(18.79)

where Hc is the value at which the magnetic energy density Hc2 /2 equals the condensation energy density fc .
If we agree to measure all lengths x in multiples of the coherence length , F
becomes simply
F =

1
1
1
d x |( ia) |2 ||2 + ||4 + h2 .
2
2
4
3

(18.80)

1077

18.5 Planar Domain Wall

In the reduced variables the equations of motion are simply


(i + a)2 = ||2,
h = 2 ( a) = 2 [2 a ( a)] = j,

(18.81)
(18.82)

where j is the current (18.67) in natural units


j=

1
||2a.
2i

(18.83)

Let us also write down these equations in the polar field parametrization
(x) = (x)ei(x) .

(18.84)

It is convenient to absorb the gradient of (x) into the vector potential, and go over
to the field
a a ,

(18.85)

This brings F to the simple form


F =

1
1
1
|( ia) |2 2 + 4 + ( a)2 ,
dx
2
2
4


(18.86)

and the field equations (18.81) and (18.82) becoeme




2 + a2 1 + 2

2

= 0,

h = 2 ( a) = 2 [2 a ( a)] = 2 a.

(18.87)
(18.88)

These will be most suitable for a upcoming discus sions.

18.5

Planar Domain Wall

As it was mentioned earlier, it has been known for a long time that superconductors
have the tendency of expelling magnetic fields (Meissner effect). Let us study what
the free energy (18.86) has to say about this magnetic phenomenon. In order to gain
some rough ideas it is useful to consider a sample with a magnetic field pointing in
x direction. Let us allow for variations of the system only along the z-coordinate.
We may choose the vector potantial a0 to point purely into the x-direction. If we
denote the x-component of a(z) a0 (z), the fields are
hy (z) = a (z),

hx = hy = 0.

(18.89)

The reduced free energy density reads


1
1
1
f = 2 2 + 4 + a2 2 + 2 a2 ,
2
2
4

(18.90)

1078

18 Scalar Quantum Electrodynamics

whose extremization yields the equations of motion


+ a2 = 3

(18.91)

2 a = a2 .

(18.92)

Differentiating (18.92) and inserting (18.89), we obtain


d

dz
2 2

1 dh
2 dz

2h
h

= h2 .

(18.93)

Inserting (18.92) into (18.91) yields

2
2h
3

= 3 .

(18.94)

The last two equations can be used as coupled differential equations for h and .
If these two field are known, the vector potential may be calculated from (18.92)
as
a=

h
a
=
.
2
2

(18.95)

This equation gives an immediate result: A constant magnetic field can exist only
for = 0, i.e., if the system has no spontaneous symmetry breakdown. If we assume,
conversely, that = const 6= 0, then Eq. (18.91) shows that also a2 is a constant,
a2 = (1 2 ) /, implying once more from Eq. from (18.95) that h is a constant.
The contradiction is resolved only for = 1, in which case (18.91) gives a = 0, and
(18.95) enforces h = 0. Then Eq. (18.93) shows that only h = 0 is a consistent
solution. Therefore, constant solutions have only the two alternatives
h = const 6= 0,
h = 0,

= 0,
= 1,

normal phase
superconductive phase.

(18.96)
(18.97)

The second alternative exhibits the experimentally observed Meissner effect that the
superconductive state does not permit the presence of a magnetic field.
Consider now z-dependent field configurations. In order to simplify the discussion let us first look at two limiting situations:
Case I :

1/ 2

This corresponds to a very short penetration depth of the magnetic field which may
propagate only over length scales (T ) = ( ) or, in natural units, z 1, i.e., it
has a unit range. For very small , Eq. (18.94) becomes


1 2 ,

(18.98)

1079

18.5 Planar Domain Wall

which can be integrated by multiplying it with to find


2

2
1
1 2 + E.
2

(18.99)

If the system is in the superconductive state for large z, the field satisfy the boundary
condition corresponding to the alternative (18.96), such that 1 for z .
Inserting this into (18.99) we see that the constant of integration E must be zero.
Integrating the resulting equation further gives
z=
or

Z
2

d
= 2 atanh ,
2
1

z
(z) = tanh .
2

(18.100)

(18.101)

The field configuration is displayed as the right-hand branch in the upper part of
Fig. 18.1.
For z < 0, the solution (18.101) becomes meaningless, since > 0 by definition.
We can continue the solution into this regime by matching it with the lower of the
trivial solution (18.97):
h = const, = 0,

(18.102)

which is shown as the left-hand branch in the upper part of Fig. 18.1.
The size of the constant magnetic field in the first case is determined by the
fact that the free energy density must be the same at large positive and negative z.
Otherwise, there would be energy flow. Inserting (18.99) for E = 0 into (18.90), it
becomes approximately
1
1 1
f + a2 2 + h2 .
4 2
2

(18.103)

For large z where h = a = 0 and = 1, this is equal to 1/4. For large negative z,
f is equal to h2 . Hence the constant h in the normal phase is equal to
1
h = hc =
2

(18.104)

Within the region around z 0, h drops to zero over a unit length scale determined
by (18.92) which is very short compared with the length 1/ over which varies
(this is precisely the coherence length). If we plot the transition region against z,
the field h jumps abruptly to zero from h = hc while has a smooth increase. The
full approximate solution plotted in Fig. 18.1 is
h(z) hc (z),
z
(z) tanh .
2

(18.105)
(18.106)

1080

18 Scalar Quantum Electrodynamics

1, type I
N

S
1, type II

Figure 18.1 Dependence of order parameter and magnetic field H on the reduced
distance z for a planar domain wall between the normal and superconductive phases. The
magnetic field points parallel to the wall.

Case II :

1/ 2

Consider now the opposite limit of a very large penetration depth. Here we approximate (18.94) by


(18.107)

a2 1 2 .

(18.108)

2 h2 4 1 2 .
implying that we may omit the gradient term in Eq. (18.91):


Let us calculate again by starting out with = 1 at z = . Then h must decrease


for positive z and we have to take the positive square root
q

(18.109)

(18.110)

h = 2 1 2 .
From Eq. (18.93), one the other hand, we obtain
d
h
dz
2

1 d
dq
1 2 .
h =
2 dz
dz

It is convenient to introduce the abbreviation


u
such that (18.110) becomes

1 2 ,

h = u ,

(18.111)

(18.112)

and (18.109) turns into the simple differential equation




2 u = 1 u2 u.

(18.113)

1081

18.5 Planar Domain Wall

After multiplying this with u , and integrating, we find


2 2

u =u

u2
1
2

+ const.

(18.114)

Inposing the asymptotic condition that at large z the order field goes against 1,
the function u(z) must vanish in this limit, fixing the constant to zero. The equation
is then solved by
z z0 =
or

du
1
1
q
= atanh q
u 1 u2 /2
1 u2 /2
u(z) =

2
.
cosh[(z z0 )/]

(18.115)

(18.116)

From (18.111) we find


=

2
.
cosh [(z z0 )/]
2

(18.117)

As z comes in from large positive values, decreases. We can arrange it to become


zero at z = 0 if we choose z0 such that
sinh(z0 /) = 1.
The magnetic field is then from Eq. (18.112):

2 sinh[(z z0 )/]
h(z) =
.
cosh2 [(z z0 )/]

(18.118)

(18.119)

It is zero at z = , and and becomes gradually stronger as approaches z = 0, where


it reaches the value
1
h = hc =
2

(18.120)

as before. From there on we may again match the trivial solution continuously by
setting
h hc , 0; z < 0.

(18.121)

The full solution shown in Fig. 18.1 is then


s

2
,
cosh [(z z0 )/]

2 sinh[(z z0 )/]
.
h(z) = (z)hc + (z)
cosh2 [(z z0 )/]
(z) = (z) 1

(18.122)
(18.123)

1082

18 Scalar Quantum Electrodynamics

The ranges over which h and vary are of equal order , or in physical units of
order = .
Case III :

= 1/ 2

Here a trial ansatz



1 
h = 1 2
2

(18.124)

can be inserted into (18.93), with the result




1 2 2 = + 2 .

(18.125)

This, in turn, happens to coincide with the other equation (18.94), thus verifying
the ansatz (18.124). For z= The fields start out with = 1 and h = 0, and go
to = 0 and h = hc = 1/ 2 for z = .
Equation (18.125) takes a particularly appealing form if we consider the auxiliary
function (z) defined by (z) e(z)/2 . Then
1
1
= e/2 + 2 e/2
2
k

1
= e/2 ,
2

(18.126)

and we may rewrite (18.125) as


(1 e ) =

2
1
2 = .

(18.127)

2
,
4

(18.128)

d
.
e 1

(18.129)

This can be integrated to


e 1 =
which is integrated to
1
z=
2

For z , the solution h (z) goes to zero like ez/ 2 , such that 1. For
z , it becomes more an more negative like (z) z 2 /4, such that (z) 0
exponentiall fast.
For intermediate values , the solutions have to be found numerically. They all
look qualitatively similar. The ratio of penetration depth to coherence length determines how far the magnetic field invades into the superconductive region relative
to the coherence length.

1083

18.6 Surface Energy

18.6

Surface Energy

Consider the energy per unit length for these classical fields as they follow from
Eq. (18.90). Inserting the equations of motion and performing one partial integration
bith in natural units:
renders the much simpler expression for the energy per area A,


Z
Z
F
1 4 1 2
= dzf = dz + h .
A
4
2

(18.130)

Obviously, the classical solution with an absolute minimum is h = 0, = 1. In


order to enforce the previously discussed configurations, an external magnetic field
is needed, and we must minimize the reduced magnetic enthalpy per unit length
area


Z
1 4 1 2
G
= dz + h h hext ,
A
4
2

(18.131)

where we have subtracted a term


m hext = h hext ,

(18.132)

A = h is coupled to the reduced


in which the reduced magnetization m h G/
external magnetic field. Such a term does change the differential equations (18.91)
and (18.92) for a and , since it is a pure surface term hext z a.
For z , where 0 and h hc , the enthalpy goes asymptotically against


Z
G
1 2

dz
h

h
h
c ext .
A
2 c

(18.133)

For z , on the other hand, where the size of the order field order field goes to
1 and h 0, the asymptotic value is


Z
G
1
.

dz

A
4

(18.134)

Thus both asymptotic regimes


have the same magnetic enthalpy for the particular
external field hext = hc = 1/ 2. If the energy was smaller in one regime, the wall
between superconductive and normal phase would start moving towards one side
such as to decrease the energy. This would go on until the system is uniform. Thus
we conclude: For h > hc , the system is uniformly normal, for h < hc uniformly
super-conductive.
We now calculate the energy stored in the finite region around z = 0 at the
critical magnetic field hext = hc . There the density deviates only slightly from the
asymptotic form, and we must evaluate the expression
G
as Z
G
1
1
1
1
.
=
dz 4 + h2 h +

A
4
2
4

2
!

(18.135)

1084

18 Scalar Quantum Electrodynamics

Note the special properties of the case = 1/ 2: Inserting (18.124), the surface
energy is seen to vanishes.
When inserting into (18.135) the 1 -solutions (18.122) and (18.123), we find,
using Eqs. (18.111) and (18.112), the negative energy

!
!
Z
2
2 1/2
G
as
1
u
G
u
2

u 1
=
dz u 1
A
2
2

1 du
2
u2
1
dz u 1
< 0.

=
2
2
3
2

2 dz

(18.136)

The same sign holds for all > 1/ 2.


For 1, on the other hand, the enthalpy (18.135) vanishes in the superconducting phase for z > 0, where = 1 and h = 0. In the normal phase, where h = hc
by Eq. (18.104), we find
Z 0
Z 1

G
as
1
1
G
4

=
dz 1 =
d (1 + 2 ),
(18.137)

A
4

2 2 0

which is positive for all < 1/ 2.

Thus we conclude: In superconductors with > 1/ 2, it is energetically more


favorable to form a wall in which the magnetic field
develops from zero up to hc than
to have a uniform field configuration. For < 1/ 2, the opposite is true. The first
case is referred to as a soft or type-II superconductor, the second as a hard type-I
superconductor.

18.7

Single Vortex Line and Critical Field Hc1

Consider now a type-II superconductor in small external magnetic field Hext , where
it is in the state of broken symmetry with a uniform order field = 0 . When
increasing Hext , there will be a critical value Hc1 where the field lines first begin to
invade the superconductor. We expect this to happen in the form of a quantized
magnetic flux tubes. vortex lines When increasing Hext further, more and more
flux tubes will perforate the superconductor, until the critical magnetic field Hc is
reached, where the sample becomes normally conducting. The regime between Hc1
and Hc is called the mixed state of the type-II superconductor.
The quantum of flux carried by each flux tube is
0 =

ch
2 107 gauss cm2 .
2e

(18.138)

Such a flux tube may be considered as a line-like defect in the uniform superconductor. It forms a vortex line of supercurrent, very similar to the vortex lines in
superfluid helium discussed in Section 17.4. The two objects possess, however, quite
different physical properties, as we shall see.

18.7 Single Vortex Line and Critical Field Hc1

1085

Suppose the system is in the superconductive state without an external voltage


so that there is no net current j. Suppose a flux tube runs along the z-axis. Then
we can use the current (18.83) to find the vector potential
a=

j
1 1
+
.
||2 2i ||2

(18.139)

Far away from the fluc tube, the state is undisturbed, the current j vanishes, and
we have the relation
a=

1 1
.
2i ||2

(18.140)

In polar decomposition of the field (x) = (x)ei(x) , the derivative of (x) cancels,
and a(x) becomes the gradient of the phase of the order parameter:
a(x) = i (x).

(18.141)

Here we can establish contact withthe discussion of superfluid helim in Section 17.4.
There the superflow velocity was proportional to the gradient of a phase angle
variable (x). The periodicity of (x) led to the quantization rule that, when taking
the integral over d(x) along a closed circuit around the vortex line it had to be
an integral multiple of 2 [recall Eq. (17.105)]. The same rule applies here:
I

d(x) =

dx (x) = 2n.

(18.142)

By Stokes theorem, this is equal to the integral dS a, where dS is a surface


element of the area enclosed by the circuit. Since h = ( a)is the magnetic field
in natural units [recall (18.85)], the integral (18.142) is directly proportional to the
magnetic flux through the area of the circuit
R

 =

SB

dS h =

dS ( a) =

dx a =

dx = 2n. (18.143)

= c = 1,
This holds in natural units, indicated by a bar on top of , in which h
the vector potential is given by (18.74), and x is measure in units of the coherence
length.
The quantization condition in physical cgs units follows from the same argument
applied to the original current (18.67). Remembering
q=

2e
.
h
c

(18.144)

from (18.63), the relation (18.140) reads


A=

2e 1
,
2ihc ||2

(18.145)

1086

18 Scalar Quantum Electrodynamics

and (18.141) becomes


A(x) =

h
c
0
i (x) =
i (x).
2e
2

(18.146)

The magnetic flux integral in cgs units is therefore

=

SB

dS H =

0
dS ( A) =
dx A =
2
B
I

0.

dx = n

(18.147)

It is an integer multiple of the fundamental flux (18.138).


It is instructive to performing the circuit integral (18.143) once more around a
circle close to theHflux tube, where the current in Eq. (18.139) no longer vanishes. The
angular integral dx still has to be equal to 2n, and we find the quantization
rule
j
dx A +
||2
B

= 2n,

(18.148)

dx j = 2n,

(18.149)

or

 + ||1 2

This shows that through the smaller circuit, which enloses fewer magnetic field
lines, there is an increasing contribution of the supercurrent around the center of
the vortex line, which keeps the sum of the two contributions in (18.149) equal
to 2n. This shows that the flux tube is indeed a vortex line with respect to
supercurrent. The circular current density must be inversely proportional to the
distance. This behavior will be derived explicitly in Eq. (18.191).
Quantitatively, we can find the properties of a vortex line by solving the field
equations (18.87) and (18.88) in cylindrical coordinates. Inserting the second into
the first equation, we find
1 d d 2

r +
r dr dr 3

dh
dr

!2

(1 2 ) = 0,

(18.150)

where h denotes the z-component of h. Forming the curl of the second equation
gives the cylindrical analogue of (18.95), i.e.,
h = 2

1 d r d
h.
r dr 2 dr

(18.151)

For r , we have the boundary condition = 1, h = 0 (superconductive state


with Meissner effect) and j = 0 (no current). Since j is proportional to h by
Eq. (18.82), the last condition implies that
h (r) = 0,

r .

(18.152)

18.7 Single Vortex Line and Critical Field Hc1

1087

In cylindrical coordinates, the flux quantization condition (18.143) can be written


in the form

= 2

dr r h(r) = 2n.

(18.153)

Inserting here Eq. (18.151) yields

= 2

"

r
h (r)
2

#
0

= 2

r
h (r)
2

(18.154)

r=0

so that the quantization condition turns into a boundary condition at the origin:
h 2

n1
,
r

for r 0.

(18.155)

Inserting this condition into (18.150) we see that close to the origin, (r) satisfies
the equation

1 d d
n2
r (r) + 2 (1 2 ) 0,
r dr dr
r

(18.156)

implying the small-r behavior of (r):


 n

(18.157)

c2 r
h(r) = h(0) n
2

(18.158)

(r) = cn

+ O r n+1 .

Reinserting this back into (18.155) we have


 2n

For large r, where 1, the differential equation (18.151) is solved by the modified
Bessel function K0 , with some proportionality factor :
h(r) K0

r
,

 

r .

(18.159)

More explicitly, the limit is


q

h(r) /2rer/ .,

r .

(18.160)

Consider now the deep type-II regime where 1/ 2. There (r) goes rapidly
to unity, as compared to the length scale over which h(r) changes, which is in
the present natural units. Therefore, the behavior (18.159) holds very close to
the origin. We can determine by matching (18.159) to (18.155) and find (since
K0 (r) = K1 (r) l/r for small r)
=

(18.161)

1088

18 Scalar Quantum Electrodynamics

In general, h(r) and (r) have to be found numerically. A typical solution for n = 1
is shown in Fig. 18.2 for = 10.
The energy of a vortex line can be calculated hy using (18.86). Inserting the
equations of motion (18.87) and (18.88), and subtracting the condensation energy,
we obtain from (18.130)]
1
Fvort = Fvort Fc =
2

1
d x (1 4 ) + h2 .
2
3

(18.162)

For 1/ 2, we may neglect the small radius r 1, over which increases


quickly from zero to its asymptotic value = 1. Above r 1 the magnetic field for
r r is given by (18.159). Inserting this into (18.150) with (18.157), we find and
for r , h is given by (18.159). Inserting this into (18.150) with (18.157), we find
(r) 1

n2
.
2r 2

(18.163)

Thus the region 1 r yields an energy per unit length


Z
1 Z
1
1
r
1
1
2
4
2
drr 2 + 2 K02
Fvort = 2
drr (1 ) + h = n
L
2
2
r

1
1


 

. (18.164)

For , the second integral becomes a constant, since 0 dx x K02 (x) = 1/2.
The first integral, on th other hand, has a logarithmic divergence so that we find
R

1
Fvort n2 [log + const.].
L

(18.165)

A more careful evaluation if the integral yields n2 (log + 0.08).


Let us now see at which external magnetic field such a vortex line can form. For
vort /L,
this we consider again the magnetic enthalpy (18.131), and subtract from G
ext
vort /L coupling hh so that, per unit length
the magnetic G
1
Gh = n2 (log + 0.08) 2
L

dr r h hext .

(18.166)

But the integral over h is simply the flux quantum (18.143) associated with the
vortex line, such that
1
Gh = n2 (log + 0.08) 2n hext .
L

(18.167)

When this is smaller than zero, a vortex line will invade the superconductor along
the z-axis. The associated critical magnetic field is
hc1 =

n log + 0.08
.
2

(18.168)

18.7 Single Vortex Line and Critical Field Hc1

1089

Figure 18.2 Order parameter , and the magnetic field h for an n = 1 vortex line in a
deep type-II superconductor with K = 10.

Hc1

n=2
n=1

Figure 18.3 Critical field hc1 at which a vortex line of strength n forms when it first
invades a type-II superconductor, as a function of the parameter . The dotted line
indicates the asymptotic result (1/2) log of Eq. (18.168). The magnetic field hc is

measured in units of 2Hc where Hc is the magnetic field at which the magnetic energy
equals the condensation energy.
Table 18.1 The different critical magnetic fields in units of gauss for various impurities.

material
Pb
0.85 Pb, 0.15 Ir
0.75 Pb, 0.25 In
0.70 Pb, 0.30 Tl
0.976 Pb, 0.024 Hg
0.912 Pb , 0.088 Bi
Nb
0.5 Nb, 0.5 Ta

Hc
Hc1 Hc2
550 550 550
650 250 3040
570 200 3500
430 145 2920
580 340 1460
675 245 3250
1608 1300 2680
252
1370

Tc /K
4.2
4.2
4.2
4.2
4.2
4.2
4.2
5.6

1090

18 Scalar Quantum Electrodynamics

For large , this field can be quite small. The asymptotic result is compared
with a numerical solution of the differential equation for n = 1, 2, 3, . . . in Fig. 18.3.
For a comparison with
experiment one expresses this field in terms of the critical
magnetic field hc = 1/ 2 and measures the ratio
n
Hc1
= (log + 0.08).
Hc
2

(18.169)

If the magnetic field is increased above Hc1 , many flux tubes will invade the typeII superconductor. These turn out to repell each other. The repulsion energy can
be minimized if the flux tubes form a hexagonal array, as shown in Fig. 18.4. The
tubes will perform thermal fluctuations, which may be so violent that the superflow
experiences dissipation. The taming of these fluctuations is one of the main problems
in high-temperature superconductors. It is usually don by introducing lattice defects
at which the votex lines are pinned.

Figure 18.4 Spatial distribution magnetization of the order parameter (x) in a typical
mixed state in which the vortex lines form a hexagonal lattice (from W.M. Kleiner et al.,
See References.)

18.8

Critical Field Hc2 where Superconductivity is Destroyed

When considering planar z-dependent field configurations we found that for H > Hc ,
the order parameter vanishes and the magnetic field perforates the superconductor.
Experimentally this is not true. The symmetric field inside the sample to increase
markedly at a larger field value Hc2 , which in a deep type-II superconductor can lie
far above Hc . We shall now see, by allowing for a more general space dependence,
that a non-zero field exists up to a magnetic field

(18.170)
Hc2 = 2Hc > Hc for > 1/ 2.
Only for H > Hc2 the field (x is forced to zero in the entire system, which then a
normal conductor.

18.8 Critical Field Hc2 where Superconductivity is Destroyed

1091

In order to calculate the size of Hc2 , we observe that for H very close to Hc2 ,
the order parameter must be very small. Hence we can linearize the field equation
(18.81), writing it approximately as
(i + a)2 .

(18.171)

For a uniform magnetic field in the z-direction, we may choose different forms of
the vector potential a which differ by gauge transformations. A convenient form is
a=

h
(0, x, 0) .

(18.172)

Then satisfies the Schrodinger equation

!2

h
2 + iy + x
x

This may be solved by a general ansatz

z2 = .

(18.173)

(x, y, z) = eikz z+iky y (x),

(18.174)

where (x) satisfies the differential equation

!2

h
d2
2 + ky + x
dx


1 kz2 (x)

= 0.

(18.175)

This is the Schrodinger equation of a linear oscillator of frequency = h/, with


the potential centered at
x0 =

ky
.
h

(18.176)

In fact, Eq. (18.175) can be written as


xx (x) + 2 (x x0 )2 (x) = (1 kz 2 )(x).

(18.177)

This equation has a solution (x), which goes to zero for x , if



1
h
1
1
1 kz2 En = n +
=
n+
.
2
2

2


(18.178)

The energy En = 12 (1 kz2 ) is bounded by 1 from above. Hence there cannot be


solutions for arbitrarily large h. The highest h is supported by the ground state
solution, where n = 0. But also this ceases to exist as h reaches the critical value
hc2 given by
h
1

=
=
2
2
2

(18.179)

1092

18 Scalar Quantum Electrodynamics

where
hc2 = .

(18.180)

Going back to physical magnetic fields, this amounts to

Hc2 = 2Hc .

(18.181)

The wave functions of these solutions are strongly concentrated around x x0 ,


implying
that the system is normal almost everywhere, except for a sheet of thickness
q
/h, in units of the coherence length.
This sheet carries a supercurrent. Let us study its properties. If we insert (x)
of Eq. (18.174) and a of Eq. (18.172) into Eq. (18.83) for the supercurrent, we find
"

h
1
||2 a = |(x)|2 (0, ky , kz ) (0, x, 0)
j =
2i

!
h
= |(x)|2 0, (x x0 ), kz .

(18.182)

The supercondicting sheet is centered around x = x0 parallel to the y z-plane. It


carries no net charge flow along the x-axis. It consists of a double layer of current
flowing in y directon for x > x0 or x < x0 , respectively.
If the current along the z-axis is nonzero, the critical magnetic field is reduced
to a lower value


hc2 1 kz2 .

(18.183)

The sheet configuration is obviously degenerate with respect to the position x0


of the sheet. Moreover, being in the regime of a linear Schrodinger equation, we can
use the superposition principle to set up different spatial structures. For example,
we may form an average over sheets in the rotated (cos
x + sin
y , z)-planes, and
obtain cylindrical structures. These have a definite angular momentum around the
z axis rather than a definite linear momentum ky . Their wave function can be found
directly from a vector potential
a=

h
(y, x, 0),
2

(18.184)

which reads in cylindrical coordinates


a =

h
,
2

ar = 0,

az = 0.

(18.185)

The linearized field equation now becomes


"

1
2
1 2

r + 2+ 2 2
r r r z
r

h2
+ 2 r 2 = (z, r, )

(18.186)

1093

18.9 Order of the Superconducting Phase Transition

and may be solved by a factorized ansatz


q

(zr) = ( h/r)eim eikz z .

(18.187)

The differential operator is the same as for a two-dimensional harmonic oscillator of


frequency = h/ and its wave functions are well known. They may be expressed
in terms of the confluent hapergeometric functions 1 F1 (a; b; z) as

(r ) = er /2 r |m|/2 1 F1 (n; |m| + 1; r ),


The corresponding energy eigenvalues where r =

(18.188)

h/r.

1
1
1
E = n + |m| m +
+ kz2 .
2
2
2


(18.189)

They have to be equated with 1/2(1 kz2 ) to obtain the highest possible magnetic
field, which is again found to be
hc2 = (1 kz2 ),

(18.190)

as in (18.183). The current carried by this cylindrical solution is found by inserting


(18.184) and (18.187) into (18.82). For n = 0 and arbitrary m > 0, we obtain
"

x
h
y
h/r m 2 , 2 , 0 + (0, 0, kz ) (y, x, 0)
j =
r
r
2
! #
q
"
h
m
= 2
e
h/r kz ez + 2
r
2
2

q

(18.191)

The wave function is non-zero within a narrow tube of diameter /h in units of


the coherence length. Within this tube, a current is rotating clockwise around the
z-axis, just as we deduced before from the qualitative discussion of Eq. (18.149). In
addition, there may be an arbitrary linear current flowing in z-direction.
The magnetic enthalpy (18.131) of a type-II superconductor is always smaller
than zero because of the negative interfacial energy. Therefore, states with a nonzero
magnetic field will laways form for H below Hc2 .
If the magnetic field is far below Hc2 , the order field (x) will be so large that
the quartic terms in the field equation (18.87) can no longer be neglected. The field
configuration becomes inaccessible to simple analytic calculations.

18.9

Order of the Superconducting Phase Transition

At the mean-field level, the superconducting phase transition is of second order. If


the supercoductor is of type II, this is no longer true. The fluctuations of the vector
potential make the transition first order. This result is obtained by studying the
statistical mechanics of the vortex lines in a superconductor. This is possible with
the help of disorder field theory, which will be discussed in Chapter 20 [1, 2].

1094

18 Scalar Quantum Electrodynamics

Notes and References


For more information on vortex lines see
H. Kleinert, Gauge Fields in Condensed Matter , Vol I, World Scientific, 1989
(kl/b1), where kl is short for http://www.physik.fu-berlin.de/~kleinert.
See also: H. Kleinert, Multivalued Fields in Condensed Matter, Electromagnetism,
and Gravitation, World Scientific, Singapore 2009, pp. 1497 (kl/b11).
The individual citations refer to:
[1] H. Kleinert, Lett. Nuovo Cimento 35, 405 (1982) (kl/97).
[2] H. Kleinert, Europhys. Letters 74, 889 (2006) (cond-mat/0509430) (kl/360)

Nobody can be exactly like me.


Sometimes even I have trouble doing it
Tallulah Bankhead (19031968)

19
Exactly Solvable O(N )-Symmetric 4-Theory
for Large N
The functional integral representation of quantum field theory introduced in Chapter
13 has an important advantage over the operator formulation: it offer a similar
freedom in changing variables of integration as ordinary integrals do. In physical
applications it may often be more economic to describe the system not in terms of
the most fundamental particles. In fact, many interesting phenomena in many-body
systems are carried by products of funcdamental fields which will be called composite
fields or collective field.

19.1

Introduction of a Collective Field

As an example we consider a 4 -theory, in which the field carries an extra index


a = 1, . . . , N allowing for an O(N)-degree of freedom. The generating functional is
Z[j] = N

De

d4 x

m2
1
(a )2 20 2a g4
2

(2a )

+j a

(19.1)

where N is a normalization factor


h

N = Tr log 2 m20

i1/2

(19.2)

We have slightly changed the convention of how the coupling constant appears in
the action, replacing g/4! by g/4, to save some numerical factors in the upcoming
calculations.
The above theory of a fluctuating a will now be tranformed into another one,
in which only an O(N)-invariant collective field appears, which describes the phenomena carried by the composite 2a . We can achieve this by means of an auxiliary
functional
Zaux =

2 2

Dei 4g (ga ) .
1

1095

(19.3)

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1096

The integral is of the pure Fresnel type and cannot depend on the field 2a . Thus
it is a constant and may be multiplied into (19.1) without changing the dynamics.
Note that the field fluctuates quadratically around 2a (up to the factor g). Thus it
contains all informations on this composite operator. In fact the classical equation
of motion is simply
cl = g2a ,

(19.4)

i.e., at the classical level is equal to g2a . The modified functional


Z[j] = Zaux
= N

De

DDei

d4 x

m2
1
(2a )2 20 2a g4
2

(2a )

+ja a

1 2
+ja a }
d4 x{ 12 a ( 2 m20 )a + 4g

(19.5)

describes exactly the same theory as before.


In the presence of the collective -field, the a -fields have a propagator
G (x, x )

i
(x, x ).
m20

(19.6)

Diagrammatically it is easy to understand why this is so. In the diagrams associated with the action in (19.5), the original four-vertex
= 3!gTabcd = 2g(ab cd + ac bd + ad bc )

(19.7)

is generated by the exchange of . If this field is represented by a wiggly line, its


interactions with the original field may be pictured by the diagram

(19.8)

ab

The quadratic part of the action in the field is


tv

A0 [] = i

1 2
.
4g

(19.9)

which corresponds to a propagator1


G =
1

= 2g.

(19.10)

The name propagator is not quite appropriate, since the -particle does not propagate in this
lowest approximation.

1097

19.1 Introduction of a Collective Field

The four-vertices (19.4) are then recovered, in the partition function (19.5), from
the diagrams

(19.11)

An important point to observe now is that the action (19.5) is quadratic in such
that we can perform the Gaussian functional integral over using formula (13.25),
obtaining
Z[j] = N

De

d4 x

1 2

+iN h
Tr log
4g
2

i
( 2 m20 )+ 2i j 2 m
2 j
0

(19.12)

where the last term is due to the quadratic completion [1].


quadratic completion
The factor N in front of the trace log arises from the N components a each of
them giving the same contribution.
The new representation of Z[j] describes the same physics in terms of a single
composite field rather than N fundamental fields a .
Since the main application of this formalism will be in many-body problems
where composite fields describe collective phenomena, the field will be referred to
as collective field and the exponent in (19.12) as collective action.
Let us derive the Feynman rules for the collective field theory. The free part
of the action is 2 /4g with the -propagator (19.10). The vertices are obtained
from the powers n in an expansion of the remaining part of the collective action.
The trace of the logarithm may be expanded into loops just as shown in (15.32). In
contrast to the diagrams (15.36), the loops now emit single wiggly lines for rather
than pairs of -lines

(19.13)
Similarly, the current piece may be expanded in as
"

1
i
1
i
i
i
j 2
j = j
+
(i) 2
2
2
2
2
2
2 m0
2 m0 m0
m20
#
i
i
i
(i) 2
(i) 2
+ ... .
(19.14)
+
2
2
2
m0
m0
m20

This corresponds to the following Feynman graphs

(19.15)
(19.16)

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1098

in which a line connects two currents and emits any number of wiggly -lines along
the way.
It is obvious that every fundamental graph can be recovered by connecting the
-ends in (19.13) and (19.15) via the propagator G = 2g.

19.2

The Limit of Large N

One of the most interesting features of the collective field theory is that it can be
used to solve the full quantum field theory in the limit N if the coupling
constnat is simultaneously decreased as 1/N by setting
g = g1 /N.

(19.17)

with fixed g1 , in order to see how this works we first rewrite, in the absence of
external currents j the functional integral over as [2]
iN Acoll []

iN

De

d4 x

1
2g1

2 i h
Tr log( 2 m20 )
2

(19.18)

But the factor N in the exponent has the same effect as previously 1/h in the
semiclassical limit. For N , the exponential becomes a -functional around the
extremum of the action which is given by
!

(x) = 4g1
2

i
(x, x).
2
m20

(19.19)

Fluctuations around this solution are suppressed by a factor 1/N (just as previously
h
). Thus, to leading order in N, the is given by
W [j]|j=0 = Acoll [] =

"

1 2
h

dx
iN Tr log( 2 m20 ) .
2g
2
4

Let us now include external currents and consider


iW [j]

1
i
D exp iNAcoll [] j 2
j .
2 m20

(19.20)

Then as N , we have to allow for the possibility that the external particles
contribute to leading order in N (for instance if there is one external line for each
index a = 1, . . . , N). Then the field has to be chosen to extremize the full exponent
which happens, say at = [j]. The extremization becomes more familiar looking
by going from W [j] to the effective action. For this we introduce the effective field
a [j] =

W [j]
i
=i 2
ja ,
ja
m20 [j]

(19.21)

whose inversion yields ja = ja [a ] and from this [] = [j[]]. Note that there
is no contribution from the /ja -dependence via the chain rule /ja W [j]/,

19.2 The Limit of Large N

1099

since W [j] is extremal in for fixed ja . Now we may calculate the effective action
from
[] = NAcoll []

1
2

d4 xa 2 m20 a ja [a ]a . (19.22)

Here is the extremal functional [], It is obvious that this functional can be
obtained directly from minimizing (19.22) at a constant . This is so since the
extremality condition
W [j]
=0

K=

determines = [j], and implies also that


[] = [W [j] j] = (Wj [j] ) j + W [j] = 0,

(19.23)

a relation which serves to determine = []. Thus it is more practical to consider


[, ] = NAcoll [] +

1
2

d4 x a 2 m20 a

(19.24)

as an effective action of two fields and . The physical field configurations are
obtained by extremizing [, K] with respect to both arguments, thus obtaining say,
0 , 0 .
The calculation of higher vertex functions proceeds as follows. One expands
[, ] around 0 , 0 by writing
h

[, ] = 0 + , 0 + = 0 , 0
o
i
i
h
i
h
h
1n
+
0 , 0 + 2 0 0 + 0 0
2
+ higher derivative terms.
(19.25)
This first derivatives vanish, the second determine propagators of and fields
h

0
0
G = i1
,

0
0
G = i1
, .

(19.26)
(19.27)

By means of the vertex functions fixed by the higher derivatives we may now compose
all tree diagrams. Only those with external lines survive since = 0. But any
number of internal -lines may occur.
Let us convince ourselves that these vertex functions are the same as those which
would be obtained from higher functional derivatives of [] of (19.22). In order to
avoid writing the two arguments of [, ] let us denote [] [, []] by F []
for the discussion to follow, i.e.,
F [] [, []]

(19.28)

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1100

As argued before, the functional [] satisfies


[, []] = 0

(19.29)

= 1
.

(19.30)

or

This relation can be used when deriving the higher vertex functions from (19.28):
F = + = 1
.

(19.31)

This may be represented graphically as


(19.32)
An open blob with two double legs stands for the second functional derivative of
[, ] with respect to , which is i times the propagator of the -field:
(19.33)

The grey blobs are the vertices of the two-argument effective action [, ]. They
have single or double legs emerging, the latter representing the -field. Differentiation this with respect to adds one more single leg:

(19.34)

A differentiation with respect to adds one more double leg:

(19.35)

19.2 The Limit of Large N

1101

3
Since 1
/ = we see that differentiating (19.33) with respect to
produces a three- vertex:

(19.36)

Differentiating (19.31) further we find


1
3
F = 1

(19.37)

which has the graphical form

(19.38)

The procedure can be continued via the chain rule of differentiation. A blob with n
legs
F =

(19.39)

represents an n-point vertex function F . A differentiation with respect to the


field adds, step by step, one more external leg, i.e.,

(19.40)

Proceeding in this way we can keep on differentiating (19.37) and construct all
higher diagrams. At the end, we discard all diagrams in which a double line ends
in a shaded blob (-line tadpole diagrams), since is extremal such that = 0.
The particular structure of [, ] in Eq. (19.24)
[, ] = NAcoll [] +

1
2

d4 xa 2 m20 a

(19.41)

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1102

can be used to simplify these graphical rules. If we set [] NAcoll [], we obtain
the simple vertex functions
a b =
a =
a b =
a1 ...an =
1 ...n a =
1 ...n =

2 m20 ab

(19.42)

a ,
ab ,
0,
n 2,
0,
n 2,
n 2.
1 ...n

(19.43)
(19.44)
(19.45)
(19.46)
(19.47)

Then (19.37) becomes


1
1
2
a b = 2 m20 a
b +
3
.

(19.48)

This has to be evaluated at the extremum where = 0. We therefore obtain the


propagator of the -field
1
G = iF
=i

h

1
b
2 m20 ab a

i1

(19.49)

Graphically, we exhibit the simplification brought about by Eq. (19.42) and (19.43)
by replacing the diagrams and as follows:
(19.50)

Then (19.36) looks as follows:

(19.51)

and (19.38):

(19.52)

19.2 The Limit of Large N

1103

A functional differentiation of with more than one -leg with respect to


gives

(19.53)

Let us finally write down the propagators and vertices of the -field explicitly.
For this we consider the Taylor series of the trace log term in (19.18):


i
N Tr log 2 m20 0
2
"
!#


i
i
2
2
0

= N Tr log m0 + Tr log 1 +
i
2
2 m20 0
"
#n

i
1
i X

(i ) .
= N
(19.54)
2 n=1 n 2 m20 0

Reading off the quadratic terms


"

N 1 2
+ iTr
4 g1

2
m20 0

!#

(19.55)

we find in momentum space the Wick rotated propagator

G (q )

i1

2
1
= (q q ) i
N g1
(4)

)1

dD k
1
1
.
2
D
2
0
2
(2) k m0 (qk) m20 0
(19.56)

This propagator is a non-perturbative object. Expanding it in powers of g1 , we may


write the interaction caused by it as a series of ordinary Feynman diagrams

N
2

(19.57)
where the wiggle with a blub represents the propagator (19.56) of the -field, a line
the simple propagator
G =

i
,
2 m20 0

(19.58)

1104

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

and a dot the fundamental 4 -vertex g1 .


It should be pointed out that the higher terms bubble sum in (19.57) do not
carry the same multiplicity factors as in the original perturbation expansion of the
4 -theory in (10.6.1). The factors in that expansion are 4g/4!, (4!4!/2)(g/4!)2 for
zero- and one-bubble diagrams. After exchanging g/4! g/4 these become 6g, 24g 2.
In the exchange between two -lines, only only one third of the contractions are
done, so that 6g goes reduces to 2g = 2g1 /N, which explains the first term in the
propagator (19.57). For N components, the second multiplicity receives a factor
TI = (N + 8)/9 [recall (11.230) and (11.239)], which in the limit N becomes
N/9. Thus the second terms carries a factor 2Ng 2 = 2g1 /N, in agreement with the
second term in (19.57).
The n-point vertex functions for the field are given by
"

i
N i
(i )
i
2
2 n m20 0

#n

(19.59)

which are the one-loop diagrams (19.13) with lines emerging from the circle. In
this way every n-point vertex function of the original field theory has been reduced,
in the limit N , to a finite sum of tree diagrams involving propagators
and -vertices, all of which consist of one-loop diagrams of the field. For the
four-point function we obtain

(19.60)

plus terms which contain at the end of one double line and vanish at the extremum. We now see that the rules derived in this algebraic fashion are indeed just
the Feynman rules for the tree diagrams of the theory with propagators
1
G = i

and shaded blobs as vertices representing . The double lines may either end in
a field a or in a Kronecker ab .

1105

19.3 Variational Equations

19.3

Variational Equations

We have seen in Section 14.9 that the effective action can be used as a basis for
variational calculations of a field theory. As an application we study here a simple
exactly solvable field-theoretic model which permits us to see the effect of all loop
corrections. The simplest model is a 4a -theory with N fields 1 , . . . , N in the limit
N . In statistical mechanics, this is known as the spherical model. In the last
section, we we have derived the effective action as a function of the classical field
a (x) and the collective field (x) as
1
m2
1
N 2
[a , ] = A[a ] + coll [, ] =
d x

(a )2 0 2a 2a +
2
2
2
4g0


i
+ N Trlog 2 m20 .
(19.61)
2
Z

"

To leading order in N, this is the exact effective action of the theory. For the
discussion to follow it will be useful to abbreviate m20 + (x) by (x):
m20 + (x) (x),

(19.62)

since this combination appears inside the functional Tr log like a spacetimedependent mass term. Then we write
[a , ] =
+

"

1
N 2
N 2
N 4
1
(a )2 2a
m0 +
+
m
d x
2
2
2g0
4g0
4g0 0
D



i
N Trlog 2 .
2

(19.63)

According to Eq. (14.79), the effective action has to be extremal for physical
values of the vacuum expectation of the fields a = ha i. We expect the ground
state to be isotropic in spacetime, and look for minima of the effective potential
as a function of constant fields , a , for which the functional Tr log reduces to a
simple integral over phase space. Dividing out the spacetime volume, we obtain the
effective potential
1 2
N 2
N 2
N 4 N
v(, ) =
a +
m0

m +
2
2g0
4g0
4g0 0
2



dD k
2
log
k
+

. (19.64)
(2)D

The momentum integral can be done using formula (11.134) and we obtain
v(, ) =

N 2
N 4 1
N 2
1
2
1 2
m0

m0 + N S
a +
D (D/2)(1 D/2) D/2 .
2
2g0
4g0
4g0
2
2
D
(19.65)

To find the extrema of this we distinguish two cases:

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1106
1. Case

V
= a = 0,
a
which allows for the two possible solutions:
a)

a = 0

b)

a 6= 0,

= 0.

2. Case

V
Nm20 N
1
= 2a +
+N S
D (D/2)(1D/2)D/21 = 0.

g0
g0
2

(19.66)

(19.67)

The extremum 1a) corresponds to a normal ground state possessing the same
O(N)-symmetry as the original action. The ground state associated with the other
extremum 1b), on the other hand, breaks the O(N)-symmetry of the action. This
phenomenon is called 0 spontaneous symmetry breakdownIt is a universal phenomenon in It is a universal phenomenon in many-body systems, where it is observed
in the form of ordered phases. The most widely known systems of this type are ferromagnets, which display a spontaneous magnetization sufficiently low temperature.
The magnetization vector may be identified with the field at an expectation a
for N = 3. The magnetized state is not rotationally invariant, and violates the
rotational invariance of the Hamilton operator of the system. Above a critical temperature called Curie or Neel temperature, the system looses its magnetization and
becomes rotationally symmetric. This is the normal state.
In the O(N)-symmetric solution 1a), the second equation (19.67) determines :
N
1
Nm20
= NS
D (D/2)(1 D/2)D/21 .
g0
g0
2

(19.68)

In the case 1b), it determines


Nm20
.
(19.69)
g0
We have seen before that the functional integral is stable only for g0 > 0. The state
with spontaneously broken O(N)-symmetry can exist only for m0 2 < 0.
2a =

19.3.1

Non-trivial Ground States

The solutions to the above equations depend on the dimension of spacetime, and
will therefore be discussed separately:
Dimension D=4
Setting D = 4 and using (11.162) to expand
1
1
(D/2)(1 D/2) = + O(),
2

1107

19.3 Variational Equations

equation (19.67) becomes


2a =


1

N
m20 + N S
D N S
D log 2 .
g0

(19.70)

This is renormalized by introducing the physical masses and coupling constants m


and g:
1
1
1

+S
D ,

g
g0

2
2
m0
m
,
=
g
g0

(19.71)
(19.72)

with an arbitrary mass parameter . Then we can take the limit 0 in (19.70)
and find
2a =


1
N

4 log 2 .
m2 N S
g
2

(19.73)

where S4 = 1/8 2 the corresponding potential is


!

1
N
N
N 4

1
1
m0 + N S
4 log 2
.
v(, ) = 2a + m2 2
2
2g
4g
4g0
8

2
(19.74)
Consider the ground state with a spontaneous symmetry breakdown, where the
minimum of n(a , M) lies at a 6= 0 with = 0. From (19.69) and (19.72), this
minimum lies at
2a =

Nm2
g

(19.75)

For positive renormalized coupling constant g, this ground state requires a regular
m2 . Let us see how v(a , M) behaves in the neighborhood of m2 = 0. For m2 > 0,
2a necessarily increases since the logarithm in (19.74) is negative. Its slope in is
initially infinite, being given by

1 2 Nm2
N
1
1
v(a , )
a +
log 2 +
=
NS
4

2
g
2g
4

2
"

!#

(19.76)

If we insert (19.73) back into the potential (19.74), we find


vmin

1
Nm2
2
= a +
4
g

NS
4

1 2
.
16

(19.77)

Then, for increasing small , the potential increases. This goes on until vmin reaches
a maximum which is determined by
2a max

Nm2
1
+
= NS
4 max .
g
2

(19.78)

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1108

Together with (19.73) this implies


log

max
2
= 1 +
.
2

gS
4

(19.79)

If we want to calculate a (M) for a value 2a > 2a max , we have to choose


complex, so that also the potential becomes complex indicating an instability of the
corresponding state. Even though this field configuration can be reached only via
an external current, there are disastrous consequences as we shall see later when
discussing the Green functions of the theory.
A similar phenomenon arises for 2a < Nm2 /g. Then has to be taken imaginary (note: real negative would give an imaginary 2a ). Also there, v(a , )
becomes complex. Contrary to the previous regime, however, this complexity is of a
simple physical nature. We shall see in Section 17.2 that it is related to a singularity of all Green functions at zero energy-momentum which arises from the massless
Nambu-Goldstone bosons.
Dimension D=3
Here the equation (19.67) reads
2a =


N 2
m m20 + N S
3 M,
g0

(19.80)

where S
3 = 1/2 2). With this regularization neither m20 nor g0 need to be renormalized, so that they can be replaced by m2 and g. The potential becomes
1
Nm2
v(a , M) =
2a +
2
g

N 2

NS
3 3/2 .
4g
6

(19.81)

Again, as increases from zero, 2a +Nm2 /g moves to positive values. The potential
at the extremum g in is
vext

N
4
1
1
Nm2

1
2

=
+
g
2 D
4g
D
!
2
1 N 2
Nm
1
+
2a +

=
6
g
12 g
!

2a

(19.82)

There is now an important difference with respect to the previous D = 4, case. The
potential increases monotonously for 2a > Nm2 /g and does not become complex
for any large positive 2 . For 2a < Nm2 /g, however, the situation is in some
respect similar to D = 4, but in an important respect it is different. To see most
easily what happens we may express from (19.67) in terms of 2a + Nm2 /g and
have

g
= S
3 +
4

"

gS
3
4

2

g
+
N

2a

Nm2
+
g

!#

(19.83)

1109

19.3 Variational Equations

As long as 2a > Nm2 /g, the solution has > 0. Below this value, the square
root becomes negative. Thus there is again a complex value of the auxiliary field ,
just as for D = 4. The difference with respect to that case lies in the fact that in
spite of being complex v(a , ) remains real, and increases with 2 < Nm2 /g.
It therefore has a proper minimum at 2 = Nm2 /g. Thus the theory is consistent
and displays a spontaneous breakdown of O(N)-symmetry.
Dimension D=2
In two dimensions everything changes drastically. Here the potential becomes for
D = 2 + [using (D/2)(1 D/2)2/D 2/ + 1]
V

1 2
N 2
N 4 1
N 2

=
m0

m0 N S
a +
2 log 2 1 ,
2
2g0
4g0
4g0
4

so that Eq. (19.67) becomes


2a =


1 1

N
m20 + N S
D + N2 S
D log 2 .
g0
2

(19.84)

We can again renormalize the mass by introducing


1
m2 = m20 g0 S
D .

(19.85)

For the coupling constant there is no need of renormalization, i.e., g = g0 . Then the
finite equation reads
2a =


N
1
m2 + N S
2 log /2
g
2

(19.86)

where S
2 = 1/2. The potential at the minimum is given by
vmin

Nm2
1
2 +
=
2
g



N 2
1
+N S
2 log /2 1 .
4g
4

(19.87)

Now 2a as a function of behaves quite different from the previous cases. There
is no finite value of 2a for which can be equal to zero since for 0, 2a to
. Thus there can be no spontaneous symmetry breakdown in two dimensions.
This condensation turns out to have an exception for N = 2, where there exists a
quasi-ordered ground state with a quasi-symmetry breakdown. The conclusion also
does not apply to N = 1. Since the O(1) symmetry is discrete, the only group
elements being 1 and -1. The O(1)-symmetric 4 -theory is equivalent to an Ising
model of magnetism which is well-known to have a second-order phase transition
into a magnetized state, which is a phase with broken symmetry.

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1110
Dimension D=1

Also in one dimension, there can be no spontaneous symmetry breakdown. This is


obvious since the system is equivalent, to quantum mechanics of a particle at the
time dependent position (t), i.e., the Hamiltonian
1
m2
g0  2 2

(19.88)
(a )2 0 2a
2
2
4N a
may be interpreted as that of a mass point in N-dimensional space moving in an
O(N)-invariant potential. The ground state is unique, and there can be no degeneracy, hence no spontaneous symmetry breakdown. Contrary to the case D = 2,
this argument also holds for any N, also for N = 2 and N = 1. Certainly, Eq.
(19.67) has to reflect this feature and it indeed does so. There is now no need for
renormalization and, so that we can omit the subscripts zero and find for D = 1
with S
1 = 1/:
H=


N
m2 ()1/2 ,
g
!
1
N
Nm2
2
= +
+ 2 .
2
g
4g

2a =
vmin

(19.89)
(19.90)

Again there is no solution for 0, since this would imply 2a .

19.4

Special Features of Two Dimensions

The impossibility of the spontaneous breakdown of a continous symmetry in two


spacetime dimensions is related to the fact such a phase would contain massless
Nambu-Goldstone bosons in two dimensions. But it can be shown that massless
fields (x) cannot exist in this reduced spacetime. The reason is basically, that the
associated propagator
G(x) = h(x)(0)i =

d2 q i iqx
e
(2)2 q 2

(19.91)

is divergent for all x, due to the singularity at q = 0. It needs a small-q cutoff, or


equivalently an infinitesimally small mass . In D dimensions, the massive propagator reads, according to (7.137) Eq. (7.135),
G(x) =

i
dD q
eiqx
D
2
2
(2) q

1
=
(2)D/2

|x|

!D/21

KD/21 (|x|).

(19.92)

where |x| (x x )2 . For |x| 1/, which due to the smallness of is almost
no restriction, we use the small-argument limit of the Bessel function
 

1 2
K (z) ()
2 z

(19.93)

1111

19.4 Special Features of Two Dimensions

to obtain the approximations

G(x)

1
(D/2 1) D/2 D2 , D > 2
4 |x|
1
x2
log
,
D = 2.
2
2e

(19.94)

where = 0.577 . . . is Eulers number. The two-dimensional expression diverges


in the limit 0. This behavior has given rise to a physical feature of twodimensional systems, formulated as a theorem, which in many-body physics is known
as the Mermin-Wagner theorem, and in quantum field theory as Colemans theorem
[3]. Nonwithstanding this theorem, the system in which one would have expected
massless Nambu-Goldstone modes do have a characteristic feature: They show what
are called quasi-long-range correlations. These lead to physical observations which
are quite similar to those of proper long-range fluctuations [4].There are experiments
which measure correlation functions of the exponential of such fields, and these have
a power-like behavior. Let us calculate this. By Eq. (19.94) we have
1

2 2 (x)2ab(x)(0)+b2 2 (0)i

C(x) heia(x) eib(0) i = e 2 h[a(x)b(0)] i = e 2 ha


1

2 G(0)

= e 2 (ab)

eab[G(x)G(0)] .

(19.95)

Now, since G(0) is positively infinite, the prefactor vanishes unless a = b, in which
case it is equal to unity, and we obtain
a2

2 [G(x)G(0)]

heia(x) eia(0) i = e 2 h[(x)(0)] i = ea

(19.96)

The exponent contains now the subtracted propagator (19.94):

G (x) G(x) G(0)



d2 q
1
x2
i
iqx
e 1 log
.
(2)2 q 2 2
2
2e

(19.97)

In contrast to (19.94), this subtracted integral is now infrared finite but ultravioletdivergent, and contains therefore an ultraviolet cutoff . The x-behavior (19.97) is
correct for |x| 1/, which is again almost everywhere, due to the large size of .
After inserting (19.97) into the correlation function (19.95) with a = b, we obtain
the pure power falloff behavior with the distance
C(x) =

!a2 /2
x2
,
2e

(19.98)

called quasi-long-range behavior. It is observed experimentally for many observables in two-dimensional systems where a perfectly ordered stae is forbidden by the
Mermin-Wagner theorem.
In two dimensions, the system with the complex pair field exhibits also a
completely new pheomenon. There are macroscopic quantum fluctuations which

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1112

produce quantum vortices and antivortices, in analogy to the vortex lines in threedimensional superfluid helium and superconductors, discussed in Section 18.7. They
attract each other by a logarithmic potential just like an electron gas with Coulomb
interaction. At a temperature Tc , the long-range correlations of this Coulomb interaction is made short range by screening, which is caused by the dissociation of
vortex pairs into free vortices. Thus there is a phase transition even though there
can be no condensate [5].

19.5

Experimental Consequences

Given the effective potentials and the associated ground states and assuming that
the N limit carries valid information also for finite N we can study the
radiative correction to field expectation values like
h2 (x)i

(19.99)

and the susceptibility


(2)D (D) (q + q )1 (q) hT a (q)b (q )ic




2 []

D (D)

2
2
=
ab , (19.100)
=
(2)

(q

q
)
m
+

q

a (q)b (q ) =0

where the subscript c indicates the connected Green function. In mean-field theory,
the first quantity behaved like
h2 (x)i = 0,
h2 (x)i =

N
Nm

,
g
g

for m2 > 0,

( > 0),

(19.101)

for m2 < 0,

( < 0).

(19.102)

The inverse susceptibility is for T > Tc


(q) = m2 + q2 q02 = 2 + q2 q02 ,

(19.103)

k (q) = q2 q02 ,
(q) = 2m2 + q2 q02 = 22 + q2 q02 .

(19.104)

and for T < Tc

These behaviors can be modified considerably by the radiative corrections calculated in the last section. Consider first the cases of one and two dimensions
where there is no spontaneous symmetry breakdown. Here the susceptibility remains isotropic and finite for all temperatures. We calculate
1 (0) =
1 (q) = + q2 q02

(19.105)

1113

19.5 Experimental Consequences

and see that far above Tc , 1 approaches the mean field value since there
m2 2 Tc . Close to Tc , however, 1
(0) goes no longer to zero. It smoothly
becomes flatter and tends to zero like
2

2 4/g
1
(0) e
g 2 1
1
(0)

44

D=2
D = 1.

This rapid falloff may be seen as a washed-out remnant of the original anisotropic
vanishing of 1
(0) due to the Nambu-Goldstone modes. Here it is due to large
isotropic pretransitional fluctuations. The expectations of h2 (x)i can be obtained
as follows: Differentiate the generating functional Z[j] with respect to the mass
square in the functional picture. Then we see that
2i

Z[j] =
m2

Z

dD x2 eiA[]+i

dD xj(x)(x)

(19.106)

But this is, up to the normalization factor Z[j], the Green function in the presence
of the external field j:
Z

dD xhT (x)(x)ij .

(19.107)

Explicitly
Z

dD xh2 (x)ij = Z 1 [j] 2i

Z[j].
m2

(19.108)

From this, the desired quantity h2 (x)i can be calculated as


h0|2(x)|0i =

1 1

Z [0]2i
Z[0].
VT
m2

(19.109)

Using the effective action this becomes


h0|2 (x)|0i =

2
[]
V T m2

(19.110)

v().
m2

(19.111)

or with the effective potential


h0|2 (x)|0i = 2

Inserting the previous results in one and two dimensions we find


h0|2(x)|0i =


1
m2 .
g

(19.112)

For large m, where m2 , this tends to zero, i.e., to the mean-field result. As
m2 goes to zero, however, the function remains smooth (and positive). As the

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1114

temperature drops far below Tc (m2 ) , tends to zero and the behavior
becomes again mean-field like.
We can summarize: In one- and two dimensional systems, the expectation value
h2 (x)i interpolates smoothly between the mean-field limits above and below Tc .
The susceptibility 1 (0) does the same thing. Thereby it remains isotropic for all
decreasing T .
Consider now the cases of 3 and 4-dimensions for which the only ground states
with real potential v(, ) were found at = 0 and 0a 6= 0. In order to calculate
the susceptibilities at q = 0 we write V in the form
1
v(, ) = 2a ()
2

(19.113)

where
N

Nm2
N
1
24 () = 2 +
+ S
4 2 log 2
2g
g
4

and
23 () =

N 2 Nm2

+
+N S
3 3/2
2g
g
3

(19.114)

in 4 and 3 dimensions, respectively. We have recorded the dimension as a subscript


on (). Let us now form the derivatives
Va = a +

1 2

D
() a ,
2 a


Va b = ab + {[a b

(19.115)

1
+ (a b)] ()a b } + 2a () a b . (19.116)
2


We may now use the equation of state 2 = 2 () and its derivative a = ()a
to simplify (19.115) and write it as
V a b

a b
a b
= ab +
= ab
()
2c

"

a b
1
+

+
2c .
2

c
()

(19.117)

Thus we can identify the transverse susceptibility as


1

(0) = .

(19.118)

and the longitudinal one


k1 (0)

"

"

1
1
2
+

=
2

+
().
c
1/ ()
()

(19.119)

Actually, we could have obtained the result (19.116) from our earlier equations
(19.48) where the full inverse Green function was derived. Inserting there
m20 + = ,

() (),

(19.120)

19.6 Correlation Functions for Large N

1115

we immediately verify (19.117).


For spontaneously broken symmetry, vanishes and we find
1

(0) = 0

(19.121)

just as in mean-field theory, since this is a rigorous consequence of the NambuGoldstone theorem. For the longitudinal component we obtain

1
k (0) = 2 (0)/ (0).

(19.122)

Inserting (19.114 ) we have in 3 and 4 dimensions


1
k (0) = 0.

(19.123)

This, contrary to the mean field case where k1 (0) = 2m2 > 0, the longitudinal
susceptibility diverges. At first sight we might be led to conclude that also the mode
in the longitudinal direction has become massless due to fluctuations. It will turn
out soon that this conclusion would be false. The nature of this singularity will be
studied in more detail in the next section (Hugenholtz-Pines Theorem).

19.6

Correlation Functions for Large N

We have argued in Chapter 14 that the set of all vertex functions can be obtained
from a finite number of tree diagrams which follow from the effective action [0 +
, 0 + ] considered as a functional of , [see the equations after (6.194)]. Let
us therefore set
0 = m20 + 0 .

(19.124)

Inserting into the action (19.61) = 0 + , we expand




1
1
1
i
()2 0 2 2 + NTr log 2 0
2
2
2
2
Z

Z
2
N 
N
N
dD x +
dD x2 +
0 m20
0 m20 .
(19.125)
+
4g0
2g0
4g0

[, 0 + ] =

The value of 0 or M0 is determined by solving


[0a , 0 ] = 0

(19.126)

This eliminates all graphs with a line ending in the vacuum (tadpole graphs). The
same rules hold for m2 < 0 in one and two dimensions where there is no spontaneous
symmetry breakdown.
Consider now the non-trivial cases of m2 < 0 in three and four dimensions. Again
we can remove the tadpoles by going to the extremum. Contrary to (19.126),
this lies now at
0 = 0 = 0

(19.127)

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

1116

with a non-zero value of given by


2a = (0) =

N 2
m.
g0 0

(19.128)

Thus we may write the effective action as


[, 0 + ] =

i
1
1 h
(19.129)
(a )2 2a (0a )2
2
2



n
X
i
N
1 N 4
1
2
+ m
+ (2) i Tr
2
2

n 4g 0
n=3

This action is free of linear terms in if instead of the fields a one uses the shifted
fields a = a 0a . The components of a parallel to 0a call them k , In order
to find the eigenmodes we have to diagonalize the matrix, written for euclidean
momenta,
q 2 0
0 (2) (q)
where 0

(19.130)

20a . Let us write down (2) explicitly as


"

(19.131)

1
dD k
.
D
2
(2) k (k + q)2

(19.132)

1
1
(q) = N
+ I(q)
2
g
(2)

with
I(q) =

In D dimensions, this was calculated in (5.135):


Z 1

1
dD x
I(q) = S
D (2 /2) (/2)
.
2
0
[q 2 x(1 x)]/2

(19.133)

Consider first the case of 3 dimensions. Then S


3 = 1/2 2 and the integral is finite
I(q) =

1
Z1 D
1
1/2

d
x
[x(1

x)]
=
.
8 (q 2 )1/2 0
8 q2

(19.134)

Using this we find for the diagonalized modes


q + g/8
+ qg/8 2m2 )
gq 2
=
.
N (m2 qg/16 q 2 /2)

Gkd kd =
Gd d

q (q 2

(19.135)
(19.136)

In the limit g 0, the first mode becomes the mean-field longitudinal particle of
mass 2m2 with the second mode disappearing.

19.6 Correlation Functions for Large N

1117

We are now in a position to understand the divergence of the longitudinal susceptibility for q 2 0. From (19.135) we see that the propagator has a square root
singularity in q 2 at this point. This does not correspond to another zero-mass particle. It is generated by the possibility of the kd particle to decay into two massless
Nambu-Goldstone bosons if fluctuations are taken into account. As a matter of fact,
the d mode does not have any proper particle nature: If we return from euclidean
to physical values of q, the mean field pole at q 2 = 2m2 , is no longer visible but
hides in the second sheet of a complex q 2 plane. This is cut from q 2 = 0 to q 2 = .
The physical region lies right above the cut, in accordance with the usual i prescription for singularities (actually, due to the i prescription, the cut has to be
displaced slightly down by i). The pole can only be reached one by passing the cut
from above into the second sheet at
g
q 2 =
16

i.e.,

2m2

g
+
16

2

(19.137)

g 2
g 2
g
2m2
q 2 = 2m2 2
i
.
16
8
16
There is another pole at the complex conjugate position above the axis but this is far
away from the physical region above the cut: It can only be reached by circulating
anticlockwise around the branch point and going into the second sheet from below.
Let us now look at the four-dimensional situation. Here I(q) becomes with
S
4 = 1/8 2 [see (11.175) and (11.171)]


1
I(q) = 2
8

q2
1 1 1
+ + log 2 .
2 2

(19.138)

The 1/-singularity is absorbed in a renormalization of 1/g0 as in (19.71), and we


have
"
!#
N 1
1
q2
(2)
(q) =
+
1 log 2
(19.139)
2 g 16 2

such that the diagonalized propagators become


G

kd kd

1
1/g + (1 log q 2 /2 ) /16 2
=
2 m2 /g + (q 2 /2) [1/g + (1 q 2 /2) /16 2 ]

(19.140)

q2
1
N m2 /g + (q 2 /2) [1/g + (1 q 2 /2 ) /16 2]

(19.141)

Gd d =

Again we observe the cut in the q 2 -plane, starting at q 2 = 0, which is now logarithmic
rather than of the square root type due to the increased dimensionality of space time.
In addition there is a conjugate pair of poles one near +2m2 but hidden in the second
sheet [see Fig. 8.1]. There is, however, an important difference with respect to the
previous case. There is another pole in both propagators which cannot correspond
to a physical particle since it lies at a large positive value of q 2 which corresponds
to a particle moving with a velocity faster than light.

1118

19 Exactly Solvable O(N )-Symmetric 4 -Theory for Large N

Notes and References


For more details see the paper by
S. Coleman, R. Jackiw, D. Politzer, Phys. Rev. D 10, 2491 (1974).
The particular citations in this chapter refer to:
[1] The functional Gaussian Integral is the Hubbard-Stratonovich transformation
(13.98). See
R.L. Stratonovich, Sov. Phys. Dokl. 2, 416 (1958);
J. Hubbard, Phys. Rev. Letters 3, 77 (1959);
B. M
uhlschlegel, J. Math. Phys. , 3, 522 (1962);
J. Langer, Phys. Rev. A 134, 553 (1964);
T.M. Rice, Phys. Rev. A 140 1889 (1965); J. Math. Phys. 8, 1581 (1967);
A.V. Svidzinskij, Teor. Mat. Fiz. 9, 273 (1971);
D. Sherrington, J. Phys. C 4, 401 (1971).
[2] K. Wilson, Phys. Rev. D 7, 2911 (1973).
[3] D. Mermin and H. Wagner, Phys. Rev. Lett. 22, 1133 (1966);
S. Coleman, Comm. Math. Phys. 31, 259 (1973).
[4] For more details see
H. Kleinert, Gauge Fields in Condensed Matter , Vol. I Superflow and Vortex
Lines, World Scientific, Singapore 1989, pp. 1744 (klnt.de/b1)
[5] J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 (1973);
V. Ambegaokar and S. Teitel, Phys. Rev. B 19, 1667 (1979).

Any intelligent fool can make things bigger, more complex, and more violent.
It takes a touch of geniusand a lot of courageto move in the opposite direction
E. F. Schumacher (19111977)

20
Non-Linear -Model
Another model which is exactly solvable in a limit N and illustrates some
typical properties of quantum field theories is the so-called non-linear -model .
It has its origin in statistical mechanics where it arises as a limit of the classical
Heisenberg spin model of ferromagnetism. There it is known under the name of
spherical model. The nonlinear -model yields a complementary approach to the
critical phenomena described previously by the O(N)-symmetric 4 -theory.

20.1

Definition of Classical Heisenberg Model

The model consists of a fluctuating field na (x) of unit vectors with N components
na (x) (a = 1, . . . , N),

n2a (x) = 1.

(20.1)

We shall discuss here the euclidean formulation in which the model is described by
a partition function
Z=

#
#
"
dN 1 na (x)
1 Z D
2
exp
d x[na (x)] .
SN
2g

"
Y Z
x

(20.2)

The constant SN is the surface of a sphere in NRdimensions covered by the directional


integral dD na . This ensures a unit integral dN 1 na /SN = 1. Explicitly, SN =
2 N/2 /(N/2) [recall (11.126)]. This model can be thought of as arising from the
euclidean action of a previous O(N)-symmetric 4 -theory in Chapter ??,
A=

1
m2 2 g  2 2
d x (a )2 +

+
2
2 a 4 a
D

"

(20.3)

by going to the limit of very large negative m2 . For negative m2 , the O(N)-symmetry
is spontaneously broken, and the N initially degenerate
fluctuation modes decomq
2
pose into a massive fluctuation of the size a of the field (radial fluctuations
in field space) and N 1 fluctuation modes of the direction of the field a (azimuthal
fluctuations in field space) described by the unit vectors
q

na a / 2a .
1119

(20.4)

20 Non-Linear -Model

1120

These are the 0Goldstone bosons arising whenever there is a spontaneous breakdown of a continuous symmetry (see Section 17.2). In the limit m2 , the size
fluctuations are completely frozen out. The action can be written approximately as
A

m2

(m2 )2 1 2
+ || (na )2 .
d x
4g
2
D

"

(20.5)

Dropping the first term with the constant condensation energy, we arrive at the
O(N) non-linear model (20.3) with a coupling constant
1
g=
.
(20.6)
||2

Readers familiar with models of statistical mechanics will recognize the intimate
relation of the model (20.2) with the so-called classical O(N)-symmetric Heisenberg
model of ferromagnetism. It is defined on a lattice, where its energy is written as
a sum over nearest-neighbor interactions between local D-dimensional spin vectors
Si , which are conventionally normalized to unit length:
J X
Si Sj with J > 0 and S2i = 1.
(20.7)
E=
2 {i,j}
If Si were not normali The index pairs i, j run over all nearest neighbor pairs.
The spin vectors Si fluctuate and represent the order field. The expectation value
M hSi i is called the magnetization. If M is nonzero, the system is said to exhibit
a spontaneous magnetization. This happens only at low temperatures. Above a
critical temperature TcMF, the system is normal with M = 0.
A special case of this model is the famous Ising model where the direction of the
vector Si is restricted to a single axis, pointing parallel or antiparallel to it. The
phase transition of magnetic systems with strong anisotropy can be described by this
model. Then Si can be replaced by a scalar field with positive and negative signs,
and the symmetry which is spontaneously broken in the low-temperature phase is
the reflection symmetry S(x) S(x).
Let us denote the lattice points by a vector x on a simple cubic lattice, i.e., we
let x take the values
x(m1 ,m1 ,...,md ) D
(20.8)
i=1 mi i,

where i are the basis vectors of a D-dimensional hypercubic lattice, and mi are
integer numbers 0, 1, 2, . . . .
The partition function associated with the generalized Heisenberg model is
Z=

eE/kB T ,

(20.9)

spin
configurations

where the sum runs over all possible spin configurations. The unit spin vectors Si
are now identified with the unit vector fields n(x) in (20.1). The sum over products
P
{i,j} ni nj in the energy (20.7) can then be rewritten as
X

{i,j}

ni nj =

{i,j}

[ni (nj ni ) + 1] = 2

XX
x

{n(x)[n(x + i) n(x)]) + 1}

1121

20.1 Definition of Classical Heisenberg Model

X
x

X
i

[n(x + i) n(x)] 2D

(20.10)

Now we introduce lattice gradients


1
[n(x + i) n(x)] ,
a
1
[n(x) n(x i)] ,
i n(x)
a
i n(x)

(20.11)
(20.12)

where a is the lattice spacing, and rewrite (20.10) as


X

{i,j}

ni nj =

X
x

[i n(x)] 2D .

(20.13)

With this, the partition function (20.9) takes the form


Z=

"

J X X 2
n(x) exp
a [i n(x)]2 2D
2kB T x
i


Y Z
x

!#

(20.14)

We can use the lattice version of partial integration1


XX
x

[i f (x)] g(x) =

XX
x

f (x)i g(x),

(20.15)

which is valid for all lattice functions with periodic boundary conditions, to express
(20.13) in terms of a lattice version of the Laplace operator i i , where repeated
lattice unit vectors are summed. Then
X
x

X
i

a [i n(x)] 2D =

Xh
x

a2 n(x)i i n(x) + 2D .

(20.16)

Ignoring the irrelevant constant term x 2D, we can then write the sum over all spin
configurations in the partition function (20.9) as a product of integrals over a unit
sphere at each lattice point
Z=

Y Z
x

Ja2 X
n(x) exp
n(x)i i n(x) .
2kB T x


(20.17)

For small a, the sums (20.16) have continuum limit


2D

a
1

2D
d x [na (x)] 2
a
D

2D

= a

2D
d x na (x) na (x) + 2 . (20.18)
a
D

For the handling of lattice gradients see Section 2.2 in the textbook H. Kleinert, Path Integrals
in Quantum Mechanics, Statistics and Polymer Physics, World Scientific Publishing Co., Singapore
1995.

20 Non-Linear -Model

1122

After dropping the irrelevant D/a2 term, the Boltzmann factor in the partition
function (20.17) agrees with the exponential in the partition function (20.2) of the
non-linear -model with a coupling constant
g=

kB T D2
a
.
J

(20.19)

In these models, g grows with the temperature, and because of the significance of the
model for statistical mechanics we shall find it convenient to replace the prefactor 1/g
in (20.2) by a reduced inverse temperature 1/T , discussing the model as a function
of the reduced temperature rather than coupling strength. As we see from (20.19),
this reduced temperature corresponds to the true one if the true one is measured in
units of J/aD2 .
As mentioned in the beginning of this chapter, the limit N of this model
is referred to as the spherical model.

20.2

Spherical Model

The N -limit of the Heisenberg model was first solved by Berlin and Kac in
1952.2 The solution of the model is quite simply. We liberate na from its constraint
n2a = 1, allowing each component to fluctuate between and . The constraint
is enforced by an extra functional integral over a Lagrange multiplier (x). Thus
we write the lattice partition function as
Z=

"
Y Z

d(x)
4iT

dN n (x)
a

SN

1
exp
2g

d x (na ) +

n2a

)
i

.(20.20)

In this form, the integrals over the components na are Gaussian, so that they can
be done, leaving a functional integral over only (x):
Z=

"
Y Z
x

h
i
d(x)
N
1
exp Tr log 2 + (x) +
4iT
2
2T


d x(x) .

(20.21)

In the limit N , this is the partition function of the spherical model. In this
limit, the first term in the exponent grows so large that the fluctuations become
frozen at the extremum of the action. The limit is non-trivial if we allow g 1 to
carry a factor N. Thus we set
g t/N.

(20.22)

Now we can go to large N, and find a partition function


Z
eF/t ,
N

T.H. Berlin and M. Kac, Phys. Rev. 86 , 82 (1952).

(20.23)

20.3 Free Energy and Gap Equation in D > 2 Dimensions

1123

with the free energy per O(N)-degree of freedom and per temperature unit
h
i
F
1
1

Tr log 2 + (x)
t N 2
t


dD x(x) ,

(20.24)

to be evaluated at the extremum where the functional derivative F/(x) vanishes.


Explicitly, this condition reads
1
= [2 + (x)]1 (x, x).
t

(20.25)

Each solution i (x) of this equation must be found, and the partition function is
given by the sum over the Boltzmann factors (20.23) for all of them. In general,
these solutions are hard to find. One solution, however, is straightforward: that
with a constant (x) , for which the free energy (20.26) becomes
1
F

D
tL N 2

"Z



1
dD p
2
log
p
+

D
(2)
t

dD x ,

(20.26)

where LD is the spatial volume of the system, and the extremality condition (20.25)
turns into the so-called gap equation:
1
1 Z dD k
=
.
D
2
t
(2) k +

(20.27)

This solution will now be discussed.

20.3

Free Energy and Gap Equation in D > 2 Dimensions

Using formulas (11.134) and (11.137), the integrals in (20.26) and (20.27) can be
done right-away in D dimensions with the result
F
D D/2
1
1
D

D
tLD N 2 D
2
2
2t


(20.28)

for the free energy and


1
D D/21
1
D
1

= S
D
t
2
2
2


(20.29)

for the gap equation, where S


D = SD /(2)D = 2/(4)D/2 (D/2) from Eq. (11.129).
It is easy to check that the free energy (20.28) is extremal for satisfying (20.29).
In order to discuss the physical consequences of these equations, consider first
the case of two dimensions. Letting D 2 approach zero, the equations become
F
1 1 1

1+/2

+ O(),
D
tLD N
2 1 + /2
2t
1
1
=

SD /2 + O().
t

(20.30)
(20.31)

20 Non-Linear -Model

1124

Note that we use for the deviations from D = 2 dimensions, in contrast to for
th deviations from D = 4.
The gap equation has a solution. only for negative t. To avoid the infinities in the
limit 0, we introduce an arbitrary mass scale and a renormalized temperature
tR via the defining equation for tR :

1
1

S
D
t
tR

(20.32)

Then we can take the limit 0 and find


!

F
1


SD log 2 1
D
N
tL
4

2tR
1

1
=

SD log 2 .
tR
2

(20.33)
(20.34)

satisfying the equation


The renormalized gap equation is solved for =

= e2/tR SD .
2

(20.35)

For this value, the free energy has a maximum (see Fig. 20.1). This corresonds to
F
TN

Maximum

Figure 20.1 Free energy as a function of for D = 2 (schematically). The gap equation
of the maximum.
determines the position

a minimum in the -integrals in (20.21) which runs along the imaginary direction.
Only the immediate neighborhood of this minimum contributes to the integral.
The value of at the minimum is a parameter characterizing the solution, having
the dimension of a square mass. From the role which plays in Eq. (20.20) and
in the trace of the logarithm of the fluctuations (20.21) we see that is the square
mass of the na -fluctuations. This implies that even though we had started out with
a model describing N 1 massless Goldstone bosons, the interacting
system is not

massless but all N fields have acquired a nonzero mass M = . What we witness

20.3 Free Energy and Gap Equation in D > 2 Dimensions

1125

here is the spontaneous restoration of the O(N)-symmetry due to the violence of the
directional fluctuations in two dimensions.
It is instructive to consider also the case of more than two dimensions and treat
it in a way which has a smooth limit for D 2 0. If is arbitrary, the free
energy in D 2 + dimensions becomes
F ()
c 1+/2

,
D
tL
2 1 + /2 2t

(20.36)

where
S
D
D
D
c
1
2
2
2


1
/2
= S
D
.
sin (/2)

(20.37)

For small this behaves like


c

1
+ O().
2

(20.38)

To have a smooth limit 0 -limit of the free energy, we write

1 c
F ()
=
D
tL
2 1 + /2
2

!/2

1+

where tR is the renormalized temperature defined by

2tR

c .

t
tR

20.3.1

(20.39)

(20.40)

High-Temperature Phase

At high temperatures, where tR is large, the free energy (20.36) has a maximum, as
of the maximum is determined by the
can be seen in Fig. 20.2. The position =

0.5

1.5

2.5

-0.25
-0.5
-0.75
-1

tR > tcR

F
tLD
tR < tcR

-1.25
-1.5

Figure 20.2 Free energy as a function of for D = 2 (schematically). The gap equation
of the maximum.
determines the position

20 Non-Linear -Model

1126
unrenormalized gap equation as
1
/2
= c
t

(20.41)

For 2 > > 0, i.e., between two and four dimensions, there is a solution for the bare
parameter t < 0. Equation (20.40) implies that in this regime the renormalization
temperature is larger than a certain critical value tcR given by
tcR = c1 > 0.

(20.42)

In terms of tR , Eq. (20.40) reads


!

1
1
1
=
c ,
t
tR tR

(20.43)

as a function of tR > tc :
and Eq. (20.43) gives the gap
R
c
= 2 1 tR

tR

2/

(20.44)

= 0, as can be seen in Fig. 20.2.


For tR < tcR , the maximum lies at =
Note that in order to make the functional integral of the theory (20.20) welldefined for tR > tcR where t < 0, the na -integrations have to be taken along the
imaginary field axis. We shall see later in Section 20.5 that this apparently unphysical condition is an artifact of the analytic continuation in the dimension D.

20.3.2

Low-Temperature Phase

In the low-temperature regime


tR < tcR ,

(20.45)

where the unrenormalized temperature is positive, the gap equation (20.41) has no
solution. The reason for this is an incorrect treatment of the fluctuations at zero
momentum. In order to see what went wrong in the above calculations we may
proceed in two ways. One is inspired by the treatment in Subsection 2.14.3 of the
ideal Bose gas at a fixed particle number N. There we observed that the equation
for the particle number containing a momentum integral over the Bose distribution
could be performed only down to a certain temperature, where the chemical potential
first hits zero (from the negative side). For lower temperatures, the state of zero
momentum accumulates a macroscopic number of states, the so-called condensate.
To account for its presence, it was essential to treat the system in a finite volume,
where the momenta have to be summed rather than being integrated over. Then,
the zero-momentum state can be isolated and treated separately. This is a typical
solution of an infrared problem of a field system. Let us do the same thing here.

20.3 Free Energy and Gap Equation in D > 2 Dimensions

1127

If the system is enclosed in a box of volume LD , we calculate the free energy


(20.24) as


F
1
1 X

2
=
log

+
log
k
+

.
D
D
D
tL
2L
2L k6=0
2t

(20.46)

which
The first term is responsible for generating a local maximum at a real =
is a minimum for the -fluctuations along the imaginary axis. The maximum lies
at
1 X 1
1 1
1
+ D
.
=
D
2

L L k6=0 k +
t

(20.47)

If the system has more than two dimensions, the sum over 1/(k2 + ) with k 6= 0 can
R
be replaced by a phase space integral dD k LD /(2)D (k2 + ). The relative error
is only of order 1/L2 , which can be calculated exactly using a D-dimensional version of the Euler-Maclaurin approximationEuler-Maclaurin approximation to sum
in Eq. (7.674). Thus the gap equation (20.41) reads, more precisely,
1
1
/2 .
= D c
t
L

(20.48)

as before in
For t < 0, this equation has for large volume LD the same solution
on the right-hand side can be ignored. For t > 0,
(20.41), since the first term 1/LD
however, this term makes an essential difference. It gives rise to a nonzero solution
which is vanishingly small in the limit of infinite volume LD . Here the second
,
term in (20.48) can be neglected with respect to the first term, so that for positive
t:
t .

LD

(20.49)

The general solution for both finite and infinite volumes LD is illustrated in Fig. 20.3.
In the limit of infinite volume, the solution is
= 0,

> 0,

t 0,
t < 0.

(20.50)

There is a phase transition at t = 0. For t > 0, all modes become massive and the
O(N) symmetry is restored. For t < 0, the original N 1 massless Goldstone modes
survive.
What does this imply for the renormalized quantities? Using (20.42) and (20.43),
we find from (20.48) the properly renormalized gap equation correcting the previous
Eq. (20.44):

tc
tc
1 R = RD +
tR
2
L

!/2

(20.51)

20 Non-Linear -Model

1128
2
1.5
1
0.5
-1

-0.5

0.5

1
2 t

1.5

Figure 20.3 Solution of the gap equation (20.48) for = 1 and large volume LD . For
L , the right-hand part of the curve approaches more and more the abscissa.

For tR tcR and large volume LD , we can neglect the first term on the right-hand
side, and (20.51) coincides with (20.44). For tR < tcR , on the other hand, we can
neglect the second term on the right-hand side, and find for the spontaneously
of the na -fluctuations:
generated square mass
tR
tR

(L) L2
tcR

!1

(20.52)

This almost zero part of the gap in the low-temperature phase is pictured by the
goes
right-hand branch in Fig. 20.3. In the thermodynamic limit of infinite volume,
to zero. In this limit, the system has a phase transition at the critical temperature
tcR > 0, where the symmetry changes as described above. The statements for t < 0
and t > 0 hold now for tR > tcR and tR < tcR , respectively. This is, of course, the same
transition found earlier in the O(N)-symmetric 4 -theory, only in quite a different
description.
is always nonzero, and there exists no
In a finite volume, the square mass
phase in which the na -fluctuations are massless. The spontaneous breakdown of
O(N)-symmetry is always restored by fluctuations.

20.4

Approaching the Critical Point

It is instructive to study the way in which the mass vanishes as tR approaches tcR
from above. From (20.44) we know that
c
= 2 1 tR

tR

2/

(20.53)

is the square mass of the na fluctuations, this implies that the correlation
Since
functions of the na fields fall of exponentially with
hna (x)na (0)i
e|x|/
x

(20.54)

1129

20.5 Physical Property of Bare Temperature

2 is the correlation length. As tR approaches the critical temperature,


where =
goes to infinity according to the power law
1

(tR tcR )

(20.55)

1
= .

(20.56)

with

This number is called the critical exponent of the correlation length. The power
behavior of as a function of temperature is shown in Fig. 20.4. A similar power
behavior will be obtained later, in Chapter 21, for finite N as well, using renormalization group techniques, although for finite N it will no longer be possible to find
an exact exact result.

1
(tR tcR )1/

tcR

tR

Figure 20.4 Temperature behavior of the correlation length.

20.5

Physical Property of Bare Temperature

The reader may rightfully wonder whether the entire treatment is consistent. We
have observed above that the phase in which the symmetry has been restored possesses a negative bare temperature. This is not the sign of t which the model was
originally defined with. Only for field integrations dna along the imaginary field
direction does the partition function make sense. Fortunately, this apparent inconsistency is not intrinsic to the theory, but merely a mathematical artefact of the
dimensional regularization of the divergent integral. A more physical regularization
would have proceeded via a cutoff in momentum space. Let us verify how this
would change the range of the bare temperature where the phase transition takes
place at a positive t.

20 Non-Linear -Model

1130

In order to be specific, consider first the cases of D = 2 and 3 dimensions. Using


the integrals
Z

|k|<



d2 k
2
log
k
+

(2)2

|k|<



d3 k
2
log
k
+

(2)3




i
1 h 2
+ log 2 + 2 log
4
 

i
1 h 2
log 2 1 + log 2 + 1 log ,
4
(20.57)
(
)


3
1
2 3

3
2
2
log + + 3 2 arctan
6 2
3


 

2
3
1
+ 3 2 ,
3 log 2
(20.58)
2
6
3

the free energy (20.24) becomes


 

i
F

1 1 h 2
2
2
log

1
+
log

+
1

log

=
,
tL2
2 4  
2t



2
1 1
F
3/2
2
3
.
+ 3
=
log
3
2
tL
2 6
3
2t

(20.59)
(20.60)

In writing these expressions, we have assumed all momenta and masses to be dimensionless quantities. For proper expressions, all these quantities should be divided by
some standard mass , which we set equal to one.
The dimensionally regularized expression (20.36), on the other hand, reads for
dimensions D = 2 + with small > 0, and for D = 3, respectively:
F
tL2 N

1 1 2
1

log 2 ,
4 2
4 2

2t

(20.61)

F
tL3 N

1 1 2 3

2 .
4 2 3
2t

(20.62)

For D = 2 + , the positively divergent cutoff expression (log 2 + 1) /8 in the


second term of (20.60) is replaced by the negatively divergent /4 in the first
term of (20.61). For D = 3, dimensional regularized expression in (20.62) shows no
trace of the infinite linearly divergent term /4 2 in (20.60).
Thus the inverse temperature 1/t in the dimensionally regularized expressions
must be thought of as being an unphysical version of the original inverse temperature
1/t in the initial bare energy, from which the constants [(log 2 + 1) + 2/]/4 for
D = 2, or /2 2 for D = 3, have been subtracted. Since these are very large
quantities, there is no wonder that the dimensionally regularized discussion yields
the phase with restored symmetry at a negative unphysical t.
The analytic continuation is taking advantage of the inobservability of the bare
value of t in a continuous field theory. In any system in which there are well-defined
physical mechanisms which regularize the short-wavelength fluctuations (for example lattice structures) and in which the bare temperature t is measurable by microscopic methods, the dimensional regularization cannot be used directly to properly

1131

20.6 Spherical Model on Lattice

represent the physical circumstances of the bare theory. This must be kept in mind
when applying this method. This is, for example, the case in O(N)-lattice models
of statistical mechanics to which the present non-linear -model is related via the
continuum limit. This will be discussed in the next section.

20.6

Spherical Model on Lattice

In order to illustrate the point made in the last section it is instructive to go through
the discussion once more in the lattice formulation of the spherical model. Then the
lattice spacing a provides us with a natural short-distance cutoff, so that there is no
need for introducing a momentum space cutoff , nor dimensional regularization.
On a simple hypercubic lattice, the partition function (20.17) reads
Z=

!
#
#
"
i
o
d(x)a2 Y Z dD na (x)
a2 X nh
2
2
exp
(i n) + n ) ,
4i
SD
2 x
x
(20.63)

"
Y Z
x

where the gradient term has been brought to the lattice form (20.14) without the
irrelevant term 2D. The dimensionless parameter

1
J
=
T
kB T

(20.64)

is called the stiffness of the directional fluctuations. The free energy per lattice site
is given by
1
f =
2

D
X
dD kaD
log
N
[2 2 cos(ki a)] + a2 a2 .
D
(2)
/a
i=1

/a

(20.65)

where a2 D
i=1 [2 2 cos (ki a)] is the Fourier transform of the lattice Laplacian

i i . Extremizing this with respect to yields the gap equation


P

= NvaD2 (0),

(20.66)

where
vaD2 m2 (x)

/a

/a

dD k aD
eikx
P
2 2
(2)D D
i=1 [2 2 cos(ki a)] + a m

(20.67)

is the dimensionless lattice Green function of mass m in D dimensions. For small ,


i.e., high temperatures, the gap equation has a solution > 0 and the fluctuations
are massive. The solution exists up to a critical c given by
c = Nv0D (0).

(20.68)

For D = 2, the momentum integral diverges at the origin: v02 (0) = . This implies
c = 0, so that a nonzero exists for all temperatures T larger than zero. Here the
approximation (20.68) becomes useless, and a special discussion is necessary [7].

20 Non-Linear -Model

1132

The Green function (20.67) is the dimensionless lattice version of the Yukawa
potential in D dimensions
VmD2 (x)

dD k
eikx
.
(2)D k2 + m2

(20.69)

The lattice version has the Fourier representation


vaD2 m2 (x)

"
Y Z
i

/a

/a

eikx
d(aki )
.
PD
2 2
2
i=1 (2 2 cos aki ) + a m
#

The denominator can be rewritten as


multiple integral
vaD2 m2 (0) =

R
0

ds es[

s(a2 m2 +2D)

dse

PD

i=1

"
D Z
Y

i=1

(22 cos aki )+a2 m2 ]

(20.70)
, leading to the

di 2s cos i
.
e
2

(20.71)

The integrations over ki can now easily be performed, and we obtain the integral
representation
Z
2 2
D
dses(a m +2D) [I0 (2s)]D ,
(20.72)
va2 m2 (0) =
0

where I0 (2s) is the modified Bessel function. Integrating this numerically, we find
for D = 3, 4, . . . , . . . the values shown in Table 20.1.3 A power series expansion of

Table 20.1 Values of the lattice Yukawa potential vlD2 (0) of reduced mass l2 at the origin
for different dimensions and l2 . The lower entries show the approximate values from the
hopping expansion (20.74).

D
3
4

v0D (0)
0.2527
0.2171
0.1549
0.1496

v1D (0)
0.1710
0.1691
0.1271
0.1265

v2D (0)
0.1410
0.1407
0.1105
0.1104

v3D (0)
0.1214
0.1214
0.0983
0.0983

v4D (0)
0.1071
0.1071
0.0888
0.0888

the Dth power of the modified Bessel function in (20.72),


[I0 (2s)]D = 1 + Ds2 + D(2D 1)

s3
s4
+ D(6D 2 9D + 4) + . . . ,
4
36

(20.73)

leads to the so-called hopping expansion for vaD2 (0):


vaD2 (0) =
3

2D
6D(2D 1) 20D(6D 2 9 D + 4)
1
+
+
+
+O(D 9 ),
2
2
3
2
5
2
7
2D + a (2D + a ) (2D + a )
(2D + a )
(20.74)

See Tables on p.178 and p. 241 of the textbook H. Kleinert, Gauge Fields in Condensed Matter,
Vol. I, World Scientific, Singapore, 1989, and Eq. (6A.41) on p. 239.

1133

20.6 Spherical Model on Lattice

which converges rapidly for large D, and yields for D = 3, 4 to the approximate
values shown in the lower entries of Table 20.1. They lie quite close to the exact
values in the upper part.
The lattice potential at the origin vaD2 (0) in the gap equation (20.66) is always
smaller than the massless potential v0D (0). A nonzero value for can therefore
only be found for sufficiently small values of the stiffness , i.e., for sufficiently high
temperatures T [recall (20.64)]. The temperature Tc at which the gap equation
(20.66) has the solution = 0 determines the Curie point. Thus we have
Tc =

J
,
c kB

(20.75)

where c is the critical stiffness (20.68):


c = Nv0D (0).

(20.76)

This result, derived for large N, turns out to be amazingly accurate even for rather
small N. As an important example, take N = 2 where the model consists of planar
spins and is referred to as XY-model. For D = 3, it describes accurately the critical
behavior of the superfluid transition in helium near the -transition. From (20.76)
and the value v0D (0) 0.2527 in Table 20.1 we estimate the critical value
c 0.5054.

(20.77)

In Monte-Carlo simulation of this model one obtains, on the other hand,4


c 0.45.

(20.78)

Thus, in three dimensions, we can us the large-N result (20.76) practically for all
N 2.
For D > 2, where c has a finite positive value, the solution for > c requires
special consideration. The situation is similar to what we encountered earlier in
the free Bose liquid. A condensate forms, consisting of Goldstone bosons of zero
momentum. In order to describe this, we enclose the system in a finite box of
volume V and replace the integral over k by the corresponding momentum sum
Z

aD X
dD kaD

.
(2)D
V k

(20.79)

Since the integrand in (20.67) is singular at k = 0, the k = 0 mode has to be treated


separately. We split the right-hand side of the gap equation (20.68) for > c as
follows
D

1=N
4

a
1
1
+
2

V
a

D
k6=0 i=1

1
.
[2 2 cos (ki a)] + a2

See pages 390 and 391 in the textbook in Footnote 3.

(20.80)

20 Non-Linear -Model

1134

This equation can now be solved for > c . For c , the very small prefactor
aD /V guarantees a solution with a very small: a2 is of the order aD /V = aD /LD ,
where L is the linear size of the system. This follows from the observation that the
first nonzero contribution in the sum is of the order
1
1

2
(2a/L) + a2
(2a/L)2 +

(20.81)

aD
LD

For D > 2, one has for large L aD /LD a2 /L2 , such that the second term in the
denominator can be ignored, and the first term in (20.94) is much larger than the
second term, which can be neglected. For > c , the gap equation in a large volume
reads therefore approximately
1N

aD
V a2

(20.82)

What happens if is only slightly larger than c2 ? Since D > 2 the sum on the
right-hand side can again be replaced by an integral and we find the gap equation,
valid for very large V but arbitrary > c :
1 =

aD 1 1
N
aD N N
c
+
v
(0)
=
+ .
0
2
2
V a

V a

(20.83)

Hence the square mass of directional fluctuations is found to be


=

1
1 aD 1
.
2
N V a c

(20.84)

This is positive and infinitely small for a large volume. As a function of temperature
t = 1/, this looks very similar to the plot in (20.3).
Let us look in some detail at the difference between the treatment of the gap
equation on the lattice and that of the dimensionally regularized gap equation in the
continuum field theory. This will allow us to understand the reason for the negative
sign of t in the ordered phase in the continuum model. To be specific, we assume
D = 3. The integral appearing in the gap equation is, in dimensional regularization,
Z


d3 k
D
1
D
1
1
mD2
= S
D
3
2
2
(2) k + m
2
2
2
D=3

  
1
3
1
m
= S
3
2
2
2
1
= m.
4


(20.85)

The corresponding lattice expression is


1
a3

/a

/a

d3 ka3
1 3
1
=
v 2 (0)
(2)3 a2 i [2 2 cos (ki a)] + m2
a m

(20.86)

20.7 Background Field Treatment of Cold Phase

1135

where for small m2


1
1
v0 (0)
m + O(m).
a
4

(20.87)

In the lattice gap equation (20.68), the first term on the right-hand side of (20.94)
goes to infinity in the continuum limit a 0. In the analytic continuation (20.67),
on the other hand, it is absent. It is this term which in the lattice model is responsible
for the fact that positive in both phases, with a positive c > 0 separating the
two phases. In contrast to this, the two phases in the continuum theory exist for
opposite signs of the bare quantity . This curious situation which is caused entirely
by the mathematics of analytic regularization is removed after renormalizing the
theory. Along the renormalized 1/tR -axis, the situation is again analogous to the
more physical one in the lattice model, where need not be renormalized.

20.7

Background Field Treatment of Cold Phase

There is another more elegant way of dealing with the low temperature phase which
does not require the delicate treatment of the k = 0 -mode. It yields a consistent
gap equation for all temperatures without the intermediate consideration of a finite
volume. It is obtained by minimizing the effective action of the model, calculated
by expanding na around a background field. Let us denote this background field
by N a (x), and the fluctuations around it by na (x) = na (x) N a (x). Then the
one-loop effective action becomes
o
NZ D n
d x (Na )2 + (x)Na2 (x)
[N ] =
2t
h
i
N
NZ D
2
+
d x ,
Tr log (x) +
2
2t
a

(20.88)

to be evaluated at the minimum with respect to N a (x) and the maximum with
respect to . In the limit N , the one-loop effective action (1/N) [N a ] is the
exact one.
The new gap equation reads
Na 2 1
+ =

t
t

dD k
1
D
2

(2) k +

(20.89)

and the background field Na satisfies the equation of motion

2 + (x)
Na (x) = 0.

(20.90)

1
This equation shows that determines the correlation length of the Na field.
The ground state is obtained for a constant N a field, which satisfies
a = 0.
N

(20.91)

20 Non-Linear -Model

1136

= 0. In the first case, the gap equation reduces


This has two solutions, Na = 0 or
to the previous one, which in dimensional regularization had a solution only for
t > 0. In the other phase with t > 0, the gap equation turns into an equation for
the size of Na

Na2 1
+ =
t
t

dD k 1
(2)D k2

(20.92)

The direction of Na is arbitrary and will be chosen by the system at will. This
shows that for t > 0, the ground state displays a spontaneous breakdown of O(N)
symmetry. It display a spontaneous magnetization.
Comparison with Eq. (20.47) shows that the gap equation (20.92) coincides precisely with the previous gap equation on the lattice (20.94) for the low temperature
we
phase, with Na2 /t playing the role of 1/LD . Given the previous solution for ,
can therefore identify for t > 0
t 1
Na2 = D .
L

(20.93)

The quantity Na2 can be considered as an order


parameter of the low temperature
q
2
phase. It specifies the magnetization M Na .
In the framework of dimensional regularization, the integral on the right-hand
side of the gap equation vanishes and the order parameter for tR < tcR has the value
Na2 = 1.

(20.94)

The background vector has the unit lngth of the original fluctuating field na (x).
The fact that Na is of unit length for all t > 0 or tR < tcR is once more an artifact of
dimensional regularization. On a lattice, the magnetization M goes to zero smoothly,
with a power law in c:
M ( c ) .

(20.95)

The power is conventionally denoted by , which should in thes context not be


confused with the inverse temperature . On a lattice, the gap equation (20.92)
reads, for > c ,
1

Na2

N
=

/a

/a

such that

dD kaD
1
,
PD
D
(2)
i=1 [2 2 cos (ki a)]

1 Na2 =

Nv0D (0)
.

(20.96)

(20.97)

From this we extract the power behavior


M=

c
Na2 = 1

!1/2

showing that the critical exponent has the value 1/2.

(20.98)

20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in D Spacetime Dimensions 1137

20.8

Quantum Statistics at Nonzero Temperature

Let us also study quantum fluctuations in the non-linear -model. For this we
assume that one of the spatial coordinates (x1 , . . . , xD ) , for instance xD , is an
imaginary time variable restricted to the interval (0, h
) with = 1/kB T ,
and let the fields be periodic in with period h
. Equivalently, we may think of
this model as a nonlinear -model on an infinitely long spatial strip with periodic
boundary conditions, whose width along the xD -axis is . In this context, we shall
write the D 1 -dimensional purely spatial vectors as fat letters, whereas the Ddimensional spacetime vectors are denoted by slim letters.
Since temperature enters the theory now via the period h
= 1/kB T , we shall
return to the original notation g for the coupling constant as in the partition function
(20.2) [which had been traded for the reduced temperature t in (20.22) and (20.19)]
In the limit N , we can study the effects of temperature exactly. The
partition function is (20.26), and theRgap equation has the same form as in (20.27),
P
except that the momentum integral dpD /2 is replaced by a sum T pD =m over
the Matsubara frequencies
m = 2T m,

m = 0, 1, 2, . . . .

(20.99)

guaranteeing the periodicity of the fields on the interval xD (0, ). The gap
equation (20.27) at finite temperature is therefore
1
= ( 2 + )1 =
g

1
dD1 p X
T
.
D1
2
2
(2)
m m + p +

(20.100)

We have gone to natural units with h


= 1, kB = 1. This equation can be renormalized with the same renormalized coupling constant gR () as in the infinite system.
We add and subtract, on the right-hand side, the T = 0 -limit of the rigth-hand side
and obtain
1

1
dD p
=
D
2
(2) p +

Z
X
1
d
1
dD1 p
. (20.101)
T

D1
2
2
2
(2)
2 + p2 +
m m + p +
!

The left-hand side is the zero-temperature gap equation (20.27).

20.8.1

Two-Dimensional Model

For the subsequent discussion we focus attention on the two-dimensional case. Then
the zero-temperature gap equation can be renormalized in D = 2 + dimensions
using Eq. (20.32). This brings the left-hand side in (20.101) to the form

1
+S
D + S
D 2
g

!/2

1
+ S
2 log 2 .
gR () 2

(20.102)

20 Non-Linear -Model

1138

The sum over m on the right-hand side of Eq. (20.177) can be performed using the
sum formula

X
1
1

T
=
coth
,
(20.103)
2
2
2
2T
m= m +

with p2 + . In the limit T 0, this becomes


Z

1
1
d
=
.
2
2
2 +
2

(20.104)

Subtracting (20.104) from (20.103), and using (20.102), we obtain the temperaturedependent gap equation in two dimensions
Z
1
dp 1

coth
+
log 2 =
1 .
gR () 4

2T
2 2


(20.105)

of this equation
Eliminating the arbitrary mass scale in favor of the solution =
for T = 0, which is
= 2 e4/gR (2 ) ,

(20.106)
this equation can be rewritten in a renormalization group invariant way as

1
log =
4

dp 1
coth
1 .
2 2
2T


(20.107)

The right-hand side is a function of the dimensionless variable T


T /(2T )2

(20.108)

and will be denoted by


Z

dp 1

1
coth
1
S1 (T ).
2 2
2T
2


(20.109)

In terms of this function, the gap equation reads simply [compare with the forthcoming fermionic gap function (23.221)]

log = 2S1 (T ).

(20.110)

The functions S1 (T ) can also be rewritten as follows


S1 (T ) =

dp

coth

2T


1 = 2

1
dp 
1 e/T
.

There exists no critical temperature Tc where vanishes.


The gap equation (20.110) is fast convergent only at low temperature.

(20.111)

20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in D Spacetime Dimensions 1139

For high temperatures, it is better to keep the original sum over m in (20.103).
Inserting this sum into (20.111), and perform the integral over p gives
S1 (T ) = T

m=

1
2

m=

dm
2

dm

1
q

2 +
m

1
.
m2 + T

(20.112)

Taking the term with m = 0 out of the sum, the right-hand side can be rearranged
as follows:
Z

X
X
1
1
1
1
1

S1 (T ) = +

+
dm q

. (20.113)
2
m
2 T m=1
0
m + T
m2 + 2T
1 m
!

The last two terms diverge. The two divergences cancel each other. To see this we
truncate the sum and the integral at some finite large integer value m = M. Then
the two terms have the limits
M
X

and

1
1
,
log M + + O
M
m=1 m

(20.114)

M
2M
1
arcsin
,
log + O
M
T
T

(20.115)

respectively. Combining them, we obtain the alternative expression for S1 (T )


1
T e2
1

S1 (T ) = S1 (T ) +
,
+ log
2
4
2 T

(20.116)

where S1 (T ) denotes the convergent sum


S1 (T )

m=1

1
1
2

.
m
m + T

(20.117)

Inserting this into (20.116), we arrive at the desired alternative form for the gap
equation which converges fast at high temperatures:
T
1
log = S1 (T ) + ,
T
2 T

(20.118)

/4.
where T e
It is straight-forward to calculate S1 (T ) as a function of T , and find from this
at
T (T ) as well as = T 4 2 T . Plotting (T ) we see that starts out with
T = 0, and grows rapidly to infinity.

20 Non-Linear -Model

1140

The gap equation is obtained by extremizing the effective potential as a function


of . The effective potential is given by
1
1
v() =
N
2

dp X

2
T
log(m
+ p2 + ) .
2 m
2g

(20.119)

We split this again into the zero-temperature equation


1 Z d2 p

1
v0 () =
log(p2 + )
2
N
2 (2)
2g
!

=
log 1 ,
8

(20.120)

and a finite temperature correction




dp
1
T v() = T
log 1 e/T
N
2
!
Z
X Z dm
1 dp
2
=
T
log(m
+ p2 + ). (20.121)

2 2
2

m
Z

The right-hand side depends only on T , and will be denoted by (/2)S0 (T ), i.e.

1
T v() =
S0 (T )
N
2

(20.122)

The integral over p in (20.120) can be performed in analytic regularization. For


D = 2, this amounts to the sequence of steps
Z

dp
2
log(m
+ p2 + )
2

dp
2
exp[ (m
+ p2 + )]
2

d
2
3/2 exp[ (m
+ )]
2
0
(1/2) q 2
m + .
=
2
=

(20.123)

Hence
!
Z

dm q 2
m +
S0 (T ) =
T

2
m=

1
=
2T

m=

!q

dm

m2 + T .

(20.124)

Note that S0 (T ) has the derivative


1
d
T S0 (T ) = S1 (T ),
dT
2

(20.125)

20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in D Spacetime Dimensions 1141

so that a differentiation of the potential


!

log 1 +
v() =
S0 (T )
N
8
2

(20.126)

leads properly to the gap equation (20.110).


At high temperatures, it is useful to do the same manipulations with S0 (T ) as
with S1 (T ), and rewriting it as follows
T e2
1
1
log
1 ,
S0 (T ) = S0 (T ) + +
4
4
2 T
!

(20.127)

where
!

q
X

2
2 +
S0 (T ) =
T
m
m
m=1
2m
!
q

1 X
T
2
.
=
m + T m
T m=1
2m

(20.128)

Then the total potential becomes

1

S0 (T ).
v() = log + +
N
4
T
2
4 T

(20.129)

In the close neighborhood of two dimensions, we set D = 2 + > 2, and split


the zero-temperature potential v0 () conveniently as follows:
1

1
v0 () = c
/2 =
c c
/2 ,
N
2g
2 1 + /2
2g
2
2
1 + /2
(20.130)
with c of Eq. (20.37) with the small- behavior (20.38).
The finite-temperature correction can be written as follows:
"

1
D/2
T vT (T ) =
aD S0 (T ),
N
2

(20.131)

with
S0 (T ) =

a1
D

X
dm
2 1 Z dD1 p
2
log(m
+ p2 + ), (20.132)
T

D/2
D1

2 (2)
2
m=
!

where the constant aD is chosen to have a convenient form for S0 (T ) at all D:


aD =

1 D

2
2

1
(4)(D1)/2

(20.133)

20 Non-Linear -Model

1142

with a2 =R 1, a4 = 1/6. By rewriting log a as the analytically regularization


integral 0 (d / )e a , Eq. (20.132) becomes
S0 (T ) =

a1
D

D/2

m=
D1

dm
2


h
i
d
p
2
2
exp

(
+
)
.
exp

p
m
(2)D1


(20.134)

The integral over p can now be performed, with the result


Z



1
dD1 p
2
=
exp

p
.
(2)D1
(4 )(D1)/2

(20.135)

Using now the integral formula


Z

d (D1)/2 E
1D
E (D1)/2 ,

e
=

2


(20.136)

we find
!
Z

(D1)/2
X
dm  2
2 1

S0 (T ) =
T

m
D/2 2
2
m=

2T D/2

m=

dm (m2 + T )(D1)/2

(20.137)

For D = 2, this reduces to the previous sum (20.124).


In going from (20.132) to (20.137), we have of course rederived formula (11.134)
in the form
Z

20.8.2

1
1 D
dD1 q
2
2
log(
+
q
)
=

m
(2)D1
(4)(D1)/2
2
2


 q

2
m

D1

(20.138)

Four-Dimensional Model

For D = 4, we obtain
!
Z

3
X

dm q 2

S0 (T ) =
T

m
2
2
m=

1
=
22T

m=

!q

dm

m2 + T .

(20.139)

By analogy with the two-dimensional expression (20.127), this can be processed


further to a convergent sum
(q
)

3
1 X
3
3 2T
3
2

S0 (T ) = 2
m + T m T m
.
T m=1
2
8m

(20.140)

20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in D Spacetime Dimensions 1143

20.8.3

Any Dimension

In any dimension D, the expression aD S0 (D ) in temperature correction (20.131) to


the potential which generalizes (20.127) is derived as follows: First we perform the
integral over m in (20.137), and rewrite aD S0 (T ) as
1

a S0 (T ) = aD

D/2

"

(D1)/2
T

+2

(m + T )

(D1)/2

m=1

+ c

2
,
D
(20.141)

the last term coming from the integral over m in (20.137) [recall (20.37)]. Then we
expand the second term in powers of T as
1

a S0 (T ) = aD

D/2

"

(D1)/2
T

+2

m=1 k=0

(D 1)/2
k

mD12k kT

+ c

2
.
D

(20.142)

Performing the sum over m in the curly brackets gives


"

(D1)/2
T

+2

k=0

(D 1)/2
k

(2k + 1

D)kT

(20.143)

the term with k = D/2


has a singularity for D close
For an even dimension D = D,
to 2, 4, . . . . This singularity cancels a corresponding one in c .
say D = D
+ . Then there are two
Let D be close to the even dimension D,
singular terms in aD S0 (T ):

aD S0 (T )

sg

1 D


+

1 D
1

1
D/2
2! 2
=

(1

)
D1 D/2


T

2
2
1
T D
4
+1
+
2
2 2
!

D
1
.
(20.144)
D2 1

2
2 D+
2 4


Expanding (1 ) = (1 ) / + . . . , the 1/-singularity in the first term is


given by

1 D
!
+

2
2
1
1
1
()D/2
1 D
!
!

  = D1
D1

D
2
2
1
D
D
4
4
+1

2
2
2 2
!

(20.145)

In the second term, we write

1 D


2 2
2

= ()

D/2

(1 + /2) 2
 (1 /2) =
+ /2
D/2


(1)D/2
!
,

sin
+
2
2
2
(20.146)

20 Non-Linear -Model

1144

and see that the 1/-singularities cancel each other. The finite- independent contribution is obtained by expanding
(k + ) = (k)[1 + (k)],

(20.147)

such that the first 1/ -singularity is accompanied by

1 D
1+
log T

2
2
2

"

1 D
+
+
2
2

2 log 4
2
 

#)

(20.148)

The second singularity has a residue

2
D

log 4
1+
2
D
2
"

Since

1 D
+

2
2

the difference is

1 D

2
2

!#)

= cot

(20.149)

D
= 0,
2

X 1

T e2 D/2
.

log

2
4
k
k=1

(20.150)

Thus, altogether, we obtain for an even numbers of dimension D = 2, 4, 6, . . . the


finite sums
S0 (T ) =

(D1)/2
T
D/2
2T

+2

k=0,6=D/2

(D 1)/2
k

(2k + 1 D)kT + S0L (T ),

(20.151)

where
S0L (T )

D/2
a1
D ()

T
1
log

D 4 D2
4e2
4 1 +
2


D/2

1
.
k=1 k
X

(20.152)

It is now easy to split (20.151) into


1
( ) + S L ( )
S0 (T ) = S0 (T ) + + S0 (T ) + S
0
T
T
0
2 T

(20.153)

with the convergent sum


S0 (T ) =

(D 1)/2
k

(2k + 1 D)kT ,

(20.154)

D/21

(D 1)/2
k

(2k + 1 D)kT .

(20.155)

D/2
T k=D/2+1

and the finite sum


( ) =
S
0
T

1
D/2

k=0

20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in D Spacetime Dimensions 1145
z
Inserting the the divergent representation
for the -function at negative
m=1 m
z, this is formally equal to the divergent sum

( ) =
S
0 T

D/2


X

D1

m=0

D1
+
T mD3 + . . .
2

(20.156)

D/21

where the last omitted term in parentheses is proportional to T


/m. The convergent sum (20.154) is obviously equal a convergent power series expansion of the
convergent infinite series
S0 (T ) =

1
D/2

2T

q
X

m2 + T

D1

m=1

mD1

D1
T mD3 + . . .
2

, (20.157)

where as many powers of T are subtracted as needed for convergence. These are the
generalizations of the two- and four-dimensional expressions (20.128) and (20.140),
respectively. The expression (20.153) written in this way converges fast for high
temperatures.
Let us now derive an expression which converges fast at low temperature T . For
this it is most convenient to use the D-dimensional generalization of Eq. (20.121)
d
X
dD1 p
m
2
log(m
+ p2 + )
T

D1
(2)
2

m
Z


dD1p
/T
= T
.
(20.158)
log
1

e
(2)D1

1
1
T v() =
N
2

and expand the logarithm in powers of e/T . This gives

X
D/2
1 Z dpD1 m p2 +/T
aD S0 (T ) = T
e
2
m

(2)D1
m=1

(20.159)


1
1 X
1
2
S
D1 D/2
dssD2em s +1/T .
2
m
0
T m=1

We now use the integral formula

K (z) =

 

z
2

1
Z

1/2

2
 2 
dss2 s2 + 1
ez s +1 ,
1 0
+
2

 

(20.160)

and see that

q
X
2D/2
1
S0 (T ) = 2 
T . (20.161)
  

D/2 KD/2 2 m
1 D
1 m=1

2 m
T

2
2
2

The same expression will be found once more in an exactly solvable fermionic model

in Eq. (23.273), except for a fermionic alternating sign ()m1


.

20 Non-Linear -Model

1146
For D = 2, the sum (20.161) reduces to
S0 (T ) = 2

q
1
K1 2 T m

2 m
T
m=1

(20.162)

we can take the limit of large T , using K1 (z) 1/z, and find
z0

S0 (T )

1 X
1
1
1
= 2 () =
.
2
2
2 T m=1
m

2 T
12T

(20.163)

This yields the free energy of black-body radiation in two dimensions

1
T v()
T 2 .
T
N
6

(20.164)

The same result could have been obtained from the divergent expression (20.119)
using the generalized Euler-Maclaurin formula [regularized by analytic continuation
as in the formal calculation of the Casimir effect in Eq. (7.708)]
1
1
T v() =
N
2
=

1
T
2

dp
T
2

"

m=

m=

dm
2
log(m
+ p2 )
2

2 = 2T 2 (1) = T 2 .
m
6

(20.165)

In the second step we have used the integral formula (11.134) in the form
Z

dD1 q
1
1 D
2
log(m
+ q2) =

D1
(D1)/2
(2)
(4)
2
2


 q

2
m

D1

(20.166)

In four dimensions, we have


S0 (T ) = 6

q
1
T
K
2
m


2 2
m=1

2 m
T


(20.167)

which becomes for large T [using K2 (z) 2/z 2 ]


z0

S0 (T )
12
The potential

m=1

12 4 4
12T 4
1
4 = 2 (4) = 2 T 90 .
2 m
T

(20.168)

2
2
2
1
4
T v =
S0 =
S
=
T
(20.169)
2
N
2
12 2
90
is the analog of the free energy density f of black-body radiation. It is related to
the internal energy density. u by f = 31 u.
Recall the analog of the 0 Stefan-Boltzmann law for the free energy density f of
hot (or massless) bosons. In this limit, f has a pure power behavior of T n , and the

20.8 Quantum Statistics at Nonzero Temperature Nonlinear -Model in D Spacetime Dimensions 1147

associated entropy density s = f /T is related to f by a factor 4/T . For the


internal energy density u E/V = f + T S and for the specific heat at constant
volume cV = u/T |V , the factors are 1 n and (1 n)n/T , respectively.
The original Stefan-Boltzmann law for black body radiation is obtained from
this by accounting for the two polarization degrees of freedom with an extra factor
2 so that
1
2T 4
T v = f = u = 2
.
(20.170)
3
90
In physical units, this becomes
f = 2

2 4
T ,
3c

(20.171)

where is the Stefan-Boltzmann constant


=

4
2 kB
g
5
3 2 5.67 10
sec K4
60h c

(20.172)

The same result could, of course have been obtained directly from (20.132), which
leads to the four-dimensional version of Eq. (20.165). Using (20.166) for D = 3 we
obtain
1
1
T v() =
N
2
=

d3 p
T
(2)3

"

m=

dm
2
log(m
+ p2 )
2

q
2
3
1 X
1 2
4 4
4
2 =
T
(1/2)
T
(3)
=
T
. (20.173)

m
2 m= (4)3/2 3
3
90

recalling Eq. (7.709).


At arbitrary D, we use the small-z behavior
 

1 2
K (z)
2 z

(),

in (20.161) to find
T 0

S0 (T )

1D/2
a1

D 2

D
2

1
D (D)
2 T

1
2
D
T D (D).

D/2
D/2
aD
2


(20.174)

Dropping the prefactor factor 2/D/2 aD gives the finite-temperature contribution


T v()/N to the free energy density in D dimensions.
D
T D (D).
T v()/N
D/2

2
T

(20.175)

The same result could, of course, have been obtained directly from (20.132), using

m=

D1
m
= (2T )D12(1 D),

(20.176)

20 Non-Linear -Model

1148
and the well-known identity for Riemanns -functions5

z
1 z
(1 z) = 1/2z
(z).

2 2
2


20.9

 

(20.177)

Criteria for Onset of Fluctuations in Ginzburg-Landau


Theories

The understanding of the nonlinear -model permits us to improve our understanding of the phase transition in 4 -theories with O(N)-symmetry discussed in Chapter 27. We shall base the discussion on the euclidean version of the Lagrangian
(17.6), which is then called Ginzburg-Landau Hamiltonian [1]:
1
m2 2 g  2 2
.
(a )2 +
+
2
2 a 4 a
In the associated partition function
H(a , a ) =

Z=

Da e

dD xH/kB T

(20.178)

(20.179)

a phase transition takes place at a temperature Tc which lies always below the mean
field temperature TcMF , where the mass term in (20.178) changes its sign. It is
possible to give a rough estimate of the shift T TcMF Tc which is caused by
fluctuations.
After an obvious renormalization of field and mass, the Ginzburg-Landau energy
density in D dimensions may be written as
i2
gh
1
H(a ,a) = D 2a2 [a (x)]2 + 2a (x)+ 2a (x) .
2a
2


(20.180)

From here on we use natural units with kB TcMF = 1. The fields have zero engineering
dimension, a denotes some microscopic length scale of the system, usually the size of
atoms or molecules, and g is some interaction strength. The parameter specifies
the
zero-temperature coherence length of the system in units of a as being 0 =
a/ 2. This can vary greatly from system to system. In superconductors, for
example, it can lie anywhere between a few thousand, and less than ten in hightemperature superconductors.
We shall here only be concerned with the destruction of the ordered state which
lies below the critical temperature where < 0: There the fields in the energy
density fluctuate around an ordered ground state with a constant vector ha i
a hi Na Na in field space, whose direction vector
Na breaks
spontaq
q
2
neously the O(N)-symmetry, and whose magnitude is = a = /g, where
the energy density is minimal, fluctuating around the condensation energy density H0 q
= H(a , 0) = 2 /4gaD . The temperature-dependent coherence length
= a/ 2| | describes the range of the size fluctuations of the order field.
5

See, for instance, I.S. Gradsteyn and I.M. Ryznik, Table of Integrals, Series, and Products,
Academic Press, New York, 1980, formula 9.635.4.

20.9 Criteria for Onset of Fluctuations in Ginzburg-Landau Theories

20.9.1

1149

Ginzburgs Criterion

The magnitude of the fluctuations is estimated by assuming the field to live in


patches on a simple cubic lattice of spacing l = l, choosing eventually a spacing
parameter between l = 1 and 2 to ensure the independence of the patches. In the lowtemperature ordered phase with < 0, the small fluctuations (x) = (x) of
the size of the order field (x) around the minimum of (20.180) have an Hamiltonian
Hsz (a ,a)

o
1 n 22
2
2

a
[(x)]
+
2|
|
(x)
.
2aD

(20.181)

Their size is therefore given by


2

h[(x)] i =

aD2 1
aD
dD k
=
(2)D 2 a2 k 2 + 2| |
lD2 2

dD q
1
,
D
2
(2) q + l2

(20.182)

where q is the dimensionless reduced momentum l k. Inserting l , the right-hand


side can be rewritten as
Z
l2D
dD q
1
D/21
(2|
|)
,
D
D
2

(2) q + l2

(20.183)

The relative size of the fluctuations can therefore be written as


h[(x)]2 i
= 2l2D (2| |)D/22 gD vlD2 (0),
2

(20.184)

where vlD2 (0) is the lattice Yukawa potential of reduced mass l of Eq. (20.67). Its size
is calculated most easily with the help of the integral formula (20.72). Mean-field
behavior breaks down if (20.184) is of the order unity, which happens at the reduced
Ginzburg temperature
|G | [l2D K vlD2 (0)]2/(4D) ,

D < 4.

(20.185)

where
K 2D/21 g/D ,

(20.186)

i.e., at a Ginzburg temperature TG TcMF(1 |G |). Ginzburg, in his original paper


[2], estimated l 1 and v1D (0) in three dimensions by an integral carried up to
|p| = :
v1D (0)

d3 p
1
1
2
3
2
(2) p + 1
2

This led him to the estimate

dp p2

p2

1
1

.
+1
4

(20.187)

1 g2
.
(20.188)
8 2 3
In old-fashioned type-II superconductors, |G | can be as small as 108 [3], which
explains why conventional superconductors are well described by mean-field theory.
|G |

20 Non-Linear -Model

1150

In modern high-Tc superconductors, on the other hand, Ginzburgs estimate leads


to |G | 0.01 [4], such that critical exponents become observable.
The length parameter l is a free parameter this criterium and Landaus estimate
based on the assumtion l 1 corresponds to the physical situation that the order
field (x is properly defined only up to a length scale of the order of the coherence
length. In an ordinary superconductor, this is indeed the case. There the order field
describes the Cooper pairs of electrons whose wave function extends over a coherence
length, such that the cutoff in all momentum integrals would be of this order. In
modern superconductors, where the phase transition occurs at higher critical temperature of the order of 100 K, however, the Cooper pairs could have a much smaller
diameter than the coherence length.6 In this case, l would be considerably smaller
than unity. If the Cooper pairs are bound very strongly, another effect appears: The
phase transition is caused by quantum fluctuations. In this limit, the Cooper pairs
form an almost free gas of almost point-like bosons which undergo Bose-Einstein
condensation of the type described in Subsection 2.14.3. The relevant length scale
is then the De Broglie wavelength of thermal motion
=

2h
,
2MkB T

(20.189)

where M is the mass of the Cooper pairs.


For D > 4, the right-hand side in (20.184) decreases when approaching the
critical point, so only mean-field behavior is observed. If D = 4 lies only slightly
below four, the right-hand side of (20.184) behaves like | |/2 , implying a good
mean-field description until | | is extremely small.

20.9.2

Kleinerts Criterion

If a 4 -theory has only a single real field, Ginzburgs criterion gives a reliable estimate
at which temperature fluctuations become important. In an O(N)-symmetric system
with Nambu-Goldstone modes, however, it grossly underestimated the temperature
shift T of the transition. A better estimate is based on the observation that the
kinetic term defines a second, completely independent, energy scale of the system.
To identify it, we split the fields according to size and direction in O(N) field space
as a = na , n2a = 1. The directions na describe the long-range fluctuations of
the Goldstone modes. Sufficiently far from the critical regime, we may neglect the
gradient term of the size (x), and approximate the energy density by
H(, na ) =
6

1
g
2a2 2 (x) [na (x)]2 + 2 (x)+ 4 (x) .
D
2a
2


See E. Babaev and H. Kleinert, Phys. Rev. B 59 , 12083 (1999) (cond-mat/9907138), and
references therein.

20.9 Criteria for Onset of Fluctuations in Ginzburg-Landau Theories

1151

The fluctuations of the Goldstone modes are controlled by the gradient term whose
magnitude depends on the size of at the minimum of the potential. The gradient
energy density is

(20.190)
Hna (na ) = D2 [na (x)]2 ,
2l
with
= () =

l
a

!D2

2 =

1 lD2
D lD2
=
.
2(2| |)D/22 g
K | |D/22

(20.191)

This is the second energy scale. It measures how much energy is spent when reversing
the direction vector na over the distance l . It is the continuous version of the
stiffness of the directional field defined for a lattice model in Eq. (20.64).
From the discussion of the spherical model in Section 20.6 we know that directional fluctuations disorder a system if the bending stiffness drops below a certain
critical value cr . For the O(N)-symmetry with large N, this critical value was found
to be, on a cubic lattice, in three, four, and large-D dimensions [see Table 20.1 and
the hopping expansion (20.74)]
cr = Nv0D (0) N 0.2527 ,

N 0.1549,

N/2D,

(20.192)

respectively. We have remarked below Eq. (20.76) that, although these values were
derived only for large N, Monte Carlo simulations show they can be trusted already
for D = 3 and N = 2 to within about 10%.
The simulations are done by putting the Heisenberg model on a lattice of unit
spacing, so that the energy density for N = 2 takes the XY-model form
HXY (na )

=1,...,D

[1 cos (x)],

(20.193)

where i denotes the lattice gradient in the ith coordinate direction, and
arctan n2 /n1 . Since the quality of the approximation increases with N and D, we
can trust Eq. (20.192) to within 10% for all N and D 3. his accuracy will be
sufficient for the criterion to be derived here.
The critical stiffness can, incidentally, be also estimated by calculating its renormalized version from a sum of an infinite number of terms in a perturbation expansion. Expanding the cosine in the energy (20.193) into a Taylor series, and
calculating the harmonic expectation values of quartic, sextic, etc. terms, we find
in a self-consistent approximation of the Hartree-Fock-Bogoliubov type that the
stiffness has a renormalized value [7] R = e1/2DR . This softens with increasing
temperature 1/, until reaches a critical value cr = e/2D, where R drops to
zero. In D = 3 dimensions, this happens at cr = 0.4530 . . . , a value which is in
MC
excellent agreement with the Monte Carlo number cr
0.45. The prediction of
such sharp drop is true only in two dimensions, as shown by Kosterlitz and Thouless
[6]. For D > 2 it is an artefact of the approximations, and the exact stiffness goes
to zero like |Tc T |(D2) , with a critical exponent 1/2 + (4 D)/10 + . . . .

20 Non-Linear -Model

1152

The estimate for the critical stiffness (20.192) leads now directly to the announced criterion: The phase fluctuations will disorder the system if the stiffness
in Eq. (20.190) drops below the critical value (20.192), which happens at a reduced
temperature
|K | [Nl2D K v0D (0)]2/(4D) ,
D < 4, N 2.
(20.194)
Thus we obtain the important result that

|K | [Nv0D (0)/vlD2 (0)]2/(4D) |G |, D < 4, N 2.

(20.195)

This implies that for all systems with N 2, directional fluctuations destroy the
order before size fluctuations become large. They cause a phase transition below
the Ginzburg temperature, at TK TcMF(1 |K |). For D = 3, and l = (1, 3/2, 2),
the relation becomes |K | N 2 |G | (2.20, 3.48, 5.56). Thus, if the critical regime
is approached in a 4 -theory with a well-formed mean-field regime, the transition
is always initiated by directional fluctuations. In particular, the estimates for the
critical regime of the high-|Tc | superconductors [4] will receive a factor 9.
The dominance of directional fluctuations is, of course, most prominent in the
limit of large N, and it is therefore not surprising that the critical exponents of the
4 -theory and the Heisenberg model have the same 1/N-expansions in any dimension
D > 2, as a pleasant demonstration of the universality of critical phenomena.
By adding to the field energy density H(, na ) the energy density of directional
fluctuations with the field-dependent stiffness = () = D 2 lD2 /(2| |)D/21 we
can study, as in Ref. [8], the combined energy density in the disordered phase where
the symmetry is restored but the average of the size of the order field is nonzero.
The directional fluctuations play a crucial role in pion physics, as pointed out in
Ref. [8] and discussed in detail in Section 23.10.

20.9.3

Experimental Consequences

How do we determine experimentally the fluctuation parameter K to estimate |G |


and
| |? In magnetic systems, one measures the susceptibility tensor AB (k)
R D Kikx
d xe ha (x)b (0)i at wave vector k, and decomposes it into parallel and perpendicular parts as AB (k) = (a b /2 )k (k)+(AB a b /2 ) (k). The mean-field
behavior of these quantities is k (k) aD /(2 a2 k 2 + 2| |) and (k) aD /2 a2 k 2 .
Combining these
at k = 0 with the mean-field behavior of the spontaneous magnetiq
zation = | |/g, and with the temperature-dependent coherence length , we see
that the size of K can immediately be estimated from a plot, versus t T /Tc 1,
of either of the dimensionless experimental quantities
Kexp |t|

2D/2

or
Kexp

|t|2D/2

k2

(k)
,

D2 kB T 2 k0
1 k (0)
,
D kB T 2

(20.196)

(20.197)

20.9 Criteria for Onset of Fluctuations in Ginzburg-Landau Theories

1153

these being written down in physical units. Note that t measures the temperature
distance from the experimental Tc , in contrast to T /TcMF 1. In the mean-field
regime, where t , Kexp is constant, and can be inserted into Eq. (20.194) to find
the temperature TK where directional fluctuations destroy the order.
In superfluid helium we may plot, in analogy to the transverse susceptibility
expression for Kexp , the quantity Kexp |t|2D/2 M 2 kB T / D2h
2 s , where M is the
atomic mass and s the superfluid mass density, which at the mean-field level is
defined by writing the gradient energy (20.190) as e gradient energy (20.190) as
(s /2kB T )(h2 /M 2 )[na (x)]2 . In the critical regime, the three expressions for Kexp
go universally to zero like |t|2D/2 , since |t| , k (0) |t|(2) , k 2 (k)|k0
|t| , 2 |t|(D2+) , s |t|(D2) , with [(N + 2)/2(N + 8)2 ](4 D)2 + . . . .
Experimentally, the superfluid density of helium for D = 3 shows no mean-field
behavior a la Ginzburg-Landau down to T Tc /4, such that the above formulas
cannot properly be applied. Let us nevertheless estimate orders of magnitude of a
would-be mean-field behavior: s / 2| | [9], where = M/a3 is the total mass
density, with a 3.59
A [10]. Then the factor kB Tc at Tc = 2.18 K can be expressed
as kB Tc 2.35h2 /Ma3 [10]. With 0 2
A, we obtain an estimate K 1.2 a/0 2.
Inserting this into Eq. (20.194) and relation (20.195), we obtain for l = 1 and 2
|K | 1, |G | 0.12 and |K | 0.255, |G | 0.03,

(20.198)

respectively. The large size of |K | reflects the bad quality of a mean-field description.
The larger l gives the more physical estimate.

[1] L.D. Landau, J.E.T.P. 7, 627 (1937); V.L. Ginzburg and L.D. Landau,
J.E.T.P. 20, 1064 (1950);
[2] V.L. Ginzburg, Fiz. Twerd. Tela 2, 2031 (1960) [Sov. Phys. Solid State 2, 1824
(1961)]. See also the detailed discussion in Chapter 13 of the textbook L.D.
Landau and E.M. Lifshitz, Statistical Physics, 3rd edition, Pergamon Press,
London, 1968.
[3] See Eq. (3.24) on p. 315 of the textbook in Footnote 3 where g/3
111(Tc /TF )2 , with TF 103 Tc .
[4] C.J. Lobb, Phys. Rev. B 36, 3930 (1987).
[5] D.B. Murray and B.G. Nickel, (unpublished). R. Guida and J. Zinn-Justin,
J. Phys. A 31, 8130 (1998); H. Kleinert, FU-Berlin preprint 1999 (condmat/9906107).
[6] J.M. Kosterlitz and D.J. Thouless, J. Phys. C 6, 1181 (1973), Prog. Low Temp.
Phys. B 7, 371 (1978).

1154

20 Non-Linear -Model

[7] See Section 7.8. in the textbook in Footnote. 3.


[8] H. Kleinert and B. Van den Bossche, Phys. Lett. B 474, 336 (2000) (hepph/9907274).
[9] See Fig. 5.3 on p. 428 in the textbook in Ref. 3.
[10] See pp. 256257 in the textbook in Ref. 3.

It was beautiful and simple, as truly great swindles are


O. Henry (18621910)

21
The Renormalization Group
In the previous chapter we have discussed some solvable models which illustrate
an important aspect of the renormalization procedure: When defining renormalized
quantities in a massless theory, we cannot avoid to introduce some mass parameter
. The value of this mass parameter is a matter of choice. If the Lagrangian
density contains a mass m, the role of can be played by m itself. But this is not
always the most convenient choice. Often we want to study the limit m 0 of a
theory. Then it is preferable to renormalize also a massive theory with the help of
an auxiliary mass . As long as m 6= 0, this amounts to an overparametrization
of the theory. This is a useful technical trick to derive the characteristic behavior
of experimental observables for an important set of physical phenomena. These
are the critical phenomena which occur as a function of temperature T in manybody systems as the T tends towards a so-called critical temperature Tc , where the
system undergoes a second-order phase transition. At Tc , fluctuations acquire an
infinitely long range corresponding to the mass m in the fields going to zero. The
mass parameter vanishes near Tc like
2

m = 0

T
1 .
Tc


(21.1)

Suppose that we want to study the temperature behavior of a field theory in the
neighborhood of m2 = 0, and that we have chosen the mass m itself as the mass scale
of renormalization. Then we are able to calculate renormalized Green functions as
long as m2 6= 0. At the limiting point m2 = 0, we have to change the renormalization
procedure and introduce some non-zero mass scale . Obviously, such a procedure
is not well adapted to study the limit m 0. A unified renormalization of massive
and the massless case with a single arbitrary mass scale is preferable.
Certainly, all physical quantities must be independent of the mass parameter
when expressed in terms of the bare quantities. Only the renormalized quantities
depend on .
1155

1156

21.1

21 The Renormalization Group

Example for Redundancy in Parametrization


of Renormalized Theory

As an example, recall the situation in the O(N)-symmetric self-interacting fourfermion theory. In the limit N , the effective potential reads at zero input
fermion mass [recall b from Eq. (23.20)]:
1
1
v() =
N
2

2
2

b 2+ =

g0
2 g0

!2+

(21.2)

This expression is trivially independent of the mass parameter . The necessity


for introducing the mass parameter arises only if we want to go to the limit of
two dimensions with 0. Then we want to define the theory with the help of
a finite renormalized coupling constant, and this cannot be done without . The
renormalized coupling constant will necessarily depend on , as the bare coupling
constant g0 is always held fixed. The -dependence is given by
1
1
=
b .
g(, ) g0

(21.3)

To emphasize the -dependence at fixed g0 , as well as the dependence on , we have


added these arguments to the renormalized coupling g.
In a more general renormalizable theory, all renormalized quantities are functions
of and . Let us also make the dependence on the bare quantities explicit by writing
m = m(m0 , g0 , .)
g = g(m0 , g0 , , ).

(21.4)
(21.5)

The fields are renormalized by a factor


1

= 0 Z 2 (m0 , g0 , , ) .

(21.6)

If we vary the mass parameter at fixed bare quantities, the Green functions remain
the same. The renormalized quantities m, g change with . This implies that they
are not directly experimentally observable. Experimental quantities such as particle
masses and scattering amplitudes cannot depend on the dummy mass parameter .
At 6= 0, these depend only on the initial bare quantities. The main observable of
the O(N)-symmetric four-fermion theory with zero initial mass is the fermion mass
Mf arising from a spontaneous symmetry breakdown. It is given by the value of
at the minimum of the potential v.
Mf =

2/D
g0 b

!1/

(21.7)

In the limit 0, the finite physical mass Mf must always be reproduced by the
constant theory. Since b diverges the bare coupling has to become infinitesimally
small.

21.1 Example for Redundancy in Parametrization of Renormalized Theory

1157

The mass scale is needed to describe the situation in the limit 0 in terms of
finite quantities. To do this, we express (21.7) in terms of the renormalized coupling
g which is defined by
1
1

b ,
(21.8)
g(, )
g0
as follows [using (21.3)]:
Mf
=

2/D
g0 b

!1/

"

1
2
1+
=
D
b g(, )

!#1/

(21.9)

Here we can easily take the limit 0, in which the right-hand side becomes:
(

!#)

"

1
1
2
Mf
1+
= exp
log

D
b g(, )
(
!#)
"

1
1

= exp
1+
log 1

2
b g(, )
!)
(

1
+

exp

2 b g(, ) b g(, )
(
!)
0
1

exp
+
2 g()

(21.10)

where
g() g(, 0).

(21.11)

This equation shows very clearly how the renormalization process leads to an overparametrisation of, the same theory by the parameter . The physical mass Mf
is fixed, and the equation specifies in which way the renormalized coupling g()
depends on to guarantee a certain mass Mf . See Fig. 21.1 for an illustration. This
relation between and g() is typical for a large variety of theories, as we shall see
in the sequel.
In general, the -dependence of the renormalized observables is governed by a
partial differential equation called the Callan-Symanzik equation. This equation
allows us to find curves in the space of renormalized coupling and mass parameters
g, m along which a field theory has identical observables. These curves are called
renormalization group trajectories, , and the movement along them if is changed
will be called renormalization flow.
Since the theory is renormalizable, the relations between observable quantities
must all be finite. The model potential (21.2) illustrates this fact. Inverting relation
(21.7), we express the bare coupling constant g0 in terms of the fermion mass Mf .
After inserting the result into (21.2), we obtain
1
Mf 2 D

v() =

b
N
2
2
Mf
"

! #

(21.12)

1158

21 The Renormalization Group

= Mf exp

1
+
2
g

Figure 21.1 Curves in the (, g)-plane corresponding to the same physical fermion mass
Mf = 1, . . . , 5, and thus to identical physical observables. These curves are the renormalization group trajectories of the O(N )-symmetric four-fermion model in the limit N .

This is a finite relation for all values of , in particular for 0 where it has the
limit
2

1
log
v() =
N
2
Mf
"

.
2

(21.13)

Note that by expressing v() in terms of the physical observable Mf rather than the
finite quantity g(), we have arrived at an expression which is finite and independent
of . This v() will be called a renormalization group invariant version of the
effective potential.
We shall see that an understanding of the renormalization group will permit
us to find an economic way of expressing all renormalized quantitties in such a independent way. Mathematically speaking, the determination of physical oservables
which do not depend on the renormalization scheme will amount to a search for
invariants under the renormalization group.
To develop the theory we shall focus our attention again on the 4 -theory in
D = 4 dimensions, renormalized via an -expansion.

21.2

Renormalization Scheme

In Chapter 11 we have seen how to obtain, for the 4 -theory in four spacetime
dimensions, all finite vertex functions (n) (p, g, m, ) order by order in perturbation
theory. Besides the physical parameters g and m, it was convenient to carry along
an auxiliary mass parameter . The finiteness of the Feynman graphs was achieved
by adding to the renormalized Lagrangian
1
m2 2 g 4
L = ()2

2
2
4!

(21.14)

1159

21.2 Renormalization Scheme

a corresponding set of counterterms, divergent for 0,


m2 m 2 g g 4
1
c
c .
Lc = c ()2
2
2
4!

(21.15)

We have replaced the original renormalized coupling constant g by g . As we


have seen before in Chapter 11, this has the advantage of making all counterterms functions of the new dimensionless renormalized coupling constant g only [call
(11.255)]. The counterterms were calculated at every order in perturbation theory
from the primitively divergent diagrams (divergent means divergent in the limit
0). Renormalizability of the theory implies that no other divergent counterterms are necessary than those which can be written in the same form as the original
Lagrangian as in Eq. (21.15). Then all divergencies can be absorbed in a multiplicative redefinition of field, mass, and coupling constant. The total Lagrangian L + Lc
is
1
m0 2 2 g0 4
2
L0 = L + Lc = (0 )
0 0 ,
2
2
4!

(21.16)

and this may be identified with the initial bare Lagrangian, in which the bare field,
mass, and coupling constant are related to the finite renormalized ones by
0 Z 1/2 ,
m0 2 Z m2 Zm2 ,
g0 Z 2 gZg ,

(21.17)

with the renormalization constants


Z 1 + c (g, ),
Zm2 1 + cm (g, ),
Zg 1 + cg (g, ).

(21.18)

The bare Green functions when calculated in 4 dimensions


G0 (n) (x1 , . . . , xn ; g0, m0 ) = h0|T 0(x1 ) . . . 0 (xn )|0i,

(21.19)

are finite objects as long as 6= 0. When multiplied by the renormalization constants


Z n/2 , they turn into the renormalized Green functions which remain finite in the
limit 0:
G(n) (x1 , . . . , xn ; g, m, ) Z n/2 (g, )G0 (n) (x1 , . . . , xn ; g0, m0 , ).

(21.20)

Correspondingly, the bare vertex functions


(n)

0 (x1 , . . . , xn ; g0 , m0 ,) i

dx1 dxn

(21.21)

G0 1 (x1 , x1 ; g0 , m0 ,) G0 1 (xn , xn ; g0 , m0 ,) G0 (n) (x1 , . . . , xn ; g0 , m0 ,)

1160

21 The Renormalization Group

become finite after being renormalized by the inverse factors Z n/2 (g, ) of (21.20):
(n) (x1 , . . . , xn ; g, m, ) = Z n/2 0 (n) (x1 , . . . , xn ; g0 , m0 , ).

(21.22)

As discussed in Chapter 11, there are different possible methods of fixing the counterterms. The most common choice is based on the normalization conditions
(2)
|p2=2 = 1,
p2
(2) |p2=2 = m2 ,
(4) |symm

point

= .

(21.23)
(21.24)
(21.25)

where the symmetry point is defined by


pi pj =

2
(4ij 1);
4

i, j = 1, . . . , 4.

(21.26)

The minimal subtraction scheme due to t Hooft requires, instead, the cancellation of all 1/, 1/2 , 1/3 , . . . pole terms in the vertex functions. This has the pleasant
property that the resulting counterterms depend only on the coupling constants and
not on the mass m.
In both schemes, the arbitrary mass parameter can be avoided by setting it
equal to m of the fields are massive, but is unavoidable in massless theories.

21.3

The Renormalization Group Equation

As discussed above, physical observables should be independent on choice of ,


and this fundamental property will now be formulated quantitatively. Suppose we
calculate perturbatively the bare n-point vertex function at a fixed bare coupling g0
and mass m0 the bare vertices in 4 dimensions:
(n)

0 (p1 , . . . , pn ; g0 , m0 , ).

(21.27)

The result is trivially independent of the choice of the mass parameter which was
introduced only to define the finite renormalized coupling strength g = g() and
mass m = m(). We therefore have the obvious relation

(n)
(p1 , . . . pn ; g0 , m0 , ) = 0,
0

(21.28)

This obvious relation turns into a nontrivial one for the renormalized vertex functions
n/2

(n) (p1 , . . . , pn ; g, m, ) Z

(n)

(g0 , m0 , )0 (p1 , . . . , pn ; g0 , m0 , ).

(21.29)

If the bare parameters g0 , m0 , are kept fixed, the dependence in (n) can come
from the renormalized quantities g and m whose values epend on the choice of ,

1161

21.4 Calculation of Coefficient Functions from Counter Terms

which is emphasized by writing them as g() and m(). By applying (21.28) to


(21.29) we see from the chain rule of differentiation that


1/2

+ g|g0,m0 , n log Z |g0 ,m0 , + log m2 |g0 ,m0 , m2 m2

(n) (p1 , . . . , pn ; g, m, ) = 0,

(21.30)

where we have written for /, to save space. The right-hand sides of (21.30)
(21.30) can be expressed as finite functions of g, m, and . One defines the functions
= g|g0,m0 , ,
=
m =

1/2
log Z |g0 ,m0 , ,
log m2 |g0 ,m0 , .,

(21.31)
(21.32)
(21.33)

and reexpressed them as functions of g, m, and . With these, Eq.(21.31) becomes


the following differential equation
( + g n + m m2 m2 )(n) (p1 , . . . , pn ; g, m, ) = 0.

(21.34)

This is the celebrated renormalization group equation (RGE) If we would not use
the minimal subtraction scheme but some cutoff in momentum space, then the
parameter in Eqs.(21.27)(21.33) would be replaced by the cutoff . The RGE is
the quantitative formulation of the invariance of the vertex functions (n) under a
change of the auxiliary mass scale , which on the one hand appears explicitly in
the vertex functions (n) (p1 , . . . , pn ; g, m, ), on the other hand in the -dependent
renormalized quantities g = g() and m = m().

21.4

Calculation of Coefficient Functions


from Counter Terms

The solution of a partial differential equation like (21.34) is, in general, awkward,
since , , m may depend on g, m and . It is an important special property of
2
tHoofts renormalization scheme, that the counterterms cg , cm , and c happen to
be independent of the masses m and depend only on the coupling constant g, apart
from . This makes their calculation simple and, moreover, produces coefficient
functions , , m in the RGE (21.34) which depend only on g and . Suppressing
the -argument, we shall therefore write (g), (g), m(g).
Consider first (g). Since Z depends only on g we certainly have
(g) = g|g0,m0 ,

d
1/2
log Z (g).
dg

With the help of (21.31), this can be rewritten as


(g) = (g)

d
1/2
log Z (g).
dg

(21.35)

1162

21 The Renormalization Group

Similarly, using the renormalization equation m2 = m2


0 Zm (g)Z (g) we see that

d
d
log Zm
log Z
m (g) = (g)
dg
dg

(21.36)

or
m (g) 2(g) m (g) = (g)

d
log Zm (g).
dg

(21.37)

Finally, we use
g0 = g Zg (g)Z2(g)

(21.38)

to find
#1

"

g0 |g
d
(g) =
=
log(g Zg Z2 )
g g0 |
dg
g
.
=

1 + g[log(g Zg )] 2g[log(Z )]

(21.39)

or, using (21.35),


(g) =

+ 4(g)
log[g Zg (g)]

(21.40)

All three functions depend only on g.


The formulas can be brought to a more practical shape by making use of the
explicit 1/-expansion structure of the counterterms as functions of g. Then we may
rewrite Eqs. (21.35), (21.37), (21.40) as follows
(g) 1 +

c (g)

=1

[m (g) + 2(g)] 1 +

=1

(g)

[gcg (g)]
=1

X
1
c (g) ,
= (g)
2
=1

cm
(g)

= (g)

= [ + 4(g)]g 1 +

(21.41)


cm
(g) ,

cg (g)

(21.42)

=1

=1

(21.43)

For a renormalizable theory, the functions (g), (g), m (g) have to remain finite
in the limit 0. Thus we can expand these quantities in a power series in with
powers 0 , 1 , 2 . . . . By inserting (21.41) into (21.43) we see that (g) can at most
contain the following powers of :
(g) = 0 (g) + 1 (g).

(21.44)

1163

21.4 Calculation of Coefficient Functions from Counter Terms

When inserting this into (21.41), (21.42), we find (g) and m (g) to be independent
of . Equating the regular parts of the three equations yields
0 + 1 + 1 (cg1 + gcg1 ) = ( + 4)g gcg1
1
=
1 c
2 1

m m 2 = 1 cm
1

(21.45)

which are solved by


1 (g) = g
0 (g) = g 2cg1 (g) + 4g

(21.46)

Thus, amazingly, the three functions (g), (g), m(g) can be expressed in terms of
g-derivatives of the three residues cg1 (g), c1 (g), cm
1 (g) of the simple 1/-poles in the
counterterms:
(g) = g + g 2 cg1 (g) + 4g,
g
(g) = c1 (g) ,
2

m (g) = m (g) 2(g) = gcm


1 (g) .

(21.47)

The finiteness of (g), (g), m(g) at = 0 forces all higher residues of in Eqs.
(21.41)(21.43) to disappear. This implies an infinite set of consistency relations
among the expansion coefficients which are useful for checking the calculations:
0 (gcg ) g(gcg+1) = 4gcg g cg+1 ,
(m

c
2)cm

=
=

(21.48)

gc+1 ,

m
0 cm
+ gc+1 .

0 c

(21.49)
(21.50)

Let us now turn to the actual calculation of (g), (g), m(g) in the O(N)symmetric 4 theory. The renormalization constants were given in Section 11.8.
Replacing g S
D by g, for brevity (so that it becomes equal to our former ) we
obtained [recall Eqs. (11.249)]
3 1
1
15
1
1
Zg = 1 + cg = 1 + g TI + g 2
,
TI 2 2 3TI4
+
2
4

22 4
1
1
(21.51)
Z = 1 + c = 1 g 2 TD3 ,
48




1
1
1
1
1 g2
Zm = 1 + cm = 1 + g TD1 +
3TD1 D2 2 TD3 2 +
.
2

2
The 1/-pole terms are


3
3
cg1 = g TI g 2 TI4 ,
2
4
1
c1 = g 2 TD3 ,
48
1
1
2
cm
= g TD1 g 2 TD3 .
1
2
8



(21.52)

1164

21 The Renormalization Group

Inserting these into (21.47), we find


3
3
(g) 4(g)g = g + g 2 TI g 3TI4 ,
2
2
1
(g) = g 2 TD3 ,
48
g
g2
m (g) 2(g) =
TD1 TD3 .
2
4

(21.53)

The reduced matrix elements T were given in (5.75). For N = 1, all T s are equal
to unity.
Explicitly, we can solve for (g), (g), m(g) as follows [making use of the explicit
reduced matrix elements T calculated in Eqs. (11.234)(11.246).
(g) =
=
(g) =
m (g) =

1
3
TI4 TD3
g + g TI g
2
2
12
N + 8 2 3N + 14 3
g +
g
g ,
6
12
1 N +2
1
,
g 2 TD3 = g 2
48
48 3 

1
5
1
5 N +2
g TD1 g 2 TD3 = g g 2
.
2
24
2
24
3
23

(21.54)
(21.55)
(21.56)

With the help of Eqs. (21.48)(21.50) we check these results for consistency:
0 (gcg1 ) g(gcg2 ) = 4gc1 g gc2 g ,
c1 = 0 c1 gc2 ,

m
(m 2)cm
= 0 cm
1
1 + gc2 .

(21.57)

Inserting (21.54) (21.56), we see that to order g 2, the two sides are indeed equal.
It is instructive to observe the cancellation of 1/n -singularities in Eqs. (21.35)
(21.40) a somewhat more explicitly. Take for instance (g) of Eq. (21.35), and the
five-loop result for the -function to be stated in (21.192) for N = 1:
#

"

17
145 3(3) 4
3
g
+
(g) = g + g 2 g 3 +
2
12
64
2
#
"
3499 4 39(3) 15(5) 5
g +
(21.58)
+

+
768
80
8
2
"
#
764621 1189 4 5 6 7965(3) 45(3)2 987(5) 1323(7) 6
+
g .

+
+
+
+
73728
23040
448
512
32
32
32
The logarithmic derivative of the renormalization constant has the five-loop expansion:
[log Z (g)]

1
3
1
95
65
1
=
g2 + 3 +
g3
g

2
2
24
16
64
32
1152
1536


21.5 Solution of Renormalization Group Equations for Vertex Functions

163
553
3709
4
(3)
(3) 4
9
g

+
+

3
+ 4+
3
2
2
64
768
3072
36864 2880 16
256


13
179
23
+
+

g5
30724 552963 163842
#

"

1165

(21.59)

When forming the product (g) [log Z (g)] to get (g) via (21.35), the only -term
in (g) reduces the highest 1/n singularities in the g n -term of [log Z (g)] by one
power of 1/. For n 2 the resulting terms are still singular, but they are canceled
by the product of the finite g 2 -term in (g) and the singular terms associated with
the lower power g n1 in [log Z (g)], etc..

21.5

Solution of Renormalization Group Equations


for Vertex Functions

The subtraction of divergencies via the -expansion has given us coefficient functions
, , m which depend only on the renormalized coupling constant g(). This makes
it easy to solve the renormalization group equation (21.34). Suppose we have calculated the renormalized vertex functions for some initial mass scale with a renormalized coupling g = g() and mass m = m(). Then the coupling g() = g()
at an arbitrary mass scale = , which we shall denote for brevity by g , is
given by
d
(21.60)
g = (g ).
d
Equation (21.60) can be rewritten as
1
d ln
=
dg
(g )

(21.61)

This is solved with the initial condition


g1 = g.
The result is
ln =

dg
.
(g )

(21.62)

(21.63)

Similarly we have for the -dependence of the renormalized mass m() the equation
m2 ()

d 2
m () = m (g ),
d

(21.64)

which can be rewritten as with the notation m m() = m( ) as


d ln m2
= m (g ),
d ln

(21.65)

and is solved with the initial condition


m1 = m

(21.66)

1166

21 The Renormalization Group

by

m2 Z d
ln 2 =
(21.67)
m (g ).
m

1
Using the functions g and m , the renormalization group equation for the vertex
functions which is initially a partial differential equation can be rephrased as an
ordinary differential equation:
"

n(g ) (n) (pi ; g , m , ) = 0.


d

(21.68)

For brevity, we have written pi collectively for p1 , . . . , pn . Equation (21.68) is immediately integrated as follows
" Z

(n) (pi ; g , m , ) = exp n

d
(g ) (n) (pi ; g, m, ),

(21.69)

which shows the full -dependence of the renormalized vertex functions.


This equation can now be used to study the critical behavior of (n) (pi ; g, m, ),
i.e.,the behavior of (n) as the renormalized mass m goes to zero while the mass scale
at which the theory is defined remains fixed. For this purpose it is useful to take
advantage of the dimensional structure of the vertex functions. In D dimensions, a
field has a (mass) dimension d = D/2 1. A Green function containing n fields
has the dimension d = n(D/2
1). When performing n 1 Fourier transforms of
R
the fields to momentum space, dD xeipx , the dimension becomes nd (n 1)D.
The Fourier transformed two-point function G(2) (p) has dimension 2 for all D.
The vertex function (n) is obtained by multiplying the Fourier transformed Green
function G(n) by n factors G(2) (p)1 . The total mass dimension of (n) is therefore
D nd = D n(D/2 1). Then, since the dimension of pi , g, m, are 1, 0, 1, 1,
respectively, the vertex function (n) must satisfy the trivial identity valid for any
scale parameter ,
n

(n) (pi ; g, m, ) = D 2 (D2) (n) (pi /; g, m/, /) .

(21.70)

Inserting on the right-hand side the left-hand side of (21.69), we obtain

(n)

(pi ; g, m, ) =

Dn(D/21)

"

exp n

(g )
d

(n) (pi /, g , m /, ) .
(21.71)

We now choose = m so that the rescaled mass m / is equal to the mass parameter :
(21.72)
m /|=m = .
Then (21.71) becomes

(n)

(pi ; g, m, ) =

Dn(D/21)
m

"

exp n

(g )
d

(n) (pi /m , gm , , ) .
(21.73)

21.5 Solution of Renormalization Group Equations for Vertex Functions

1167

This equation relates the renormalized vertex functions (n) (pi , g, m, ) of an arbitrary mass to those of a fixed mass equal to the mass parameter at rescaled
momenta pi / and a running coupling constant coupling constant g . Apart from a
Dn(D/21)
trivial overall rescaling factor m
due to the naive dimension, there is also
a nontrivial exponential function.
Explicitly, the parameter m in (21.72) is found by using Eq. (21.67) to calculate
the ratio
m d
m2
m2
=
exp
[m (g )] .
2
2

0
)
(Z
m d
m2
=
exp
[2 m (g )]
2 2

0
)
(Z
m d
[2 m (g )] .
= exp

(Z

(21.74)

Our goal is to study the behavior of the vertex functions on the left-hand side
of (21.73) in the critical region where m 0. On the right-hand side, the mass
dependence resides in the rescaling parameter m .
In many-body systems undergoing a second-order phase transition, the mass m
approaches zero proportional to T Tc . Equation (21.74) specifies how the auxiliary
scale parameter m which rules the scaling relation (21.73) behaves for T Tc .
Near the critical point T = Tc , experimental correlation functions show a simple
scaling behavior. Such a behavior can be found from Eqs. (21.73) and (21.74), if
in the limit m 0 the coupling constant g runs into a fixed point g , at which the
running coupling constant g becomes independent of , i.e.,
"

d g
d

= 0,

(21.75)

g=g

Assuming that m
m (g ) < 2, which will be true in the present field theory, the
integrand in Eq. (21.74) is singular at = 0, and the asymptotic behavior of m for
m 0 can immediately be found:

m2 m0
exp
2

(Z

2m
[2

(g
)]
=

.
m
m

(21.76)

we see that m goes to zero for m 0 with the power law


m

m2
2

)
!1/(2m

(21.77)

In applications to second-order phase transitions we shall identify m2 /2 with the


relative distance from the critical temperature
T Tc
m2

,
2

Tc

(21.78)

1168

21 The Renormalization Group

and write (21.77) as

m 1/(2m )

(21.79)

In this limit, the exponential prefactor in (21.73) has the power behavior
"

exp n

m
1

d
(g )

m 0

/(2 )
m

(21.80)

where (g ). The n-point vertex function behaves therefore like


m0

Dn(D/21)n
(n) (pi ; g, m, ) m
(n) (pi /m ; g , , ).

(21.81)

For the two-point vertex function this implies a momentum dependence


m0

(p; g, m, ) p22 f(p/m ).

(21.82)

with some function f(). When going over to the two-point function in x-space
G(x; g, m, ) =

1
dD p ipx
e
D
(2)
(p; g, m, )

(21.83)

this amounts to an x-dependence


1

m0

G(x; g, m, )

g(x/m ) =
D2+2

g x(m2 /2 )1/(2m ) .
D2+2

(21.84)

This is of a general form discovered by Kadanoff fpr connected correlation functions:


G(2)
c (r) =

f (r/t )
,
r D2+

(21.85)

With the identification (21.78, we see that the critical exponent in (21.85) is given
by
1
.
(21.86)
=

2 m
For T Tc , the length scale = diverges like
(t) = 0 |t| .

(21.87)

By comparing the power behavior at the critical point in (21.84) with that in (21.85)
we identify the critical exponent as
= 2 :

(21.88)

21.6 Renormalization Group for Effective Action and Potential

21.6

1169

Renormalization Group for Effective Action


and Potential

The same methods can be applied to find the one parameter set of identical theories
for the full effective action [] for different values of the mass parameter . Consider
the power series expansion of the effective action in terms of vertex functions
[; g, m, ] =

1
n=0 n!

dD p
d D p1
.
.
.
(p1 ) (pn )(n) (p1 , . . . , pn ; g, m, ).
(2)D
(2)D
(21.89)

and apply the RGE in Eq.(21.34) to every coefficient. This effectively replaces the
factor n in front of (g) as follows
n

dD p
(p)
.
D
(2)
(p)

(21.90)

Hence we find immediately the RGE


"

+ (g)g (g)

dD p

(p)
+ m (g)m2 m [; g, m, ] = 0.
D
(2)
(p)
(21.91)
#

A corresponding result holds for the effective potential


v() = LD [; g, m, ]|(x)=const= ,

(21.92)

where L is the size of the D-dimensional box under consideration. The effective
potential satisfies the differential equation
"

+ (g)g (g)
+ m (g)m2 m v(; g, m, ) = 0.

(21.93)

Because of its relevance to physical applications we shall first solve only this latter
equation along the lines of Eqs.(21.63)(21.68). We introduce an additional running
field strength () satisfying the differential equation
1
d
() = (g()),
() d

(21.94)

with the initial condition


(1) = .

(21.95)

This is solved by
"

()
= exp

d
(g ) .

(21.96)

1170

21 The Renormalization Group

Using , we have the identity along the renormalization trajectory


v(; g , m , ) v(; g, m, ).

(21.97)

Note that there is no prefactor in contrast to (21.73). All m-dependence is contained


in m and () via (21.96). In the limit 0 the field behaves like
()

(21.98)

Since [] is dimensionless and v() is related to [] by (21.92), the dimensional


transformation corresponding to (21.70) becomes simply
v(; g, m, ) = D v(/ D/21 ; g, m/, /)

(21.99)

Combining this with (21.97), we find


v(; g, m, ) = D v( / D/21 , g, m /, ).

(21.100)

At the mass-dependent value m of Eq. (21.72), we obtain the analog of (21.73) for
the effective potential
D/21
D
; gm , , ).
v(; g, m, ) = m
v(m /m

(21.101)

In the critical regime, where 0, Eq. (21.96) becomes


m

m
,

(21.102)

and the effective potential has the power behavior


m0

D
+D/21
v(; g, m, ) m
v(/m
; g , , ).

(21.103)

this goes over onto For applications to systems below Tc it is most convenient to
consider, instead of [; g, m, ], the vertex functions in the presence of an external
magentization. They are obtained by expanding [; g, m, ] functionally around
(x) 0 :

(n)

n
[; g, m, ]|0 .
(p1 , . . . , pn ; 0 , m, )
(x1 ) . . . (xn )

(21.104)

This gives

(n)

(p1 , . . . , pn ; 0 , g, m, ) =

n+n

0
(n+n ) (p1 , . . . pn , 0, . . . , 0; g, m, ),

n!
n =0

(21.105)

where the zeros later the arents p1 , . . . , pn indicate that there are n more momentum
arents pn+1 , . . . , pn+n which have been set zero since each constant field 0 has a

1171

21.7 Approach to Scaling

momentum dependence (D) (p)0 . Thus the renormalization group equation for the
vertex function at a non-zero field, (n) (p1 , . . . , pn ; 0 , , ), can be obtained from

those at zero field (n+n ) (p1 , . . . pn , pn+1 , . . . , pn+n ; ) with the last n momenta set
equal to zero, i.e., from
h

+ (g)g (n + n )(g) + m (g)m2 m (n+n ) (p1 , . . . , pn , 0, . . . 0; g, m, ) = 0.

(21.106)

Inserting these equations into (21.105), we obtain the RGE


"

+ (g)g (g) 0
0

+ m (g)m m (n) (p1 , . . . , pn ; 0 , g, m, ) = 0.


(21.107)

When treated as above, this leads to the scaling relation


(n) (pi ; 0 , g, m, ) = en

R
1

d
(g )

(n) (pi ; 0 , g , m , )

(21.108)

Together with the trivial scaling relation


(n) (pi ; 0 , g, m, ) = Dn(D/21) (n) (pi /; 0 / D/21 , g, m/, /), (21.109)
we find
R

R m

(n) (pi ; 0 , g, m, ) = Dn(D/21) en 1 (g )


(n) (pi /; 0 / D/21 , g , m /, ),

(21.110)

which becomes at = m of Eq. (21.72):

(n)

(pi ; 0 , g, m, ) =

d
(g )
Dn(D/21) n 1

m
e
D/21
, gm , , ),
(n) (pi /m ; 0 m /m

(21.111)

and thus, near the critical point,


Dn(
(n) (pi ; 0 , g, m, ) = m

21.7

+D/21)

+D/21
(n) (pi /m ; 0 /m
, g , , ). (21.112)

Approach to Scaling

In Eq. (21.84) we have extracted from the scaling relation (21.81) for the two-point
vertex function Kadanoffs scaling law (21.85). From this we read off the critical

exponents = 1/(2 m
) for the temperature nehavior of the coherence length and

the exponent = 2 for the critical power behavior of the green function.
There exists a further important critical exponent which governs the approach
to the scaling law (21.84). In order to find this we expand the right-hand side of
Eq. (21.73) around g and write

(n)

(pi ; g, m, ) =

Dn(D/21)
m

(n)

"

exp n

(pi /m ; gm , , )

(g )
d

(n) (pi /m , g , , ) .
(21.113)

1172

21 The Renormalization Group

Using Eq. (21.7), we obtain the correction factor


C (n) (pi ; m , g , , ) = 1 + [gm g ] (n) (pi /m ; g , , ) + . . . .

(21.114)

where (n) is th dimensionless function


(n) (pi /m ; g , , ) =

log (n) (pi /m ; g , , ).


g

(21.115)

The way in which g() tends to g when approaching the critical point can readily
be deduced from Eq. (21.63). Let be the slope of the -function at the zero:

d(g)

.

dg g=g

Then we can expand near g

(21.116)

(g) = (g g ) + O((g g )2 ),

(21.117)

and deduce from (21.63) the behavior


log

1
log(g g ) + . . . ,

(21.118)

implying for small m the approach to g :

+ ... .
gm g m

(21.119)

Inserting this into the correction factor (21.114), we see that


(n)
C (n) (pi ; m , g , , ) = 1 + m
(pi /m ; g , , ) + . . . ,

(21.120)

In the scaling limit m 0, a finite correction to scaling is observed if (n) is


homogenous in the variables pi /m of degree . For the two-point vertex function:
]
m0
C (2) (pi ; m , g , 1, 1) 1 + const (p/m ) + . . . .
(21.121)
Then
(p; g, m, ) =

2
m

"

exp 2

m d

(g ) (p/m , g , , ) C (2) (p; m ; gm , , )


(21.122)

behaves for t 0 like


(p; g, m, )

m0

"

p
f (p/t ) 1 + (g g ) const

+ ...

(21.123)

21.8 Explicit Solution of the RGE close to D = 4 Dimensions

1173

Thus the approach to scaling is controlled the exponent , the slope of the beta
function at the fixed point g .
From the above discussion it is obvious that for an infrared stable fixed point,
is positive.
At this point one may wonder about the universality of this result since, in principle, other corrections to scaling might arise if higher powers of the field were included
in the energy functional, for example 6 or 2 ()2 . Those terms are found to be
irrelevant by dimensional considerations and by studying the renormalization group
equations in the presence of such terms. The scaling behavior and the approach to
it remain unchanged.

21.8

Explicit Solution of the RGE close to D=4 Dimensions

Up to the two-loop level, the -function was found in (21.54) to have in D = 4


dimensions the simple form
(g) = g + bg 2
(21.124)
with b = (N + 8)/6. This function starts out with negative slope and has a zero at
g = /b.

(21.125)

If is small enough, this statement is reliable even if we know (g) only to order g 2
in per turbation theory. Inserting (21.124) into equation (21.63) we may calculate
with 1 / and rewriting g() briefly as g(),
Z

log =

g()

dg
.
g + bg 2

(21.126)

Here we realize the important consequence of the zero in the -function: If g is


sufficiently close to the zero g then, in the limit 0, the value g() always runs
into g independent on whether g = g(1) lies slightly above or below g . The point
g is called a fixed point of the renormalization group equation. Since g is reached
in the limit 0 one speaks of an infrared stable fixed point. The flow of ()
when sending 0 is illustrated in Fig. (21.2) by an arrow. Using the variable
1/g instead of g, Eq. (21.126) becomes
1
log =

1/g()

1/g

dx
.
1/g x

(21.127)

This can be integrated directly to


=

|1/g 1/g()|1/

g() =

g
.
1 + (g /g 1)

|1/g 1/g|1/

(21.128)

so that
(21.129)

1174

21 The Renormalization Group

(g)

g/g

Figure 21.2 Flow of the coupling constant g() as the scale parameter approaches zero
(infrared limit). For the opposite direction (ultraviolet limit), the arrows are reversed.

In general, the -function may behave in many different ways for larger g. In
particular, there may be more zeros to the right of g . We can see from Eq. (21.126)
that for negative (g), the coupling constant g() will always runs towards zero from
the right. The opposite statement holds for positive (g).
In general, the initial coupling g() g can flow only into the zero which lies
in its range of attraction. In the present case this is guaranteed, for small , if g is
sufficiently small.
In the limit , we see from Eq. (21.129) that g tends to zero which is the
other trivial zero of the beta function. This happens for any zero with a positive
slope of (g). The limit corresponds to m2 , and for this reason such
zeros are called ultraviolet stable.
In the limit , the coubling constant goes to zero, g() 0. In this limit
(g()) 0 and m (g()) 0. Then scaling relation (21.77) ensures m = m/,
and the correlation functions behave like
m0

(n) (pi ; g, m, )

!Dn(D/21)

(n) (pi /(m/); 0, , ).

(21.130)

This is the behavior of a free-field theory where the fields fluctuate in a trivial purely
Gaussian way. The zero in (g) at g = 0 is therefore called the trivial or Gaussian
fixed point. In the 4 -theory, the Gaussian fixed point is ultraviolet-stable (UVstable). Since the theory tends in th limit m against a free theory, one also
says that it is ultraviolet-free. Note that this is true only in less than 4 dimensions.
In D = 4 dimensions where = 0, the beta function has only one fixed point, the
trivial Gaussian fixed point at the origin. The theory is UV unstable and IR trivial.
It is interesting to observe in which way the renormalization group equation
(21.126) sums up an infinite number of diagrams of the perturbation expansion. To

21.8 Explicit Solution of the RGE close to D = 4 Dimensions

1175

At the one-loop level, the perturbative relation (21.38) between bare and renormalized coupling constant is according to (21.51) [and using g = /b from (21.125)]:
"

g
N +8
+ O(g 2 ) = g 1 + + O(g 2 ) .
g0 = g 1 + g
9
g

(21.131)

This may be rewritten as


1
1
1
= + O(g).

g0
g g

(21.132)

The solution (21.71), on the other hand, can be brought to the form
!

1
1
1
1
=
.
g g
g g

(21.133)

This solution may be compared with the perturbative one in the limit of large
where g is small. In this limit we identify
1
1
1
=
.

g0 ()
g g

(21.134)

1
1
1
= .

g0
g g

(21.135)

With this, Eq. (21.71) yields

This differs from the perturbative equation (21.136) by having no hiher order correction terms collected in O(g). The renormalization group equation has resummed
the first-order result (21.136) to (21.135). By taking the inverse of (21.135) we find
the geometric series
g0

g
= g 1 +
g

!2

g
+
g

!3

+ O(g 4 ) .

(21.136)

Since the first correction is due to a single loop, this series corresonds to summing
up arbitrarily long chains of loops.
A relation of the type (21.135) was encountered before in the context of exactly
solvable O(N)-symmetric 4 -theories. Indeed, that relation being exact must also
be an exact solution of the renormalization group equation. The difference between
that and the present equation is a factor 3 which is due to the fact that for N
only one of the three Wick contractions contributes [recall the discussion after
Eq. (19.58)].
The power of the renormalization group is to allow us to take the perturbative
equation (21.131) to the scaling limit which corresponds to replacing that equation
by by
"
#
g

2
g0 = g () 1 + + O(g ) ,
(21.137)
g

1176

21 The Renormalization Group

and letting 0 in that equation where the left-hand side becomes large and the
expansion on the right-hand side is useless.
The problem appears even more dramatically by using the opposite equation
g = g0 ()

"

g0 ()
+ O(g02) ,
1
g
#

(21.138)

where the terms on the right-hand side diverge increasingly fast for higher orders.
The resummed form of this equation is from (21.135):
1
1
1
=
+ ,

g
g0 ()
g

(21.139)

from which one deduces directly that g g in the scaling limit 0.


The critical coupling g where (g) vanishes can be calculated for the O(N)symmetric 4 theory in 4 dimensions as a power series expansion in starting
from Eq. (21.54) [recalling the reduced matrix elements TD1 etc. calculated in
Eqs. (11.234)(11.246)]:
1
2 3 2
3
2
+ ...
+
TI4 TD3
3TI
2
12
3TI
"
#
6
3N + 14 2
=
+3
+ ... .
N +8
(N + 8)2


g =

(21.140)

Inserting this into (21.55) and (21.56) we find the expansions for , m
up to this
order:

(g ) =

1 TD3 2 1 N + 2 2
,
=
108 TI2
4 (N + 8)2
2

m
2 m (g )



 2
1 TD1
1
4 3
1
TD1

=
+
TD3
TI4 TD3
3 TI
27 2
12
TI
9
TI2
"
#
N +3
N +2
1+6
= 2 1 ,
=
N +8
(N + 8)2

so that

"

N +2
13N + 44
=
1+
= 2 1 .
2
N +8
2(N + 8)

(21.141)

(21.142)

(21.143)

The last equation yields the -expansion for the critical exponent :
=

1
N +2
(N + 2)(N + 3)(N + 20)
+
+
+ ....
2 4(N + 8)
8(N + 8)3

(21.144)

For = d/dg|g=g we obtain from (21.54) and (21.140)


=+

1
TD 6TI4
3 3



3TI

2

= 3

3N + 14 2
+ ... .
(N + 8)2

(21.145)

1177

21.9 Further Critical Relations

21.9

Further Critical Relations

We shall now derive some more consequences of the above general scaling relations
(21.73), (21.101), (21.112). To have some specific physical system in mind, we may
think of (x) for N = 3 as describing the local fluctuating magnetization in some
magnetic system. fluctuating magnetization field. At the mean-field level we see
that if the square mass m2 is larger than zero, the field expectation value = hi
vanishes. This corresponds to being above the transition temperature for magnetic
q
ordering. For m2 < 0, on the other hand, the value becomes non-zero 6m2 /g.
This corresponds to the spontaneous magnetization of a ferromagnet. Certainly,
in a proper magnetic system, the magnetization can take any spatial directions,
while in the model with a real field allows only for two opposite directions. An
improvement can be made by letting be a vector field and this will be done in
a later section. It should, however, be noted that experimentally it is possible to
find special magnetic systems in which the lattice anisotropy inhibits strongly the
rotational degrees of freedom and favors only two directions of magnetization. Such
systems are referred to as Ising like. The following discussions will apply to these
if we identify with the Ising like magnetization M Mz . A linear coupling to an
external current j in the effective action
[; g, m2 , ] +

dD xj [M, g, m2 , ] +

dD xMH

(21.146)

is then to be interpreted as coupling to an external magnetic field.


The mass square is identified with the relative temperature distance from critical
value as in
From what we have learned in the Section (21.6), we may conclude that the fixed
point at g = g governs the leading behaviour of all vertex functions close to the
critical point.
The critical exponents , , determine the critical behavior of all observables
and the approach to this behavior.
Let us derive the scaling relations of several important thermodynamic quantities
and correlation functions.

21.9.1

Scaling Relations Above Tc

Consider at first the regime T Tc with a vanishing external field H.


Specific Heat Consider the specific heat as a function of temperature. The
ground state energy above Tc is given by the effective potential at zero average
field v( = 0), which according to (21.103) and (21.79) has the scaling behavior
m0

D
v( = 0) m
const. D/(2m ) = D .

(21.147)

Forming twice the derivative with respect to we find for the specific heat at constant volume
t0

C D/(2m )2 = D2 , > 0.
(21.148)

1178

21 The Renormalization Group

This behaviour has been observed experimentally and the critical exponent has been
named :
t0
C t ; t > 0.
(21.149)
Thus we can identify

= 2 D/(2 m
) = 2 D.

(21.150)

Susceptibility Suppose now that the system at T > Tc is brought into an external
field j 6= 0 which we shall call magnetic field settin j = M (indicating applications
to magnetism). The equilibrium value of the internal magnetic field M P hi is
no longer zero but MH j . It is determined by the equation of state

From (21.103) we see that


0

j m D

v()
j=
.
j

(D/21)

(21.151)

+D/21
v (/m
; g , , ),

> 0.

(21.152)

The susceptibility is obtained by an additional differentiation


1

2 v 0 D(D2) 2
+D/21
m
m v (/m
; g , , ),

2

=j

> 0. (21.153)

For > 0 and zero field H j one has = 0, and finds

22
1 m
= 2(1

)/(2 )
m

(21.154)

Experimentally, this critical exponent is called :


0

1 ,

> 0.

(21.155)

Hence we can identify


=2

1
= (2 ).
2

(21.156)

Critical Magnetization At the critical point, the proportionality of H and M


is destroyed by quantum fluctuations. Experimentally one observes
H M ;

t = 0.

(21.157)

Equation (21.152) shows that if the effective potential is to be finite at the critical
point where m = 0, the derivative v must behave like some power
t0

v const. /

+D/21

t>0

(21.158)

1179

21.9 Further Critical Relations


D
then, since j m
satisfy

(D/21)

v , the power which makes j finite for m 0 must

D (D/2 1) + [ + (D/2 1)] = 0.

(21.159)

From this we obtain


=

21.9.2

D + 2 2
D+2
=
.

D 2 + 2
D2+

(21.160)

Scaling Relations Below Tc

Let us now turn to scaling results below Tc . Since all individual vertex functions in
the expansion of the effective action (21.89) can be calculated for m2 < 0 just as
well as for m2 > 0, the main difference lies in v() not having a minimum at = 0
but at = 0 6= 0 for vanishing external fields. In addition, the scaling variable m
is not fixed by the condition (21.72), but by a similar relation with an opposite sign
m2 / 2 |=m = 2 ,

(21.161)

so that we obtain via (21.74) the relation


m2 m0
2 exp

(Z

2m
[2 m (g )] = m
.

(21.162)

rather than (21.76), leading to

m ( )1/(2m )

(21.163)

instead of (21.79).
Spontaneous Magnetization Consider first the behaviour of the spontaneous
magnetization M0 0 as the temperature approaches Tc from below. The equilibrium value of is determined by the minimum of the effective potential v().
According to the general scaling form (21.103), the minimum must be situated at a
fixed

+D/21
0 /m
= const .

(21.164)

Hence 0 depends on m2 and thus on the reduced temperature t as follows:

+D/21
M0 0 m
= ( )(

+D/21)/(2 )
m

(21.165)

Thus we derive the experimentally observable relation


M0 0 ( )

(21.166)

with the critical exponent [not to be confused with the -function (21.31)]
=

+ D/2 1

=
(D 2 + ).

2 m
2

(21.167)

1180

21 The Renormalization Group

Coherence Length Consider now the temperature dependent coherence length


below Tc . From (21.112) we read off that the two-point function satisfies
0

+D/21
22
, g , , ,
(2) (p; 0 , g, m2, ) m
(2) p/m ; 0 /m

> 0. (21.168)

As in the previous case T > Tc , this is a function of


p/m = ( )p,

(21.169)

with the same temperature behavior as in (21.87). The same critical exponent
governs the divergence of the coherence lengtk below and above Tc .
This above-below symmetry is found for the critical indices and of specific
heat and susceptibility. When deriving the scaling behavior for < 0, we have to
keep track of the change of the abverage field 0 with temperature.
Specific Heat The exponent of the specific heat, , follows now from the potential
t0

+D/21
v(0 ) m D v 0 /m
, g , , ,

> 0.

(21.170)

Since the temperature change of 0 takes place at a constant combination (21.164),


the presence of 0 6= 0 can be ignored and we obtain the same result as in (21.147),
implying a temperature behaviour
D
v(0 ) m
D .

(21.171)

This agrees with the T > Tc behavior (21.147), leading to the same critical exponent
of the specific heat as in (21.150).
Susceptibility Suppose now that an external magnetic field is switched on. The
equilibrium value shifts by j 0 from 0 to j , which is determined by
(21.151), (21.152). For we obtain the scaling relation
D(
j m

+D/21)

+D/21
v ((0 + )/m
, g , , ).

(21.172)

Expanding this to first order in gives

22
+D/21
j m
v (0 /m
, g , , ).

(21.173)

Since 0 changes with t according to (21.166), the last factor is independent of


temperature, and the susceptibility (t) has the same functional form as in (21.153),
exhibiting the same critical exponent as in (21.155).

1181

21.10 Comparison of Scaling Relations with Experiment

Widoms Relation Finally it is worth noting that Eq. (21.152) corresponds


exactly to a scaling relation proposed and investigated first by Widom (1965) for
the free energy at a fixed magnetization M:
F (t, M) = 2 ( /M 1/ ),

(21.174)

where is some smooth function of its arguments.


Differentiating this with respect to M, we find the magnetic equation of state
H = 2 M 11/ ( /M 1/ ).
This can be rewritten as

(21.175)

(21.176)
=f

M
M 1/
with some function f (x), as is easily proved by inserting on the left-hand side the
scaling relation = 1 + (2 )/. In terms of the variables of our field theory,
the equation of state (21.176) may be rewritten as
j=

1/

f
1/


= g ,

(21.177)

where g(x) is some other function. By comparing this with (21.152) we see that
= [D (D/2 1)]

21.10

= (D + 2 ).
2 m
2

(21.178)

Comparison of Scaling Relations with Experiment

For a comparison with experiment we may pick three two sets of critical data and
extract the values of , , and . The remaining critical indices can then be found
from the above-derived scaling relations which we summarized, for convenience:
= 2 2 ,
= (2 )
(2 ),
2 .
D+2
=
.
D2+

=
=
=
=

(21.179)
(21.180)
(21.181)
(21.182)
(21.183)

As an example take the magnetic system CrBr3 where one measures


0.368,

4.3,

1.215.

Inserting these into the scaling relation


= /( 1)

(21.184)

1182

21 The Renormalization Group

which folows (21.179) and (21.183), we see that it is satisfied excellently. Inserting
into the relation
D + 2 (D 2)
(21.185)
=
+1
we find for D = 3
0.132,
(21.186)
and from the relation = (2 ),
0.65.

(21.187)

Results from -Expansion


Let us compare these numbers with the results from the expansion. Using Eqs.
(21.140) - (21.142) and the above formulas for the critical indices in terms of , m
we find the series for the two independent critical indices.
N +2 2
+ ...
2(N + 8)2

(21.188)

N +2
(N + 2)(N 2 + 23N + 60) 2
1
+
+
+ ... .
2 4(N + 8)
8(N + 8)3

(21.189)

These results are valid only for infinitesimal . For applications, one assumes that
the expansions can be used also in D = 3 dimensions, i.e. for = 1, where
physics takes place. The justification of this rather ad hoc procedure lies in an
approximate numerical decrease of the and 2 terms in (21.140,21.142). Thus we
find for N = 1, 2, ,
1
7
203
= 21 + 12
+ 992
2 + . . . = 994
+ . . . 0.627,
1
1
11 2
131
= 2 + 10 + 200 + . . . = 200 + . . . 0.655,
= 21 + 14 + 81 2 + . . . = 87 + . . . 0.875,

N = 1,
N = 2,
N = .

(21.190)

The other critical index is


1
,
= 54
1
= 15 ,
= 21 ,

N = 1,
N = 2,
N = ,

(21.191)

in the three cases.


The -expansion for has decreasing contributions of the higher orders. The value
up to order 2 , 0.627, is in reasonable agreement with the experimantal value
0.65. 1 2 = 13 , 0.6 and this is in good agreement with the experimental
value (21.187) here from p.30.
There is only one term to order so no convergence can be judges. The agreement

1183

21.11 Higher-Order Expansion

with experiment reasonable. The value to order 2 at = 1 is 0.019 which,


via the scaling relation = /(2 ), corresponds to 0.61, quite close to the
experimental value 0.65.
It is not advisable to calculate to higher order than 2 since the agreement to
order 3 usually becomes worse. In fact, the series does not converge and it is
meaningless to carry the expansion any further. The rough agreement up to order
is a consequence of the asymptotic convergence of the series. the

21.11

Higher-Order Expansion

Let us state the presently known higher powers in up to order 4 one finds1 after
much work the beta function
(g) =

g2
6 [N

g4
+ 432
g5
7776

+ 8]

g3
6 [3N

+ 14]

[33N 2 + 922N + 2960 + (3) 96(5N + 22)]

[5N 3 + 6320N 2 + 80456N + 196648


+ (3) 96(63N 2 + 764N + 2332)

(4) 288(5N + 22)(N + 8)

+ (5) 1920(2N 2 + 55N + 186)]

g
+ 124416
[13N 4 + 12578N 3 + 808496N 2 + 6646336N + 13177344

+ (3) 16(9N 4 + 1248N 3 + 67640N 2 + 552280N + 1314336)

+ 2 (3) 768(6N 3 59N 2 + 446N + 3264)

(4) 288(63N 3 + 1388N 2 + 9532N + 21120)

+ (5) 256(305N 3 + 7466N 2 + 66986N + 165084)

(6)(N + 8) 9600(2N 2 + 55N + 186)

+ (7) 112896(14N 2 + 189N + 526)] ,

(21.192)

and the critical indices


() =

(N +2)2
2(N +8)2
2

1+

[N 2
4(N +8)2

+ 56N + 272]

16(N +8)4 [5N 4 + 230N 3 1124N 2 17920N 46144


3

+ (3)(N + 8) 384(5N + 22)]

64(N +8)6 [13N 6 + 946N 5 + 27620N 4 + 121472N 3


262528N 2 2912768N 5655552

(3)(N + 8) 16(N 5 + 10N 4 + 1220N 3 1136N 2 68672N 171264)

+ (4)(N + 8)3 1152(5N + 22)

(5)(N + 8)2 5120(2N 2 + 55N + 186)] ,


1

(21.193)

See H. Kleinert, J. Neu, V. Schulte-Frohlinde, K.G. Chetyrkin, and S.A. Larin, Phys. Lett. B
272, 39 (1991).

1184

21 The Renormalization Group

and
1/() = 2 +
2

(N +2)
N +8

2(N +8)2 (13N

+ 44)

+ 8(N+8)4 [3N 3 452N 2 2672N 5312


3
+ 32(N +8)6

+ (3)(N + 8) 96(5N + 22)]

[3N 5 + 398N 4 12900N 3 81552N 2 219968N 357120

+ (3)(N + 8) 16(3N 4 194N 3 + 148N 2 + 9472N + 19488)

+ (4)(N + 8)3 288(5N + 22)


4

(5)(N + 8)2 1280(2N 2 + 55N + 186)]

+ 128(N +8)8 [3N 7 1198N 6 27484N 5 1055344N 4

5242112N 3 5256704N 2 + 6999040N 626688

(3)(N + 8) 16(13N 6 310N 5 + 19004N 4 + 102400N 3


381536N 2 2792576N 4240640)

2 (3)(N + 8)2 1024(2N 4 + 18N 3 + 981N 2 + 6994N + 11688)

+ (4)(N + 8)3 48(3N 4 194N 3 + 148N 2 + 9472N + 19488)

+ (5)(N + 8)2 256(155N 4 + 3026N 3 + 989N 2 66018N 130608)

(6)(N + 8)4 6400(2N 2 + 55N + 186)

+ (7)(N + 8)3 56448(14N 2 + 189N + 526)] ,

(21.194)

The higher orders in start growing rapidly for = 1 after the 0 , 1 terms there is
no convergence. In fact, it can easily be shown that the series is only an asymptotic
expansion. Resummation techniques are needed to extract reliable informations.2

21.12

Mean-Field Results for Critical Indices

In mean-field theory fluctuations are neglected and the effective action is equal to
the bare action with the quantum field replaced by the expectation value = hi:
[] =

1
m2 2 g 4
d x ()2
+ j .
2
2
4
D

"

(21.195)

The mass is assumed to behave as


2

m =

T
1 2 ,
Tc


(21.196)

and and j are interpreted as magnetization M and field H, respectively. The


two-point function is
(2) (p) = p2 m2 ,
(21.197)
2

For a detailed discussion and the many references see the textbook W. Janke and H. Kleinert,
Borel Resummation of Divergent Perturbation Series, World Scientific Publ. Comp., Singapore,
1997. This book contains many references to the original papers.

1185

21.12 Mean-Field Results for Critical Indices

from which we read off, at m2 = 0 or T = Tc , that the critical index which


determines the power behaviour of the correlation function vanishes:
= 0.

(21.198)

Above Tc , we rewrite (2) in the rescaled form (21.70):


i

(2) (p; g, m, ) = 2 (p/)2 m2 / 2 .

(21.199)

Thus we identify m = m/ and the coherence length has the temperature behaviour
( ) =

1
1
=
1/2 .
m
m

(21.200)

This implies the critical index

1
= .
(21.201)
2
We now turn to the regime T < Tc . Then the effective potential has a nontrivial
minimum at
s
m2
( )1/2 .
(21.202)
0 =
g

The critical exponent is therefore determined to be


1
= .
2

(21.203)

Inserting = 0 + into the effective action (21.195), we find the two-point


function below Tc
(2)

2
=

= p2 + 2m2 .
=0

(21.204)

This shows that the coherence length is equal to


( ) =

1
1/2 .
2
2m

(21.205)

Hence the index below Tc is the same as above Tc :


1
= .
2

(21.206)

Consider now the specific heat. The equilibrium energy above Tc and below T is:

0,
m4
,
4g

T Tc ,
T < Tc ,

(21.207)

1186

21 The Renormalization Group

respectively. Differentiating this twice with respect to m2 we see that


C = const ,

T < Tc ,

(21.208)

which implies that


= 0.

(21.209)

Note that the index above Tc is undetermined at the mean-field level since C 0.
Let us now allow for a nonzero magnetic field. Then for T < Tc , the equilibrium
value of 0 i given by
m2 0 + g30 = j.
(21.210)
For T = Tc , this has the solution

j = g30

(21.211)

= 3.

(21.212)

so that we can read off [see (21.157)]


For T > Tc , we see that for small j

1
j
m2

(21.213)

so that the susceptibility is


(m2 )1 1 ,

(21.214)

= 1.

(21.215)

giving the critical index [see (21.155)]

Below Tc , on the other hand, we insert = 0 + and find


(m2 + 3g20 ) = j
or

1
j
2m2

(21.217)

1
( )1 .
2
2m

(21.218)

=
which amounts to
=

(21.216)

Thus is the same as :


= = 1.

(21.219)

Note that measures the size of the fluctuations of the magnetization field M(x)
around the equilibrium value. It is determined by the curvature of the effective
potential at the minimum. Above Tc this is m2 , below Tc it is twice the opposite of
m2 (which is again positive).
The critical indices satisfy the general scaling relations (21.179)(21.183). Indeed, with = /2 = 0, wee see that the relations are
D
D+2
1 D
= = 2 , =
, = = 1, =
1 .
2
D2
2 2
For D = 4 they yield once more the above derived mean field exponent.

(21.220)

21.13 Effective Potential in the Critical Regime to Order

21.13

1187

Effective Potential in the Critical Regime to Order

From (21.103), (21.77), (21.86), and (21.167), the effective potential has for small
mass values the universal form
m2
2

m0

v(; g, m, )
D

D D

!D

/(D2)/2
w
(m2 /2 )( +D/21)

/(D2)/2
w
,

(21.221)

where w is some dimensionless function. In the presence of an external magnetic


field H j, this corresponds to an equation of state in the Widom form (21.177):
v
/(D2)/2
j=
= D/2+1 D w

= g .

(21.222)

We shall now demonstrate how the function w can be calculated at the one-loop
level. The renormalization group analysis of the previous sections made statements
only on the numerical values for the critical indices. Nothing was known on the
specific shape of the effective action v and thus of the general Widom function f (x).
To lowest order in it is possible to calculate v from the one-loop effective potential.
Consider the Lagrangian density
m2 2 g0 4
1
.
L = ()2
2
2
4

(21.223)

We have taken the factor out of g0 to make it g0 a dimensionless number in


D = 4 dimensions. The one-loop correction to the effective action is calculated
in the usual way. The partition function reads
Z=

DeiA[]/h =

De(i/h)

dD xL(x)

(21.224)

Expanding around the classical solution


(x) = + (x),

(21.225)

the exponent has the lowest terms


Z

d xL()

1
g0
d x (x) 2 + m20 + 2 (x) + . . . .
2
2


(21.226)

Performing the path integral over the quadratic fluctuations gives


(i/
h)

Z e

dD xL()

i
= exp
h

Z

det

1/2

m20

g0
+ 2 2
2

i
g0
d xL()+ h
Tr log 2 + m20 + 2
2
2
D



eiA1 loop [] .(21.227)

1188

21 The Renormalization Group

As was shown before one can replace directly by the extremal value in the oneloop corrected action A1 loop [], thus obtaining the correct effective action [] to
order h
. The evaluation is easiest at a constant field where [] determines the
effective potential as
v() = LD []|=const ,
with LD being the volume of the system in D dimensions. Then the trace of the
logarithm is simply the integral
X
p

log p +

m20

Z
dD p
g0
g0 2
D
log p2 + m20 + 2 ,
+ =L
D
2
(2)

and the effective potential at the one-loop level reads


v() =

m20 2 g0 4 i Z dD p
g0 2
2
2
+ h

log
p
+
m
+
.
2
4!
2
(2)D

(21.228)

Performing a Wick rotation to euclidean momenta p with p2E = p2 , dD p idD pE ,


the third term becomes
1
2

g0
d D pE
log p2E + m20 + 2 .
D
(2)
2


(21.229)

This integral was calculated in Eq. (11.134) for all dimensions D as


Z

d D pE
1
2
log(p2E + m20 ) = S
D (D/2)(1 D/2) (m20 )D/2 .
D
(2)
2
D

(21.230)

This brings the the potential to the form


v=

m2 4 g 4 1 1
2
g0
+ + S
D (D/2)(1 D/2) (m0 + 2 )D/2(21.231)
.
2
4!
2 2
D
2

It will be more convenient to consider directly the equation of state rather than v
itself and derive
g0
1
g0 2 1/2
v
.
= m20 + 3 + S
D (2/2)(1+/2)hg0 m20 +

j=

3!
4
2
(21.232)


Close to = 0, the last term can be expanded as


1
g0
S

+ O()

g0 2
m20 g0 2
1 log
m20 +

+
2
2
2
2 2



"

+ O() .
(21.233)

According to the minimal substraction scheme in dimensional regularization we may


simply drop the singular 1/-piece and arrive at the renormalized expression
v
g
S
D g 2 g 2
m2 g 2
j=
= m2 + 3 + h

(m + ) log
+ 2 . (21.234)

3
4
2
2
2
!

21.13 Effective Potential in the Critical Regime to Order

1189

The potential is
g 2
S

m2 2 g 3
+
+h
m2 +

v=
2
3!
8
2


2 "

m2 g 2
log
+ 2
2
2

1
.(21.235)

Let us now see how these expressions can be brought to the universal forms (21.221),
(21.222). Neglecting, for a moment, the logarithmic h
-correction, we rewrite the first
two terms in (21.234) as
v

j=
= 3+/2 1+/2

= 3+/2

m2
2

m2
g
1+
2

3!

!3/2

1+/2

s

1+/2

s

m2
2

m2
g
1+
2

3!

!2

1+/2

s

m2
2

!2
.(21.236)

Comparing this with (21.222), we can identify the mean-field critical indices [as in
(21.203), (21.212)]
1
= 3, = .
(21.237)
2
We shall now calculate the lowest-order corrections to thse critical indices by
considering two specific scaling relations. First, we consider the limit 0. Since
for > 0, the derivative v/ has to carry a factor , equation (21.222) shows
that v/ should behave as a function of as follows
j 2 D2 = 2 (1) = 2 .

(21.238)

From (21.236) we see, on the other hand, that


m2
S
D g
m2
v
= 2 2 1 + h

log 2 .
j=

(21.239)

This is a first-order expansion in g of the power law


=

m2
2

!1+h SD g/4

+ O(g 2 ).

(21.240)

From this, we read off the one-loop corrected critical exponent which rules the
divergence of the magnetic susceptibility near the critical point via :

h
S
Dg
+ O(g 2 ).
(21.241)
4
For a second independent exponent we work directly at the critical point m = 0.


By Widoms relation, v/ behaves like /(D2)/2 . From (21.236) we find the
power law
(2 ) = 1 +

v
g 3 5 2
=
2

3!
g
g 3 5 2
=
2
3!
g

!3/2

g
2 1/2
r

!3 "

g
2 1/2

S
D g
1+h

log
2

!3+h3 SD g/2

g
2 1

+ O(g 2 ).

!#

(21.242)

1190

21 The Renormalization Group

We can therefore identify the critical index with which the critical magnetization
responds to an external field (in the scaling law H M )
3
D g + O(g 2 ).
=3+h
S
2

(21.243)

Using (21.179)(21.183) we can calculate the exponent


=

D
1+

(21.244)

which rules the way in which the spontaneous magnetization vanishes for 0
[in the scaling law M ( ) ]. Further
=

+1
D1

(21.245)

governing the divergence of the coherence length (in ). Note that satisfies
the scaling relation

=
,
(21.246)
1
as it should according to (21.184). Inserting (21.241) and (21.243) we find
1
3
D g) = h
S
g/4 + O(g 2 ).
= (1 + h
S
g/4)/(2 + h
S
2
2

(21.247)

We can use these results to explicitly find the function g(x) in the universal Widom
expressions (21.222), (21.177) at the one-loop level. We define the dimensionless
field variable
r
g
y=
(21.248)
2 1/2
and the combined variable occuring in (21.242)
x /y 1/ .

(21.249)

Then v/ in (21.242) is expressible in the form


v
= 2 y 1 f (x),

(21.250)

with a function f (x) related to the potential v by


f

2
x = x v (x).
g

(21.251)

By rewriting (21.236) in terms of y we find


v
S
D g

= 2 ( + y 2 /3) + h
( + y 2) log( + y 2 ) .

4


(21.252)

21.13 Effective Potential in the Critical Regime to Order

1191

Upon identifying
x=

y 1/

y 2+h SD g

S
log y 2 ,
2 1h
y
2


(21.253)

this can obviously be brought to the form


3
g
1
v
S
g log y 2
(21.254)
= 2 y 1 1 h
+x 1+h
S
log y 2

4
3
2






g
g
+hS
g/4 (1 + x) 1 + h
S
log y 2 log y 2 + log(1 + x) 1 + h
S
log y 2
.
2
2


 



Since we are working at the one-loop level, the g 2 corrections can be neglected and
we remain with
v
= 2 y 1



1
S
D g
+x +h

(1 + x) log(1 + x) .
3
4


(21.255)

To the same order in g, this is the same as


v
2
= 2 y 1 + (1 + x) .

3


(21.256)

We have identified the leading power of 1 + x with since it determines the behaviour of the susceptibility for 0, x . The expression (21.256) is of
the desired universal form (21.250).
Note that at the one-loop level, this form has been obtained by working purely
in 4 dimensions and calculating the correction to order g. Nowhere have we needed
the -expansion and the critical indices , etc. as functions of the running coupling
constant g(). Continuation to 4 dimensions bcomes only necessary as we are
trying to approach the physical situation D = 3. If we move only an infinitesimal
piece away from D = 4, the running coupling stabilizes at h
S
D g = 23 . This is
extracted from the coupling constant renormalization
1
SD h

=
g()
g0
2

(21.257)

by going to the infrared limit 0. This is what finally determines the critical
indices as functions of :
= 3 + ,
= 0,
1
=
, = 1 + /6,
2 6

1
+ .
=
2 12

(21.258)

If the function f (x) in the universal form (21.250) is desired to higher orders in ,
the calculation becomes quite cumbersome already at the next order in .

1192

21 The Renormalization Group

21.14

O(N)-Symmetric Theory

For completeness, let us generalize the previous calculation to the full O(N)symmetric case with the Lagrangian density
m20 2 g0 2 2
1
2
(a ) .
L = (a )
2
2 0 4!

(21.259)

Now the quadratic fluctuations are given by


2A =

g0
1
(ab 2 + 2a b ) b
d4 x a 2 m20
2
6


(21.260)

so that the one-loop corrected effective action becomes


1
m20 2 g0 2 2
2
[] = (a )

(a )
2
2a
4

i
g0
2
2
2
h
tr log ( m )ab
(ab + 2a b ) ,
2
6

(21.261)

with an effective potential


g0 2 2 1
m2
(a ) + tr
v[] = 0 2a +
2
4!
2

d D pE
g0
2
2
log
(p
+
m
)
+
(ab 2 +2a b ) .
ab
E
(2)D
6
(21.262)


The trace of the logarithm of the matrix calculated as the sum of the logarithms of
the eigenvalues which are, obviously,
2

p +

m20

g0 2
+

6N

(21.263)

for the N 1 unit vectors orthogonal to and


p2 + m2 +

g0 2

2N

(21.264)

for the Nth unit vector parallel to a : Hence


m20 2 g0
+
(a )2
(21.265)
2 a 4!N
"
#


Z

2
h
d D pE
g

0
0
+
(N + 1) log p2E + m20 +
2 + log(p2E + m20 +
2 ) .
2
(2)
6N
2N

v() =

Performing the momentum integral, this becomes


v(a ) =

m20 2 g0 2 2 h
1
2
a +
(a ) + S
(D/2)(1 D/2)
2
4!N
2 2
D



g
0
0
2
2 D/2
2
2 D/2
(N 1)(m0 +
.
))
+ (m0 +
)
6N
2

(21.266)

21.14 O(N )-Symmetric Theory

1193

The corresponding equation is


g0 2
g0
v(a )
= a m20 +
a + h

S
(D/2)(1 D/2)
ja =
a
3!
4N


1
g0 2
g0 2 1/2
(N 1)(m20 +
.
) + (m20 +
)
3
6N
2N


(21.267)

Expanding this to lowest order in = 4 D, we find after renormalization


m2 2 g 2 2
g 2 2
S

g 2
v() =
(N 1)(m2 +
a +
(a ) + h

) log(m2 +
)
2
4!
8
6
6

g
g 2
+ (m2 + 2 )2 log(m2 +
) .
(21.268)
2
2


with an equation of state


v
g 2
g 1
g 2
g 2
ja =
= a m2 +
a + h
S
(N 1)(m2 +
a ) log(m2 +
)
a
3!
4 3
6
6 a

g 2
g 2
a ) log(m2 +
a ) .
(21.269)
+ (m20 +
2
2


Again we may identify the susceptibility by taking a 0




ja = m2 1 + h

g
N+2
g 1
2
(N 1) + 1 log m2 (m2 )1+h 4 S 3 +O(g ) (21.270)
4D 3

so that [compare (21.241)]


= 1+h

g N + 2
S
D
.
4
3

(21.271)

The index is found by comparing the m2 = 0 -relation (21.250) [recall also (21.242)]
g
N 1
g 2
a 1 +
+3 h

log 2
ja a
3!
3
4


(21.272)

with the defining relation




ja a 2a
This gives

(1)/2

(21.273)

S
D g N 1
1
=1+h

+3 .
2
4
3
Using the same reduced variables as before, i.e.,


y =

(21.274)

g 2a
2 1/2

x = /y 1/ ,

m2 = 2 ,

(21.275)

1194

21 The Renormalization Group

the equation of state can be brought to the universal form


ja =

v
1
= a 2 y 1
+x
(21.276)
a
3






x
x
g 1
log 1 +
+ (1 + x) log(1 + x) .
(N 1) 1 +
+h
S

4 3
3
3


Using again , this can be rewritten, in analogy with (21.256), as


v
2
x
3
N 1
ja =
= a 2 y 1
1+
(N 2) +
(1 + x) +
a
3
N +2
N +2
3


 

. (21.277)

At this level, the indices and do not yet have their infrared fixed point values.
From (21.140) we know that the fixed-point lies at
g h
gS
=

6
.
N +8

(21.278)

This allows us to express in terms of as


= 3 + ,

(21.279)

valid for any N. Actually, this would have followed directly from the general scaling
relations (21.179)(21.183) and the fact that the anomalous dimension vanishes
to order [see (21.141).
Finally, comparing (21.271) with (21.278) we obtain
=1+

1N +2
.
2N +8

(21.280)

Thus we can calculate the other critical indices as

1
N +2
= +
,
2
2 4(N + 8)
3
1
.
= ( 1) =
2 2(N + 8)
=

21.15

(21.281)
(21.282)

Direct Scaling Form in the Limit of Large N

In the limit N , effective action and potential are known exactly and it is
possible to retrieve the scaling properties most directly. Consider the potential in
D = 4 dimensions as calculated in (8.4): (here we use gN0 as a coupling constant
rather than g0 )
v() =

m2
N 2 1
1
2
1
2a + 0
+ NS
(2 /2)(1 + /2)
2/2 ,
2
2g0 4g0
2
2
4
(21.283)

21.15 Direct Scaling Form in the Limit of Large N

1195

where is the function of 2a :


2a =

N
Nm20
1

NS
(2 /2)(1 + /2)1/2 .
g0
g0
2

(21.284)

We see that m20 and 2a scale in the same way so that v is a function of 2a /m20 and
m20 . If we let m20 0 with 2a /m20 fixed, the last term in Eq. (21.284) dominates
the first and we find the limit
m0 0
1
N
2a
m20
N S
(2 /2)(1 + /2)1/2 .
+
2
m0 g0
2

(21.285)

Similarly, in v the last term dominates the third so that we can write
1
N
1
2
m20 2a
+ NS
+
(2 /2)(1 + /2)
2/2
v()
2
2 m0 g0
2
2
4
"

m2
0
2

2a
N
+
2
m0 g0

2
N
2a
const m20
+
2
4
m0 g0
"

!#1+1/(1/2)

. (21.286)

Comparing this with Widoms parametrization


v()

(2a )(+1)/2 f

m20
(2a )1/

(21.287)

we see that
1
= ,
2

(21.288)

2 /2
+1
=
,
2
1 /2

(21.289)

and

or

6
D+2
=
.
(21.290)
2
D2
This agrees with the N limit of (21.279)(21.282).
Note that we have done the calculations in terms of the unrenormalized quantities
2
m0 and g0 . If we go over to the renormalized ones m and g, the changes affect only
the terms which were dropped in the limit m20 0 [i.e. the third term in (21.280)
and the first one in (21.281)]. The result remains therefore the same.
=

How much easier it is to be critical than to be correct


Benjamin Disraeli (18041881)

22
Critical Properties of Non-linear -Model
22.1

Introductory Remarks

The 4 -expansion of O(N)-symmetric 4 -theories is not the only technique to


study critical phenomenon of three-dimensional systems with O(N)-symmetry. A
great deal can be learned about such systems by ignoring the size fluctuations of the
fields and focusing attention only upon the Goldstone modes of the system, which
are present in the ordered phase. We have explained before in Chapter 9 that in the
limit of very low temperature or strongly negative mass parameter m2 , the fields
2
a , a = 1 . . . N in a (2a ) -theory
Z

A=

m2 2 g  2 2
1

(a )2 +
+
d x
2
2 a 4 a
d

"

(22.1)

acquire an infinite mass


M 2 = 2m2

(22.2)

and are practically frozen at a fixed size


2a = ||2 .

(22.3)

Only the italic of a are fluctuating easily. If these directions are denoted by the
unit vectors na = a /, the action describing these fluctuations can be written as
A=

1
2t

dD x (na )2 ,

n2a = 1

(22.4)

where 1/t = ||2 . The partition function is defined by functionally integrating eA


over all directions at each space point. The measure of integration is
Z

Dn =
=

YZ
x,a

YZ
x,a


dN na (x) (D)  2
na 1

SN

dN 1 na (x)
,
SN
1196

(22.5)

1197

22.1 Introductory Remarks

where SN is the surface of the sphere in the N-dimensional space, i.e., SN =


2 N/2 /(N/2). The partition function is therefore
Z=

Z 

Dn A[n]
e
.
SN


(22.6)

As mentioned above, this is the continuum version of the so-called classical O(N)Heisenberg model of magnetism on a lattice
Z=

"
Y Z

1
dn(x)
exp

SN
2T

x,a

Xh

(i na (x + i))2 + 2D

(22.7)

x,i

where i n(x) = n (x) + i n(x) are the differences of na (x) between sites. Because
of this connection, we shall rename g as t and speak about it as a temperature. In
Chapter 9 we have solved this model exactly in the limit N , and the reader is
referred to this chapter to recapitulate some of the results.
We shall now calculate the partition function (22.6) perturbativey for arbitrary
N. For this we select a particular orientation of the direction vector na , say Na =
(0, 0, 0, 0, 1), as a reference direction state, and consider the first N 1 components
a (x) na (x);

a = 1, . . . , N 1

(22.8)

as independent fields of the system. The Nth component is then determined from
the others by the unit length:
nN (x) =

1 a2 (x) (x),

(22.9)

and we can write the action of the non-linear model as


1
A=
2t

"

 q

d x (a ) + 1 a2

2 #

(22.10)

This action is scale invariant for D = 2 dimensions in, where the field a has a
vanishing naive dimension. The theory is renormalizable by power counting. Below
D = 2 dimensions, the fluctuations are so violent that massless excitations do not
exist. All Goldstone modes become massive at all temperature. The dimension
D = 2 is called the lower critical dimension.
For the action (22.4), the measure of integration is found by performing in (22.5)
the integral over nN (x) after which remain with
Z

NY
1
D N 1 

=
1 2 x,a=1

This can also be written as


NY
1

x,a=1

N 1

1
a exp D
2a


da (x)
q

1 a2 (x)

d x log 1

a2 (x)

(22.11)



(22.12)

22 Critical Properties of Non-linear -Model

1198

where we have assumed the path integral to be set up on an underlying simple


lattice of spacing a, such that we may write the products over all space points as
exponentials of sums, which are turned into integrals:
Y

"

f (x) = exp

X
x

1 Z D
log f (x) = exp
d x log f (x) .
aD


(22.13)

Then the partition function (22.6) of the non-linear model becomes


Z =

NY
1

x,a=1

Z

N 1

1
exp
2t

"

 q

d x (a ) + 1 a2

2 #

)
1
h
i
1 Z D NX
2
D d x
log 1 a (x) .
2a
a=1

(22.14)

Note that the original O(N)-invariance of the Lagrangian (22.232) under transformation
na = ab nb ,

(22.15)

with ab = ba , is present in (22.4) for transformations


q

1 c2 = aN a .

22.2

a = ab b + aN 1 c2 ,
(22.16)

Perturbation Theory

For a perturbation calculation of (22.14) we identify the free action as


1
dD x(a )2 ,
2t
and consider the remaining terms as interactions:
A0 =

Aint =

dD x

(22.17)

1 1  2  2
3  2   2 2
a b
a b + . . .
(22.18)
2t 4
16

1 2
1  2 2 1  2 3

D a (x) D a
...
2a
4a
6aD a


The functional
the regime 2 < 1, to be covered twice, once
integration is limited to
2
for = + 1  and once for = 1 a2 . The perturbation expansion is a
power series in t. For small t, the fluctuations will take place in the neighborhood
of 2 0. One may therefore replace the integrations over each field component
a (x) by ba one from to , committing an error of the order e1/t . With this
replacement, the free propagator is
dD k t ik(xy)
e
(2)D k 2
ab G0 (x y).

ha (x)b (y)i = ab

(22.19)

1199

22.2 Perturbation Theory

Observe that the 2 -term coming from the measure is not considered as being part
of the free-field action, but is treated as an interaction term since it carries a factor
t.
In order to avoid problems with singularities at k = 0 (infrared singularities),
it is useful to place the system into an external field pointing along the preferred
a = N direction:
Aext =

h
t

dD x =

h
t

dD x 1 a2 .

(22.20)

This corresponds to putting the Heisenberg model of magnetism into an external


magnetic field. This field serves to stabilize the selected preferred direction along
na = (0, 0, 0, . . . , 1) which was chosen to define the fields a . Expanding also Aext
in powers of a , this replacement produces a pion mass term in the action
h
Am =
2

dD x a2 (x),

(22.21)

t
dD k ik(xy)
e
D
2
(2)
k + m2

(22.22)

which changes the free propagator into


G0 (x y) =

where
m2 = h.
It also adds a further set of interactions


Z
1  2 2
1  2 3
b +
b + . . . ,
Aint = m2 dD x 1 +
8
16

(22.23)

(22.24)

the first being a trivial shift in the energy. In momentum space, the lowest interactions in (22.18) can be written as follows
2-point
=

1
4aD

dD q a (q)a (q),

(22.25)
(22.26)

4-point
=

1
8t

d-D q1

dD q2

dD q(q 2 + m2 )a (q1 )a (qq1 )b (q2 )b (q2 q),


(22.27)

1
= D
4a

d-D q1

d-D q2

d-D q a (q1 )a (q q1 )b (q2 )b (q2 q),


(22.28)

22 Critical Properties of Non-linear -Model

1200
6-point



1
(22.29)
d-D q4 d-D q q 2 + m2
16t
a (q1 ) a (qq1 ) b (q2 ) b (q3 ) c (q4 ) c (q2 + q3 + q4 q) ,
Z

1
(22.30)
= D d-D q1 d-D q4 d-D q
6a
a (q1 )a (q q1 ) b (q2 ) b (q3 ) c (q4 ) c (q2 + q3 + q4 q) .
Z

The dashed line stands for a factor q 2 + m2 , while the wiggly lines represent a local
interaction. In addition, these lines indicate the contraction of the O(N)-exponents
of the pairs of -lines emerging from its ends.
From the magnetic interaction (22.20) we obtain once more the lower two-point
interaction with a prefactor m2 /8t, and the lower four-point interaction with a prefactor m2 /16t, etc.
We shall now study the renormalization of the propagator to lowest order in perturbation theory. The self-energy contains two contractions involving the four-point
function:
(p) =

(22.31)

Let us calculate the first two diagrams using (22.27) and closing two lines by a
propagator. Alternatively, we may calculate the expectation
1
8t

d-Dq1 d-Dq2 d-Dq (q 2 + m2 )hc (p)a (q1 ) a (q q1 ) b (q2 ) b (q2 q) d (p)i

form all Wick contractions, remove the disconnected diagrams, and amputate the
external legs. This gives
(D)(q)
(D)
(D)
2 d q1 d q2 d q (q 2 + m2 ) (N 1)ca - (p + q1 ) ad - (q q1 p) 2
q2 + m2

- (D) (q q1 q2 )

(D)
(D)

(q

p)
+2ca - (p + q1 ) ab
2

(q q1 )2 + m2
Z

-D

-D

-D

= cd

1
2

- D (N 1)
dq

m2
q 2 + m2
+
2
.
q 2 + m2
(p + q)2 + m2
)

(22.32)

By rewriting
Z

d-D q

q 2 + m2
=
(q + p)2 + m2

(q + p)2 + m2 2p(q + p) + p2
,
d-D q
(q + p)2 + m2

(22.33)

1201

22.2 Perturbation Theory

we can drop the p(q + p)-term, since it is odd in (q + p), and (22.33) becomes
Z

d-D q + p2

d-D q

1
.
q 2 + m2

(22.34)

On a lattice, the first integral is equal to


1 X
1
=
,
D
Na q
aD

(22.35)

where N is the number of cells in the underlying lattice. Hence we can identify
1
d-D q = D .
a

(22.36)

This term in (22.34) is precisely cancelled by the contribution of the two-point


interaction (22.26), and we arrive at
cd (p) = cd

1
(N 1) m2 + p2 L
2


(22.37)

where L is the one-loop integral


L

d-D q

1
.
q 2 + m2

(22.38)

The two-point vertex function is therefore at the one-loop level


(2)

cd (p) = cd


1
1 2
p + m2 + (N 1)m2 + p2 L .
t
2


 

(22.39)

A one-loop effective action would have a free-field part


Aeff [] =

(2)

dd x c (x)cd (i)d (x)

(22.40)

The resulting two-point vertex function is thus


(2)
cd (p)

= cd


1
1 2
p + m2 + (N 1)m2 + p2 L + O(t).
t
2


(22.41)

The corresponding effective action has a free-field part


Aeff [] =

(2)

dD x c (x)cd (i)d (x),

(22.42)

and reads in x-space


Aeff

1
=
2

d x

(

1
1
h
1 + (N 1) tL 2a .
+ L (a )2
t
t
2


(22.43)

22 Critical Properties of Non-linear -Model

1202

These are the same field expressions as in the original free-field action, except that
t and h2 are replaced by the renormalized values,
1
1

+L
t
t

1
h h 1 + (N 1)tL .
2


(22.44)

The surprising feature of this theory is that due to the O(N)-symmetry, the renormlized action to all orders can be described in terms of only two renormalization
constants, a wave function renormalization
R = Z1/2

(22.45)

and a temperature renormalization


tR = Zt1 t.

(22.46)

The symmetry forces us to introduce the renormalized magnetic field h as


hR = Zt1 Z1/2 h Zh h.

(22.47)

This will be soon be demonstrated to any order on general grounds. Supposing,


for a moment, that this is true we write the renormalized free part of the effective
action as
Aeff

1
=
2tR

dD x (R )2 + hR R2 .

(22.48)

Comparison with (22.44) gives


Zt Z1 = (1 tL)


N 1
tL
Z1/2 = 1 +
2

(22.49)

and hence
Z = 1 (N 1) tL + . . . ,
Zt = 1 (N 2) tL + . . . ,
Zh = Zt1 Z1/2 = 1 +

N 3
tL + . . . .
2

(22.50)

(22.51)

Thus the temperature is renormalized as follows


1
1
= (N 2) L + . . . .
tR
t

(22.52)

1203

22.3 Symmetry Properties of Renormalized Effective Action

For the simplest case of an O(2)-model, this implies no first-order renormalization


of the temperature at all, i.e., the coupling strength of the a -fields is unchanged by
fluctuations. This property turns out to hold at all orders in t.
Obviously, this result depends crucially on the way the two renormalization factors in (22.43) are distributed over wave function and temperature. It is therefore
worthwhile to go through the general renormalization in some detail.
It is interesting to point out a property of this field theory in D = 1 dimensions
where we may interpret x as an imaginary time and a (x) as a positionvariable
xa ( )/R of a point particle on the surface of a sphere of radius R = 1/ t in N
dimensions. The comparison will be presented in Section 22.8

22.3

Symmetry Properties of Renormalized


Effective Action

The O(N)-symmetry of the theory implies the existence of a Ward-Takahashi identity which has to be fulfilled by the generating functional of all correlation functions:
Z[ja , h] =

D
1

exp
2
t
1
(

 q

dD x (a ) + 1 a2

2

ja a h 1 a2 .
(22.53)

Under a rotation in each of the (N, a)-planes for a = 1, . . . , N 1, by an angle


aN a :
q

a = a 1 c2 ,

(22.54)

1 a2 = a a ,

(22.55)

the source terms go over into


Z

d x ja a h 1
=

a2

 q

+ ja 1

c2

ha a


q

d x (ja ha ) a + (h + ja a ) 1

c2

(22.56)

This implies that Z[ja , h] must be invariant under the replacement


ja ja + ha ,
h h ja a ,

(22.57)
(22.58)

which changes Z[ja , h] by


Z[ja , h] =

"

h
a
Z a .
ja
h

This change has to vanish for any a .

(22.59)

22 Critical Properties of Non-linear -Model

1204

In order to find the consequences for the connected correlation functions, we go


over to their generating functional
W [ja , h] = t log Z[ja , h].

(22.60)

Defining the effective fields as usual by the functional derivatives of W [ja , h] with
respect to the currents,
a

W
,
ja

(22.61)

the effective energy at a fixed magnetic field h is given by the Legendre transform
[a , h] = W [ja , h] +

dD x a (x)ja (x),

(22.62)

which contains the information on the currents in the derivative


ja =

.
a

(22.63)

The symmetry property (22.59) of Z[ja , h] can then be expressed as


Z

"

=0
d x h(x)a (x) +
h(x) a (x)
D

(22.64)

We now separate [a , h] into an h-free part (0) [a ], and a part containing h. Since
h is small, we may keep only the linear term in h, and continue with
[a , h] = (0) [a ] +

dD x h(x)(1) [a ](x) + . . . .

(22.65)

For the first two expansion terms, Eq. (22.64) implies


(0) [a ]
= 0
a (x)
(
)
Z
(1)

(1)
[a ](x)
D
d x a (x ) + [a ](x )
= 0.
a (x )
Z

dD x (1) [a ](x)

(22.66)
(22.67)

For the further discussion we shall consider only the case of two dimensions, where
the action is scale invariant and renormalizable by power counting. We focus attention upon the divergent part of the effective action, which also satisfies the differential equations (22.66), (22.67). Since the theory is renormalizable, the infinities
appear only in quite restricted functional forms, the terms which have the same
R
R
form as those in the original, i.e., d2 x(a )2 and d2 x(2a )n . Thus we only need
to search for solutions of (22.66) among (22.67) quite restricted local functionals. In
fact, the search for the expansion coefficient (1) [a ](x) can be restricted even more.

1205

22.3 Symmetry Properties of Renormalized Effective Action

Since h has dimension two, (1) [a ](x) must be dimensionless. This implies that in
a functional expansion of (1) [a ](x),
(1)

(1)

[a ](x) = [0](x) +

d2 y (1) [0](x, y)a(y)

1 Z 2 2 (1)
d yd z [0](x, y)a (y)a (z) + . . . , (22.68)
+
2
the only coefficients permitted on dimensional grounds are (2) -functions, i.e.,
(1) [a ](x) is not only a local functional but even a function of a (x):
(1) [a ](x) = (1) (a (x)) .

(22.69)

(1) (1)
+ a = 0
a

(22.70)

Thus Eq. (22.63) implies

which can be integrated right-away to give


(1) (a (x)) =

B 2a (x).

(22.71)

The h-free term (0) [a ] must reflect the O(N)-invariance. Hence it must have the
form
(0)

[a ] =

1 q
1
C 2a
(a )2 +
d xA
2
2
D

2 )

(22.72)

with some arbitrary constant A and C.


The point is now that due to the identity (22.66), the constants C and B are
not independent. Indeed, since
(0)
1
= 2 a +
a
a 2

d2 x

(c i c )2
C 2d

a
a
c i c
i a i c
q
+
(c i c )2 i q
= 2 a +
2
2
C d
(C d )
C 2 C 2
a

= 2 a + q
i2 C d2 ,
2
C d

(22.73)

we can write (22.66) as


Z

d x B

2a

"

(22.74)

(22.75)

q
a
2
a +
C 2d = 0.

C 2
2

Partial integrations on the first term brings this to


Z

d2 x a 2 B 2c

v
u
uB
+t

2c 2 q
C 2c = 0,
C 2d

22 Critical Properties of Non-linear -Model

1206

and we conclude that the two constants B and C must be identical. The terms in
the effective action with divergences have therefore the general form
1
[a ] =
t
(0)

q
A
A
dx
i B 2a
(i a )2 +
2
2
2

2

h B 2a .

(22.76)

This can be brought to the original form by introducing renormalized quantities


(
)
 q
2
Z
q
1
1
1
(0) [a ] =
d2 x
i 1 2R a hR 1 2R a (22.77)
(i R a )2 +
tR
2
2

If we rewrite (22.76) as
i2
AB Z 2 1 h
dx
(B 1/2 a )
[a ] =
t
2
)
2
 q
q
1
1
1/2
1 (B 1/2 a )2 , (22.78)
+ 1 (B 1/2 a )2 hA B
2


(2)

we identify
R a = B 1/2 a ,
AB
1
=
,
tR
t
hR = A1 B 1/2 h,

(22.79)
(22.80)

implying a renormalizability of the theory in terms of the two renormalization constants:


Zt = AB,
Z = B.

22.4

(22.81)

Critical Behavior in 2+ Dimensions

Just as in the 4 -theory where we studied the renormalization group in 4 dimensions, we shall now consider 2 + dimensions for > 0. The naive scale dimension
of t is , where is a mass. We therefore go over to a dimensionless renormalized
temperature by replacing in Eq. (22.52)
1

.
tR
tR

(22.82)

Then the renormalization equations for tR reads

1
1
= (N 2)L = Zt .
tR
t
t

(22.83)

1207

22.5 Critical Exponents

We now calculate L in 2 + dimensions. According to Eqs. (11.139) and (11.137),


the integral diverges near = 0 as follows
L=

1
dD q
h/2
= S
D 1 +
D
2
(2) q + h
2
2
2
1
=
SD h/2 + O(1 ).

(22.84)

We use once more the mass scale and remove from this integral the divergent
expression
1
Ldiv =
SD .

(22.85)

1 1

= + (N 2)
SD ,
tR
t 2

(22.86)

Then (22.84) can be written

or

t = tR

1
1 + (N 2)
SD tR

(22.87)

From (22.86) we calculate the -function


2
1
(tR ) t1
SD t2R + . . . .
R tR = tR tR |t, = tR (N 2)

(22.88)

This has an ultraviolet-stable fixed point at


tcR =

22.5

+ ...
(N 2)

(22.89)

Critical Exponents

We now use the above results to calculate the critical exponents at the one-loop
level. We first consider the exponent governing the way in which the correlation
length goes to infinity as tR approaches the critical point tcR from above:
(tR tcR ) .

(22.90)

The following analysis is very similar to that in Chapter 21, except for one important
difference: the physical variable which depends directly on the temperature of the
system is now tR (). In the 4 -theory it was m2 () T /Tc 1 . By analogy,
we set here = t/tc 1, ao that the behavior (22.90) amounts to .
The most direct way to calculate the critical exponent is based on the observation that physically observable quantities cannot have anomalous dimensions.
Since the coherence length is such a quantity, it satisfies (t) = 0, and after
expressing (t) in terms of in terms of the renormalized temperature tR , which

22 Critical Properties of Non-linear -Model

1208

depends itself on , we obtain the renormalization group equation analog of the


renormalization group equation (21.34) without a -term (recall Eq. (22.18)
"

(tR , ) = 0,
+ (tR )
tR

(22.91)

this being an analog of the renormalization group equation (21.34) without the terms. Now, dimensional analysis implies that must depend on the two arguments
as follows
R ),
(tR , ) = 1 (t

(22.92)

R ), the differential equation becomes


so that for (t
#

"


(tR ) = 0.
1 + (tR )
tR

(22.93)

This is solved by
R ) = const exp
(t

(Z

1
.
dt
(t)

tR

(22.94)

We may choose the correlation length at zero temperature to coincide with 1 ,


which amounts to the initial condition
= 1.
(0)

(22.95)

The integral is singular at t = 0, and a subtraction is necessary to extract a finite


result. For this we make use of the explicit form of the singularity, (t) 1/t fund
in (22.84), and rewrite (22.94) as follows
R ) = const
(t

1/
tR

exp

(Z

tR

"

1
1
dt

(t) t

#)

(22.96)

For tR tcR , this equation behaves like


R ) const
(t
const

(tcR )1/

exp

(tcR )1/ (tR

(Z

tR

1
dt c
(tR )(t tcR )

c
tcR )1/ (tR ) .

(22.97)

From this we identify the critical exponent of the coherence length as


= 1/ (tcR ).

(22.98)

In Eq. (22.88) we found


(tcR ) = Z(N 2)
SD tR ,

(22.99)

1209

22.5 Critical Exponents

which is equal to at tR = tcR . Thus


= 1 .

(22.100)

This result is independent of N, which is the reason why it coincides the exact result
for N found in Chapter ??. This is to be compared with the O(N)-symmetric
4 -theory where we obtained in 4 dimensions [recall Eq. (21.189)]
=

1
N +2
+
.
2 4(N + 8)

(22.101)

The relation between the two expansion parameters is = 2 . Since both models
describe the same physical system, they should have the same critical exponents.
However, the expressions (22.98) and (22.101) have quite a different N-dependence.
This shows that none of the two results can really be trusted in three dimensions.
Apparantely, the parameter = 1 is too large in either case to justify using the
lowest-order term in the -expansion.
Let us calculate the critical exponent of the spontaneous magnetization M
in the ordered phase. It isdefined by the power law in which the renormalized
expectation of the field = 1 2 goes to zero as tR approaches the critical point
[recall Eq. (21.157)]:
M hR i R ( ) ;

tR tcR 0

(22.102)

Since = 1 2 has obviously a vanishing engeneering dimension, the magnetization M = can depend only on tR (). Moreover, due to the O(N)-symmetry,
is renormalized in the same way as in Eq. (22.45), so that
R = Z1/2

(22.103)

Explicitly,
this follows from
q Eq. (22.76). There the field has gone over into into
q
1/2
B 2a which is B
1 2Ra , thus confirming (22.103) [recalling (22.81)].
Writing the magnetization M(tR , ) as M(tR ()), we see that it satisfies the
renormalization equation
"

(tR )
+ (tR ) M(tR ) = 0
tR

(22.104)

(tR ) = Z1/2 Z1/2 .

(22.105)

where

According to (22.50) and (22.86), the lowest-order result is


(tR ) =

(N 1)
.
2
SD tR

(22.106)

22 Critical Properties of Non-linear -Model

1210
Equation (22.104) is then solved by
"

M(tR ) = exp

(t )
.
dt
(t )
#

(22.107)

Near the critical point tcR , this behaves like


c

M(tR ) ( ) (tR )/ (tR ) (tR ) ,

(22.108)

Recalling the definition (21.166) we identify the the critical exponent of spontaneous
magnetization [not to be confused with the -function (22.88)] as
=

(tcR )
.
(tcR )

(22.109)

We shall denote (tcR ) and (tcR ) by and , respectively, so that = / .


Inserting the lowest-order -expansions, (N 1)/2, /(N 2), we
obtain
=

N 1
+ O()
2(N 2)

(22.110)

This is to be compared with the -exponent of the O(N)-symmetric 4 -theory.


Takong the scaling expression (21.167) and inserting D = 4 , 1/2 +
O( ), O(2 )] we see that
= 1 + O( )

(22.111)

Only the N = -values agree with each other, the low-N exponents are quite
different, indicating again that in three dimensions, = = 1 is somewhat too
large in either expansion.
Let us finally calculate the anomalous dimension of the fields. The renormalized
two point function of fields G(2) (p) satisfies the renormalization group equation
"

+ (tR )
+ 2 G(2) (p, tR , ) = 0.
tR

(22.112)

This can be rewritten as


"

+ 2 G(2) (p, tR (), ) = 0


d

(22.113)

with the running coupling constant satisfying


"

+ 2 G(2) (p, tR (), ) = 0


d

(22.114)

with the running coupling constant satisfying

d
tR () = (tR ).
d

(22.115)

1211

22.5 Critical Exponents

The solution reads


G

(2)

"

(p, tR (), ) = exp 2

d
( ) G(2) (p, tR (1 ), 1 ).

(22.116)

!2

(22.117)

Near the critical point, this behaves like


(2)

G (p, tR (), )

G(2) (p, tcR , 1 ).

We now observe that the Fourier transformed Green function


(2)

G (p) =

dD xeipx hR (x)R (0)i

(22.118)

contains now field R (x) with a vanishing naive dimensions. Then G(2) (p) has the
naive dimension D, in contrast to the case of an ordinary scalar field (x) where
the dimension of G(2) (p) is 2 for all D. Thus (22.117) becomes
G(2) (p, tR (), )

!2 +D

G(2) (p/1, tcR , )

(22.119)

Going to the critical point, tR () = tcR , we see that G(2) (p) has to behave like a pure
power in

G(2) (p, tcR , ) pD (p/)2 .

(22.120)

From this we extract the exponent of the critical correlation function


= + 2 ,

(22.121)

so that to lowest order in :


=

+ O(2 ).
N 2

(22.122)

Recall that is defined by the anomalous behaviour of the Green function in x


space, i.e.
G(2) (x) = hR (x)R (0)i

1
x

(22.123)

corresponding to G(2) (p) pD . This is in contrast to the general form for the
O(N)-symmetric 4 -theory in 4 dimensions, where has the naive dimension
(D 2)/2 and G(2) (x) x(D2+) . In that theory, the exponent was [recall
Eq. (21.188)]
=

N +2 2
+ O(3 ).
2(N + 8)

(22.124)

22 Critical Properties of Non-linear -Model

1212

The different N-dependence of (22.122) and (22.124) show once again that in three
dimensions, = = 1 is somewhat too large for either expansion.
It should be realized that due to the vanishing naive dimensions of the Goldstone
fields (x) for any D, all the scaling relations are different from those derived in
Chapter 21 for the 4 -theory. In order to find all scaling rules, we can start out
from the renormalization group equation obeyed by the general N-point function
(

+ (tR )
+ n (tR ) + h (tR )hR hR (n) (p, tR , hR , ) = 0
tR

(22.125)

where [compare with (21.30) and (21.33)]




hR
h
h

=
Zt1 Z1/2 |t,h
hR
h
hR




1 1 1/2
h
tR
1/2

Z + Zt Z log Z
=
hR
t
2




h
1
1 1/2
1 1/2
=
Zt Z + (tR ) + Zt Z
hR
tR
"
#
(tR )
= h +
+ .
tR

(22.126)

Near the critical point where (tR ) vanishes, this implies


hR ()

hR (1 )

(22.127)

Together with

d
tR () (tR tcR ),
d

(22.128)

we have
tR () tcR

!1/

(22.129)

from which we conclude that 1 goes to zero near the critical point as follows

(22.130)

This implies that hR (1 ) vanishes according to the relation


hR ()

( ) ()
hR (1 )

(22.131)

1213

22.5 Critical Exponents

The general vertex function (n) satisfies therefore the scaling relation

(n)

(p, tR (), hR (), ) = exp n

d
( )(n) (p, tR (1 )hR (1 ), 1) .

(22.132)
)

Since the naive dimension of (n) and hR are D and 2, respectively, we can rescale
the right-hand side and get in the critical regime
1

!D+n

(n) p , tcR , hR (1 )
1
1

thus arriving at a scaling relation

!2

, ,

(22.133)

(n) (p, tR , hR , ) D ()n (n) p, tcR , hR ()()D , .

(22.134)

In a similar fashion we derive the scaling relation for the vertex functions in the
presence of background fields and [in analogy with (21.93) and (21.94)]. Since
both fields appear O(N)-symmetrically, we can suppress one of them

(n) (p, tR , (), (), ) = Dn (n) p, tcR , /, , .

(22.135)

From this we may derive all critical exponents of the system, and relate them to
and . The exponent of magnetization calculated in (22.109) satisfies the scaling
relations

= = (D 2 + ),
(22.136)
2
wich is the same as that derived in 4 -expansion in Eq. (21.167).
Let us also find the critical exponent of the specific heat C [recall (21.148)].
Since (0) D D , and C measures the second temperature derivative of (0)
we see that
= D 2

(22.137)

just as before in (22.112).


The critical exponent of the susceptibility [see (21.155)] is obtained by differentiating (0) twice with respect to the field strength
1

2 (0)

D+2 (2 D) .
2

(22.138)

Hence
= (2 D) = ( 2)

(22.139)

as before in (21.156). Alternatively, we could have used d2 (0) /dh2R , with the
same result.

22 Critical Properties of Non-linear -Model

1214

For the exponent of the critical magnetization [see [21.157)], we differentiate


with respect to and see that hR must behave like
(0)

hR

(0)

D+ ( ) .

(22.140)

If this relation is to remain nontrivial for 0, we have the condition


D + + = 0,

(22.141)

or
= 1 +

D+2
D
=
,

D2+

(22.142)

just as in (21.160). The same critical exponent would of course follow from the
relation
R

h0
1/
hR .
hR

(22.143)

Thus, even though the derivation was completely different, all scaling relations are
exactly the same as those found in the O(N)-symmetric 4 -theory. The basic reason

is that since and D+ , , the equation of state

j = D+ v (tcR , , )
 

= f

(22.144)

follows Widoms general scaling form (21.176)


h

=f

M
M 1/


22.6

(22.145)

Two- and Three-Loop Results

For completeness, let us also write down the result up to the three loop level. The
two-loop diagrams are shown in Fig. 22.1. We have omitted the corresponding
diagrams involving the 1/aD terms, for brevity. The three loop diagrams are quite
numerous. In order to display them we drop the dashed and wiggly lines and show
only the topological structure in Fig. 22.2. The two renormalization constants have
the following 1/-singularities
Zt

(N 2)2 N 2 2
N 2
tR
tR +
+
= 1+

2
2
"
#
(N 2)3 7 (N 2)2 (N 2)(N + 2) 3
+
tR + . . . , (22.146)
+
+
3
6
2
12
"

22.6 Two- and Three-Loop Results

1215

(N 1)(N 32 ) 2
N 1
tR +
tR
(22.147)

2
"
#


N 1
(N 1)(N 2) (N 2)(N 1) 3
19
15
2
+
+
tR + . . . .
3N + N
+
3
3
2
2
32
4

Z = 1 +

From this one finds the functions (tR ) and (tR ) [recalling (22.88) and (22.105)]
(tR ) =

tR
1 + tR tR log Zt

1
= tR (N 2)t2R (N 2)t3R (N 2)(N + 2)t4R + . . . ,(22.148)
4
d log Z
3
2 (tR ) = (tR )
= (N 1)tR + (N 1)(N 2)t3R + . . . . (22.149)
dtR
4
The critical coupling tcR defined by (tcR ) = 0 has the -expansion
tcR =

2
6N
3

+
+ ... ,
N 2 (N 2)2
4 (N 2)3

(22.150)

and the critical exponents are


3
2
+
+ ... ,
(22.151)
N 2 2(N 2)
2
N(N 1)
3

(N 1)
+
+ . . . . (22.152)
= + 2 (tcR ) =
N 2
(N 2)2
2
(N 2)3

1 = (tcR ) = +

Figure 22.1 Two-loop diagrams

Figure 22.2 Three-loop diagrams

22 Critical Properties of Non-linear -Model

1216

From the scaling law = (D 2 + )/2 we obtain the critical exponent magnetization
=

(N 1)
3 N 1 2
N 1

+
+ ...
2
2(N 2) (N 2)
2 (N 2)3

(22.153)

These -expansion show a very poor convergence at = 1, where they are supposed
to describe the three-dimensional situation. Moreover, the series for 1 has no
alternating signs, indicating that it may not even be Borel-summable. One might
nevertheless attempt a Borel-Pade resummation.
Consider, for example, the case of N = 3 (vector model) where
1
1 = + 2 + 3 + . . . .
2

(22.154)

It is the Borel transformation of the function


t2
t3
+ 3
+ ... ,
2!
2 3!

B (t) = t + 2

(22.155)

i.e.
1 =

et B (t).

(22.156)

Let us take a simple [1, 2]-Pade approximation for B (t). This is a ratio of a polynomial of degree one in the numerator and of degree two in the denominator, which
has the same small-t expansion as in (22.155). It reads
B (t) =

t
2 2

1 t+ t
2
6

(22.157)

and it is easy to verify that it has the expansion (22.154). Inserted into (22.156) we
obtain at = 1:
1 =

et

t
1.25.
t t2
1 +
2
6

(22.158)

Note that the [1, 1] approximant would have a pole for t > 0 preventing its use.
To resum similarly the series (22.152) for , we may use a [2,1] Pad`eapproximation:
t3
22 t2
+ 33 + . . .
2!
3!
1 t/2
,
= t
1 + t/2

B (t) = t

(22.159)

22.7 Variational Resummation of -Expansions

1217

to find at = 1:
=

et t

1 t/2
= 0.11.
1 + t/2

(22.160)

These results may be compared with the well-known critical exponents of the threedimensional classical Heisenberg model:
1 1.39,
0.04.

(22.161)
(22.162)

The numbers for 1 agree reasonably well with each other considering the arbitrariness of the resummation procedure. The numbers for agree insofar as that
both are small.
Recently, the -expansions have been driven to order 4 with an entirely different
technique using perturbation expansions in 1/N.1 The results for the forth-order
contributions to the series (22.148), (22.149), (22.151) and (22.152) are

i
N 2 h 2
N 22N + 34 18 (N 3) (3) t5R ,
12
(N 1)(N 2)
[4 (N 5) + 3 (N 3) (3)] t4R ,
=
12

i
4
1 h 2
N 7N + N 2 18 (N 3) (3)
,
=
12
(N 2)4
h
i
4
= 30 14N + N 2 + (54 18N)(3)
,
4(N 2)3

(22.163)

5 =
2 4
tcR4
41

4 = (N 1) 6 + 2 N + N 2 + (12 + N + N 2 ) (3)

(22.164)
(22.165)
(22.166)
i

4
. (22.167)
4 (N 2)4

The additional terms do not improve the critical exponents if only Pade-Borel resummation techniques are used. Another, more powerful method is needed to extract
precise numbers.

22.7

Variational Resummation of
-Expansions

Having seen the difficulties in extracting accurate critical exponents from the above
-expansions, we decide to resum the above series using information from the 4 expansions of the 4 -theories. This becomes possible by rewriting the latter as
power series in the variable 2(4 D)/(D 2). The initial five expansion
coefficients of these functions are known from -expansions of critical exponents
1

W. Bernreuther and F.J. Wegner, Phys. Rev. Lett. 57, 1383 (1986); I. Jack, D.R.T. Jones,
and N. Mohammedi, Phys. Lett. B 220, 171 (1989), Nucl. Phys. B 322, 431 (1989); N.A. Kivel,
A.A. Stepanenko, A.N. Vasilev, Nucl.Phys. B 424, 619 (1994) (hep-th/9308073).

1218

22 Critical Properties of Non-linear -Model

in O(N)-symmetric 4 -theory in D = 4 dimensions. The first four expansion


coefficients of the strong-coupling expansions are determined from (22.151), (22.152),
(22.166), and (22.167). Recall that the critical exponent of the classical Heisenberg
model with N = 3, which governs the divergence of the coherence length as
|T Tc | , the renormalization group treatment of seven-loop expansions in three
dimensions yields = 0.7073 0.0030,2 whereas five-loop expansions in D = 4
dimensions [3] extrapolated to = 1. lead to = 0.7050 0.00553. Apart from
the expansion coefficients, the latter results incorporate information on the largeorder growth contained in the tip of the left-hand cut in the complex couplingconstant plane. Results very close to these were recently obtained from a novel
strong-coupling 4 -theory4 using the six-loop expansions available in the literature.
It is generally accepted that, as a consequence of the universality hypothesis of
critical phenomena of all systems with equal Goldstone bosons, the same critical exponents should be obtainable from renormalization group studies of O(N) nonlinear
-models in D = 2 + dimensions at = 1, unless the second-order character of the
transition is destroyed by fluctuations. The fact that the latter expansions have remained, up to now, rather useless for any practical calculation due to their non-Borel
character has led some authors to doubt the use of such expansions around the lower
critical dimension altogether5 . This would be quite unfortunate, since it would jeopardize other intreresting theories such as Andersons theory of localization6 . Basis
of these doubts is the increasing relevance of higher powers of the derivative term.7
Hopes for the utility of these expansions are however not completely lost, since the
argument involves an interchange of limits of continuing in and of increasing the
number of derivatives.8 .
In this Section we can confirm these hopes by extracting from the series (22.151),
(22.152), (22.166), and (22.167) supplemented by the five-lopp -expansions around
four dimensions very accurate critical exponents for the classical Heisenberg model
as well as for the universality classes O(N) with N = 0, 1, 2.
When evaluated the series for 1 at = 1, the first series yields for the threedimensional O(3)-model the diverging successive values 1 = (1, 2, 2.5 , 3.25). Also
Pade approximations do not help, the best of them, the [1,2]-approximation, giving
2

R. Guida and J. Zinn-Justin, preprint (cond-mat/9803240). This paper paper uses unpublished
expansions of the renormalization group functions in three dimensions up to seven loops by D.B.
Murray and B.G. Nickel.
3
H. Kleinert, J. Neu, V. Schulte-Frohlinde, K.G. Chetyrkin, and S.A. Larin, Phys. Lett. B 272,
39 (1991) (hep-th/9503230), Erratum ibid. B 319 (1993). See also H. Kleinert and V. SchulteFrohlinde, Phys. Lett. B342, 284 (1995) (cond-mat/9503038).
4
H. Kleinert, Phys. Rev. D 57, 2264 (1998) (E-Print aps1997jun25 001); addendum (condmat/9803268); See also (cond-mat/9801167).
5
G.E. Castilly and S. Chakravarty, Is the phase transition of the Heisenberg model described by
the (2+)-expansion of the Nonlinear -model? Nucl Phys B 485, 613 (1997) (cond-mat/9605088).
6
P.W. Anderson, Phys. Rev. 109, 1429 (1958).
7
F.J. Wegner, Z. Phys. B 78, 36 (1990).
8
E. Brezin and S. Hikami (cond-mat/9612016).

22.7 Variational Resummation of -Expansions

1219

the too large value = 2. So far, the only result which is not far from the true value
has been obtained via the Pade-Borel transform [9]
P [1,2](, t) =

t
,
1 t/2 + 2 t2 /6

(22.168)

from which one obtains the -dependent critical exponent

() =

dt et P [1,2](, t)

(22.169)

whose value at = 1 is 1.252, corresponding to 0.799, still too large to


be useful. The other Pade-Borel approximants are singular and thus of no use, as
shown in Fig. 22.3.
The full -dependence of the power series (22.166) and the Pade-Borelapproximation (22.169) can be seen in Fig. 22.4.
A direct evaluation of the series for the anomalous dimension yields the successive the even worse values (2, 2, 4, 5). Here the nonsingular Borel-Pade approximations [2, 1], [1, 2], and [1, 1] yield 0.147, 0.150, and 0.139, rather than the
correct value 0.032. The full -dependence of the power series (22.167) and the
Pade[2,1]-Borel-approximation are also shown in Fig. 22.4.

22.7.1

Strong-Coupling Theory

The remedy for these problems comes from our theory developed in the reference
quoted in Footnote 4 which allows us to extract the strong-coupling properties of
a 4 -theory from perturbation expansions. In particular, we found power behavior for the renormalization constants at large couplings, and from this all critical
exponents of the system. By using the known expansion coefficients of the renormalization constants in three dimensions up to six loops, we derived extremely accurate
values for the critical exponents. The method is a systematic extension to arbitrary
orders9 of the Feynman-Kleinert variational approximation to path integrals.10 For
an anharmonic oscillator, this so-called variational perturbation theory 11 yields expansions which converge uniformly in the coupling strength and exponentially fast,
1/3
like econstL in the order L of the approximation, as was observed in various
works12 and proved recently.13 . The extension to field theory showed to same type
of convergence, with the power 1/3 replaced by 1 , where is the critical exponent governing the approach to scaling. This field-theoretic development will
9

H. Kleinert, Phys. Lett. A 173, 332 (1993) (klnrt.de/213/213.pdf)


R.P. Feynman and H. Kleinert, Phys. Rev. A34, 5080 (1986) (klnrt.de/159/159.pdf).
11
H. Kleinert, Path Integrals in Quantum Mechanics, Statistics and Polymer Physics, World
Scientific, Singapore 1995, Second extended edition, pp. 1850 (klnrt.de/b5).
12
W. Janke and H. Kleinert, Phys. Lett. A 199, 287 (1995), Phys. Rev. Lett. 75, 2787 (1995)
(quant-ph/9502019).
13
H. Kleinert and W. Janke, Phys. Lett. A 206, 283 (1995) (quant-ph/9509005). A convergence
proof for the anharmonic oscillator which is completely equivalent our results in [16] was given by
R. Guida, K. Konishi, and H. Suzuki, Annals Phys. 249, 109 (1996) (hep-th/9505084).
10

22 Critical Properties of Non-linear -Model

1220

now be combined with an earlier interpolation procedure14 to resum the non-Borel


expansion (22.151) and (22.166).
Let us briefly recall the interpolation procedure by which a divergent weakP
coupling expansion of the type EL (g0 ) = Ln=0 an g0n can be combined with a strong2/q m
p/q PM
coupling expansion EM (g0 ) = g0
) . Previously treated examples
m=0 bm (g0
[18] were the anharmonic oscillator with parameters p = 1/3, q = 3 for the energy
eigenvalues, and the Frohlich polaron with p = 1, q = 1 for the ground-state energy
and p = 4, q = 1 for the mass. As described in detail in the textbook in Footnote 11,
the first step is to rewrite the weak-coupling expansion with the help of an auxiliary
scale parameter as
 n
L
X
g0
p
an q
EL (g0 ) =
(22.170)

n=0
where is eventually set equal to 1. We shall see below that the quotient p/q
parametrizes the leading power behavior in g0 of the strong-coupling expansion,
whereas 2/q characterizes the approach to the leading power behavior. In a second
step we replace by the identical expression

K 2 + 2 K 2
(22.171)
containing a dummy scaling parameter K. The series (22.170) is then reexpanded
in powers of g0 up to the order L, thereby treating 2 K 2 as a quantity of order
g0 . The result is most conveniently expressed in terms of dimensionless parameters
g0 g0 /K q and (1
2 )/
g0 , where
/K. Then the replacement (22.171)
amounts to

K(1
g0 )1/2 ,
(22.172)
so that the reexpanded series reads explicitly
WL (
g0 , ) = K

L
X

n () (
g 0 )n ,

(22.173)

(22.174)

n=0

with the coefficients:


n () =

n
X

j=0

aj

(p qj)/2
nj

()nj .

For any fixed g0 , we form the first and second derivatives of WL (g0 , K) with respect
to K, calculate the K-values of the extrema and the turning points, and select the
smallest of these as the optimal scaling parameter KL . The function WL (g0 )
WL (g0 , KL ) constitutes the Lth variational approximation EL (g0 ) to the function
E(g0 ). It is easy to take this approximation to the strong-coupling limit g0 .
For this we observe that (22.173) has the scaling form
WL (g0 , K) = K p wL (
g0 ,
2 ).
14

H. Kleinert, Phys. Lett. A 207, 133 (1995) (quant-ph/9507005).

(22.175)

22.7 Variational Resummation of -Expansions

1221
1/q

For dimensional reasons, the optimal KL increases with g0 like KL g0 cL , so that


g0 = cq
g0 = cqL remain finite in the strong-coupling limit, whereas
2
L and = 1/
goes to zero like 1/[cL (g0 /q )1/q ]2 . Hence
p/q

WL (g0 , KL ) g0 cpL wL (cq


L , 0).

(22.176)

Here cL plays the role of the variational parameter to be determined by the optimal
extremum or turning point of cpL wL (cq
L , 0). The full strong-coupling expansion is
2
obtained by expanding wL (
g0 ,
) in powers of
2 = (g0 /q g0 )2/q . The result is
p/q

WL (g0 ) = g0

b0 (
g0 ) + b1 (
g0 )

g0
q

2/q

+ b2 (
g0 )

g0
q

4/q

+ . . . (22.177)

with

1 (n)
(2np)/q
wL (
g0 , 0)
g0
,
(22.178)
n!
(n)
where wL (
g0 ,
2 ) is the nth derivatives of wL (
g0 ,
2 ) with respect to
2 . Explicitly:
bn (
g0 ) =

L
ln
X
X
1 (n)
aj
wL (
g0 , 0) = (1)l+n
n!
j=0
l=0

(p qj)/2
lj

lj
n

(
g0 )j . (22.179)

The optimal expansion of the energy (22.177) is obtained by expanding


g0 = 0 + 1

g0
q

2/q

+ 2

g0
q

4/q

+ ... ,

(22.180)

where 0 = cq
L , and finding the optimal extremum (or turning point) in the resulting
polynomials of 1 , 2 , . . . . In this way we obtain a systematic strong-coupling coupling expansion in powers of (g0 /q )2/q . This is done as follows: We first optimize
the leading strong-coupling coefficient b0 (
g0 ) in g0 , and identify the optimal position
by 0 . Optimizing WL (g0 ) with the expansion (22.180), in 1 , 2 , . . . , yields for the
parameters p = 2, q = 2 at the coefficients 1 , 2 , . . . and optimal bn (
g0 )s by the
equations listed in Table 22.1.
It was demonstrated in [18] how one can now find a variational perturbation
series for functions for which one knows L weak-coupling and M strong-coupling
expansion coefficients. We must merely extend the set of of coefficients a1 , . . . , aL by
M unknown ones aL+1 , . . . , aL+M , and determine the latter via a fit of the resulting
strong-coupling coefficients b0 , . . . , bM 1 to the known ones.

22.7.2

Interpolation

This interpolation procedure will now be applied to the perturbation expansion


(22.166) in 2 + dimensions, considering it as the strong-coupling expansion of a
series in the variable = 2(4 D)/(D 2) = 4(1 /2)/ = /(1 /2):
1 = 4 1 8
n

N 4 2
N 4 3
+ 16

N 2
N 2

32 [52 + 108(3)] [16 + 36(3)] N + N 2

3
+ . . . . (22.181)
(N 2)3

22 Critical Properties of Non-linear -Model

1222

The weak-coupling expansion of this series is [3]


1 = 2

 (2 + N )2
N +2
+ 20 + 3 N + N 2
N +8
2 (8 + N )3

 (2 + N )3

+ 2240 624 N 212 N 2 9 N 3 2 N 4 + 8 12 (8 + N )(22 + 5 N )(3)

8 (8 + N )5

+ 5 568576 + 382144 N + 103920 N 2 + 9532 N 3 + 1142 N 4 + 21 N 5 + 4 N 6

+80 249600 148960 N 42912 N 2 6516 N 3 350 N 4 + 3 N 5 (3) + 80 18 (8 + N )3 (22 + 5n) (4)

80 80 (8 + N )2 186 + 55n + 2n2 (5)

(2 + N )
160 (8 + N )7

+ 945 105091072 106771456 N 47635968 N 2 11768576 N 3 1835504 N 4 122812 N 5 6270 N 6 45 N 7 8 N 8

+15120 57911296 + 46323968 N + 17913728 N 2 + 3869024 N 3 + 514592 N 4 + 46900 N 5 + 1902 N 6 37 N 7 (3)


+15120 47874048 40615936 N 11928064 N 2 1525888 N 3 89408 N 4 3200 N 5 128 N 6
+945 101376 + 61056 N + 13392 N 2 + 1278 N 3 + 45 N
2

+256 945 (8 + N )


4

(4)

2 (3)

345552 + 193822 N + 48749 N + 6506 N 3 + 235 N 4 (5)

+945 56448 (8 + N )3 526 + 189 N + 14 N 2 (7)

(N +2)
5
120960 (8+N )9

+ ... ,

(22.182)

whose numerical forms are for N = 3, 4, 5, 1


N = 3 : 1 = 2 0.45455 + 0.071375 2 + 0.15733 3 0.52631 4 + 1.5993 5
N = 4 : 1 = 2 0.5 + 0.0833333 2 + 0.147522 3 0.499944 4 + 1.47036 5,

N = 5 : 1 = 2 0.538462 + 0.0955849 2 + 0.135442 3 0.469842 4 + 1.34491 5,


N = 1 : 1 = 2 0.333333 + 0.0493827 2 + 0.158478 3 0.539937 4 + 1.78954 5.

(22.183)
(22.184)
(22.185)
(22.186)

Extending these series by four more terms a6 6 + a7 7 + a8 8 + a9 9 , we calculate


the strong-coupling coefficients (22.178) by extremizing (22.177) with (22.180), after
identifying g0 with and the parameters (p, q) with (2, 2). Then the coefficients
a6 , a7 , a8 , a9 are determined to make b0 (
g0 ), b1 (
g0 ), b2 (
g0 ), b3 (
g0 ) agree with (22.187).
The technique of doing this is described in detail in Ref. [18].
In order to see how the result improves with the number M of additional
terms in (22.180), we go through this procedure successively for M = 1, 2, 3, 4.
The successive additional expansion coefficients for the O(N) universality classes
with N = 3, 4, 5, 1 are listed in Tables 22.222.5, respectively. The four resulting curves for 1 () are shown in Figs. 22.422.7. For N = 3, the successive critical exponents at = 1 taken form Fig. 22.4 are (1 , 2 , 3 , 4 ) =
(0.87917, 0.75899, 0.731431, 0.712152). Their M-dependence is plotted in Fig. 22.8
as a function of the variable x = M 2 which makes them lie approximately on a
straight line intercepting the -axis at = 0.705. This extrapolated value is in
excellent agreement with the previous determination [4] from 4 -theories, and somewhat smaller than the seven-loop result in [2]. The other O(N) universality classes
are displayed analogously.
Since the -expansion (22.166) is singular at N = 2, we expect difficulties with
our procedure for N < 2. Indeed, the determination of the expansion coefficients in
Table 22.5 shows a large irregularity for M = 3, and the successive approximations
for 1 in Fig. 22.7 show no sign of convergence with increasing M.

22.7 Variational Resummation of -Expansions

1223

Finally, we plot our highest (M = 4) approximations for N = 3, 4, 5 together


with the large-N approximations for N = , 20, 10, 6 in Fig. 22.12 to see the trend
for increasing N, which shows that the latter for N = 6 is still far from the exact
curve. This can also be seen in Fig. 3 of Ref. [19].
The same calculation is done for the critical exponent . In the variable , the
series (22.167) reads
=

1
2
3
2n
+
8n(N

1)
N 2
(N 2)2
(N 2)3
+16(N 1)

h

6 2n2 + 12 n n2 (3)

4
+ . . . (22.187)
,
(N 2)4

whereas the weak-coupling expansion is [3]


/2 =

 (2 + N )
2+N
+ 272 + 56 N N 2
2
2(N + 8)
8(N + 8)4

+ 46144 + 17920 N + 1124 N 2 230 N 3 5 N 4 32 12 (22 + 5 N ) (3)

(2 + N ) 2
32(N + 8)6

+ 5655552 + 2912768 N + 262528 N 2 121472 N 3 27620 N 4 946 N 5 13 N 6

+16(N + 8) 171264 68672 N 1136 N 2 + 1220 N 3 + 10 N 4 + N 5 (3)


+128 9 (N + 8)3 (22 + 5 N ) (4) + 128 40 (N + 8)2 186 + 55 N + 2 N 2

whose numerical forms are for N = 3, 4, 5, 1

(5)

(2 + N ) 3
+ . . .(22.188)
.
128(N + 8)8

n = 3 : /2 = 5/242 + 0.0183987 0.0166488 2 + 0.032432 3,


n = 4 : /2 = 1/48 + 0.0173611 0.0157657 2 + 0.029057 3,

N = 5 : /2 = 7/338 + 0.0161453 0.0148734 2 + 0.0259628 3,


2

N = 1 : / = 1/54 + 0.01869 0.0176738 + 0.0386577 .

(22.189)
(22.190)
(22.191)
(22.192)

Extending these series by four more terms a6 6 + a7 7 + a8 8 + a9 9 , we calculate


the strong-coupling coefficients (22.178) by extremizing (22.177) with (22.180), after
identifying g0 with and the parameters (p, q) with (2, 2). Then the coefficients
a6 , a7 , a8 , a9 are determined to make b0 (
g0 ), b1 (
g0 ), b2 (
g0 ), b3 (
g0 ) agree with (22.187).
The technique of doing this is described in detail in Ref. [18].
In order to see how the result improves with the number M of additional
terms in (22.180), we go through this procedure successively for M = 1, 2, 3, 4.
The successive additional expansion coefficients for the O(N) universality classes
with N = 3, 4, 5, 1 are listed in Tables 22.222.5, repsectively. The four resulting curves for 1 () are shown in Figs. 22.422.7. For N = 3, the successive critical exponents at = 1 taken form Fig. 22.4 are (1 , 2 , 3 , 4 ) =
(0.87917, 0.75899, 0.731431, 0.712152). Their M-dependence is plotted in Fig. 22.8
as a function of the variable x = M 2 which makes them lie approximately on a
straight line intercepting the -axis at = 0.705. This extrapolated value is in
excellent agreement with the previous determination [4] from 4 -theories, and somewhat smaller than the seven-loop result in [2]. The other O(N) universality classes
are displayed analogously.

22 Critical Properties of Non-linear -Model

1224

Since the -expansion (22.166) is singular at N = 2, we expect difficulties with


our procedure for N < 2. Indeed, the determination of the expansion coefficients in
Table 22.5 shows a large irregularity for M = 3, and the successive approximations
for 1 in Fig. 22.7 show no sign of convergence with increasing M.
Finally, we plot our highest (M = 4) approximations for N = 3, 4, 5 together
with the large-N approximations for N = , 20, 10, 6 in Fig. 22.12 to see the trend
for increasing N, which shows that the latter for N = 6 is still far from the exact
curve. This can also be seen in Fig. 3 of Ref. [19].

Table 22.1 Equations determining the coefficients bn (


g0 ) in the strong-coupling expanq
sion (22.177) and the associated i cn i in (22.180) from the functions bn bn (0 ) and
their derivatives. For brevity, we have suppressed the argument 0 in these functions.

n bn
2 b2 + 1b1 + 12 12b0
(3)
3 b3 + 2b1 + 1b2 + 1 2b0 + 21 12b1 + 61 13b0
4 b4 + 3b1 + 2b2 + 1b3 + ( 21 22 + 1 3 )b0
(3)
(3)
1 4(4)
+1 2b1 + 21 12b2 + 21 12 2b0 + 16 13b1 + 24
1 b0

N 1
b /b
1 0
(3)

(b2 + 1b1 + 21 12b0 )/b0


(3)
(b3 + 2b1 + 1b2 + 1 2b0
(3)
(4)
+ 1 2b1 + 1 3b0 )/b
2 1

6 1

Table 22.2 Coefficients of the successive extension of the expansion coefficients in


Eq. (22.186) for n = 3 determined from M = 1, 2, 3, 4 strong-coupling coefficients
(4, 8, 16, 160) of Eq. (22.187).

n
1
2
3
4

a6
203.827
5.67653
4.25622
3.80331

a7

a8

a9

17.6165
9.04109 15.7331
6.87304 10.0012 12.3552

Table 22.3 Coefficients of the successive extension of the expansion coefficients in


Eq. (22.186) for n = 4 determined from M = 1, 2, 3, 4 strong-coupling coefficients
(4, 0, 0, 221.096) of Eq. (22.187).

n
1
2

a6
a7
a8
147.508
7.91064 37.1745

a9

22.7 Variational Resummation of -Expansions

1225

Table 22.4 Coefficients of the successive extension of the expansion coefficients in


Eq. (22.186) for n = 5 determined from M = 1, 2, 3, 4 strong-coupling coefficients
(8, 8/3, 16/3, 106.131) of Eq. (22.187).

n
1
2
3
4

a6
108.648
10.1408
4.75598
3.57909

a7

a8

a9

60.7217
15.1045 38.9689
7.84272 14.1142 21.6045

Table 22.5 Coefficients of the successive extension of the expansion coefficients in


Eq. (22.186) for n = 1 determined from M = 1, 2, 3, 4 strong-coupling coefficients
(4, 24, 48, 3825.54) of Eq. (22.187).

n
1
2
3
4

a6
413.921
5.25285
442759
5.7343

2
1.5
1
0.5
-0.5
-1
-1.5

a7

a8

a9

12.1104
12450066 196950675
13.7134
25.226
38.0976

Pade-Borel
for 1/
at = 1
at n = 3
[1, 2]

4
[1, 1]

8
[2, 1]

10

[3,1]

4
2

[1, 3]

Pade-Borel
for 1/
at = 1
at n = 3

[2, 2]

10 t

-2
-4

Figure 22.3 Integrands of the Pade-Borel transform (22.169) for the Pade approximants
[1,1], [2,1], [1,2] and for [1,3], [3,1], [2,2] at = 1 Only the last is integrable, yielding
1 1.25183 1/.79883.

22 Critical Properties of Non-linear -Model

1226

2
1.75
1.5
1.25
1
0.75
0.5
0.25

1
2
0
0.5
1.5
Figure 22.4 The inverse of the critical exponent for the classical Heisenberg model
in the O(3)-universality class is plotted as a function of = 4 D The solid curve is
obtained from a variational interpolation of the five-loop -expansion of the 4 -theory
around D = 4 dimensions and the four-loop -expansion (22.166) of the nonlinear -model
around 2 dimensions. The long-dashed curves are the successive approximations which
make use of the strong-coupling expansion (22.187) only up to the first, second, and third
order. The short-dashed curves display the first three and four terms of the -expansion and
its Pade [1,2]-Borel approximations (the lowest curve). The dot corresponds to the sevenloop result in D = 3 dimensions, = 0.7096 The values from our successive interpolation
are (1 , 2 , 3 , 4 ) = (0.87917, 0.75899, 0.731431, 0.712152). These are extrapolated in
Fig. 22.8 to infinite order, yielding = 0.705.

22.8

Relation of -Model to Quantum


Mechanics of a Point Particle on a Sphere

In D=1 dimensions, the coordinate x may be identified as a time t and the field
variable (t) as a normalized position variable x (t) of a massive point particle on
the surface of a sphere. Then the action (22.232) for t = 2/R describes the quantum
mechanics of this system. It has the general geometric form
A=

1Z
dt g (x)x (t)x (t),
2

, = 1, . . . , D.

(22.193)

with a metric


g = R2 +

x x
.
1 x2


(22.194)

The Christoffel symbols defined via the inverse metric g = (g 1) as


1
g ( g + g g )
2

(22.195)

has for (22.194) the simple form


=

1
g x ,
R2

(22.196)

22.8 Relation of -Model to Quantum Mechanics of a Point Particle on a Sphere 1227

2
1.75
1.5
1.25
1
0.75
0.5
0.25

1
2
0
0.5
1.5
Figure 22.5 The inverse of the critical exponent for the O(3)-universality class is
plotted as a function of = 4 D The solid curve is obtained from a variational interpolation of the five-loop -expansion of the 4 -theory around D = 4 dimensions and
the four-loop -expansion (22.166) of the nonlinear -model around 2 dimensions. The
long-dashed curves are the successive approximations which make use of the strongcoupling expansion (22.187) only up to the first, second, and third order. The shortdashed curves display the first three and four terms of the -expansion and its Pade
[1,2]-Borel approximations (the lowest curve). The dot corresponds to the seven-loop
result in D = 3 dimensions, = 0.7096 The values from our successive interpolation
are (1 , 2 , 3 , 4 ) = (0.88635, 0.810441, 0.786099 0.768565). These are extrapolated in
Fig. 22.9 to infinite order, yielding = 0.738.

2
1.75
1.5
1.25
1
0.75
0.5
0.25

1
2
0
0.5
1.5
Figure 22.6 The inverse of the critical exponent for the O(5)-universality class is plotted as a function of = 4 D The solid curve is obtained from a variational interpolation
of the five-loop -expansion of the 4 -theory around D = 4 dimensions and the four-loop
-expansion (22.166) of the nonlinear -model around 2 dimensions. The long-dashed
curves are the successive approximations which make use of the strong-coupling expansion
(22.187) only up to the first, second, and third order. The short-dashed curves display the
first three and four terms of the -expansion. There is no Pade-Borel approximation. The
dot corresponds to the seven-loop result in D = 3 dimensions, = 0.767 The values from
our successive interpolation are (1 , 2 , 3 , 4 ) = (0.89278, 0.842391, 0.820491, 0.802416).
These are extrapolated in Fig. 22.10 to infinite order, yielding = 0.767.

22 Critical Properties of Non-linear -Model

1228

2
1.75
1.5
1.25
1
0.75
0.5
0.25

1
2
0
0.5
1.5
Figure 22.7 The inverse of the critical exponent for the O(1)-universality class (of
the Ising model) is plotted as a function of = 4 D The solid curve is obtained from
a variational interpolation of the five-loop -expansion of the 4 -theory around D = 4
dimensions and the four-loop -expansion (22.166) of the nonlinear -model around 2
dimensions. The long-dashed curves are the successive approximations which make use
of the strong-coupling expansion (22.187) only up to the first, second, and third order.
The short-dashed curves display the first three and four terms of the -expansion. There
is no Pade-Borel approximation. The dot corresponds to the seven-loop result in D = 3
dimensions, = 0.6304 The values from our successive interpolation are (1 , 2 , 3 , 4 ) =
(0.862357, 0.665451, 2.08686, 0.802416). Their position with respect to the accurate value
= 0.767 is shown in Fig. 22.11 showing the failure to converge up to order M = 4.

0.9
0.85

M =1

0.8
0.75

2
4

0.2

0.4

0.6

x = M 2

0.8

Figure 22.8 The three successive approximations (1 , 2 , 3 , 4 )= (0.87917, 0.75899,


0.731431, 0.712152) for n = 3 (Heisenberg model) plotted as a functione of x = M 2
which makes them lie a straight line with the intercept = 0.705.

22.8 Relation of -Model to Quantum Mechanics of a Point Particle on a Sphere 1229

0.9

0.85
0.8

M =1

2
4

0.75
0.2

0.4

0.6

0.8

x = M 1.3

Figure 22.9 The three successive approximations (1 , 2 , 3 , 4 ) = (0.88635, 0.810441,


0.786099 0.768565) for n = 4 plotted as a function of x = M 1.3 which puts them closest
to a straight line with the intercept = 0.738.

0.92
0.9
0.88
0.86
0.84
0.82
0.8
0.78

M =1
2
3
4

0.2

0.4

0.6

x=M

0.8

Figure 22.10 The three successive approximations (1 , 2 , 3 , 4 ) = (0.89278, 0.842391,


0.820491, 0.802416) for n = 5 plotted as a function of x = M 1 which puts them closest
to a straight line with the intercept = 0.767.

0.95
3
0.9
0.85
0.8
2
0.75 4
0.7
0.65
0.2

M =1

0.4

0.6

x = M 2

0.8

Figure 22.11 The three successive approximations (1 , 2 , 3 , 4 ) = (0.862357, 0.665451,


2.08686, 0.802416) for n = 1 plotted as a function of x = M 2 which puts them closest to
a straight line with the intercept = 0.767.

22 Critical Properties of Non-linear -Model

1230
2
1.5

n = 5, 4, 3

n = , 20, 10, 6

0.5
1

0.5

1.5

Figure 22.12 The highest approximations (M = 4) for n = 3, 4, 5 (counting from the


top), and the 1/n-expansions to order 1/n2 for n = , 20, 10, 6 (counting from the bottom). The sixth-order result is obviously still far from the exact one.

and the curvature tensor


R +

(22.197)

is found to be
R = g R =

1
( )
R2

1 



.(22.198)
x
x

+
x
x
x
x

x
x
+

1 x2


Using (22.194), this can be expressed in the form typical for homogeous spaces:
R =

1
(g g g g ) .
R2

(22.199)

The Ricci tensor and curvature scalar are [recall (4.371)]


R = R = (D 2)g ,

R = g R =

(D 1)(D 2)
.
R2

(22.200)

The path integral associated with the action (22.232) can be solved exactly by
time slicing [20]. The resulting time evolution amplitude has an energy spectrum
given by the eigenvalues of the Hamilton operator
H=

h
2
L2 ,
2MR2

(22.201)

where h
2 L2 is th square of the angular momentum operator in N dimensions under
which the sphere is rotationally invariant. The eigenvalues of L2 are l(l + N 2)
with l = 0, 1, 2, . . . . The spectrum of H coincides with that of curved-space
Schrodinger operator
h
2
H=
,
(22.202)
2MR2

22.8 Relation of -Model to Quantum Mechanics of a Point Particle on a Sphere 1231

where is the Laplace-Beltrami operator


1

= gg
g

(22.203)

on surface of the sphere.


Let us investigate whether the above formal evaluation of the Feynman diagrams
in the nonlinear -model produces the same energy spectrum. The answer will be
negative: the spectrum emerging from the -model will turn out to be the spectrum
of a particle near the surface of a sphere. That spectrum is obtained by time-slicing
naively the continuous path integral
Z=

Dna De 2

d {n 2a +[(na )2 1]}

(22.204)

in which a particle moves initially through an N-dimensional Euclidean space, and is


restricted to a D = N 1-dimensional sphere by the -functions. In the time-sliced
version, the particle lies on the surface only at the points where the -function forces
it to do so. The intermediate sections of paths, however, run straight through the
N-dimensional space to the next point on the surface, thereby leaving the surface.
This is why this path integral descibes a particle, near the surface of a sphere. The
difference between the two energies is explained in Ref. [20].
The spectrum near the surface of a sphere in N dimensions is
El =

h
2
(L2 )l ,
2R2 2

(22.205)

with eigenvalues (L22 )l = (l + N/2 1)2 1/4 for l = 0, 1, 2, . . . which have their
origin in a large-argument expansion of a Bessel function
Il+N/21 (z/)

(L2 )l
2z z/
1 2 + ...
e

2z
"

(22.206)
R

1
2
a 2
These appears in the time slices of width of the exponential e 2 d {n a +[(n ) 1]}
when transforming the cartesian cordinates na the spherical coordinates in an Ndimensional space. On the sphere, the Hamiltonian operator (22.201) gives the
energy spectrum
h
2
El =
(L2 )l ,
(22.207)
2
2R
where (L2 )l = l(l + N 1) for l = 0, 1, 2, . . . . The difference between the energies
near and on the sphere has the l-independent value

El =

h
2 (N 1)(N 3)
.
2R2
4

(22.208)

The lowest-order calculation of the renormalization constants in the nonlinear model shows that this calculation procedure produces a spectrum of the type
(22.205).

22 Critical Properties of Non-linear -Model

1232

To see this we consider the renormalized action (22.48) and observe that the
squared mass of the -field is modified by the factor Zh of Eq. (22.51). For the mass
m itsself, this implies
!

h
N 3
L+ ... ,
mR = m 1 + 2
R
4

(22.209)

where we have replaced t by 1/R2 and reinserted h


, thus returning to proper physical
units. The Feynman
integral
(22.38)
for
L
can
easily
be performed for D = 1 using
R
2
2
the formula dp/(p + m ) = , yielding L = 1/2m [compare also (11.123)]. The
mass receives therefore an explicit correction
h
N 3
+ ... .
(22.210)
2R2 4
The quantum mechanical ground state energy of a harmonic oscillator of frequancy
mR is mR h
/2. The ground state energy of th N 1 harmonically fluctuating variables
xa ( ) is therefore
mR = m +

mh
h
2 (N 1)(N 3)
mRh

= (N 1)
+
+ ... .
E = (N 1)
2
2
2R2
8

(22.211)

In the limit m 0, we remain with the ground state energy (22.205) near the
sphere.

22.9

Generalization of the Model

There exists a simple generalization of a point particle on a surface of a sphere: The


spinning top. In fact, the ordinary spinning top in three dimensions is equivalent
to a point particle on the surface of a sphere in four dimensions.15 The action of a
symmetric spinning top with moments of inertia I is
A=

1
2I

dt(12 + 22 + 32),

(22.212)

where i are the angular velocities of the spinning top. If g = ei'(t)/2 denote the
2 2 -rotation matrices specifying the orientation of the top, they are defined by
i = itr(i g 1t g).

(22.213)

Alternatively we can write


1
(22.214)
dt tr[(g 1 t g)2 ],
I
This system has a field theoretic analog in a generalized nonlinear -model defined
by the action
A=

1Z D
A=
d x tr[g 1 g g 1 g].
t
15

See Chapter 8 in the textbook cited in Ref. [1] on p. 80.

(22.215)

1233

22.9 Generalization of the Model

The partition function (22.14) including an external magnetic field of the type
(22.20) becomes for this action
Z =

Dg exp

1
t

dD x tr[g 1 g g 1 g] + h tr[D(g) + D (g)]


1
2aD

dD x log det g ,

(22.216)

where D(g) is some matrix representation of the group elements g. In general,


we may assume g to be elements of any Lie group which can be represented as
exponentials
g = eia Ga ,
(22.217)
where the generators Ga close under commutations relations, with antisymmetric
structure constants fabc :
[Ga , Gb ] = ifabc Gc .
(22.218)
From the exponential form we find the partial derivatives g 1 g as
g 1 g = eia Ga

i
1 h i(a +da )Ga
a
e
eia Ga
.
da

(22.219)

The exponential eia Ga ei(a +da )Ga is expanded in a the Neumann-Liouville or Dyson
series [recall (1.246)]. Abbreviating a Ga by
/ , this gives
i
/ i(
/ a +d
/)

e e

= 1i

d ei / d/
ei / + O(d/
2)

i
i
, d/
] + [/
, [/
, d/
]] + . . . .. (22.220)
= 1 i d/
+ [/
2!
3!


In this way we find


1

g g = iGa

1
1
a + b c fabc + b c d fbde fef a + . . . .
2
6

(22.221)

Normalizing the generators to


1
tr (Ga Gb ) = ab ,
2

ab facd fbcd ,

(22.222)

1
tr[g 1 g g 1 g] = gab () a b ,
2

(22.223)

we obtain

with the metric in field space


gab () = ab

1
1
facd fbce d e
facd fcpe fphf fbhk d e f k + . . . . (22.224)
12
360

22 Critical Properties of Non-linear -Model

1234

From this we calculate the connection


1
abc = fcad fbde e + . . . ,
12
and the curvature tensor (22.197) as
1
Rabcd = (fach fbdh fbch fadh ) + . . . .
12
The expansion (22.224) starts therefore out like
g = Radbe d e + . . . .

(22.225)

(22.226)

(22.227)

The commutation relations (22.218) are satisfied by the adjoint representation


(Ga )bc = ifabc ,

(22.228)

which allows us to rewrite (22.229) as

1
fabe fecd + . . . .
12
In the tensor form of the adjoint representation, the generatos are
Rabcd =

(Gab )cd = ifabe fecd ,

(22.229)

(22.230)

so that we may identify Rabcd = i(Gab )cd . In the case of the Lorentz group, the
generators (22.230) become (Gab )cd = i(gac gbd gbc gad ) as in Eq. (4.65), and (22.229)
coincides with (22.199).
The perturbation expansion and critical behavior of this theory can be discussed
in complete analogy with the previous O(N)-sigma model. One separates the action
(22.232) in the coordinates a (t) into free and interacting parts as
1Z
dt ab a (t) b (t), a, b = 1, . . . , N 1,
(22.231)
A =
2Z
1
Aint =
dt Radbe a (t) b (t) d (t) e (t),
(22.232)
2
and obtains the first-order correction to the free-particle energy density
1
E (1) = hAinti = Radbe h a (t) b (t)ih d (t) e i(t)
(22.233)
2
With a small infrared regulator mass m, the correlation function is
1
ab m|tt |
dq iqt
e 2
=
e
,
(22.234)
q + m2
2m
2
and yields the expectation values at equal times
m
1
,
h a (t) b (t)i = ab ,
(22.235)
h a (t) a (t)i = ab
2m
2
R
the latter in dimensional regularization where dk/2 = 0 that allows us to omit
the Jacobian in the measure of path iintegration. Inserting this into (22.236), we
obtain
1
E (1) = hAint i = R.
(22.236)
8
a

h (t) (t )i =

ab

Notes and References

1235

Notes and References


For a derivation of the critical exponents of the O(N) nonlinear -model see
S. Hikami, E. Brezin, J. Phys. A 11, 1141 (1978). The present knowledge on the
critical exponents is compiled in Table VI of
J.C. LeGuillou and J. Zinn-Justin, Phys. Lett. B 21, 3976 (1980).
For the perturbative analysis of the critical behavior of the generalized -model defined in Section 22.9 see
A. McKane nad M. Stone, Nucl. Phys. B 163, 169 (1980).
For a more recent discussion of critical phenomena see the textbook
H. Kleinert and V. Schulte-Frohlinde, Critical Properties of 4 -Theories, World Scientific, Singapore 2001, pp. 1489 (klnrt.de/b8).

Bibliography
[1] For the more recent determinations with various method see
S.A. Antonenko and A.I. Sokolov, Phys. Rev. E 51, 1894 (1995);
S.A. Antonenko, A.I. Sokolov, and B.N. Shalaev, Fiz. Tverd. Tela (Leningrad)
33, 1447 (1991) [Sov. Phys. Solid State 33, 815 (1991)]; S.A. Antonenko and
A.I. Sokolov, Phys. Rev. B 49, 15901 (1984);
S.A. Antonenko, K.B. Varnashev, Phys. Rev. B 59, 8363 (1994).
[2] R. Guida and J. Zinn-Justin, J. Phys. A 31, 8103 (1998) (cond-mat/9803240).
This paper paper uses unpublished expansions of the renormalization group
functions in three dimensions up to seven loops by D.B. Murray and B.G.
Nickel.
[3] H. Kleinert, J. Neu, V. Schulte-Frohlinde, K.G. Chetyrkin, and S.A. Larin,
Phys. Lett. B 272, 39 (1991) (hep-th/9503230), Erratum ibid. B 319 (1993).
See also
H. Kleinert and V. Schulte-Frohlinde, Phys. Lett. B 342, 284 (1995) (condmat/9503038).
[4] H. Kleinert, Phys. Rev. D 57, 2264 (1998) (E-Print aps1997jun25 001); addendum (cond-mat/9803268); See also (cond-mat/9801167).
[5] G.E. Castilly and S. Chakravarty, Is the phase transition of the Heisenberg
model described by the (2 + )-expansion of the Nonlinear -model? Nucl. Phys.
B 485, 613 (1997) (cond-mat/9605088).
[6] P.W. Anderson, Phys. Rev. 109, 1429 (1958).
[7] F.J. Wegner, Z. Phys. B 78, 36 (1990).
[8] E. Brezin and S. Hikami (cond-mat/9612016).
[9] S. Hikami and E. Brezin, J. Phys. A 11, 1141 (1978).
[10] W. Bernreuther and F.J. Wegner, Phys. Rev. Lett. 57, 1383 (1986);
I. Jack, D.R.T. Jones, and N. Mohammedi, Phys. Lett. B 220, 171 (1989),
Nucl. Phys. B 322, 431 (1989);
N.A. Kivel, A.A. Stepanenko, A.N. Vasilev, Nucl.Phys. B 424, 619 (1994)
(hep-th/9308073).
1236

Notes and References

1237

[11] R.P. Feynman and H. Kleinert, Phys. Rev. A34, 5080 (1986)
(klnrt.de/159/159.pdf).
A similar approach has been pursued independently by
R. Giachetti and V. Tognetti, Phys. Rev. Lett. 55, 912 (1985); Int. J. Magn.
Mater. 54-57, 861 (1986);
R. Giachetti, V. Tognetti, and R. Vaia, Phys. Rev. B 33, 7647 (1986).
[12] H. Kleinert, Phys. Lett. A 173, 332 (1993) (klnrt.de/213/213.pdf).
[13] H. Kleinert, Path Integrals in Quantum Mechanics, Statistics and Polymer
Physics, and Financial Markets World Scientific Publishing, Singapore, 2009,
Second extended edition, pp. 1850 (klnrt.de/b5).
[14] W. Janke and H. Kleinert, Phys. Lett. A 199, 287 (1995).
[15] W. Janke and H. Kleinert, Phys. Rev. Lett. 75, 2787 (1995) (quantph/9502019).
[16] H. Kleinert and W. Janke, Phys. Lett. A206, 283 (1995) (quant-ph/9509005).
[17] A convergence proof for the anharmonic oscillator which is completely equivalent our results in [16] was given by
R. Guida, K. Konishi, and H. Suzuki, Annals Phys. 249, 109 (1996) (hepth/9505084).
Predecessors of these works which did not explain the exponentially fast convergence in the strong-couplings limit observed in Ref. [15] are
I.R.C. Buckley, A. Duncan, and H.F. Jones, Phys. Rev. D 47, 2554 (1993);
C.M. Bender, A. Duncan, and H.F. Jones, Phys. Rev. D 49, 4219 (1994);
A. Duncan and H.F. Jones, Phys. Rev. D 47, 2560 (1993);
C. Arvanitis, H.F. Jones, and C.S. Parker, Phys. Rev. D 52, 3704 (1995) (hepth/9502386); R. Guida, K. Konishi, and H. Suzuki, Annals Phys. 241, 152
(1995) (hep-th/9407027).
[18] H. Kleinert, Phys. Lett. A 207, 133 (1995) (quant-ph/9507005).
[19] H. Kleinert, Strong-Coupling Behavior of 4 -Theories and Critical Exponents,
Phys. Rev. D 57, 2264 (1998) (klnrt.de/257/257.pdf). Addendum in condmat/9803268.
[20] See Sections 8.9 and 10.4 in Re. [13].

It is the nature of all greatness not to be exact


Edmund Burke (17291797)

23
Exactly Solvable O(N )-Symmetric
Four-Fermion Theory in 2 + Dimensions
As in the O(N)-symmetric 4 theory in Section 19.2, there exists also a class of
fermionic field theories which can be solved exactly by introducing collective quantum fields of the type discussed in Section 19.1. Consider a large number N of
identical fermion fields, coupled by an O(N)-symmetric four-field interaction.

23.1

Scalar Self-Interaction

The Lagrange density reads at finite N:


g0  2
a a .
L = a (i/
m0 ) a +
2N

(23.1)

where the index a runs from 1 to N. This model has first been studied in slightly
different forms by Anselm, by Vaks and Larkin, and by Nambu and Jona-Lasinio
to exhibit the phenomenon of spontanoeus breakdown of chiral symmetry.1 In a
so-called chiral version of th model, the latter authors were the first to point out
the existence of a Nambu-Goldstone boson for spontaneously broken continuous
symmetry. In fact, an extended version of the model can be written down which
illustrates the properties of all low-energy phenomena in hadron physics.2 But recent
research has shown that the approximations used in that model are too crude so
that the result is not really true.3 The chiral fluctuations are so strong that the
spontaneously broken symmetry is restored. This will be discussed in Section 23.10.
The inspiration for studying this model came from the microscopic theory of
superconductivity which had just been found by Bardeen, Cooper, and Schrieffer.4
A satisfactory definition of the model is only possible in D = 2 + spacetime
dimensions, where it is renormalizable and called the Gross-Neveu model .
1

See Notes and References.


H. Kleinert, On the Hadronization of Quark Theories, Lectures presented at the Erice Summer
Institute 1976, in Understanding the Fundamental Constituents of Matter , Plenum Press, New
York, 1978, A. Zichichi ed., pp. 289-390. http://klnrt.de/53 .
3
H. Kleinert and B. Van den Bossche, (hep-ph/9907274).
4
J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957).
2

1238

1239

23.1 Scalar Self-Interaction

At the mean-field level, the effective action becomes


h

, = A , =

2 
g0 
d x (i/
m0 ) +
a a
2N


(23.2)

and yields an equation of motion


g0
a a b (x) = 0.
N

i/
m0 +

(23.3)

This equation differs in an important physical aspect from the Bose case. It can
only have a trivial solution a = 0 for the following reason. In fact, a non-vanishing
fermion field expectation in the ground state
a = h0|a |0i

(23.4)

would imply the state |0i to contain a coherent mixture of bosonic and fermionic
quantum numbers. Such a state has never been observed in nature. There seems to
exist a superselection rule forbidding such super-positions. Thus, unlike the Bose
field expectation, a can never become non-zero. Nevertheless, the following discussion will show that the model does exhibit a 0spontaneous symmetry breakdown,

even though the broken symmetry will not be O(N)


but a discrete one. As before,
we shall carry on the discussion in any dimension D.
The generating functional contains linear terms in which the fields are coupled
to fermionic anticommuting sources (x) and (x):
iW [,
]

Z[, ] = e

iA[,]+i(+c.c.) .
DD e

(23.5)

as in
We introduce a collective field fluctuation on the average around g ,
Eq. (19.4), and rewrite this as
Z[, ] =

DD De

dD x a (i / m0 )a +(+c.c.)
2
2g
0

(23.6)

The fields can be integrated out according to the rule (13.97), and yield a generating functional containing only the collective field :
Z[, ] =

DeiAcoll []G

(23.7)

where
(

N
Acoll [] = N
2g0

d x iTr log [i/


m0 ] ,

and
G (x, x )

i
(x, x ).
i/
m0

(23.8)

(23.9)

1240

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

is the propagator of the fermions in the presence of the collective -field, by analogy
with (19.6). In the limit N , the field is squeezed into the extremum of the
action and we obtain the effective action [compare (19.24)]
1 Z D 2
1

d x (x) iTr log [i/


m0 (x)]
[, , ] =
N
2g0
Z
1
+
dD x a (x) [i/
m0 (x)] a (x)
N

(23.10)

yields the equations of motion,


The extremum of [, , ]
[i/
m0 (x)] a (x) = 0,

(23.11)

and
!

1
i
(x) = g0 tr
(x, x) g0
a a (x),
i/
m0
N

(23.12)

where tr indicates a trace over the Dirac indices. As argued above, the field expectation a (x) must vanish, so that the only equation to be solved is
!

i
(x, x).
(x) = g0 tr
i/
m0

(23.13)

It is called gap equation, because of its first analogous appearance in the theory of
superconductivity.
Thus, as far as the extremum is concerned, we may study only the purely collective part of the exact action
1
1
[] =
N
2g0

dD x 2 iTr log (i/


m0 ) .

(23.14)

Let us seek for an extremal constant solution 0 Then the gap equation reduces to
dD p p/ + m0 + 0
(2)D p2 (m0 + 0 )2
Z
d D pE
m0 + 0
= tr(1) g0
.
(2)D p2E + (m0 + 0 )2

0 = ig0 tr

(23.15)

The Dirac matrices have disappeared, except for the unit matrix inside the trace.
This makes it possible to use this equation in any desired number dimensions. We
only need to know the dimension of the Dirac matrices, which is 2D/2 for even D.
In this form, Eq. (23.4) may be extrapolated analytically to any non-integer value
of D.5
5

For details the reader is referred toy Bernard de Wits lecture notes on this subject in
http://www.phys.uu.nl/~bdewit/ftip/AppendixE.pdf.

1241

23.1 Scalar Self-Interaction

For a constant , the effective action gives rise to an effective potential


1
1
1 2
1
v() =
[] =
tr(1)
N
NV
2g0
2

i
h
d D pE
2
2
,
log
p
+
(m
+
)
0
E
(2)D
(23.16)

where V dD x is the total volume of the system. The last term is obtained from
the Tr log in (23.10) by the following calculation
R

1
dD p
log (/
p m0 ) =
D
(2)
2
1Z
=
2
Z
i
=
2
Z

dD p
[log (/
p m0 ) + log (/
p m0 )]
(2)D
i
h
dD p
2
2
log
p
+
(m
+
)
0
(2)D
i
h
d D pE
2
2
.
log
p
+
(m
+
)
0
E
(2)D

The integral is performed with the help of formula (11.134), and yields
1
1 2
2
1
v() =
2D/21 D2 S
D (D/2)(1 D/2)
N
2g0
2
D

m0 +

!D

2 . (23.17)

We have introduced, as usual, and arbitrary mass scale , which will enable us to
study the theory in the limit m0 = 0.
We now focus attention upon the dimensional neighborhood of D = 2 spacetime
dimensions, setting
D = 2 + ,

with > 0.

(23.18)

Then

2
m0 +
1
v() =
b

N
2 g0

where the constant b stands for

!2+

2 ,

2
1
2 /2
2 S
D (D/2)(1D/2) =
(1D/2)
D
D (2)D/2
2
1
=
(/2).
2+ (2)1+/2

(23.19)

b =

(23.20)

For small , the constant b behaves like


b




1
1 log 2e + O().

2


(23.21)

Therefore, the bare parameters must somehow diverge if the theory is to remain
finite in the limit 0.

1242

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

For simplicity, consider at first the massless case, m = 0. Then a renormalized


coupling constant can be defined via
1
1

b
=
.

g0
g

(23.22)

The limit 0 can now be taken at a finite g and we obtain the renormalized
potential
1
1 2 2

v()
+
log
N
2 g

"

!#

(23.23)

For arbitrary 0 < < 4, we may write a renormalized v as


1

v() =
N
2

+ b 2 1
g

"

! #)

(23.24)

which reduces to (23.23) for 0 .


Let us now study the possibility of a nontrivial solution = 0 for the gap
equation (23.15) at m0 = 0. There is no need to perform the integral in that
equation since the result can immediately be obtained from the minimum of v().
Thus we simply differentiate (23.19) for m0 = 0, and obtain
2+
1 = g0 b
2

(23.25)

In the renormalized version (23.24) this reads


1 = gb

"

1+
2

and becomes in the limit 0:

1 ,

(23.26)

(23.27)

0
1
1
.
+ log
1= g

The vacuum expectation of the collective field is therefore given by


0 = e(1/2 + /g) .

(23.28)

The question arises whether this non-trivial solution corresponds to the true ground
state of the problem. For this, we differentiate v() once more, and find
1
D
1
v () =
b (D 1) .
N
g0
2

(23.29)

Inserting (23.26), this becomes


1
1
v () = (D 2) = b (1 + /2) 0 ,
N
g0

(23.30)

1243

23.2 Spontaneous Symmetry Breakdown

which is positive for D > 2 if


g0 < 0.

(23.31)

What does this condition mean for the renormalized coupling g? Using (23.22), we
see that g0 < 0 and g0 > 0 amount to g > g and g < g , respectively, with
g b1
= .

(23.32)

This is a critical coupling constant above which the model has a phase with a nonvanishing field expectation 0 . If the renormalized coupling lies below g , the bare
coupling constant is positive and only the trivial solution 0 = 0 has a positive
v (0 ) indicating stability.
What are the physical properties of the two solutions? Looking back at the
effective action (23.107), we see that 0 increased the fermion mass term to M =
m0 + 0 . In the present zero-m0 case,
M = 0 .

(23.33)

We therefore conclude that for g < g , the massless input fermions remain massless. For g > g , on the other hand, the massless fermions acquire a mass M = 0
from the fluctuations. We observe a spontaneous generation of a fermion mass.
insmass,spontaneous+generation+of+fermionThe result may also be phrased differently: In the weak-coupling phase with g < g -phase, the fermions keep their
initial long-range correlations, in the strong-couping phase with g > g , however,
the spontaneously generated mass limits their correlation functions to a range 1/M.

23.2

Spontaneous Symmetry Breakdown

The spontaneous mass generation is closely related to the fact that the model displays, for zero initial mass, the phenomenon of spontaneous symmetry breakdown
(recall Chapters 27, 19). Indeed, for m0 = 0, the Lagrangian (23.1) possesses an
additional symmetry called 5 -invariance. In two dimensions we may choose the
following -matrices:
0

0 1
1 0

= ,

0 1
1
0

= i 2

(23.34)

which satisfy

{ , } = 2g

=2

1
0
0 1

(23.35)

The hermitian 5 -matrix is defined in analogy with the four dimensional case in
Eq. (4.538) as
5 =

1
= 0 1 ,
2!

(23.36)

1244

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

where is the completely antisymmetric tensor with 01 = 1. In the representation


(23.34), 5 is equal to the Pauli matrix 3 :
5

1
0
0 1

(23.37)

We now introduce 5 -transformations T5 as follows:


5

(23.38)

5.
5 0

(23.39)

T5

which satisfies T25 = 1. Under T5 :


T5

Hence:

(23.40)

If the bare mass m0 in (23.1) is zero, the Lagrangian is invariant under T5 .


Also the action in the exponents of (23.6) and (23.7) are invariant, if we assign
in accordance with (23.36), (23.38) the transformation
to g ,
.
T5

(23.41)

Thus the m0 = 0 collective action (23.8) is symmetric in . It is precisely this


5 symmetry which is broken by the non-vanishing expectation value hi = 0 .
We may compare this result with our Bose discussion in Section 19.4 of the last
Chapter. There we found that the continuous O(N) symmetry could not be broken
in less or equal two dimensions due to fluctuations, due to the fact that such a
phase would contain massless Nambu-Goldstone bosons in two dimensions, whose
correlation functions diverge at any distance. Here we see, in contrast, that a discrete
symmetry can be broken in two dimensions.

23.3

Dimensionally Transmuted Coupling Constant

It is worth pointing out an important structural property of the final result: The
characteristic parameter of the two-dimensional system is the fermion mass which
is determined from the coupling strength and the arbitrary mass parameter via
(23.28), (23.33) as
M = e(1/2 + /g) .

(23.42)

The original theory with m0 = 0 had only a single free parameter, namely the
coupling strength g0 . The auxiliary mass parameter was only introduced for the

1245

23.4 Scattering Amplitude for Fermions

purpose of renormalizing the theory in the massless case. The renormalized coupling
g depends on the choice of , and (23.22) should more explicitly be written as
1
1
b =
.

g0
g()

(23.43)

In this way, a system with a single parameter g0 has been recharacterized by two
parameters and g().
This increase of parameters is certainly an artefact. There must be a relation
between and g() such that different pairs (, g()) correspond to the same set of
Greens functions, i. e., to the same physical theory. Indeed, such a relation follows
from (23.22). At a fixed g0 , we can plot curves in the (, g) plane which correspond
to one and the same theory. In the limit 0 this relation between and g has a
subtlety. It goes over into the mass relation (23.42).
The fermion mass is a physically observable finite quantity. There are infinitely
many pairs of parameter and coupling g() which lead to the same fermion mass,
and this mass is the most economic single parameter by which all properties of the
theory can be expressed. Let us illustrate this by reexpressing the potential (23.23)
in terms of M. For this we write the renormalized gap equation in the form (with
0 = M)
1
M
+ + log
= 0.
g 2

(23.44)

Multiplying this by 2 /2, and subtracting the result from the potential (23.23), we
find
1

2
1
log
.
v() =

N
2
M
2


(23.45)

This has indeed the desired property that neither nor g() appear but only the
single parameter M. The same property can, of course, be verified in D > 2 dimensions. Here we may combine (23.26) with (23.24) and find

1
v() =
b

N
2 g0

!2+

=
b 2 1 +
2
2
M
"

 #

(23.46)

For 0, this reduces to (23.45).

23.4

Scattering Amplitude for Fermions

We calculate the scattering amplitude for fermions. As we know from the discussion
in the Bose case, this is given entirely by the exchange of -propagators. These can
be extracted from the effective action (23.10) with m0 = 0 by forming the second

1246

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

functional derivative at 0 [compare the discussion leading to (19.135) and (19.136)].


The quadratic piece in 0 = M is
"

N 1
=
2 g0
2

d x + i Tr

i
i

i/
M i/
M

!#

(23.47)

From this we extract the propagator of the -fluctuations in momentum space


G (q) =

i
1
,
1
N g0 + (q)

(23.48)

where the self-energy of the field is given by the integral


dD k
i
i
D
(2) k/ M + i (/
k q/ ) M + i
Z
D
d k [(/
k + M)(/
k q/ + M)]
= i tr
.
(2)D (k 2 M 2 ) [(k q)2 M 2 ]

(q) = i tr

(23.49)

with the momenta illustrated in the Feynman diagram shown in Fig. 23.1

.
Figure 23.1 One-loop Feynman diagram in the inverse propagator of the -field.

Forming the trace and going to euclidean momentum space yields


k(k q)E M 2
d D kE
,
(2)D (kE2 + M 2 ) [(k q)2E + M 2 ]

(23.50)

d D kE Z 1
k(k q)E M 2
dx
2.
(2)D 0
(kE2 2kE qE x + qE2 x + M 2 )

(23.51)

(q) = 2D/2

In the last line we have gone to the euclidean form. The denominator can be
treated with the help of the Feynman formula (11.156), and we may write [compare
(11.159)]
(q) = 2D/2

The integrand can be rearranged to

(k qx)2E + (k qx)E qE (2x 1) qE2 x(1 x) M 2


.
2
[(k qx)2E + qE2 x(1 x) + M 2 ]

(23.52)

Upon integration over the shifted momentum k k qx, the second term in the
numerator will vanish since it is odd in k . We therefore remain with the integral
(q) = 2

d D kE
(2)D
0
)
(
1
M 2 + qE2 x(1 x)
2
2 2
2 .
kE + qE2 x(1 x) + M 2
[kE + qE2 x(1 x) + M 2 ]

D/2

dx

(23.53)

1247

23.4 Scattering Amplitude for Fermions

This can be integrated using formulas (11.123), leading to


(q) = 2
Z

= 2
=

D/2

S
D

"

1 (D/2)(2 D/2)
1 (D/2)(1 D/2)
2
2
(1)
2
(2)

dx qE2 x(1 x) + M 2
D/21

iD/21

(23.54)

S
D (D1)(D/2) (1 D/2) M

D(D 1)
b M
2

dx

"

qE2
x(1 x) + 1
M2

dx

#/2

"

qE2
x(1x)+1
M2

#D/21

(23.55)

where we have introduced the parameter b of (23.20). Inserting this into (23.47)
we obtain the propagator in terms of the renormalized coupling g:

1
D(D1)
M
i

b
G =

g0
N
2

i
D(D1)
1
= + b 1

N
g
2

! Z

g2
x(1x) + 1
dx
M2

!Z

"

#/2 1

q2
dx E2 x(1x) + 1
M
0
1

"

The expression in the last curly brackets behaves for small like

#/2 1

.(23.56)

Z1
q2
M

(1 + ) 1 +
dx log E2 x(1 x) + 1 + log
1 1+
2
2 0
M

("Z
)
!#
1

qE2
M
=
dx log
,
x(1 x) + 1 + 3 + 2 log
2
M2

0


"

(23.57)

so that the big parenthesis of (23.48) become, in D = 2 dimensions,


"
(
)!
#
qE2
1 Z1
M
1
dx log
.
+
x(1 x) + 1 + 3 + 2 log
g 2 0
M2

(23.58)

This result can be expressed completely in terms of M. For this purpose, we subtract
again the gap equation (23.27) and find
1
2

q2
dx log E2 x(1 x) + 1 + 2
M
(

"

(23.59)

and the -propagator takes the desired form


1
i
1
N g0 + (q)
2
i
#
"
= Z1
.
2
N
qE
x(1 x) + 1 + 2
dx log
M2
0

G (q) =

(23.60)

1248

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

The integral in the denominator has been solved before in Eqs. (11.171)(11.174)],
where we found, with z qE2 /M 2 ,
J(z) =

dx log [zx(1 x) + 1] = 2 + 2 coth ,

(23.61)

where
v
u

u
z
q2
= atanh
= atanht 2 E 2 , sinh =
z+4
qE + 4M

z
=
4

qE2
.
4M 2

(23.62)

The function J(z) is monotonously increasing in qE2 , with the minimum lying at the
origin, as shown in Fig. 23.2. The propagator itself starts out, in momentum space,
3
2.5
2

2
J(qE
/M 2 ) + 2

1.5
1
0.5
-10

-5

10

2
qE
/M 2

Figure 23.2 The function J(z) + 2 in the denominator pf the -propagator (23.60). FOr
z < 4, the curve shows the real, the dashed curve the imaginary part.

with
G (q)|qE2 =0 = i

,
N

(23.63)

and has a monotonously decreasing size for growing euclidean momentum qE .


We may now ask whether there exists a scalar ground state in the fermion antifermion scattering amplitude, which is usually called -particle in the analogy
with a resonance of in the proton proton scattering amplitude which is seen
at roughly 700 MeV (and which is the origin of using the name for the collective
This particle would have to manifest itself in a pole in the propagator
field .
2
G (qE ) at timelike q 2 , i. e., at a negative value of qE2 = s = M2 . Indeed, the
denominator (23.61) is seen to vanish for
s = 4M 2

(23.64)

as can be seen by continuing (23.61) to 0 < s < 4M 2 using


coth = coth

(23.65)

In terms of s, the parameter is given by


= arctan

s
, sin =
s 4M 2

s
.
4M 2

(23.66)

1249

23.4 Scattering Amplitude for Fermions

Close to 4M 2 the propagator behaves like


G

2i
=
N

s
.
s 4M 2

(23.67)

Thus we see that there is no proper particle pole at s = 4M 2 , but only a branch
cut which runs from s = 4M 2 to infinity. Such a cut is present in every scattering
amplitude and commonly referred to as the elastic cut. This is an artifact of the
leading large-N approximation investigated in the present discussion. For finite N,
the Green function does have a proper bound-state pole before the cut starts.
As in the effective potential, there is only the fermion mass M which characterizes the theory, instead of the bare coupling g0 , since the latter becomes undefined
for 0. The fermion mass M is independent on the particular renormalization
procedure. It may come as a surprise that a dimensionless quantity g0 has been
replaced by quantity with the dimension of a mass. This process is often referred to
as dimensional transmutation. It was first observed in the microscopic theory of superconductivity. There are many superconductors with different coupling strengths
g and mass parameters (which are a characteristic for the phonon spectrum; see
Chapter 18), but there is only one quantity specific for the superconductivity properties which is the critical temperature Tc . Theories with the same Tc are identical
superconductors independent on what g or they were derived from. There is one
important difference, however, between the different cases: In the fundamental Lagrangian (23.1), the mass parameter and the associated g() are not detectable
separately by any physical experiment. Only their combination M is. In a superconductor, on the other hand, both quantities are properties of the microscopic
substructure and can both be measured. This points at an important physical aspect of the renormalization procedure: Every theory which requires renormalization
of the coupling constant has a redundancy in its parameterization with a mass parameter and a renormalized coupling constant g(). This redundancy cannot be
resolved at the level of the theory itself. But there may be a more microscopic theory in which both parameters and g() acquire a separate physical significance.
Until now, theoretical physics has gone precisely this way. Every theory which was
initially considered to be microscopic turned later out to be a phenomenological
description of even more microscopic substructures.
The reader may wonder how this description applies to the phase at small coupling constants g < g or g 0 > 0, where the fermions remain massless and there
are long-range correlations. The potential may still be parametrized in the form
(23.19), (23.24) with m0 = 0:

1
2

v() =
b

N
2 g0

2
=
2

2
b 2
g

"

!2+

2
2

#)

(23.68)

1250

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

There remains the arbitrary mass parameter with the renormalized coupling g
depending on the choice of . There is now no fermion mass in terms of which the
result can be expressed in a renormalization independent fashion. Nevertheless, it
is still possible to substitute the pair of parameters , g(), by a single one whose
dimension is mass. For this we may simply turn the sign of Eq. (23.25) and define
M by
D
1 g0 b
2

(23.69)

Now v() = can be rewritten as


1
M 2 D

b
v() =
+
N
2
2
M
"

 #

(23.70)

and M is a measure for the deviation of the potential from its quadratic shape. The
minimum lies at the origin corresponding to the massless fermions.
Note that this potential exists only for truly larger than zero. For 0, there
is no finite limit. This is due to the fact that only for negative g0 , 1/g0 b can
be compensated to become a finite quantity 1/g in the limit 0.
If we calculate (23.58) for vanishing fermion mass, we obtain

1
D(D 1)
qE2
i

b
G (q) =

g0
N
2
2

!/2 Z

1
0

dx [x(1 x)]/2

i
D(D 1) 2 (1 + /2)
1

b
N
g0
2
(2 + )

qE2
2

!/2 1

(23.71)

This may be expressed in terms of the auxiliary mass parameter (23.69) as


i
G (q) =
N

D
b M
2

1

2 (1 + /2)
1 +
(1 + )

qE2
M2

!/2 1
.

(23.72)

It should be pointed out that if we had calculated propagator G (q) by expanding


the action around the wrong ground state solution, say 0 = 0 for g0 < 0, g > g , the
resulting propagator would show this mistake. This is seen directly in Eq. (23.71),
which is singular at euclidean momentum by having an unphysical tachyon
qE2

"

(2 + )
2
1
=
2
g0 b (1 + /2) D(D 1)

#2/

(23.73)

This may also be expressed in terms of renormalized quantities as


qE2
=
2

"

1
1+
gb

(2 + )
2
(1 + /2) D(D 1)

#2/

(23.74)

1251

23.4 Scattering Amplitude for Fermions

When going to Minkowski space this amounts to a particle pole at


q 2 = qE2 .
This is similar to the situation in 4 theory in four dimensions. Also there we found
such a particle with an imaginary mass which travels faster than the speed of light
and is therefore unphysical (tachyon). There it appeared for very large q 2 , here for
very small q 2 . Since a tachyon can have states with arbitrary negative energy, there
must be another ground state for the theory which lies lower than the zero-field
configuration.
It can be argued that for finite N, positive couplings g0 correspond to another
interesting physical phase for which the collective field (g0 /N)a a is no longer
appropriate. Instead, a field proportional to a a allows now for an economic description. This will becomes clearer after the next section.
There is one more observation we should make in the massive phase. We may
express the potential (23.24) also in terms of g , and find the form
1
1
v() =
N
2 g

"

g
g
1 2 +
g
g

! #

2 .

(23.75)

This form exhibits very nicely the unstable origin for g > g and the stabilization
due to the term 2+ . The potential looks very similar to the previously discussed
4 theory. In fact, for 2 (D 4) it takes exactly this form. The minimum lies
at = 0 = M, where M is the fermion mass
2 g
M
=

D g
"

g
1
g

!#1/

(23.76)

in terms of which the potential may be written in a natural parametrization


1

1
v() = M 2
N
2
g

g
1
g

!"

M


2

2
+
D

D #

(23.77)

The mass scale can be replaced by the physical mass at infinite g


M

2
=
D


1/

(23.78)

Then (23.76) reads


M
g
= 1
1
M
g

!1/

(23.79)

We argued before that g, g play the role of temperature T and critical temperature
Tc in surface layers. There the mass goes with
T
M

Tc


1/

g
1
=

!1/

(23.80)

1252

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

It vanishes at the critical point in which case v takes on a pure power behavior

1
v() D
0
N

!D

(23.81)

This power can also be seen at arbitrary T if is increased to be much larger than
the mass scale M [ultraviolet (UV) limit of the theory].
Note that in the opposite limit of small [infrared (IR) limit], the power behavior
is
1
v() 2
0
N

(23.82)

which corresponds to the g 0, i.e., the free-field limit of the theory. One says,
the theory behaves IR free. Such UV and IR power behaviors are typical at a
critical point. They have been the subject of extended experimental and theoretical
investigation over the past decade. It will be worthwhile to dedicate the next chapter
to the corresponding physical phenomena.

23.5

Nonzero Bare Fermion Mass

Before we come to that, let us shortly indicate what happens to this model if there
is a fermion mass from the beginning, a case which we discarded for the sake of simplicity. We may assume m0 to be positive, since otherwise its sign can be changed by
m0 .
Differentiating (23.19),
a simple 5 transformation under which m0
we obtain the gap equation
0
m0 + 0

b
= 1+

g0
2

!1+

(23.83)

which has a solution 0 > 0 for g0 < 0 and m0 < 0 < 0 for g0 > 0. In
other words, for repulsive interaction the mass becomes larger and for attractive
interaction smaller. The second derivative is

1
D
1
m0 +
v () =
b (D 1)

N
g0
2

!D2
,

(23.84)

and at the solution 0 of the gap equation (23.83):


1
1
0
v (0 ) =
1 (1 + )

N
g0
m0 +
1 ( + m0 ) (1 + )m0
.
=
g0
m0 +


From this we deduce the stability regions for negative and positive g0 .

(23.85)

1253

23.6 Pairing Model and Dynamically Generated Goldstone Bosons

Let us renormalize the effective potential (23.19). Introducing the sum tot
m0 + , whose equilibrium value tot
0 = m0 + 0 is the total finite fermion mass M.
Then

1
tot 2tot m0 + m20
tot
v() =

N
2
g0

!2+

2 .

(23.86)

Here we renormalize the coupling again using Eq. (23.22). The term 2tot m0 /g0
is made finite by defining the renormalized mass as
m0
m
= ,

g0
g

(23.87)

g0
m0
=
= 1 g0 b = (1 + gb )1 .
m
g

(23.88)

i.e.,

The term m20 /g0 in (23.86) is not finite for 0, It needs a trivial additive
renormalization of the vacuum energy. Assuming that this has been supplied, we
find the renormalized potential [compare (23.24)]
1

v() =
N
2

(tot m)2
2
b tot
g

"

tot

#)

(23.89)

in which g and m depend on . The renormalized gap equation becomes


1 = gb

"

1+
2

M
,
M m

(23.90)

where tot
has been replaced by the fermion mass M. At tot
0
0 = M, the effective
potential has the value
1

m(m tot )
2
.
v() =
b
+ b tot
N
2
g
2+
"

(23.91)

Note that the potentials (23.24) and (23.89) for the massless and the massive models
differ by a term
(tot m)2
.
(23.92)
v() =
2
g

23.6

Pairing Model and Dynamically Generated


Goldstone Bosons

The model discussed in the last section is somewhat uninteresting, since the symmetry which is broken is discrete. It is instructive to consider a slightly modified
situation in which there is a spontaneous breakdown of a continuous symmetry.

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

1254

From the Nambu-Goldstone theorem we then expect the occurrence of a massless


particle. Again we consider N fields a in D = 2 + dimensions, but now we
take the Lagrangian to be

g0  T   T
a C a b Cb .
L = a (i/
m0 ) a +
2N

(23.93)

Here C is the matrix of charge conjugation which is defined by [recall (4.596)]


C C 1 = T .

(23.94)

In two dimensions, where the -matrices have the explicit form (23.34), we may use
C = 1:
C = 1 = i 2 .

(23.95)

It is the same matrix which was introduced in the four-dimensional discussion in


Eq. (4.598) as the 2 2-submatrix c of the 4 4 charge conjugation matrix C.
Due to the antisymmetry of C, we have


a C aT

= aT Ca .

(23.96)

As a consequence, the interaction potential in (23.93) is negative for g0 < 0, amounting to an attractive potential.
Now we introduce a collective field by adding to L the term
2
N
g0 T

b ,
2g0
N b

leading to the partition function


Z[, ] =

( Z

exp i

DD D

(23.97)

"

#)


1 T
+ N ||2
a Ca + c.c. +
d x a (i/
m0 )a +
2
2g0
D

In order to integrate out the Fermi fields, we rewrite the free part of Lagrangian in
the matrix form

1 T
C,
2

0
i/
m0
i/
m0
0

C T ,

(23.98)

which is the same as (i/


m0 ) since

T CC T = T T = ,

T C / C T = T /

T = / .

(23.99)

1255

23.6 Pairing Model and Dynamically Generated Goldstone Bosons

But then the interaction with can be combined with (23.81) in the form
1 T 1
iG
2 i

(23.100)

where
!

C T

1
, T T , C


(23.101)

denotes the doubled fermion field and


iG1

C 0
0 C

i/
m0
i/
m0

(23.102)

is the inverse propagator in the presence of the external field . Observe that is
a quasi-real field since related to by the similarity transformation

C T

0 C 0
C 0 0

C T

0 C 0
C 0 0

. (23.103)

For a quasi-real field, G1


must be an antisymmetric matrix in the combined spinor
and functional space, and this can easily be verified:
C
C(/
m0 )
C(/
m0 )
C

!T

C T
(/
T m0 )C T
(/
T m0 )C T
CT

=
=

C
C(/
m0 )
C(/
m0 )
C

(23.104)

, where the transposition applies to the combined space,


sinceC / T C 1 = T = /
T
which is the origin of the negative sign with respect to the relation C C 1 =
of Eq. (23.94). In spinor space, partial integration makes the derivative equivalent

to the manifestly antisymmetric functional matrix 12 ( ). In momentum space,


1

(2)
the kinetic part of iG1
(p , p) = (p + p)iG (p) is antidiagonal in momentum
space, satisfying GT (p, p ) = G (p, p ), since (D) (p + p)p is an antisymmetric
functional matrix. This is necessary to have a nonzero kinetic part in the fermionic
Lagrangian, which reads in terms of the quasi-real field (p):
Z

D D

d p d p (p

) iG1
(p , p)(p)

dD p (p) iG1
(p)(p).

(23.105)

We can now perform the functional integral over the fermion fields according to
the rule (13.96), leading to
Z[j] =

1 T

DD eiN A[]+ 2 ja G ja ,

(23.106)

where A[] is the collective action


1
A[] =
2g0

i
dD x ||2 Tr log iG1
,
2

(23.107)

1256

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

and ja is the doubled version of the external source in analogy to (23.86)


j=

T
C 1

(23.108)

This is chosen so that




+ = 1 j T T j ,

(23.109)

and a quadratic completion gives


1
1
i T
1 T 1
iG + (j T T j) = (T + j T iGT )iG1
( + iG j) j G j. (23.110)
2
2
2
2
Note the sign change in front of 21 j T G j in Eq. (23.106) with respect to the Bose
case, in accordance with the negative relative sign of the source term 12 (j T T j).
In the limit N we obtain from (23.107) the effective action
1
1
i
1 T 1
[, ] =
||2 Tr log iG1
iG a
+
N
2g0
2
2N a

(23.111)

in the same way as in the last chapter for the simpler model with a real -field.
The ground state has = 0, whereas = 0 satisfies the gap equation
1
0
= trG0
g0
2

C 0
0 C

(23.112)

where we may assume 0 to be real, as we shall show later.


As before, we shall consider first the case of zero initial mass m0 . Then the
Greens function is inverted as follows
G0 (x, y) =

dD p ip(xy)
i
e
D
2
(2)
p 0

0
p/
p/
0

C 1 0
0 C 1

,(23.113)

as we can verify by multiplying with (23.102). Thus the gap equation (23.112) is
simply
Z
dD p
1
1
= 2D/2
,
D
2
g0
(2) p + M 2

(23.114)

where we have introduced the notation


M 0 ,

(23.115)

to indicate the significance of 0 as a spontaneously generated fermion mass. Also,


we have taken the trace in Dirac space to be 2D/2 in D dimensions.
The integral in (23.113) can be performed just as before and we find
1
D
= b

g0

(23.116)

1257

23.6 Pairing Model and Dynamically Generated Goldstone Bosons

The effective potential is now for a = 0:


1
i
1
v() =
||2 +
N
2g0
2

dD p
log
(2)D

p/
p/

(23.117)

such that we obtain, after a Wick rotation,


1
1
1
v() = ||2 2D/2
N
g0
2



d D pE
2
2
.
log
p
+
||
E
(2)D

Performing the integral gives


1
1
D
1
v() =
||2 2D/21 S
D (D/2) 1
N
2g0
2
2

||
||
b

2 g0

!2+

2
||D
D

2 ,

(23.118)

from which the gap equation (23.116) can again be recovered by differentiation.
Stability is insured for g0 < 0, i.e., for attractive interactions. For 0 = 0, only the
trivial solution 0 = 0 is stable.
For 0 6= 0, we may use (23.116) and express the potential in terms of M rather
than the bare coupling constant g0 as
1
M
D
||
v() =
b ||2

N
2
2
M
"

! #

(23.119)

As before in (23.22), v() can be expressed in terms of the renormalized coupling


1
1

b
=
,

g0
g

(23.120)

in the alternative form

||2
||
1
v() =
+ b 1

N
2
g

!D2

From either expression, we find in the limit 0:

||2

(23.121)

1
|| 1
1
log
||2
v() =

N
2
M
2
"
#
2
1 ||
1
||
2
=
,
+ || log
2 g

(23.122)

in analogy to (23.23) and (23.45) .


Let us now study the propagator of the complete -field. For small deviations

0 from the ground state value of , we find from (23.111) the quadratic
term
1 2
1
=
N
2

(Z

||2
i
d x
+ Tr
g0
2
D

"

GM

GM

#)

. (23.123)

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

1258

In momentum space, the trace term may be written more explicitly as


dD k
i
i
i
M 2 [ (q) (q) + (q) (q)] 2D/2
D
2
2
2
(2) k M (k q)2 M 2
)
Z
dD k
i
i

+ [ (q) (q) + (q) (q)]


tr[/
k (/
k q/ )] .
(2)D k 2 M 2 (k q)2 M 2
(

In a Wick-rotated form, this becomes


1n 2
2 /M 2 )
M [ (q) (q) + (q) (q)] (q
E
2
h
io
E2 /M 2 ) ,
+ [ (q) (q) + (q) (q)] (qE2 /M 2 ) M 2 (q

(23.124)

where (qE2 /M 2 ) is the previous self-energy (23.49) calculated in (23.55). The


(q 2 /M 2 ) stands for
slightly simpler quantity
E


q 2 /M 2

= i 2D/2

dD k
i
i
,
D
2
2
(2) k M (k q)2 M 2

and is calculated as follows:




qE2 /M 2

d D kE Z 1
1
=2
dx 2
2
D
2
(2) 0
[kE + qE x(1 x) + M 2 ]
Z 1
iD/22
h
1
dx qE2 x(1 x) + M 2
= 2D/2 S
D (D/2)(2 D/2)
2
0
Z 1
iD/22
h
D
= b (1 D/2)
dx qE2 x(1 x) + M 2
.
2
0
D/2

As a result, the action for the quadratic deviations from 0 can be written as
(
!
1 2
V Z dD q
1
=
+ A [ (q) (q) + (q) (q)]
N
4
(2)D
g0

+ B [ (q) (q) + (q) (q)] , (23.125)


with the coefficients
2 /M 2 )
A = (qE2 /M 2 ) (q
E
h
i
D
= b M (D 1)J1 (qE2 /M 2 ) + (1 D/2)J2 (qE2 /M 2 ) ,
2

B = (qE2 /M 2 )
D
=
b (1 D/2)M J2 (qE2 /M 2 ),
2

(23.126)

(23.127)

and the integrals


J1 (z) =

dx [zx(1x) + 1]D/21 ,

J2 (z) =

dx [zx(1x) + 1]D/22 . (23.128)

23.6 Pairing Model and Dynamically Generated Goldstone Bosons

1259

Thus the propagators of real and imaginary parts of the field are
1
i
1
N g0 + A + B
1
i
=
1
N g0 + A B

Gre re =
Gim im

(23.129)
(23.130)

and for the complex fields , :


1
i
(B),


N g 1 + A 2 B 2
0


i
1
g01 + A .
= 2 
2
N g 1 + A B 2

G = 2
G

(23.131)
(23.132)

The expressions (23.129)(23.132) can be made finite by using the gap equation
(23.116). The term involving 1/g0 ,
o
i

nh

D
1
+ A = b M 1 (D 1)J1 qE2 /M 2 (1 D/2) J2 qE2 /M 2 ,
g0
2
(23.133)

depends then only on the parameter M, and remains finite for 0, where it
becomes
i
o
1
1 nh 2
+A
J(qE /M 2 ) + 2 J20 (qE2 /M 2 ) ,
g0
2

(23.134)

where J(z) = 2dJ2 (z)/d|=0 is the function (23.61). The function B needs no
renormalization, and has the 0 -limit
1 0 2
J (q /M 2 ).
0 2 2 E

(23.135)

We now observe that there is a zero mass excitation in the imaginary part of the
-field, the component of which points orthogonal to the real ground state value
0 = M. To show this we consider the denominator of the propagator (23.130):
nh
i
o
1
D
+ A B = b M 1 (D 1)J1 (qE2 /M 2 ) (2 D)J2 (qE2 /M 2 ) .
g0
2
(23.136)

By expanding in powers of z qE2 /M 2 ,


D2
z + O(z 2 ),
12
D4
z + O(z 2 ),
1+
12

J1 1 +
J2

(23.137)

1260

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

we find for small momenta


1
D2
D
+AB =
b M 1(D1) 1+
z
g0
2
12
D2
D
z + O(z 2 ),
= b M
2
4




(2D) 1+

D4
z
12



+O(z 2 )
(23.138)

such that the propagator of (23.130) becomes, expressed in terms of Minkowski


square momentum q 2 = qE2 ,
Gim im =

1
i
2
4M 2 2 + regular part at q 2 = 0.
N D(D 2)b
q

(23.139)

Since b < 0 for D > 2, the residue is positive


Res Gim im

1
4 2
2
4M 2
M ,
N D(D 2)b
N

(23.140)

such that the propagator exhibits a proper particle pole at q 2 = 0. The positive sign
is necessary for a positive norm of the corresponding particle state in the Hilbert
space.
In the limit 0, expression (23.136) becomes
1
1
+AB
[J(z) + 2 2J20 (z)].
0
g0
2

(23.141)

The integral J(z) was calculated in Eq. (23.61). For the integral J20 (z), we find by
a similar calculation
J20 (z) =

2
2
coth
=
=
[J(z) + 2],
2
sinh 2
z+4
cosh

(23.142)

such that we obtain in Minkowski space with with z qE2 /M 2 = q 2 /M 2 :


i
1
2
N
2 tanh
i
4M 2 q 2
1
=
2
.
2
2
N
q
J(q /M 2 ) + 2

Gim im =

(23.143)

The real part of the fluctuating field has for 0 the propagator
Gre re =

23.7

i
1
2
.
N
J(q 2 /M 2 ) + 2

(23.144)

Spontaneously Broken Symmetry

When finding the pole at q 2 = 0, we have said that this particle was a NambuGoldstone boson. In order to justify this association, we have to exhibit the continuous symmetry which has been spontaneously broken by the ground state solution.

1261

23.7 Spontaneously Broken Symmetry

Looking back at the original Lagrangian (23.93), we see that it is invariant under
global gauge transformations
ei , = const.
ei .

(23.145)

Similarly, the collective action (23.97) remains invariant, if the collective field
which is, on the average, equal to a pair of -fields is transformed with twice the
phase angle:
e2i

(23.146)

This invariance has been used before when we chose a real ground state expectation
0 . Any other phase would have given the same physical result. Of course, once this
phase is chosen, the invariance (23.145) is destroyed. Thus the zero-mass particle
is indeed a Nambu-Goldstone particle. It corresponds to excitation whose longwavelength limit reduces to a pure global gauge transformation.
Strictly speaking, this zero mass boson can only exist in dimensions D > 2, as
follows from a very general theorem of Mermin, Wagner, and Coleman (see Footnote 3). Indeed, we have seen before in the Bose case that fluctuations prevent the
spontaneous breakdown of a continuous symmetry, which might be resent at the
mean-field level. Thus we may conclude that if fluctuations were included in the
collective field, the theory would also exhibit this general feature in two space-time
dimensions. In the limit N there are no fluctuations in . Thus Colemans
theorem should be satisfied after including all 1/N corrections. Things are more
subtle, however. In two dimensions, there is a critical coupling strength where a
quasi-ordered state exists. This will be discussed in Subsection 23.8.
The physical interpretation of the field is the following: Due to the attraction
for g0 < 0, the fermions form bound state pairs, called, Cooper pairs, which are
bosons and can form a condensate, just as before the bosonic fields in 4 theory.
In fact, the effective potential for the -field looks qualitatively very similar to
that of the bosonic potential v() of the O(N)-symmetric theory for negative m2 .
The origin is unstable and there is a new minimum at 0 6= 0 with an arbitrary
phase [see (23.100) with m0 = 0]. Just as in the previous model with an interaction

2
(g0 /2N) a a , the opposite sign g0 > 0 does not lead to a spontaneous symmetry
breakdown, and massless fermions remain massless.
Finally we must justify why we have called the vacuum expectation 0 = M
the spontaneously generated fermion mass. Looking back at the collective effective
action (23.111), we see that the 0 fields appear in the form


1 T
1 
T
C,
i/
+ M T C + C =
2
2

M i/

i/
M

T
C

. (23.147)

There is a simple transformation which brings this to the canonical Dirac form.
With the two-dimensional 5 -matrix (23.36), we see that
1 + 5
1 5
T
= 1 + 5 + T C 1 5 , (23.148)
+
C ,

=
2
2
2
2

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

1262
and hence



1 T
1 T
T
T
C + C
C5 5 C ,
2
2

(23.149)

where we have used the projection property


P2 = P
of the chiral projection matrix

(23.150)

1 5
,
2

(23.151)

P = P ,

(23.152)

P
and the fact that
which implies that
P 0 P = 0,

P 0 P = 0 ,

(23.153)

But the 5 -terms in (23.149) vanish due to the relations




C5T C 1 = C 0 1

T

C 1 = 0 1 = 5 .

(23.154)

Thus the mass term in Eq. (23.147) becomes simply M .


The gradient term in (23.147) is invariant under the transformation (23.148):

i/
= i/
.

(23.155)

This follows again from (23.152), implying here that


P 0 P = 0 ,

P 0 P = 0,

(23.156)

In this context it should be mentioned that the whole model could have been
written in terms of -fields defined in terms of -fields as in (23.148) from the
outset. If we supplement the relation (23.149) by

5 =



1 T
1 T
T
C C +
C5 + 5 CT , (23.157)
2
2

where the second parenthesis is again zero, we see that the fermion field part in the
effective action (23.97) can be written for zero sources and mass as




1
i
T
T
a i/
a + re Ta Ca + a Ca im Ta Ca a Ca
2
2
N

( 2 + 2 ),
(23.158)
= a (i/
i5 ) a
2g0

where we have identified


re ,

= im .

(23.159)

1263

23.7 Spontaneously Broken Symmetry

The invariance under global gauge transformation (23.145) becomes, in terms of


-fields, an invariance under the transformation

1
2

ei 0
0
ei

ei 1
ei 2

(23.160)

= ei5 .

(23.161)

Such transformations involving 5 are referred to as chiral transformations , and


play an important role in theories of weak interactions (See Chapter 26). Under the

chiral transformation, and i5 behave like a vector in a plane

e2i5 = cos 2 + sin 2 i5

i5 ei5 5 ei5 = sin 2 + cos 2 i5 . (23.162)


Thus, the transformation (23.146) becomes with (23.149) and (23.157):

cos 2 sin 2
sin 2 cos 2

(23.163)

which leaves the transformed effective action (23.158) chirally invariant.


The ground state breaks chiral invariance since acquires an expectation value
0 = M. The Nambu-Goldstone boson generated by this phase transition is the
massless field . It is for this reason that chiral invariance is believed to be an
important principle of strong interactions among elementary particles. There is a
particle in nature, the pion, whose mass is a particle in nature, the pion, whose
electrically neutral version has a mass of 135 MeV, and lies much lower than any
other strongly interacting particle. One therefore interprets the pion as an almost
Nambu-Goldstone particle of the underlying Lagrangian. It was really in this context
that Nambu initiated the study of chiral symmetry in particle physics.
Finally, let us remark that the inclusion of an initial fermion mass m0 6= 0 is
possible but will not be done here, since it merely makes the discussion more involved
while adding little to the understanding of the model.
In D = 2 + dimensions, the model with the Lagrangian

2 
g0  2 

L = a (i/
m0 i5 ) a +
a a + a i5 a
,
2N

(23.164)

which after the introduction of collective fields and as in Eq. (23.6) corresponds
to a Lagrangian

N  2
+ 2 ,
L = a (i/
m0 i5 ) a +
2g0

(23.165)

and thus to the effective action (23.158), is called the chiral Gross-Neveu model.

1264

23.8

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

Relation between Pairing and Gross-Neveu Model

Both models discussed in the last chapter showed spontaneous mass generation in
the limit of N only for one sign of the bare coupling constant: the Gross-Neveu
model (23.6) only for an interaction
g0  2
a a ,
N

with g0 < 0,

(23.166)

with g0 < 0.

(23.167)

and the pairing model (23.97) only for


g0 T
Ca b C bT ,
N a

For the forces between the fermions described by the fields, the latter case implies
attraction, the first repulsion. In either case, the opposite sign of g0 leaves the
fermions massless for N .
These exactly soluble models may be an idealized limiting descriptions of two
phases of the not exactly solvable N = 1 model


+ g0
2
L = i/
2

(23.168)

for D > 2.6 As before, we can certainly introduce a collective field as in the
partition function (23.6). But for N = 1, the field will fluctuate strongly, such
that the phase properties of the model cannot be derived in this way. Suppose
now that fluctuations do not completely destroy the fact that for g0 < 0 there is
a solution in which a mass is generated spontaneously, i.e., that the N limit
gives at least a qualitatively correct description of the system for g0 < 0.
Then we would tend to believe that also the N = 1 version of the pairing model
T
+ g0 T C C
L = i/
2

(23.169)

should have a solution for g0 < 0 which resembles that of the N limit, i.e., in
which there are massive fermions. But the associated Cooper pair fields would then
describe bound states which carry massless Nambu-Goldstone bosons.
The interesting observation is no that these solutions of the two models are two
different phases of one and the same model model. The point is that the interactions
in the Lagrangians (23.168) and (23.169) are really identical, apart from an opposite
sign of the coupling constant, and a factor two. This follows directly by rewriting the
!
1
interactions in terms of spin up and spin down components of the field =
.
2
For the interaction in (23.168), we obtain
2 = ( 1 + 2 )2 = 2 1 2 ,
()
2
1
2
1
6

I. Ojima and R. Fukuda, Progr. Theor. Phys. 57, 1720 (1977)

(23.170)

23.9 Comparison with O(N )-Symmetric 4 -Theory

1265

We have omitted terms containing squares of the fields 12 = 22 and their conjugates,
since these vanish for Grassmann variables. Proceeding similarly with the interaction
in (23.169), we obtain
| T C|2 = | 1 2 + 2 1 |2 = |22 1 |2
= 4(2 1 ) 2 1 = 41 2 2 1 = 42 1 1 2 ,

(23.171)

such that, indeed,

2

1
= | T C|2 .
2

(23.172)

Thus we expect the following two phases for the N = 1 -model: for g0 < 0 or
g > g , a phase with massive fermions and a spontaneously broken 5 invariance.
Here the system is symmetric under global gauge transformations. The other phase
for g0 > 0, g < g has again massive fermions but, in addition, massless NambuGoldstone modes due to a spontaneously broken global gauge symmetry. Physically,
the second phases will be distinguished from the first by the strong long-range
fluctuations which do not exist in the first phase.
For N , either of the two phases forms the only existing solution of the
two models which differ by the distributes the indices over the four fermion fields in
(23.166) and (23.167).
We pointed out before the analogy between the coupling constant in this model
and the temperature in the euclidean formulation of the model. With this interpretation of g, the behavior of this model looks very similar to that found experimentally
in thin films of 4 He. There is a phase transition at a certain temperature Tc . Above
Tc there are only short-range correlations (the system is normal). Below there are
long-range correlations (the system is super-fluid) due to Goldstone excitations of
the condensate. By the Mermin-Wagner theorem discussed on page 1111, there can
be no Nambu-Goldstone in exactly two dimensions, but there do exist quasi-longrange fluctuations with power-like correlation functions of the type (19.98).

23.9

Comparison with O(N)-Symmetric 4 -Theory

After having observed the possibility of spontaneously generating a mass in a massless theory via fluctuations we may look once more back at the scalar 4 version of
the O(N) model in D = 4 dimensions. Note the different notation for 4 D
in contrast to D 2 in the other sections. In the massless case the potential is
1
N 2 1
1
2
v(, ) = 2a
+ NS
D (D/2)(1 D/2) D/2
2
4g0
2
2
D
(23.173)
which may be written as
1
1 2 b D/2
1
2
v(, ) =


N
2N a 4g0
4

(23.174)

1266

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

with
b

4
1
1 1
2
S
D (D/2)(1 D/2) = S
D 2
D
D
8

(23.175)

Note the opposite, bosonic, sign of the last term (23.173) in comparison with the
fermionic equation (23.17), and the absence in b of a factor 2D/21 with respect to
the fermionic b in Eq. (23.20), which came from the Dirac trace in D dimensions,
and a factor 1/2 caused by the linearity of the Dirac operator in p. The bosonic
b is related to the bosonic constant introduced in the discussion of the nonlinear
-model in Eq. (20.37) by b = 4c /D = b .
The two gap equations are
a = 0,
(23.176)
to be solved by = 0 or a = 0. and
1 2
1
D
a b D/21 = 0.
N
g0
4

(23.177)

A renormalized coupling constant may be introduced by setting


1
1
+ b =

g0
g

(23.178)

and v(, ) becomes

!/2

1
1
2
2

2 .
v(, ) =
a
+ b

N
2N
4 g
2

(23.179)

In the renormalized form, the different signs of the bare coupling g0 > or g0 < 0
correspond to g < g or g > g with g = b 1 8 2 .
There is a O(N)-symmetric phase with a = 0 and 0 6= 0 for g0 < 0, where the
bosons acquire a mass. For g0 < 0 there is only the solution a = 0, 0 = 0 which
is again a symmetric phase, but contrary to the previous one this is massless.
Both phases are stable, since the determinant of the second-derivative matrix
vab (a , (2a )) is nonnegative, where a subscript a abbreviates the derivative /a ,
and (2a ) is the solution of the gap equation (23.177)
Consider, however, the excitations: In the massless phase with g < g , we calculate the propagator G from the quadratic variation
1
2 [, ] = (2)
2

(23.180)

"

(23.181)

where in euclidean space

(2)

N 1
=
+ I(q)
2 g0

23.9 Comparison with O(N )-Symmetric 4 -Theory

1267

with [see (11.154), (11.163), and (11.167)]


I(q) =

D
1
dD k
= 1
2
2
D
(2) kE (k + q)E
2


2 (1 /2)
q2

= c E2
(2 )

!/2

D qE2
b
4
2

!/2

(23.182)

1
D
b 2 .
4
8

(23.183)

where we have replaced b by


D
c 1
2


Thus we obtain a propagator for the collective field :


G =

1
2i

N 1/g0 + c (qE2 /2 )/2

(23.184)

For g0 > 0 with g < g , this is a physically acceptable quantity. For g0 < 0 with
g > g , however, there is a tachyon pole at
qE2
=
2

1
g0 c

!2/

(23.185)

as an indication that we have expanded around the wrong vacuum 0 = 0. We must


insert the spontaneously generated mass and find that this solution must be rejected
since it has a tachyon pole. The same observation was made in Subsection 19.3.1
for D = 4, in the phase with g0 > 0, which here is equivalent to g > g .
Alternatively, we can see that in the gap equation near four dimensions
1
1
1
= S
4 ,

g0
g

(23.186)

a finite renormalized coupling g can only be achieved in the limit 0 for negative
g0 < 0 in which case the 4 potential turns the wrong way around. In this case
the only consistent solution for 0 is the free one with g = 0. We now realize
the difference with the N Fermi case. There were two possible consistent
phases, one in which a mass was spontaneously generated which was the phase
which was the one with g > g , g0 < 0 and another one for g < g where the fermions
remain massless. Here only the phase which remains massless with < g , g0 > 0 is
acceptable.
The four-dimensional theory has no consistent ground state, with = 0, except
for g = 0. In contrast, the three-dimensional does have one for g < g 8 2 .
In the Gross-Neveu model we pointed out the existence of certain power laws
in the massive phase. One concerned the physical mass as a function of g g or
T T c [see (23.77)], the other was a power law for v() at the critical point
g = g , i.e., at T = Tc . For T 6= Tc , this power was still found in the UV limit
M, while for M we found the IR free point law v() 2 .

1268

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

We now show that the O(N)-symmetric 4 theory shows quite a similar power
behavior which, however, is opposite as far as IR and UV limits are concerned. For
this the mass parameter m20 has to be set proportional to T /Tc 1, and the critical
theory will be obtained in the limit m0 0. In order to see this, consider the
potential (23.173):
1
m2
1 2 b D/2 m40
1
v(, (2a )) =
2a + 0

,
N
2N
2g0
4g0
4
4g0

(23.187)

where (2a ) is the function of 2a for which v (, ) = 0:


1 2
m2 b D
1
a = 0 + 1/2 .
N
g0
g0
2 2

(23.188)

Suppose we are in the normal phase a = 0 and 0 =


6 0, and consider the critical
regime with vary small m20 . Then from (23.167), we see to behave as a function
of m20 0 as
(m20 )1/(1/2) .

(23.189)

Inserting this into (23.187), we obtain for the minimal value of v(, (2a )) the power
behavior
vmin (m20 )1+1/(1/2) .

(23.190)

At the critical point with m0 = 0, we have


1
D
1 2
a = + b 1/2 .
N
g0
4

(23.191)

Contrary to the Gross-Neveu model there is no pure power behavior. Only if also
g0 = 0, g = 0 (free theory) or g0 = (g = g ), a pure power behavior is obtained:
2a ,
g0 = 0 g = 0
2
1/2
a
, g0 = g = g

(or ),
(or 0).

(23.192)

The same behavior is found at any coupling strength for 0 (ultraviolet limit)
or (infrared limit), respectively. In the renormalized form of (23.192)

1
1 2
D
a =
b 1
g
N
4

!/2

(23.193)

the two limits are separated by the scale parameter . For small , the potential
itself behaves like
v() (2 )1+1/(1/2) ,

(23.194)

23.9 Comparison with O(N )-Symmetric 4 -Theory

1269

as determined by the first and last term in (23.173). The small- behavior can be
collected in the single formula valid for m20
v()

"

m20

N
2
+
2
m0 g0

!#1+1/(1/2)

(23.195)

which follows from writing (23.188) as


1 2 m20
D
a +
b 1/2 ,
N
g0
4

(23.196)

valid for small and vanishing and reinserting, this into (23.187).
If is interpreted as magnetization M and m20 (T /Tc 1) as the deviation of
the temperature from the critical value, this corresponds to a general power law
v(M) M

+1

T /T c 1
M 1/

(23.197)

which was first observed experimentally by Widom in magnetic systems. In the


present model:
1
= ,
2

+1
2 /2
=
.
2
1 /2

(23.198)

For large (UV-limit), there is again free field power behavior


1 2 m20

+
N a
g0

2a m20
+
N
g

!2

(23.199)

(23.200)

Note that there is great similarity with the Gross-Neveu model, as far as power
behaviors are concerned. But the scaling limits are opposite. In the Gross-Neveu
model, there is only one scale, the fermion mass M or the distance, of g from g (i.e.,
T from Tc ). The mass M separates UV and IR limits. Once at Tc where M = 0
there are pure powers. Away from Tc there are different powers in the UV and IR
limit with M = 0 agreeing with the UV limit. In the 4 theory, the massless theory
at the point m20 = 0 still has freedom in the coupling. It can be anywhere between
0 and g . At both ends there are pure powers. In between there are powers in
the IR and UV limit. If m0 is taken away from zero, there are more power laws in
m20 (T /T c 1) for fields and potential.

1270

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

23.10

Two Phase Transitions in Chiral Gross-Neveu Model

We shall now demonstrate that the chiral Gross-Neveu model in 2 + dimensions


has for a small number N of fermions two phase transitions corresponding to pair
formation and pair condensation.7
In the first transition, fermions and antifermions acquire spontaneously a mass
and are bound to pairs which behave like a Bose liquid in a chirally symmetric
state. In the second transition, the Bose liquid condenses into a coherent state
which breaks chiral symmetry. This suggests the possibility that in particle physics,
the generation of quark masses may also happen separately from the breakdown of
chiral symmetry.8
Starting point of our discussion is Eq. (23.139) for the propagator of the massless
Goldstone modes of the pair field in D = 2 + dimensions:
Gim im

14
=
N

1
1

g
g

!1

M 2

i
+ regular part at q 2 = 0. (23.201)
q2

The sign of the pole term guarantees a positive norm of the corresponding particle
state in the Hilbert space.
The residue of the pole term will allow us to conclude that the chiral model has
two phase transitions. Consider first the case = 0 where the collective field theory
consists of complex field with O(2)-symmetry ei . From the discussion of
the Kosterlitz-Thouless transition we know9 that a complex field system possesses
macroscopic excitations of the form of vortices and antivortices. These attract each
other by a logarithmic Coulomb potential, just like a gas of electrons an positrons
in two dimensions. At low temperatures, the vortices and antivortices form bound
pairs. The grand-canonical ensemble of pairs exhibits quasi-long-range correlations.
At some temperature Tc , the vortex pairs break up, and the correlations becomes
short-range. The phase transition is of infinite order.
We have shown in the above-cited textbook that this transition is most easily
understood in a model field theory involving of a pure phase field (x), with a
Lagrange density

(23.202)
L = [(x)]2 ,
2
where is the stiffness of the -fluctuations. The important feature of the phase field
is that it is a cyclic field with = + 2. In order to ensure that such jumps by
2 carry no energy, the gradient in the Lagrange density needs a modification which
allows for the existence of vortices and antivortices. For details see H. Kleinert,
Proceedings of a NATO Advanced Study Institute on Formation and Interactions of
7

H. Kleinert and E. Babaev, Phys. Lett. B 438 , 311 (1998) (hep-th/9809112).


This has been shown in the paper cited in Footnote 3.
9
See the textbook H. Kleinert, Gauge Fields in Condensed Matter , World Scientific, 1989 Section 11.9 (klnrt.de/b1/contents.html). Or see the lecture H. Kleinert, Proceedings of a NATO
Advanced Study Institute on Formation and Interactions of Topological Defects at the University
of Cambridge, England (cond-mat/9503030).
8

1271

23.10 Two Phase Transitions in Chiral Gross-Neveu Model

Topological Defects at the University of Cambridge, England (cond-mat/9503030)


and the textbook H. Kleinert, Gauge Fields in Condensed Matter , World Scientific,
1989 ((klnrt.de/b1/contents.html).
After including vortices and antivortices
at positions xi , xj , their partition function can be written as [recall (??)]

1
qi qi
Z=
exp 4 2
log(|xi xj |/r0 ) ,

2
gas
i<j
X

(23.203)

where r0 characterizes the size of the vortices. For a single vortex-antivortex pair,
the average square distance r 2 diverges as the stiffness falls below [recall (??)]
KT = 1/TKT = 2/ 0.63662.

(23.204)

The large-stiffness state with bound vortex pairs has a coherent phase field (x),
the low-stiffness state with separated vortex pairs exhibits incoherent phase fluctuations. The same situation is found in three dimensions, only that the excitations
are vortex lines. These become infinitely long and prolific in a second-order phase
transition at a critical point c 0.33.
The result (23.201) for = 0 can now be used to estimate a critical value of
the number of field components N = Nc below which the phase fluctuations of the
complex field become so violent that the system has a phase transition. For this
we write im = M and find from (23.201) a propagator of the -field
G

i 4
+ regular terms.
N q2

(23.205)

Comparing this with the propagator for the model Lagrange density (23.202)
G =

1 i
q2

(23.206)

we identify the stiffness = N/4. The pair version of the chiral Gross-Neveu
model has therefore a vortex-antivortex pair breaking transition if N falls below the
critical value Nc = 8.
Consider now the model in 2 + dimensions where pairs form at g = g . A
comparison between the propagator (23.201) and (23.206) yields a stiffness of phase
fluctuations
!
N
g
=
.
(23.207)
M 1
4
g
The linear vanishing of the stiffness with the distance of the coupling constant g
from the critical value g is in agreement with a general scaling relation, according
to which the critical exponent of bending rigidity should be equal to (D 2).
The propagator for the real part of the pair field has, by (23.138), a correlation
length


D 1 1/2
=
.
(23.208)
12M 2

1272

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

70
M 6= 0, broken chiral symmetry
60
N 50 M = 0
Pair Formation
40
chiral
Phase Disordering Transition

30
symm

20
10
M 6= 0, chiral symmetry
1

3
g/g

Figure 23.3 The two transition lines in the N g-plane of the chiral Gross-Neveu model
in 2 + dimensions. For = 0, the vertical transition line coincides with the N -axis, and
the solid hyperbola degenerates into a horizontal line at Nc = 8. The quark masses and
chiral properties are indicated.

Inserting the g-dependence of M from (23.79), we see that


1
=
M

D1
12

1/2

g
1
g

!1/

(23.209)

so that the coherence length diverges for g g with a critical exponent = 1/.
The stiffness (23.207) implies the existence of a phase transition in the neighborhood of two and in three dimensions at roughly
g
Nc 8 1
g

!1

D 2,

g
Nc 4.19 1
g

!1

D = 3. (23.210)

As N is lowered below these critical values, the phase fluctuations of the pair field
become incoherent and the pair condensate dissolves. The different phases are
indicated in Fig. 23.3. In the chiral formulation of the same model, the intermediate
phase has chiral symmetry in spite of a nonzero spontaneously generated quark
mass M 6= 0. The reason why this is possible is that the quark mass depends only
on |0 |, thus allowing for arbitrary phase fluctuations preserving chiral symmetry.
The sceptical reader may wonder whether the solid hyperbola in Fig. 23.3 is not
simply the proper (albeit approximate) continuation of the vertical line for smaller
N. There are two simple counterarguments. One is formal: For infinitesimal
the first transition lies precisely at g = g = for all N, so that the horizontal
transition line is clearly distinguished from it. The other argument is physical. If
N is lowered at some very large g, the binding energy of the pairs increases with
1/N {in two dimensions, the binding energy is 4M sin2 [/2(N 1)]}. It is then
impossible that the phase fluctuations on the horizontal branch of the transition
line, which are low-energy excitations, unbind the strongly bound pairs. This will
only happen in the limit N where the binding energy becomes zero and the

1273

23.11 Finite-Temperature Properties

two transition curves merge into a single curve. This is the situation in the theory
of superconductivity, where Cooper pair binding and pair condensation coincide.
In the ordinary Gross-Neveu model, the analog of the phase disordering transition
is an Ising transition, in which the vacuum expectation value of jumps between
0 and 0 in a disorderly fashion. In two dimensions, this occurs at some critical
value Nc . In 2 + dimensions, this transition should again exist independently of
the transition at which the system enters into a state of nonzero 0 . It will be
interesting to see these two transitions in either model confirmed by Monte-Carlo
simulations.
There exists a four-dimensional version of this discussion in which it is shown
that, very probably, a pion condensate cannot form in the Nambu-Jona-Lasinio
model, due to directional fluctuations of the pion field.10

23.11

Finite-Temperature Properties

It is useful to study also the behavior of the Gross-Neveu model at a finite temperature. The thermal properties of this model will closely resemble those of a
superconductor. For this we confine the imaginary-time variable to the interval
(0, h
) with = 1/kB T , and take the fields to be antiperiodic under +h.
Equivalently, we may think of this model as a nonlinear -model on an infinitely
long spatial strip with antiperiodic boundary conditions, whose width along the axis is . In the limit N , we can study the effects of temperature exactly. For
simplicity, we consider the model only for a vanishing initial bare mass m0 . The
corresponding effective potential of the -field in Eq. (23.16)
i
h
1
1 2
1 Z d D pE
1
2
2
v() = [] =
tr(1)
log
p
+

E
N
N
2g0
2 (2)D

(23.211)

is generalized to finite temperatures T by exchanging the momentum integral by a


sum over Matsubara frequencies m = (2m + 1)T /h, m = 0, 1 , 2, . . . :
Z

d D pE
dD1 p 1

(2)D
(2)D1 h

(23.212)

m =

thus becoming (in natural units with kB = 1 and h


= 1)
1 2
1
1
v() =
2D/2
N
2g0
2

X
dD1 pE
2
T
log(m
+ p2E + 2 ).
(2)D1 m=

(23.213)

The gap equation is obtained by minimizing this action and becomes [compare
(23.15)]
1
= 2D/2
g0
10

X
1
dD1 pE
T
.
2
D1
2
2
(2)
m= m + pE +

(23.214)

See H. Kleinert and B. Van den Bossche, Phys. Lett. B 474, 336 (2000) (hep-ph/9907274).

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

1274

Using the well-known summation formula


T

1
=
tanh
2
2
2
2T
m= m +


(23.215)

the gap equation becomes


1
= 2D/2
g0

dD1 pE 1
tanh
.
(2)D1 2
2T

(23.216)

We shall renormalize this by adding and subtracting, on the right-hand side, the
zero-temperature limit:
Z
Z

dD1 pE
d D pE 1
1
1
D/2
D/2
tanh
= 2
+
2
1 . (23.217)
2
D1
2
D
g0
(2)
pE +
(2) 2
2T


Near two dimensions, the first integral can be written as


2

D/21

2 D/21

S
D (D/2)(1 D/2)( )

D
1

= b b
log +
+ O().
2

2
(23.218)
!

where we have used Eqs. (23.17), (23.20). The renormalized gap equation reads
then, at = 0,

1
1 1 Z dp
1
tanh
= log + +
1 .
2
gR ( )

2 0
2T


(23.219)

It is convenient to express the zero-temperature part of this equation without the


arbitrary scale parameter , using the renormalization invariant mass
"

1
M = exp

gR () 2

which solves the zero-temperature gap equation. Then we arrive at the finite equation



1
,
(23.220)
log
= S1

2T
with the function [compare with the bosonic version of this in Eq. (20.112)]
S1

2T

Z
1
dp

1  /T
tanh
e
+1
1 = 2
dp
. (23.221)

2T

0


This function accounts for all finite-temperature effects. It depends only on the
dimensionless ratio /T . We have divided this by one more factor 2 for later

convenience. The dimensionless ratio /2T will in the following be denoted by ,


i.e.,
= .
(23.222)

2T

1275

23.11 Finite-Temperature Properties

The solution (T ) M(T ) of (23.220) is now the temperature-dependent fermion


vanishes and (0) = 0 = M. As the
mass. For T = 0, the function S1 ()
temperature rises, the fermion mass M(T ) decrease, until it vanishes at a certain
critical temperature Tc . The value of Tc is found by assuming (T ) to be small, and
approximating the right-hand side of (23.221) by
2

"

dp 1
p
tanh
2 p
2T


1
.
2
p + 2

(23.223)

Integrating the first term by parts gives


1

p
log
2T



Z
p
2
dx log x cosh x .

tanh
2T 0
0


(23.224)

The integral is convergent and equal to log (4e /), where = 0.577 . . . is Eulers
number. This follows at once from the formula
Z

4 
2
, ()( 1)
(23.225)
1

2
dx x1 cosh2 ax =
(2a)
0
in the limit 1, where
(z)

1
z
k=0 k

(23.226)

is Riemanns zeta function. Using the property (0) = 1/2 and (0) =
21 log(2) (1) = . The second term
in (23.223) can be integrated directly
q
i
h
2
with the result asinh (p/) = log p/ + p /2 + 1 . Hence we find for (23.223)

p
1
log

2T


p
p
tanh
log +
2T


 

p2
2e
1
4e

,
+1
+log
= log
2

T

(23.227)

and (23.220) determines the critical temperature by the equation


log
or

= log

e
T

(23.228)

e
e
=M .
(23.229)

At this temperature, the fermion mass M(T ) vanishes.


In order to study the full temperature behavior of M(T ), the right-hand side
of the gap equation (23.220) has to be evaluated numerically. For this purpose, we
shall derive a more useful form of the gap equation. Let us go back once more to
the original form (23.214) and rewrite it for 0 as
Tc = 0

1
2D/2
g0

2
1
1
dD p
1

log
=

+
+1

(2)D p2 + 2
g0
2
2
Z

X
dm
1
1
1

q
q

= S1 () = T
2 + 2
2 + 2

2 m
0
m
m=0
!

(23.230)

1276

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

[compare (23.44)]. This can again be written as in (23.220):

= S1 (),
0

log

(23.231)

where S1 is given by the sum minus integral

=
S1 ()

m=0

1/2

dm r

1
m+

It is useful to reorganize the sum as follows:

=
S1 ()

m=0

r


1
m+

1
2

1
m=0 m +

1
2

2

+ 2T

1
2

2

.
+

m + 12

1
.
dm r
2
1/2
1
2

m+ 2 +

(23.232)

2T

(23.233)

log(2m/
The sum
The integral up to some large m = m
gives asinh(m/
)
).
1
over 1/(m + 2 ) is + log 4 + (m
+ 3/2) = + log 4 + (m
+ 1/2) + 1/(m
+ 1/2).
2
Using the large-z behavior (m
+ 1/2) log 1/2m
1/4m
, we find for the sum
the convergent sum
the limit + log 4 + log m,
and for S1 ()
=
S1 ()

m=0

r


1
m+

1
2

2

2
+

+ log(2e ).

S1 ()



1
+ log 2e

m + 21

(23.234)

cancels a similar term on the left-hand side of the gap equation


The logarithm of
(23.220), and using the connection (23.229) between 0 and the critical temperature,
and substituting M for 0 , we obtain the gap equation in a form most suitable for
a numerical evaluation:
T

log
= S1 ().
(23.235)
Tc
as a function of .
The
This can be used to calculate T /Tc and M(T ) = = 2T
resulting function M(T ) is plotted in Fig. 23.4.
It is quite easy to calculate the way in which M(T ) vanishes as T approaches
the critical temperature. We simply expand
=
S1 ()

k=1

1/2
k

2k

m=0

1


2k+1

1
2

m+

1 2 X
3 4
1
1
=

3 +

 ... .
2 m=0 m + 1
8 m=0 m + 1 5
2

(23.236)

1277

23.11 Finite-Temperature Properties


e 1.75
1.5
1.25
1
0.75
0.5
0.25

(T )

0.2

0.6

0.4

0.8

T /Tc

Figure 23.4 Solution of the temperature dependent gap equation, showing the decrease
of the fermion mass M (T ) = (T ) with increasing temperature T /Tc .

Expanding near Tc the logarithm as log (T /Tc ) T /Tc 1, we find in a first approximation
s
s
8
T
M(T ) = Tc
(23.237)
1 .
7(3)
Tc
In the opposite limit of low temperatures, the series (23.236) converges very slowly.
It is, however, easy to find out how behaves near T = 0 by expanding in (23.221)
tanh
Using the integral

1=2
()m em/T
.
2T
m=1

(23.238)

2
2
em p + /T
dp 2
= 2K0 (m/T

),
p + 2

(23.239)

where K0 (z) is the associated Bessel function, we find the alternative expression
log

0


X

=2
.
= S1 ()
()m K0 2 m

0
m=1

(23.240)

For small T , K0 (m/T

) has the asymptotic behavior


K0

T


e/T ,
2m/T

(23.241)

so that we can expand


= 0 1

2 /T
e
0 /T

+ O e2/T

(23.242)

and see that approaches its T = 0 -value 0 exponentially fast from below (see
Fig. 23.4).
Note that the properties of the gap equation (23.220) are quite similar to those
of the corresponding equation (20.110) in the nonlinear -model.

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

1278

It is instructive to go through the same discussion once more in D dimensions.


For this it is convenient to rewrite the gap equation (23.230) in accordance with the
general procedure of dimensional regularization in Section 11.5 as
Z
Z
1
dD p (p2 +2 ) D2 1D/2 (1D)/2
D/2

d
2
e
=
2

(3/2 D/2)S1 ().


g0
(2)D

0
(23.243)

The left-hand side corresponds to the zero-temperature gap equation and is integrated directly to


D

D
1

b =
(23.244)
b
,
g0
2
g()
2
while
3 D 1
2D D1 (D1)/2

S1 () =
2

(23.245)
2
2
!
Z
Z
Z

i
h

X
dm
dD1 p
2
2
2
.
exp

+
p
+

d
m
(2)D1
2
0
m=
 



The momentum integrations can now be done, with the result


1
1
2D/2 D D/2
g0
2

D
d 2
3 D
(23.246)
S1 (),
e
= D2 21D/2 1/2 2

D/2

2
2

where
= 22D
S1 ()

3 D

2
2

 

1 Z

D/21

m=0

dm (m
2 +2 )
e
.
2
(23.247)

By performing the -integral on the right-hand side, we find


= 2
S1 ()

2D

m=0

2D
=

m=0

dm
2
(m
+ 2 )(D3)/2
2

1/2

! "

dm

1
m+
2

2

2
+

#(D3)/2

(23.248)

Its its contributions


Let us expand the sum over m formally in a power series of .
is
to S1 ()


S1 ()

sum part

2D
=

k=0

(D 3)/2
k

where

(z, a)

2k (2k + 3 D, 1/2) ,

1
z
k=0 (k + n)

(23.249)

(23.250)

1279

23.11 Finite-Temperature Properties

is Riemanns zeta function with (z, 1) = (z) of Eq. (23.226). The integral over m
in (23.248) subtracts from this the T = 0 -limit of (23.249):
2

D/21 (D1)/2

3 D

2
2

 

1

 



3 D 1
D

.(23.251)
b =

1
2
2
2
2
2

say D = D
+ , the kth term with k = D/2
1
For D near an even dimension D,
has an 1/-singularity coming from the zeta function


D, 1/2
1+D

1
[1 + (1/2)] .

(23.252)

This is canceled by a singularity of opposite sign in b . We now expand up to first


order in :
2D

(D 3)/2
k

D2
(1, 1/2) =

k=D/2

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

(D/2)(1/2 + /2)


1
1


+ (1/2)



1
+
2
log

1
(3/2 D/2) (D/2)(1/2)
2
cos D/2



i

h
(1)D/2

1 + 2 log + (1/2) + 3/2 D/2

(3/2 D/2) (D/2)

2


Similarly we expand the T = 0 -term (23.251):


 




D
3 D 1

2
2
2
2


D/2

i
()

=
1 (D/2) 3/2 D/2
. (23.253)

(D/2)(3/2
D/2)

2
Subtracting the two terms from each other yields
h


i
()D/2

(1/2) D/2


2
log

(D/2)(3/2 D/2)
2
h

D/2


i

()
D/2

2 log(2e )
. (23.254)
=
(D/2)(3/2 D/2) 2

= 2 by
On the right-hand side we may replace the Digamma function (D/2)
for D

> 2, (D/2)

+ . . . + 2/(D
2). Going now
(D/2)
= , and for D
= + 2/D

to the limit 0, we obtain for even D = D:


h
i

()D/2
(D/2)
2 log(2e )
(D/2)(3/2 D/2) 2
!

X
(D 3)/2 2k+2D
+

(2k + 3 D, 1/2) .
k
k=0,k6=D/21

=
S1 ()

(23.255)

1280

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

2 out of the sum, we split


By taking the negative powers of
h
D/2


i
()

(D/2)
=
2 log 2e
S1 ()
(D/2)(9/2 D/2) 2
D/22

(D 3)/2
k

X
k

where


S1 ()

k=D/2

2k+2D (2k + 3 D, 1/2) + S1 (),

(23.256)
!

(D 3)/2
k

2k+2D (2k + 3 D, 1/2) .

(23.257)

can also be calculated from the sum part in (23.248) by performing


The sum S1 ()
D/2 subtractions:
2D

=
S1 ()

"

X

m=0

1
m+
2

2

1
m+
2


D3

#(D3)/2

1
D 3 2
m+

2
2


D5

+ ... .

(23.258)

This sum is convergent and starts out with S1 (0) = 0. It is the generalization of
in Eq. (23.234) to arbitrary even D.
S1 ()
in Eqs. (23.221) and (23.240)
Let us also generalize the other expression for S1 ()
to arbitrary D, such that it gives the finite-temperature correction to the gap equation
# "
#
"
Z

X
1
1
dD1 pE
1
D/2
iT tr
...
(0) = 2
T
2 + p2 + 2
i/

(2)D1 m= m
E
T 0
=

D2 1D/2 (1D)/2
3 D

S1 (),
2

2
2


(23.259)

where the notation


T X(T ) X(T ) X(0)

(23.260)

denotes the finite-temperature correction to any quantity X(T ). By comparison


with (23.245), we can identify

dD1 p 1
3 D 1
tanh

1
2
2
(2)D1
2T
 
 



Z
3 D

D 1 1 2D
D2 1
=

tanh

1 .

dp p
2
2
2
2

2T
0
(23.261)


 

= 2D2 (D+1)/2 2D
S1 ()

Writing tanh (/2T ) 1 as 2(1 + e/T )1 , and expanding this in powers of e/T ,
we have the alternative expression
= 2 3 D D 1
S1 ()
2
2
2
2
 

1

2D

()

m=1

dp pD2 1 em/T
.

(23.262)

1281

23.11 Finite-Temperature Properties

Using the integral representation for the modified Bessel function


 

(1/2)
( + 1/2)

z
K (z) =
2

dss2 (s2 + 1)1/2 ez

s2 +1

(23.263)

we arrive at the expansion


 1

= 3D 1
S1 ()
2
2
2


 

2D/2

1D/2

()m 2 m

m=1

KD/21 (2 m
),
(23.264)

which for D = 2 reduces properly to (23.240).


Let us also calculate the finite-temperature correction to the free energy for all
dimensions D. By integrating the gap equation (23.243) with (23.244) and (23.245)
in 2 /2, we obtain
1
2
v() =
+ iTr log(i/
) + iT log(i/
),
N
2g

(23.265)

with
iT Tr log(i/
) = 2

D/2

= 2

D/2



X
dm
dD1 pE 1
2
2
T

log

m
(2)D1 2
2
m=
!



dD1 pE
/T
,
log
1
+
e
(2)D1

(23.266)

so that
We now introduce a function S0 (),
iT Tr log(i/
)

D D/2 1/2D/2
1 D
1
(23.267)
S0 (),
T v() =
2

2
2


the following expressions


and find for S0 ()
= 4D 1 D D 1
S0 ()
2
2
2
2
 

1

dp pD2 log 1 + e/T ,


(23.268)

or, alternatively,
=
D
S ()
0

m=0

1/2

! "

dm

1
m+
2

2

2
+

#(D1)/2

(23.269)

is related to S1 ()
by an integration
The new function e S0 ()
= D 1
D
S0 ()
2

,
d2T D2
S1 ()
T

(23.270)

1282

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

we can integrate (23.255)-(23.258), and see that the latter expression is separated
into a convergent sum
=
D
S0 ()

"

X

m=0

1
m+
2

2

1
m+
2


+ 2T

#(D1)/2

(D1)/2

1
D 1 2

m+
2
2


D3

+ . . . ,(23.271)

with D/2 + 1 subtractions, an additional logarithmic term


h
i

()D/2
(D/2 + 1/2) ,
2 log(2 )
(D/2)(1/2 D/2) 2

(23.272)

from the integral over the second term in


and a sum over negative powers of
(23.256).
Alternatively we find from an integral (23.270) over the expansion (23.264):
 

= 2 2D/2
S0 ()

 1 X

1 D
1

2 2
2


D/2

()m 2 m

m=1

KD/2 (2 m
).(23.273)

According to Eq. (23.270), this has to satisfy


D S0
1 D D2
d
=

S1 ,
2
2
d

(23.274)

and a comparison with (23.264) shows that it does, since


i

z D/2 KD/2 (z) = z D/2 KD/21 (z).

(23.275)

In the high-temperature limit, we use the small-z behavior


D
1
KD/2 (z)
2
2


  D/2

z
2

(23.276)

to find
1D
D
S0 ()

 X

(D/2)
1

(1/2 D/2)
2

(1)m (m)
D .

(23.277)

m=1

The sum can be expressed in terms of Riemanns zeta function (23.226) as

(1)m (m)
D = 1 21D (D).

m=1

(23.278)

For D = 2, we thus obtain


2

T (2).
S0 ()
2

(23.279)

1283

23.11 Finite-Temperature Properties

Recalling (23.267), this yields the low-temperature behavior of the effective potential
in two dimensions
1 2
T 2
1
T v()
T (2) =
.
N

(23.280)

This is the well-known free energy of a hot (or massless) Fermi Gas in two dimensions. It could have been obtained directly by dimensional regularization from the
= 0 -expression
1
T v() =
N

= T

dp
2

m=

q
X

dm
2
log(m
+ p2 )
2

2 = 4T 2 (1, 1/2) = T 2 ,
m
6

m=

(23.281)

the last equation following from (1, 1/2) = (21 1)(1) = 1/24.
In D = 4 dimensions, the result is

i1 X
1 4 h

(3/2)

()m m
4
3
m=1

h
i1 7

(4),
= 4 24 T 4 (3/2)
8

=
S0 ()

(23.282)

and for the potential


7 2
1
T v() 4T 4 D .
N
8g

(23.283)

The latter which follows directly from


1
T v() = 2
N
=2

d3 p
(2)3

d Z d p
T

(2)3

m=

3
X

= (4 3/2 )1 ( 23 ) T



dm
2
log m
+ p2
2
2

e (m +p

2)

m=

2 3/2
(m
) = T 4 4 3/2 ( 23 ) (3, 12 ) . (23.284)

m=

It is the fermion equivalent to the Stefan-Boltzmann law for the free energy density
f of hot (or massless) fermions. Compare the discussion for the bosonic model in
the paragraph below Eq. (20.169), where we show that for a pure power behavior of
T n of f , the entropy density s = f /T is related to f by a factor 4/T , whereas
the internal energy density u E/V = f + T S and for the specific heat at constant
volume cV = u/T |V , carry the relative factors 1 n and (1 n)n/T , respectively.
Thus, Eq. (23.284) yields for these three thermodynamic quantities
1
T
T
2 2 7 4
f = u = s = cV = 4
T .
3
4
12
3 60 8

(23.285)

1284

23 Exactly Solvable O(N )-Symmetric Four-Fermion Theory in 2 + Dimensions

In proper physical units


1
T
T
2 7 4
f = u = s = cV = 4
T ,
3
4
12
3c 8

(23.286)

where is the Stefan-Boltzmann constant (20.172). The prefactor 4 accounts for


the two polarization degree of freedom of both particles and antiparticles, while the
factor 7/8 is typical for fermions.
Recall that the black-body radiation of photons has a free energy
T
T
2 4
2 4
1
T = 2
T ,
f = u = s = cV = 2
3
4
12
90
3c

(23.287)

where the prefactor 2 on the right-hand side accounts now for the two polarization
degrees of photons. In contrast to the fermion case, particles and antiparticles
coincide.

Notes and References


The earliest studies of the model were published by
A.A. Anselm, JETP (Sov. Phys.) 9, 608 (1959).
and by V.G. Vaks and A.I. Larkin, JETP (Sov. Phys.) 13, 979 (1961).
In four spacetime dimensions, the model has been applied to particle physics by
Y. Nambu and G. Jona Lasinio, Phys. Rev. 122, 345 (1961); 124, 246 (1961);
See also:
J.D. Bjorken, Ann. Phys. 24, 174 (1963);
I. Bialynicke-Birula, Phys. Rev. 130, 465 (1963);
G.S. Guralnik, Phys. Rev. B136, 1404, 1417 (1963);
H. Umezawa, Nuovo Cimento XL, 450 (1965);
Y. Freundlich and D. Lurie, Phys. Rev. D 8, 2386 (1974);
H. Pagels, Phys. Rev. D 7, 3689 (1973);
K. Lane, Phys. Rev. D10, 2605 (1974);
R. Jackiw and K. Johnson, Phys. Rev. D 8, 2386 (1974);
P. Langacker and H. Pagels, Phys. Rev. D 9, 3413 (1974); D 10, 2904 (1974);
P. Langacker, Phys. Rev. Letters 34, 1592 (1975);
H. Pagels, Phys. Rev. D 14, 2747 (1976);
H. Matsumoto, H. Umezawa, N. J. Papastamatiou, Nucl. Phys. B 68, 236 (1974),
B 82, 45 (1974);
L. Leplae, H. Umezawa, and F. Mancini, Phys. Rev. C 10 (1974).
The model in D = 2 + dimensions is named after
D. Gross and A. Neveu, Phys. Rev. D 10 3235 (1974),
A comprehensive explanation of the low-energy properties of hadron physics was
given on the basis of this model by
H. Kleinert, On the Hadronization of Quark Theories, Lectures presented at the
Erice Summer Institute 1976, in Understanding the Fundamental Constituents of
Matter , Plenum Press, New York, 1978, A. Zichichi ed., pp. 289-390.

Notes and References

1285

The Dirac algebra in arbitrary dimensions D is discussed in


B. Rosenstein, B.J. Warr, and S.H. Park, Phys. Rep. 205 (1991).
For calculations beyond the leading 1/N approximation see
K. Akama, Nambu-Jona-Lasinio model at the next-to-leading order in 1/N, Nucl.
Phys. A 629, 37C (1998) (nucl-th/9709001),
M. Huang, P.F. Zhuang, W.Q. Chao, Massive meson fluctuation in NJL model ,
Commun. Theor. Phys. 34, 91 (2000); The mesonic fluctuations and corrections in
the chiral symmetry breaking vacuum, Phys. Lett. B 514, 63 (2001).

In science one tries to tell people something


that no one ever knew before. But in poetry, its the exact opposite
Paul Dirac (19021984)

24
Internal Symmetries of Strong Interactions
If we open up a standard table of the presently known elementary particles, we see a
confusing variety of many different objects. Before trying to develop a more detailed
theory of their interactions it is useful to find certain organizing principles which
correlate their quantum numbers and mass spectra.

24.1

Classification of Elementary Particles

The most fundamental distinction concerns the statistics of the particles which can
be of the Fermi-Dirac or Bose-Einstein type in which case the particles are called
fermions or bosons, respectively. Historically, the fermions were subdivided into
light and heavy fermions. The former are called leptons and comprise particles such
as electrons, muons, and neutrinos. Recently, also heavier objects have been found
which should be considered as leptons, due to the similarity of their interactions
with that of the other leptons. In fact, the characteristic property of leptons is now
that they are fermions with no strong interactions. The distinction according to
mass is no longer significant.
The most important of the traditional heavy fermions are the particles which make
up the atomic nucleus, the so-called nucleons which can be protons and neutrons.
These particles have strong interactions. They can be excited in nuclear collisions producing short-lived resonances which are also classified as elementary heavy
fermions. The set of all heavy fermions with strong interactions are called baryons.
All baryons can be produced in collision or decay processes also as in the form of
antiparticles. These look very similar to the corresponding particles, they have exactly the same mass and spin, but other quantum numbers such as the charge are
reversed.
The particles with Bose-Einstein statistics are separated into those with dont
and those which do possess strong interactions. The former play an important role
in the theoretical description of the gravitational, the electromagnetic, and the weak
interactions. They are called graviton, photon, and intermediate vector bosons W
and Z. The bosons with strong interactions are called mesons. The set of all
particles with strong interactions, baryons and mesons, are called hadrons.
1286

24.1 Classification of Elementary Particles

1287

Only few of all these particles are completely stable. The most prominent ones
make up the stable matter of the universe. These are the protons, electrons, photons,
neutrinos, and gravitons. The proton is the only stable baryon. The other important
nuclear constituent, the neutron, lives on the average only 898 16 seconds. After
this it decays into a proton, an electron, and an antineutrino.
Among the leptons, only electrons and neutrinos are stable particles. The muon,
a particle which is very similar to the electron but much heavier, decays in about
106 seconds into an electron, a neutrino, and an antineutrino.
The proton is not only the only stable heavy fermion, it is the only stable strongly
interacting particle. The mesons are all unstable.
The decays proceeding over a long time as in the case of the neutron are called
weak. Even lifetimes of 1013 seconds are still considered as long in elementary
particle physics and the decay is called weak. A particle with such a lifetimes down
to 1011 sec leaves an observable trace in a bubble chamber.
Most of the hadrons which decay via strong interaction processes live only for a
much shorter time than that, so short that they do travel any visible distance in a
bubble chamber. They are observable only as resonances in scattering cross sections.
The most prominent example is the first resonance in pion nucleon scattering shown
in Fig. 24.2. The resonance is called (1232). It lies at a pion beam momentum

Figure 24.1 The total and elastic + -proton cross section showing clearly the first
nucleon resonance called (1232).

1288

24 Internal Symmetries of Strong Interactions

Figure 24.2 The total and elastic -proton cross section showing again clearly the
first nucleon resonance called (1232), but with only a third of the height. This is in
agreement with the isospin 3/2 assignment and the amplitude relations (24.39) if one
invokes the optical theorem, according to which the total cross section is proportional to
the imaginary part of the forward scattering amplitude.

p = 0.34 GeV/c.

(24.1)

In the center of mass coordinate frame, the energy is equal to the mass of the
resonance M,
ECM = M =
q

(E + mp )2 p2 =

m2 + m2p + 2mp E

(24.2)

where E = p2 + m2 is beam energy and m , mp are the masses of pion and


proton. From p = 0.34 GeV/c one finds M = 1232 MeV/c2 which is the number
given in the parenthesis of (1232). The large width in the total cross section shows
the rather short lifetime. The peak is of the Lorentz shape and can be parametrized
as

tot
.
(24.3)
(ECM M)2 + (/2)2

24.2 Isospin in Nuclear Physics

1289

At the energies ECM = M /2 the cross section has fallen to half its peak value.
The parameter is therefore identified with the width of the resonance. From the
experimental curve we extract the width
115 MeV.

(24.4)

In the quantum mechanical theory of resonance scattering it is known that such a


cross section indicates an exponentially decaying wave function of the form
ei(M i/2)t/h .

(24.5)

Indeed, such a wave function leads to a scattering amplitude which contains a pole
in the energy plane below the real axis, proportional to
1
.
ECM M + i/2

(24.6)

Its absolute square gives the above Lorentzian cross section. The lifetime of such a
wave function is
= 2h/.
(24.7)
This is why the inverse width is sometimes called the half-life. In physical units, the
right-hand side is to be multiplied with h
if is measured in MeV and in seconds.
The (1231) resonance has therefore a lifetime
= 2h/ = 0.6583 1021
5.72 1024 sec.

MeV
sec

(24.8)

It decays mostly into a pion and a nucleon. Such a short lived decay is called strong
decay. A small fraction, however, about 0.6% of the decays goes into a nucleon and
a photon. This is called the radiative decay channel.
Since the (1232) resonance can decay into a nucleon and a photon, we should
expect that it can also be produced in a collision of the final states, i.e., when a
photon hits a nucleon. Indeed, the cross sections in Fig. 24.3 show clearly this
resonance.
In the past twenty years, various simple organizing principles have been found
according to which the variety of particles and resonances can be classified and remembered and their interactions be related. These principles are based on symmetry
groups associated with certain characteristic invariances of the different interactions.
They have helped in understanding the quantum numbers, the energy spectra, the
various decay channels, and the different scattering cross sections.

24.2

Isospin in Nuclear Physics

If one compares binding energies of so-called mirror nuclei (which are obtained
from each other by the interchange of protons and neutrons) with each other one

1290

24 Internal Symmetries of Strong Interactions

Figure 24.3 The photon-proton and photon-deuteron total cross sections showing clearly
the first nucleon resonance (1232) as well as a second resonance called N (1520).

finds that the difference may be attributed entirely to the different electromagnetic
interactions of protons and neutrons. For instance, 3 H= nnp and 3 He= ppn have
binding energies BH3 = 8.452 MeV and B3 He = 7.728 MeV, respectively, with the
difference being accounted for by the proton-neutron mass difference mn mp =
1.293323.000016 MeV and by the different Coulomb field energy around the nuclei.
Indeed, if the Coulomb energy is approximated by that of a uniformly charged sphere
of charge Z and radius R, the field energy is
ECoul

6 e2
6 hc
= (1/2)(Z 1)
= (1/2)Z(Z 1)
.
5 4R
5 R

(24.9)

With = e2 /4hc 1/137 being the fine-structure constant. Using for the nuclear
radius for the atomic number A the estimate R 1.45 A1/3 1013 cm, we can write
6
ECoul Z(Z 1) A1/3 MeV.
5

(24.10)

For A = 3, Z = 2 we estimate ECoul 0.57 MeV and find for the difference in
binding energy
BH3 BHe3 1.29 MeV 0.57 MeV 0.72 MeV,
in rough agreement with the observed 0.76 MeV.

(24.11)

24.2 Isospin in Nuclear Physics

1291

Figure 24.4 Mirror nuclei 5 B11 and 6 C11 with their excited states (the numbers are the
excitation energies in MeV). The ground state of 6 C11 decays into that of 5 B11 by positive
-decay (i.e., emission of a positron and a neutrino) as indicated by the arrow. An estimate
of the Coulomb energy via formula (24.9) explains roughly the energy difference of 1.98
between the two level schemes.

Not only the binding energies, also the excitation spectra of mirror nuclei differ
only by common electromagnetic shift, as seen in the example in Fig. 24.4. This
property of mirror nuclei has led nuclear theorists to postulate that if it were possible
to switch off the electromagnetic interactions the potentials Vpp and Vpn would turn
out to be exactly equal. This symmetry is called charge symmetry.
The charge symmetry possesses a natural extension. When analyzing the data
of nucleon-nucleon scattering at low energy, it was found that also also Vpp Vpn .
This led Heisenberg to postulate that within purely nuclear interactions, protons and
neutrons should be two indistinguishable states of the same particle, the nucleon N.
He described the two states in analogy with the two spin states of a spin 1/2 object
by a two-component object
N=

p
n

(24.12)

and called this object an isospinor state. Heisenberg considered the states as a
representation of a rotation group in a fictitious space, the space of isotopic spin.
The generators are denoted by Ii and obey the commutation rules of the rotation
group
[Ii , Ij ] = iabc Ik .

(24.13)

The two component isospinors are rotated by the 2 2 representation of Ii , given


by Ii = i /2. Proton and neutron are eigenstates of the third component I3 = 3 /2
with eigenvalue 1/2 and 1/2.

1292

24 Internal Symmetries of Strong Interactions

The exchange of protons and neutrons corresponds to a rotation around the 2 axes
of isospin by 1800
eiI2

p
n

i2 /2

=e

p
n

0 1
1 0

p
n

n
p

(24.14)

The postulate of indistinguishability of protons and neutrons can then be formulated mathematically as the commutativity of the generators of isospin Ii with the
Hamiltonian of strong interactions
[Ii , Hstrong ] = 0.

(24.15)

This is called isotopic spin symmetry or isospin invariance of strong interactions.


If |Ai is a nuclear n body wave function, then the isospin acts additively on each
nucleon just as ordinary angular momentum (compare Appendix 3B)
Ii |Ai = (Ii 1 . . . 1 + . . . + 1 1 1 . . . Ii )|Ai.

(24.16)

Nuclei which possess a large difference in their number of protons, np , and neutrons,
nn , have a third component of isospin
1
I3 |Ai = (np nn )|Ai.
2

(24.17)

Their total isospin I must be at least equal to I3 . The ground state has I = I3 , the
excited states have a higher I. If the Coulomb energies are taken into account, they
can be seen to be members of isospin multiplets formed together with neighboring
nuclei of equal total number of nucleons and different np nn (see Fig. 24.5). With
the help of isospin rotations, the charge symmetry operation may be represented as
a rotation around the second axis by an angle : exp(iI2 ).

Figure 24.5 Singlets and triplets of isospin in the nuclei 6 C14 , 7 N14 , 8 O14 (with the excitation energies in MeV). The dotted lines connect levels are roughly degenerate after
subtracting the Coulomb energy according to the estimate (24.9).

The postulate of isopsin symmetry leads immediately to interesting selection


rules for decay process via nuclear interactions, such as decay or emission of

24.2 Isospin in Nuclear Physics

1293

deuterons. Consider a nucleus |Ai that it has an equal number of protons and
neutrons, i.e. a nucleus that is charge symmetric to itself, a so-called self-conjugate
nucleus. Since it consist of an even number of isospin 1/2 nucleons, its total isopin is
necessarily integer, and the third componment of the isopsin is zero. The operation
exp(iI2 ) must therefore produce the same state up to a phase :
exp(iI2 )|Ai = |Ai.

(24.18)

Aplying this operation twice must lead back to the original state. Hence the phase
can only be 1. It is called the intrinsic charge parity of the nucleus. Every
charge-symmetric nucleus has a definite charge parity.
The charge parity is calculable uniquely as a function of the total isospin.
We use the property of the eigenfunctions of angular momentum Ylm (, ) with
m = 0 to pick up a phase ()l , when rotated around the y-axis by an angle [see
the matrix elements of the rotation matrix (4.337)]. This is also obvious from the
representation [(I + I3)!(I I3 )!]1/2 ap(I+I3 ) an(II3 ) |0i of an isospin I, I3 state in terms
of creation operators ap , an of isospin up and down, as constructed in Eq. (4.322).
Under an isospin rotation around the 2-axis by , then ap and an go over into an and
ap , respectively. A charge symmetric state with I3 = 0 has therefore the charge
parity
= ()I .
(24.19)
The charge parity gives rise to interesting section rules for decay processes via nuclear
interactions. If the Hamiltonian commutes with all generators of isospin, the initial
and final states must have equal charge parities. Typical forbidden processes are
10
5B

6
3 Li
4 Be

(24.20)

where 5 B10 can be the 4.5 MeV or the 6.5 MeV resonance of 5 B10 . Now, the excited
nucleus 5 B10 of excitation energy 4.5 MeV has is a member an I = 1 isospin triplet
formed together with ground states of the isotopes 4 Be10 and 6 C10 . It therefore has
a charge parity = 1. The nuclei 3 Li6 and 4 Be8 , on the other hand, have isospin 0
and therefore a positive charge parity = +1. They are hence forbidden processes,
which have indeed never been observed in the laboratory.
Let us now study the consequences of the larger isotopic spin symmetry in nuclear
reaction. A deuteron in the ground state has the orbit in a symmetric swave and
the spin states are given by the three symmetric triplet states
| i,

1
| + i,
2

| i.

The antisymmetry of the wave function requires the isotopic wave function to be
antisymmetric, i.e. it must be
1
|pn npi.
2

1294

24 Internal Symmetries of Strong Interactions

This is an I = 0 state of isospin.


If a deuteron collides with 4 Be9 , which has I = 12 , I3 = 12 , the isospin of the
two-particle system is necessarily I = 21 , I3 = 12 . We can now compare two possible
final states
10
5 B1.7 MeV + n
and
10
4 Be

+ p.

If we denote the isospins of the two particles by I (1) , I (2) , the isospin states are in
the first case
(1)
(2)
(24.21)
|I (1) , I3 ; I (2) , I3 i = |1, 0; 21 , 12 i,

and in the second

(1)

(2)

|I (1) , I3 ; I (2) , I3 i = |1, 1; 1, 12 i.

(24.22)

The strength with which these two final states are coupled to the initial state | 21 , 12 i
are proportional to the Clebsch-Gordan coefficients
1
h 21 , 12 |1, 0; 21 , 12 i = ,
3
s
2
,
h 21 , 21 |1, 1; 1, 12 i =
3

(24.23)
(24.24)

respectively (recall Table 4.2). Hence the first reaction cross section should be half
as big as the second, a fact born out by experiments.

24.3

Isospin in Pion Physics

In 1935 Yukawa1 postulated the existence of a bose particle which should mediate
the strong interactions between nucleons in a similar way as the photons do between
atoms. From the size of nuclei being of the order of 1013 cm Yukawa concluded that
instead of a Coulomb potential
1
v(x) =
(24.25)
4r
which solves the field equation
x 2 v(x) = (3) (x).

(24.26)

The potential between nucleons should have a finite range and read
v Y (x) =

er
,
4r

(24.27)

thus solving the differential equation


(x 2 + 2 )v Y (x) = (3) (x).
1

H. Yukawa, Proc. Phys. Math. Soc., Japan 17 , 48 (1935).

(24.28)

24.3 Isospin in Pion Physics

1295

He identified the inverse range with the mass of the particle. This particle was
discovered experimentally around 1946 and is now called the pion. It exists in three
charge states + , 0 , . The neutral pion has a mass
m0 = 134.9642 .0038 MeV,

(24.29)

from which the masses of the charged pions differ by


m m0 = 4.6043 .00037 MeV.

(24.30)

The -mesons exist only for a small time. The neutral 0 meson decays mainly into
two photons with a lifetime
0 (0.87 .04) 1016 sec.

(24.31)

This relatively fast decay is caused purely by electromagnetic interactions. The


charged states decay mainly into , with a lifetime of
(2.6030 0.0023) 108 sec.

(24.32)

This decay is so slow that it proceeds via weak interactions.


Since nuclear forces are supposed to arise mainly from pion interactions, the
isosymmetry of nuclear forces implies that the pion-nucleon coupling should be symmetric under isospin rotations. The pion itself is obviously an isospin 1 object. If it
interacts with a nucleon of isospin 1/2, the total isospin can be either 32 or 21 . The
Clebsch-Gordan coefficients for the different channels of total isospin |I, I3 i are the
following (see Table 4.2)
| + pi = |1, 1; 12 , 21 i
| + ni = |1, 1; 21 , 12 i
| 0 pi

| 0 ni

= |1, 0; 21 , 21 i

= |1, 0; 21 , 12 i

| pi = |1, 1;
| ni = |1, 1;

1 1
, i
2 2
1
, 12 i
2

= | 32 , 23 i,
q
= 13 | 32 , 21 i + 23 | 21 , 21 i,
=

=
=
=

2 3 1
| 2 , 2 i 13 | 12 , 21 i,
2 3
| , 12 i + 13 | 12 , 21 i,
3 2
q
2 1
1 | 3 , 1 i
| , 12 i,
2
3 2
3 2
| 23 , 32 i.

q3

From these we find the amplitudes for scattering processes to be all given in
terms of two amplitudes, one for total isospin I = 23 and one for total isospin I = 21 .
The amplitudes are given by the so-called scattering matrix whose matrix elements
are obtained by evaluating the scattering operator S between incoming and outgoing
states. The detailed theory of the scattering matrix will be given in Chapters 8 and
9. There we shall see that if all generators Ii of isospin rotations commute with
the Hamiltonian operator of strong interactions Hstrong , they also commute with the
scattering operator S. Then the amplitudes do not depend on the I3 components of
the total isospin.
[Ii , S] = 0.
(24.33)

1296

24 Internal Symmetries of Strong Interactions

We therefore can use I+ inside the matrix elements hI, I3 |S|I, I3i to change I3 to
any allowed value |I3 | I. From
hI, I3 |I+ S|I, I3 1i = hI, I3|SI+ |I, I3 1i

(24.34)

it follows that
q

hI, I3 1|S|I, I3 1i (I + I3 )(I I3 + 1)


q

= hI, I3 |S|I, I3i (I + I3 )(I I3 + 1)

(24.35)

[see Eqs. (4.370), (4.374)]. Thus the diagonal matrix elements hI, I3 |S|I, I3 + 1i
are independent of I3 . Similarly one sees that the non-diagonal elements vanish.
We therefore define the reduced scattering amplitude S I of a given isospin I by the
equation
hI, I3 |S|I, I3i I3 I3 S I .
(24.36)
This statement is a special case of the Wigner-Eckart theorem (4.871) for the
matrix elements of an arbitrary spherical tensor operator TI,I3 of isospin I. Such a
tensor operator is defined by the transformation law [compare (4.867)]

[Ii , TI,I3 ] = TI,I3 D I (Ii )I3 I3 ,

(24.37)

where D I (Ii )I3 I3 are the representation matrices of the generators Ii of the isospin
rotation group. According to the Wigner-Eckart theorem, the matrix elements of
a tensor operator between the representation states |I, I3 i are related by ClebschGordan coefficients:
hI , I3 |TI ,I3 |I, I3i = hI , I3 |I , I3 ; I, I3 ihI ||T ||Ii,

(24.38)

which account for the dependence on the quantum numbers I3 , I3 , I3. The matrix
elements depend only on a few reduced matrix elements hI ||T ||Ii which are nonzero
if the isospins I and I can be coupled to I . Within this definition, the scattering
operator is trivial tensor operator of isospin zero where (24.38) reduces to (24.36).
Note that (24.36) is also a manifestation of Im Schurs Lemma, according to
which a matrix commuting with an irreducible set of matrices must be proportional
to the unit matrix.
As an application of (24.36), the scattering amplitudes for various pion-nucleon
scattering processes are given by the following combinations of only two amplitudes,
S 3/2 and S 1/2 :
S+ p+ p = S 3/2 ,
S+ n+ n = (1/3)S 3/2 + (2/3)S 1/2 ,
S0 p0 p = (2/3)S 3/2 + (1/3)S 1/2 ,

S+ n0 p = ( 2/3)S 3/2 ( 2/3)S 1/2 ,


S0 n0n = (2/3)S 3/2 + (1/3)S 1/2 ,

24.4 SU(3)-Symmetry

1297
S+ p p = (1/3)S 3/2 + (2/3)S 1/2 ,

S0 n p = ( 2/3)S 3/2 ( 2/3)S 1/2 ,

S n n = S 3/2 .

(24.39)
(24.40)

The symbols in the charge exchange reactions exhibit time-reversal invariance.


By taking the absolute squares of the amplitudes, we see that the cross sections
must satisfy
pn0 + 20 p0 p = + p+ p + p p .

(24.41)

This is borne out by experiment.

24.4

SU(3)-Symmetry

In 1944, the set of fundamental particles was enriched by a particle which was a
heavy neutral meson of mass 500 MeV with a lifetime of 1010 sec. It was
called a strange particle 2 and is now called K 0 . It is now known to decay mainly
into + , and has
mK 0 497.72 0.07 MeV ,
K 0 (0.8323 0.0022) 1010 sec.

(24.42)

In 1951, a charged partner of this particle was discovered which decays mostly into
a + and a neutrino of the meson type, namely the K + meson. It has a mass and
a lifetime
mK + 493.667 0.014 MeV ,
K + (1.2371 0.0026) 108 sec.

(24.43)

One later found also a K decaying into negatively meson and a neutrino with
the same lifetime. In 1955 one discovered, moreover, that there are actually two K 0
mesons, now called K 0 and K0 . These mesons could soon be produced abundantly in
particle accelerators and it became possible to study their interactions with nucleons.
In the course of this it was found that not only the mesons but also the nucleons
possess strange partners. Moreover, if one assigned to the strange mesons a quantum
number S = 1, called strangeness, and to the strange partner f the baryons in the
associate production S = 1, then strangeness is a conserved quantum number in
a nuclear scattering process. If an ordinary non-strange meson such as a pion hits
a nucleon, then out comes either a pion and a nucleon both of strangeness zero, or
a strange meson with S = 1 together with a strange partner of the nucleon with
S = 1. For example
p p
(24.44)
2

M. Gell-Mann, Phys. Rev.


10 , 581 (1953).

92 , 833 (1952); T. Nakano, K. Nishijima, Progr. Theor. Phys.

1298

24 Internal Symmetries of Strong Interactions

or
p

K 0 0 ,
K + ,

(24.45)

There are also associate production processes where two Ks are produced in addition to a nucleon:
p K K 0 p,
(24.46)
+ p K + K0 p.
(24.47)

These allow us to deduce the strangeness for K + , K have the strangeness S = 1, 1,


respectively.
By scattering further a strange meson K on a proton,
K p K + ,

(24.48)

one is able to produce a new particle of strangeness S = 2. This particle is


called cascade since it decays in a cascade like process with the first step
0

(24.49)

and the second steps


0 p ,
.

(24.50)
(24.51)

Note that in contrast to the production process, the decay of the strange particles
violates strangeness. For instance:
n
S : 1
0
0.

(24.52)

This violation, however, can be blamed on another interaction, the weak interaction,
since the decay proceeds quite slowly, with a lifetime of roughly 1010 sec.
While non-strange particles have a charge
Q = N/2 + I3 ,

(24.53)

where N is the total nucleon-number, the charge of strange particles is given by


Q = Y /2 + I3 ,

(24.54)

with Y being defined as a convenient combination of N, and S, called hypercharge


Y N + S.

(24.55)

Gell-Mann plotted the known ordinary and strange pseudoscalar meson states in

24.4 SU(3)-Symmetry

1299

Figure 24.6 Pseudoscalar meson octet states associated with the pions. The same picture
holds for the vector meson octet states with the replacement (24.58).

Figure 24.7 Baryon octet states associated with the nucleons.

the I3 , Y plane and found the multiplet of eight particles shown in Fig. 24.6 (there
and in subsequent similar plots the ket symbols | . . .i of the states are omitted). The
center contains, beside the 0 -meson, another pseudoscalar neutral meson called the
meson. Its mass is
m = 548.8 0.6 MeV

(24.56)

1.05 0.15 keV

(24.57)

and it decays with a width

principally into and 3 0 . A ninth much heavier pseudoscalar meson (958) of


width 8.5 MeV known at that time did not fit into the scheme by having a larger
mass and no partners.

1300

24 Internal Symmetries of Strong Interactions

Gell-Mann did the same thing with the spin 1 resonances of negative parity of the
nine pseudoscalar particles, replacing
K

K (892)
(770)
(783)
(1020)

(24.58)

He continued to organize the nucleons and their strange partners in this way and
found again an octet shown in Fig. 24.7 with the masses and lifetimes given in
Table 24.1. Gell-Mann further noticed that there exists no similar regularity also

Figure 24.8 Baryon decuplet states associated with the first resonance of the nucleons.

Table 24.1 Masses and lifetimes of the octet states associated with the nucleons.

p
n
+

0
0
0

m( MeV)
(1010 sec)
938.2796 0.0027
stable
939.5731 0.0027 (898 16) 1010
1189.37 0.06
0.8 0.004
1197.34 0.05
1.482 0.011
1192 0.08 (5.8 1.3) 1010
1115.60 0.05
2.632 0.020
1314.9 0.6
1.642 0.015
1321.32 0.13
1.642 0.015

for the most prominent excited states of the nucleon, the resonance (1232) of
mass 1232 MeV and width 115 5 MeV and with charge states , 0, +, ++ together
with their strange partners (1385) of width 37 MeV and (1530) of width

24.4 SU(3)-Symmetry

1301

10 MeV. When he looked at the masses of these excited states he observed an about
equal spacing when plotted in proportional to Y . When plotted in an I3 , Y system,
the particles fill an equilateral triangle except for a missing lower corner. Starting
from these observation he put forward the hypothesis that just as nuclear forces
are independent under isospin rotations, which form the group of unitary matrices
in two dimensions, they should also be approximately invariant under the most
direct extension of this group, which includes the additional quantum number of
strangeness. So he postulated an approximate invariance of the strong interactions
under the the group of unitary matrices in three dimensions, SU(3). Since the masses
splittings are much larger than in the former SU(2) multiplets, this symmetry is
much more broken in nature then the isotopic symmetry.
In field theory, the group theoretic implications of SU(3) symmetry are studied
by equipping the relativistic fields describing all paricles involved with an extra
SU(3) subscript. First one intruduces a Dirac field with subscripts = 1, 2, 3:

1 (x)

(x) > (x) = 2 (x) ,


3 (x)

(24.59)

which changes under SU(3)-transformations by group elements from the fundamental three dimensional representation of SU(3), called 3. This fundamental property
will be emphasized by calling such Dirac fields,and denoting them by q (x). Omitting the spacetime arguments, the transformation law is:
3 : q U q ,

U SU(3).

(24.60)

The hermitian adjoint of a quark field, (q ) (q ) transforms according to the


complex conjugate representation, called 3 . To emphasize the different transformation behavior we write [U ] as (U ) so that
3 : q U q ,

U SU(3).

(24.61)

The unitary scalar product q q (q ) q is an invariant and the same thing is true
for the contraction of q with any other particle field which transforms like (q ) .
For Dirac fields it is more convenient to work with the fields q(x) = q (x) 0 rather
than q (x) because of their more favorable Lorentz transformation properties, in
particular the Lorentz in variance of the scalar product q(x)q(x). Thus we shall
work with conjugate quark fields
q = (
q 1 , q2 , q3 ).

(24.62)

The contraction q q is obviously invariant under Lorentz- and SU(3)transformations:


q q = invariant
(24.63)
We have seen in Eq. (7.298) that the charge-conjugate Dirac field is directly related
to q by a similarity transformation of the Dirac indices, which for quark fields reads
qC (x) = C qT (x).

(24.64)

1302

24 Internal Symmetries of Strong Interactions

The fields q annihilate antiparticles, and will therefore be referred to as antiquark


fields, ignoring at this point the similarity transformation.
The traceless tensor field obtained from the direct product of q and q
M q q 31 q q

(24.65)

is, in general, a bilocal field containing q (x, t) and q (x , t) at different space points
which are irrelevant to the present discussion and therefore omitted. This field forms
an invariant 8-dimensional representation space of SU(3) which Gell-Mann identified
with the 8 meson states described above. This octet representation is irreducible.
This means that every state of it can be reached by performing group operations on
an arbitrary single fixed state. Gell-Mann called the fields up, down, and strange
quark fields and denoted them by u, d, s, i.e.

q = d .
s

(24.66)

Similarly for the antiquarks:


s).
q = (
u, d,

(24.67)

He associated the octet fields (24.65) with the pseudoscalar mesons as shown in
Fig. 24.9. The field M annihilates the corresponding particle, for example +
annihilates the particle + . In contrast to isospin, the SU(3) symmetry is
M 1 2 = du

Figure 24.9 Quark content of the pseudoscalar meson octet fields. The particle and
quark symbols denote the annihilation parts of the corresponding fields.

strongly broken. The pion and the K mesons have quite different masses, 135 MeV
and 490 MeV.
The operator which measures the third component of isospin, I3 , is the number
operator
1

N
d)].
I3 = [(N
(24.68)
u Nd ) (Nu
2

24.4 SU(3)-Symmetry

1303

u,d,s the total number operator of the u, d, s


Here and in the sequel we denote by N
quarks and their antiparticles, respectively. Thus we have the commutation rules
u,d , q ] = nu,d q ,
[N

(u, d, s)

(24.69)

with the eigenvalues


nu = (1, 0, 0),

nd = (0, 1, 0),

(24.70)

and similar relations between the number operators and the antiquarks q =
s). The hypercharge of the mesons is given by
(
u, d,
1
s)],

+ N
d 2N
Y = [(N
u + Nd 2Ns ) (Nu
3

(24.71)

where Ns counts the number of strange quarks, i.e., it is defined to have, for (u, d, s)
s), the eigenvalues ns = (0, 0, 1) and (0, 0, 1), respectively, with correand (
u, d,
sponding eigenvalues of Ns for the antiquarks.
The group SU(3) is a 3 3 1 = 8 parameter group. Its Lie algebra possesses
8 traceless hermitian generators which are conventionally denoted by a /2 (a =
1, , 8) with the 3 3 traceless hermitian matrices

1 =

4 =

6 =

0
1
0
0
0
1
0
0
0

1
0
0
0
0
0
0
0
1

0
0
0
1
0
0
0
1
0

2 =

5 =

7 =

0
i
0
0
0
i
0
0
0

i 0
0 0
0 0
0 i
0 0
0 0
0 0
0 i
i 0

1
0 0

3
, = 0 1 0 ,
0
0 0

(24.72)

1 0
0

1
8
0
, = 3 0 1
.
0 0 2

Note that they have the same trace orthonormality relations as the Pauli matrices:
tr(a b ) = 2ab .

(24.73)

It is easy to see that the 3 3 -matrices U = eia /2 span the space to unitary 3 3
-matrices of unit determinant. The first three a s are the direct extension of the
i -matrices of isospin into three dimensions.
The group structure is specified by the commutation rules of the generators
[a , b ] = 2ifabc c

(24.74)

where the structure constants fabc are totally antisymmetric in abc, due to the Jacobi
identity satisfied by the commutators. Their values are given in Table 24.4. Let Ga
be the operators in Hilbert space associated with these Lie algebra elements. They
act on the three quark field operators of the fundamental representation as follows:
h

a , q = (a /2) q .
G

(24.75)

1304

24 Internal Symmetries of Strong Interactions

Table 24.2 Structure Constants of SU(3). All elements not given in the table follow
from the total antisymmetry.

abc
fabc
123
1
147
1/2
156
-1/2
246
1/2
257
1/2
345
1/2
367 -1/2
458 3/2
678
3/2
An explicit operator expression for the generators can be given by assigning to the
q-fields canonical commutation or anticommutation relations
[q , q ] =

(24.76)

i simply as follows [recall Section 2.5]


Then we can express the generators G
Ga = q (a /2) q .

(24.77)

a have indeed the same commutation relation


It is easy to see that these operators G
as the matrices a /2:
a, G
b ] = ifabc G
c.
[G
(24.78)
Besides these commutators, some calculations will involve the anticommutators of
the a -matrices. They are given by
n

a , b = 2dabc c + (4/3)ab

(24.79)

where dabc are completely symmetric in abc and have the nonzero matrix elements
shown in Table 24.3.
Consider now antiquark fields transforming according to the the representation

a the antiquark fields behaves as follows


3 of SU(3). Under commutation with G
a , q ] =
[G
q (a /2) = (a /2) q .

(24.80)

The antiquark field have the same commutation rules as the quark fields, i.e., with
the index notation (
q ) (
q ) ,
[
q , q ] = .

(24.81)

The generators of SU(3) have for quark and antiquark fields therefore an operator
form:
a = q (a /2)q q (a /2)
G
q.
(24.82)

24.4 SU(3)-Symmetry

1305

Table 24.3 The symmetric couplings dabc . All values not given explicitly are obtained
from the total symmetry in abc.

abc
118
146
157
228
247
256
338
344

dabc

1/ 3
1/2
1/2
1/ 3
-1/2
1/2
1/ 3
1/2

abc
dabc
355
1/2
366
-1/2
377
-1/2
448 -1/2 3
558 -1/2 3
668 -1/2 3
778 -1/2 3
888 -1/ 3

The generator of the third component if isospin is


3 = q (3 /2)q q (3 /2)
I3 = G
q
1
1

u u d d).
(24.83)
= (u u d d) (
2
2
Similarly, the hypercharge is given by
i
2
2 h 8
Y = G
q ( /2)q q (8 /2)
q
8 =
3
3
i
1h
(u u + d d 2s s) (
u u + d d 2
s s) ,
(24.84)
=
3
and the charge by
= Y /2 + I3
Q
i
1h
(2u u d d s s) (2
u u d d s s) .
(24.85)
=
3
The raising and lowering of the third component is done with the operators
h
i
I = 21 q 1i2 q q (1i2 ) q
=

u d d u
d u u d

s u s u
u s u s

(24.86)

where 1i2 is a short notation for 1 i2 . A pair of operators which raise and
lower the hypercharge by a unit (thereby also raising and lowering I3 by half a unit)
is3
h
i
V = 12 q 4i5 q q (4i5 ) q
=

(24.87)

u u
s s),
Note that there exists a third component to the operators Vpm , V3 = 21 (u us s) 21 (
which commutes with V in the same way as the isospin operators. The three operators form the
so-called V -spin subgroup of SU(3). A similar thing holds for the U operators of (24.88) with
U3 = 12 (d d s s) 21 (d d s s).
3

1306

24 Internal Symmetries of Strong Interactions

The effect of these operators on the quark and antiquark states |q i (created by the
operators (q ) when acting on the vacuum |0i ) is illustrated in Fig. 24.10 [this
being the analog to the illustration in Fig. 4.3]. Note that there exists a related pair

Figure 24.10 Effect of raising and lowering operators on quark and antiquark states.
Here and in the subsequent weight diagrams the ket symbols of the states are omitted.

of operators which raise and lower hypercharge by a unit while acting oppositely to
V in the isospin:
U =

q 6i7 q q (6i7 ) q
(
)
s d s d
=
.
d s d s
1
2

(24.88)

These, however, will play no role in the future discussion.


Consider now the composite states associated with the product representation
3 3 of a quark and an antiquark assigned to the pseudoscalar mesons in (24.65).
The quantum numbers I3 and Y are additive so that the quantum numbers of the
mesons octet can be found by vector addition, i.e. by placing the 3 diagram with
the (I3 , Y ) origin once on each state of the 3 diagram and adding up the vectors as
shown in Fig. 24.11. In the theory of group representations the vectors |I3 , Y i of
the quarks and the antiquarks are referred to as the fundamental weights. Using
the raising and lowering operators I , V it is easy to find all states of an irreducible
representation starting out from any fixed state. For instance, the operators I+ to
the particle state | i = |d
ui, we find

2| 0 i,
I+ | i = |
uu ddi
(24.89)
and further

2| i.
I+ | 0 i = 2|dui

(24.90)

24.4 SU(3)-Symmetry

1307
sd

su

dd

u
d

uu

du

ss

ds

us

Figure 24.11 Adding the fundamental weights in product representation space of 3 and
3 vectors.

Similarly, applying V+ to | i = |
udi gives
V+ | i = |
ssi = |K 0 i.

(24.91)

By applying V+ to |K0 i we find


0 i = |
V+ |K
uu ssi.

(24.92)

This state has the same (I3 , Y ) quantum numbers as | 0 i but is not an eigenstate
of total isospin I 2 . Such a state is given by
|i =

1
6

2
|
uu + dd
ssi

(24.93)

0 i can be decomposed into a 0 and an state as follows:


so that V+ |K
0i =
V+ |K

1
2

+ 21 |
2
|
uu ddi
uu + dd
ssi

1
2

| 0 i +

3
2

|i.

(24.94)

It has become customary (de Swart convention4 ) to change the phases to the q
states so that the matrix elements of I and V are all positive. This is achieved by
changing the phases of the states of the representation 3 as shown in Fig. 24.12. For
the SU(3) representation matrices U , these phase changes on the states correspond

to a similarity transformation. The new matrices will be denoted by U:


U U

(24.95)

Correspondingly, we shall say that the antiquark fields q after the phase changes
transform according to the 3 representation.
The phases of all higher representations are adjusted accordingly. The meson
octet states associated with the product of 3 and 3 states are shown in Fig. 24.13.
As a check we calculate a couple of phases:
4

J.J. de Swart, Rev. Mod. Phys. 35 , 916 (1963).

1308

24 Internal Symmetries of Strong Interactions

|
si

2
3

31 |
ui

+1
+1

|di

21

+ 12

0
T3

Figure 24.12 The states of the 3-representation with de Swart phases.


K 0 =
sd

+1

K + =
su

S 0

=u
d
=

dd)

+ = du
2
(
uu + dd
ss)

1 (
uu
2

0 =

1
6

0 = ds

K
+1

K = u
s
0
T3

Figure 24.13 The quark-antiquark content of the meson octet meson states with de Swart
phase convention.

V+ | 0 i = V+

1
2

1
2

|
uu ddi

|
sui =

1
2

|K + i

(24.96)

and
= |dui
= | 0 i,
V+ |K i = V+ | dsi

(24.97)

which are indeed positive.


Note that these quark-antiquark states can also be obtained by forming certain
combinations of a -matrices and sandwiching them between the creation operator
parts of the quark and antiquark fields
q = (u , d , s ), (
q ) = (
u, d , s ),

(24.98)

24.4 SU(3)-Symmetry

1309

and applying them to the vacuum. The creation operators of the meson octet states
are

M a = q (a / 2)
q,
(24.99)
with the combinations indicated in Fig. 24.14. These combinations of a indices are

K 0 = (6 + i7)/ 2

+1

= (1 i2)/ 2

K + = (4 + i5)/ 2

0 = 3

+ = (1 + i2)/ 2

=8

K = (4 i5)/ 2

0
T3

0 = (6 i7)/ 2
K
+1

Figure 24.14 Combination


of indices a in thepseudoscalar octet field M a of (24.99).

Here (1 i2)/ 2 stands for (M 1 iM 2 )/ 2.

often used to specify octet states instead of the quark content. Any hermitian
octet

operator Oa (for instance Ga itself and M a ) with the indices (1 + i2)/ 2 adds to
a given state the quantum numbers of a + . Let us verify the de Swart phases of
the assignments in Fig. 24.14 by applying I , V to the meson states |Ma i = M a |0i
with the above index combinations. Using the commutation rules
a , M b ] = ifabc M c
[G

(24.100)

we have, for example,


[I+ , M ] =

2f123 M 3 = 2M 0 ,
q

f123 (M 2 + iM 1 ) = M + ,

[I+ , M K 0 ] = (f174 M 4 f165 M 5 ) = 2M K + .


[I+ , M

] = i

1
2

(24.101)

There is another simple way of constructing an octet representation of SU(3)


from quark states without the use of antiquarks. By applying the product of three
fields q q q at three different places5 to the vacuum one obtains 27 states. If these
5

This assumption of different place points is necessary to get all 27 states for otherwise some
states will coincide due to particle identity.

1310

24 Internal Symmetries of Strong Interactions

are decomposed into the irreducible contents of SU(3) they lead to the following
multiplets
3 3 3 = 10 + 8 + 8 + 1.
(24.102)
The octet states can be identified with the nucleon octet states. In the decuplet
there are doubly charged non-strange states which can be identified with the first
nucleon resonance (1236). It is therefore suggestive to identify the other 10 states
with the strange partners of this resonance. At the time when Gell-Mann proposed
this identification he could only match 9 of the decuplet states with known particles
and resonances. There was one more state of strangeness 2 at the bottom of the
weight diagram to be constructed in Fig. 24.17 which closes it to a triangle. Since
the masses within the two rows are almost degenerate and since they grow from row
to row by about the same amount
m m 150,
m m 145,
with the masses given in units MeV, Gell-Mann extrapolated the mass of the tenth
particle to be
M M + 3(M M )
1232 + 3 1567 1682.

(24.103)

He therefore postulated the existence of a particle with this mass, a negative charge,
and a hypercharge 2 as it resulted from the three quark content of the states
and called the particle . This particle was found in 1964 at a mass of 1675 8
Mev. It was a triumph of the approximate SU(3) symmetry hypothesis and the 10
assignment of the excited nucleons and their strange partners.
Let us explicitly construct the irreducible representations contained in the product 3 3 3 and assign them to physical states. We perform the multiplication
successively
3 3 3 = 3 (3 3).
(24.104)

The product of two states 3 3 is easily reduced to


3 3 = 6 + 3.

(24.105)

To see this we write down the antisymmetric and symmetric combinations as shown
in Fig. 24.15. The first is a 6-representation, the second transforms in the same
way as an antiquark representation 3. As before we have assigned the phases to
satisfy the de Swart convention. After forming these products we continue the
multiplication.
3 3 3 = 3 (6 + 3) = 3 6 + 3 3
(24.106)
Let us now form 3 6 and 3 3. The product 3 3 has been calculated before in
terms of quark-antiquark states
3 3 = 8 + 1.

(24.107)

24.4 SU(3)-Symmetry

1311

us, su

ds, sd

dd (ud + du)/ 2 uu
s
3

uu

ud, du

dd

ss

(ud du)/ 2

(us + su)/ 2 +

(ds sd)/ 2

(ds + sd)/ 2

(us su)/ 2

ss
3

Figure 24.15 The quark content in the reduction of the product 3 3 = 6 + 3.

(ud du)/ 2

(us su)/ 2

(ds sd)/ 2
3

d(ud du)/ 2

u(ud du)/ 2

= d(ds sd)/ 2

u(us su)/ 2

s(ds sd)/ 2

s(us su)/ 2
8

1

6

uds
usd
+dsu
dus
+sud
sdu

Figure 24.16 The octet and singlet states obtained from 3 3 in the product space of
tree-quarks.

Now the octet states contain three quarks as shown in Fig. 24.16. The octet states
can be identified with the nucleons and their strange partners.

1312

24 Internal Symmetries of Strong Interactions

The singlet state |s i is the completely antisymmetric combination of uds states:


1
|s i = (|udsi + |dsui + |sudi + |dusi + |sdui + |usdi).
3!
(24.108)
The proton state is

1
|pi = |u(ud du)i.
(24.109)
2
The state |0 i in the octet is found by applying I to the right most state | i =
|u(us su)i thereby obtaining
1
|0 i = |d(us su) + u(ds sd)i.
2

(24.110)

The isosinglet state |0 i is obtained by applying V to |pi = |u(ud du)i/ 2 and


separating the result into a linear combination of |0 i and an orthogonal state which
must be |0 i. First we have
V |pi =

1
2 |sud + usd sdu udsi.
2

(24.111)

The scalar product with |0 i gives


1
h0 |V |pi = .
2
Hence we rewrite
V |pi =

and identify

 0
2 12 | i +

3
2

(24.112)

|0 i

|0 i 1/12|u(ds sd) d(us su) 2s(ud du)i.

(24.113)

(24.114)

Consider now the product 3 6. It gives the states shown in Fig. 24.17. The state
|uuui is identified with the 1232 MeV resonance ++ (1232) of isospin and parity
3/2 . By applying the operators I and V to |uuui we find the ten decuplet states
|++ i
|+ i
|0 i
| i
|+ i
|0 i
| i
|0 i
| i
| i

=
=
=
=
=
=
=
=
=
=

|uuui
1
|uud + udu + duui
31
3 |udd + dud + uddi
|dddi
1
|uus + usu + suui
31
|uds + usd + dsu + dus + sud + sdui
61
3 |dds + dsd + sddi
1
|uss + sus + ssui
31
6 |dss + sds + ssdi
|sssi

24.4 SU(3)-Symmetry

1313
dd

(ud + du)/ 2 uu

(us + su)/ 2

(ds + sd)/ 2
s
ss

uuu

ddd

suu, u(us + su)/ 2

uss, s(us + su)/ 2

sss
10 + 8

Figure 24.17 The irreducible three-quark states 10 and 8 in the product 3 6 (the
symbol (. . .)s denotes complete symmetrization).

In addition to this decuplet, there is a further octet of particle states which can
alternatively be assigned to the nucleon octet. Its quark states are found by constructing one state, say the proton, and applying the operators I ,
V to it. The
proton state is found by linearly composing |uudi and |u(ud + du)i/ 2 so that the
result is orthogonal to |+ i in (24.115) and to |pi of (24.109):
|p i =

1
6

|u(ud + du) 2uudi.

(24.115)

The prime in |p i indicates the different quark wave function with respect to (24.109).
All other octet states associated with |p i are then obtained by applying to it I , V ,
for instance
q
|n i = I |p i = 12 |d(ud du)i.
(24.116)
As a first test of the octet assignment we calculate the mass differences within
the nucleon octet assuming that the mass operator transforms like an SU(3) singlet
plus a form which behaves like a neutral isosinglet member of an octet, i.e., like .
Fortunately, the Glebsch-Gordan coefficients for the matrix elements
h8a |8b |8c i

(24.117)

are already in our position. They are of two types and given by the traces
tr([a , b ]c ) = fabc and tr({a b }c ) = dabc . The first is antisymmetric, the second
symmetric in abc. We therefore expand
h8a |8b |8c i = F ifabc + Ddabc ,

(24.118)

1314

24 Internal Symmetries of Strong Interactions

where F and D are two unknown irreducible matrix elements. To relate these to
particles, we only have to multiply them with the SU(3)-polarization vectors a (m)
of the particles whose assignments are clear from Fig. (24.14). For instance
1
a ( + ) = (1, i, 0, 0, 0, 0, 0, 0).
2

(24.119)

Then we have between particles


h8m2 |8m|8m1 i = a (m2 )b (m)(F ifabc + Ddabc )c (m1 ).

(24.120)

We may also write


h8m2 |8m|8m1 i = F ifm2 m m1 + Ddm2 m m1 ,

(24.121)

with obvious notation. The matrix associated with the b = 8 =


component are
supposed to explain three mass differences. Two of them may be used to determine
the two irreducible matrix elements F and D. The symmetry breaking will then
allows us to derive one relation between the masses. Using the values of fabc and
dabc with b = 8 from Table 24.3 we have for the distances of the masses from some
common SU(3) singlet mass msg

mN = hp|8 |pi
= F f485 + Dd484 = F 3/2 D/23,
m = h+ |8 |+ i =
Dd181 =
D/3,
(24.122)
0
0
m = h |8 | i =
Dd888 =
D/3,

m = h0 |8 |0 i = F f687 + Dd686 = F 3/2 D/2 3.


Using these equations we readily find that the common singlet mass is
msg = 21 (m + m ) = 12 (mN + m ) + 14 (m m ),

(24.123)

so that the masses have to satisfy the famous Gell-Mann Okubo relation
1
2

(mN + m ) = 14 (3m + m ).

(24.124)

It holds reasonably well, the left-hand side being approximately equal to 1.128 GeV
the right-hand side 1.134 GeV.
For the decuplet, the same ansatz reproduces the equal mass splittings between
the rows which had led Gell-Mann to his prediction of the -meson.
For the pseudoscalar meson octet the relative mass splitting is very large and
the mass relation cannot be expected to be as good as for the baryon octet. If we
use, however, the square masses, it is nevertheless in surprisingly good agreement
with experiment:
m2K = 14 (3m2 + m2 ).
(24.125)
The left-hand side is equal to 0.25 GeV2 , the right-hand side 0.23 GeV2 . An argument for using square-masses is that, in contrast to the fermion Lagrangian, the
boson Lagrangian has a mass term proportional to m2 .

24.4 SU(3)-Symmetry

1315

The vector meson octet has a smaller relative mass splitting and the mass relation
should be good. If we insert, however, the experimental masses
m2K = 14 (3m2 + m2 ).

(24.126)

there is a bad surprise: The left-hand side is equal to 0.80 GeV2 while the right-hand
side 0.61 GeV2 . This disagreement can be resolved6 by postulating that the symmetry breaking part of the Lagrangian mixes the isosinglet octet state with the SU(3)
singlet state. This is always possible with the symmetry breaking transforming like
an -meson. Specifically, if 0 , 0 denote the SU(3) states, the physical particles
can be
= cos 0 sin 0
= sin 0 + cos 0 .

(24.127)

with some unknown mixing angle . The symmetry breaking Lagrangian can contain
the mass terms
Lsymm

br

= m28 02 m21 20 m218 0 0 + . . . ,

(24.128)

with some SU(3) octet, singlet, and mixing mass parameters and the associated
SU(3) fields. The octet mass parameter is determined from the Gell-Mann Okubo
relation (24.126) as
m8 = 930 MeV.
(24.129)
Between the physical states (24.127) this becomes diagonal if the mixing angle is
given by
1
m2
= arctan 2 18 2 .
(24.130)
2
m9 m1
and the diagonal masses are
m2 = 21 (m28 + m21 ) 21 [(m28 m21 )2 m418 ]
m2 = 21 (m28 + m21 ) + 12 [(m28 m21 )2 m418 ]

(24.131)
(24.132)

From the experimental numbers we determine the parameters


m18 0.65 MeV,

cos 0.77.

(24.133)

The mixing angle can be tested by experiment since it determines the decay rate of
the processes
K +K .
(24.134)

which occurs with a width 2.2 MeV. Since the vector meson has spin 1, the pseudoscalar kaons must come out in a p-wave. Their orbital wave function is therefore
antisymmetric. But for bosons the total wave function must be symmetric which implies that the SU(3) quantum numbers have to be coupled antisymmetrically. This,
6

J.J. Sakurai, Phys. Rev. Letters 9 , 472 (1962).

1316

24 Internal Symmetries of Strong Interactions

however, excludes the SU(3)-singlet part of from the decay (since its coupling
would be proportional to ab in the SU(3)-indices of the pseudoscalar mesons. The
amplitude for the decay is therefore proportional to the SU(3) Clebsch-Gordan
co

efficient cos (80 |8K + |8K )cos h0, 0| 21 , 21 , 21 if845 = cos (1/ 2) ( 3/2). This
is to be compared with the decay 0
+ whose amplitude is proportional to
(80 |8 |8+ )h1, 0|1, 1; 1, 1if123 = (1/ 2) 1. Taking the squares we find the
ratio


S( K + K ) 2
3



=
cos2 0.44.
(24.135)

S(0 + )
4
The decay rates contain a phase space factor

where q is the momentum of the


of the decaying vector meson. In
1020, q 359, M 770 MeV so
0.078 MeV, respectively. Together

The experimental ratio is

q3
,
M2

(24.136)

emerging pseudoscalar mesons and M the mass


the processes at hand we have qK 127, M
that the phase space factors are 0.0020 MeV and
with (24.135) this gives the ratio of decay rates



( K + K ) 2




(0 + )


( K + K ) 2




(0 + )

0.011.

exp

0.014,

(24.137)

(24.138)

in reasonable agreement with SU(3) symmetry.


For the pseudoscalar mesons the same type of mixing can occur between - and

- mesons. However, since their mass difference is relatively large compared to that
between and their mixing angle is much smaller which explains why no mixing
was needed to satisfy their Gell-Mann Okubo relation (24.125).
As a further test of the octet assignment we may calculate branching ratios for
the decay of the decuplet resonances in the nucleon and the meson octets. Also here
the agreement with experimental data is reasonably good.

24.5

New Quarks

Recently discovered particles and the theoretical attempts to explain their interactions and decay properties suggest the existence of more quarks, a charmed quark
called c, a top quark called t, and a bottom quark called b. It is tempting to try and
incorporate these quarks into the above broken symmetry considerations and extend the basic quark antiquark SU(3) multiplet by these additional quarks to higher
multiplets with an even larger broken symmetry group. This has led to reasonable
particle classifications only if one includes the lightest of the additional quarks, the

24.6 Tensor Representations and Young Tableaux

1317

charmed quark c. The four quarks u, d, s, c may be considered as the fundamental


representation 4 of an internal SU(4) symmetry group, with the antiquarks transforming according to 4. The weight diagram, is three-dimensional and contains an
additional charm axis labeled by C in Fig. 24.18.

Figure 24.18 The four quarks u, d, s, c and their position in the three-dimensional weight
space with the new quantum number charm. By combining a quark and an antiquark one
finds the 16 states on the right-hand side, of which 15 form an irreducible representation
of SU(4). The new states D 0 , D + , F + and their antiparticles have been found in the
laboratory.

24.6

Tensor Representations and Young Tableaux

A systematic construction of the tensor representation in direct product spaces such


as 3 3 3 is possible with the help of Young tableaux. The irreducible representation are obtained by forming tensors which are irreducible under permutations of
the indices. They are of a definite symmetry type specified by the Young tableau
introduced in App. 2A. The tableau 1 2 3 indicates a tensor of complete symme1
corresponds to a complete
try transforming like a 10 representation, while 2
3
antisymmetric tensor transforming like an SU(3) singlet. For the mixed tableau
1 2
we adopt the first convention to first symmetrize |q q q i in the first two
3
indices, then antisymmetrize in the first and the third index. Equivalently we can
1 2
2 1
use another tableaux, say
or
. Each gives an octet representation.
3
3
For instance, the proton wave function (24.109) arises from the symmetry operations
2 1
applied to the state |uudi. The symmetrization in the
of the Young tableau
3
first two indices gives
|uudi 2|uudi

(symmetry in 21),

(24.139)

1318

24 Internal Symmetries of Strong Interactions

the antisymmetrization in the second and third index (of the original tensor |q q q i)
|uudi 2|u(ud du)i

(antisymmetry in 23).

(24.140)

Hence:

2 1
|uudi = 2|n(nd du)i = 2 2|pi
3

(24.141)

Equivalent proton state could have been obtained from the tableaux
1 2
|uudi = 2|uud duui,
3

(24.142)

1 2
|udui = 2|udu duui,
3

(24.143)

or the sum of the two which gives


|u(ud + du) 2duui.

(24.144)

This is the proton state |p i of Eq. (24.115) found in the product 6 3.


Which mixture of the two independent three-quark wave functions is verified by the
proton in nature cannot be decided by SU(3) symmetry alone but must emerge from
a dynamical model. We shall see later that the larger symmetry group SU(6) makes
definite predictions.
In general, the dimensionality of the SU(n) representation of a tableaux is obtained from the following formula
dSU(n) =

D(l1 , l2 , . . . , ln )
,
D(n 1, n 2, . . . , 0)

(24.145)

where
D(x1 , . . . , xn )

(xi xj ),

(24.146)

i = 1, 2, 3, . . . , n,

(24.147)

i<j

and
li n + mi i,

with mi being the number of boxes in the ith row of the tableau. Note that also
rows with no boxes have to be counted, up to i = n.
A useful alternative formula for the dimension is
dSU(n) =

1
l1 !l2 ! ln !
,
Q
0!1! (n 1)! i,j hij

(24.148)

where hij is the sum of the number of boxes to the right of the element ij in the
tableau plus the number of boxes below ij plus 1.

24.6 Tensor Representations and Young Tableaux

1319

Take some examples. The trivial Young tableau


has m1 = 1, m2 = 0, m3 =
0, . . . , mn = 0 and thus l1 = n, l2 = n 2, . . . , ln = 0. Hence
d

= n.

(24.149)

The first non-trivial one


has m1 = 2, m2 = 0, . . . , mn = 0 and thus l1 =
n + 1, l2 = n 2, l3 = n 3, . . . , ln = 0. Hence
d

= n(n + 1)/2,

(24.150)

= n(n 1)/2,

(24.151)

which for SU(3) gives 6. Further


d

which for SU(3) gives 3, the dimension of the representation 3.


For 3-quark states, the dimensions are
d
d
d

= n(n + 1)(n + 2)/6,

(24.152)

= (n 1)n(n + 1)/3,

(24.153)

= n(n 1)(n 2)/6,

(24.154)

which for SU(3) take the values 10, 8, 1, respectively. Somewhat more generally, the
dimensions of some simple tableaux are
d

1 2 3

1 2 3

n+k1
k
n+k1
k+1

n
.
d
=
1
k
2
..
.

(24.155)

k,

(24.156)

(24.157)

k
A column with n boxes is completely antisymmetric in the corresponding indices
and thus proportional to the invariant antisymmetric tensor i1 i2 ,...,in . This is why

1320

24 Internal Symmetries of Strong Interactions

its dimensionality is 1. In constructing all Young tableaux any such column can be
omitted. For instance, in SU(3),
=

(24.158)

One often indicates this cancellation by crossing out any complete column with a
vertical line:

(24.159)

The representation associated with each Young tableaux occurs as often as dimensionality of the associated representation of the permutation group. This follows
from the fact that the representation matrices of the permutation group commute
with those of SU(3). They can therefore be brought simultaneously to a block form
along the diagonal

D1
D1
D2
..

(24.160)

Along the diagonal there are ni identical irreducible representation blocks Di of dimensionality di , etc. Now, according to Schurs lemma, any matrix commuting with
such a block matrix must be itself a block matrix with the dual form which consists
of di identical blocks of dimensionality ni . The formula for the dimensionality of the
irreducible representations of the permutation group were given in App. 2A. If the
Young tableau has r boxes the dimensionality is given by
r!
,
ij hij

d Sr = Q

(24.161)

where the numbers hij are the same as in (24.145).


Using this formula we find, for instance, that the product of three-quark states

decompose into one SU(3) decuplet

and one singlet

, two octets

:
3 3 3 = 10 + 8 + 8 + 1

(24.162)

24.6 Tensor Representations and Young Tableaux

1321

The Young tableau can also be used to find the irreducible contents in a product
of two or more irreducible representations. For instance, suppose we want to know
the irreducible contents of 8 8. We take the two Young tableaux

and distinguish the rows of boxes in the second factor by a, b, c, . . .. In the example
a a
. Then we add the lettered boxes to the first tableau in the following way:
b
a) add all the as, so as to get all proper tableaux (i.e. with mi+1 mi ) which has
no more than one a in each column.
b) add the bs once more in according to the same rule
c) add the cs . . . .
In this way we obtain in our example the expansion

a a
b

a a

88

a a
+
b

27

10

a
a b

a
a

+
b

10

(24.163)

8
a

a
8

a
a b
1

In accordance with the rule given earlier we have dropped all complete columns.
Note that because of the antisymmetry of any complete column, an incomplete
column is equivalent to the complex conjugate representation associated with the
Young tableau obtained from the missing boxes necessary to complete it (indicated
ny a box with a circle), i.e.

3
This is how we can see immediately that the representation associated with the
tableau

is a 10



10

1322

24 Internal Symmetries of Strong Interactions

The octet representation, being the adjoint representation of SU(3) , is real [its
generators are (Ga )bc = ifabc with real antisymmetric matrices (fa )bc ; recall (4.83)].
this is obvious since the missing boxes

From the Young tableau of the octet


form once more the same tableau
, so the representation is equal


to the complex conjugate of itself and thus real.
in

24.7

Effective Interactions among Hadrons

There exists a great variety of strongly interacting particles which many possible
interactions between them. Let us study a few of them.

24.7.1

The Pion-Nucleon Interaction

The most important strong interaction is that between pions and nucleons since
it provides for the dominant part of the forces which keep the nuclei together. If
we neglect electromagnetic interactions, this interaction exhibits isospin symmetry
pions nucleons, so that a system of and has the following effective action
A = A + AN + AN N
o
1n
[a (x)]2 m2 a2 (x)
2
Z
4

=
d x N(x)(i/
MN )N(x)

A =
AN

d4 x

AN N = gN N

(x)i5 a N(x)a (x)


d4 x N

(24.164)
(24.165)
(24.166)
(24.167)

Here a (x) is an isovector field and N(x) a Dirac spinor field with an additional
isospinor index which is not explicitly written down. The matrices a are 2 2 Pauli
matrices acting on the isospinor indices of N(x). Studies of pion nucleon scattering
amplitudes determine the size of the pseudoscalar coupling constant gN N to be
2
gN
N
14.4 .4
4

(24.168)

In some analyses, the pion nucleon interaction in (24.167) is parametrized in the


axial vector form
Z
fN N
d4 xN (x) 5 a N(x) a .
(24.169)
AN N =
M
For physical nucleons on the mass shell, the two interactions are the same with the
relation between the two coupling constants
fN N =

M
g.
2MN

(24.170)

24.7 Effective Interactions among Hadrons

1323

Thus
2
fN
N = 0.081 0.002.

(24.171)

In either form, the coupling strength is too large to perform perturbative calculations
with the interaction AN N .
It is useful to find the SU(3)-symmetric extension of the pion nucleon interaction. In SU(3), there are two ways of coupling three octets with each other, an
antisymmetric one fijk a symmetric one, dijk . The easiest way to do calculation is
by writing the nucleon octet as a 3 3 -matrix.

+ 16 0
+

12 0 +

1 0
2

B=

1 0
6

p
n
26 0

(24.172)

and the pseudoscalar octet as

1 0
2

+ 16 0
+

12 0 +
0
K
K

P =

1 0
6

K+

K0
.
26 0

(24.173)

In this context the phases of are chosen as (1 i2 )/ 2 so that P is a


hermitian matrix. Then the interaction Lagrangian has a simple explicit form
LPBB = g

F
B]P ) + g D tr({B,
B}P )
tr([B,
F +D
F +D

(24.174)

where g gN N is the pion nucleon coupling constant and F/D the so called F over-D ratio, which gives the relative strength of antisymmetric and symmetric
coupling. Working out the different components, this has the expansion in terms of
isospinors

 + (1 2)( )  + 3 34 NN
3 2
K ) 
K)  (
(1 2)(N

3
2
(K N) 
(K ) 3
N
(1 2)
K
3


N)
Lint = g (N

3 4
3 2
3 4
c

+ K
K N
+ K c
3
3
3
2
2
)
+i(1 )( ) + ( ) + (
3
3

2
2 0
+ ( ) 0
3
3

  


 

 
(24.175)

1324

24 Internal Symmetries of Strong Interactions

Here we have used the isospinors


p
n

(24.176)

and
K+
K0

0
K
K

(24.177)

. The parameter is defined by


Note that K c is the charge conjugate spinor C K
D
F +D

(24.178)

F/D = (1 )/.

(24.179)

p 0 K +

(24.180)

so that

The two numbers g = gN N and F/D parametrize all couplings.


The pion nucleon coupling was written down in Eq. (24.168). The F/D-ratio is
determined experimentally from analysis of the production process

which contains the coupling

2g(1 2)
gn K + =
F D
.
= 2g
F +D

(24.181)

From the experimental data one finds


F
2

D
3

(24.182)

For the coupling between pions and -particles, the interaction Lagrangian
(24.175) yields
F
g = g(1 ) = g
,
(24.183)
F +D
The SU(3)-relations between coupling constants in the Lagrangian (24.175) are,
of course, in agreement with standard SU(3) Clebsch-Gordan coefficients. The coefficients of the SU(2) subgroup can be taken from Table 4.2 in Chapter 4.1. The
isoscalar factors form Table 24.4.
Take, for
instance, the matrix element hp| 0 |pi. The SU(2) Clebsch-Gordon
coefficient is1/ 3. The antisymmetric
and symmetric isoscalar factors are 3/ 12 =

1/2 and 9/ 20 = 3/2 5, respectively. Hence, with the normalization factors of


(24A.13),
F +D
.
(24.184)
hp| 0 |pi =
2

24.7 Effective Interactions among Hadrons

1325

+ 0
+
The matrix element
h
|
|
i
has
an
SU(2)
Clebsch-Gordan
1/
2 and the
q
isoscalar factors 2/3 and 0, respectively, so that
h+ | 0 |+ i = F,

(24.185)

just before in (24.183).


Let usqalso calculate the coupling gn K + for which the Clebsch-Gordan coeffi
cient is 2/3 and the isoscalar factors are 1/2 and 3/2 5, respectively. Thus
1
hn|K + | i = (F D)
2

(24.186)

F D
gn K + = 2
F +D

(24.187)

and

as in (24.181).

24.7.2

The Decay (1232) N

The interactions between the various members of the nucleon octet with pseudoscalar
mesons and the decuplet resonances of spin-parity J P = 3/2+ are all determined
by only a single SU(3)-invariant matrix element. This, in turn, may be chosen to
coincide with the coupling between the proton, the + and the ++ (1232) resonance.
The Lagrangian for this specific interaction is conventionally written as
L

++ pn+


g  ++ +
=
+ + c.c.
MN

(24.188)

where is a Rarita Schwinger spinor for the spin 3/2 particle ++ (1232). Some
people write 4h/M instead of g /MN .
The isospin structure can be incorporated by multiplying the Lagrangian by the
Clebsch-Gordan factor
h 23 c| 12 a1bi
h 23 23 | 21 21 11i

(24.189)

where c, a, b are the isospin orientations of , N and . Alternatively, we can


use Rarita Schwinger isospinors a (each of the three components 1 , 2 , 3 is
a two-component isospinor) for the particle and write in isospace (suppressing
the Lorentz indices)
L=

g 

N N
MN

 


(24.190)

Indeed, the isospinor for the ++ resonance has only a +-component with isospin
++ p + as in
up so that N contains the particles ++ , p, and + in the form
(24.188).

 

1326

24 Internal Symmetries of Strong Interactions

For calculations it is useful to record the completeness relation of RaritaSchwinger isospinors


X
s3

b = ab 1 1 a b = 2 ab 1 1 [a , b ],
a
3
3
6

(24.191)

where 1 stands for the 2 2 unit matrix in isospace.


The coupling g can be determined by an analysis of the strength of the ++
resonance in the + p scattering amplitude. This gives
g 12.43 0.92

(24.192)

h 0.457

(24.193)

corresponding to
If we assume the resonance to be very narrow we can immediately calculate the
decay rate from the interaction (24.188). The decay amplitude is found as follows.
We consider a ++ resonance in its rest frame and use helicity amplitudes along the
direction of the momentum of the emerging proton, which we assume to run in the
3-direction. The helicity can be 3/2 or 1/2. Only the latter two orientations
allow to conserve angular momentum with the helicities of the proton being 1/2.
The two orientations decay with the same rate, so we only need to study one case, say
the helicity +1/2. The Rarita-Schwinger spinor is a Clebsch-Gordan combination
of spin J, M = 1, 0 with 1/2, 1/2 and of J, M = 1, 1 with 1/2 1/2:
u (0; 23 , 21 ) =
Between the state

0
0
0
1

u(0,

1
2

)+
6

0
1
i
0

u(0, 12 ).

a+ ap |0i

(24.194)

(24.195)

and a state at rest h0|a with unit normalization in a finite volume V , the wave
functions are for the pion and the nucleon

1
1
eiq x u(pN , s3 )
2E V
V

(24.196)

and for the resonance


1
u (p , s3 ).
V

(24.197)

We therefore find the decay amplitude (the t-matrix)


t =

1
g
1
1
q
u u q 3
MN
2E EN /MN
V

g
=
MN

1
1
1
2
,
cosh pN
3q
3
2
2E V
EN /MN

(24.198)
(24.199)

24.7 Effective Interactions among Hadrons

1327

where
u(pN , 21 )u(0, 12 ) = cosh

1
=
(EN + MN )
2
2MN

(24.200)

with EN , E being the energies of proton and pion in the rest frame of the decaying
resonance and pN the momentum of the nucleon emerging in the z-direction. The
decay rate is found from formula (9.336), for the helicity h = 1/2 state:
1
EN E
d|t|2
2
3
(2)
M
pN g 2 2 2 1
4
2
p (EN + MN )
=
(2)3 M MN2 3 N 4
g 2 2 p3N
=
(EN + MN ).
4MN2 3 M
Z

h=1/2 N =

(24.201)

Since the is with probability 1/2 in the helicity 3/2 state where it cannot decay,
the final decay rate is only one half of (24.201), i.e.,
g 2 1 p3N
=
4 3 M

EN + MN
MN2

(24.202)

In terms of the three particle masses involved, the energy EN is given by


M 2 + MN2 M2
EN =
2M

(24.203)

and the momentum p of the outcoming proton (and pion) by


pN =

[M2 (MN + M )2 ][M2 (MN M )2 ]/2M .

(24.204)

Using the experimental width


115MeV,

(24.205)

we obtain
g 14.43 1.07g

(24.206)

h 0.53,

(24.207)

corresponding to

in reasonable agreement with independent determination (refx9.21), (24.193) from


the pion nucleon scattering amplitudes.
It should be noted that the ratio g /g 1.07 is a manifestation of SU(4) symmetry. Since the resonance and the nucleon are in the same SU(4) multiplet 20
(corresponding to the SU(6) multiplet 56), all pion couplings are controlled by only

1328

24 Internal Symmetries of Strong Interactions

one irreducible matrix element. In calculating them we have to take account of the
fact that, in order to preserve parity, the pseudoscalar pion must emerge in a p wave.
The matrix element between a + resonance and a proton
1
u (0, 21 )u(pN , s3 )iq
MN

(24.208)

has for small momentum pN in the z-direction the limiting form

1
2 pN
i
(p1N ip2N )u(0, 12 ) .
u(0, 21 )u(0, 12 ) +
3 MN
6MN

(24.209)

Thus it can be written in terms of two component spinors as


2

2 pN 1
( ) ( 21 ).
3 MN 2

(24.210)

The p-wave emission leads to the appearance of a spin matrix i acting upon the
helicity index of the nucleon. The same thing is observed for the matrix elements
in the pseudoscalar pion nucleon coupling
u(0, 12 )i5 u(pN , 21 ).

(24.211)

For small momentum, this becomes


i (( 21 ), ( 12 ))

1
0
0 1

e/2 ( 21 )
e/2 ( 21 )

pN 1
( ) ( 12 ). (24.212)
MN 2

We now add the internal SU(2) structure and the SU(4)-matrix associated with
the emission of the pion in the z-direction of + and the nucleon transforms like
the tensor operator
a =
a3.
(24.213)
The ratio of the couplings p 0 p and + 0 p can therefore be found quite simply by
applying 3 3 to the proton state

2h
p =
u u d + u d u + d u u
(24.214)
3

1





(u u d + u d u + u d u + u u d + d u u + d u u ) ,
2
which gives
3 3p =

2h
3(u u d + u u u + d u u )
(24.215)
3

1





+ (u u d + u d u + u d u + u u d + d u u + d u u ) .
2

24.7 Effective Interactions among Hadrons

1329

This gives the matrix elements


2
1
p p =
33
g
4

2
33+
+ 3 3 p =
g
3 3

5
6 = ,
3

1
4
6 =
2.
2
3


(24.216)

For the ratio between g and g we have to go from + 3 3 p to the matrix element
++ + 3 p which involves a Clebsch-Gordan factor
4
1 4
(24.217)
2= .
++ + 3 p =
3
23
Thus we obtain
g
45
4
=
= = 0.8
(24.218)
g
33
5
in reasonable agreement with the experimental ratio 0.92 from the analysis of pion
nucleon scattering amplitudes.
By a similar quark calculation, we can predict from SU(6) symmetry the F/D
ratio of the coupling between nucleon octets and pseudoscalar mesons. According
to the interaction Lagrangian (24.175), the coupling + + 0 is given by
2F
.
(24.219)
F +D
But this matrix element is easily calculated. The SU(6) state is obtained from +
by applying V to it which changes the d quark into an s quark. Hence

2
+
(u u s + u s u + s u u )
(24.220)
=
3

1
u u s + u s u + s u u + u u s + u s u + s u u .
2
Applying 3 3 gives

2
+
2(u u s + u s u + s u u ).
(24.221)
3 3 =
3
The scalar product with + gives
4
+ 3 3 + = .
(24.222)
3
In comparison with the pion nucleon matrix element gN N = g, this amounts to the
relation
4
g0 + + = g.
(24.223)
5
We therefore find the F/D ratio the SU(6) value
g+ + 0 = g(1 ) = g

F/D = 2/3.

(24.224)

in good agreement with the experimental determination. In App. A we have collected


a few useful formulas for the SU(3) calculations involved.

1330

24.7.3

24 Internal Symmetries of Strong Interactions

Vector Meson Decay (770)

Among mesons, the most directly observable coupling is the coupling. It governs
the strength of the most prominent resonance in the S(770) pion pion scattering
amplitude. Since is a vector meson of isospin 1 it is coupled to the two pions in a
p- wave. The Lagrangian interaction is given by
g
L=
2

  

(24.225)

For the decay + 0 , the invariant matrix element t is given by


1
1
q
2p (p , s3 )
h 0 + |t|+ i = g q
2
2V Epi 2V M

(24.226)

where (p , s3 ) is the polarization vector of the meson and the decay rate is,
according to formula (9.336),
2
+ 0 + = g

4 pCM 1
4|p |2 .
(2)2 M2 23

(24.227)

The average over the initial polarization gives


p p
1 X
1
p p
(p , s3 ) (p , s3 ) = p p g 2
3
3
p
s3

1
= p2CM ,
3

(24.228)

where pCM is the momentum of the pions in the rest frame of the decaying -meson:
pCM =

1/2
1 2
M 4m2
.
2

(24.229)

This brings (24.227) to the form

+ 0 +

2
2
2
2
g
g
2 p3CM
2 M 4m
=
=
4 3 M2
4 3
8M2

3/2

(24.230)

The experimental decay width is


153MeV.

(24.231)

From this one finds the coupling constant


2
g
2.85.
4

(24.232)

24.7 Effective Interactions among Hadrons

24.7.4

1331

Vector Meson Decays (783) and (783)

Another important coupling is that between , , and :


L = g

 .

(24.233)

It is responsible for the decay


(24.234)

= 9.8 .3 MeV

(24.235)

which proceeds at a rate

The channels are dominated by the resonance. One can therefore assume a
sequential decay and finds that


3 = (M 3M )4 M2 4M2
10.3 MeV

2
2
g
g
,
4 4

2

2
2
g
1 1 g
M M2
3.56
27 4 4 4

(24.236)

where the number 3.56 is due to the numeric integration of a phase space integral7 .
The experimental rate gives
2
g
0.95
,

4
M2

(24.237)

and hence
2
g
1
0.117

6.14

.
g 2
M2
GeV2

24.7.5

(24.238)

Vector Meson Decays K (892) K

In K scattering one observes the SU(3) partner of the meson, the strange vector
meson K (892). Its coupling is written as

(24.239)

2 p3CM K
,
3 MK2

(24.240)

LK K = gK K (K K)i
.
2
Its decay width is
K K = 3
7

1
g
2 K K

2

For details see M. Gell-Mann, D. Sharp, W.G. Wagner, Phys. Rev. Lett. 8, 261 (1962)

1332

24 Internal Symmetries of Strong Interactions

where
pCM K

1
=
2MK

rh

MK2 (MK + M )2

ih

MK2 (MK M )2

(24.241)

is the momentum of the decay products in the rest frame of K . This implies the
ratio with respect to the -width
2
2
2
3
K K
3 gK
3 gK
K M pCM K
K

0.385.
=
2
3
2
2

4 g MK pCM
4 g

(24.242)

The experimental width of K is


K K = (51.8 0.8) MeV,

(24.243)

so that its ratio with respect to the -width (24.231) is 0.3406, implying a ratio of
coupling constants
gK K
1.086.
(24.244)
g
By SU(3) symmetry, this ratio is predicted to be unity, in good agreement with
experiment.

24.7.6

Axial Vector Meson Decay a1 (1270)

A more involved coupling governs the decay of the axial vector meson a1 (1270). It
is seen as a resonance of width
a, 316 45MeV

(24.245)

in the scattering amplitude. Angular momentum and parity allow for s- and
d-wave interactions, so that there exist two independent coupling constants with a
Lagrangian density
L = ga1 (a

 )+h

a1

( a

 ) .

(24.246)

If (p), (q) denote the polarization vectors of the a1 and meson, respectively, and
k the momentum of the outgoing pion, the invariant matrix element for A+ +
reads
tba =

1
2V

h
i
1
ga1 (p) (q) ha1 q (p)p (q) .
p0 q0 k0

(24.247)

For the calculation of decay rates it is better to use a more complicated looking
decomposition into a longitudinal and a transverse helicity amplitude
tL
tT

p p
q q
1
= 3
g 2 (p) (q),
gL P q g 2
q
p
p0 q0 k0
2V
gT
(24.248)
q p q p (p) (q).
= 2

ma

24.7 Effective Interactions among Hadrons

1333

The relation between them is


ha1

m2a1 + m2 m2
= gL + gT
,
2m2a1

ga1

gT m2a1 + m2 m2
=
m2a1
2

!2

m2a1 m2 .

(24.249)

The advantage of these amplitude is that they are decoupled in the decay process.
A somewhat tedious calculation (see Appendix 24B) gives
a1 = La1 + Ta1 ,

(24.250)

gL2 1 p5
,
n 3 m2

(24.251)

where
La1 =

Ta1 =

gT2 2 p5
.
n 3 m2a1

Since the experimental masses having the ratio


m2a1
2.72,
m2

(24.252)

we have


a1 5.5 105 gL2 + 2.735 gT2 GeV,

(24.253)

and find for the coupling constants the relation


gL2 +

2m2 2
g gL2 + 0.68gT2 75.1.
m2a1 T

(24.254)

The ration between the coupling constants gL and gT has to be determined by some
other method.

24.7.7

Coupling of (770) to Nucleons

The coupling is defined by the nonminimal Lagrangian density


L = gN N

+ N N (

2M

.
)

(24.255)

An analysis of the phase shifts in N N yields the coupling strengths [1]


2
gN
N
= 0.6 0.1,
4

NN 6.6 1.0.

(24.256)

1334

24 Internal Symmetries of Strong Interactions

Appendix 24A

Useful SU(3) Formulas

The matrix elements of an octet operator between two octets labeled by the real
indices i = 1, 2, 3, . . . , 8 of the adjoint representation 8 can be written as follows
k )ij
h8i |8k 8j i = F (ifkij ) + Ddkij = F (Fk )ij + D(D

(24A.1)

where fijk are the structure constants of SU(3) and dijk those of the anticommutators
4
{i , j } = dijk k + ij .
3
The SU(3) vector ei associated with a specific particle is found from the assignment
of the nucleon octet to the 3 3 -matrix elements

N =

1 0
2

1
6

+
12 0 +
0

1
6

p
n
16 2

By contraction this with the -matrices and forming



1 
ei = tr i N .
2

(24A.2)

(24A.3)

The proton, for example, corresponds to the octet vector

0 0 1
1 i

i
e (p) = tr 0 0 0 .
2
0 0 0

(24A.4)

It has the non-zero components

1
e4 (p) = ,
2

i
e5 (p) = .
2

(24A.5)

0ij between two proton


For the 0 couplings we calculate the matrix elements F0ij D
states as follows: This gives
1
hp|F0 |pi = ,
2

0 |pi = 1 .
hp|D
2

(24A.6)

F +D
.
2

(24A.7)

Hence
hp|0 |pi =

Similarly, the SU(3) vector of the + particle is

0 1 0
1

ei (+ ) = tr i 0 0 0 .
2
0 0 0

(24A.8)

Decay rate for a1

Appendix 24B

1335

It has the non-zero components




1
e1 + = ,
2



i
e2 + = .
2

(24A.9)

0 )ij between
The coupling of 0 to particles we need the matrix elements (F0 )ij , (D
+
-states and find
h+ |F0 |+ i = 1,

0 |+ i = 0.
h+ |D

(24A.10)

Hence
h+ | 0 |+ i = F.

(24A.11)

The matrix elements of + between and pp have therefore the ratio


+ + +
2F
=
,
pp
F +D

(24A.12)

just as in (24.183), (24.185), and (24.219).


The SU(3)-matrix elements may also be calculated by using tables of the SU(3)
i correspond to the coefficients
Clebsch-Gordan coefficients. The matrices Fi D
Fi =

3h8a |88i,

i =
D

5
h8s |8 8i.
3

(24A.13)

The Clebsch Gordan coefficients can be decomposed into a product of an isospin


SU(2) Clebsch Gordan coefficients and a so-called isoscalar factor
h8a i|8j; 8ki = hh8a Yi |8Yj ; 8Yk ii(T i , T3i |T j , T3j ; T k , T3k ),

(24A.14)

where T i , T3i is the isospin contents of the SU(3) index i. The isoscalar factors are
listed in Table 24.4.

Appendix 24B

Decay rate for a1

Denoting the polarization tensors of the a1 and mesons (p) by , (q) by , and
using the transversality properties p =
0, q = 0, respectively, the amplitudes
are (dropping the normalization factors 1/ 2Nq0 associated with the three particles)
t = gL (p )(q) +
= gL (p )(q) +

gT
Ma21

q p
q q 2 qp
p qp p2

i
gT h 2 2

2
.
M
M
(
)
+
(p
)qp

(
)(qp)
a1

Ma21

(24B.1)

1336

24 Internal Symmetries of Strong Interactions

Table 24.4 Isoscalar Factors of SU(3). We have omitted the


square roots of the number
in the arrays on the right-hand sides, i.e., -3 stands for 3. This is indicated by the
superscript 1/2 on the arrays.

188


1/2
K) = 1
2 3 1 2
() (N K
8
81 8 8

N N K K
N
NK

K

K
NK


K K

83 8 8

9 1

1 6 0

20 2 12
9 1

9 1
4 4 6
4 2
9 1

1/2

1/2

N N K K
N
3 3 3 3

NK

2 8 0 0 2

NK

6
12 6 0 0
K

K

3 3 3 3
10 8 8

1/2

6 6
N N

NK
K
1

2 2 3 3 3

=
K

3 3 3 3
K
12

12
K
8 10 8
1/2

12 3
K
N

K

K
1
8 2 3 2


9 6
K
15

3 3 3 6
K K

10 10 8

1/2

K
15 3 6

K

K
1

8 8 0 8

K
K

24 12 3 3 6

K
12 12

Notes and References

1337

This expression shows directly the relation of the couplings gL and gT with the
couplings ga1 , and ha1 . We now put a1 in its rest frame so that p = (Ma1 000)
and have the scalar products
q0
(q, 12 ) (p, 12 ) = 1,
(q, 3)(p, 3) =
M
(p (q, 12 )) = 0,
pCM
(p (q, 3)) =
Ma1 ,
M

(q(p, 12 )) = 0,
(q(p, 3)) = pCM ,

(24B.2)

where
pCM =

1 q 2
[Ma1 (M M )2 ][Ma21 (M M )2 ]
2Ma1

(24B.3)

is the center-of-mass momentum of the -meson. We therefore find the helicity


amplitudes (writing the helicities as superscripts in parentheses)
h(0) 1 |t|a1 (0)i = gL p2CM
1

h 2 |t|a12 i =

Ma1
,
M

(24B.4)


gT 
2
2
2
2
2
M
M
+
E
M
a1

CM a1 = gT pCM ,
Ma21

where

Ma21 + M2 M2
ECM =
(24B.5)
2Ma1
is the energy of the meson in the center-of-mass frame. From this we obtain
directly the longitudinal width
La1 =

gL2 2 p5CM
,
4 6 M2

(24B.6)

having inserted a factor 2 due to isospin, and a transversal width


Ta1 =

gT2 2 p5CM
2
.
4 6 Ma21

(24B.7)

Notes and References


For discussion of internal symmetries see the book by
S. Gasiorowicz, Elementary Particle Physics, John Wiley & Sons, N.Y. 1966.
The individual citations refer to:
[1] G.E. Brown, R. Machleidt, Strength of the -Meson Coupling to Nucleons,
Phys. Rev. 50, 1731 (1994).

Paradise is exactly like where you are right now.


...only much, much better
Laurie Anderson (1947)

25
Symmetries Linking Internal and Spacetime
Structure. Supersymmetry
Besides space-time symmetries under the Lorentz and Poincar`e group and approximate internal symmetries under SU(2), SU(3), etc., there exists a further class of
approximate symmetries which combine external and internal indices. Historically,
the first example for this class of symmetries arose in nuclear physics where isospin
was combined with spin to form the larger group SU(4).

25.1

Approximate SU(4)-Symmetry of Nuclear Forces

Soon after the discover of the approximate isospin symmetry in nuclear physics, the
search went on for even higher approximate symmetries. It turned out that the
nuclear forces had also an approximate independence of spin. Moreover, this could
be combined with isospin to form an approximate invariance under the larger group
SU(4). The different terms in the potential between nucleons are conventionally
classified by their behavior under certain interchange operators acting upon twonucleon wave functions. These are defined as follows:
P M (x1 a1 , x2 a2 ) (x2 a1 , x1 a2 ) Majorana Operator,
P B (x1 a1 , x2 a2 ) (x1 a2 , x2 a1 ) Bartlett Operator,
P H (x1 a1 , x2 a2 ) (x2 a2 , x1 a1 ) Heisenberg Operator,

(25.1)

where x denotes the positions and a the third components of spin, s3 , of the particles.
The effect of the interchange operators on two-nucleon wave functions of definite
orbital angular momentum l and spin (singlet or triplet) is shown in Table 25.1.
Using these operators, the most general two-particle potential can be written as
follows
V = VW (r) + VM (r)P M + VB (r)P B + VH (r)P H
+ VT W (r)S12 + VT M (r)S12 P M ,

(25.2)

where S12 is the spin orbit coupling operator


S12 3(1r)(2r) (1 2 ).
1338

(25.3)

25.1 Approximate SU(4)-Symmetry of Nuclear Forces

1339

Table 25.1 Action of the different interchange operators.

PM
PB
PH

even l
triplet singlet
1
1
1
1
1
1

odd l
triplet singlet
1
1
1
1
1
1

Here r r/|r| is the direction of the distance vector r between the two particles.
The -matrices 1 , 2 act on the spin indices of the particles 1, 2, respectively. This
is, in fact, the most general way of describing noncentral forces which can be derived
from a potential. The potential () is called Wigner potential and is purely central.
The potential VT W proportional to the operator S12 provides for the so-called tensor
force which is responsible for mixing the 31 D wave into the 31 S ground state wave
function1 of the deuteron ( 2 6 %). Note that in a spin singlet states S12 = 0.
The potential VM is called Majorana potential . The Bartlett operator PB can be
expressed in terms of the Pauli spin matrices as follows
1
P B = (1 + 1 2 ).
2

(25.4)

In this parametrization, each of the six V -terms in (25.7) is proportional to the unit
matrix in the space of charge indices.
Let us now define, in this space, the operator
1
P (1 + 1 2 ),
2

(25.5)

where  denotes a three-component vector of Pauli matrices,  = ( 1 , 2 , 3 ), which,


instead of spin, act now on the isospin indices of the nucleons. It is easy to verify
that P H = P for wave functions which satisfy the Pauli principle and that
P M = P BP H .

(25.6)

The action of the spin isospin operators is shown in Table 25.2. It is then possible
to rewrite the potential (25.1) in terms of the -matrices as follows:
V

= Va (r) + V (r)(1 2 ) + V (r)(1 2 ) + V (1 2 )(1 2 )


+ VT d (r)S12 + VT (1 2 )S12 .

(25.7)

Its coefficients are related to the earlier ones by


Va = VW + 12 VB 12 VH 14 VM ,
V = 21 VH 41 VM ,
VT d = VT 12 VT M ,

V = 21 VB 14 VM ,
V = 14 VM ,
VT = 21 VT M .

(25.8)

1340

25 Symmetries Linking Internal and Spacetime Structure

Table 25.2 Action of spin and isospin operators in the expansion (25.7).

even l
triplet singlet

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

3
1
3

odd l
triplet singlet

1
3
3

3
3
9

1
1
1

In 1937, Wigner2 observed that if the spin-orbit forces were small and Wigner
and Majorana forces dominant, the potential could be written as
V = V1 (r) + V2 (r)(1 + 1 2 )(1 + 1 2 )

(25.9)

and would be invariant under spin rotations, isospin rotations, and the extension of
these two SU(2) groups to SU(4). Experimentally,
VH
0.4 0.5,
VM

(25.10)

so there is relly quite large symmetry breaking. Still, the group has turned out to
be quite useful in understanding nuclear spectra.
The group SU(4) is generated by the 15 4 4 matrices acting on a nucleon field
N with isospin index ( = 1, 2) and spin index ( = 1, 2):

1, 1

.
(25.11)
2
2 2
2
The commutation rules are given by the known ones for , , an additional set of
obvious commutation rules between the s and s and the combined generators
(/2) (/2) expressing their vector nature with respect to each subgroup, plus the
new ones
#
"
k
c
a i b j
= iab ijk 1
,
+ iabc ij 1.
(25.12)
2
2
2
2
If SU(4) is a good approximate symmetry of nuclear forces then nuclei should occur
in irreducible multiplets of SU(4). Moreover, since both isospin and  are invariant
under space reflections, each multiplet should all have the same parity.
The fundamental representation of SU(4) is four-dimensional and denoted by 4.
Its four components are identified with the four states of the nucleon as follows

N =

p
p
n
n

(25.13)

Notation: JS L where L is the orbital angular momentum, S the combined spin, and J the total
angular momentum.
2
E.P. Wigner and E. Feenberg, Phys. Rev. 51 , 106 (1937) and Rep. Progr. Phys. 8 , 274
(1941).
1

25.1 Approximate SU(4)-Symmetry of Nuclear Forces

1341

The higher representations are obtained as usual. One forms an appropriate multinucleon state of highest spin and isospin and applies the lowering operators

1,
2

,
2

3
,
4

3
,
4


.
4

(25.14)

to generate irreducible multiplets.


and take the state of highest weight
Consider the symmetric Young tableau
in isospin and spin

E

(25.15)
|1, 1; 1, 1i p p ,

where the notation on the left-hand side specifies the isospins and spins T, S and
their third components as |T, T3 ; S, S3 i.3 From the state (25.15) we run through the
entire isovector triplet of spin 1 by repeated application of the lowering operator .
For these the other spin components are obtained by applying .
The application of /2 to (25.15) gives the state

and /2 leads to

E
1
|1, 0; 1, 1i = n p + p n
2

|1, 0; 1, 0i =

E
1
n p + n p + p n + p n .
2

(25.16)

(25.17)

Since the two operators , are taken from the SU(2)T SU(2)S subgroup, they
do not change T, S. This is only possible by applying a proper SU(4) generator such
as /4. It creates the mixed state
E
1
n p + p n .
2

(25.18)

Combining this with (25.17) we now form the state


E
1
|0, 0; 0, 0i = n p p n n p + p n
4

(25.19)

which is orthogonal to (25.17) and has the property that

1
n n n n n n + n n = 0,
|0, 0; 0, 0i =
2
2
E

1
n n p n n p + p n = 0,
|0, 0; 0, 0i =
2
2

(25.20)
(25.21)

showing that it is a singlet under isospin and spin operations which was anticipated
by the zeros in the ket on the left-hand side of (25.19). Thus we have found an
irreducible decuplet of SU(4) with 10 states and the [T, S] decomposition
10 = [1, 1] + [0, 0].
3

(25.22)

In accordance with the names for the isospin and spin matrices , we denote here the iospin
operators by T rather than I, as is customary in nuclear physics.

1342

25 Symmetries Linking Internal and Spacetime Structure

In addition, there is a six-dimensional representation 6 with the states


E
1
|0, 0; 1, 1i = p n n p ,
2
E

|0, 0; 1, 1i = p n + p n n p n p = 2 |0, 0; 1, 0i ,
2
2

E



|0, 0; 1, 0i =
p n n p = 2 |0, 0; 1, 1i
2

E
1

|0, 0; 1, 1i = n n n n = |1, 1; 0, 0i
4
2
+
E

|1, 1; 0, 0i = p n + n p p n n p = 2 |1, 0; 0, 0i
2
2
+

E


p p p p = 2 |1, 1; 0, 0i .
|1, 0; 0, 0i =
(25.23)
2

These have the [T, S] content

6 = [1, 0] + [0, 1].

(25.24)

These are all irreducible contents in the tensor product, i.e.,


4 4 = 10 + 6.

(25.25)

The deuteron is a state in the irreducible representation 6. It has [T, S] = [0, 1],
i.e., it is an isosinglet vector particle transforming like the SU(4) generator /2 1.
If SU(4) is to be a good approximate symmetry, there should also be a partner [1, 0],
i.e., a spin-zero isotriplet transforming like the SU(4) generator 1 /2. This does
not really exist. However, the singlet NN scattering length is very large in the T = 1
state indicating a bound state just around the cut of the two-particle continuum in
the energy plane. The symmetry breaking prevents these SU(4) partners to appear
as physical particles.4
In elementary-particle physics, the group SU(4) combining isospin and spin has
played quite an important historical role. There, up and down quarks with spin up
and down are identified with the representation 4 of SU(4):

4=

u
u
d
d

(25.26)

For mesons, the representation 15 in 4 4 = 15 + 1 extends the isospin triplet


pseudoscalars + , 0 , , which transform in the same way as the generators of the
isospin subgroup a , by the vector mesons + , 0 , , which transform like a i ,
4

For more detail see the paper A. Franzini, A. Radicati, Phys. Lett 6, 322 (1963).

25.1 Approximate SU(4)-Symmetry of Nuclear Forces

1343

Figure 25.1 Pole position of the would-be SU(4) partner of the deuteron with spin-1 and
isospin-0. It is close to the D-pole but not a particle since it hides in the complex energy
plane just below the two-particle cut.

plus an isosinglet i which is identified with the meson. The symmetry breaking,
however, is extremely large since
m 135 MeV,
m 770 MeV,
m 783 MeV.

(25.27)

For the nucleons p, n, the SU(4) symmetry is much better. Taking three up and
down quarks with spin up and down, the three Young tableaux
,

(25.28)

correspond, according to the general dimension formula,


n(n + 1)(n + 2)
,
6

(n 1)n(n + 1)
,
3

n(n 1)(n 2)
6

(25.29)

to one symmetric 20, two mixed 20, and one antisymmetric 4 representation.
Actually, the latter 4 is equivalent the complex conjugate 4 of the fundamental
representation. To see this we note that the completely antisymmetric representation

of rank 4 is one-dimensional which means that the antisymmetric 4 must have met
with a 4 to form an invariant.
In order to find the isospin and spin contents of the representations (25.28), we
write the SU(4) index as , where are the two isospin indices and the spin
indices. The representation 4 is a doublet with respect to each subgroup
a

= 2 2.

(25.30)

1344

25 Symmetries Linking Internal and Spacetime Structure

The Young tableau


= 10 is symmetric under the simultaneous exchange of
both indices. Hence, with respect to the two subgroups it can become a product of
symmetric tensors or antisymmetric tensor
a b

10

(25.31)

Similarly
a

6 =

(25.32)

To obtain all irreducible tensors of rank 3 we multiply

and

by another

box and find within SU(4) the decomposition

10

20

(25.33)

20

and

(25.34)
4 .

20

As far as the isospin and spin contents are concerned, we multiply the decomposition
(25.31) of the 10 by
=

.
(25.35)
4
2
2
Thus the product 10 4 contains the isospin spin products

(25.36)

(25.37)

2
is one-dimensional. Hence

In the second case there is only one product since


we can decompose 10 4 as follows

10

=
4

+
4

+
4

(25.38)

25.1 Approximate SU(4)-Symmetry of Nuclear Forces

1345

(25.39)

On the left-hand side, the product reduces to


+

(25.40)

20

20

It is easy to read off which products on the right-hand side belong to these two
representations.
=

20

(25.41)

20

4
(25.42)

Thus the symmetric representation 20 consists of a spin- 12 isospin- 12 and a spin- 23


isospin- 23 multiplet. They are associated with the following particles. The spin- 12
isospin- 21 states are the two isospin states of the nucleon p, n. The spin 32 isospin
3
states contain the first pion nucleon reson (1230) plus its isospin partners with
2
the charge states , 0, +, ++. All particles have intrinsic parity +.
The mixed 20 consists of a spin- 23 isospin- 12 , a spin- 21 isospin- 32 , and spin- 12 isospin- 12
multiplet. These states are also found experimentally in pion-nucleon scattering as
resonances of negative parity. The isotriplet spin- 12 states are the P wave resonances

of spin-parity S P = 12 called (1620), the isodoublet spin- 32 states are the P wave

3
N(1520) resonances, and the isodoublet spin 21 states are the P wave 12 N(1650)
2
resonances.
The content of the antisymmetric SU(4) representation
= 4

is obviously 2 2 , i.e. an isospin 12 spin 21 multiplet. It is associated with the


+
N(1440) pion nucleon resonance of spin-parity 12 .

1346

25 Symmetries Linking Internal and Spacetime Structure

25.2

Approximate SU(6)-Symmetry in Strong Interactions

With the success of the approximate symmetry SU(4) in nuclear and particle the
question arose whether there exists an extension of the symmetry group which includes also the internal symmetry SU(3). The three quarks with spin up and down
form a sextet of particles

u

u

d

(25.43)
q=
d .


s
s
Hence the most straightforward extension would be SU(6). This group is generated
by the 35 matrices
a
i
a i
1, 1 ,
.
(25.44)
2
2
2
2
The only non-trivial commutation rules are
"

a i b j
1
1
2
1
= idabc ijk c k + ifabc ij c 1 + i ab ijk 1 k . (25.45)
,
2
2
2
2
3
2
#

Since in SU(4), the representation 4 4 contained the pseudoscalar and vector


mesons , , we expect the representation 6 6 to extend these multiplets by the
missing SU(3) partners. The representation 6 6 decomposes into 35 + 1 with the
particles in the 35 representation having the transformation properties of the 35
group generators (25.44). Thus there is now an SU(3) octet of pseudoscalar mesons
which transform like the generators a 1/2. There is a further SU(3) octet of
vector mesons , , K . They form 24 states which transforms like the generators
a i /2, and there is an SU(3) singlet vector meson (1020) which transforms like
the three generators 1 i /2. These meson multiplets are quite widely separated in
mass and the symmetry SU(6) is strongly broken. Still, the approximate symmetry
group organizes nicely all the most important low-lying meson states as illustrated
in Fig. 25.2.
As in the case of SU(4) symmetry, the situation is much better for the baryons.
The nucleons and their strange partners as well as the first nucleon resonances and
their strange partners lie all in one irreducible multiplets of three-quark states. They
decompose as follows:

=
6

56

+ 2

+
70

20

(25.46)

25.2 Approximate SU(6)-Symmetry in Strong Interactions

1347

K
J P = 0

J P = 1

J P = 1

Figure 25.2 The pseudoscalar and vector mesons of the 35 representation of SU(6).

In order to find the SU(3) and spin contents of these multiplets we decompose, as
in the case of SU(4),
a

(25.47)

and form the symmetric and antisymmetric products of two representations 6. They
are 21- and 15-dimensional and their SU(3) and spin contents are given by
=

a b

a b

21

a
b

15

Multiplying the 21 by a further 6 we find

21

56

10

+
70
+
8

+
2

(25.48)
2

1348

25 Symmetries Linking Internal and Spacetime Structure

J P = 1/2+

J P = 3/2+

Figure 25.3 The SU(3) content of the particles in the 56 representation of SU(6) containing the J P = 1/2+ nucleons and the J P = 3/2+ resonances.

It is then easy to read off the SU(3) and spin contents on the right-hand side

56

=
70

10

10

+
8

8
+

8
+

2
4

(25.49)
2

Thus the 56 consists of a decuplet of spin 32 and an octet of spin 21 states. These can
be associated with the decuplet of nucleon resonances and the nucleon octet, all of
positive parity as shown in Fig. 25.3.
The 70 consists of a decuplet of spin 21 , an octet of spin 23 , an octet of spin 12 , thus
extending the isospin multiplets of negative parity, assigned in the SU(4) case, to
the SU(3) multiplets shown in Fig. 25.4.
Figure 25.4 The nucleon resonance of negative parity in the 70 representation of SU(6).

The SU(3) SU(2) contents of the antisymmetric SU(6) representation

20

25.2 Approximate SU(6)-Symmetry in Strong Interactions

1349

are easily written down as

20

+
2

(25.50)

The octet states may be the spin 21 particles of positive parity shown in Fig. 25.5.
The SU(3) singlet 23 + partner of them has not yet been found.
N (1440)

(1660)
(1660)
(?)

Figure 25.5 Octet of spin-parity

1+
2

baryons.

Let us now study the states of the 56 representation in more detail. The state
of highest weight is the ++ (1235) resonance of spin 23

++ (1235) = u u u .

(25.51)

By lowering the third components of spin and isospin once, we obtain the normalized
state
1
+ (1235) =
d u u + u d u + u u d
3
+d u u + u d u + u u d
E

+d u u + u d u + u u d .

(25.52)

Applying to this the operator 3 3 , we find


1
3 3 + (1235) = 3(d u u + u d u + u u d )
3
E
(u d u + u u d + d u u + u u d + d u u + u d u ) .(25.53)
This should be a mixture of + (1235) and a proton p .
Since the proton is orthogonal to + (1235) , we can immediately guess a normalized proton state:

2

p =
d u u + u d u + u u d
(25.54)
3
+
1





(u d u + u u d + d u u + u u d + d u u + u d u ) .
2

1350

25 Symmetries Linking Internal and Spacetime Structure

To be sure it is a pure proton state we simply apply the isospin raising operator and
find that the state is annihilated, thus guaranteeing the isopin to be 1/2. Applying
the isospin lowering operator, we obtain the neutron state

2

d d u + d u d + u d d
(25.55)
n =
3
+
1





(d u d + d d u + u d d + d d u + u d d + d u d ) .
2
As an interesting consequence of SU(6) we calculate the ratio between the magnetic moment of neutron and proton. For this we assumes that the magnetic moment
operator transforms under SU(3) like the charge operator which is given by the generators
!
1 8
1 3
Q
+ .
(25.56)
Q= =
2
3
Under rotations, it must transform like a spin operator . Thus the magnetic
moment operator should be proportional to the following generator in SU(6)
1
3

1
2

 = (3 + 8 ) .

(25.57)

We now apply the third component of this,


1
1
3 = (3 + 8 ) 3 ,
2
3

(25.58)

to the quark states and find the eigenvalues shown in Table 25.3. Specifically, the
Table 25.3 Eigenvalues of charge and other operators on quark states.

q
3
3

(8 / 3) 3
Q 3

u
1
1
3
2
3

u
1

31
32

d
1
1
3
1
3

d
1

13

1
3

s s
0 0
32

13

2
3
1
3

application of 3 gives

2 5

(d u u + u d u + u u d )
(25.59)
3 |p i =
3 3
+


1
1





(u d u + u u d + d u u + u u d + d u u + u d u ) ,
2
3

2 4
3 |n i =
(d d u + d u d + u d d )
(25.60)
3 3
+
 
1 2





(d u d + d d u + u d d + d d u + u d d + d u d ) ,
2 3

25.2 Approximate SU(6)-Symmetry in Strong Interactions

1351

so that we find the matrix elements


2 5
1 6
=
p
= 1,
3
9 3
3 4


D E
2
4
2
2 6
n 3 n =
3+
= .
9
3
3 4
3
D

E

3 p

(25.61)
(25.62)

This yields the ratio of magnetic moments

2
n
= .
p
3

(25.63)

Experimentally, the magnetic moments are


p =
2.7928444 0.0000011,
n = 1.91304308 0.00000054,

(25.64)

in units of ch/2mp c = 505 1024 erg/Gauss. Thus the ratio is 2.055/3, in


excellent agreement with the SU(6) prediction (25.63).
By the same method we can calculate all rates of the radiative decays of the
decuplet resonances 10 8+ in terms of only one unknown overall size parameter.
The possibility of predicting the ratio of the magnetic moments is due to the
fact that in the irreducible multiplet 56 of the symmetry group SU(6), the SU(3)
octet operator a 3 /2 has, between the octet states of nucleons, a specific ratio of
symmetric and antisymmetric couplings (which within SU(3) would be arbitrary).
Let us express the two couplings in terms of the irreducible matrix elements F and
D as follows
*
a

b
2

+

c

= F (ifab c ) + D (dab c )
F (Fa )b c + D (Da )b c .

(25.65)


Then we find between protons with the octet components (4 + i 5)/ 2
1
(F3 )pp = f345 = ,
2

1
(D3 )pp = d344 = ,
2

and between neutrons with the components (6 + i7)/ 2


1
(F3 )nn = f367 = ,
2

1
(D3 )nn = d366 = .
2

(25.66)

(25.67)

The eighths components of Fa , Da have the matrix elements


(F8 )pp = f845 =
(F8 )nn = f867 =

q2

3
,
2

(D8 )pp = d866 = 21 3

(D8 )nn = d866 = 21 3 .

(25.68)

1352

25 Symmetries Linking Internal and Spacetime Structure

The matrix elements of the magnetic moment operator


1
1
3 = Q 3 (3 + 8 ) 3
2
3

(25.69)

are therefore given in terms of the irreducible matrix elements F and D by


1
1 1
1 1
+D
= F+ D ,
+

2 2
2 6
3




1 1
2
1 1
+D
= D.
hn|3 |ni = F +
2 2
2 6
3
hp|3 |pi = F

The ratio is

F/D + 31
hp|3 |pi
.
=
hn|3 |ni
23

(25.70)

(25.71)

This holds on the basis of only SU(3) symmetry. Inserting the SU(6) result (25.63)
we see that SU(6) fixes the F/D ratio
2
F
= .
D
3

(25.72)

The absolute values of the irreducible matrix elements of the operators


3
3 ,
2

8
3 ,
2

(25.73)

are also known from the above reasults (25.61) and (25.62):
F +D

1
= 1,
3

2
2
D=
3
3

(25.74)

so that

2
(25.75)
F = , D = 1.
3
It is straight-forward to calculate the magnetic
moment of any other members

+
of the octet, for instance of =
(1 + i2)/ 2:
D

+ |3|+

= F (ifQ12 ) + D dQ11
"

1
1
= F
f312 + f812
2
3
1
1
1
=
F+ D= .
2
6
2

!#

"

1
1
+D
d311 + d811
2
3

!#

(25.76)

This implies the ratio


D E

p 3 p

E
D

+ 3 +

= 2.

(25.77)

25.2 Approximate SU(6)-Symmetry in Strong Interactions

1353

In fact, the fastest way to find the F/D ratio is by calculating from the SU(6) wave
functions the matrix elements of 3 3 /2. Between protons we find immediately
*
3

p
2

+

p

5
= .
6

(25.78)

Using the wave function





E
2 
+
(25.79)
u u s + u s u + s u u
=

3
+
1
(u u s + u s u + s u u + uu s + u s u + s u u ) .
2
we obtain


+

+

2
= .
(25.80)
2
3
In terms of the irreducible matrix elements F, D, the corresponding matrix elements
would be in general
*
3

p
2

+

p

F
D
= + ,
2
2

Hence we identify, once more,

+

p

= F.

(25.81)

2
F = , D = 1.
(25.82)
3

The
E magnetic moment for the transition p is found by multiplying

+
3 p from the left by h |. This gives


E

D
2 5
1
2

+
3+ 6 =
3 p =
2.
(25.83)
9 3
6
3
Comparison with the proton matrix element (25.61) yields the SU(6) prediction
2
2p .
=
(25.84)
3
Experimentally one finds from the (1236) resonance in the photoproduction cross
section
+ p 0 + p
(25.85)
the value

exp (1.28 0.01) .

(25.86)

This is not in quite as good agrement as the proton-to-neutron ratio (25.63). We


must, however, keep in mind that the symmetry SU(6) is significantly broken and
that the comparison with the data requires taking matrix elements of between
states whose normalization factors can depend on the masses of the particles. Thus
one is bound to commit errors of the order of the relative mass differences within the
experimental multiplets, here between the (1232) and the proton p(938). These
can easily accounts for the 28 % discrepancy in the magnetic moment prediction
(25.86).

1354

25.3

25 Symmetries Linking Internal and Spacetime Structure

From SU(6) to Current Algebra

Historically, the SU(6) symmetry had an important impact upon the development
of the quark model. Since the symmetry is only approximate there was need to find
a way of controlling the symmetry breaking corections. For the group SU(6) this has
remained impossible, not so, however, for an important subgroup of SU(6) generated
by a /2, a 3 /2. It is called SU(3)R SU(3)L and its generators are conveniently
taken in the combinations
a (1 3 )

.
(25.87)
Qa =
2
2
These generators are intimately related to the SU(3) SU(3) current algebra which
was discussed before in the context of chiral symmetries. The charges and axial
charges observed in electromagnetic and weak interactions are related to Qa as
follows:
Qa = Qa+ + Qa , Qa5 = Qa+ Qa .
(25.88)
This relation makes it possible to calculate the symmetry breaking corrections in
terms of experimentally observable quantities. The good SU(6) results for the magnetic moments were based
on the prediction of the F/D ratio 2/3 and the ratio
+
hp|3 |pi /h |3|pi = 2 2/3. It is gratifying to observe that these predictions are
already contained in the SU(3) SU(3)R subgroup of SU(6). Within the framework
of current algebra it will, moreover, be possible to calculate symmetry breaking
corrections to these ratios.
The representations 56 and 70 of SU(6) have the following irreducible contents
with respect to this subgroup
3
2

1
2

56 =
(10, 1) (6, 3)10
(6, 3)8
70 =

3
2

(8, 1)

1
2

(33)]8
(3, 3)8
(3, 3)1

1 [(6, 3)
2

12
32
(3, 6)10 (1, 10)

(3, 6)8

21
32
1
[(3, 6) + (3, 5)]8 (1, 8)
2
(3, 3)8
(3, 6)1

(25.89)

(25.90)

The columns in these matrices ordered according to the third component of the spin
s3 along the direction of the momentum of the particle which is indicated in the
upper row and the subscripts denote the coupling of the two parts of SU(3)R SU(3)L
with respect to the ordinary SU(3) symmetry group generated by Qa+ + Qa . Within
this subgroup, the F/D ratio 2/3 is the F/D ratio found for the operator Qa+ Qa
between two (6, 3)8 representations and the transition p comes from the matrix
element of Qa+ Qa = a 3 /2. The representation (6, 3) contains an SU(3) octet and
a decuplet which can be identified with the nucleon octet and the first resonance
decuplet at helicity s3 = 1/2. between the octet and the decuplet parts of the
irreducible representation (6, 3).

25.3 From SU(6) to Current Algebra

1355

At this point it is useful to realize that the commutation rules of SU(3)L SU(3)R
can be rephrased as sum rules for the irreducible matrix elements of Qa , Qa5 . Take,
for example, the commutation rules between the axial charges
[Qa5 , Qb5 ] = ifabc Qc .

(25.91)

Taking the commutators for a = 1 + i2, b = a

3
[Q+
5 , Q5 ] = Q ,

(25.92)

and sandwiching it between proton states, it gives


E
E D
D
+



Q
Q

p

p
Q
p Q+
5
5 p
5
5

= p Q3 p .

(25.93)

Inserting between the axial charges a complete set of intermediate states, we obtain
E
D
2
n Q5 p

E
E 2
D
D
E
D
2

3

++ +

p

p


Q

Q

p
.
+ 0 Q


=
p
5
5

(25.94)

We now introduce irreducible matrix elements of the axial charge Qa5 with respect
to the SU(2) subgroup by defining between protons
D E

p Q35 p

Then

E
D


n Q
5 p

1
GA .
2

(25.95)

GA .

(25.96)

This can either be derived directly by using the generator 3 /2 or via the WignerEckart theorem for SU(2) vector operators. We only have to keep
in mind
that Q35

2 + , 2 if we
creates the quantum numbers of 0 while Q
5 creates those of
take + , 0 , to satisfy the Condon-Shortley phase convention5 We define further
between a proton and a -resonance
D


E
1


+ Q3T p G
2

(25.97)

and have, by the Wigner-Eckart theorem,


E



++ Q+
5 p

0 Q
5 p

3
G,
2

1
= G .
2

(25.98)

With the right-hand side of (25.94) being the isospin value of the proton 1/2 we
obtain the sum rule
G2A G2 = 1.
(25.99)
5

Verify this by applying Q to Q3 giving Q .

1356

25 Symmetries Linking Internal and Spacetime Structure

Similarly, we can take the commutator (25.92) between two ++ states and have
the sum rule


D

D
E
E

++ 2
+ ++
= 3.
(25.100)
+ Q5
p Q5

Introducing the irreducible matrix element GA of the axial charge between two states via


E
D
1


+ Q05 + = GA
(25.101)
2
so that, by the Wigner Eckart theorem,


E
D


++
=

+ Q

3GA ,
(25.102)
5
and (25.100) becomes

1 2
G + GA 2 = 1.
2
+
Between p and , finally, the commutator (25.92) becomes
D

E
ED
E D

ED



0
+ + 0


Q
n Q

p
+

5 p
5
5
E



ED
D
++

++ +

+ Q
5 p =
5

+ Q+
5 n

(25.103)

0,

(25.104)

which amounts to the sum rule

G GA 5GA G = 0.

(25.105)

The unique solution to these SU(2) SU(2) sum rules is


5
4
1
GA = , G = , GA = .
3
3
3

(25.106)

A more covariant looking derivation of these sum rules is possible by rewriting


the matrix elements of Qa5 for a = 1, 2, 3 between isospin 21 and 23 spinors as follows

a
3 1

a 11

NN : (Q5 )mm = GA (m) (m ) = G


h 2 , m |1, a; 12 , m i
2
2


E
D
3

3 3

1
,
(25.107)
,
m

1,
a;
,
m

(m)(m
)
=
G
=
G
N : (Qa5 )31

2
2
mm
2 a
2
E
15 D 3

3
3
.
: (Qa5 )33
2 , m 1, a; 2 , m
mm = GA iabc b (m)c (m ) = GA
2
2
D

Here SM s1 m1 s2 m2 are the usual SU(2) Clebsch-Gordan coefficients and a are


Rarita-Schwinger isospinors. They are obtained by coupling an ordinary isospinor
with an isovector va by a Clebsch-Gordan coefficient
a (m) =
They satisfy the constraint

3
2

(25.108)

a a (m) = 0

(25.109)

, m 1, m1 ; 21 , m2 va (m1 )(m2 ).

25.3 From SU(6) to Current Algebra

1357

and the identities


iabc b c = a , a b = iabc c .

(25.110)

Inserting the matrix elements into the commutation relations (25.91) between nucleon and -states, inserting a complete set of intermediate states, and using the
covariant matrix elements (25.107) together with the completeness relations
X

(m) (m) = , ,

a (m)b (m) =

1
2
ab [ a , a ],
3
6

(25.111)

we find once more the sum rules (25.99), (25.103), (25.105).


So far, these relations hold for the SU(2)L SU(2)R subgroup of SU(3)L SU(3)R .
Extending the procedure from SU(2) to SU(3), we decompose the irreducible matrix
of Q35 between protons, remembering (25.66), as
GA = F + D,

(25.112)

and consider once more the same commutators (25.92) between + states, giving
D

E
0 + 2
Q5

E 2

E 2

+
+
+ h| Q
+ 0 Q
= 2,
5
5

(25.113)

where 0 is the strange partner of the 0 resonance. The matrix element Q35
between two + states is, using the SU(3) decomposition as in Eq. (25.65),
D

+ Q05 + = F.

(25.114)

Hence, by the Wigner-Eckart theorem for the SU(2) subgroup,



E
D

+

=

0 Q
2F.
5

(25.115)

Using once more the SU(3) decomposition (25.65), we calculate the matrix element
D

+
0 Q
5

= F if1i2,8,(1+i2)/2 + D d1i2,8,(1+i2)/2

= D 2d118 =

2
D.
3

(25.116)

Similarly, the matrix element of


E

1


0 Q
5 p = G
2

(25.117)

+
is related to h0 | Q
5 | i by SU(3) symmetry. Using the Clebsch-Gordan coefficients of the Particle Data Table we have

D
D

E



0 Q
5 p

+
0 Q
5


3
1
,

2
2 1, 1;

1
2

1
a,
, 21
2
E

= h1, 0|1, 1; 1, 1i
1
=
2

2
1
a = a,
12
3 2

(25.118)

1358

25 Symmetries Linking Internal and Spacetime Structure

where a is some number. Using these equations we find from (25.117)




E
1
+
= G .
0 Q
5
2

(25.119)

The sum rule (25.113) becomes therefore

1
1
F 2 + D 2 + G2 = 1.
3
8

(25.120)

Given the solution of the SU(2)L + SU(2)R algebra (25.106), we find from this
additional relation the F/D ratio 23 .
For completeness, let us also look at the commutation rule (25.92) between +
particles, which gives the sum rule
E 2
D
+



0 Q
5

E 2

+
= 1.
(25.121)
+ 0 Q
5

Within SU(3), + , 0 have the vector indices 6 i7/ 2 and 4 i5/ 2 so that the
decomposition (25.65) reads

+
Q
5

= F (if1i2,(4+i5)/2,(6i7)/2 ) + D d1i2,(4+i5)/2,(6i7)/2
= F 2f147 +D (2d146 )
= F D.

(25.122)

The matrix element between 0 (1530) and + , on the other hand, has the SU(3)
irreducible matrix elements

E
D

1


2 h 21 , 21 |1, 1; 12 , 21 i a
0 Q + =
2
s

2 1
=
a,
(25.123)
2
3 2

to be compared with
D

E


Q
5 p

Hence

and the sum rule reads


This is satisfied by

1
=
2 h 32 , 12 |1, 1; 12 , 21 i a
2
!
1
1
2 a.
=
3
2

E

1
+
= G ,
0 Q
5
2

1
(F D)2 + G2 = 1.
2

2
F = ,
3

1
D= ,
3

4
G = .
3

(25.124)

(25.125)

(25.126)

(25.127)

25.4 Supersymmetry

25.4

1359

Supersymmetry

In 1974, a new type of symmetry was discovered,6 group in a nontrivial way. It


permits irreducible multiplets which contain simultaneously fermions and bosons.
The basic example is given by the free action
A=

1
1
1
1
1
M) ,
d x (A)2 + (B)2 + F 2 + G2 + M(F A + GB) + (/
2
2
2
2
2
(25.128)
4

where A, B, C, D are four real scalar fields, and is a four-component real Majorana
spinor. The four scalar fields can be transformed into the four components of the
Majorana field by the following transformations
=

[F + i5 G i/
(A + i5 B)] ,

= [F + i5 G + i(A + i5 B) / ],
A = ,
B = i5 ,

(25.129)

F = (i/
),
G = [i/
(i5 )].

(25.130)

Here is a four-component real parameter with spinor indices. The spinor character
makes it necessary to require and to be Grassmann variables which anticommute
with each other
= .
(25.131)
It is easy to verify that these transformations leave the action invariant. For the
mass terms we calculate,
(F A) = (i/
)A + F ,
(CB) = [i/
(i5 )]B + Gi5

(25.132)
(25.133)


= [F
+ i5 G i/
(A + i5 B)] + [F + i5 G + i(A + i5 B) / ]. (25.134)
The real Majorane spinor satisfies
T

= 0 = C =

0 2
2 0

(25.135)

so that

= T 0T = T 0 = T 0T = ,

5 = T 0T 5 0 T = T 5 T = 5T = 5 .
6

(25.136)

D.V. Volkov, V.P. Akulov, Zh. Eksp.Teor. Fiz. (Pisma) 16 , 621 (1972); Phys.Lett. B 46 , 109
(1973); V.P. Akulov, D.V. Volkov, Teor. Mat. Fiz. 18 , 39 (1974); J. Wess and B. Zumino, Phys.
Lett. B 49 , 521 (1974), B 59 , 163 (1975), Nucl. Phys. B 70 , 39 (1974), B 76 , 310, B 78 , 1 (1974).

1360

25 Symmetries Linking Internal and Spacetime Structure

Recall that in the Majorana representation


5 =

2
0
0 2

so that 5T = 5 .
Hence the F and G terms in the transformation (25.134) can be combined to
()
= 2(F + i5 G),

(25.137)

and cancel in the transformation of the total mass term


1

F A + GB .
2
What about the A, B terms. We write these terms in (25.134) as
()

= {[i

i(A + i5 B) }
(A + i5 B)] +


/ i(A + i5 B) i(A + i5 B) / .

(25.138)

The first term is a pure surface term and does not contribute to the total action. In
the third term we use

= ,

5 = 5 ,

(25.139)

and see that it is equal to the fourth term. Both together cancel the AB terms in
(F A + GB). Thus the total action is invariant. The transformations are called
supersymmetry transformation and the action supersymmetric.
Note that the fields F, G carry no gradients in the action and are therefore
not dynamical. They could in principle be eliminated from the action using the
equations of motion, but then the verification of the supersymmetry would be much
more involved. For this they are very useful since only if they are explicitly present
the four components of have four real scalar counterparts A, B, G, F . In a free field
theory, the equations of motion make F, G vanish. When interactions are included,
they are still non-dynamical, but no longer zero.
In particles physics, no indication of some broken supersymmetry has so far
been discovered. There exists, however, a supersymmetry between the unphysical
content in gauge fields and the Fadeev-Popov ghost fields in QED. Thus is essentail
for removing the vacuum energy of the two unphysical modes. Recall the Lagrangian
(7.515):
1
C. (25.140)
Ltot = L+LGF +Lghost = F 2 D(x) A (x)+D 2(x)/2i C
4
It is is invariant under the following two sets of supersymmetry transformations:
A = i C,
C = D,
C = D = 0,

(25.141)

25.4 Supersymmetry

1361

and
= i C,

C
= D,
= 0.
C = D

(25.142)

These are known as the BRS transformations (see Footnote 25 in Chapter 7).
The BRS supersymmetry is generated by charges
Q=

k (ak,l ck + h.c.),

=
Q

k (ak,l ck + h.c.).

(25.143)

Together with the subsidiary conditions (7.513), the conditions


Q|0i = 0,

=0
, Q|0i

(25.144)

are seen to enforce precisely the Gupta-Bleuler subsidiary condition (7.492).7


The full power of the BRS supersymmetry is needed in the renormalization of
nonabelian gauge theories. This is necessary for extracting finite results from higherorder perturbative calculations of weak interaction processes within the standard
model of electroweak interactions.
Note that by postulating the absence of ghosts from the Hilbert space at a specific
time, they are excluded for all times since the Lagrangian preserves ghost charge.
This a consequance of the invaraice
C e C,

C e C.

(25.145)

For details see T. Kugo and I. Ojima, Suppl. Prog. Theor. Phys. 66 , 1 (1979) and J. Ambjorn
and R.J. Hughes, Nucl. Phys. B 217 , 336 (1983).

People who have no weaknesses are terrible;


there is no way of taking advantage of them
Anatole France (18441924)

26
Weak Interactions
The decay of nuclei proceeds via , , emission or by fission. The first and last
processes are governed by nuclear forces, the third by electromagnetic forces. The
-emission has an entirely different origin. It is caused by weak interactions. Initially, this process posed the puzzle that the observed emerging particles, electrons
or positrons, did not come out with a definite energy equal to the mass difference
between initial and final nuclei. Instead, they had an energy distribution with a
maximal energy at the expected value. This led Nils Bohr to speculate that energy conservation could possibly be violated in quantum physics. However, in a
historic letter written in 1930 to colleagues at a meeting in T
ubingen, Wolfgang
Pauli suggested that the missing energy was carried off by a neutral particle, now
called neutrino. Its spin had to be 1/2, to preserve angular momentum.
The simplest nuclear particle which shows -decay is the neutron itself which
decays, after about 900 seconds, as follows
n p + e + e

(26.1)

where e denotes a right-handed version of Paulis particle which is nowadays called


the antineutrino associated with the electron.

26.0.1

Fermi Theory

Fermi was the first to describe -decay to lowest order in perturbation theory [1].
He suggested that the interaction is due to a local current-current interaction term
L = C p(x) n(x) e(x) (x) + c.c. .

(26.2)

Here we used the standard short notation for the fields: e (x) e(x), p (x) p(x),
etc. . This interaction looks like the electric interaction in Eq. (12.83), except that
the Coulomb potential describing the electric long-range interaction is replaced by
a short-range -function potential. Fermi also realized that of one only requires
Lorentz invariance, there could be, in principle, five similar interactions
L

= CS p(x)n(x) e(x)(x)
+ CP p(x)i5 n(x) e(x)i5 (x)
n(x) e(x) (x) + CA p(x) 5 n(x) e(x) 5 (x)
+ CV p(x)
+ 21 CT p(x) n(x) e(x) (x) + c.c. .
(26.3)
1362

1363
Fermi postulated time reversal invariance (which to our present knowledge is fulfilled
to a high degree of accuracy, so that its breakdown is observable only in very few
particularly sensitive processes), the coefficients Ci would have to be real. If i
denotes the five Dirac matrices, one may may write the interaction short as
L=

p(x)i n(x) e(x)i (x) + c.c. .

(26.4)

i=S,P,V,A,T

For studying the decay of an initial nucleus at rest |i i into a final nucleus |f i, one
has to evaluate the matrix elements
hf |
p(x)i n(x)|i i.

(26.5)

Due to the small energies involved, the final state will be almost at rest (if the nucleus
is big enough), and one may use the nonrelatistic limit for the matrix elements
between nucleons. Taking the Dirac matrices in Diracs original representations,
one can use the direct-product expressions (26.6),
D0 = 3 1,

D = i 2 , D5 = 1 1,

(26.6)

to rewrite the matrix elements of p(x)i n(x) as follows:


p (x)n(x),

S:

p(x)n(x)

P :

p(x)iD5 n(x)

V :
A:
T :

p (x) 2 1 n(x) 0,

p(x)D0 n(x) p (x)n(x),


p(x)Di n(x) p (x) 1 i n(x) 0,

p(x)D0 D5 n(x) p (x) 1 1 n(x) 0,


p(x)Di D5 n(x) p (x) 1 i n(x),

p(x) 0i n(x) p (x) 2 i n(x) 0,


p(x) ij n(x) ijk p (x) 1 k n(x).

(26.7)

The lines containing mixtures of upper and lower components are very small for
small momenta, and can be omitted. Thus one is left with only two types of nuclear
matrix elements:
hf |p (x)n(x)|i i h 1 i,
hf |p (x)  n(x)|i i h  i.

(26.8)

The S-matrix elements are then approxiamtely given by


S = 1 2i(Ef + Ee + E Ei )R,
with the R-matrix
R =

h
i
2m n
h 1 i CS u(pe )v(p ) + CV u(pe ) 0 v(p )
V
+h  i [ 21 CT u(pe )v(p ) + CA u(pe ) 5 v(p )]} ,

(26.9)

(26.10)

1364

26 Weak Interactions

where  in front of a Dirac spinor means 1 . The spin labels s3 for the electron and antineutrino have been omitted, for brevity. We have also normalized the
antineutrino states in the same way as photon states like
hp |p i = 2p0 (3) (p p )

(26.11)

to allow for a smooth limit m 0.


When calculating decay rates, we may sum over the antineutrino polarizations
which are hard to measure. The polarization sum yields
X

2m
v(p )v(p ) = p/ m ,
(26.12)
spin pol

and m is set equal to zero et the end. For unpolarized nuclei, we have
1
h i j i = ||2 ,
3

h i 1 i = 0.

(26.13)

If also the electrons are unpolarized, we obtain


X

spin pol

|R|2 =
+

1 h 2
C |h1i|2tr(/
p p/ e ) + |CV |2 |h1i|2tr(0 p/ 0 p/ e )
2me V 2 S
1
3

CA2 |hi|2 tr( i 5 p/ 5 i p/ e ) +

1
12

CT2 |hi|2 tr( i p/ 5 i p/ e )

+ 2CS CV |h1i|2 me tr(/


p 0 ) + 31 CS CV |hi|2 me tr( i p/ 0 p/ e ) , (26.14)
or more explicitly
X

spin pol

|R|

4E E
pe p
pe p
=
+ |CV |2 |h1i|2 1 +
CS2 |h1i|2 1
2
2me V
Ee E
Ee E


1 pe p
1 pe p
1
i| 1
+
+ CT2 |hi|2 1 +
3 Ee E
4
3 Ee E

2m
m
e
e
,
(26.15)
CA CT |hi|2
+ 2CS CV |h1i|2me
Ee
Ee
CA2 |h

The interference terms CS CV and CT CA show a strong threshold dependence proportional to 1/Ee . Since this is not seen experimentally one must have
CS CV 0,

CT CA 0.

(26.16)

One distinguishes now Gamow-Teller transitions for which |hi|2 6= 0, and Fermi
transitions for which |h1i|2 6= 0 (both without nuclar parity change). For transitions
in which initial and final nuclear spins are the same, only CS2 and CV2 contribute.
These are distinguishable by the electronic-antineutrino correlation in momentum
space. In the first case, electron and antineutrino come out preferably in opposite
direction (pe p < 0), in the second case in the same direction (pe p > 0). An
experiment done for the decay
Ne19 F19 + e + ,

(26.17)

1365
where the nuclear spin is 1/2 before and after the decay, showed that the latter is
true, thus eliminating CS .
Transitions which change the nuclear spin by one unit are sensitive to CA2 and
2
CT , with similar momentum correlations. The decay
He6 Li6 + e +

(26.18)

involves a change from nuclear spin 0 to 1. The electron and neutrino emerge
antiparallel, so that CT2 0. Thus only CV and CA are appreciable.
The antineutrinos which appear in these decay have been found to interact very
weakly with matter. Only many years later, in 1953, was it possible to detect them
directly via the inverse reaction
e + p n + e+

(26.19)

using a strong antineutrino source of a large nuclear reactor filled with a high density
of thermal neutrons.
After many precise measurements of angular correlations and electron and nuclear polarizations, the interaction was quite well determined to have the form
"
! #
h
i
gA
G

L = p 1
5 n e (1 5 ) + c.c. ,
gV
2
with the coupling constant

and the ratio

G = (1.14730 0.000641) 105 GeV2 ,


gA /gV

= 1.255 0.006.

(26.20)

(26.21)
(26.22)

The e part of the interaction violates space reflection symmetry in the maximal
possible way. Under parity, the left-helicity field (10 )(x) which possesses nonzero
components only in the first two entries is multiplied by
0 =

0 1
1 0

which moves the nonzero entries to the lower two components producing a righthelicity field (1+5)0 . This is not present in the Lagrangian density. This maximal
parity violation is responsible for the asymmetry of the decay electrons with respect
to the neutron spin as observed in Madame Wus experiment[2].
The ratio gA /gV can be measured in neutron decay via the up-down asymmetry
by analogy with Madame Wus experiment. The expectation hsn pe i is proportional
to
|CA |2 + Re CV CA (1 gA /gV ),
(26.23)
The deviation of the ratio gA /gV from unity can be explained by strong interactions. That this is possible was first suggested by the weak decay of the 1937
discovered muon (intially mistaken with the pion). The muon decays weakly as
follows
e + e + .

(26.24)

1366

26.0.2

26 Weak Interactions

Lepton Conservation

It was shown 1962 in an experiment in Brookhaven, that the second antineutrino


which emerges in the decay has to be distinguished from that emerging in -decay
associated with the muon. The two neutrinos are referred to as muon antineutrino
and electron antineutrino, respectively. With this assignment, the total number of
electrons plus its electron neutrinos and the total number of muons plus muon neutrinos are both separately conserved (antiparticles are counted negatively). These
laws are called lepton number conservation laws. The muon has no strong interaction
and its decay is described by the interaction Lagrangian density
ih
i
G h
L = (1 5 ) e (1 5 )e (x) + c.c. + . . . .
2

(26.25)

The value of G is given by


G = 1.16637(1) 105 GeV2 .

(26.26)

It is larger than G of the -decay in Eq. (26.21) by 2%.


The parameter (26.26) rather than (26.21) is defined as the Fermi constant for
two reasons. First, the experimental value of the lifetime of the is exteremely
accurately known:
= 2.197035(40) 106 s1 .
(26.27)
Second, the relation between G and is very simple [3]:
1

G2 m5
=
f
192 3

3 m2
m2e
(1
+
R.C.)
1
+
+ ... ,
m2
5 m2W
!

(26.28)

where R.C. denotes radiative corrections


25
2
m
R.C. =
2 1 +
log
3.7
2 4
3
me
#
 2 


4
m
2 m
+
log
2.0 log
+ C + ...

9
me
me




(26.29)

This is found by solving the renormalization group equation


(me me + () )R.C. = 0,

26.0.3

() =

2 2 1 3
+
+ ... .
3
2 2

(26.30)

Cabibbo Angle

In order to explain that the fact G is smaller than G , Cabibbo, in 1964, put
forward the hypothesis, that the reduction is due to the existence of a weak decay
of strange particles [4]. Studying the weak decays of
p + e + e
+ + e+ + e

(26.31)

1367
he observed that their weak coupling constant Gstrange was considerably smaller
than that of the neutrons. He noticed that he could explain the data quite well by
assuming that

G = cos c G
= sin c G

strange

(26.32)

where
c = 0.21

(26.33)

is the so-called Cabibbo angle.


Nowadays we believe that the weak interactions of all hadrons proceed via the
decay of the quarks inside their wave functions. The non-strange hadrons decay by
the elementary transition
d
d
u
u

u + e + e
u + +
d + e+ + c
d + + + .

(26.34)

The strange hadrons decay by the similar transition with the down quark d replaced by the strange quarks. Cabbibos hypothesis amounts to the ansatz for the
interaction Lagrangian density
h
i
G
L = u (15)(cos c d + sin c s) e (15 ) + c.c.
2
h
i
h
io h
i
G n
e (1 5 ) +h.c. (26.35)
= cos c u (15) d +sin c u (15 ) s
2

The first expression implies that the weak interactions couple only to an isospinrotated combination of down and strange quarks. Moreover, they appear only in a
special combination the vector and axial vector currents which we found previously
as conserved (or partially conserved) Noether currents of QCD. In fact we can write

u d =
2j(1+i2)/2 2j+ ,

2j5 (1+i2)/2 = 2j5+ ,


u 5 d =

2j(4+i5)/2 = 2jK
u s =
+,

(26.36)
2j5 (4+i5)/2 = 2j5K + ,
u 5d =
where the subscripts + , K + indicate the SU(3) quantum numbers which can be
created by the currents. In terms of these currents we can rewrite (26.35) as
n

L = G cos c j+ j5+ + sin c [jK + j5 K + ]

oh

e (1 5 ) +h.c. . (26.37)

1368

26.0.4

26 Weak Interactions

Cabibbo Mass Matrix

The Cabbibo theory can be phrased in a different way. The combination of quarks
cos c d + sin c s entering the interaction Lagrangian density (26.35) may be though
of as a result of a diagonalization of an nondiagonal mass matrix of the down and
strange quarks
(d, u ) M

d
u

= (d , u )

md mds
mds ms

d
u

(26.38)

This matrix is diagonalized by the physical down- and strange quarks


d
s

Vud Vus
Vus Vud

d
s

cos sin
sin
cos

d
s

(26.39)

with a Cabibbo angle solving the equation


tan =
where

md m

mds
=
,
ms m

mds

(26.40)

q
1
(md + ms ) + (md + ms )2 + 4m2ds .
(26.41)
2
The weak interactions couple only to the bare quark field d = cos d + sin s, i.e.,
the weak quark current entering the Lagrangian density (26.35) is

m
=

u (1 5 )d

26.0.5

(26.42)

Heavy Vector Bosons

The fact that the coupling constant of weak interactions carries a dimension of an
inverse square mass was historically a great obstacle in calculating higher order
correction to weak processes: The fact that the only successful field theory of fundamental interactions, quantum electrodynamics, has only a three-particle interaction
with the photon mediating the interactions, led to the hypothesis that weak interactions should be due to a similar coupling of a massive charged vector meson called
W . As we shall see later in the analysis of perturbation theory, this would generate
a weak coupling of the same type as in (26.20), but with G replaced by
G(q 2 ) =

e2W
2
q 2 MW

(26.43)

where q is the momentum carried off by the leptons and eW the coupling strength
of the heavy vector meson W to the leptons and nucleons. If MW is much larger
than 1 BeV, the q-dependence could be ignored and one would have
Gp

e2W
2
MW

(26.44)

1369
Since eW is dimensionless it was suggestive to assume it to be equal to the coupling strength of the photons. Then weak interactions would carry only one more
additional parameter, the mass of the vector meson. Using
e2 = 4

(26.45)

(26.46)

we find
MW =

4
90GeV.
G

A charged meson which has the desired properties and this was found near this
predicted mass in 1983. Its precise mass is
MW = 80.423 0.039 GeV.

(26.47)

With the usual weak decays carring off a non-zero charge via the lepton, the
question has been raised whether weak interactions could also lead to a decay into
e e , , e
e, or
. The first two could have easily escaped experimental detection,
the last two would be hard to find since they would hide under the similar but much
larger electromagnetic interaction. Such decays would come from couplings such as
h

L = G0 u (1 5 )u

inh

io

e e + e (1 5 ) + e .

(26.48)

The particle combinations u (1 5 )u, e (1 5 )e do not change the charge


of the decaying particles and are called neutral currents. They appear in the same
combination V A of vector and axial vector currents as in the interaction (26.37).
Experimental evidence for such neutral currents has been found in 1973 in form
of the purely leptonic reaction
+ e + e ,

(26.49)

as well as in interactions
u
+
d
u
+
+
d

u
,
d
u
.
d

(26.50)

Also a heavy neutral vector meson has been found which in analogy with W , mediates the neutral weak interactions. It is called Z boson and its mass is given by
[5]
MZ = 92.9 1.6GeV.

(26.51)

1370

26.1

26 Weak Interactions

Standard Model of Electroweak Interactions

All these interactions plus the electromagnetic interactions wave unified in one field
theory by Glashow, Weinberg, and Salam in the standard model of electromagnetic
and weak electroweak interactions. We shall formulate this theory here in a reduced
version in which the only leptons are electrons and electron neutrinos. Alle terms
in involving these leptons will have to be extended by similar terms involving the
other two families of leptons discovered so far:
!

(26.52)

m = 1777 MeV.

(26.53)

The charged lepton masses are well known:


me = 0.511 MeV,

m = 105.7 MeV,

The free lepton Lagrangian density for the electron part is


L = (
eL , eL ) /

eL
eL

+ eR i/
eR

(26.54)

It is invariant under an SU(2) group SUL (2) of the left-handed leptons and an
SU(1) group UR (1) of the right-handed leptons where neutrino R is absent. This
Lagrangian density is made locally SUL (2) UR (1) invariant by means of two types
of gauge fields. For the SUL (2) symmetry of the left-handed particles one uses a 2x2
gauge field
W = Wa

a
2

(26.55)

where Ta acts on the two SUL (2) isospin indices. Since these indices are mathematically equivalent to isospin, one speaks of a weak isospin. For the SUR (1) symmetry of
the right-handed electrons one uses a single vector field B . The ordinary derivatives
are replaced by the covariant derivatives
D = + igW
D = + ieB

(26.56)

After this one introducies the dynamics to the gauge fields via the Lagrangian density
1
1 2
LGF = tr [W W ] F
2
4

(26.57)

W = W W + ig[W W ]

(26.58)

where

is the SU(2)L covariant curl and


F = B B

(26.59)

1371

26.1 Standard Model of Electroweak Interactions

the usual abelian curl.


Introducing the charged W meson fields

1 
W = W1 iW2 ,
2

(26.60)

the new interactions are


Lint = g(
eL eL )

= g(
eLeL )

a
Wa

eL
eL

g eR B eR
2
!
!

W3 /2
W
/
2

eL

g eR B eR . (26.61)
eL
W+ / 2 W3 /2

Actually, there is some freedom in choosing the coupling to the vector meson B,
since we can always absorb an equal coupling to (ec , eL ) into the term involving the
field W 3 . Hence we may choose instead of g eR B eR the interaction

eL

g (
eL , eL , eR ) Y eL
eR

(26.62)

where the 3 3-matrix Y is called the weak hypercharge and has the form

Y =

yL
yL
yR

(26.63)

The parameter yL can be chosen at will since it can be absorbed into g, for instance
yL = 1.

(26.64)

Then the interaction reads



g 
Lint = W+ eL eL + c.c.
2

1
gW 3 g B eL eL
2
1
g
+ (gW3 + g B )
eL eL yR B eR eR .
2
2

(26.65)

Now we observe that the two fields W3 and B being both massless allow for an
arbitrary mixing. We introduce the physical photon field as
A = cos W B + sin W W3

(26.66)

and an orthogonal field which represents the neutral vector meson of weak interactions:
Z = +sinW B cos W W3 .

(26.67)

1372

26 Weak Interactions

The mixing angle W is called weak angle. It is chosen so that the photon couples to
the usual electromagnetic current which contains an equal amount of right-handed
and left-handed electrons

jem
= eL eL + eR eR .

(26.68)

This fixes

sin W

yR = 1,
g
g
= 2
,
cos W = 2
,
2
g +g
g + g 2
gg
.
e= 2
g + g 2

(26.69)

and the interaction becomes



g 
Lint = eL eL W+ + c.c. e e eA
(26.70)
2 2

i
eh
tan W (2
eR eR + eL eL + eL eL ) cot W eL eL eL eL Z .

2
Expressed in terms of the electromagnetic current

jem
= eL eL + eR eR ,

(26.71)

the left-handed weak currents written as vectors in the weak isospace


jL
jL3

1
eL
= (eL eL )
eL
2
!
3
eL

=
= (eL , eL )
eL
2

eL eL
eL eL

(26.72)

1
(eL eL eL eL ),
2
(26.73)

and the weak neutral current.


jnc = jL3 + sin2 W jem

(26.74)

we can write
Lint = e

jem
A

1
1
(j+ W+ + c.c.) +
j Z .

sin W
sin W nc


(26.75)

where
jnc = jL3 + sin2 W jem

(26.76)

is the neutral current. The weak angle describes the admixture of the ordinary
electric current to the neutral weak isospin current.

1373

26.1 Standard Model of Electroweak Interactions

Up to this point, the model had been constructed at the phenomenological level
by Glashow in 1961. He added to the Lagrangian mass terms for the W and the Z
meson
2 

M2
MW
W +2 + W 2 + Z Z 2
2
2

(26.77)

choosing the value for MW which gives the correct weak coupling constant not
knowing what to choose for MZ . In this way he obtained a model which parametrized
the known weak interactions and predicted the unknown neutral currents.
The weak interaction with quarks are incorporated by adding the Lagrangian
density
Lint =


g 
W+ jh + + c.c. + eA jem
2 2


q
3
2
em

2
2
g + g Z qL qL sin W j h ,
2

(26.78)

where
jh + = uL (dL cos c + sL i c )

2
1 
jemh =
d d + s s
u u
3
3

(26.79)

When the theory arrived at this stage it was realized that it contained an unpleasant
contradiction with experiments. The neutral vector meson Z was coupled to a quark
current
qL 3 qL = uL uL cos2 c dL dL

i sin c cos c dL sL + sL dc i sin2 c qL cL .

(26.80)

The third term changes the strangeness by one unit. Processes with a strangenesschanging neutral current are, however, strongly suppressed in nature. Otherwise the
particle decays
K
K 0 +

(26.81)

would have a much larger branching ratio than the experimentally observed 6
107 , 108, respectively (see Fig. 26.1). In order to achieve the desired suppression,
Glashow, Iliopolous, and Maiani [6] postulated the existence of a further quark,
called charmed quark and denoted by c, which was suppressed to form a weak
isodoublet with s in the same way as u does with d. They then considered the two
doublets
u
d

,
L

c
s

(26.82)
L

1374

26 Weak Interactions

Figure 26.1 Quark diagrams for K + and K 0 decays involving strangeness changing
neutral currents.

as two families of the same sort, both with weak hypercharge 1/3, and wrote the
left-handed neutral current as


u, d

u
d

2
+ (
c, s )
2
L

c
s

i sin2 W jem

(26.83)

Now the strangeness changing currents cancel exactly. The cancellation in the above
K decays goes as follows. Take, for instance, K + , where the compensating
diagrams are shown in Fig. 26.2. To the electromagnetic current, the charmed quark

Figure 26.2 Diagrams for the K 0 + decay with compensating strangenesschanging neutral currents.

contributes an additional

2
c c.
3
Unfortunately, the model as it stood still had an important handicap which prevented it from becoming a bona fide quantum field theory of electronical interactions.
Just as the four-fermi theory, it does allow for higher order perturbation correction.
The infinites appearing in higher order cannot be renormalized away. The reason is
the presence of the explicit mass terms. We shall see later that further symmetry is
necessary to allow for a renormalization. The action must be assumed to be invariant under the local SU(2)L U(1)R gauge group which was used to construct the
interactions. It is this gauge group which permits the renormalization of the theory.
The initial Lagrangian density contains only massless W and Z mesons To have
massive W and Z mesons it is necessary to generate these masses spontaneously,

1375

26.2 Quantum Oscillations

by an effect which was known from superconductivity as the Meissner effect. The
vector potential of magnetism in a superconductor becomes massive when entering
the superconductive phase. These aspects will be discussed later.

26.2

Quantum Oscillations

The standard model of weak interactions explains some interesting novel phenomena
in particle physics. The nature of a fundamental particle is sometimes not fixed but
can oscillate as a function of time. The first set of particles where this phenomenon
0 mesons.
was observed was in the decay of K 0 and K

26.2.1

Kaon Oscillations

In our discussion of the SU(3)-symmetry of strong interactions in Section 24.4 it was


0 have strangeness 1 and +1 respectively. It
observed that the mesons K 0 and K
is believed that strong interactions strictly conserve these quantum numbers. The
weak interactions described by the standard model, however, conserve the symmetry
0
CP. If we want to describe the weak decay of the strongly produced K 0 and K
mesons, we must decompose them into eigenstates of CP. These are the states

1 
0i ,
|K10 i |K 0 i + |K
2


1 
0i .
|K20 i |K 0 i |K
2

(26.84)

These two states have different decay channels. In its rest frame, the first state is
even under CP (recall Subsections 7.1.6 and 7.1.6), the second is odd. Hence the
first state can decay into two pions, while the second cannot. The lifetime of K10 is
0.8922 0.0020) 1010 sec producing almost exclusively + and 0 0 at a ratio
2:1.
The CP-odd state K20 must at least decay into three pions and has therefore a
much longer lifetime. This is why K10 and K20 are also called KS0 and KL0 .
Let mS , mL be their masses and S , L their decay rates. A beam of K 0
produced in a strong-interaction process will then evolve in its rest frame as an
oscillating superposition of KS and KL :
i
1 h
|K 0 i = e(imS +S /2) |KS i + e(imL +L /2) |KL i ,
2

(26.85)

where is the proper time. Since the two components decay with a different rate,
0 . These can be detected if the
the beam will contain oscillating admixtures of K
beam is directied towards a strongly interacting target, say a slab of copper. In it
0 . Thus, if one changes
the forward scattering amplitude for K 0 is smaller than for K
the thickness of the slab, the outcoming beam will have different admixtures of the
short-lived component KS0 . In this way one was able to regenerate KS0 from a beam
which had turned almost completely into KL0 by the decay into + or 0 0 .
The process was analyzed [8], and the experiment showed clearly the regenration
effect [9, 10].

1376

26 Weak Interactions

When experimentalists tried to set an upper limit to the decay of KL0 into two
pions they found that this decay was not completely forbidden [11]. This implies
that the weak decay of K 0 mesons contains a small term violating CP-symmetry.
The correct mixtures of short- and long-lived K 0 mesons must be slightly different
from (26.84):
h
i
1
0i ,
|KS0 i q
(1 + )|K 0 i + (1 )|K
2(1 + 2 )
h

1
0i ,
|K20 i q
(1 )|K 0 i (1 + )|K
2(1 + 2 )

(26.86)
(26.87)

where is of the order of 103 .


The parameter can be measured by studying a beam of K 0 as a function of time.
Due to the decay of the KS0 content, the beam will perform a damped oscillation
0 = sd and back. The content of the two components can be
from K 0 = d
s to K
0 e e arising from
studied by looking at the decay rates K 0 e e and K
the decay of the quarks s and s, respectively, in accordance with the Q = S rule.
This permits measuring directly from the asymmetry
NK 0 NK 0
N N+
Ne+ Ne
2Re
=
=
=
(3.320 0.074) 103 .
NK 0 + NK 0
N + N+
Ne+ + Ne
1 + ||2
(26.88)
0
0
Thanks to CP-violation, not only KS but also KL decays into two pions, and
this makes it possible to observe oscillations of the intensity in the decay of a K 0
beam into + . The relative intensity of decay rate changes with proper time
as follows:
I+ ( ) =

|h + |K 0 ( )i|
= eS +2|+ |e(S +L ) /2 cos(m + )+|+ |2eL ,
0
+

|h |KS ( )i|
(26.89)

where
+ |+ |ei+ =

|h + |KL0 ( )i|
,
|h + |KS0 ( )i|

m mL mS .

(26.90)

This damped oscillation was indeed observed experimentally [12], as shown in


Fig. 26.3.
The present best fits to the data yields the parameters
S
L
m
|+ |
+

=
=
=
=
=

1/S = (0.8926 0.0012) 108 sec ,


1/L = (5.17 0.04) 108 sec ,
(0.5333 0.0027) 1010h
/sec ,
(2.269 0.023) 103 ||
(44.3 0.8)o .

(26.91)
(26.92)
(26.93)
(26.94)
(26.95)

1377

26.2 Quantum Oscillations

(NK 0 NK
0 )/(NK 0 + NK
0)
(a)

(b)

[1010 sec]

[1010 sec]

mesons with respect to


Figure 26.3 Left-hand: Asymmetry of the number of
= ds
0
= sd mesons as a function of time [13]. The asymmetry is measured by the ratio
the K
0 e e . After a long time, only K 0 survives which
of the decays K 0 e e K
L
0 as a signal of CP-violation. Right-hand:
does not have an equal content of K 0 and K
Oscillation in decay rate into + of K 0 beam. Curve (a) shows the histogram of the
raw data. Curve (b) show the theoretical decay curve without oscillations, and the insert
the beste fit to the oscillations if the nonoscillating background is subtracted. Figure is
from [12].
K0

1
0.8

(NB0 NB 0 )/(NB0 + NB 0 )

0.6
0.4
0.2
0
-0.2
-0.4
-0.6
-0.8
-1
0

6
[1012 sec]

10

12

0 = bd
Figure 26.4 Asymmetry of the number of Bd0 = db mesons with respect to the B
d
mesons as a function of time [13]. Here the long- and short-lived combinations have almost
the same lifetime.

1378

26 Weak Interactions

d0 = bd the asymmetry measurement yields


For the neutral mesons Bd0 = db and B
(see Fig. 26.4) [14]:
NBd0 NBd0

N N+
Ne+ Ne
2Re Bd
=
=
(0.0035 0.0103 0.0015).
NBd0 + NBd0
N + N+
Ne+ + Ne
1 + |Bd |2
(26.96)
0
0

Similar asymmetries can be observed for Bs = sb and Bs = b


s.
All such neutral particle mixings can simply be understood by assuming that
the Lagrangian density for these particles contains a nondiagonal mass matrix
=

i
i
M11 11 M12 12
2
2

M=

i
i
M21 21 M22 22
2
2

(26.97)

This matrix must be hermitian, implying that

M21 = M12
,

21 = 12 .

(26.98)

Moreover, as a consequence of CPT invariance, the diagonal elements must satisfy


M11 = M22 = M,

11 = 22 = .

(26.99)

The most general mass matrix is therefore


i
i
M12 12
M
2
2
.
M=

i
i

M12 12 M
2
2

26.2.2

(26.100)

General Flavor Mixing

The Cabibbo theory was set up at a time when only the three quarks u, d, s were
known. After the discovery of the heavier three quarks it turned out that all quarks
can be grouped into three families of charge 2/3 and 1/3, respectively, and that
the weak current (26.42) has the generalization

d
Vud Vus Vub

(u, c, t) (1 5 ) Vcd Vcs Vcb s .


b
Vtd Vts Vtb

(26.101)

The 3 3-matrix V is called CKM-matrix after the intials of the authors Cabibbo,
Kobayashi, and Maskawa [4, 7]. It is often expressed in terms of four angles
12 , 13 , 23 , and 13 as
c12 c13
s12 c13
s13 ei13

i13
i13
c12 c23 s12 s23 s13 e
s23 c13
V = s12 c23 c12 s23 s13 e
, (26.102)
i13
i13
s12 s23 c12 c23 s13 e
c12 s23 s12 c23 s13 e
c23 c13

1379

Notes and References

where s12 sin12 , c12 cos12 . The angle 13 serves to explain the small CPviolation observed in the decay of KL0 discussed in Subsection 26.2.1. The mixing
matrix is being permanently updated as new experimental data are available. The
presently used numbers are [15]

0.974 to 0.9756 0.219 to 0.226 0.0025 to 0.0048

V = 0.219 to 0.226 0.9732 to 0.9748 0.038 to 0.044 ,


0.004 to 0.014
0.037 to 0.044 0.9990 to 0.9993

(26.103)

The most precise constraints on the size of the elements of the CKM matrix
are extracted from the low-energy s u and d u semileptonic transitions.
The determination is specialized part of phenomenological particle physics on which
excellent reviews are availabe from the Particle Data Group [7].

Notes and References


[1] E. Fermi, Nuovo Cimento 11, 1 (1934); Z. Phys. 88, 161 (1934).
[2] C.S. Wu, E. Ambler, R.W. Hayward, D.D. Hoppes, and R.P. Hudson, Phys. Rev. 105, 1413
(1957).
[3] W.J. Marciano, Phys. Rev. D 60, 093006 (1999).
[4] N. Cabibbo, Phys. Rev. Lett. 10, 531 (1963).
[5] F.J. Habert et al, Phys. Lett. B 46, 138 (1973); Nucl. Phys. B 73, (1974).
[6] S.L. Glashow, J. Iliopolous, L. Maiani, Phys. Rev. D 2, 1285 (1970).
[7] F.J. Gilman, K. Kleinknecht, B. Renk, Phys. Rev. D 66, 010001 (2002).
See also the conference summary by
F.J. Gilman, The Determination of the CKM Matrix, (hep-ph/0102345),
and the Review of
D. Goom et al., Particle Data Group (2002): Eur. J. of Phys. C 15, 110 (2000).
[8] M.L. Good, Phys. Rev. 106, 591 (1957); 110, 550 (1958).
[9] R.H. Good R.P. Matsen, F. Muller, O. Piccioni, W.M. Powell, H.S. White, W.B. Fowler,
and R.W. Birge, Phys. Rev. 124, 1223 (1961).
[10] V.L. Fitch, in Procedings of the XIII International Conference on High-Enegy Physics, ed.
by M. Alson-Garnjost (University of California Press, Berkeley, 1967), p. 63.
M.M. Nieto, Hyperfine Interact. 100, 193 (1996) (hep-ph/9509370).
[11] J.H. Chistensen, J.W. Cronin, V.L. Fitch, and R. Turlay, Phys. Rev. Lett. 13, 138 (1964).
[12] C. Geweniger, S. Gjesdal, G. Presser, P. Steffen, J. Steinberger, F. Vannucci, H. Wahl, F.
Eiself, H. Filthuth, K. Kleinknecht, V. L
uth, and G. Zech, Phys. Lett. B 48, 487 (1974).
[13] S. Gjesdal et. al., Phys. Lett. B 52, 113 (1974). The figure is from the Ph.D. thesis of V.
Luth.
[14] K.M. Ecklund et al., CLEO Collaboration,
[15] F.J. Gilman, Nucl. Instrum. Meth. A 462, 301 (2001) (hep-ph/0102345).

A little more moderation would be good.


Of course, my life hasnt exactly been one of moderation
Donald Trump (1946)

27
Nonabelian Gauge Theory of Strong Interactions
Elementary particles which interact strongly with each other are called hadrons.
According to their statistics, one distinguishes baryons and mesons. The most
prominent among these are the spin-1/2 particles in nuclear matter, protons and
neutrons. The forces between them arise, to lowest approximation, from the existence of a spin-0 meson called pion. The names baryons (heavy particles) and mesons
(middle-heavy particles) are due to the fact that the initially known baryons were
the heaviest elementary particles, the mesons were in mass between the baryons and
the electrons, which are the basic leptons.
Hadrons exhibit rich spectra, as we have seen in Chapters 24 and 25. Approximately, these spectra were explained by quark models. Just as nuclei are composed
of protons and neutrons, baryons are composed of three quarks, mesons of quarks
and antiquarks.

27.0.3

Local Color Symmetry

In spite of its success, the initial quark model exhibited several fundamental inconsistencies. Most importantly it was not compatible with the spin-statistics theorem
derived in Section 7.8. For any reasonable potential between quarks, the ground
state orbital wave function should always be without zeros implying vanishing relative angular momenta for each pair of quarks. Any higher angular momentum
would have at least one zero in the wave function which would increase the gradient
and thus the kinetic energy of the Schrodinger field via the centrifugal barrier. The
orbital wave function of the ground state of three quarks, the proton, must therefore
be symmetric under the exchange of two quarks. On the other hand, the SU(6) wave
function involving internal SU(3) and spin has the Yang tableau
. But this
implies that the nucleon wave function is completely symmetric under exchange of
all positions, all SU(3), and all spin variables. This seems to contradict the fact
that quarks have spin 1/2 so that, by the spin-statistics theorem, they should be
fermions and therefore have a completely antisymmetric wave function.
1380

1381
To remedy this contradiction, Han and Nambu suggested that quarks should be
triplet under a further SU(3) group of transformations now called color SU(3) [1].
Color appears as a further label to the quark fields:
u(x)
d(x)
s(x)
c(x)
t(x)
b(x)

q(x) =

(27.1)

which may thus be written as


u (x)
d (x)
s (x)
c (x)
t (x)
b (x)

q (x) =

(27.2)

with a label = 1, 2, 3 specifying the three colors. By postulating that all hadron
states are completely antisymmetric in color, i.e. color singlets, the contradictions
disappears and the spin-statistic relation is again valid.
After this somewhat artificial postulate the question arose how nature manages
to enforce the color antisymmetry, i.e., how it prevents color non-singlet states to
be excited. The answer suggested by Fritzsch and Gell-Mann was that color SU(3)
was a local gauge symmetry of hadronic physics. The action had to be invariant
under arbitrary nonabelian SU(3)-transformations of the quark fields
a (x)a /2

q(x) ei

q(x),

(27.3)

where the matrices a now act on the three color labels of the quark field q(x) in
SU(3). The tripling of the quarks saved not only the validity of the spin-statistics
relation for quarks. It also led to the correct rate of the particle decay
0
which is observed experimentally at a rate = 8.4 1017 s. In addition, it gave
the correct total interaction cross section observed in e+ e collisions, which require
that each quark occur in three color versions.
Once it was postulated that the quark action be invariant under local symmetry
transformations, it became necessary to introduce a gauge field to maintain the
invariance of the gradient term in the action. This led to the quark Lagrangian
L(x) = q(x)i/
D q(x) M q(x)q(x)

(27.4)

1382

27 Nonabelian Gauge Theory of Strong Interactions

where q(x) carries flavor indices distinguishing the quarks u, d, c, s, . . . and three
color indices. The gradient term contains the covariant derivative
D
/ = D = ( + igA ) ,

(27.5)

where A is a 3 3 matrix in color space to guarantee that D transforms under


local color SU(3) transformations in the same way as the field (x) itself:
D q(x) U(x)D q(x) = eia (x)

a /2

D q(x).

(27.6)

The covariance property is a bit harder to show than in electromagnetism. The


derivative of the field goes over into:
q(x) U(x)q(x) = U(x)[U 1 (x) U(x)q(x)]
= U(x) q(x) + U(x)[U 1 (x) U(x)]q(x).

(27.7)

In order to remove the second term, the gauge field A has to transform like
1
[ U(x)]U 1 (x)
ig
1
= U(x)A (x)U 1 (x) + U(x) U 1 (x).
ig

A (x) U(x)A (x)U 1 (x)

(27.8)

Then
h

[ + igA (x)] q(x) U(x) + iqA (x) + U (x) U(x)


= U(x) [ + igA (x)] q(x),

(27.9)

which is the desired covariant transformation law.

27.0.4

Gluon Action

After the introduction of such a gauge field, an interaction has to be found describing
the dynamics of this gauge field itself. If color is never observed, the action of the
gauge field should be locally SU(3)-invariant as well. The only Lagrangian which
has this property and contains, at most, first derivatives in A is given by
1
Lgluon = tr (F F )
2

(27.10)

F = A A + ig[A , A ],

(27.11)

where

is the field tensor, which is the nonabelian version of the covariant curl (4.335) of the
vector potential A . Just as A , the field tensor is a 3 3 matrix in color space. It
is easy to verify that the matrix F transforms under SU(3) covariantly as follows
F (x) U(x)F (x)U(x).

(27.12)

1383

27.1 Quantization in the Coulomb gauge

The theory is described by the Lagrangian


1
L(x) = q(x)i [ + igA (x)] q(x) tr [F (x)F (x)] .
2

(27.13)

This is a complete nonabelian analog of the gauge field Lagrangian (12.1)(12.3)


in quantum electrodynamics. For this reason it has been given the similar name
quantum chromodynamics (QCD).
Instead of 3 3 matrices A , F , one can also use an octet of vector and tensor
fields defined by
A =

a a
A ,
2

F =

a a
F .
2

(27.14)

Thus we can write the Lagrangian of QCD as


L = q(x)i

a
1 a
a
+ ig Aa (x) q(x) F
(x)F
(x),
2
4

"

(27.15)

where
a
F
(x) = Aa (x) Aa (x) gf abc Ab (x)Ac (x).

(27.16)

In terms of the eight components Aa (x), the Lagrangian (27.10) reads


1 a a
F .
LFgluon = F
4

(27.17)

There exists an equivalent way of expressing it in terms of two independent fields


a
Aa (x) and F
, often used by Schwinger:

1 a a 1 a 
Lgluon = F
F
F Aa Aa .
4
2

(27.18)

The spacetime integral over this is the canonical action corresponding to the mechanical action of a free particle [recall (1.14)]
A=

p2
dt + pq .
2
!

(27.19)

In the last twenty years, evidence has accumulated that the gauge field Aa is indeed capable of describing the forces which bind together the quarks inside hadrons.
The field quanta carried by Aa are called gluons.

27.1

Quantization in the Coulomb gauge

Let us quantize the theory for the simplest nonabelian symmetry SU(2), where the
gluons Aa (a = 1, 2, 3) are vectors in color space. We shall follow the description of

1384

27 Nonabelian Gauge Theory of Strong Interactions

the theory in Ref. [?]. In SU(2)-symmetry, the structure constants f abc reduce to
abc and the covariant curl (27.16) can be written in vector notation as
F (x) = A (x) A (x) gA(x) A (x).

(27.20)

The Lagrangian (27.18) becomes [?]


1
1
LFgluon = F F F ( A A gA A ) .
4
2

(27.21)

This is invariant under the nonabelian gauge transformations (27.8) whose infinitesi . If u
mal form is a rotation in isospace by a small angle around the direction n
n
denotes the associated infinitesimal rotation vector, the gauge transformations take
the form
1
A (x) Au (x) = A (x) + u(x) A (x) + u(x)
g
F Fu = F + u F .

(27.22)

The first Euler-Lagrange equation


LFgluon
=0
F a

(27.23)

a
reproduces the relation (27.20) between the curl and the auxiliary tensor field F
.
The second Euler-Lagrange equation

LFgluon
L
=
( Aa )
Aa

(27.24)

yields the field equation


D F F gA F = 0.

(27.25)

The combination of Eqs. (27.20) and (27.25) coincides with the field equation that
would be obtained from the second-order formulation with the Lagrangian Lgluon of
Eq. (27.17) expressed directly in terms of A and A via Eq. (27.11).
In the present first-order formulation, one is given an initial configuration of
fields Ai and F0i at some time t. From these one determines the fields at any later
time by solving the first-order field equations
0 Ai = F0i + (i gAi ) A0 ,
0 F0i = (j gAj ) Fji + gA0 F0i .

(27.26)
(27.27)

As in the abelian case of Maxwell electromagnetism [recall Eq. (7.327)], the field A0
is not a dynamical variable since the canonical momenta
L/ (0 A ) = F0 = F0

(27.28)

1385

27.1 Quantization in the Coulomb gauge

vanish for = 0. There are only three independent field momenta. The Euler
equation for A0 is not an equation of motion but a constraint equation analogous
to the Coulomb law (7.329):
(k gAk ) Fk0 = 0.

(27.29)

As in the Abelian case, it tells us that not all of the conjugate momenta F0k are
independent.
Note that in the first-order formulation, the field equation obtained from varying
Fij ,
Fij = i Aj j Ai gAi Aj ,

(27.30)

is also a constraint equation which allows us to calculate Fij for given Ai at the
same time. In addition we see from Eq. (27.30) that not all field components Ak
can be treated as independent.
In order to remove the redundancy we choose the Coulomb gauge:
k Ak = 0.

(27.31)

This is always possible because of the gauge invariance of the second kind of the
Lagrangian density [recall (4.252)]. The gauge (27.31) implies that the vector field
Ai must be transverse. Therefore, the longitudinal components FL0i of the canonical
momentum F0i are not independent, but depend on the other degrees of freedom
through the constraint (27.29). The splitting of F0i into longitudinal FL0i and transverse parts FT0i is defined by the equations
F0i = FT0i + FL0i ,

i F0i = i FL0i ,

ijk j FL0k = 0.

(27.32)

Our task is now to express A0 and FL0i in terms of the independent fields and
construct the Hamiltonian. As usual we identify the transverse components FT0i as
the electric field strengths Ei . By rewriting the longitudinal components FL0i as a
gradient of a scalar isovector f,
FL0i = i f,

(27.33)

i F0i = 2 f.

(27.34)

we have

The independent variables are the transverse vector fields Ai and its canonically
conjugate field momentum Ei .
Inserting (27.33) into the constraint (27.29) we obtain the differential equation
for f:


2 + gAk k f = gAi Ei .

(27.35)

1386

27 Nonabelian Gauge Theory of Strong Interactions

This equation can be solved formally by introducing a Green function D ab (x, x ; A),
defined as a solution of the inhomogeneous differential equation


2 ab + gacb Ack k D bd (x, x ; A) = ad (3) (x x ).

(27.36)

With the help of this Green function, Eq. (27.35) can be solved for the components
of f by the integral
a

f (x, t) = g

d3 x D ab (x, x ; A)bcdAck (x , t)Ekd (x , t).

(27.37)

This may be abbreviated as


f = g D Ak Ek .

(27.38)

The solution D bd (x, x ; A) of Eq. (27.36) cannot be found explicitly, but only via
a perturbation expansion. The lowest approximation coincides with ab times the
Green function 1/2 of electrostatics. A first iteration yields additional terms up
to the order g:
D ab (x, x ; A) =

ab
+g
4|x x |

d3 x

1
1
acb Ack k
+ . . . . (27.39)

4|x x |
4|x x |

This first-order approximation is easily verified by inserting it into Eq. (27.36).


A similar equation for A0 is obtained by taking the divergence of Eq. (27.26)
and using (27.31) and (27.32) to find


2 + gAi i A0 = 2 f,

(27.40)

which can be solved using once more D ab (x, x ; A), since the operator in brackets is
the same as in Eq. (27.35):
Aa0 (x, t)

d3 x D ab (x, x ; A)2 f b (x , t),

(27.41)

A0 = D2 f.

(27.42)

or in short form:

We can now construct the Hamiltonian density H by the Legendre transformation


of the Lagrangian density (27.18):
H = Ei

Ai
LFgluon .
t

(27.43)

From (27.26), (27.32), (27.33), and (27.42) we find that


h

t Ai = Ei i (i + gAi )D 2 f.

(27.44)

1387

27.1 Quantization in the Coulomb gauge

Because of (27.36), the operator in brackets acting on f is explicitly transverse.


Combining further (27.44) and (27.35) or (27.37), we obtain that
Z

d3 x Ei t Ai =

d3 x E2i + g(Ei Ai ) D 2 f
3

dx

E2i

f f =

d3 x E2i + (i f)2 . (27.45)

If we now rewrite the Lagrangian density as


1
1
F F F ( A + gA A )
4
2
1 2
1
1
=
(F0k B2i ) = (Ek k f)2 B2k ,
2
2
2

LFgluon =

(27.46)

and identify the magnetic field strength as


1
Bi = ijk Fjk .
2

(27.47)

we find the Hamiltonian


1
H=
2

d3 x E2i + B2i + (i f)2 .

(27.48)

The last term is like the familiar instantaneous Coulomb interaction discussed in
QED in Section 12.3 [see the last term in Eq. (12.79)].
We now express the generating functional WC [j] in the Coulomb gauge in terms
of the independent coordinates and momenta, Ai and Ei as the functional integral
WC [j] =

DETi DATi

 Z

exp i

k 1 E2 1 B2 A jk
d x Ek A
2 k 2 k




(27.49)

where the superscript T stands for Transverse, and f is a function of ETi and ATi
as in Eq. (27.37). Note that the source term here has a negative sign, so that the
covariant version below in (27.63) will read A j .
The transverse field ETi is somewhat awkward to handle. Therefore we introduce
an initially dummy variable EL by
Z

DETi

DETi DEL (3) [EL ],

(27.50)

where (3) [EL ] x (3) (EL (x)) is the -functional in four-dimensional spacetime.
Then we define three independent transverse and longitudinal components Ei by
Q

ETi

1
= ij i 2 j Ei ,

E L i

1
j E j ,
2

(27.51)

in terms of which we can rewrite the measure (27.50) as


Z

DETi =

DEi J[j Ej ],

(27.52)

1388

27 Nonabelian Gauge Theory of Strong Interactions

where J is a field-independent, and thus irrelevant, Jacobian of the transformation


from the three Ei to ETi , EL , and
DEi

3 Y
3
YY

x i=1 a=1

dEia (x).

The same decomposition is applied to the gauge fields ATi , so that the generating
functional has the functional integral representation
WC [j] = const
 Z

exp i

DEi DAi [k Ek ][k Ak ]

k 1 E2 1 B2 1 (k f)2 Ak jk
d x Ek A
2 k 2 k 2
4



. (27.53)

From this expression we can derive the Feynman diagram rules in the Coulomb
gauge. As in the abelian case, these are not covariant, and the Lorentz covariance
of the emerging S-matrix would is not obvious. The Coulomb gauge is only useful
for deriving the generating functional from the canonical formalism.
In order to find the Feynman rules for the covariant and gauge invariant S-matrix,
one should start out with a covariant-looking form of the generating functional
instead of (27.53). There, f is a function of E and A given by Eq. (27.37). It is
possible to introduce f as a variable of integration and fix its value to satisfy (27.37)
by a -functional
Z

Df [f gD Ak Ek ] = 1.

(27.54)

If the generating functional (27.53) is multiplied by this expression, it obviously


remains unchanged. Let Det M be the Jacobian of the transformation from f to
(2 + gAi i )f. The factor in front of f is a functional matrix in isospin space:


M ab (x, x ) =

2 ab + gabc Aci (y)i (4) (x x )


h

= 2 ab (3) (x x ) + gabc G(x, x )Aci (x )i (x0 x0 ), (27.55)


where 2 G(x, y) = (3) (x x ). With the help of Eq. (27.35), we can now rewrite
(27.54) as
Z

Df [f gD Ak Ek ] = Det M

Df [(2 + gAi i )f gAi Ei ], (27.56)

and the generating functional Eq. (27.53) becomes


WC [j] = Det M

DAi DEi Df [i Ai ] [i Ei ][(2 + gAi i )f gAi Ei ]

 Z 

exp i

Ek Ak

i2
1h 2
fk + B2k + (k f) ji Ai d4 x .
2


(27.57)

Next we change variables from Ei to F0i given by


F0i = Ei i f,

(27.58)

1389

27.1 Quantization in the Coulomb gauge

using (27.32) and (27.33). Then we rewrite the measure of integration in (27.57) as
DEi Df[i Ei ][(2 + gAi i )f gAi Ei ]
= DF0i Df[i F0i + 2 f][2 f gAi F0i ]
= DF0i Df[i F0i + gAi F0i ][2 f gAi F0i ].

(27.59)

Now we perform the integration over Df using the last -function in (27.59). The
Jacobian is just Det 2 , i.e., an irrelevant infinite constant which shall absorb into
the definition of M. Thus we obtain
WC [j] = Det M

Z

exp i

DAi DF0i [i Ai ] [i F0i + gAi F0i ]

(27.60)

1
1
d x F0i 0 Ai F20i (i Aj j Aj +gAi Aj )2 ji Ai
2
4
4



The exponent has been found by setting in the exponent of (27.57)


E2k + (k f)2 = (Ek k f)2 = F20i ,
and omitting the mixed term since it vanishes upon integration over x, due to the
transversality of Ek .
Next we express the -functional as a functional integral, using A0 as a dummy
variable:
[i F0i + gAi F0i ] =

YZ
x

dA0
exp {iA0 (0 F0i gAi F0i )}
2
 Z

DA0 exp i

d4 x F0i (gA0 Ai i A0 ) . (27.61)

Finally, we write the term 41 (i Aj j Ai + gAi Aj )2 in the exponent of (27.60)


as
Z

 

DFij exp i

1
1
Fij Fij Fij (i Aj j Ai + gAi Aj )
4
2



(27.62)

which is a standard Gaussian integral. Inserting (27.62) and (27.61) into (27.60),
and restricting j0 to zero, we obtain
WC [j] = Det M

DA DF [i Ai ]

 Z

exp i

1
1
d4 x F0i F0i + Fij Fij
2
4


1
Fij (i Aj j Ai + gAi Aj ) + F0i (0 Ai i A0 + gA0 Ai )
2 Z
 Z


= Det M

DA DF [i Ai ] exp i

d4 x[LFgluon + j A ] .



(27.63)

Were it not for the factor Det M, this would directly define covariant Feynman
rules. In order to calculate its effect, it is useful to reexpress it in terms of an effective

1390

27 Nonabelian Gauge Theory of Strong Interactions

Lagrangian density. Recalling the explicit form of the functional matrix (27.55), we
factorize it as

Det M = Det 2 Det [ 1 + M],

(27.64)

where
= gabc G(x, y)Ac (y) i (x0 y0 ),
M
i

1 = ab 4 (x y).

(27.65)

The determinant Det 2 is again an irrelevant infinite constant. The second factor
is expanded as
= exp Tr log( 1 + M
)
Det ( 1+ M)
(27.66)
Z

n1
h
i
X (1)
(x1 , x2 )M(x
2 , x3 ) . . . M
(xn , x1 ) .
= exp
d4 x1 . . . d4 xn tr M
n
n=0
The trace symbol tr runs only over isospin indices. Inserting (27.65), this becomes
"

= exp (0)
Det ( 1+ M)

gn
n=0 n

d x1 . . . d xn

dt tr [T Ai1 (x1 , t)i1 G(x1 , x2 )

T Ai2 (x2 , t)i2 G(x2 , x3 ) T Ain (xn , t)in G(xn , x1 )



,(27.67)

where (T a )bc = abc and Tr includes the trace over isospin indices.
Since (27.67) is a power series in the exponent, it is an effective correction in
each order to the Feynman rules obtained from LFgluon alone.

27.2

Direct Functional Quantization of Gauge Fields

Equation (27.63) can be further simplified. We can perform the functional integraa
tion over F
and obtain
W [j] = Det M

 Z

DA [i Ai (x)] exp i

d4 x[Lgluon (x) + j (x) A (x)] , (27.68)

where L(x) is the second-order Lagrangian density (27.17). Except for the factor Det M [i Ai (x)], this expression looks the same as for a standard scalar field
theory:
W [j]

 Z

D exp i

d4 x[L(x) + j(x)(x)] .

(27.69)

For the abelian gauge theory QED we have shown in Section 13.13 how to derive
such a factor following an intuitive argument due to Faddeev and Popov. Here we
may do the same for the nonabelian case. Recall how the argument went in the
abelian case. There we expressed the quadratic part of the action in the bilocal
form
Z
Z
1
1
A0 = d4 x ( A A )2 = d4 xd4 x A (x)D (x, x )A (y), (27.70)
4
2

1391

27.2 Direct Functional Quantization of Gauge Fields

with a functional matrix




D (x, x ) = 2 g 4 (x x ).
This cannot be inverted, since it involves only the transverse components of A ,
while ignoring the longitudinal components. Hence the Euclidean version of the
functional integral of Eq. (27.68) contains no Gaussian exponential involving the
longitudinal components of A , and therefore the functional integral over these
diverges. In the abelian case this was a consequence of the invariance of the exponent
in (13.279) under the gauge transformation (13.277). Here it is a consequence of the
invariance of the exponent of (27.68) under the gauge transformation
A Ag .

(27.71)

The right-hand side emerges as a result of applying the element g of the gauge group
G to the field A :
Ag

"

1
L = U(g) A L + U 1 (g) U(g) U 1 (g).
ig

(27.72)

The exponent in the functional integral (27.72) is constant on the orbits of the gauge
group, which are formed by all Ag for fixed A , while d g runs ocer the entire group
G. This causes a divergence of the functional integral for the generating functional
W [j]. The amplitude W [j = 0] is therefore proportional to the volume of orbits
Q
x dg(x), and this factor should be extracted before defining W [j = 0]. Thus the
functional integral should not be performed over all fluctuations of the gauge fields,
but only over the different orbits of A defined by the symmetry transformations of
the gauge group.
To implement this idea, we choose a hypersurface in the manifold of all fields
which intersects each orbit only once. Let
fa (A ) = 0,

a = 1, 2, . . . N

(27.73)

describe such hypersurface, where N is the dimension of the group. We shall assume
that the equation
fa (Ag ) = 0
has a unique solution g for any given field A . We are going to integrate over all
different hypersurfaces of this kind, instead of integrating over the manifold of all
fields. The conditions fa (A ) = 0 define a particular gauge, the Coulomb gauge
fa (A ) = i Aai being just a particular example.
Before proceeding further, let us recall briefly some simple facts about group
representations. For any two group elements g, g G, the product gg is also G,
and their representations satisfy the same mutiplication law
U(g)U(g ) = U(gg ).

1392

27 Nonabelian Gauge Theory of Strong Interactions

The invariant Hurwitz measure over the group G is invariant under this operation,
so that
dg = d(gg ).

(27.74)

If we parametrize U(g) in the neighborhood of the identity as


U(g) = 1 + iu L + O(u2 ),
then in the neighborhood of the identity we may choose
dg =

dua ,

g 1.

(27.75)

[fa (Ag (x))] = 1,

(27.76)

Let us define the functional f [A ] by


f [A ]
where

Z
Z

Dg

Y
a

Dg

Y Z

dg(x) .

(27.77)

Without the gauge invariance, the vacuum-to-vacuum amplitude would be given by


Z

 Z

DA exp i

d x Lgluon (x) .

(27.78)

We now insert the left-hand side of Eq. (27.76) into the funtional integral (27.78)
without changing the result:
Z

DgDAf [A ]

[fa (Ag (x))] exp

 Z

d x Lgluon (x) .

(27.79)

Now we perform a gauge transformation A (x) Ag (x) defined by (27.72) on


A (x). Under this, the action in (27.78) is invariant. We may also convince ourselves
that the functional f [A ] defined by (27.76) is gauge invariant:
g
1
f [A ]

Z Y
x

Z Y

dg (x)

[fa (Agg
(x))]

x,a

d(g(x)g (x))

x,a

Z Y
x

dg (x)

Y
x,a

[fa (Agg
(x))]

[fa (Ag (x))],

so that indeed
f [Ag ] = f [A ].

(27.80)

1393

27.2 Direct Functional Quantization of Gauge Fields

Hence we may write Eq. (27.79) also as


Z

Dg

DA f [A ]

 Z

[fa (A )] exp i

x,a

d x Lgluon (x) ,

and we find that the integrand of the group integration is independent of g(x). This
R
was the observation of Fadeev and Popov, who saw that the functional integral Dg
is simply an infinite factor independent of the fields. It can therefore be dropped
from the amplitude, so that the generating functional W [j] may be defined as
Wf [j] =

DA f [A ]

 Z

[f (A)] exp i

d x[L(x) + j (x) A (x)] . (27.81)

Faddeev and Popov also gave the canonical derivation of Eq. (27.63) as discussed in
the preceeding section.
Before demonstrating the eqivalence of Eqs. (27.68) and (27.81), we shall comQ
pute f [A ]. Since the factor f [A ] is multipled by a [f a (A (x))] in Eq. (27.81),
it suffices to compute f [A ] only for vector fields A which satisfy Eq. (27.73).
Let us define the functional matrix Mf by
f a (Ag (x)) = f a (A (x)) +

d4 y

X
b

[Mf (x, y)]ab ub (y) + O(u2 ).

(27.82)

Then we find from Eq. (27.76) that


1
f [Af ] =

Z Yn

dua (x)[f a (Ag (x))] =

x,a

Z Y
x,a

{dua (x)[Mf u]} .

The integral receives a contribution only from A-fields satisfying fa (A ) = 0, so


that
f [A ] = Det Mf = exp {Tr log Mf } .

(27.83)

The hypersurface equation fa = 0 is just the gauge condition, and for the
Coulomb gauge adopted in the preceeding section, we had
f a (A ) = i Aai = 0,
which becomes after an infinitesimal gauge transformation
f a (Ag ) = i Aai +


1  2 ab
gabc Aci i ub (x) + O(u2 ).
g

(27.84)

From this we identify


1
1
[Mf (x, x )]ab 2 ab g abc 2 Aci i 4 (x x ) [M(x, x )]ab ,
g

(27.85)

which shows that Eq. (27.68) is indeed a special case of Eq. (27.81) for f a = i Aai .

1394

27 Nonabelian Gauge Theory of Strong Interactions

Being in the possession of Eq. (27.81), we are free to use many different gauges
other than the Coulomb gauge. If we choose, for example, the manifestly covariant
Landau gauge condition A = 0, then Eq. (27.82) takes the form
1
Ag (x) = A (x) + [ 2 u + g (A u)] + O(u2 )
g
so that Mf is given by
1
[ML (x, x )]ab = ( 2 ab gabc Ac ) 4 (x x )
g

(27.86)

when A is restricted to A = 0. Removing the trivial factor (1/g) 2 from


ML (x, x ), we have
n

L) ,
L Det ML exp Tr ln(1 + M

(27.87)

where
Lab (x, x ) = gabc
M

DF (x z)Ac (z)

4
(z y)d4z,
z

(27.88)

and DF (xz) is the usual Feynman propagator satisfying 2 DF (xy) = 4 (xy).


More explicitly we can write
(

L = exp

h
(g)n
d4 x1 . . . d4 xn tr DF (x1 x2 )T A (x2 ) DF (x2 x3 )
n
n=0
. . . DF (xn x1 )T A (x1 )]} .
(27.89)
Z

The need to have the extra factor f [A ]


We can write Eq. (27.81) as
Wf [j] =

DA

[f a (A (x)] was first noted by Feynman.

 

[f (A (x))] exp i Seff +

where
Aeff =

d x j (x) A (x)

d4 x Lgluon (x) iTr ln ML .



(27.90)

(27.91)

At this point it is useful to observe that the additional term iTr ln ML in the
effective action may be thought of as arising from loops generated by a fictitious
isotriplet of complex scalar fields c obeying Fermi statistics, whose presence and
interactions can be described by the following action
Ac =

d4 x c (x) A (x) c(x)

d4 xd4 x

ca (x) [ML (x.x )]ab cb (x ).

a,b

(27.92)

1395

27.2 Direct Functional Quantization of Gauge Fields

With this, Eq. (27.90) may be written as


WL [j] =

DA

[f (A (x))]

 

Dc Dc exp i A + Ac +

d x j A (x)

(27.93)

Since c and c appear quadratically in the action (27.92), the path integral
can trivially be done with the result
Z



Dc Dc exp(iAc ) Det ML = exp {Tr ln ML } ,

Dc Dc
(27.94)

and we can expand


h

L)
exp {Tr ln ML } exp Tr ln(1 + M


L 1 Tr M
2 + . . . ()n 1 Tr M
n + . . . . (27.95)
= exp Tr M
L
L
2
n
The terms in the exponent may arise from loop diagrams of the Fermi fields c.
Let us now specify the Feynman rules resulting from WL [j] of Eq. (27.93) following the rules of Chapter 13. The gauge boson propagator is determined from the
free-field functional Z
R 4
1

2
W 0 [j] =
DA [ A ]ei d x{ 4 ( A A ) + j (x)A (x)} .
(27.96)

A convenient way of evaluating Eq. (27.96) is to express the -functional as




Z
i
(27.97)
d4 x[ A (x)]2 ,
[ A (x)] lim exp
0
2
Q
where we have discarded an irrelevant infinite constant x 2. Then we arrive
at the generating functional of free vector boson Green functions:




Z
Z
1
i
4
2

0
d xA (x) g + 1
A (x)
WL [j] = lim DA exp
0
2


Z
d4 x j (x) A (x)

+i

i
= lim exp
d4 xd4 x j (x) DF (x x ; ) j (x ) ,
0
2

where DF (x x ; ) is the free vector boson propagator




DF (x

x ; ) =

(27.98)

k k
1
d4 k
g + 2 (1 ) . (27.99)
exp {ik (x x )} 2
4
(2)
k + i
k
"

In the limit 0, this becomes


DF (x

x) =

d4 k
k k
1
g + 2
exp {ik (x x )} 2
4
(2)
k + i
k

, (27.100)

which is transverse in spacetime. The rest of the Feynman rules can be derived as
usual. They are recorded in Figs. 27.1 and 27.2. In addition, the following rules
must be kept in mind: the ghost-ghost-vector vertex mey carry a dot, where a dot
indicates that a ghost line is differentiated. Note that a ghost line cannot be dotted
at both ends and that a ghost loop carries an extra minus sign.

1396

27.3

27 Nonabelian Gauge Theory of Strong Interactions

Equivalence of Landau and Coulomb gauges

Formally, the S-matrix computed in the Landau gauge is the same as that computed in the Coulomb gauge [4]. An element of the unrenormalized S-matrix is
obtained from the corresponding Green functions by removing single particle propagators corresponding to external lines, taking the Fourier transform of the resulting
amputated Green function and placing external momenta on the mass shell. The
demonstration to be presented is basically correct, except that the S-matrix of a
gauge theory is plagued by infrared divergences and may not even be defined. In
fact, this may be the reason why massless Yang-Mills particles are not seen in nature. The point of presenting this demonstration is at this point purely pedagogical.
The technique will be useful in the discussion of spontaneously broken versions of
gauge theories.
Let us first establish a relation between WC [j] and WL [j]. From Eq. (27.68) we
obtain for the first:
WC [j] =

DA C [A ][i Ai ] exp iA[A ] + i

d4 x j A ,

(27.101)

where C = Det M satisfies


C [A ]

Dg[ Ag (x)] = 1.

(27.102)

Figure 27.1 Propagators in the Yang-Mills theory. Wavy lines are vector mesons. Dashed
lines are scalar ghosts.

Figure 27.2 Vertices in the Yang-Mills theory. Note that the ghost with index c is
pictures by a dotted line.

1397

27.3 Equivalence of Landau and Coulomb gauges

Inserting the left-hand side of Eq. (27.102) in the integrand of the functional integration in Eq. (27.101), we write
Z

WC [j] =

Dg

DA C [A ]C [A ][i Ai ]


[ Ag ] exp iA[A ] + i

d4 xj A .

Now we perform a gauge transformation of the integration variables A (x):


A (x) Ag (x).
Recalling the gauge invariance of the action A, the functional f , and the metric
DA (x), we find that
Z

WC [j] =

DA C [A ][ A ] exp(iA[A ])

C [A ]
Z

Z Y

dg(x)

DA L [A ]

(i Aig ) exp i

(27.103)
 Z

d4 xj Ag

( A (x)) exp iA[A ] + i

d xj

Ag0

As before, Ag0 is the gauge transform of A , which satisfies A = 0, such that


L

i Agi 0

= i

"

1
U(g0 ) L Ai + U 1 (g0 )i U(g0 ) U 1 (g0 ) = 0. (27.104)
ig

In deriving Eq. (27.103), we have used the fact that


C [A ]

Z Y

dg(x)

1
(i Aig

= C [A ]

Z (Y Y
x

i Agi 0

dua (x)

1
g0
M[Ag0 ]u C [A ]1
C [A ] = 1.
g

It is possible to solve Eq. (27.104) for Ag0 as a power series in A . This can be done
as a power series, beginning with
Agi 0

1
= ij i 2 j Aj + O(A2 ).

The source j in the Coulomb gauge will be restricted to


j0 = 0,

i ji = 0,

so that we may write


Z

d4 x j Ag0 =

d4 x j F (x; A ),

(27.105)

1398

27 Nonabelian Gauge Theory of Strong Interactions

where
F (x; A ) = A (x) + O(A2 ).

(27.106)

Carryinf this costruction to high orders we can finally express WC in terms of WL


as follows:
!)#
"
( Z
1
4
WL [j]|j=0 .
(27.107)
WC [j] = exp i d xj (x) F x;
i j
It is helpful to visualize the generating functionals (27.103) or (27.107) with the
help of Feynman diagrams. The two expressions imply that the Green functions
in the Coulomb gauge are the same as those in the Landau gauge, if the source
is suitably restricted by equations like those in (27.105). We only must take into
account extra vertices between source and field, represented by the term
Z

d4 x j (F A ).

(27.108)

Then one may construct Green functions in the Coulomb gauge from the Feynman
rules of the Landau gauge. This relationship becomes much simpler if we go to
the mass shell. In this case, we ought to compare only the terms having a pole in
each of the external momenta, pi , when p2i 0. Of all the diagrams generated by
the extra couplings of (27.108), only those will survive in this limit in which the
whole effect of the extra vertices can be reduced to a type of self-energy insertion
into the corresponding external line. The other corrections introduced by (27.108)
will not contribute to poles of the Green functions at p2i = 0, and therefore not to
the S-matrix. Hence in the limit p2i 0, the unrenormalized S-matrix elements in
Coulomb gauge C and Landau gauge L with propagators will differ
ZC
ZL

(p;
C)
=
(p;
L)
=
(g
+
.
.
.),
lim
D
(g + . . .).
lim
D

p2 0
p2 0
p2 + i
p2 + i
In particular, the ratio
2 = ZC /ZL
(27.109)
will be different from unity. In general, the unrenormalized S-matrix elements in
the two gauges C and L are related to each order by
SC = n SL = (ZC /ZL )n/2 SL ,
so that the renormalized S-matrix element
n/2

Sren ZC

n/2

SC = Z L

SL

is independent of the gauge in which the calculation was done.


As a consequence, WC [j] is equal to the expression (27.103) which would be
WL [j] except that the coefficient of j is Ag0 , instead of A . For the S-matrix, the
only consequence of this difference is that the renormalization constants attached
to each external line depend on the gauge. Thus we have shown that ultimately the
S-matrix can be calculated from WL [j].
As pointed out earlier, the only flaw in the above argument is that the singularity
at p2i = 0 is not in general a simple pole.

27.3 Equivalence of Landau and Coulomb gauges

27.3.1

1399

Approximate Chiral Symmetry

At first, the masses of the quarks are approximated to be equal to zero. Then
the Lagrangian density has an additional invariance. If we restrict our attention
only to the three quarks u, d, s, the flavor group is SU(3). For massless quarks, the
symmetry is extended to SU(3) SU(3), where the extension involves SU(3) flavor.
Its Noether current densities are
a
q(x),
2
a

j5a
(x) = q(x) 5 q(x),
2
ja (x) = q(x)

(27.110)

where the a matrices act on the flavor SU(3) indices. The current densities are
color singlets. The corresponding color octet currents are not observable, due to
local SU(3) color invariance.
After field quantization the field components satisfy the local SU(3) SU(3)
commutation rules derived in (8.281) and discussed in Section 25.3. The charges Qa
and axial charges Q5a , defined by
Qa =
Q5a =

d3 xj 0 ,
0
d3 xj5a
,

(27.111)

form the Lie algebra


[Qa , Qb ] = ifabc Qc ,
[Qa , Q5b ] = ifabc Q5c ,
[Q5a , Q5b ] = ifabc Qc .

(27.112)
(27.113)
(27.114)

From these we may form the chiral charges


QLa = (Qa Q5a )/2,

QR
a = (Qa + Q5a )/2,

(27.115)

which generate two commuting groups SU(3)L and SU(3)R , respectively. Due to
the chiral invariance of the massless quark gluon Lagrangian density (27.15), the
currents are conserved:
j (x) = 0,
j5 (x) = 0.

(27.116)

The ground state of the theory breaks the axial part of this symmetry spontaneously.
This gives rise to non-zero quark masses. The spontaneous breakdown is accompanied by massless pseudoscalar Nambu-Goldstone bosons, which are identified with
the pion and its flavor octet partners.
In nature, quarks are not massless and the axial charges are not conserved. The
masses of non-strange quarks are, however, very small so that the axial charges

1400

27 Nonabelian Gauge Theory of Strong Interactions

with the SU(2)-indices a = 1, 2 are approximately conserved. This is the basis of


the PCAC hypothesis (Partial Conservation of Axial vector Current). The nonzero
masses of mu , md , ms . . . raise the mass of the pion and the other pseudoscalar mesons
to the experimental nonzero values. The quark masses which give a consistent
picture of experimental data are [5]

mu
md
ms
mc
mt
mb

4.5
7.9
155
1270
40000
4250

1.4
2.4
50
50
10000
100

MeV.

(27.117)

In recent years, much insight into the theory has been gained from computer
simulations of lattice models of the theory. They have confirmed that the theory
has the desired properties to explain the many strongly interacting particles observed
in the laboratory.

Notes and References


In Monte Carlo simulations of lattice gauge theories, the mass spectrum of hyperons with b-quarks
has been calculated by
C. Alexandrou, A. Borelli, S. G
usken, F. Jegerlehner, K. Schilling, G. Siegert, R. Sommer, et al.,
hep-lat/9407027
The individual citations refer to:
[1] H. Fritzsch, M. Gell-Mann, Intern. conf. on Duality and Symmetry in Hadron Physics,
Weizmann Science Press, 1971; H. Fritzsch, M. Gell-Mann, H. Leutwyler, Phys. Lett. B 47,
365 (1973).
[2] The presentation in this section follows
E.S. Abers and T.D. Lee, Phys. Rep. C 9, 1 (1973);
V.N. Popov and L.D. Faddeev, Perturbation Theory for Gauge Invariant Fields, Kiev ITP
report (unpublished); Phys. Letters 35B (1967) 29;
B.S. DeWitt, Phys. Rev. 162, 1195 (1967).
[3] In addition to the references in [?] see
N.P. Konopleva and V.N. Popov, Kalibrovochnye Polya (Atomizdat, Moscow, 1972), in
Russian.
R.P. Feynman, Acta Phys. Polonica 26, 697 (1963);
B.S. DeWitt, Phys. Rev. 162, 1195, 1239 (1967);
S. Mandelstam, Phys. Rev. 175, 1580 (1968);
E.S. Fradkin and I.V. Tuytin, Phys. Letters B 30, 562 (1969); Phys. Rev. D 2, 2841 (1970);
M.T. Veltman, Nucl. Phys. B 21, 288 (1970);
G. T Hooft, Nucl. Phys. B 33, 173 (1971).
[4] This section is taken from V.N. Popov and L.D. Faddeev, Perturbation Theory for GaugeInvariant Fields, Kiev ITP report (unpublished).
A similar discussion was given for quantum electrodynamics in the operator field theory
language by
B. Zumino, J. Math. Phys. 1, 1 (1960).

Notes and References

1401

[5] H. Kleinert, Collective Quantum Fields, Lectures presented at the First Erice Summer School on Low-Temperature Physics, 1977, in Fortschr. Physik 26, 565-671 (1978)
(http://klnrt.de/55); J. Gasser and H. Leutwyler, Phys. Rep. 87, 77 (1982);
G. Harrison et al. Phys. Lett. B 47, 493 (1984).

1402

Index
approxiation
Hartree-Fock-Bogoljubov . . . . . . 1021

Lie . . . . . . . . . . . . . . . . . . . . 58, 249, 251


rotation group . . . . . . . . . . . . . . . . 59
algebra of charges . . . . . . . . . . . . . . . . . . 680
Alsing, P.M. . . . . . . . . . . . . . . . . . . . . . 601
Ambegaokar, V. . . . . . . . . . . . 997, 1000
Ambjorn, J. . . . . . . . . . . . . . . . . . . . . . 1403
Ambler, E. . . . . . . . . . . . . . . . . . 314, 1379
Amp`ere law . . . . . . . . . . . . . . . . . . . . . . . . 273
Ampere
law . . . . . . . . . . . . . . . . . . . 479, 519, 832
amplitude . . . ,see also time evolution44
fixed-energy . . . . . . . . . . . . . . . . . 47, 51
free-particle . . . . . . . . . . . . . . . . . . . . . 49
reflection . . . . . . . . . . . . . . . . . . . . . . 697
scattering . . . . . . . . . . . . .553, 664, 677
time evolution . . . . . . . . . . 44, 47, 150
transmission . . . . . . . . . . . . . . . . . . . 697
amputated
4-point function . . . . . . . . . . . . . . . . 759
2-point function . . . . . . . . . . . . . . . . 766
analogy, Napier . . . . . . . . . . . . . . . . . . . . 376
analytic regularization . . . . . . . . . . . . . . 801
Anderson, M.H. . . . . . . . . . . . . . 174, 237
Anderson, P.W. .997, 998, 10001002,
1260, 1278
Andrews, M.R. . . . . . . . . . . . . . . 174, 237
angle
Cabibbo . . . . . . . . . . . . . . . . . . . . . . 1367
Euler . . . . . . . . . . . . . . . . . . . . 62, 64, 66
angular momentum . . . . . . . . . . . . . . . . . . 58
four-dimensional
total . . . . . . . . . . . . . . . . . . . . 278, 310
intrinsic . . . . . . . . . . . . . . . . . . . . . . . . 242
anomalous magnetic moment . . 472, 481,
881
anomaly . . . . . . . . . . . . . . . . . . . . . . . . . . . . 797
chiral . . . . . . . . . . . . . . . . . . . . . . . . . . 797
conformal . . . . . . . . . . . . . . . . . . . . . . 797
Anselm, A.A. . . . . . . . . . . . . . . . . . . . . 1326
antenna formula . . . . . . . . . . . . . . . . . . . . 388

Aarseth, J.B. . . . . . . . . . . . . . . . . . . . . 601


Abbott, J. . . . . . . . . . . . . . . . . . . . . . . . . 420
Abers, E.S. . . . . . . . . . . . . . . . . . . . . . . 1400
Ablowitz, M.J. . . . . . . . . . . . . . 999, 1002
Abo-Shaeer, J.R. . . . . . . . . . . . .174, 237
Abraham, R. . . . . . . . . . . . . . . . . . . . . . . .80
Abramowitz, M. . . . 51, 486, 599, 600,
682, 699, 703, 704, 1003
Abrikosov, A.A. . . . 174, 175, 996, 999
absorptive part
of Green function . . . . . . . . . . . . . . 158
acceleration field . . . . . . . . . . . . . . . . . . . 387
action . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
at a distance . . . . . . . . . . . . . . . . . . . 257
canonical . . . . . . . . . . . . . . . . . . . . . . . . . 3
classical . . . . . . . . . . . . . . . . . . . . . . . . . . 1
collective . . . . . . . . . . . . . . . . . . . . . . 1139
Einstein-Hilbert . . . . . . . . . . . . . . . . 289
euclidean . . . . . . . . . . . . . . . . . . . . . . . 932
local . . . . . . . . . . . . . . . . . . . . . . 106, 257
Adam, C. . . . . . . . . . . . . . . . . . . . . . . . . . . 920
Adamchik, V. . . . . . . . . . . . . . . . . . . . .1003
adiabatic . . . . . . . . . . . . . . . . . . . . . . . . . . . 654
process . . . . . . . . . . . . . . . . . . . . . . . . 126
adjoint
Dirac spinor . . . . . . . . . . . . . . . . . . . 307
Hermitian operator . . . . . . . . . . . . . 17
representation . . . . . . . . . . . . . . . . . . 250
Adkins, G.S. . . . . . . . . . . . . . . . . . . . . . . 921
Adrianov, V.A. . . . . . . . . . . . . 997, 1000
advanced Green function . . . . . . . . . . . 155
affine connection
Riemann . . . . . . . . . . . . . . . . . . . . . . . . 79
Akama, K. . . . . . . . . . . . . . . . . . . . . . . . 1327
Akulov, V.P. . . . . . . . . . . . . . . . . . . . . 1401
Alexandrou, C. . . . . . . . . . . . . . . . . . 1400
algebra
Clifford . . . . . . . . . . . . . . . . . . . . . . . . 325

1403

1404
anti-instanton . . . . . . . . . . . . . . . . . . . . . 1097
anticausal . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
time evolution operator . . . . . . . . . 39
anticommutator . . . . . . . . . . . . . . . . . . . . . 99
Dirac matrices . . . . . . . . . . . . . . . . . 318
anticommuting variables . . . . . . . . . . . 324
antikink . . . . . . . . . . . . . . . . . . . . . . . . . . . 1097
antilinear . . . . . . . . . . . . 267, 268, 270, 478
antineutrino . . . . . . . . . . . . . . . . . . . . . . . 340
antiparticle . . . . . . . . . . 475, 480, 481, 500
antiperiodic
boundary condition . . . . . . . . . . . . 153
Green function . . . . . . . . . . . . . . . . . 153
antiquark fields . . . . . . . . . . . . . . . . . . . 1302
antiunitary . . . . . . . . . . . . . . . . . . . . . . . . 268
Antonenko, S.A. . . . . . . . . . . . . . . . . 1278
approximation
Born . . . . . . . . . . . . . . . . . . . . . . . . . . 665
dipole . . . . . . . . . . . . . . . . . . . . . . . . . 849
Arvanitis, C. . . . . . . . . . . . . . . . . . . . . 1279
Ashmore, J.F. . . . . . . . . . . . . . . . . . . . . 920
associated Legendre polynomial . . . . 381
associativity of Lie Algebra . . . . . . . . . 250
asymmetric
spinning top . . . . . . . . . . . . . . . . . . . . .78
asymptotic convergence . . . . . . . . . . . 1225
Atkins, G.S. . . . . . . . . . . . . . . . . . . . . . . 920
atom
hydrogen . . . . . . . . . . . . . . . . . . . . . . . 848
super . . . . . . . . . . . . . . . . . . . . . . . . . . 138
atomic number . . . . . . . . . . . . . . . . . . . . .314
automorphism . . . . . . . . . . . . . . . . . . . . . . 256
average
energy . . . . . . . . . . . . . . . . . . . . . . . . . . 71
particle number . . . . . . . . . . . . . . . . . 71
axial
gauge . . . . . . . . . . . . . . . . 275, 276, 281
vector
field . . . . . . . . . . . . . . . . . . . . . . . . . 279
axialvector current density . . . . . . . . . 325
azimuthal fluctuations . . . . . . . . . . . . . 1119
Babaev, E. . . . . . . . . . . . . . . . . 1150, 1312
Bachmann, M. . . . . . . . . . . . . . . . . . . . 1050
Bailey, J. . . . . . . . . . . . . . . . . . . . . . . . . . 921
Baker-Campbell-Hausdorff formula . . 44,
249, 369

Index
Ballow, D.D. . . . . . . . . . . . . . . . . . . . . 377
Bardeen, J. . . . . . . . . . . . . 237, 999, 1280
Bardeen, W.A. . . . . . . . . . . . . . . . . . . . 921
Bargmann, V. . . . . . . . . . . . . . . . . . . . . 474
Barnes, E.W. . . . . . . . . . . . . . . . . . . . 1003
Barnich, G. . . . . . . . . . . . . . . . . . . . . . . 542
Barton, G. . . . . . . . . . . . . . . . . . 997, 1000
Barut, A.O. . . . . . . . . 174, 684, 920, 921
baryons . . . . . . . . . . . . . . . . . . . . . . . . . . . 1380
basis
complete in Hilbert space . . . . . . . 21
functions . . . . . . . . . . . . . . . . . . . . . . . . 21
local . . . . . . . . . . . . . . . . . . . . . . . . . . 19
occupation number . . . . . . . . . . . . 116
Baym, G. . . . . . . 175, 238, 959, 996, 999
Bealmonod, M.T. . . . . . . . . . . 998, 1001
beam
colliding . . . . . . . . . . . . . . . . . . . . . . . .689
direct . . . . . . . . . . . . . . . . . . . . . . . . . . 661
Becchi, C. . . . . . . . . . . . . . . . . . . . . . . . . 543
Bednorz, J.G. . . . . . . . . . . . . . . . 237, 999
Belavin, A. . . . . . . . . . . . . . . . . . 997, 1001
Belinfante
energy-momentum tensor . . 674676
Belinfante, F. . . . . . . . . . . . . . . . . . . . 685
Bender, C.M. . . . . . . . . . . . . . . . . . . . 1279
Bennett, G.W. . . . . . . . . . . . . . . . . . . . 921
Berlin, T.H. . . . . . . . . . . . . . . . . . . . . .1122
Bernard, V. . . . . . . . . . . . . . . . . . . . . . . 383
Bernoulli
number . . . . . . . . . . . . . . . . . . . . . . . . 598
Bernoulli polynomial . . . . . . . . . . . . . . . 597
Bernreuther, W. . . . . . . . . . 1259, 1278
Bes, D.R. . . . . . . . . . . . . . . . . . . . 998, 1001
Bessel function . . . . . . . 51, 485, 682, 1101
modified . . . . . . . . . . . . . . . . . . . . 51, 484
spherical . . . . . . . . . . . . . . . . . . . . . . . 703
Bessel-Hagen, E. . . . . . . . . . . . . . . . . 685
Beta function . . . . . . . . . . . . . . . . . . . . . . 173
-decay
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1366
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1366
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1366
He6 . . . . . . . . . . . . . . . . . . . . . . . . . . . 1365
inverse . . . . . . . . . . . . . . . . . . . . . . . . 1365
Ne19 . . . . . . . . . . . . . . . . . . . . . . . . . . 1364

1405
neutron . . . . . . . . . . . . . . . . . . . . . . 1362
parity violation . . . . . . . . . . . . . . . . 314
Bethe, H.A. . . . . . . . . . . . . . 488, 921, 922
Bethe-Heitler cross section . . . . . . . . . .875
Bialynicke-Birula, I. . . . . . . . . . . . 1326
Bianchi identity . . . . . 272, 292, 683, 964
bilocal operator . . . . . . . . . . . . . . . 102, 107
biparticle state . . . . . . . . . . . . . . . . . . . . .504
Birge, R.W. . . . . . . . . . . . . . . . . . . . . . 1379
bispinor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
canonical . . . . . . . . . . . . . . . . . . . . . . 331
helicity . . . . . . . . . . . . . . . . . . . . . . . . 340
Bjorken, J.D. . . . . . . . . . . . . . . 920, 1326
black hole . . . . . . . . . . . . . . . . . . . . . . . . . . 402
black-body radiation . . . . . . . . . . . . . . 1326
Blaizot, J.-P. . . . . . . . . . . . . . . . . . . . . 238
Blanchet, L. . . . . . . . . . . . . . . . . . . . . . 427
Bleuler, K. . . . . . . . . . . . . . . . . . . . . . . 538
Bloch theorem . . . . . . . . . . . . . . . . . . . . . 131
Blount, E.I. . . . . . . . . . . . . . . . . 997, 1001
Boerner, H. . . . . . . . . . . . . . . . . . . . . . . 174
Bogoliubov transformation . . . . . . . . . 198
Bogoliubov, N.N. . . . . . . 197, 237, 999
Bohr
magnetic moment . . . . 469, 843, 849
radius . . . . . . . . . . . . . . . . . . . . . . . . . . 849
Bohr, N. . . . . . . . . . . . . . . . . . . . . . . . . . 1362
Bollini, C. . . . . . . . . . . . . . . . . . . . . . . . . 920
Boltzmann
constant . . . . . . . . . . . . . . . . . . . . . . . . 69
factor . . . . . . . . . . . . . . . . . . . . . . 69, 933
Borelli, A. . . . . . . . . . . . . . . . . . . . . . . 1400
Born approximation . . . . . . . . . . . . . . . 665
Bose
-Einstein
condensate . . . . . . . . . . . . . . . . . . . . 81
condensate, trap . . . . . . . . . . . . . 141
distribution . . . . . . . . . . . . . . . . . . 155
condensation . . . . . . . . . . . . . . . . . . 134
symmetry . . . . . . . . . . . . . . . . . . . . . . 865
Bose or Fermi distribution function . 951
boson
Goldstone . . . . . . . . . . . . . . . . . . . . . 1120
bosons
Green function . . . . . . . . . . . . . . . . . 118
nonequilibrium Green functions 154

number or particles . . . . . . . . . . . . 122


thermal Green function . . . . . . . . .778
bottom quark . . . . . . . . . . . . . . . . . . . . . 1317
ckner, K. . . . . . . . . . . . . . . . . . . . . 194
Bru
Br
ezin, E. . . . . . . . . . . . . . . . . . 1260, 1278
bra-ket, Dirac . . . . . . 18, 21, 25, 110, 359
Braaten, E. . . . . . . . . . . . . . . . . . . . . . . 238
bracket
Lagrange . . . . . . . . . . . . . . . . . . . . . . 7, 7
Poisson . . . . . . . . . . . . . . 4, 7, 8, 40, 58
Bradley, C.C. . . . . . . . . . . . . . . . 174, 237
Bragg reflection . . . . . . . . . . . . . . . . . . . . . 11
Brandenberger, R. . . . . . . . . . 601, 780
Brans, C. . . . . . . . . . . . . . . . . . . . . . . . . . 427
Breitenlohner, P. . . . . . . . . . . . . . . . 920
Bremsstrahlung . . . . . . . . . . . . . . . . . . . . 870
Bressi, G. . . . . . . . . . . . . . . . . . . . . . . . . . 601
Bretin, V. . . . . . . . . . . . . . . . . . . . . . . . . 174
Brevik, I. . . . . . . . . . . . . . . . . . . . . . . . . . 601
Brezin, E. . . . . . . . . 503, 996, 1000, 1277
Bridge, H.S. . . . . . . . . . . . . . . . . . . . . . 1338
Brink, D.M. . . . . . . . . . . . . . . . . . . . . . . 921
Brinkman, W.F. . 997, 998, 1000, 1001
Brittin, W.E. . . . . . . . . . . . . . . . 920, 921
Broglia, R.A. . . . . . . . . . . . . . . 998, 1001
Brown, F.R. . . . . . . . . . . . . . . . . . . . . . . 920
Brown, G.E. . . . . . . . . . . . . . . . . . . . . 1338
Brown, H.N. . . . . . . . . . . . . . . . . . . . . . 921
Brown, L.M. . . . . . . . . . . . . . . . . . . . . . 921
Brown, L.S. . . . . . . . . . . . . . . . . . . . . . . 383
Bruckner, K.A. . . . . . . . . . . . . . . . . . . 237
Buchholtz, L.J. . . . . . . . . . . . . 998, 1001
Buckley, I.R.C. . . . . . . . . . . . . . . . . . 1279
Bulanov, S.V. . . . . . . . . . . . . . . . . . . . 1003
Bunster, C. . . . . . . . . . . . . . . . . . . . . . . 384
Burgay, M. . . . . . . . . . . . . . . . . . . . . . . . 427
Burnett, K. . . . . . . . . . . . . . . . . . . . . . . 705
Butler, D. . . . . . . . . . . . . . . . . . .997, 1000
Cabibbo
angle . . . . . . . . . . . . . . . . . . . . . . . . . .1367
mass matrix . . . . . . . . . . . . .1368, 1378
Cabibbo, N. . . . . . . . . . . . . . . . . . . . . . 1379
Cage, M.E. . . . . . . . . . . . . . . . . . . . . . . . 920
Calarco, T. . . . . . . . . . . . . . . . . . . . . . . 680
calculus
Weyl . . . . . . . . . . . . . . . . . . . . . . . . . . . 362

1406
Callan-Symanzik equation . . . . . . . . . 1199
Callen, H.B. . . . . . . . . . . . . . . . . . . . . . . 80
Camilo, F. . . . . . . . . . . . . . . . . . . . . . . . . 427
canonical
action . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
bispinor . . . . . . . . . . . . . . . . . . . . . . . 331
commutation relations . . 16, 40, 522
ensemble . . . . . . . . . . . . . . . . . . . . . . . . 69
Laplacian . . . . . . . . . . . . . . . . . . . . . . . 57
quantization . . . . . . . . . . 40, 5759, 67
transformation . . . . . . . . . . . . . . . . . 6, 8
generating function . . . . . . . . . . . . . 9
Cartesian
coordinates . . . . . . . . . . . . . . . . . . . . . . 40
Carugno, G. . . . . . . . . . . . . . . . . . . . . . .601
cascade . . . . . . . . . . . . . . . . . . . . . . . . . . . 1298
Casimir
effect . . . . . . . . . . . . . . . 563, 814, 1146
force . . . . . . . . . . . . . . . . . . . . . . 469, 568
operator . . . . . . . . . . . . . . . . . . . . . . . 352
Castilly, G.E. . . . . . . . . . . . . 1260, 1278
Cauchy residue theorem . . . . . . . 147, 669
causal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
ordering . . . . . . . . . . . . . . . . . . . . . . . . . 36
propagator . . . . . . . . . . . . . . . . . . . . . 119
time evolution
amplitude . . . . . . . . . . . . . . . . . . . . . 45
operator . . . . . . . . . . . . . . . . . . . . . . 45
center-of-mass
energy . . . . . . . . . . . . . . . . . . . . . . . . . 687
frame . . . . . . . . . . . . 685, 686, 688, 689
momentum . . . . . . . . . . . . . . . . . . . . . 687
theorem . . . . . . . . . . . . . . . . . . . 653, 668
center-of-mass frame . . . . . . . . . . . . . . . 688
chain rule for operators . . . . . . . . . 96, 112
Chakravarty, S. . . . . . . . . . . 1260, 1278
champaign bottle potential . . . . . . . . . 142
Chao, W.Q. . . . . . . . . . . . . . . . . . . . . . 1327
charge
algebra . . . . . . . . . . . . . . . . . . . . . . . . 680
conjugation . . . . . . . . . . . . . . . 270, 271
operator . . . . . . . . . . . . . . . . . . . . . 480
density . . . . . . . . . . . . . . . . . . . . . . . . . 658
electric . . . . . . . . . . . . . . . . . . . . . 83, 888
electron . . . . . . . . . . . . . . . . . . . . . . . . 832
form factor . . . . . . . . . . . . . . . . 886, 901

Index
funcamental . . . . . . . . . . . . . . . . 83, 888
Noether . . . . . . . . . . . . . . . . . . . 646, 667
number . . . . . . . . . . . . . . . . . . . . . . . . 314
parity . . . . . . . . . . . . . . . . . . . 271, 1293
radius . . . . . . . . . . . . . . . . . . . . . . . . . . 887
symmetry . . . . . . . . . . . . . . . . . . . . . 1291
charge-conjugate
Pauli spinor . . . . . . . . . . . . . . . . . . . 333
charm quark . . . . . . . . . . . . . . . . . . . . . . 1317
chemical potential . . . . . . . . . . . . . . . . . . . 70
Cheng, K.S. . . . . . . . . . . . . . 81, 998, 1001
Chervyakov, A. . . . . . . . . . . . . . . . . . . 503
Cheston, W.B. . . . . . . . . . . . . . . . . . . . 693
Chetyrkin, K.G. . . . . . 824, 1260, 1278
Chevy, F. . . . . . . . . . . . . . . . . . . . . . . . . . 174
Chibisov, G.V. . . . . . . . . . . . . . . . . . . . 271
chiral
anomaly . . . . . . . . . . . . . . . . . . . . . . . 797
Gross-Neveu model . . . . . . . . . . . 1305
representation . . . . . . . . . . . . 317, 331
chirality . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
matrix . . . . . . . . . . . . . . . . . . . . 317, 340
Chistensen, J.H. . . . . . . . . . . . . . . . . 1379
Christoffel symbol . . . . . . . . . . 11, 79, 290
Clark, D.L. . . . . . . . . . . . . . . . . . . . . . . 693
classical
electron radius . . . . . . . . . . . . . . . . . 858
fields, collective . . . . . . . . . . . . . . . 1028
mechanics . . . . . . . . . . . . . . . . . . . . . . . . 1
orbit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
partition function . . . . . . . . . . . . . . . 69
path . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2
solution . . . . . . . . . . . . . . . . . 1096, 1097
statistics . . . . . . . . . . . . . . . . . . . . . . . . 68
Clebsch-Gordan coefficients . . . 296, 356,
361, 377379, 381, 489, 502, 511,
851
Clifford algebra . . . . . . . . . . . . . . . . . . . . 325
coefficients
Clebsch-Gordan . . . . . . 296, 356, 361,
377379, 381, 489, 502, 511
Fourier . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Racah recoupling . . . . . . . . . . . . . . 505
coherent state . . . . . . . . . . . . . . . . . . . . . .531
Coleman
theorem . . . . . . . . . . . . . . . . . . . . . . . 1153

1407
Coleman, S. . . 685, 920, 996, 999, 1160
collective
action . . . . . . . . . . . . . . . . . . . . . . . . . 1139
classial fields . . . . . . . . . . . . . . . . . . 1028
field . . . . . . . . . . . . . . . . . . . . . 1137, 1139
collective quantum fields . . . . . . . . . . 1280
collective phenomena . . . . . . . . . . . . . . . . 91
colliding beam . . . . . . . . . . . . . . . . . . . . . .689
commutation rules
canonical . . . . . . . . . . . . . . . 16, 40, 522
equal-time . . . . . . . . . . . . . . . . . . 40, 522
commutator function . . . . . . . . . . . . . . .470
commuting observables . . . . . . . . . . . . . . . 4
complete basis . . . . . . . . . . . . . . . . . . . . . . 21
completeness
relation . . . . . . . . . . . . . . . . . . . . . . . . 377
relation, semi . . . . . . . . . . . . . . . . . . 336
completeness relation 19, 22, 23, 28, 29,
32, 47, 49, 522
basis dyads . . . . . . . . . . . . . . . . . . . . . . 53
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
component
spherical . . . . . . . . . . . . . . . . . . . . . . . 851
components
Cartesian . . . . . . . . . . . . . . . . . . . . . . . 40
spherical . . . . . . . . . . . . . 282, 360, 362
transverse . . . . . . . . . . . . . . . . . . . . . 826
composite field . . . . . . . . . . . . . . . . . . . . 1137
composition law for time evolution operator . . . . . . . . . . . . . . . . . . . . . . . . 38,
39
Compton
relation . . . . . . . . . . . . . . . . . . . . . . . . 862
scattering . . . . . . . . . . . . . . . . . . . . . .857
wavelength . . . . . . . . . . . . . . . . . . . . . 904
condensate . . . . . . . . . . . . . . . . . . . 134, 1126
condensation
Bose-Einstein . . . . . . . . . . . . . . 81, 134
energy . . . . . . . . . . . . . . . . . . 1096, 1102
condensed phase . . . . . . . . . . . . . . . . . . . .762
condition
Fermi-Dirac . . . . . . . . . . . . . . . . . . . 524
Gupta-Bleuler . . . . . . . . . . . . . . . . . . 538
Schwarz integrability . . . . . . . . . . . . . 6
Condon, E.U. . . . . . . . . . . . . . . . . . . . . .354
Condon-Shortley

phase convention 281, 286, 354, 378


confluent hypergeometric functions . 483
conformal anomaly . . . . . . . . . . . . . . . . . 797
conjugation
charge . . . . . . . . . . . . . . . . . . . . 270, 271
hermitian . . . . . . . . . . . . . . . . . . . . . . . 32
connection
Christoffel, or affine . . . . . . . . . . . . 290
Riemann . . . . . . . . . . . . . . . . . . . . . . . . 79
conservation . . . . . . . . . . . . . . . . . . . . . . . 645
current . . . . . . . . . . . . . . . . 18, 262, 658
energy . . . . . . . . . . . . . . . . . . . . . . . . . . 14
global . . . . . . . . . . . . . . . . . . . . . 117, 658
law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 833
local . . . . . . . . . . . . . . . . . . . . . . . . . 658
local . . . . . . . . . . . . . . . . . . . . . . 116, 661
probability . . . . . . . . . . . . . . . . . . . . . . 16
conserved
current . . . . . . . . . . . . . . . . . . . . 661, 833
quantity . . . . . . . . . . . . . . . . . . . . . . . .646
constant
Boltzmann . . . . . . . . . . . . . . . . . . . . . . 69
Euler-Mascheroni . . . . . . . . . . . . . . 984
fine-structure . . . . . . . . . . . . . . . . . . .470
gravitational . . . . . . . . . . . . . . . . . . . 292
of motion . . . . . . . . . . . . . . . . . .646, 649
Planck . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Stefan-Boltzmann . . . . . . . 1147, 1326
structure . . . . . . . . . . . . . . . . . . . . . . 250
constraint
primary . . . . . . . . . . . . . . . . . . . 514, 548
secondary . . . . . . . . . . . . . . . . . . . . . .514
continuity law . . . . . . . . . . . . . . . . . . . . . . . 18
continuous spectrum . . . . . . . . . . . . . . . . .49
contraction . . . . . . . . . . . . . . . . . . . . . . . . . 580
group . . . . . . . . . . . . . . . . . . . . . . . . . . 348
Wick . . . . . . . . . . . . . . . . . . . . . . . . . . . 776
contravariant
spherical tensor operator . . . . . . . 362
convention, Einstein summation . . . . 2, 4
Cooper, L. N. . . . . . . . . . . . . . . . . . . . 1280
Cooper, L.N. . . . . . . . . . . . . . . . . 237, 999
Cooperstock, F.I. . . . . . . . . . . . . . . . 395
coordinate
generalized . . . . . . . . . . . . . . . . . . . . . . . 1
transformation

1408
general . . . . . . . . . . . . . . . . . . . . . . . 292
Cornell, E.A. . . . . . . . . . . . . . . . 174, 237
Cornwall, J.M. . . . . . . . . . . . 1029, 1049
correlation
energy . . . . . . . . . . . . . . . . . . . . . . . . . 194
function
4-point . . . . . . . . . . . . . . . . . . . . . . . 759
n-point . . . . . . . . . . . . . . . . . . . . . . 748
2-point . . . . . . . . . . . . . . . . . . . . . . . 757
correlation function . . . . . . . . . . . . . . . . 953
correspondence principle . 15, 17, 31, 56,
58, 64, 68
group . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Heisenberg . . . . . . . . . . . . . . . . . . 41, 42
Coulomb
gauge . . . . . . . 275, 276, 281, 515, 519
Hamiltonian . . . . . . . . . . . . . . . . . . . . . 16
interaction . . . . . . . . . . . . . . . . . . . . . 837
law 273, 275, 285, 479, 515, 825, 832
potential . . . . . . . . . . . . . . . . . . . . . . . 895
Dirac equation . . . . . . . . . . . . . . . 485
external . . . . . . . . . . . . . . . . . . . . . . 481
Klein-Gordon equation . . . . . . . 484
coupling
reduced . . . . . . . . . . . . . . . . . . . . . . . . 812
spin-orbit . . . . . . . . . . . . . . . . . . . . . . 470
covariant
derivative . . . . . . . . . . . . . . . . . 477, 830
critical
exponent . . . 1037, 1087, 1210, 1213,
12181220
index . . . . . . . . . . . . . . . . . . . . . . . . . 1087
magnetization . . . . . . . . . . . . . . . . . 1220
phenomena . . . . . . . . . . . . . . . . . . . 1197
temperature . . . . . . . . . . . . . . . . . . . . 208
theory . . . . . . . . . . . . . . . . . . . . . . . . 1038
Cronin, J.W. . . . . . . . . . . . . . . . . . . . . 1379
cross section
differential . . . . . . . . . . . . . . . . . . . . . 674
laboratory . . . . . . . . . . . . . . . . . . . . . 686
crossing symmetry . . . . . . . . . . . . . 859, 865
cumulant expansion . . . . . . . . . . . . . . . 1008
Curie temperature . . . . . . . . . . . . . . . . 1133
current
conservation . . . . . . . . . . . . . . . . . . . 658
Euler-Lagrange type equation 648

Index
conservation law . . . . . . . . . . . 18, 262
conserved . . . . . . . . . . . . . . . . . 661, 833
density . . . . . . 18, 307, 325, 478, 831
external . . . . . . . . . . . . . . . . . . . . . . . 587
neutral . . . . . . . . . . . . . . . . . . 1369, 1372
Noether . . . . . . . . . . . . . . . . . . . . . . . 658
current density
axialvector . . . . . . . . . . . . . . . . . . . . .325
vector . . . . . . . . . . . . . . . . . . . . . . . . . 325
Curtright, T. . . . . . . . . . . . . . . . . . . . . 806
curvature
scalar . . . . . . . . . . . . . . . . . . 68, 80, 290
of spinning top . . . . . . . . . . . . . . . . 80
tensor
Riemann . . . . . . . . . . . . . . . . . . . . 289
curved spacetime . . . . . . . . . . . . . . . . . . . . 10
cutoff
function . . . . . . . . . . . . . . . . . . . . . . . 566
ultraviolet . . . . . . . . . . . . . . . . . . . . . . 782
UV . . . . . . . . . . . . . . . . . . . . . . . . . . . . 782
cyclic
coordinate . . . . . . . . . . . . . . . . . . . . . . . .9
Czarnecki, A. . . . . . . . . . . . . . . . . . . . . 921
DAmico, N. . . . . . . . . . . . . . . . . . . . . . . 427
Dahmen, H. . . . . . . . . . . . . . . . . . . . . . . 1029
Dalibard, J. . . . . . . . . . . . . . . . . . . . . . . 174
Dalitz plot . . . . . . . . . . . . . . . . . . . . . . . . . 688
Dalvit, D.A.R. . . . . . . . . . . . . . . . . . . . 601
Damour, T. . . . . . . . . . . . . . . . . . . . . . . . 427
Davier, M. . . . . . . . . . . . . . . . . . . . . . . . . 921
Davies, P.C.W. . . . . . . . . . . . . . . . . . . . 602
Davis, A-C. . . . . . . . . . . . . . . . . . . 601, 780
Davis, K.B. . . . . . . . . . . . . . . . . . . 174, 237
Davis, M. . . . . . . . . . . . . . . . . . . . . . . . . . 427
Dayhoff, E.E. . . . . . . . . . . . . . . . . . . . . 922
De Celles, L. . . . . . . . . . . . . . . . . . . . . 601
de Dominicis, C. . . . . . . . . . . . . . . . . 1029
De Gennes, P.G. . . . . . 997, 1000, 1001
de Gennes, P.G. . . . . . . . . . . . . 998, 1002
De Witt, B.S. . . . . . . . . . . . . . . 998, 1001
De Dominicis, C. . . . . . . . . . . . . . . . . 1049
DeBroglie, L.V. . . . . . . . . . . . 133, 1150
decomposition formula
Gordon . . . . . . . . . . . . . . . . . . . . . . . . 845
defect
line . . . . . . . . . . . . . . . . . . . . . . . .483, 756

1409
degenerate
Fermi gas . . . . . . . . . . . . . . . . . . . . . . 126
Dekker, H. . . . . . . . . . . . . . . . . . . . . . . . . 81
Delrieu, J.M. . . . . . . . . . . . . . . 998, 1001
-function
and Heaviside function . . . . . . . . . . 45
Dirac . . . . . . . . . . . . . . . . . . . . . . . .24, 45
transverse . . . . . . . . . . . . . . . . . . . . . . 515
-functional . . . . . . . . . . . . . . . . . . . . . . . . 924
density
axialvector current . . . . . . . . . . . . . 325
charge . . . . . . . . . . . . . . . . . . . . . . . . . 658
current . . . . . . 18, 307, 325, 478, 831
Hamiltonian . . . . . . . . . . . . . . . . . . . 102
Lagrangian . . . . . . . . . . .106, 258, 462
matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 34
normal . . . . . . . . . . . . . . . . . . . . . . . . . 182
of states . . . . . . . . . . . . . . . . . . . . . . . 122
operator . . . . . . . . . . . . . . . . . . . . . . . . 33
probability . . . . . . . . . . . . . . . . . . . . . . 18
states . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
vector current . . . . . . . . . . . . . . . . . 325
depletion . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
derivative, covariant . . . . . . . . . . .477, 830
Deser, S. . . . . . . . . . . . . . . . . 384, 420, 427
detailed balance, principle . . . . . . . . . 693
determinant
functional . . . . . . . . . . . . . . . . . . . . . 926
Slater . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
DeWitt, B.S. . . . . . . . . . . . 81, 602, 1400
diagram
disconnected . . . . . . . . . . . . . . . . . . . 757
Feynman . . . . . . . . . . . . . . . . . . . . . . . 755
vacuum . . . . . . . . . . . . . . . . . . . . . . . . 756
diagrams
vacuum . . . . . . . . . . . . . . . . . . . . . . . . 814
Dicke, R.H. . . . . . . . . . . . . . . . . . . . . . . . 427
differential
cross section . . . . . . . . . . . . . . . . . . . 674
differential cross section in the center of
mass . . . . . . . . . . . . . . . . . . . . . . . 686
differential equation
Hamilton-Jacobi . . . . . . . . . . . . . . . . 10
diffraction pattern . . . . . . . . . . . . . . . . . . . 30
dimensional regularization . . . . . 801, 930
dimensional transmutation . . . . . . . . 1291

dimensionality theorem . . . . . . . . . . . . 168


dipole
approximation . . . . . . . . . . . . . . . . . 849
ghost . . . . . . . . . . . . . . . . . . . . . . . . . . 545
Dirac
action . . . . . . . . . . . . . . . . . . . . . . . . . . 307
algebra . . . . . . . . . . . . . . . . . . . . . . . . . 318
bra-ket notation 18, 21, 25, 110, 359
-function . . . . . . . . . . . . . . . . . . . 24, 45
and Heaviside function . . . . . . . . 45
equation . . . . 307, 312, 318, 320, 329
in Coulomb potential . . . . . . . . . 485
momentum space . . . . . . . . . . . . .329
field . . . . . . . . . . . . . . . . . . . . . . . 481, 559
parity . . . . . . . . . . . . . . . . . . . . . . . . 510
Lorentz transformations . . . . . . . . 308
matrices . . . . . . . . . . . . . . . . . . . . . . . 307
picture . . . . . . . . . . . . . . . . . . . . . . . . 646
spinor . . . . . . . . . . . . . . . . . . . . . . . . . 306
adjoint . . . . . . . . . . . . . . . . . . . . . . 307
vs-Maxwell-equations . . . . . . . . . . .320
Dirac, P.A.M. . . 80, 317, 383, 514, 524
direct
beam . . . . . . . . . . . . . . . . . . . . . . . . . . 661
disconnected diagram . . . . . . . . . . . . . . .757
discontinuity
fixed-energy amplitude . . . . . . . . . . 49
discrete
spectrum . . . . . . . . . . . . . . . . . . . . . . . 672
hydrogen . . . . . . . . . . . . . . . . . . . . . 483
symmetry
transformation . . . . . . . . . . . . . . . 558
discrete hydrogen pectrum
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . . 487
relativistic . . . . . . . . . . . . . . . . . . . . . 485
disorder
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 494
field theory . . . . . . . . . . . . . . . 494, 756
disorder field theories . . . . . . . . . . . . . . . 149
dispersive part of Green function . . . 158
dissipation
-fluctuation theorem . . . . . . . 158, 163
dissipative part
of Green function . . . . . . . . . 158, 158
distribution . . . . . . . . . . . . . . . . . . . . . . . . 529
Bose-Einstein . . . . . . . . . . . . . . . . . . 155

1410
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Fermi-Dirac . . . . . . . . . . . . . . . . . . . . 155
Gibbs . . . . . . . . . . . . . . . . . . . . . . 69, 928
Heaviside . . . . . . . . . . . . . . . . . . . . . . . .25
distributions . . . . . . . . . . . . . . . . . . . . . . . 490
distributions (generalized functions) 25,
46
divergence
ultraviolet . . . . . . . . . . . . . . . . . . . . . . 768
Dodo, T. . . . . . . . . . . . . . . . . . . . . . . . . . . 531
Dohm, V. . . . . . . . . . . . . . . . . . . . . . . . . 1052
Domb, C. . . . . . . . . . . . . . . . . . . . . 996, 1000
dominance
vector
meson . . . . . . . . . . . . . . . . . . . . . . . 555
Dominguez, A.G. . . . . . . . . . . . . . . . . . 920
doubl
well . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
double
-slit experiment . . . . . . . . . . . . . . . . . 12
-well potential . . . . . . . . . . . . . . . . .1098
-well potential . . . . . . . . . . .1096, 1097
pulsar . . . . . . . . . . . . . . . . . . . . . . . . . . 406
Dreitlein, J. . . . . . . . . . . . . . . . . . . . . . 601
Drell, S.D. . . . . . . . . . . . . . . . . . . . . . . . 920
dual
field tensor . . . . . . . . . . . . . . . . . . . . . 272
transform . . . . . . . . . . . . . . . . . . . . . . 570
Dulong-Petit law . . . . . . . . . . . . . . 125, 137
Duncan, A. . . . . . . . . . . . . . . . . . . . . . . 1279
Dunne, G.V. . . . . . . . . . . . . . . . . . . . . . 1003
Durand, L. . . . . . . . . . . . . . . . . . . . . . . . 601
Durbin, R. . . . . . . . . . . . . . . . . . . . . . . . . 693
Durfee, D.S. . . . . . . . . . . . . . . . . 174, 237
dynamical
metric . . . . . . . . . . . . . . . . . . . . . . . . . .106
dynamical metric . . . . . . . . . . . . . . . . . . . . 15
Dyson series . . . . . . . . . . . . . . . . . . .36, 1275
Dyson, F. . . . . . . . . . . . . . . . . . . . . . . . . . 582
Dzyaloshinski, I.E. . . . . . 174, 175, 999
Dzyaloskinski, I.E. . . . . . . . . . . . . . . 996
Ebisawa, H. . . . . . . 997, 999, 1000, 1002
Ecklund, K.M. . . . . . . . . . . . . . . . . . . 1379
Edmonds, A.R. . . . . . . . . . . . . . . . .81, 383
effect
Casimir . . . . . . . . . . . . . . . . . . . 563, 814

Index
frame-dragging . . . . . . . . . . . . . . . . 426
Lense-Thirring . . . . . . . . . . . . . . . . . 426
Paschen-Back . . . . . . . . . . . . . . . . . . 846
Zeeman . . . . . . . . . . . . . . . . . . . . . . . . 846
anomalous . . . . . . . . . . . . . . . . . . . 846
effective
range . . . . . . . . . . . . . . . . . . . . . . . . . . 670
effective action . . . . . . . . . . . . . . . . . . . 1030
Ehrenfest theorem . . . . . . . . . . . . . . . . . 269
Einstein
-Bose distribution . . . . . . . . . . . . . . 155
-Hilbert action . . . . . . . . . . . . . . . . . 289
parameter . . . . . . . . . . . . . . . . . . . . . . 241
summation convention . 2, 4, 98, 243
tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 291
Eiself, F. . . . . . . . . . . . . . . . . . . . . . . . . 1379
elastic cut . . . . . . . . . . . . . . . . . . . . . . . . . 1291
electric
charge . . . . . . . . . . . . . . . . . . . . . . 83, 888
electrodynamics, quantum (QED) . . 825
electromagnetic
coupling
minimal . . . . . . . . . . . . . . . . 830, 1108
field . . . . . . . . . . . . . . . . . . 386, 416, 421
electron
charge . . . . . . . . . . . . . . . . . . . . . . . . . 832
radius
classical . . . . . . . . . . . . . . . . . . . . . 858
electroweak interactions . . . . . . . . . . . 1370
elliptic theta function . . . . . . . . . 138, 169
energy
-momentum tensor
Belinfante . . . . . . . . . . . . . . . . . . . . 676
average . . . . . . . . . . . . . . . . . . . . . . . . . 71
center-of-mass . . . . . . . . . . . . . . . . . . 687
condensation . . . . . . . . . . . . 1096, 1102
conservation law . . . . . . . . . . . . . . . . .14
correlation . . . . . . . . . . . . . . . . . . . . . 194
free . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
internal . . . . . . . . . . . . . . . . . . . . . . . . . 71
level shift . . . . . . . . . . . . . . . . . . . . . . 657
shell . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
vacuum . . . . . . . . . . . . . . . . . . . 533, 814
energy conservation . . . . . . . . . . . . . . . 1098
energy-momentum tensor
Belinfante . . . . . . . . . . . . . . . . . 674, 675

1411
Engelsberg, A. . . . . . . . . . . . . .997, 1001
ensemble
canonical . . . . . . . . . . . . . . . . . . . . . . . 69
grand-canonical . . . . . . . . . . . . . . 71, 74
Ensher, J.R. . . . . . . . . . . . . . . . . . 174, 237
equal-time commutation rules . . . 40, 522
equation
Dirac . . . . . . . 307, 312, 318, 320, 329
in Coulomb potential . . . . . . . . . 485
momentum space . . . . . . . . . . . . .329
Euler-Lagrange . . 2, 3, 5, 6, 10, 285,
463, 472, 514, 518, 548, 554
gap . . . . . . . . . . . . . . . . . . . . . . 202, 1282
Hamilton-Jacobi . . . . . . . . . . . . . . . . 10
Heisenberg . . . . . . . . . . . . . . . . . . . . . 470
Helmholtz . . . . . . . . . . . . . . . . . . . . . 564
Klein-Gordon . . . 261, 274, 295, 317,
463, 471, 487, 522, 541, 544, 547
in Coulomb potential . . . . . . . . . 484
Liouville . . . . . . . . . . . . . . . . . . . . . . . . 34
Lippmann-Schwinger . . . . . . 651653,
660664, 668, 673, 680, 701
Maxwell . . . . 320, 467, 479, 564, 871
of motion . . . . . . . . . . . . . . . . . . . . . . . 42
Hamilton . . . . . . . . . . . . . . . . .3, 4, 42
Heisenberg . . . . . . . . . . . . . . . . . . . . 42
of state . . . . . . . . . . . . . . . . . . . 123, 137
Pauli . . . . . . . . . . . . . . . . . . . . . 320, 480
Schr
odinger
relative . . . . . . . . . . . . . . . . . . . . . . 680
Schroedinger . . . . . 15, 1618, 25, 26,
3436, 3941, 45, 53, 55, 150
time-independent . . . . . . . . . . . . . 16
Thomas . . . . . . . . . . . . . . . . . . .475, 475
equation of motion
Heisenberg . . . . . . . . . . . . . . . . . . . . . 656
equivalence principle . . . . . . . . . . . . . . . 289
Erdelyi, A. . . . . . . . . . . . . . . . . . . . . . . 1003
Erickson, G.W. . . . . . . . . . . . . . . . . . . 921
Esposito, G. . . . . . . . . . . . . . . . . . . . . . . 601
Euclidean
group . . . . . . . . . . . . . . . . . . . . . . 58, 348
euclidean
action . . . . . . . . . . . . . . . . . . . . . . . . . . 932
four-momentum . . . . . . . . . . . . . . . . 952
spacetime coordinates . . . . . 932, 952

Euler
-Lagrange equations . . . 2, 3, 5, 6, 10
angles . . . . . . . . . . . . . . . . . . . 62, 64, 66
formula . . . . . . . . . . . . . . . . . . 599, 600
number . . . . . . . . . . . . . . . . . . . . . . . . 600
polynomial . . . . . . . . . . . . . . . . . . . . 600
relation . . . . . . . . . . . . . . . . . . . . . . . . 135
relation, thermodynamic . . . . . . . . 74
Euler, H. . . . . . . . . . . . . . . . . . . . . . . . . 1002
Euler-Heisenberg Lagrangian . . . . . . . 974
Euler-Lagrange . . . . . . . . . . . . . . . . . . . . . 833
euation . . . . . . . . . . . . . . . . . . . . . . . . .967
Euler-Lagrange equation . 285, 463, 472,
514, 518, 548, 554, 646
type of for current conservation 648
Euler-Maclaurin approximation . . . .1127
Euler-Maclaurin formula . . . . . . . . . . . 566
Euler-Mascheroni constant . . . . . . . . . 984
even permutation . . . . . . . . . . . . . . . . . . . 166
evolution . . . . . . . . . . . . . . ,see also time35
exclusion principle
Pauli . . . . . . . . . . . . . . . . . . . . . . . . . . 112
expansion
fugacity . . . . . . . . . . . . . . . . . . . . . . . 138
Jacob-Wick . . . . . . . . . . . . . . . . . . . . 601
Lie . . . . . . . . . . . . . . . . . . . . . 44, 62, 477
Neumann-Liouville 36, 647, 647, 651
partial-wave . . . . . . . . . . . . . . 666, 700
perturbation
Schwinger-Dyson . . . . . . . . . . . . . 553
Robinson . . . . . . . . . . . . . . . . . 136, 977
Taylor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
weak-coupling . . . . . . . . . . . . . . . . . . 176
expectation value . . . . . . . . . . . . . . . . . . . 32
experiment
double-slit . . . . . . . . . . . . . . . . . . . . . . . 12
Stern-Gerlach . . . . . . . . . . . . . . . . . . 299
exponent
critical . . . . . . . . . . . . . . . . . . . . . . . . 1087
exponent, critical . . . . . 1037, 1210, 1213,
12181220
exponential integral . . . . . . . . . . . . . . . . 984
external
Coulomb potential . . . . . . . . . . . . . 481
current . . . . . . . . . . . . . . . . . . . . . . . . 587
potential . . . . . . . . . . . . . . . . . . . . . . . . 83

1412
factor
Boltzmann . . . . . . . . . . . . . . . . . 69, 933
Lande . . . . . . . . . . . . . . . . . . . . . . . . . 470
Faddeev, L.D. . . . . . . . . . . . . . 1400, 1401
Faddev-Popov determinant . . 1112, 1113
Fadeev, L.D. . . . . . . . . . . . . . . . . . . . . . 542
Fadeev-Popov . . . . . . . . . . . . . . . 1109, 1111
determinant . . . . . 967, 968, 970, 972
ghost . . . . . . . . . . . . . . . . . . . . . . . . . . 542
ghost field . . . . . . . . . . . . . . . . . . . . . . 971
Fadeev-Popov ghost . . . . . . . . . . . . . . . 1402
far zone . . . . . . . . . . . . . . . . . . . . . . . . . . . 1103
Faraday law . . . . . . . . . . . . . . . . . . . . . . . 273
Farley, F.J.M. . . . . . . . . . . . . . . . . . . . 921
Fermi
-Dirac distribution . . . . . . . . . . . . . 155
Fermi transitions . . . . . . . . . . . . . . . . . . 1364
Fermi gas
degenerate . . . . . . . . . . . . . . . . . . . . . 126
Fermi, E. . . . . . . . . . . . . . . . 520, 524, 1379
Fermi-Dirac
condition . . . . . . . . . . . . . . . . . . . . . . 524
fermions
Green function . . . . . . . . . . . . . . . . . 118
nonequilibrium Green functions 154
number of particles . . . . . . . . . . . . . 127
thermal Green function . . . . . . . . .778
Feshbach-Villar Hamiltonian . . . . . . . .321
Fetter, A. . . . . . . . . . . . . . . . . . . . . . . . . 174
Fetter, A.L. . .175, 194, 996, 998, 999,
1001
Feynman
diagram . . . . . . . . . . . . . . . . . . . . . . . . 755
disconnected . . . . . . . . . . . . . . . . . 757
gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 522
propagator . . . . . . . . . . . . . . . . . . . . 481
Feynman, R.P. 237, 315, 542, 921, 996,
999, 1261, 1279, 1400
field
quark . . . . . . . . . . . . . . . . . . . . . . . . 1301
acceleration . . . . . . . . . . . . . . . . . . . . 387
axial vector . . . . . . . . . . . . . . . . . . . . 279
collective . . . . . . . . . . . . . . . . 1137, 1139
composite . . . . . . . . . . . . . . . . . . . . . 1137
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . . 481
parity . . . . . . . . . . . . . . . . . . . . . . . . 510

Index
disorder . . . . . . . . . . . . . . . . . . .494, 756
electromagnetic . . . . . . . 386, 416, 421
fluctuations 923, 940, 944, 964, 965,
1015, 1087
ghost . . . . . . . . . . . . . . . . . . . . . . . . . . 273
Klein-Gordon . . . 246, 383, 481, 521,
559, 578, 589
local . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
mean . . . . . . . . . . . . . .1085, 1087, 1094
operator
Schr
odinger . . . . . . . . . . . . . . . . . . 268
order . . . . . . . . . . . . . . . . . . . . . . . . . . 494
pseudotensor . . . . . . . . . . . . . . . . . . 294
quantization . . . . . . . . . . . . . . . . . . . 104
Rarita-Schwinger . . . . . . . . . . . . . . . 368
scalar . . . . . . . . . . . . . . . . . . . . . 246, 257
Schr
odinger . . . . . . . . . . . . . . . . . . . . 462
tensor . . . . . . . . . . . . . . . . . . . . . . . . . 272
dual . . . . . . . . . . . . . . . . . . . . . . . . . 272
theory
disorder . . . . . . . . . . . . . . . . . . . . . 494
renormalizable . . . . . . . . . . . . . . . 781
transverse . . . . . . . . . . . . . . . . . . . . . 826
velocity . . . . . . . . . . . . . . . . . . . . . . . . 387
Filthuth, H. . . . . . . . . . . . . . . . . . . . . .1379
fine-structure constant . . . 470, 886, 889,
903, 920
finite-temperature
Green function . . . . . . . . . . . . . . . . . 143
propagator . . . . . . . . . . . . . . . . . . . . . 143
Fitch, V.L. . . . . . . . . . . . . . . . . . . . . . . 1379
fixed point . . . . . . . . . . . . . . . . . . 1209, 1215
Gaussian . . . . . . . . . . . . . . . . . . . . . . 1216
infrared stable . . . . . . . . . . . . . . . . 1215
range of attraction . . . . . . . . . . . . 1216
ultraviolet stable . . . . . . . . . . . . . . 1216
fixed-energy amplitude . . . . . . . . . . 47, 51
discontinuity . . . . . . . . . . . . . . . . . . . . 49
fixing gauge . . . . . . . . . . . . . . . . . . . . . . . .274
Flannery, B.P. . . . . . . . . . . . . . . . . . . . 175
fluctuation
-dissipation theorem . . . . . . . 158, 163
field . . 923, 940, 944, 964, 965, 1015,
1087
gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 965
part of Green function . . . . . . . . . 158

1413
quantum . . . . . . . . . . . . . . . . . . . . . . 149
vacuum . . . . . . . . . . . . . . . . . . . 467, 563
flux
quantum . . . . . . . . . . . . . . . . . . . . . . 1126
tube . . . . . . . . . . . . . . . . . . . . 1126, 1128
Fock, V. . . . . . . . . . . . . . . . . . . . . . . . . . . 383
force
Casimir . . . . . . . . . . . . . . . . . . . 469, 568
Lorentz . . . . . . . . . . . . . . . . . . . . . . . . 276
nonlocal . . . . . . . . . . . . . . . . . . . . . . . 835
tensor . . . . . . . . . . . . . . . . . . . . . . . . . 1381
van der Waals . . . . . . . . . . . . . . . . . . 568
Van-der-Waals . . . . . . . . . . . . 469, 563
form factor
charge . . . . . . . . . . . . . . . . . . . . . 886, 901
magnetic . . . . . . . . . . . . . . . . . . . . . . . 886
formalism
Hamilton . . . . . . . . . . . . . . . . . . . . . . . . . 3
Lagrange . . . . . . . . . . . . . . . . . . . . . . . . . 1
Lehmann-Symanzik-Zimmermann
770
Schwinger, proper-time . . . . . . . . . 483
formula
antenna . . . . . . . . . . . . . . . . . . . . . . . . 388
Baker-Campbell-Hausdorff . 44, 249,
369
Euler . . . . . . . . . . . . . . . . . . . . 599, 600
Euler-Maclaurin . . . . . . . . . . . . . . . 566
Fresnel integral . . . . . . . . . . . . . . . . . . 50
Gell-Mann Low . . . . . . . . . . . . 775, 782
Heron . . . . . . . . . . . . . . . . . . . . . . . . . . 377
LHuillier . . . . . . . . . . . . . . . . . . . . . . 377
Larmor . . . . . . . . . . . . . . . . . . . . . . . . 388
Lie expansion 44, 62, 252, 265, 304,
305, 322, 477
Mott . . . . . . . . . . . . . . . . . . . . . . . . . . 856
Poisson . . . . . . . . . . . . . . . . 29, 570, 572
reflection . . . . . . . . . . . . . . . . . . . . . . . 574
Rutherford . . . . . . . . . . . . . . . . . . . . . 856
Sochocki . . . . . . . . . . . . . . . . . . . .48, 491
four-curl . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
four-dimensional
angular momentum
total . . . . . . . . . . . . . . . . . . . . 278, 310
four-vector . . . . . . . . . . . . . . . . . . . . . . . . . 239
of matrices . . . . . . . . . . . . . . . . . . . . 302

Fourier
coefficients . . . . . . . . . . . . . . . . . . . . . . 30
transform . . . . . . . . . . . . . . . . 14, 29, 33
inverse . . . . . . . . . . . . . . . . . . . . 29, 46
4-point function . . . . . . . . . . . . . . . . . . . . 759
amputated . . . . . . . . . . . . . . . . . . . . . 759
Fowler, W.B. . . . . . . . . . . . . . . . . . . . 1379
hlich, H. . . . . . . . . . . . . . . . . . . . . . 237
Fro
Fradkin, E.S. . . . . . . . . . . . . . . . 175, 1400
frame
center-of-mass . . . 685, 686, 688, 689
Lorentz . . . . . . . . . . . . . . . . . . . . . . . . 240
frame-dragging effect . . . . . . . . . . . . . . .426
Fredkin, D.R. . . . . . . . . . . . . . . 998, 1001
free thermal Green function . . . . . . . . 950
free energy . . . . . . . . . . . . . . . . . . . . . . . . . . 71
free-particle
amplitude . . . . . . . . . . . . . . . . . . . . . . . 49
propagator . . . . . . . . . . . . . . . . 117, 118
Freire, P.C.C. . . . . . . . . . . . . . . . . . . . 427
French, J.B. . . . . . . . . . . . . . . . . . . . . . .921
frequency
Matsubara . . . . . . . . . . . . . . . . . . . . . 146
of wave . . . . . . . . . . . . . . . . . . . . . . . . . 12
Fresnel integral . . . . . . . . . . . . . . . . . 50, 924
Freundlich, Y. . . . . . . . . . . . . . . . . . . 1326
Fritzsch, H. . . . . . . . . . . . . . . . . . . . . . 1400
Fry, M.P. . . . . . . . . . . . . . . . . . . . . . . . . . 503
fugacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
expansion . . . . . . . . . . . . . . . . . . . . . .138
function
basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
Bessel . . . . . . . . . . . 51, 682, 703, 1101
modified . . . . . . . . . . . . . . . . . .51, 484
Beta . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
commutator . . . . . . . . . . . . . . . . . . . 470
cutoff . . . . . . . . . . . . . . . . . . . . . . . . . . 566
elliptic theta . . . . . . . . . . . . . . 138, 169
generalized zeta . . . . . . . . . . . . . . . . . 76
generating for canonical transformations . . . . . . . . . . . . . . . . . . . . . . . . . 9
Green . . . . . . . . . . . . . . . . . . . . .118, 119
Hankel . . . . . . . . . . . . . . . . . . . . . 52, 485
Heaviside . . . 45, 117, 488, 490, 507,
508, 518, 551, 552, 589, 596
hypergeometric . . . . . . . . . . . . . 65, 699

1414
confluent . . . . . . . . . . . . . . . . . . . . . 483
Kummer . . . . . . . . . . . . . . . . . . . . . . . 483
n-point . . . . . . . . . . . . . . . . . . . . . . . . 578
operator zeta . . . . . . . . . . . . . . . . . . . . 76
Polylogarithmic . . . . . . . . . . . . . . . . 170
Riemann zeta . . . . . . . . . . . . . . . . . . . 76
test . . . . . . . . . . . . . . . . . . . . . . . . . 25, 46
vertex . . . . . . . . . . . . . . . . . . . . . . . . . . 766
wave . . . . . . . . . . . . . . . . 12, 47, 48, 115
Wigner . . . . . . . . . . . . . . . . . . . . . . . . . 34
functional
delta . . . . . . . . . . . . . . . . . . . . . . . . . . . 924
determinant . . . . . . . . . . . . . . . . . . . 926
Fourier transform . . . . . . . . . . . . . . 927
gauge-fixing . . . . . . . . . . . . . . . . . . . . 965
generating . . . . . . . . . . 589, 751, 1140
integra
Gaussian . . . . . . . . . . . . . . . . . . . . . 927
integral . . . . . . . . . . . . . . . . . . . . . . . . 923
Gaussian . . . . . . . . . . . . . . . . . . . . 1139
imaginary-time . . . . . . . . . . . . . . . 932
matrices . . . . . . . . . . . . . . . . . . . . . . 1004
matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 40
vector notation . . . . . . . . . . . . . . . . . 927
vectors . . . . . . . . . . . . . . . . . . . . . . . . 1004
fundamental
charge . . . . . . . . . . . . . . . . . . . . . . 83, 888
flux . . . . . . . . . . . . . . . . . . . . . . . . . . . 1126
representation . . . . . . . . . . . . . . . . . 300
weight . . . . . . . . . . . . . . . . . . . . . . . . 1307
sken, S. . . . . . . . . . . . . . . . . . . . . . . . 1400
Gu
Gailey, J. . . . . . . . . . . . . . . . . . . . . . . . . .921
Galilei
invariance . . . . . . . . . . . . . . . . . . . . . . 657
transformation . . . . . . . . . . . . . . . . . 657
5 invariance . . . . . . . . . . . . . . . . . . . . . 1285
Gamow-Teller transitions . . . . 315, 1364
gap equation . . . . . . . . . . . 202, 1123, 1282
gas
ideal . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
of stars . . . . . . . . . . . . . . . . . . . . . . . . .407
Gasiorowicz, S. . . . . . . . . . . . . 780, 1338
Gasser, J. . . . . . . . . . . . . . . . . . . . . . . . 1401
Gastmans, R. . . . . . . . . . . . . . . . . . . . . . 921
gauge
axial . . . . . . . . . . . . . . . . . 275, 276, 281

Index
Coulomb . . . . . . . . .276, 281, 515, 519
Feynman . . . . . . . . . . . . . . . . . . . . . . 522
fixing . . . . . . . . . . . . . . . . . . . . . . . . . . 274
functional . . . . . . . . . . . . . . . . . . . . 965
term . . . . . . . . . . . . . . . . . . . . . . . . . 520
Hilbert . . . . . . . . . . . . . . . . . . . 294, 296
invariance
nonholonomic . . . . . . . . . . . . . . . . 830
Lorentz . . . . . . . . . 274, 295, 385, 544
radiation . . . . . . . . . . . . . . . . . . . . . . 275
transformation
second-kind . . . . . . . . . . . . . . . . . . 274
transverse . . . . . . . . . . . . . . . . . . . . . 283
Gauss
law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887
theorem . . . . . . . . . . . . . . . . . . . . . . . . 262
Gaussian
fixed point . . . . . . . . . . . . . . . . . . . . 1216
functional integral . . . . . . . . 927, 1139
integral . . . . . . . . . . . 50, 925, 926, 965
Gavazzi, G.M. . . . . . . . . . . . . . . . . . . . . . 81
Gegenbauer polynomials . . . . . . . . . . . . 699
Gelfand, I.M. . . . . . . . . . . . . . . . . . . . . . 81
Gell-Mann
Low formula . . . . . . . . . . . . . . .775, 782
SU(3) symmetry . . . . . . . . . . . . . . 1301
Gell-Mann, M. . . . . 194, 315, 651, 655,
695, 1400
general relativity
coordinate transformation . . . . . . 292
Einstein equation
free-field . . . . . . . . . . . . . . . . . . . . . 292
generalized
coordinates . . . . . . . . . . . . . . . . . . . . . . . 1
functions (distributions) . . . . . 25, 46
zeta function . . . . . . . . . . . . . . . . . . . . 76
generating
function for canonical transformations . . . . . . . . . . . . . . . . . . . . . . . . . 9
functional . . . . . . . . . . . 589, 751, 1140
generator . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
geodesic . . . . . . . . . . . . . . . . . . . . . . . . 11, 289
geodetic precession . . . . . . . . . . . . 425, 426
geometric quantization . . . . . . . . . . . . . . 81
geometry
Lobatschewski . . . . . . . . . . . . . . . . . 376

1415
Minkowski . . . . . . . . . . . . . . . . . . . . . 289
Geweniger, C. . . . . . . . . . . . . . . . . . . 1379
Ghandour, G. . . . . . . . . . . . . . . . . . . . . 806
ghost
dipole . . . . . . . . . . . . . . . . . . . . . . . . . 545
Fadeev-Popov . . . . . . . . . . . . . . . . . 542
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 273
ghosts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 973
Giachetti, R. . . . . . . . . . . . . . . . . . . . 1279
Giambagi, J. . . . . . . . . . . . . . . . . . . . . . . 920
Gibbs
distribution . . . . . . . . . . . . . . . . .69, 928
Giddings, S. . . . . . . . . . . . . . . . . . . . . . . 420
Gies, H. . . . . . . . . . . . . . . . . . . . . . . . . . . 1002
Gilman, F.J. . . . . . . . . . . . . . . . . . . . . . 1379
Ginzburg, V.I. . . . . . . . . . . . . . . . . . . 1003
Ginzburg, V.L. . . . . . . . . . . . . . . . . . . 1153
Ginzburg-Landau . . . . . . . . . . . . . . . . . 1148
Ginzburg-Landau theory . . . . 1107, 1116
Girard theorem . . . . . . . . . . . . . . . . . . . . .377
Gjesdal, S. . . . . . . . . . . . . . . . . . . . . . . 1379
Glashow, S.L. . . . . . . . . . . . . . . . . . . . 1379
global
conservation law . . . . . . . . . . 117, 658
representations . . . . . . . . . . . . . . . . . 301
Goldberger, M. . . . . . . . . . . . . . . . . . 651
Goldberger, M.L. . . . . . . . . . . . . . . . 705
Goldberger, W. . . . . . . . . . . . . . . . . . 427
golden
rule . . . . . . . . . . . . . . . . . . . . . . . . . . . .675
golden rule . . . . . . . . . . . . . . . . . . . . . . . . . 848
Goldstein, H. . . . . . . . . . . . . . . . . . . . . . 80
Goldstone bosons . . . . . . . . . . . . . . . . . 1120
Good, M.L. . . . . . . . . . . . . . . . . . . . . . . 1379
Good, R.H. . . . . . . . . . . . . . . . . . . . . . . 1379
Goom, D. . . . . . . . . . . . . . . . . . . . . . . . . 1379
Gordons decomposition formula . . . . 886
Gordon decomposition formula . . . . . 845
Gordon, W. . . . . . . . . . . . . .383, 920, 921
Gorkov, L.P. . . . . . . 174, 175, 996, 999
Gould, C.M. . . . . . . . . . . . . . . . . 998, 1001
Goursat, E. . . . . . . . . . . . . . . . . . . . . . . 803
Gradshteyn, I.S. . . . 51, 129, 170, 208,
226, 483, 484, 569, 575, 699, 701,
941, 984, 1003
grand-canonical

ensemble . . . . . . . . . . . . . . . . . . . . 71, 74
Hamiltonian . . . . . . . . . . . . . . . . . . . . 71
quantum-statistical partition function . . . . . . . . . . . . . . . . . . . . . . . . . 70
Grassmann
integral . . . . . . . . . . . . . . . . . . . . . . . . 935
variables . . . . . . . . . . . . 113, 324, 1307
gravitational constant . . . . . . . . . . . . . . 292
gravitons . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
gravity
Einstein-Hilbert action . . . . . . . . . 289
harmonic approximation . . . . . . . 291
Green
function
thermyl . . . . . . . . . . . . . . . . . . . . . . 143
Green function
bosons . . . . . . . . . . . . . . . . . . . . . . . . . 118
fermions . . . . . . . . . . . . . . . . . . . . . . . 118
Green function . . . . . . . . . . . . . . . . 118, 119
advanced . . . . . . . . . . . . . . . . . . . . . . . 155
antiperiodic . . . . . . . . . . . . . . . . . . . . 153
finite-temperature . . . . . . . . . . . . . . 143
4-point . . . . . . . . . . . . . . . . . . . . . . . . . 759
imaginary-time . . . . . . . . . . . . . . . . . 153
mnemonic rule . . . . . . . . . . . . . . . . . 472
n-point . . . . . . . . . . . . . . . . . . . . . . . . .748
real-time for Tungzero . . . . . 150, 153
retarded . . . . . . . . . . . . . . . . . . . . . . . 151
thermal
bosons . . . . . . . . . . . . . . . . . . . . . . . 778
time-ordered . . . . . . . . . . . . . . . . . . . 156
2-point . . . . . . . . . . . . . . . . . . . . . . . . . 757
Green, M.S. . . . . . . . . . . . . . . . . 996, 1000
Gross, D. . . . . . . . . . . . . . . . . . . . . . . . . 1326
Gross-Neveu model . . . . 1280, 1305, 1315
group
contraction . . . . . . . . . . . . . . . . . . . . 348
correspondence principle . . . . . . . . 58
Euclidean . . . . . . . . . . . . . . . . . . 58, 348
little . . . . . . . . . . . . . 242, 243, 341, 346
Lorentz
inhomogeneous . . . . . . . . . 264, 310
proper . . . . . . . . . . . . . . . . . . . . . . . 242
special . . . . . . . . . . . . . . . . . . 241, 256
orthogonal
special . . . . . . . . . . . . . . . . . . . . . . . 243

1416
permutation . . . . . . . . . . . . . . . . . . . . 165
Poincare . . . 264, 279, 293, 310, 475,
503, 504, 601
quantization . . . . . . . . . . . . . . . . . 58, 61
reflection
space . . . . . . . . . . . . . . . . . . . . . . . . 265
renormalization . . . . . . . . . . . . . . . . 802
representation . . . . . . . . . . . . 299, 657
rotation . . . . . . . . . . . . . . . . . . . . . . . . 240
symmetric . . . . . . . . . . . . . . . . . . . . . 165
symmetry . . . . . . . . . . . . . . . . . . . . . . 645
Wigner . . . . . . . . . . . . . . . . . . . . . . . . 242
Gubernatis, J.E. . . . . . . . . . . . . . . . . . 175
Guida, R. . . . . . . 1260, 1261, 1278, 1279
Gupta, S.N. . . . . . . . . . . . . . . . . . . . . . . .538
Gupta-Bleuler . . . . . . . . . . . . 837, 965, 968
condition . . . . . . . . . . . . . . . . . . . . . . . 538
quantization . . . . . . . . . . 539, 540, 542
theory . . . . . . . . . . . . . . . . . . . . . . . . . 538
Gupta-Bleuler formalism . . 963, 968, 973
Gupta-Bleuler quantization . . . . . . . . . 826
Gupta-Bleuler subsidiary condition . 973
Guralnik, G.S. . . . . . . . . . . . . . . . . . . 1326
gyromagnetic ratio . . 470, 472, 475, 481,
499, 843, 846
H. Press, W. . . . . . . . . . . . . . . . . . . . . . 407
H. Price, R. . . . . . . . . . . . . . . . . . . . . . . 407
hler, G. . . . . . . . . . . . . . . . . . . . . . . . 383
Ho
Habert, F.J. . . . . . . . . . . . . . . . . . . . . .1379
hadrons . . . . . . . . . . . . . . . . . . . . . 1286, 1380
Hamel, G. . . . . . . . . . . . . . . . . . . . . . . . . . .80
Hamermesh, M. . . . . . . . . . . . . . . . . . . 174
Hamilton
-Jacobi differential equation . . . . . 10
equation of motion . . . . . . . . . 3, 4, 42
formalism . . . . . . . . . . . . . . . . . . . . . . . . 3
Hamiltonian . . . . . . . . . . . . . . . . 2, 649, 656
Coulomb . . . . . . . . . . . . . . . . . . . . . . . . 16
density . . . . . . . . . . . . . . . . . . . . . . . . 102
Feshbach-Villar . . . . . . . . . . . . . . . . 321
grand-canonical . . . . . . . . . . . . . . . . . 71
interaction . . . . . . . . . . . . . . . . . . . . . 553
local . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
operator . . . . . . . . . . . . . . . . . . . . . . . 657
second-quantized . . . . . . . . . . . . . . . . 96
unperturbed . . . . . . . . . . . . . . . . . . . 644

Index
Hankel function . . . . . . . . . . . . . . . . . . . . 485
Hankel function . . . . . . . . . . . . . . . . .52, 485
hard-photon regime . . . . . . . . . . . . . . . . 907
harmonic
approximation to gravity . . . . . . . 291
hyperspherical . . . . . . . . . . . . . . . . . 698
spherical . . . . . . . . . . 60, 358, 381, 666
spinor . . . . . . . . . . . . . . . . . . . . . . . . 502
Harriman, J.M. . . . . . . . . . . . . . . . . . . 922
Harrison, G. . . . . . . . . . . . . . . . . . . . . 1401
Hartree-Fock-Bogoljubov approxiation
1021
Hayward, R.W. . . . . . . . . . . . . 314, 1379
Heaviside function 45, 45, 117, 488, 490,
507, 508, 518, 551, 552, 589, 596
Hedrick, E.R. . . . . . . . . . . . . . . . . . . . . 803
Heisenberg
-Euler Lagrangian . . . . . . . . . . . . . . 974
correspondence principle . . . . 41, 42
equation . . . . . . . . . . . . . . . . . . . . . . . 470
equation of motion . . . . . . . . . 42, 656
matrices . . . . . . . . . . . . . . . . . . . . . 4042
model . . . . . . . . . . . . . . . . . . . . . . . . . 1120
operator . . . . . . . . . . . . . . . . . . . . . . . . 41
picture . . . . 40, 41, 42, 109, 150, 645
in nonequilibrium theory 151, 160
state . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
uncertainty principle . . . . . . . . . . . . 14
Heisenberg, W. . . . . . . . . 524, 545, 1002
helicity . . . . . . . . . . . . . . . . . . . . . . . . 338, 349
bispinor . . . . . . . . . . . . . . . . . . . . . . . 340
matrix . . . . . . . . . . . . . . . . . . . . . . . . . 283
spinor . . . . . . . . . . . . . . . . . . . . . . . . . 339
helium
liquid . . . . . . . . . . . . . . . . . . . . . . . . . . 134
superfluid . . . . . . . . . . . . . . . . . . . . . . 138
Helmholtz equation . . . . . . . . . . . . . . . . 564
Henneaux, M. . . . . . . . . . . . . . . . . . . . . 384
Hermitian
-adjoint operator . . . . . . . . . . . . . . . . 17
conjugate . . . . . . . . . . . . . . . . . . . . . . . 32
operator . . . . . . . . . . . . . . . . . . . . . . . . 17
Heron formula . . . . . . . . . . . . . . . . . . . . . . 377
Hessian metric . . . . . . . . . 2, 15, 55, 66, 79
Hibbs, A.R. . . . . . . . . . . . . . . . . . . . . . . . 999
Higuchi, A. . . . . . . . . . . . . . . . . . . . . . . . 602

1417
Hikami, S. . . . . . . . . . . . .1260, 1277, 1278
Hilbert
gauge . . . . . . . . . . . . . . . . . . . . . 294, 296
space . . . . . . . . . . . . . . . . . . . . . . . . 16, 18
second-quantized . . . . . . . . . . . . . . 95
Ho, T.L. . . . . . . . . . . . . . . . . . . . . 998, 1001
Hye, J.S. . . . . . . . . . . . . . . . . . . . . . . . . 601
hole
black . . . . . . . . . . . . . . . . . . . . . . . . . . . 402
positron . . . . . . . . . . . . . . . . . . . . . . . . 499
Hoppes, D.D. . . . . . . . . . . . . . . . 314, 1379
Huang, K. . . . . . . . . . . . . . . . 174, 237, 680
Huang, M. . . . . . . . . . . . . . . . . . . . . . . . 1327
Hubbard, J. . . . . . . . . . . . . 996, 999, 1160
Hubbard-Stratonovic transformation 956
Hubbard-Stratonovich
transformation
937
Hudson, R.P. . . . . . . . . . . . . . . . 314, 1379
Hughes, R.J. . . . . . . . . . . . . . . . . . . . . 1403
Hulet, R.G. . . . . . . . . . . . . . . . . . 174, 237
Hurwitz
zeta function . . . . . . . . . . . . . . . . . . 575
hydrodynamic limit . . . . . . . . . . . . . . . 1092
hydrogen atom . . . . . . . . . . . . . . . . . . . . . 848
hypercharge
weak . . . . . . . . . . . . . . . . . . . . . . . . . . 1371
hyperfine-splitting . . . . . . . . . . . . . . . . . 488
hypergeometric functions . . . . . . . 65, 699
confluent . . . . . . . . . . . . . . . . . . . . . . . 483
ideal gas . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
identity
Bianchi . . . . . . . . . 272, 292, 683, 964
Jacobi . . . . . . . . . . . . . . . . . . . . . . . 4, 250
Ward . . . . . . . . . . . . . . . . . . . . . . 797, 842
Ward-Takahashi . . . . . 840, 842, 1092
identity permutation . . . . . . . . . . . . . . . 165
Idziaszek, Z. . . . . . . . . . . . . . . . . . . . . . . 680
i-prescription . . . . . . . . . . . . . . . . . . . . . . . 48
Iliopolous, J. . . . . . . . . . . . . . . . . . . . 1379
imaginary-time
functional integral . . . . . . . . . . . . . . 932
Green function . . . . . . . . . . . . . . . . . 153
Wick rotation . . . . . . . . . . . . . . . . . . 482
Imamura, T. . . . . . . . . . . . . . . . . . . . . . . 531
impact parameter . . . . . . . . . . . . . . . . . . 852
inequality

for nonequilibrium Green functions


159
infrared-stable fixed point . . . . . . . . . 1215
inhomogeneous
Lorentz group . . . . . . . . . . . . . . . . . 310
inhomogeneous Lorentz group . . . . . . 264
initial- and final-state interactions . 691,
692
insertions
mass . . . . . . . . . . . . . . . . . . . . . . . . . . . 789
wave function . . . . . . . . . . . . . . . . . . 789
instanton . . . . . . . . . . . . . . . . . . . . . . . . . .1097
integrability condition
Schwarz . . . . . . . . . . . 6, 272, 274, 292
integral
exponential . . . . . . . . . . . . . . . . . . . . 984
Fresnel . . . . . . . . . . . . . . . . . . . . . 50, 924
functional . . . . . . . . . . . . . . . . . . . . . . 923
Gaussian . . . . . . . . . . . . . . . . . . . . 1139
Gaussian . . . . . . . . . . . . . . 50, 925, 965
Grassmann . . . . . . . . . . . . . . . . . . . . . 935
loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . 769
path . . . . . . . . . . . . . . . . . . . . . . . . . . . 937
principal-value . . . . . . . . . . . . . . 48, 158
integral rule, Veltman . . . . 574, 799, 822
interaction
Hamiltonian . . . . . . . . . . . . . . . . . . . .553
picture . . . . . . . . . . . . . . . . . . . . . . . . 646
picture (Dirac) . . . . . . . . . . . . . . . . . . 43
interactions
initial- and final-state . . . . . . . . . . 692
weak . . . . . . . . . . . . . . . . . . . . 1362, 1371
internal
symmetry . . . . . . . . . . . . . . . . . . . . . 677
internal energy . . . . . . . . . . . . . . . . . . . . . . 71
interraction
Coulomb . . . . . . . . . . . . . . . . . . . . . . . 837
intrinsic
angular momentum . . . . . . . . . . . . 242
intrinsic parity . . . . . . . . . . . . . . . . . . . . . 266
invariance
5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1285
Galilei . . . . . . . . . . . . . . . . . . . . . . . . . 657
isospin . . . . . . . . . . . . . . . . . . . . . . . 1292
rotationally . . . . . . . . . . . . . . . . . . . . 650
translational . . . . . . . . . . . . . . . . . . . 650

1418
invariant representation space rotation
group . . . . . . . . . . . . . . . . . . . . . . 352
inverse
Fourier transform . . . . . . . . . . . . . . . 46
inversion
space . . 255, 265, 311, 312, 314, 315,
364
time . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
irreducible
one-particle . . . . . . . . . . . . . . . . . . . . 765
Ising
model . . . . . . . . . . . . . . . . . . 1099, 1120
Ising like . . . . . . . . . . . . . . . . . . . . . . . . . . 1219
isospace
weak . . . . . . . . . . . . . . . . . . . . . . . . . . 1372
isospin invariance . . . . . . . . . . . . . . . . . 1292
isospinor . . . . . . . . . . . . . . . . . . . . . . . . . . 1291
isotopic spin . . . . . . . . . . . . . . . . . . . . . . 1291
isotopic spin symmetry . . . . . . . . . . . 1292
Itzykson, C. 503, 920, 1002, 1029, 1049
Jack, I. . . . . . . . . . . . . . . . . . . . . 1259, 1278
Jackiw, R. . 920, 921, 1029, 1049, 1160,
1326
Jackson, J.D. . . . . . . . . . . . 479, 684, 920
Jackson, R. . . . . . . . . . . . . . . . . . . . . . . .420
Jacob, M. . . . . . . . . . . . . . . . . . . . . . . . . . 601
Jacob-Wick expansion . . . . . . . . . . . . . 601
Jacobi
identity . . . . . . . . . . . . . . . . . . . . . . 4, 250
polynomials . . . . . . . . . . . . . . . . 65, 699
Janke, W. . . . . . . . . . . . . . . . . . 1261, 1279
Jegerlehner, F. . . . . . . . . . . . . 921, 1400
Jentschura, U.D. . . . . . . . . . . . . . . . 1002
Johnson, K. . . . . . . . . . . . . . . . . . . . . . 1326
Jona Lasinio, G. . . . . . . . . . . . . . . . . 1326
Jona-Lasinio, G. . . . . . . . . . . . . . . . . 1029
Jones, D.R.T. . . .997, 1000, 1259, 1278
Jones, H.F. . . . . . . . . . . . . . . . . . . . . . . 1279
Joos, H. . . . . . . . . . . . . . . . . . . . . . . . . . . . 601
Jordan rule . . . . . . . . . . . . . . . . . . . . . . . . . 15
Joshi, B.C. . . . . . . . . . . . . . . . . . . . . . . . .427
K-meson, oscillations . . . . . . . . . . . . . .1376
Kac, M. . . . . . . . . . . . . . . . . . . . . . . . . . . 1122
Kadanoff, L. . . . . . . . . . . . . . . . 996, 1000
Kadanoff, L.P. . . . . . . . . . . . . . .996, 999

Index
Kalogera, V. . . . . . . . . . . . . . . . . . . . . 427
Kalos, M.H. . . . . . . . . . . . . . . . . . . . . . . 238
Kamenshchik, A.Y. . . . . . . . . . . . . . . 601
Kamo, H. . . . . . . . . . . . . . . . . . . . . . . . . . . .81
Kapka, A. . . . . . . . . . . . . . . . . . . . . . . . . 112
Kastening
, B. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
Kastening, B. . . . . . . . . . . . . . 1050, 1052
Kaup, D.J. . . . . . . . . . . . . . . . . . . 999, 1002
Kawai, T. . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Kay, I. . . . . . . . . . . . . . . . . . . . . . . 999, 1002
Kazanskii, A.K. . . . . . . . . . . . . . . . . . 1029
Ketterle, W. . . . . . . . . . . . . . . . 174, 237
Kim, C. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 427
Kim, W.J. . . . . . . . . . . . . . . . . . . . . . . . . . 601
kink . . . . . . . . . . . . . . . . . . . . . . . . . 1097, 1099
Kinoshita, T. . . . . . . . . . . . 543, 920, 921
Kirsten, K. . . . . . . . . . . . . . . . . . . . . . . . 601
Kittel, C. . . . . . . . . . . . . . . . . . . . . . . . . 194
Kivel, N.A. . . . . . . . . . . . . . . . . 1259, 1278
Klein, O. . . . . . . . . . . . . . . . . . . . . . . . . . 383
Klein-Gordon
equation . . . 261, 274, 295, 317, 463,
471, 487, 522, 541, 544, 547
in Coulomb potential . . . . . . . . . 484
field . . .246, 383, 481, 521, 559, 578,
589
Klein-Gordon equation
Coulomb potential . . . . . . . . . . . . . 484
Klein-Nishina cross section . . . . . . . . . 875
Klein-Nishina formula . . . . . . . . . . . . . . 862
Kleinert, A. . . . . . . . . . . . . . . . . . . . . . . .vii
Kleinert, H. . . . . . . . . . . . . . . . . . . vii, 68,
80, 81, 174, 237, 238, 427, 490,
493, 494, 503, 568, 571, 601, 674,
685, 705, 780, 824, 920, 921, 923,
977, 996, 998, 1000, 1001, 1029,
1049, 1050, 1052, 1099, 1106,
1121, 1132, 1136, 1150, 1153,
1154, 1160, 12601262, 1277
1280, 1312, 1313, 1315, 1326,
1401
Kleinknecht, K. . . . . . . . . . . . . . . . . 1379
Kleman, M. . . . . . . . . . . . . . . . . . 998, 1002
Konishi, K. . . . . . . . . . . . . . . . . 1261, 1279
Konopleva, N.P. . . . . . . . . . . . . . . . . 1400

1419
Kopnin, N.B. . . . . . . . . . . . . . . . 998, 1002
Kramer, M. . . . . . . . . . . . . . . . . . . . . . . 427
Kroll, N.M. . . . . . . . . . . . . . . . . . . . . . . 921
Krusius, N. . . . . . . . . . . . . . . . . . 998, 1001
Kubo, R. . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Kuchar, K. . . . . . . . . . . . . . . . . . . . . . . . 420
Kumar, P. . . . . . . . .997, 999, 1001, 1002
Kummer function . . . . . . . . . . . . . . . . . . 483
Kurasuji, H. . . . . . . . . . . . . . . . . . . . . .1029
Kurn, D.M. . . . . . . . . . . . . . . . . . . 174, 237
Kuroda, Y. . . . . . . . . . . . . . . . . . 998, 1001
Kusius, M. . . . . . . . . . . . . . . . . . . 998, 1001
LHuillier formula . . . . . . . . . . . . . . . . . . 377
th, V. . . . . . . . . . . . . . . . . . . . . . . . . . 1379
Lu
laboratory
cross section . . . . . . . . . . . . . . . . . . . 686
momentum . . . . . . . . . . . . . . . . . . . . . 686
Lagrange
brackets . . . . . . . . . . . . . . . . . . . . . . . 7, 7
formalism . . . . . . . . . . . . . . . . . . . . . . . . 1
Lagrange, J.L. . . . . . . . . . . . . . . . . . . . . 80
Lagrangian . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
density . . . . . . . . . . . . . . . 106, 258, 462
Heisenberg-Euler . . . . . . . . . . . . . . . 974
Laguerre
polynomials . . . . . . . . . . . . . . . . . . . . 483
Lamb
shift . . . . . . . . . . . . . . . . . . . . . . . . . . . 487
Lamb constant . . . . . . . . . . . . . . . . . . . . . 905
Lamb shift . 888, 896, 903, 905907, 914,
917
Lamb, W.E. . . . . . . . . . . . . . . . . . . 921, 922
lambda transition . . . . . . . . . . . . . . . . . 1133
Lamm, D.R. . . . . . . . . . . . . . . . . . . . . . . 1002
Lamoreaux, S.K. . . . . . . . . . . . . . . . . . 601
Lande factor . . . . . . . . . . . . . . . . . . . . . . . 470
Land`e factor . . . . . . . . . . . . . . . . . . . . . . . 846
Landau, L.D. . . . .80, 81, 174, 187, 483,
513, 664, 684, 999, 1002, 1153
Landau-Pitajevski theory . . . . . . . . . .1100
Lane, K. . . . . . . . . . . . . . . . . . . . . . . . . . 1326
Langacker, P. . . . . . . . . . . . . . . . . . . 1326
Langer, J. . . . . . . . . . . . . . 996, 999, 1160
Laplace
-Beltrami operator 55, 57, 58, 61, 67
Laplace-Beltrami . . . . . . . . . . . . . . . . . . 1273

Laplacian . . . . . . . . . . . . . . . . . . . . 53, 54, 58


canonical . . . . . . . . . . . . . . . . . . . . . . . . 57
largest subgroup . . . . . . . . . . . . . . . . . . . .242
Larin, S.A. . . . . . .824, 1052, 1260, 1278
Larkin, A.I. . . . . . . . . . . . . . . . . . . . . . 1326
Larmor
formula . . . . . . . . . . . . . . . . . . . . . . . . 388
Larmors formula . . . . . . . . . . . . . . . . . . . 858
lattice
Wigner . . . . . . . . . . . . . . . . . . . . . . . . . 194
Lautrup, B.E. . . . . . . . . . . . . . . . . . . . . 921
law
Amp`ere . . . . . . . . . . . . . . . . . . . . . . . .273
Ampere . . . . . . . . . . . . . . 479, 519, 832
conservation . . . . . . . . . . . . . . 645, 833
local . . . . . . . . . . . . . . . . . . . . . . . . . 661
continuity . . . . . . . . . . . . . . . . . . . . . . . 18
Coulomb . . . 273, 275, 285, 479, 515,
825, 832
current conservation . . . . . . . . 18, 262
Dulong-Petit . . . . . . . . . . . . . . . . . . . 125
Dulong-Petit law . . . . . . . . . . . . . . 137
energy conservation . . . . . . . . . . . . . 14
Faraday . . . . . . . . . . . . . . . . . . . . . . . 273
Gauss . . . . . . . . . . . . . . . . . . . . . . . . . . 887
minimal substitution 466, 469, 476,
829
Newtons third . . . . . . . . . . . . . . . . . . 83
probability conservation . . . . . . . . . 16
Stefan-Boltzmann . . . . . . . 1146, 1325
Layzer, A.J. . . . . . . . . . . . . . . . . . . . . . . 921
Le Guillou, J.C. . . . . . . . . . . . 996, 1000
Lebesgue
lemma . . . . . . . . . . . . . . . . . . . . . . . . . 821
Lee, D.M. . . . . . . . . 997, 998, 1000, 1001
Lee, M.D. . . . . . . . . . . . . . . . . . . . . . . . . . 705
Lee, T.D. . . . . . . . . . . 174, 237, 314, 1400
left-handed
neutrino . . . . . . . . . . . . . . . . . . . . . . . 340
Legendre
polynomial . . . . . . . . . . . . . . . . . . . . . 358
polynomials . . . . . . . . . . . . . . . . . . . 382
transform . . . . . . . . . . . . . . . . . . . . . . 469
Legendre transform . . . . . . . . . . . . . . . . .828
Legendre polynomial
associated . . . . . . . . . . . . . . . . . . . . . 381

1420
Legendre transform . . 466, 498, 550, 649
Leggett, A.L. . . . . . . . . . . . . . . 997, 1000
LeGuillou, J.C. . . . . . . . . . . . . . . . . . 1277
Lehmann-Symanzik-Zimmermann
formalism . . . . . . . . . . . . . . . . . . 770
Leibnitz rule
for operators . . . . . . . . . . . . . . . . . . . . 96
Lemma
Riemann-Lebesgue . . . . . . . . 260, 383
Schur . . . . . . . . . . . . . . . . . . . . . . . . 1296
lemma
Lebesgue . . . . . . . . . . . . . . . . . . . . . . . 821
Wick . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
length
Planck . . . . . . . . . . . . . . . . . . . . 292, 468
scattering . . . . . . . . . . . . . . . . . 670, 680
thermal . . . . . . . . . . . . . . . . . . . . . . . . 123
Lense-Thirring effect . . . . . . . . . . . . . . . 426
Leplae, L. . . . . . . . . . . . . . . . . . . . . . . . 1326
lepton number conservation laws . . 1366
Leutwyler, H. . . . . . . . 420, 1400, 1401
level
shift formula . . . . . . . . . . . . . . . . . . . 657
level splitting
hyperfine . . . . . . . . . . . . . . . . . . . . . . 488
Levesque, D. . . . . . . . . . . . . . . . . . . . . . 238
Levi-Civita
symbol . . . . . . . . . . . . . . . . . . . . . . . . . 378
Levi-Civita tensor . . . . . . . . . . . . . . . . . . 247
Levine, M.J. . . . . . . . . . . . . . . . . . . . . . . 921
Levinson theorem . . . . . . . . . . . . . . . . . . 671
Levinson, N. . . . . . . . . . . . . . . . . . . . . . . 705
Lienard-Wiechert potential . . . . 386, 490
Lie
algebra . . . . . . . . . . . . . . . . 58, 249, 251
representation . . . . . . . . . . . 251, 300
rotation group . . . . . . . . . . . . . . . . 59
expansion formula . 44, 62, 252, 265,
304, 305, 322, 477
Lifshitz, E.M. . . 80, 81, 174, 483, 664,
684, 999, 1002
Lifshitz, Z. M. . . . . . . . . . . . . . . . . . . . 187
light
cone . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
velocity . . . . . . . . . . . . . . . . . . . . . . . . . 13
Lim, P.H. . . . . . . . . . . . . . . . . . . . . . . . . . 395

Index
limit
hydrodynamic . . . . . . . . . . . . . . . . . 1092
thermodynamic . . . . . . . . . . . . . . . . 768
line
defect . . . . . . . . . . . . . . . . . . . . . 483, 756
vortex . . . . 138, 483, 756, 1126, 1127
linear
response theory . . . . . . . . . . . 150, 152
space . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Liotta, R. . . . . . . . . . . . . . . . . . . 998, 1001
Liouville
Wigner equation . . . . . . . . . . . . . . . . 34
Liouville equation . . . . . . . . . . . . . . . . . . . 34
Lippmann-Schwinger equation . 651653,
660664, 668, 673, 680, 701
liquid helium . . . . . . . . . . . . . . . . . . . . . . . 134
little
group . . . . . . . . . . . . . . . . . . . . . . . . . . 346
little group . . . . . . . . . . . . . . . 242, 243, 341
Loar, H. . . . . . . . . . . . . . . . . . . . . . . . . . . 693
Lobatschewski Geometry . . . . . . . . . . . 376
local
action . . . . . . . . . . . . . . . . . . . . 106, 257
basis functions . . . . . . . . . . . . . . . . . . 19
conservation law . 18, 116, 658, 661
field theory . . . . . . . . . . . . . . . . . . . . .257
Hamiltonian . . . . . . . . . . . . . . . . . . . 102
oscillator algebra . . . . . . . . . . . . . . . . 93
symmetry
transformation . . . . . . . . . . 647, 659
transformation . . . . . . . . . . . . . . 4, 647
local Lagrangian . . . . . . . . . . . . . . . . . . . 462
locality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
longitudinal
polarization vector . . . . . . . 286, 287
loop
integral . . . . . . . . . . . . . . . . . . . . . . . . 769
Lorentz
covariant matrices . . . . . . . . . . . . . 302
force . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
frame . . . . . . . . . . . . . . . . . . . . . . . . . . 240
gauge . . . . . . . . . . . 274, 295, 385, 544
group
inhomogeneous . . . . . . . . . . . . . . 264
proper . . . . . . . . . . . . . . . . . . . . . . . 242
special . . . . . . . . . . . . . . . . . . 241, 256

1421
transformations . . . . . . . . . . . . . . . . 657
proper . . . . . . . . . . . . . . . . . . . . . . . 242
Lorentz group . . . . . . . . . . . . . . . . . . . . . . 278
Lorentz group
inhomogeneous . . . . . . . . . . . . . . . . 310
Lorentz transformations
bispinors . . . . . . . . . . . . . . . . . . . . . . . 308
Lorimer, D.R. . . . . . . . . . . . . . . . . . . . . 427
Love, A. . . . . . . . . . . . . . . . . . . . . 997, 1000
Low, F. . . . . . . . . . . . . . . . . . . . . . . . . . . . 655
LSZ reduction formulas . . . . . . . . . . . . . 770
luminosity . . . . . . . . . . . . . . . . . . . . . . . . . . 689
Lurie, D. . . . . . . . . . . . . . . . . . . . . . . . . 1326
Luttinger, J.M. . . . . . . . . . . . . . . . . . . 237
Lyne, A.G. . . . . . . . . . . . . . . . . . . . . . . . . 427
M. Ryzhik, I. . . . . . . . . . . . . . . . . . . . . . 941
nnel, M. . . . . . . . . . . . . . . . . . . . . . . . 238
Ma
nnigmann, M. . . . . . . . . . . . . . . . . 1052
Mo
hlschlegel, B. . . . . . . . . . . . 996, 999
Mu
Mller operator . . . . . . . . . . . . . . . . . . . . 649
Mller, C. . . . . . . . . . . . . . . . . . . . . . . . 649
Macfarlane, A. . . . . . . . . . . . . . . . . . . 601
Machleidt, R. . . . . . . . . . . . . . . . . . . .1338
Macke, W. . . . . . . . . . . . . . . . . . . . . . . . .194
magnetic
form factor . . . . . . . . . . . . . . . . . . . . . 886
magnetic enthalpy . . . . . . . . . . . . . . . . . 1125
magnetic moment
anomalous . . . . . . . . . . . . 472, 481, 881
magnetic trap, Bose-Einstein condensation . . . . . . . . . . . . . . . . . . . . . . . . 141
magnetization
critical . . . . . . . . . . . . . . . . . . . . . . . . 1220
spontaneous . . . . . . . . . . . . . 1120, 1221
magneton, Bohr . . . . . . . . . . . . . . . 469, 843
Maiani, L. . . . . . . . . . . . . . . . . . . . . . . . 1379
Majorana
matrices . . . . . . . . . . . . . . . . . . . 328, 329
representation . . . . . . . . . . . . . . . . . 328
spinor . . . . . . . . . . . . . . . . . . . . . . . . . 329
Majorana potential . . . . . . . . . . . . . . . . 1381
Majorana spinor . . . . . . . . . . . . . . . . . . 1401
Maki, K. . . . . . . . . . . . . . . . . . . . . 9971002
Manchester, R.N. . . . . . . . . . . . . . . . 427
Mancini, F. . . . . . . . . . . . . . . . . . . . . . . 1326
Mandelstam

variables . . . . . . . . . . . . . . 688, 877, 879


Mandelstam triangle . . . . . . . . . . . . . . . 688
Mandelstam, S. . . . . . . . . . . . . . . . . . 1400
many-body
system . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Marciano, W.J. . . . . . . . . . . . . 921, 1379
Marr, L. . . . . . . . . . . . . . . . . . . . . . . . . . . 601
Marsden, J.E. . . . . . . . . . . . . . . . . . . . . . 80
Marshak, R.E. . . . . . . . . . . . . . . 315, 693
Martin, P.C. . . . . . . . . . . . . . . . . . . . . 1029
mass
insertions . . . . . . . . . . . . . . . . . . . . . . 789
Planck . . . . . . . . . . . . . . . . . . . . . . . . . 292
reduced . . . . . . . . . . . . . . . . . . . . . . . . 680
shell . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
mass matrix
Cabibbo . . . . . . . . . . . . . . . . . . . . . . 1368
Cabibbo, Kobayashi, Maskawa 1378
material waves . . . . . . . . . . . . . . . . . . . . . . 11
Mathews, J. . . . . . . . . . . . . . . . . . . . . . . 427
Mathews, P.T. . . . . . . . . . . . . . . 996, 999
matrices
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . . 307
four-vector of . . . . . . . . . . . . . . . . . . 302
functional . . . . . . . . . . . . . . . . . . . . . 1004
Lorentz covariant . . . . . . . . . . . . . . 302
Majorana . . . . . . . . . . . . . . . . . 328, 329
Pauli . . . . . . . . . . . . . . . . . . . . . . . . . . 301
matrix
S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 650
T . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
chirality . . . . . . . . . . . . . . . . . . 317, 340
density . . . . . . . . . . . . . . . . . . . . . . . . . . 34
functional . . . . . . . . . . . . . . . . . . . . . . . 40
Heisenberg . . . . . . . . . . . . . . . . . . . 4042
helicity . . . . . . . . . . . . . . . . . . . . . . . . 283
Hessian . . . . . . . . . . . . 2, 15, 55, 66, 79
Pauli . . . . . . . . . . . . . . . . . . . . . . 302, 304
Pauli spin . . . . . . . . . . . . . . . . . . . . . . . 64
representation . . . . . . . . . . . . . . . . . 300
scattering . . . . . . . . . . . . . . . . . . . . . . 650
symplectic unit . . . . . . . . . . . . . . . . . . . 7
matrix element
reduced . . . . . . . . . . . . . . . . . . . . . . . . 361
Matsas, G.E.A. . . . . . . . . . . . . . . . . . . 602
Matsen, R.P. . . . . . . . . . . . . . . . . . . . . 1379

1422
Matsubara
sum . . . . . . . . . . . . . . . . . . . . . . . . . . . 1137
Matsubara frequencies . . . . . . . . . . . . . . 941
Matsubara frequencies . . . . . . . . . . . . . 146
Matsumoto, H. . . . . . . . . . . . . . . . . . 1326
Matthews, M.R. . . . . . . . . . . . . 174, 237
maximal parity violation in -decay 314
Maxwell
equation . . . . . . . . . . . . . . . . . . .564, 871
equations . . . . . . . . . . . . . 320, 467, 479
Maxwell equations . . . . . . . . . . . . . 827, 832
Maxwells field equations . . . . . . . . . . . 832
McKane, A. . . . . . . . . . . . . . . . . . . . . . 1277
McLaughlin, M.A. . . . . . . . . . . . . . . . 427
mean field
approximation . . . . 1085, 1087, 1094
mechanics
classical . . . . . . . . . . . . . . . . . . . . . . . . . . 1
quantum . . . . . . . . . . . . . . . . . . . . . . . . 11
quantum-statistical . . . . . . . . . . . . . . 68
statistical . . . . . . . . . . . . . . . . . . . . . . . 68
Meissner effect . . . . . . . . . . . . . . . . . . . . 1375
Meissner, U-G. . . . . . . . . . . . . . . . . . . . 383
Mermin, D. . . . . . . . . . . . . . . . . . 175, 1160
Mermin, N.D. 959, 997, 998, 1000, 1001
Mermin-Wagner theorem . . . . . . . . . . 1307
Mermin-Wagner theorem . . . . . . . . . . 1153
Merzbacher, E. . . . . . . . . . . . . . . . . . . . 80
meson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
mesons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1380
Messiah, A. . . . . . . . . . . . . . . . . . . . . . . . . 80
metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
dynamical . . . . . . . . . . . . . . . . . . . . . . 106
Hessian . . . . . . . . . . . . 2, 15, 55, 66, 79
Minkowski . . . . . . . . . . . . . . . . . . . . . 243
tensor . . . . . . . . . . . . . . . . . . . . . . . . . . .54
Mewes, M.-O. . . . . . . . . . . . . . . . 174, 237
Mexican hat potential . . . . . . . . . . . . . . 142
Michel, L. . . . . . . . . . . . . . . . . . . . . . . . . 474
Miller, A.I. . . . . . . . . . . . . . . . . . . . . . 1002
Milonni, P.W. . . . . . . . . . . . . . . . . . . . . 601
Milton, K.A. . . . . . . . . . . . . . . . . . . . . . 601
Mineev, V.P. . . . . . . . . . . . . . . . 998, 1002
minimal
coupling . . . . . . . . . . . . . . . . . . 830, 1108
substitution . . . . . . . . . . 466, 469, 476

Index
substitution rule . . . . . . . . . 830, 1108
subtraction scheme . . . . . . . . . . . . 1202
minimal substitution . . . . . . . . . . . . . . . 829
Minkowski
geometry . . . . . . . . . . . . . . . . . . . . . . 289
metric . . . . . . . . . . . . . . . . . . . . . . . . . .243
space . . . . . . . . . . . . . . . . . . . . . . . . . . 243
mirror
reflection . . . . . . . . . . . . . . . . . . . . . . . 255
Mitter, H. . . . . . . . . . . . . . . . . . . . . . . . . 920
mixed state . . . . . . . . . . . . . . . . . . . . . . . 1126
mixing angle . . . . . . . . . . . . . . . . . . . . . . 1315
mnemonic rule . . . . . . . . . . . . . . . . . . . . . 475
for Green function . . . . . . . . . . . . . 472
model
Heisenberg . . . . . . . . . . . . . . . . . . . . 1120
Ising . . . . . . . . . . . . . . . . . . . . . . . . . . 1099
quark . . . . . . . . . . . . . . . . . . . . . . . . . 1380
Schwinger . . . . . . . . . . . . . . . . . . . . . . 897
spherical . . . . . . . . . . 1119, 1122, 1147
model,
Ising . . . . . . . . . . . . . . . . . . . . . . . . . 1120
modified
Bessel function . . . . . . . . . . . . . 51, 484
Mosla ller operator . . . . . . . . . . . . . . . . . 650
Mosla ller, C. . . . . . . . . . . . . . . . . . . . 649
Mohammedi, N. . . . . . . . . . . . .1259, 1278
Mohr, P.J. . . . . . . . . . . . . . . . . . . . 921, 922
moment
magnetic
anomalous . . . . . . . . . . . . . . 472, 881
momentum
angular . . . . . . . . . . . . . . . . . . . . . . . . . .58
center-of-mass . . . . . . . . . . . . . . . . . . 687
laboratory . . . . . . . . . . . . . . . . . . . . . 686
reference . . . . . . . . . . . . . . . . . . . . . . . 242
rest . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
space
wave functions in . . . . . . . . . . . . . 28
transfer . . . . . . . . . . . . . . . . . . . . . . . . 844
Moore, M.A. . . . . . . . . . . . . . . . 997, 1000
Morawetz, K. . . . . . . . . . . . . . . . . . . . . 238
Morgan, S.A. . . . . . . . . . . . . . . . . . . . . 705
Moser, J.K. . . . . . . . . . . . . . . . . . . . . . . . .80
Moses, H.E. . . . . . . . . . . . . . . . . 999, 1002
motion

1423
equation of . . . . . . . . . . . . . . . . . . . . . . 42
Mott
formula . . . . . . . . . . . . . . . . . . . . . . . . 856
Mott scattering . . . . . . . . . . . . . . . . . . . . .874
Mottelson, B.R. . . . . . . . . . . . 998, 1001
Mourou, G. . . . . . . . . . . . . . . . . . . . . . 1003
Mueller, K.G. . . . . . . . . . . . . . . 237, 999
Muller, F. . . . . . . . . . . . . . . . . . . . . . . 1379
multispinor
operator . . . . . . . . . . . . . . . . . . . . . . . 100
Mulvey, J.H. . . . . . . . . . . . . . . . . . . . . . 921
Murray, D.B. . . . . . . . . . . . . . . . . . . . 1153
Mustepanenko, M.V. . . . . . . . . . . . . 601
Neel temperature . . . . . . . . . . . . . . . . . .1148
Nagi, A.D.S. . . . . . . . . . . . . . . . . 998, 1001
Nakanishi, N. . . . . . . . . . . 543, 996, 1000
Nambu, Y. . . . . . . . . . . . . . . . . . . . . . . . 1326
Nambu-Goldstone theorem 181, 185, 217
Napier
analogy . . . . . . . . . . . . . . . . . . . . . . . . 376
natural units . . . . . . . . . . . . . . . . . . . . . . . 260
Naturforsch, Z. . . . . . . . . . . . . . . . . . 194
Nelson, D.R. . . . . . . . . . . . . . . . . 174, 237
Neu, J. . . . . . . . . . . . . . . . . 824, 1260, 1278
Neumann functions . . . . . . . . . . . . . . . . . 485
Neumann-Liouville . . . . . . . . . . . . . . . . 1275
Neumann-Liouville expansion . . 36, 647,
647, 651
neutral current . . . . . . . . . . . . . . 1369, 1372
neutrino . . . . . . . . . . . . . . . . . . . . . . 312, 1362
left-handed . . . . . . . . . . . . . . . . . . . . 340
neutron
-decay . . . . . . . . . . . . . . . . . . . . . . 1362
Nevell, A.C. . . . . . . . . . . . . . . . 999, 1002
Neveu, A. . . . . . . . . . . . . . . . . . . . . . . . .1326
Newton
gravitational constant . . . . . . . . . .292
Newtons
third law . . . . . . . . . . . . . . . . . . . . . . . . 83
Newton, I. . . . . . . . . . . . . . . . . . . . . . . . . . 80
Nickel, B.G. . . . . . . . . . . . . . . . . . . . . 1153
Nieto, A. . . . . . . . . . . . . . . . . . . . . . . . . . 238
Nieto, M.M. . . . . . . . . . . . . . . . . . . . . . 1379
Noether
charge . . . . . . . . . . . . . . . . . . . . 646, 667
current . . . . . . . . . . . . . . . . . . . . . . . . 658

Noether, E. . . . . . . . . . . . . . . . . . . . . . . 685
non-linear -model . . . . . . . . . . . . . . . . 1119
nonequilibrium
Green function
bosons . . . . . . . . . . . . . . . . . . . . . . . 154
fermions . . . . . . . . . . . . . . . . . . . . . 154
inequalities . . . . . . . . . . . . . . . . . . . 159
spectral representation . . . . . . . 153
Heisenberg picture . . . . . . . . .151, 160
quantum statistics . . . . . . . . . . . . . 150
Schroedinger picture . . . . . . . . . . . 151
nonholonomic
gauge invariance . . . . . . . . . . . . . . . 830
nonlocal
force . . . . . . . . . . . . . . . . . . . . . . . . . . . 835
normal
component . . . . . . . . . . . . . . . . . . . . 134
order . . . . . . . . . . . . . . . . . . . . . 527, 531
particle
density . . . . . . . . . . . . . . . . . . . . . . . 182
phase . . . . . . . . . . . . . . . . . . . . . . . . . . 762
product . . . . . . . . . . . . . . . . . . . . . . . . 579
normal ground state . . . . . . . . . . . . . . . 1148
normal product . . . . . . . . . . . . . . . . . . . . 467
normalization conditions . . . . . . . . . . 1202
Norton, W.W. . . . . . . . . . . . . . . . . . . . 427
notation
Weyl . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
Nozi`
eres, P. . . . . . . . . . . . . . . . . . 174, 237
Noziers, P. . . . . . . . . . . . . . . . . . . . . . . . 194
n-point
function . . . . . . . . . . . . . . . . . . . . . . . 578
propagator . . . . . . . . . . . . . . . . . . . . 578
n-point-function . . . . . . . . . . . . . . . . . . . . 748
nucleon
number . . . . . . . . . . . . . . . . . . . . . . . . 314
nucleus
self-conjugate . . . . . . . . . . . . . . . . .1293
number
atomic . . . . . . . . . . . . . . . . . . . . . . . . . 314
Bernoulli . . . . . . . . . . . . . . . . . . . . . . 598
bosons . . . . . . . . . . . . . . . . . . . . . . . . . 122
charge . . . . . . . . . . . . . . . . . . . . . . . . . 314
Euler . . . . . . . . . . . . . . . . . . . . . . . . . . 600
fermions . . . . . . . . . . . . . . . . . . . . . . . 127
nucleon . . . . . . . . . . . . . . . . . . . . . . . . 314

1424
number operator . . . . . . . . . . . . . . . . . . . . 92
observables
commuting . . . . . . . . . . . . . . . . . . . . . . . 4
operators . . . . . . . . . . . . . . . . . . . . . . . . 31
occupation number
basis . . . . . . . . . . . . . . . . . . . . . . . . . . .116
odd permutation . . . . . . . . . . . . . . . . . . . 166
off-shell T -matrix . . . . . . . . . . . . . . . . . . .671
Ohanian, H. . . . . . . . . . . . . . . . . . . . . . . 427
one-body-potential . . . . . . . . . . . . . . . . . . 83
one-particle
irreducible . . . . . . . . . . . . . . . . . . . . . 765
reducible . . . . . . . . . . . . . . . . . . . . . . . 765
Onofrio, R. . . . . . . . . . . . . . . . . . . . . . . 601
Onsager, L. . . . . . . . . . . . . . . . . . .174, 237
operator
antilinearr . . . . . . . . . . . . . . . . . . . . . 478
bilocal . . . . . . . . . . . . . . . . . . . . . . . . . 102
Casimir . . . . . . . . . . . . . . . . . . . . . . . . 352
chain rule . . . . . . . . . . . . . . . . . . 96, 112
charge
conjugation . . . . . . . . . . . . . . . . . . 480
density . . . . . . . . . . . . . . . . . . . . . . . . . .33
Hamiltonian . . . . . . . . . . . . . . . . . . . 657
Heisenberg . . . . . . . . . . . . . . . . . . . . . . 41
Hermitian . . . . . . . . . . . . . . . . . . . . . . . 17
Laplace-Beltrami . 55, 57, 58, 61, 67
Leibnitz rule . . . . . . . . . . . . . . . . . . . . 96
Mosla ller . . . . . . . . . . . . . . . . . . . . . . 650
multispinor . . . . . . . . . . . . . . . . . . . . 100
observable . . . . . . . . . . . . . . . . . . . . . . . 31
ordering problem . . . . . . . . . . . . 17, 56
parity . . . . . . . . . . . . . . . . . . . . . . . . . 476
representation . . . . . . . . . . . . . . . . . . .98
resolvent . . . . . . . . . . . . . . . . . . . . . . . . 47
spinor . . . . . . . . . . . . . . . . . . . . . . . . . 100
tensor . . . . . . . . . . . . . . . . . . . . . . . . . 100
spherical . . . . . . . . . . . . . . . . . . . . . 361
spherical, contravariant . . . . . . .362
time evolution . 35, 36, 3841, 44, 70
interaction picture . . . . . . . . . . . . 43
retarded . . . . . . . . . . . . . . . . . . . . . . 39
time-ordering . . . . . . . . . . . . . . . .36, 38
vector . . . . . . . . . . . 100, 251, 309, 850
zeta function . . . . . . . . . . . . . . . . . . . . 76
orbital

Index
transformation . . . . . . . . . . . . . . . . . 670
orbits
classical . . . . . . . . . . . . . . . . . . . . . . . . . . 1
order
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 494
normal . . . . . . . . . . . . . . . . . . . . 527, 531
of operators, causal . . . . . . . . . . . . . . 36
parameter . . . . . . . . . . . . . . . . . . . . . 494
problem for operators . . . . . . . . 17, 56
orthogonal
group
special . . . . . . . . . . . . . . . . . . . . . . . 243
orthogonality
relation . . . . . . . . . . . . . . . . . . . . 19, 263
basis dyads . . . . . . . . . . . . . . . . . . . . 53
theorem
group representations . . . . . . . . 168
orthogonality relation . . . . . . . . . . . . . . 522
orthogonality theorem
great . . . . . . . . . . . . . . . . . . . . . . . . . . 168
orthonormality
relation . . . . . . . . . . . . . . . . . . . . . . . . . 19
oscillator
local algebra . . . . . . . . . . . . . . . . . . . . 93
strength . . . . . . . . . . . . . . . . . . . . . . . 849
Osheroff, D.D. . . . . . . . . . . . . 997, 1001
Oteo, J.A. . . . . . . . . . . . . . . . . . . . . . . . . 371
packet, wave . . . . . . . . . . . . . . . . . . . . . . . . 14
Pagels, H. . . . . . . . . . . . . . . . . . . . . . . . 1326
pair
potential . . . . . . . . . . . . . . . . . . . . . . . . 83
Papastamatiou, N. J. . . . . . . . . . . . 1326
parameter
Einstein . . . . . . . . . . . . . . . . . . . . . . . . 241
impact . . . . . . . . . . . . . . . . . . . . . . . . .852
order . . . . . . . . . . . . . . . . . . . . . . . . . . 494
Pardee, W.J. . . . . . . . . . . . . . . . . . . . . . 383
parity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310
charge . . . . . . . . . . . . . . . . . . . . . . . . . 271
Dirac particle . . . . . . . . . . . . . . . . . . 510
intrinsic . . . . . . . . . . . . . . . . . . . . . . . 266
of charge . . . . . . . . . . . . . . . . . . . . . 1293
operator . . . . . . . . . . . . . . . . . . . . . . . 476
transformation . . . . . . . . . . . . . . . . . 255
violation in -decay . . . . . . . . . . . . 314
Park, S.H. . . . . . . . . . . . . . . . . . . . . . . . 1327

1425
Parker, C.S. . . . . . . . . . . . . . . . . . . . . 1279
partial-wave
expansion . . . . . . . . . . . . . . . . . 666, 700
scattering amplitude . 667, 700, 701
unitarity . . . . . . . . . . . . . . . . . . . . . . . 667
particle
near the sphere . . . . . . . . . . . . . . . 1273
normal density . . . . . . . . . . . . . . . . . 182
number operator . . . . . . . . . . . . . . . . 92
number, average . . . . . . . . . . . . . . . . 71
on surface of sphere . . . . . . . . . . . . . 58
pseudoscalar . . . . . . . . . . . . . . . . . . . 266
scalar . . . . . . . . . . . . . . . . . . . . . . . . . . 266
charged . . . . . . . . . . . . . . . . . . . . . . 472
neutral . . . . . . . . . . . . . . . . . . . . . . 466
spin-0 . . . . . . . . . . . . . . . . . . . . . . . . . . 463
spinless
charged . . . . . . . . . . . . . . . . . . . . . . 472
partition function
classical . . . . . . . . . . . . . . . . . . . . . . . . . 69
grand-canonical
quantum-statistical . . . . . . . . . . . 70
quantum-mechanical . . . . . . . . . . . . 70
quantum-statistical . . . . . . . . . . . . . 69
Paschen-Back effect . . . . . . . . . . . . . . . . 846
path
classical . . . . . . . . . . . . . . . . . . . . . . . . . . 2
integral . . . . . . . . . . . . . . . . . . . . . . . . 937
pattern, diffraction . . . . . . . . . . . . . . . . . . 30
Pauli
equation . . . . . . . . . . . . . . . . . . 320, 480
exclusion principle . . . . . . . . . . . . . 112
matrices . . . . . . . . . . . . . . . . . . . . . . . 301
matrix . . . . . . . . . . . . . . . . . . . . 302, 304
principle . . . . . . . . . . . . . . . . . . . . . . . 266
spin matrices . . . . . . . . . . . . . . . . . . . . 64
spinor . . . . . . . . . . . . . . . . . . . . . 301, 490
charge-conjugate . . . . . . . . . . . . .333
Pauli, W. . . . . . . . . . . . . . . . . . . . 524, 1362
Pauli-Villars regularization . . . . . . . . . 899
Paves, M. . . . . . . . . . . . . . . . . . . . . . . . . . 407
Peccei, R.D. . . . . . . . . . . . . . . . . . . . . . . 383
Pelster, A. . . . . . . . . . . . . . . . . . 238, 1050
Penrose, O. . . . . . . . . . . . . . . . . . .174, 237
permutation
even or odd . . . . . . . . . . . . . . . . . . . . 166

group . . . . . . . . . . . . . . . . . . . . . . . . . . 165
identity . . . . . . . . . . . . . . . . . . . . . . . . 165
perturbation . . . . . . . . . . . . . . . . . . . . . . . 644
expansion
Schwinger-Dyson . . . . . . . . . . . . . 553
theory
variational . . . . . . . . . . . . . . . . . . . 176
Petermann, A. . . . . . . . . . . . . . . . . . . . 921
Peters, P.C. . . . . . . . . . . . . . . . . . . . . . .427
Petley, B.W. . . . . . . . . . . . . . . . . . . . . . 383
Pfahl, E.D. . . . . . . . . . . . . . . . . . . . . . . . 920
phase
condensed . . . . . . . . . . . . . . . . . . . . . . 762
convention
Condon-Shortley . . 281, 286, 354,
378
normal . . . . . . . . . . . . . . . . . . . . . . . . . 762
shift . . . . . . . . . . . . . . . . . . . . . . . . . . . 668
space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
phenomena
collective . . . . . . . . . . . . . . . . . . . . . . . .91
critical . . . . . . . . . . . . . . . . . . . . . . . . 1197
photoelectric-effect . . . . . . . . . . . . . . . . . . 13
photons . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Picasso, E. . . . . . . . . . . . . . . . . . . . . . . . 921
Piccioni, O. . . . . . . . . . . . . . . . . . . . . . 1379
picture
Heisenberg 40, 41, 42, 109, 150, 645
in nonequilibrium theory 151, 160
interaction . . . . . . . . . . . . . . . . . . . . . 553
interaction (Dirac) . . . . . . . . . . . . . . 43
interaction or Dirac . . . . . . . . . . . . 646
Schr
odinger . . . . . . . . . . . . . . . . . . . . 109
Schroedinger . . . . . . . . . . . . . . . . 41, 42
in nonequilibrium theory . . . . . 151
Pines, D. . . . . . . . . . . . . . . . . 174, 194, 237
pion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1295
Pitaevski, L.P. . . . . . . . . . . . . . . . . . . . . 80
Planck
constant . . . . . . . . . . . . . . . . . . . . . . . . .13
length . . . . . . . . . . . . . . . . . . . . 292, 468
mass . . . . . . . . . . . . . . . . . . . . . . 292, 292
plane wave . . . . . . . . . . . . . . . . . . . . . . . . . . 13
plot
Dalitz plot . . . . . . . . . . . . . . . . . . . . . 688
Plunien, G. . . . . . . . . . . . . . . . . . . . . . . . 922

1426
Podolsky, B. . . . . . . . . . . . . . . . . . . . . . . 81
Poenaru, V. . . . . . . . . . . . . . . . . 998, 1002
Poincare
group . . 293, 310, 475, 503, 504, 601
Poincare group . . . . . . . . . . . . . . . . 264, 279
point
fixed . . . . . . . . . . . . . . . . . . . . 1209, 1215
transformation . . . . . . . . . . . . . . . . . . . 4
Poisson
brackets . . . . . . . . . . . . . 4, 7, 8, 40, 58
summation formula . . . 29, 148, 570,
572
polarization . . . . . . . . . . . . . . . . . . . . . . . . 280
vector
longitudinal . . . . . . . . . . . . 286, 287
scalar . . . . . . . . . . . . . . . . . . . . . . . .288
polarization vector
transverse . . . . . . . . . . . . . . . . . . . . . 283
Politzer, D. . . . . . . . . . . . . . . . . . . . . . 1160
Polyakov, A. . . . . . . . . . . . . . . . 997, 1001
Polylogarithmic functions . . . . . . . . . . 170
polynomial
Bernoulli . . . . . . . . . . . . . . . . . . . . . . 597
Euler . . . . . . . . . . . . . . . . . . . . . . . . . . 600
Gegenbauer . . . . . . . . . . . . . . . . . . . . 699
Jacobi . . . . . . . . . . . . . . . . . . . . . . 65, 699
Laguerre . . . . . . . . . . . . . . . . . . . . . . . 483
Legendre . . . . . . . . . . . . . . . . . .358, 382
Popov, V.N. 542, 997, 999, 1000, 1400,
1401
Porto, R.A. . . . . . . . . . . . . . . . . . . . . . . 427
positron . . . . . . . . . . . . . . . . . . . . . . . 500, 856
hole . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
Possenti, A. . . . . . . . . . . . . . . . . . . . . . . 427
potential
champaign bottle . . . . . . . . . . . . . . .142
chemical . . . . . . . . . . . . . . . . . . . . . . . . 70
Coulomb . . . . . . . . . . . . . . . . . . . . . . . 895
Dirac equation . . . . . . . . . . . . . . . 485
external . . . . . . . . . . . . . . . . . . . . . . 481
Klein-Gordon equation . . . . . . . 484
double-well . . . . . . . . . 142, 1097, 1098
external . . . . . . . . . . . . . . . . . . . . . . . . . 83
Lienard-Wiechert . . . . . . . . . 386, 490
Mexican hat . . . . . . . . . . . . . . . . . . . 142
pair . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

Index
Wigner . . . . . . . . . . . . . . . . . . . . . . . 1381
Poulson, D.N. . . . . . . . . . . . . . . 998, 1001
Powell, W.M. . . . . . . . . . . . . . . . . . . 1379
precession
geodetic . . . . . . . . . . . . . . . . . . . 425, 426
Lense-Thirring (frame-dragging) 426
Thomas . . . . 343, 345, 346, 371, 377,
470, 473, 843
Wigner . . . . . . . . . . . . . . . 343, 371, 372
prescription
i . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
Press, W.H. . . . . . . . . . . . . . . . . . 175, 427
Presser, G. . . . . . . . . . . . . . . . . . . . . . . 1379
pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Price, R.H. . . . . . . . . . . . . . . . . . . . . . . . 427
primary constraint . . . . . . . . . . . . 514, 548
principal-value . . . . . . . . . . . . . . . . . . . . . 491
principal-value integral . . . . . . . . . 48, 158
principle
correspondence . . 15, 17, 31, 56, 58,
64, 68
detailed balance . . . . . . . . . . . . . . . 693
equivalence . . . . . . . . . . . . . . . . . . . . 289
Pauli . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
probability
conservation law . . . . . . . . . . . . . . . . .16
probability current density . . . . . . . . . 262
problem
operator-ordering . . . . . . . . . . . . 17, 56
process
adiabatic . . . . . . . . . . . . . . . . . . . . . . 126
product
normal . . . . . . . . . . . . . . . . . . . . 467, 579
scalar . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Prokhorov, L.P. . . . . . . . . . . . . . . . . . 538
Prokhorov, L.V. . . . . . . . . . . . . . . . . . 601
propagator . . . . . . ,see also Green45, 577
causal . . . . . . . . . . . . . . . . . . . . . . . . . . 119
Feynman . . . . . . . . . . . . . . . . . . . . . . 481
finite-temperature . . . . . . . . . . . . . . 143
free-particle . . . . . . . . . . . . . . . 117, 118
n-point . . . . . . . . . . . . . . . . . . . . . . . . 578
retarded . . . . . . . . . . . . . . . . . . . . . . . 486
proper Lorentz group . . . . . . . . . . . . . . . 242
proper Lorentz transformations . . . . . 242
proper time . . . . . . . . . . . . . . . . . . . . 10, 483

1427
proper-time formalism, Schwinger . . 483
pseudo-physical state . . . . . . . . . . . . . . 538
pseudoscalar
particle . . . . . . . . . . . . . . . . . . . . . . . . 266
pseudotensor field . . . . . . . . . . . . . . . . . . 294
pulsar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
double . . . . . . . . . . . . . . . . . . . . . . . . . 406
QED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825
quantization
canonical . . . . . . . . . . . . . 40, 5759, 67
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
field, Hamiltonian . . . . . . . . . . . . . . . 96
geometric . . . . . . . . . . . . . . . . . . . . . . . 81
group . . . . . . . . . . . . . . . . . . . . . . . 58, 61
Gupta-Bleuler . . . . . . . . 539, 540, 542
second . . . . . . . . . . . . . . . . . . . . . . . . . 111
quantum
-statistical
partition function . . . . . . . . . . . . . 69
chromodynamics (QCD) . . . . . . 1383
electrodynamics . . . . . . . . . . . . . . . . 825
electrodynamics (QED) . . 825, 1383
fluctuation . . . . . . . . . . . . . . . . . . . . . 149
mechanics . . . . . . . . . . . . . . . . . . . . . . . 11
partition function . . . . . . . . . . . . . 70
number
radial . . . . . . . . . . . . . . . . . . . . . . . .483
of flux . . . . . . . . . . . . . . . . . . . . . . . . 1126
statistics . . . . . . . . . . . . . . . . . . . . . . . . 68
nonequilibrium . . . . . . . . . . . . . . . 150
quantum-statistical Gibbs distribution
function . . . . . . . . . . . . . . . . . . . . . 70
quark
bottom . . . . . . . . . . . . . . . . . . . . . . . 1317
charm . . . . . . . . . . . . . . . . . . . . . . . . .1317
field . . . . . . . . . . . . . . . . . . . . . . . . . . 1301
model . . . . . . . . . . . . . . . . . . . . . . . . . 1380
top . . . . . . . . . . . . . . . . . . . . . . . . . . . 1317
quasi-Cartesian . . . . . . . . . . . . . . . . . . . . . . 15
quasi-long-range correlations . . . . . . 1153
Racah recoupling coefficient . . . . . . . . 505
Raczka, R. . . . . . . . . . . . . . . . . . . . . . . . 174
radial fluctuations . . . . . . . . . . . . . . . . . 1119
radiation
gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 275

radiative, decay . . . . . . . . . . . . . . . . . . . 1289


radius of charge . . . . . . . . . . . . . . . . . . . . 887
Rainer, D. . . . . . . . 997, 998, 1000, 1001
Raman, C. . . . . . . . . . . . . . . . . . . . 174, 237
range
of attraction (fixed point) . . . . . 1216
effective . . . . . . . . . . . . . . . . . . . . . . . 670
rank . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .249
rapidity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
Rarita-Schwinger
field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
Rarita-Schwinger isospinors . . . . . . . 1398
ratio
gyromagnetic . . . 470, 472, 475, 481,
499, 843
real-time Green function
for Tungzero . . . . . . . . . . . . . . 150, 153
redial
quantum number . . . . . . . . . . . . . . 483
reduced
coupling constant . . . . . . . . . . . . . . 812
matrix element . . . . . . . . . . . . . . . . 361
surface . . . . . . . . . . . . . . . . . . . . . . . . . 800
reduced mass . . . . . . . . . . . . . . . . . . . . . . . 680
reduced matrix elements . . . . . . . . . . 1296
reducible
one-particle . . . . . . . . . . . . . . . . . . . . 765
reduction formulas, LSZ . . . . . . . . . . . . 770
reference
momentum . . . . . . . . . . . . . . . . . . . . . 242
reflection
mirror . . . . . . . . . . . . . . . . . . . . . . . . . .255
space . . . . . . . . . . . . . . . . . . . . . . . . . . 255
reflection amplitude . . . . . . . . . . . . . . . . 697
reflection, Bragg . . . . . . . . . . . . . . . . . . . . 11
reflections
group . . . . . . . . . . . . . . . . . . . . . . . . . . 265
regime
ultraviolet . . . . . . . . . . . . . . . . . . . . . . 769
regular
part . . . . . . . . . . . . . . . . . . . . . . . . . . . . 787
regularization
analytic . . . . . . . . . . . . . . . . . . . . . . . . 801
dimensional . . . . . . . . . . . . . . . . . . . . 801
zeta function . . . . . . . . . . . . . . . . . . 572
Reiner, A.S. . . . . . . . . . . . . . . . . . . . . . . 705

1428
Reinhardt, H. . . . . . . . . . . . . . . . . . . . 1029
relation
canonical
commutation . . . . . . . . . . . . . 40, 522
completeness . 19, 22, 23, 28, 29, 32,
47, 49, 377, 522
basis dyads . . . . . . . . . . . . . . . . . . . . 53
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . 21
Compton . . . . . . . . . . . . . . . . . . . . . . . 862
Euler . . . . . . . . . . . . . . . . . . . . . . 74, 135
orthogonality . . . . . . . . . . 19, 263, 522
basis dyads . . . . . . . . . . . . . . . . . . . . 53
orthonormality . . . . . . . . . . . . . . . . . . 19
uncertainty . . . . . . . . . . . . . . . . . . . . . . 33
unitarity . . . . . . . . . . . . . . . . . . . . . . . 645
Remiddi, E. . . . . . . . . . . . . . . . . . . . . . . . 921
Renk, B. . . . . . . . . . . . . . . . . . . . . . . . . . 1379
renormalizable
field theories . . . . . . . . . . . . . . . . . . . 781
renormalizable theries . . . . . . . . . . . . . . 468
renormalization . . . . . . . . . . . . . . . . 184, 975
group . . . . . . . . . . . . . . . . . . . . . . . . . . 802
group equation . . . . . . . . . . 1203, 1213
group invariant . . . . . . . . . . . . . . . 1200
group trajectory . . . . . . . . . . . . . . 1199
representation
adjoint . . . . . . . . . . . . . . . . . . . . . . . . . 250
chiral . . . . . . . . . . . . . . . . . . . . . 317, 331
fundamental . . . . . . . . . . . . . . . . . . . 300
global . . . . . . . . . . . . . . . . . . . . . . . . . . 301
group . . . . . . . . . . . . . . . . . . . . . 299, 657
Lie algebra . . . . . . . . . . . . . . . . 251, 300
Majorana . . . . . . . . . . . . . . . . . . . . . . 328
matrix . . . . . . . . . . . . . . . . . . . . . . . . . 300
operator
second-quantized . . . . . . . . . . . . . . 98
spectral . . . . . . . . . . . . . . . . . . . . . . . . . 47
nonequilibrium Green functions
153
Weyl . . . . . . . . . . . . . . . . . . . . . . . . . . 317
residue theorem, Cauchy . . . . . . . . . . . 147
resolvent . . . . . . . . . . . . . . . . . . . . . . . .47, 652
rest momentum . . . . . . . . . . . . . . . . . . . . 242
retarded . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Green function . . . . . . . . . . . . . . . . . 151
propagator . . . . . . . . . . . . . . . . . . . . 486

Index
time evolution
amplitude . . . . . . . . . . . . . . . . . . . . . 45
operator . . . . . . . . . . . . . . . . . . 39, 45
Retherford, R.C. . . . . . . . . . . . . . . . 921
reversal
time . . . . . . . . . . . . . . . . . . . . . . 256, 324
time, second quantization . . . . . . 268
Reynolds, J. . . . . . . . . . . . . . . . . . . . . . 427
Ricci tensor . . . . . . . . . . . . . . . . . . . . 80, 290
Rice, T.M. . . . . . . . . . . . . . 996, 999, 1160
Richardson, R.C. . . . . . . . . . . 997, 1000
Richardson, R.W. . . . . 997, 998, 1000,
1001
Riemann
-Lebesgue Lemma . . . . . . . . . 260, 383
connection . . . . . . . . . . . . . . . . . . . . . . 79
spinning top . . . . . . . . . . . . . . . . . . .80
curvature tensor . . . . . . . . . . . . . . . 289
zeta function . . . . . 76, 124, 209, 599
Riemanns zeta function . . . . . 1317, 1321
right-handed
antineutrino . . . . . . . . . . . . . . . . . . . 340
Ritus, V.I. . . . . . . . . . . . . . . . . . . . . . . . 1003
Roberts, A. . . . . . . . . . . . . . . . . . . . . . . 693
Robinson expansion . . . . . . . . . . . 136, 977
Robinson, J.E. . . . . . . . . . . . . . . . 237, 977
Rohrlich, F. . . . . . . . . . . . . . . . . . . . . . .920
Rosenfeld, A.H. . . . . . . . . . . . . . . . . . 695
Rosenfeld, L. . . . . . . . . . . . . . . . . . . . . 685
Rosenstein, B. . . . . . . . . . . . . . . . . . . 1327
Roskies, R.Z. . . . . . . . . . . . . . . . . . . . . . 921
rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
group . . . . . . . . . . . . . . . . . . . . . . . . . . 240
invariance . . . . . . . . . . . . . . . . . . . . . 650
Wick . . . . . . . . . . . . . . . . . 482, 894, 896
Wigner . . . . . . . . . . . . . . . . . . . . . . . . 341
Rothstein, I. . . . . . . . . . . . . . . . . . . . . . 427
Rouet, A. . . . . . . . . . . . . . . . . . . . . . . . . . 543
Rubinstein, J. . . . . . . . . . . . . . . 999, 1002
Ruffini, R. . . . . . . . . . . . . . . 407, 427, 503
rule
golden . . . . . . . . . . . . . . . . . . . . . . . . . 675
Jordan . . . . . . . . . . . . . . . . . . . . . . . . . . 15
minimal substitution . 466, 469, 476
mnemonic . . . . . . . . . . . . . . . . . . . . . . 475
Veltman . . . . . . . . . 574, 799, 800, 822

1429
running coupling constant . . . . . . . . . 1209
Ruoso, G. . . . . . . . . . . . . . . . . . . . . . . . . . 601
Rutherford
formula . . . . . . . . . . . . . . . . . . . . . . . . 856
scattering . . . . . . . . . . . . . . . . . . . . . . 851
Ryzhik, I.M. . . . 51, 129, 170, 208, 226,
483, 484, 569, 575, 699, 701, 984,
1003
Rzewuski, J. . . . . . . . . . . . . . . . . . 996, 999
S-matrix . . . . . . . . . . . . . . . . . . . . . . . . . . .650
Sackett, C.A. . . . . . . . . . . . . . . . 174, 237
Saharian, A.A. . . . . . . . . . . . . . . . . . . . 601
Saint-James, D. . . . . . . . . . . . . 997, 1000
Sakurai, J.J. . . . . . . . . . . . . . . . . .315, 695
Salam, A. . . . . . . . . . . . . . . . . . . . . 996, 999
Salpeter, E.E. . . . . . . . . . . . . . . 488, 922
Sarkissian, J.M. . . . . . . . . . . . . . . . . . . 427
Sarma, G. . . . . . . . . . . . . . . . . . . .997, 1000
Sawada, K. . . . . . . . . . . . . . . . . . . . . . . . 237
scalar . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 526
curvature . . . . . . . . . . . . . . .68, 80, 290
field . . . . . . . . . . . . . . . . . . . . . . 246, 257
particle
charged . . . . . . . . . . . . . . . . . . . . . . 472
neutral . . . . . . . . . . . . . . . . . . . . . . 466
polarization vector . . . . . . . . . . . . . 288
product . . . . . . . . . . . . . . . . . . . . . . . . . 19
scalar polarization vector . . . . . . . . . . . 525
scalar quantum electrodanymics . . . 1107
scalar particle . . . . . . . . . . . . . . . . . . . . . .266
scalar polarization vector . . . . . . . . . . 288
scattering
amplitude . . . . . . . . . . . . 553, 664, 677
partial-wave . . . . . . . .667, 700, 701
Compton . . . . . . . . . . . . . . . . . . . . . . 857
length . . . . . . . . . . . . . . . . . . . . 670, 680
matrix . . . . . . . . . . . . . . . . . . . . . . . . . 650
Mott . . . . . . . . . . . . . . . . . . . . . . . . . . . 874
Rutherford . . . . . . . . . . . . . . . . 851, 856
Thomson . . . . . . . . . . . . . . . . . . 858, 863
Schafer, G. . . . . . . . . . . . . . . . . . . . . . . 427
Schiff, L.I. . . . . . . . . . . . . . . . . . . . . . . . . 80
Schilling, K. . . . . . . . . . . . . . . . . . . . . 1400
Schmidt, S. . . . . . . . . . . . . . . . . . . . . . . . 238
Schouten, J.A. . . . . . . . . . . . . . . . . . . . . 81
Schr
odinger

field . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
field operator . . . . . . . . . . . . . . . . . . 268
picture . . . . . . . . . . . . . . . . . . . . . . . . . 109
dinger, E. . . . . . . . . . . . . . . . . . 383
Schro
Schr
odinger equation
second-quantized . . . . . . . . . . . . . . . . 96
Schreiber, M. . . . . . . . . . . . . . . . . . . . . 238
Schriefer, J.R. . . . . . . . . . . . . . . . . . . .999
Schrieffer, J. R. . . . . . . . . . . . . . . . 1280
Schrieffer, J.R. . . . . . . . . . . . . . . . . . 237
Schr
odinger
equation
relative . . . . . . . . . . . . . . . . . . . . . . 680
Schroedinger
equation . . 15, 1618, 25, 26, 3436,
3941, 45, 53, 55, 150
time-independent . . . . . . . . . . . . . 16
picture . . . . . . . . . . . . . . . . . . . . . . 41, 42
in nonequilibrium theory . . . . . 151
wave function . . . . . . . . . . . . . . . . . . . 17
Schubert, C. . . . . . . . . . . . . . . . . . . . . . 780
Schulte-Frohlinde, V. . vii, 174, 237,
780, 824, 1029, 1052, 1106, 1260,
1277, 1278
Schur
Lemma . . . . . . . . . . . . . . . . . . . . . . . 1296
Schwartz, A. . . . . . . . . . . . . . . . 997, 1001
Schwartz, L. . . . . . . . . . . . . . . . . . . . . . . 81
Schwarz
integrability condition 272, 274, 292
lemma . . . . . . . . . . . . . . . . . . . . . . . . . 272
Schwarz integrability condition . . . . . . . 6
Schwarz, H.A. . . . . . . . . . . . . . . . . . . 6, 81
Schweber, S. . . . . . . . . . . . 684, 705, 920
Schwinger
proper-time formalism . . . . . . . . . . 483
term . . . . . . . . . . . . 552, 553, 554, 578
Schwinger model . . . . . . . . . . . . . . . . . . . 897
Schwinger, J. 492, 503, 601, 920, 921,
999, 1002
Schwinger-Dyson, perturbation series 553
seagull diagram . . . . . . . . . . . . . . . . . . . 1109
seagull term . . . . . . . . . . . . . . . . . . . . . . . . 908
second quantization . . . . . . . . . . . . . . . . 111
and time reversal . . . . . . . . . . . . . . . 268
of spin . . . . . . . . . . . . . . . . . . . . . . . . . 351

1430
Hamiltonian . . . . . . . . . . . . . . . . . . . . . 96
Hilbert space . . . . . . . . . . . . . . . . . . . 95
Schr
odinger equation . . . . . . . . . . . . 96
second sound . . . . . . . . . . . . . . . . . . . . . . . 186
second-order phase transition . . . . . . 1197
secondary constraint . . . . . . . . . . . . . . . 514
Seff, G. . . . . . . . . . . . . . . . . . . . . . . . . . . . 922
Segur, H. . . . . . . . . . . . . . . . . . . . 999, 1002
self-conjugate nucleus . . . . . . . . . . . . . 1293
self-energy . . . . . . . . . . . . . . . . . . . . . . . . 1016
semi-completeness . . . . . . . . . . . . . . . . . . 336
Serene, J.W. . . . . . . . . . . 175, 998, 1001
series
Dyson . . . . . . . . . . . . . . . . . . . . 36, 1275
Taylor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Seurin, Y. . . . . . . . . . . . . . . . . . . . . . . . . 174
Shalaev, B.N. . . . . . . . . . . . . . . . . . . . 1278
shell
energy . . . . . . . . . . . . . . . . . . . . . . . . . 120
mass . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
Sherrington, D. . . . . . . .996, 999, 1160
She-Sheng Xue . . . . . . . . . . . . . . . . . . . 503
Shifman, M. . . . . . . . . . . . . . . . . . . . . . 1003
shift
Lamb . . . . . . . . . . . . . . . . . . . . . . . . . . 487
level . . . . . . . . . . . . . . . . . . . . . . . . . . . 657
phase . . . . . . . . . . . . . . . . . . . . . . . . . . 668
Shilov, G.E. . . . . . . . . . . . . . . . . . . . . . . . 81
short-distance singularity . . . . . . . . . . . 768
Shortley, G.H. . . . . . . . . . . . . . . . . . . .354
Siegel, C.L. . . . . . . . . . . . . . . . . . . . . . . . .80
Siegert, G. . . . . . . . . . . . . . . . . . . . . . . 1400
Silver, R.N. . . . . . . . . . . . . . . . . . . . . . . 175
singularity, short-distance . . . . . . . . . . 768
Sivia, D.S. . . . . . . . . . . . . . . . . . . . . . . . . 175
Slater determinant . . . . . . . . . . . . . . . . . . 88
Smith, H. . . . . . . . . . . . . . . . . . . . 997, 1001
Smorodinskij, J.A. . . . . . . . . . . . . . . . 377
Sochocki formula . . . . . . . . . . . . . . . 48, 491
soft symmetry breaking . . . . . . . . . . . . 679
soft-photon regime . . . . . . . . . . . . . . . . . 907
Sokolov, A.I. . . . . . . . . . . . . . . . . . . . 1278
solution
classical . . . . . . . . . . . . . . . . . 1096, 1097
Sommer, R. . . . . . . . . . . . . . . . . . . . . . . 1400
Sommerfeld, A. . . . . . . . . . . . . . . 80, 377

Index
Sommerfield, C.M. . . . . . . . . . . . . . . 921
sound, second . . . . . . . . . . . . . . . . . . . . . . 186
space
-time
curved . . . . . . . . . . . . . . . . . . . . . . . . 10
Minkowski . . . . . . . . . . . . . . . . . . . 243
Hilbert . . . . . . . . . . . . . . . . . . . . . . 16, 18
inversion . . . 255, 265, 311, 312, 314,
315, 364
linear . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
reflection . . . . . . . . . . . . . . . . . . . . . . 255
group . . . . . . . . . . . . . . . . . . . . . . . . 265
special
Lorentz group . . . . . . . . . . . . .241, 256
orthogonal group . . . . . . . . . . . . . . 243
specific heat . . . . . . . . . . . . . . . . . . . . . . . 1219
spectral
function sum rule . . . . . . . . . . . . . . 159
representation . . . . . . . . . . . . . . . . . . .47
dissipative part . . . . . . . . . . . . . . 158
nonequilibrium Green functions
153
spectrum, continuous . . . . . . . . . . . . . . . . 49
spectrum, discrete . . . . . . . . . . . . . . . . . . 672
hydrogen . . . . . . . . . . . . . . . . . . . . . . . 483
Dirac . . . . . . . . . . . . . . . . . . . . . . . . 487
relativistic . . . . . . . . . . . . . . . . . . . 485
sphere
particle near . . . . . . . . . . . . . . . . . . 1273
particle on surface . . . . . . . . . . . . . . . 58
surface in D dimensions . . . . . . . . . 72
spherical
components . . . . . 282, 360, 362, 851
harmonics . . . . . . . . 60, 358, 381, 666
hyper . . . . . . . . . . . . . . . . . . . . . . . . 698
spinor . . . . . . . . . . . . . . . . . . 489, 502
model . . . . . . . . . . . . . 1119, 1122, 1147
tensor operator . . . . . . . . . . . 360, 361
contravariant . . . . . . . . . . . . . . . . . 362
spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
Pauli matrices . . . . . . . . . . . . . . . . . . . 64
second quantization . . . . . . . . . . . . 351
three-vector . . . . . . . . . . . . . . . . . . . 347
transformation . . . . . . . . . . . . . . . . . 670
spin-orbit coupling . . . . . . . . . . . . . . . . . 470

1431
spinless particle . . . . . . . . . . . . . . . . . . . . 463
charged . . . . . . . . . . . . . . . . . . . . . . . . 472
spinning top . . . . . . . 58, 61, 6668, 70, 78
curvature scalar . . . . . . . . . . . . . . . . . 80
Ricci tensor . . . . . . . . . . . . . . . . . . . . . 80
Riemann connection . . . . . . . . . . . . . 80
spinor
Dirac . . . . . . . . . . . . . . . . . . . . . . . . . . 306
adjoint . . . . . . . . . . . . . . . . . . . . . . 307
helicity . . . . . . . . . . . . . . . . . . . 339, 340
Majorana . . . . . . . . . . . . . . . . . . . . . . 329
operator . . . . . . . . . . . . . . . . . . . . . . . 100
Pauli . . . . . . . . . . . . . . . . . . . . . . 301, 490
spherical
harmonics . . . . . . . . . . . . . . . . . . . . 502
spherical harmonics . . . . . . . . . . . . 489
Weyl . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
spontaneous
generation of fermion mass . . . . 1285
magnetization . . . . . . . . . . . 1120, 1221
symmetry breakdown . . . 1085, 1120,
1148, 1281, 1285
stable
ultraviolet . . . . . . . . . . . . . . . . . . . . 1216
Stanley, H.E. . . . . . . . . . . . . . . 996, 1000
Stare, G. . . . . . . . . . . . . . . . . . . . 997, 1000
state
biparticle . . . . . . . . . . . . . . . . . . . . . . 504
coherent . . . . . . . . . . . . . . . . . . . . . . . 531
density . . . . . . . . . . . . . . . . . . . . . . . . . .75
density of . . . . . . . . . . . . . . . . . . . . . . 122
equation of . . . . . . . . . . . . . . . . . . . . 123
Heisenberg . . . . . . . . . . . . . . . . . . . . . 109
pseudo-physical . . . . . . . . . . . . . . . . 538
Schroedinger . . . . . . . . . . . . . . . . . . . . 17
stationary . . . . . . . . . . . . . . . . . . . 16, 34
vacuum . . . . . . . . . . . . . . . . . . . . . . . . 531
statistical mechanics . . . . . . . . . . . . . . . . . 68
statistics
classical . . . . . . . . . . . . . . . . . . . . . . . . . 68
quantum . . . . . . . . . . . . . . . . . . . . . . . . 68
Steen, F.H. . . . . . . . . . . . . . . . . . . . . . . . 377
Stefan-Boltzmann
constant . . . . . . . . . . . . . . . . 1147, 1326
law . . . . . . . . . . . . . . . . . . . . . 1146, 1325
Steffen, P. . . . . . . . . . . . . . . . . . . . . . . 1379

Stegun, I. . 51, 486, 599, 600, 682, 699,


703, 704, 1003
Stehn, J.R. . . . . . . . . . . . . . . . . . . . . . . . 921
Steinberger, J. . . . . . . . . . . . . 693, 1379
Stepanenko, A.A. . . . . . . . . . 1259, 1278
Stern-Gerlach experiment . . . . . . . . . . 299
stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . 1131
Stock, V.S. . . . . . . . . . . . . . . . . . . . . . . . 174
Stone, M. . . . . . . . . . . . . . . . . . . . . . . . 1277
Stora, R. . . . . . . . . . . . . . . . . . . . . . . . . . 543
sser, M. . . . . . . . . . . . . . . . . . . . . 1052
Stro
strange particle . . . . . . . . . . . . . . . . . . . 1297
Stratonovich, R.L. . . . . . . . . 996, 1160
Streater, R.F. . . . . . . . . . . . . . . . . . . . 543
strength
oscillator . . . . . . . . . . . . . . . . . . . . . . 849
Strocchi, F. . . . . . . . . . . . . . . . . . . . . . . 543
structure constants . . . . . . . . . . . . . . . . 250
subgroup
largest . . . . . . . . . . . . . . . . . . . . . . . . . 242
subtraction . . . . . . . . . . . . . . . . . . . . 884, 886
minimal . . . . . . . . . . . . . . . . . . . . . . . 1202
Sudarshan, E.C.G. . . . . . . . . . . . . . . . 315
Sudarsky, D. . . . . . . . . . . . . . . . . . . . . . 602
sum rule
current algebra . . . . . . . . . . 1397, 1398
spectral function . . . . . . . . . . . . . . . 159
Thomas-Reiche-Kuhn . . . . . . . . . . 849
summation
convention
Einstein . . . . . . . . . . . . . . . . . . . . . . 98
convention Einstein . . . . . . . . . . . . 243
convention, Einstein . . . . . . . . . . . 2, 4
formula, Poisson . . 29, 29, 148, 570,
572
Matsubara . . . . . . . . . . . . . . . . . . . . 1137
Sunakawa, S. . . . . . . . . . . . . . . . . . . . . . 531
superatom . . . . . . . . . . . . . . . . . . . . . . . . . .138
superfluid
component . . . . . . . . . . . . . . . . . . . . 134
density . . . . . . . . . . . . . . . . . . . . . . . . 1104
helium . . . . . . . . . . . . . . . . . . . . . . . . . 138
superfluidity . . . . . . . . . . . . . . . . . . . . . . . .134
superselection rule . . . . . . . . . . . . . . . . 1281
supersymmetry . . . . . . . . . . . . . . . . . . . . 1402
surface

1432
of sphere
in D dimensions . . . . . . . . . . . . . . . 72
particle on . . . . . . . . . . . . . . . . . . . . 58
reduced . . . . . . . . . . . . . . . . . . . . . . . . 800
terms in partial integration . . . . . . . 2
susceptibility . . . . . . . . . . . . . . . . 1037, 1220
Suskov, A.O. . . . . . . . . . . . . . . . . . . . . . 601
Susskind, L. . . . . . . . . . . . . . . . . . . . . . . 920
Suzuki, H. . . . . . . . . . . . . . . . . . 1261, 1279
Suzuki, T. . . . . . . . . . . . . . . . . . . . . . . . 1029
Svidzinskij, A.V. . . . . . . 996, 999, 1160
symbol
Christoffel . . . . . . . . . . . . . . 11, 79, 290
Levi-Civita . . . . . . . . . . . . . . . . . . . . . 378
symmetric group . . . . . . . . . . . . . . . . . . . 165
symmetry
Bose . . . . . . . . . . . . . . . . . . . . . . . . . . . 865
breaking
soft . . . . . . . . . . . . . . . . . . . . . . . . . . 679
breaking, spontaneous . . . . . . . . . 1085
charge . . . . . . . . . . . . . . . . . . . . . . . . 1291
crossing . . . . . . . . . . . . . . . . . . . 859, 865
group . . . . . . . . . . . . . . . . . . . . . . . . . . 645
internal . . . . . . . . . . . . . . . . . . . . . . . . 677
isotopic spin . . . . . . . . . . . . . . . . . . 1292
spontaneous breakdown 1120, 1148,
1281, 1285
transformation . . . . . . . . . . . . 645, 664
discrete . . . . . . . . . . . . . . . . . . . . . . 558
spacetime-dependent . . . . . . . . . 659
variation . . . . . . . . 645, 646, 655, 662
symplectic
coordinate transformations . . . . . . . 7
unit matrix . . . . . . . . . . . . . . . . . . . . . . . 7
T -matrix
off-shell . . . . . . . . . . . . . . . . . . . . . . . . 671

Tauber , U.C. . . . . . . . . . . . . . . . . . . . . 174


uber, U.C. . . . . . . . . . . . . . . . . . . . . . 237
Ta
Tajima, T. . . . . . . . . . . . . . . . . . . . . . . . 1003
Takahashi, Y. . . . . . . . . . . . . . . . 920, 921
Takeda, G. . . . . . . . . . . . . . . . . . . . . . . . 696
Taylor expansion . . . . . . . . . . . . . . . . . . . . . 1
Taylor, J.H. . . . . . . . . . . . . . . . . . . . . . . 427
Teitelboim, C. . . . . . . . . . . . . . . . . . . . 384
Telegdi, V.L. . . . . . . . . . . . . . . . . . . . . .474
temperature

Index
critical . . . . . . . . . . . . . . . . . . . . . . . . . 208
Curie . . . . . . . . . . . . . . . . . . . . . . . . . 1133
Unruh . . . . . . . . . . . . . . . . . . . . . . . . . 577
Templeton, S. . . . . . . . . . . . . . . . . . . . . 420
tensor
curvature
Riemann . . . . . . . . . . . . . . . . . . . . 289
Einstein . . . . . . . . . . . . . . . . . . . . . . . . 291
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
dual . . . . . . . . . . . . . . . . . . . . . . . . . 272
force . . . . . . . . . . . . . . . . . . . . . . . . . . 1381
Levi-Civita . . . . . . . . . . . . . . . . . . . . . 247
metric . . . . . . . . . . . . . . . . . . . . . . . . . . 54
operator . . . . . . . . . . . . . . . . . . . . . . . 100
Ricci . . . . . . . . . . . . . . . . . . . . . . . 80, 290
tensor operator
spherical . . . . . . . . . . . . . . . . . . . . . . . 361
contravariant . . . . . . . . . . . . . . . . . 362
spherical components . . . . . . . . . . 360
term, Schwinger . . . . . 552, 553, 554, 578
test function . . . . . . . . . . . . . . . . . . . . .25, 46
Teukolsky, S.A. . . . . . . . . . . . . . . . . . 175
Theis, W.R. . . . . . . . . . . . . . . . . . . . . . . 315
theorem
Bloch . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Cauchy . . . . . . . . . . . . . . . . . . . . . . . . 669
center-of-mass . . . . . . . . . . . . . 653, 668
Coleman . . . . . . . . . . . . . . . . . . . . . . 1153
dimensionality . . . . . . . . . . . . . . . . . 168
Ehrenfest . . . . . . . . . . . . . . . . . . . . . . 269
Gauss . . . . . . . . . . . . . . . . . . . . . . . . . . 262
Griard . . . . . . . . . . . . . . . . . . . . . . . . . 377
Levinson . . . . . . . . . . . . . . . . . . . . . . .671
Mermin-Wagner . . . . . . . . . . . . . . . 1153
Wick . . . . . . . . . . . . . . . . . . . . . . 582, 591
Wigner-Eckart . . . . . . . . . . . . 361, 850
theorem.Nambu-Goldstone 181, 185, 217
theory
critical . . . . . . . . . . . . . . . . . . . . . . . . 1038
Gupta-Bleuler . . . . . . . . . . . . . . . . . . 538
linear response . . . . . . . 150, 150, 152
of fields, renormalizable . . . . . . . . 781
V-A . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
thermal
Green functions . . . . . . . . . . . . . . . . 143
length . . . . . . . . . . . . . . . . . . . . . . . . . 123

1433
thermodynamic
limit . . . . . . . . . . . . . . . . . . . . . . . . . . . 768
relation, Euler . . . . . . . . . . . . . . . . . . 74
theta function, elliptic . . . . . . . . .138, 169
Thomas
equation . . . . . . . . . . . . . . . . . . 475, 475
precession . .343, 343, 345, 346, 371,
375, 377, 470, 473, 843
Thomas, E.J. . . . . . . . . . . . . . . . 997, 1000
Thomas, L.T. . . . . . . . . . . . . . . . . . . . . . 475
Thomas-Reiche-Kuhn sum rule . . . . .849
Thomson scattering . . . . . . . . . . . 858, 863
t Hooft, G. 806, 920, 997, 1001, 1401
3 j-symbols, Wigner . . . . . . . . . . . . . 378
time
-ordered
Green function . . . . . . . . . . . . . . . 156
-ordering
operator . . . . . . . . . . . . . . . . . . 36, 38
inversion . . . . . . . . . . . . . . . . . . . . . . .256
proper . . . . . . . . . . . . . . . . . . . . . . . . . 483
reversal . . . . . . . . . . . . . . . . . . . 256, 324
second quantization . . . . . . . . . . 268
time evolution
amplitude . . . . . . . . . . . . . . .44, 47, 150
causal . . . . . . . . . . . . . . . . . . . . . . . . 45
retarded . . . . . . . . . . . . . . . . . . . . . . 45
operator . . . . . . .35, 36, 3841, 44, 70
anticausal . . . . . . . . . . . . . . . . . . . . . 39
causal . . . . . . . . . . . . . . . . . . . . . . . . 45
composition law . . . . . . . . . . . 38, 39
interaction picture . . . . . . . . . . . . 43
retarded . . . . . . . . . . . . . . . . . . .39, 45
Titchmarsh, E.C. . . . . . . . . . . . . . . . . 383
T -matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 662
Tognetti, V. . . . . . . . . . . . . . . . . . . . . 1279
Tollet, J.J. . . . . . . . . . . . . . . . . . 174, 237
Tomboulis, E. . . . . . . . . . . . . . 1029, 1049
top quark . . . . . . . . . . . . . . . . . . . . . . . . . 1317
top, spinning . . . . . . 58, 61, 6668, 70, 78
asymmetric . . . . . . . . . . . . . . . . . . . . . .78
curvature scalar . . . . . . . . . . . . . . . . . 80
Ricci tensor . . . . . . . . . . . . . . . . . . . . . 80
Riemann connection . . . . . . . . . . . . . 80
Toulouse, G. . . . . . . . . . . . . . . . 998, 1002
tracelog . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

transfer, momentum . . . . . . . . . . . . . . . . 844


transformation
Bogoliubov . . . . . . . . . . . . . . . . . . . . . 198
canonical . . . . . . . . . . . . . . . . . . . . . . 6, 8
coordinate
general . . . . . . . . . . . . . . . . . . . . . . . 292
duality . . . . . . . . . . . . . . . . . . . . . . . . 570
Fourier . . . . . . . . . . . . . . . . . . .14, 29, 33
inverse . . . . . . . . . . . . . . . . . . . . . . . . 29
Galilei . . . . . . . . . . . . . . . . . . . . . . . . . 657
gauge
second-kind . . . . . . . . . . . . . . . . . .274
Hubbard-Stratonovich . . . . . . . . . . 937
Legendre . . . . 466, 469, 498, 550, 649
Lorentz . . . . . . . . . . . . . . . . . . . . . . . . 657
proper . . . . . . . . . . . . . . . . . . . . . . . 242
orbital . . . . . . . . . . . . . . . . . . . . . . . . . 670
parity . . . . . . . . . . . . . . . . . . . . . . . . . 255
point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
spin . . . . . . . . . . . . . . . . . . . . . . . . . . . . 670
symmetry . . . . . . . . . . . . . . . . . 645, 664
discrete . . . . . . . . . . . . . . . . . . . . . . 558
spacetime-dependent . . . . . . . . . 659
transition
Gamow-Teller . . . . . . . . . . . . . . . . . 315
lambda . . . . . . . . . . . . . . . . . . . . . . . 1133
translation . . . . . . . . . . . . . . . . . . . . . .58, 310
invariance . . . . . . . . . . . . . . . . . . . . . . 650
transmission amplitude . . . . . . . . . . . . 697
transposition . . . . . . . . . . . . . . . . . . . 85, 165
transverse
-function . . . . . . . . . . . . . . . . . . . . . . 515
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 826
gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 283
polarization vector . . . . . . . . . . . . . 283
trap, Bose-Einstein condensation . . . 141
triangle, Mandelstam . . . . . . . . . . . . . . . 688
Triebwasser, S. . . . . . . . . . . . . . . . . . . 922
trivial fixed point . . . . . . . . . . . . . . . . . 1216
Trunov, N.N. . . . . . . . . . . . . . . . . . . . . . 601
tube
flux . . . . . . . . . . . . . . . . . . . . . 1126, 1128
Turlay, R. . . . . . . . . . . . . . . . . . . . . . . 1379
Tuytin, I.V. . . . . . . . . . . . . . . . . . . . . . 1400
two-body-potential . . . . . . . . . . . . . . . . . . 83
2-point function . . . . . . . . . . . . . . . . . . . . 757

1434
amputated . . . . . . . . . . . . . . . . . . . . . 766
Tyupkin, Y. . . . . . . . . . . . . . . . . .997, 1001
Uehling, E.A. . . . . . . . . . . . . . . . . . . . . 922
ultraviolet
cutoff . . . . . . . . . . . . . . . . . . . . . . . . . . 782
divergence . . . . . . . . . . . . . . . . . . . . . 768
regime . . . . . . . . . . . . . . . . . . . . . . . . . 769
stable . . . . . . . . . . . . . . . . . . . . . . . . . 1216
ultraviolet-stable fixed point . . . . . . . 1216
Umezawa, H. . . . . . . . . . . . . . . . . . . . . 1326
uncertainty
principle . . . . . . . . . . . . . . . . . . . . . . . . 14
relation . . . . . . . . . . . . . . . . . . . . . . . . . 33
unit matrix, symplectic . . . . . . . . . . . . . . . 7
unitarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
of partial waves . . . . . . . . . . . . . . . . 667
relation . . . . . . . . . . . . . . . . . . . . . . . . 645
units, natural . . . . . . . . . . . . . . . . . . . . . . 260
Unruh temperature . . . . . . . . . . . . . . . . . 577
Unruh, W.G. . . . . . . . . . . . . . . . . . . . . . 601
Utiyama, R. . . . . . . . . . . . . . . . . . . . . . . 531
UV cutoff . . . . . . . . . . . . . . . . . . . . . . . . . . 782
V-A theory . . . . . . . . . . . . . . . . . . . . . . . . 315
vacuum
diagram . . . . . . . . . . . . . . . . . . . . . . . . 756
diagrams . . . . . . . . . . . . . . . . . . . . . . . 814
energy . . . . . . . . . . . . . . . . . . . . .533, 814
fluctuations . . . . . . . . . . . . . . . 467, 563
state . . . . . . . . . . . . . . . . . . . . . . . . . . . 531
Vaia, R. . . . . . . . . . . . . . . . . . . . . . . . . . . 1279
Vainshtein, A. . . . . . . . . . . . . . . . . . . .1003
Vaks, V.G. . . . . . . . . . . . . . . . . . . . . . . . 1326
Valatin, J.G. . . . . . 197, 237, 997, 1000
Valluri, S.R. . . . . . . . . . . . . . . . . . . . . 1002
Van den Bossche, B. 1154, 1280, 1315
van Druten, N.J. . . . . . . . . . . . . . . . . 174
Van Dyck
Jr., R.S. . . . . . . . . . . . . . . . . . . . . . . 920
van Winter, C. . . . . . . . . . . . . . . . . . . . . 81
van der Waals force . . . . . . . . . . . . . . . . 568
van Druten, N.J. . . . . . . . . . . . . . . . . 237
Van Leeuwen, J.M.J. . . . . . . . . . . . . 705
Van-der-Waals force . . . . . . . . . . . . . . . . 563
van-der-Waals force . . . . . . . . . . . . . . . . 469
Vannucci, F. . . . . . . . . . . . . . . . . . . . . 1379

Index
VanVleck, J.H. . . . . . . . . . . . . . . . . . . 685
Vanzella, D.A.T. . . . . . . . . . . . . . . . . 602
variable
anticommuting . . . . . . . . . . . . . . . . 324
Grassmann . . . . . . . . . 113, 324, 1307
Mandelstam . . . . . . . . . . 688, 877, 879
variation
in action principle . . . . . . . . . . . . . . . . 1
symmetry . . . . . . . 645, 646, 655, 662
variational perturbation theory . . . . . 176
Varnashev, K.B. . . . . . . . . . . . . . . . . 1278
Vasilev, A.N. . . . . . . . . . . . . . 1259, 1278
Vasilev, A.N. . . . . . . . . . . . . . . . . . . . . 1029
vector
current density . . . . . . . . . . . . . . . . 325
field
axial . . . . . . . . . . . . . . . . . . . . . . . . .279
four-dimensional
coordinates . . . . . . . . . . . . . . . . . . 245
momenta . . . . . . . . . . . . . . . . . . . . 239
functional . . . . . . . . . . . . . . . . . . . . . 1004
meson dominance . . . . . . . . . . . . . . 555
operator . . . . . . . . .100, 251, 309, 850
polarization
longitudinal . . . . . . . . . . . . 286, 287
transverse . . . . . . . . . . . . . . . . . . . 283
spin . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
velocity
field . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
light . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Veltman integral rule 574, 799, 800, 822
Veltman, M.T. . . . . . . . . 806, 920, 1401
Verlet, L. . . . . . . . . . . . . . . . . . . . . . . . . 238
vertex function . . . . . . . . . . . . . . . . . . . . . 766
Vetterling, W.T. . . . . . . . . . . . . . . . .175
Vidberg, H.J. . . . . . . . . . . . . . . . . . . . . .175
vNeumann, J. . . . . . . . . . . . . . . . . . . . . . . 16
Vogels, J.M. . . . . . . . . . . . . . . . . 174, 237
Volkov, D.V. . . . . . . . . . . . . . . . . . . . .1401
Volovik, G.E. . . . . . . . . . . . . . . 998, 1002
vortex line . . . . 138, 483, 756, 1126, 1127
Vuorio, M. . . . . . . . . . . . . . . . . . 998, 1001
Wagner, H. . . . . . . . . . . . . . . . . . . . . . . 1160
Wagoner, R.V. . . . . . . . . . . . . . . . . . . . 427
Wahl, H. . . . . . . . . . . . . . . . . . . . . . . . . 1379
Walcka, J.D. . . . . . . . . . . . . . . . . . . . . . 194

1435
Walecka, J.D. . . . . . 174, 175, 996, 999
Walker, M. . . . . . . . . . . . . . . . . . . . . . . .427
Ward identity . . . . . . . . . . . . 797, 842, 901
Ward, J.C. . . . . . . . . . . . . . . . . . . . . . . . 920
Ward-Takahashi identity 840, 842, 901,
1092
Warr, B.J. . . . . . . . . . . . . . . . . . . . . . . 1327
Watson, G. . . . . . . . . . . . . . . . . . . . . . . 1003
Watson, K.N. . . . . . . . . . . . . . . . . . . . . 705
wave
frequency . . . . . . . . . . . . . . . . . . . . . . . 12
function . . . . . . . . . . . . 12, 47, 48, 115
momentum space . . . . . . . . . . . . . 28
Schroedinger . . . . . . . . . . . . . . . . . . 17
material . . . . . . . . . . . . . . . . . . . . . . . . 11
packet . . . . . . . . . . . . . . . . . . . . . . . . . . 14
plane . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
wave function
insertions . . . . . . . . . . . . . . . . . . . . . . 789
wavelength . . . . . . . . . . . . . . . 389, 390, 395
Compton . . . . . . . . . . . . . . . . . . . . . . . 904
weak
angle . . . . . . . . . . . . . . . . . . . . . . . . . .1372
hypercharge . . . . . . . . . . . . . . . . . . 1371
interactions . . . . . . . . . . . . . 1362, 1371
isospace . . . . . . . . . . . . . . . . . . . . . . 1372
weak-coupling
expansion . . . . . . . . . . . . . . . . . . . . . . 176
Wegner, F.J. . . 996, 1000, 1259, 1260,
1278
Weierstrass, K. . . . . . . . . . . . . . . . . . . .81
weight
fundamental . . . . . . . . . . . . . . . . . . 1307
Weinberg, S. . . . . . . . . . . . . . 81, 427, 921
Weisberg, J.M. . . . . . . . . . . . . . . . . . . .427
Weisskopf, V. . . . . . . . . . . . . . . 921, 1002
Weizel, W. . . . . . . . . . . . . . . . . . . . . . . . . 80
Weniger, E.J. . . . . . . . . . . . . . . . . . . . 1002
Wess, J. . . . . . . . . . . . . . . . . . . . . . . . . . 1401
Weyl
calculus . . . . . . . . . . . . . . . . . . . . . . . . 362
notation . . . . . . . . . . . . . . . . . . . . . . . .355
representation . . . . . . . . . . . . . . . . . 317
Weyl spinors . . . . . . . . . . . . . . . . . . . . . . . 303
Weyl, H. . . . . . . . . . . . . . . . . . . . . . . . . . 314
Wheater, J. . . . . . . . . . . . . . . . . . . . . . 1003

Wheatley, J.C. . . 997, 998, 1000, 1001


White, H.S. . . . . . . . . . . . . . . . . . . . . . 1379
Whittaker, E. . . . . . . . . . . . . . . . . . . 1003
Wick
contraction . . . . . . . . . . . . . . . . . . . . . 776
expansion . . . . . . . . . . . . . . . . . . . . . . 783
lemma . . . . . . . . . . . . . . . . . . . . . . . . . 540
rotation . . . . . . . . . . . . . . 482, 894, 896
theorem . . . . . . . . . . . . . . . . . . . 582, 591
Wicks theorem . . . . . . . . . . . . . . . . . . . . .930
Wick, G.C. . . . . . . . . . . . . . . . . . . 582, 601
width . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1289
Wiegel, F.W. . . . . . . . . . . . . . . . . . . . . 999
Wieman, C.E. . . . . . . . . . . . . . . . . 174, 237
Wightman, A.S. . . . . . . . . . . . . . . . . . . 543
Wigner
3j-symbols . . . . . . . . . . . . . . . . . . . . 378
function . . . . . . . . . . . . . . . . . . . . . . . . 34
group . . . . . . . . . . . . . . . . . . . . . . . . . . 242
lattice . . . . . . . . . . . . . . . . . . . . . . . . . . 194
Liouville equation . . . . . . . . . . . . . . . 34
potential . . . . . . . . . . . . . . . . . . . . . . 1381
precession . . . . . . . . . . . . 343, 371, 372
rotation . . . . . . . . . . . . . . . . . . . . . . . 341
Wigner, E.P. . . . . . . . . . . . . . . . . . . . . . 194
Wigner-Eckart theorem . 361, 850, 1296
Will, C.M. . . . . . . . . . . . . . . . . . . . . . . . 427
Wilson, K. . . . . . . . . . . . . . . . . . . . . . . 1160
Wilson, R. . . . . . . . . . . . . . . . . . . . . . . . . 693
Woodhouse, N.M.J. . . . . . . . . . . . . . . . 81
Wu, C.S. . . . . . . . . . . . . . . . . . . . . 314, 1379
Yang, C.N. . . . . . . . . . . . . . . 174, 237, 314
Yennie, D.R. . . . . . . . . . . . . . . . . . . . . . 921
Yukalov, V. I. . . . . . . . . . . . . . . . . . . . 238
Yukalova, E. P. . . . . . . . . . . . . . . . . . 238
Zech, G. . . . . . . . . . . . . . . . . . . . . . . . . . 1379
Zeeman effect . . . . . . . . . . . . . . . . . . . . . . 846
anomlaous . . . . . . . . . . . . . . . . . . . . . 846
zeta function . . . . . . . . . . . . . . . . . . 124, 209
generalized . . . . . . . . . . . . . . . . . . . . . . 76
Hurwitz . . . . . . . . . . . . . . . . . . . . . . . 575
operator . . . . . . . . . . . . . . . . . . . . . . . . .76
reflection formula . . . . . . . . . . . . . . 574
regularization . . . . . . . . . . . . . . . . . . 572
Riemann . . . . . . . . . . . . . . . . . . . 76, 599

1436
Zhuang, P.F. . . . . . . . . . . . . . . . . . . . . 1327
Zichichi, A. . . . . . . . . . . . . 996, 999, 1000
Zinn Justin, J. . . . . . . . . . . . . . . . . . . 1000
Zinn-Justin, J. . . 238, 996, 1260, 1277,
1278
Zuber, J.-B. . . . . . . . . . 1002, 1029, 1049
Zuber, J.B. . . . . . . . . . . . . . . . . . . 503, 920
Zumino, B. . . . . . . . . . . . . . . . . . . . . . . 1401

Index

Vous aimerez peut-être aussi