You are on page 1of 905

A.I.

Akhiezer
V.B. Berestetskii
Quantum
Electrodynamics
INTERSCIENCE MONOGRAPHS AND TEXTS
IN PHYSICS AND ASTRONOMY
Edited by R. E. MARSHAK

Volum e I: E. R. Cohen, K. M. Crowe, and J. W. M. DuMond


THE FUNDAMENTAL CONSTANTS OF
PHYSICS
Volum e II: G. J. Dienes and G. H. Vineyard
RADIATION EFFECTS IN SOLIDS
Volum e III: N. N. Bogoliubov and D. V. Shirkov
INTRODUCTION TO THE THEORY OF
QUANTIZED FIELDS
Volum e IV: J. B. Marion and J. L. Fowler, Editors
FAST NEUTRON PHYSICS In two parts
Part I: Techniques
Part II: Experiments and Theory
Volum e V: D. M. Ritson, Editor
TECHNIQUES OF HIGH ENERGY PHYSICS
Volum e VI: R. N. Thomas and R. G. Athay
PHYSICS OF THE SOLAR CHROMOSPHERE
Volum e VII: L. H. Aller
THE ABUNDANCE OF THE ELEMENTS
Volum e VIII: E. N. Parker
INTERPLANETARY DYNAMICAL PROCESSES
Volum e IX: C. L. Longmire
ELEMENTARY PLASMA PHYSICS
Volum e X: R. Brout and P. Carruthers
LECTURES ON THE MANY-ELECTRON
PROBLEM
Volum e XI: A. I. Akhiezer and V. B. Berestetskii
QUANTUM ELECTRODYNAMICS
Volum e XII: H. Panofsky and J. Lumley
STRUCTURE OF ATMOSPHERIC TURBULENCE
Volum e XIII: R. D. Heidenreich
INTRODUCTION TO TRANSMISSION
ELECTRON MICROSCOPY
Additional volumes in preparation
INTERSCIENCE MONOGRAPHS AND TEXTS
IN PHYSICS AND ASTRONOMY
Edited by R. E. MARSHAK
University of Rochester, Rochester, New York
VOLUME XI

E d ito r ia l A d v iso ry B o a rd

A. ABRAGAM, College de France, Paris, France


H. ALFVEN, Royal Institute of Technology, Stockholm, Sweden
L. V. BERKNER, Southwest Research Conference, Dallas, Texas
H. J. BHABHA, Tata Institute for Fundamental Research, Bombay, India
L. BIERMANN, Max Planck Institut fur Physik und Astrophysik, Munich,
Germany
N. N. BOGOLIUBOV, J.I.N.R., Dubna, U.S.S.R.
A. BOHR, Institute for Theoretical Physics, Copenhagen, Denmark
S. CHANDRASEKHAR, Enrico Fermi Institute, Chicago, Illinois
J. W. DuMOND, California Institute of Technology, Pasadena, California
J. FRIEDEL, University of Paris, Orsay, France
L. GOLDBERG, Harvard University, Cambridge, Massachusetts
M. GOLDHABER, Brookhaven National Laboratory, Upton, New York
H. A. GOVE, Institute for Theoretical Physics, Copenhagen, Denmark
C. HERRING, Bell Telephone Laboratories, Murray Hill, New Jersey
J. KAPLAN, University of California, Los Angeles, California
C. M0LLER, Institute for Theoretical Physics, Copenhagen, Denmark
W. K. H. PANOFSKY, Stanford University, Stanford, California
R. E. PEIERLS, Oxford University, Oxford, England
F. PRESS, California Institute of Technology, Pasadena, California
B. ROSSI, Massachusetts Institute of Technology, Cambridge, Massachusetts
A. SCHAWLOW, Stanford University, Stanford, California
R. A. SMITH, Massachusetts Institute of Technology, Cambridge, Massachusetts
L. SPITZER, Jr., Princeton Observatory, Princeton, New Jersey
B. STROMGREN, Institute for Advanced Study, Princeton, New Jersey
G. TORALDO di FRANCIA, University of Florence, Florence, Italy
G. E. UHLENBECK, Rockefeller Institute, New York, New York
L. VAN HOVE, C.E.R.N., Geneva, Switzerland
V. F. WEISSKOPF, C.E.R.N., Geneva, Switzerland
H. YUKAWA, Institute for Theoretical Physics, Kyoto, Japan
QUANTUM ELECTRODYNAMICS

A. I. AKHIEZER
Physico-Technical Institute, Academy of Sciences,
Khar’kov, U.S.S.R.

V. B. BERESTETSKII
Institute for Theoretical and Experimental Physics,
Academy of Sciences, Moscow, U.S.S.R.

Authorized English Edition Revised and Enlarged by the Authors

Translated from the Second Russian Edition by


G. M. Volkoff
University of British Columbia,
Vancouver, B.C., Canada

INTERSCIENCE PUBLISHERS 1965


a division of John Wiley & Sons, New York • London • Sydney
English Translation and All Additions to the Russian Edition
Copyright (c) 1965, John Wiley & Sons, Inc.
ALL RIGHTS RESERVED. This book or any
part thereof must not be reproduced in any
form without the written permission of the
publisher.

Library of Congress Catalog Card Number 63-17766

Printed in the United States of America


Translator’s Note

This translation follows in the main the second Russian edition of


1959, but in addition incorporates several new sections (in particular
§§ 9.4, 26.2, 26.6, 48.4, 48.5, 51.5, 52.5) provided by the authors
specially for this edition, as well as a number of minor revisions, and
corrections of numerous typographical errors in the formulas. The
translator is grateful to the authors for their cooperation.
G. V o l k o f f
Vancouver, Canada

v
P r e fa c e to th e S eco n d E d itio n

In preparing the second edition we have subjected the book to


extensive revision. The prineipal aim and the contents of the book
have not been altered; the book is devoted to the systematic presen­
tation of electromagnetic processes only. Only some general theorems
and methods go beyond the framework of electrodynamics proper.
In the second edition the space devoted to these has been increased
(reflection properties, Green’s functions, functional methods, etc.).
In the presentation of the principles of quantum electrodynamics
the theory of renormalizations has been subjected to the most exten­
sive revision. Without claiming complete mathematical rigor we have
attempted to present the concept of renormalization from a single
simple physical point of view, avoiding purely prescription-like
methods of eliminating divergences, and making maximum use of
the general properties of quantum mechanical systems. In connection
with this the structure of the book has been somewhat altered. The
investigation of the scattering matrix together with the theory of
radiation corrections has been segregated into a separate chapter
(Chapter VII). The investigation of electrodynamic processes in the
first nonvanishing approximation, which involves neither the removal
of divergences nor renormalizations, is carried out in Chapters V and
VI, while higher order approximations are considered in Chapter VIII.
In presenting specific effects we have aimed, within reasonable
limits, at the greatest possible degree of completeness of results and
of detail of calculations. The number of different electrodynamic
phenomena described has been increased. In particular, the theory of
processes involving polarized particles, the method of impact param­
eters, etc., have been included.
We wish to express our sincere gratitude to V. Aleksin, V.
Bar’iakhtar, V. Boldyshev, D. Volkov, S. Peletminskii, R. Polovin,
and P. Fomin, who have aided us significantly in the preparation of
the second edition of this book.
A. I. A k h iezer , V. B. B erestetsk ii
vi
F ro m th e P r e fa c e to th e F irst E d itio n

At the present time a number of particles is known which corre­


spond to various quantum fields interacting with each other. However,
of the many types of physical interactions existing in nature the only
one, apart from gravitation, that has been studied in sufficient detail
is the electromagnetic interaction. The theory of the latter interaction
is the subject of quantum electrodynamics, to which the systematic
exposition of this book is devoted.
Since the electromagnetic interaction is the fundamental one in
the case of electrons and photons, quantum electrodynamics enables
us to explain and to predict a wide range of phenomena related to
the behavior of these particles. As for the application of quantum
electrodynamics to other particles (nucleons and mesons), it is con­
siderably restricted because of the essential role played by other types
of interactions (nuclear or meson interactions) in the case of those
particles. Therefore, problems relating to mesons are not treated in
this book, and the interaction of nucleons with the electromagnetic
field is treated only in the limit of low velocities.
The formulation of the fundamental equations of quantum electro­
dynamics, and even the very possibility of separating the interacting
fields into the electromagnetic and electron-positron fields, is based
on the fact that the interaction between these fields is a weak one.
This circumstance finds its expression in the smallness of the constant
a = e1/h e which characterizes the interaction. Therefore the inter­
action between the fields is treated in quantum electrodynamics as a
small perturbation, and the mathematical method employed in quan­
tum electrodynamics is perturbation theory in which all quantitative
results are expressed in terms of power series in a.
Since both the electromagnetic and the electron-positron fields are
systems with an unlimited number of degrees of freedom, the ap­
plication of perturbation theory gives rise to divergent expressions
characteristic of the present theory, which are absent only in the first
nonvanishing approximation of perturbation theory. The development
vii
Vlll FROM THE PREFACE TO THE FIRST EDITION

of quantum electrodynamics in recent years has permitted the estab­


lishment of principles for regularizing divergent expressions, so that
it has then become possible to calculate higher order approximations
(the so-called radiation corrections).
This progress is to a great extent associated with the new invariant
formulation of perturbation theory. Invariant perturbation theory has
made it possible to present the results in a compact and relativistically
invariant form, and this has allowed the rules for regularization to be
formulated. On the other hand, the use of invariant perturbation
theory has significant practical advantages over the earlier methods
even in the case of first-order calculations. Therefore the whole pre­
sentation in this book is based on invariant perturbation theory.
Although it is a completely satisfactory theory within a definite field
of physical phenomena, modern quantum electrodynamics has the
important drawback that in order to remove the divergences which
arise in the theory additional concepts must be introduced which are
neither contained in the fundamental formulation of the theory, nor
reflected in its basic equations. This state of affairs is apparently due
to profound causes. They are implicit in the fact that it is frequently
impossible to construct a closed theory of a limited set of phenomena
(in the present case, the purely electromagnetic ones) without taking
into account a wider class of interactions existing in nature.
We wish to express our gratitude to Academician L. D. Landau and
Professor I. Ia. Pomeranchuk and to the participants in the seminars
directed by them for discussions of a number of problems presented
in this book.
Contents

CH A PTER I

QUANTUM MECHANICS OF THE PHOTON


§ 1. The Photon Wave Function 1
1. Introduction. 2. The Photon Wave Function in
k-Space. 3. Energy. 4. Normalization of the Photon
Wave Function.

§ 2. Photon States of Definite Momentum 9


1. Photon Momentum Operator. 2. Impossibility of In­
troducing a Photon Wave Function in the Coordinate
Representation. 3. Plane Waves. 4. Polarization Density
Matrix for the Photon.

§ 3. Angular Momentum. Photon Spin . 17


1. Angular Momentum Operator. 2. Photon Spin Oper­
ator. 3. Photon Spin Wave Functions.

§ 4. Photon States of Definite Angular Momentum and


Parity . . . . 24
1. Eigenfunctions of the Photon Angular Momentum
Operator. 2. Longitudinal and Transverse Vector Spher­
ical Harmonics. 3. Parity of Photon States. 4. Expan­
sion in Spherical Waves. 5. Expressions for the Electric
and Magnetic Fields.

§ 5. Scattering of Photons by a System of Charges 36


1. Incoming and Outgoing Waves. 2. Effective Scattering
Cross Section. 3. The Optical Theorem. 4. Dispersion
Relations.

§ 6. The Photon Field Potentials . . . 46


1. Transverse, Longitudinal, and Scalar Potentials.
2. Longitudinally Polarized “Photon.” 3. Potentials for
Plane and Spherical Waves.
ix
X CONTENTS

§ 7. System of Photons 52
1. Wave Function for a System of Two Photons. 2. Even
and Odd States of Two Photons. 3. Classification of the
States of Two Photons of Definite Angular Momentum.
4. Wave Function for a System of an Arbitrary Number
of Photons.
§ 8. L-Vectors and Spherical Harmonics 62
1. Irreducible Tensors. 2. The Algebra of L-Vectors.
3. Spherical Harmonics.

CH A PTER II

RELATIVISTIC QUANTUM MECHANICS OF


THE ELECTRON
§ 9. The Dirac Equation 73
1. Spinors. Pauli Matrices. 2. Dirac Equations. Dirac
Matrices. 3. Unitary Transformations of Bispinors.
4. The Necessity for Four-Component Electron Wave
Functions. 5. Symmetric Form of the Dirac Equation.
Equation of Continuity. 6. Invariance of the Dirac
Equation. 7. Bilinear Combinations of the Components
of the Wave Function.

§ 10. Electron and Positron States. States of Definite


Momentum and Polarization 86
1. Solutions with Positive and Negative Frequencies.
2. The Charge Conjugation Transformation. 3. The Posi­
tron Wave Function. 4. Plane Waves. 5. Polarization of
a Plane Wave. 6. Polarization Density Matrix for the
Electron. 7. Averaging over Polarization States.

§ 11. Electron States of Definite Angular M omentum and


P a rity ...................................... 105
1. Orbital and Spin Functions. Spherical Spinors. 2. Wave
Function of a State of Definite Angular Momentum.
3. Parity of a State. 4. Expansion in Spherical Waves.

§ 12. Electron in an External Field . 112


1. The Dirac Equation with an External Field. 2. Sepa­
ration of Variables in a Central Field. 3. Asymptotic Be­
havior of the Radial Functions. 4. Behavior of Energy
Levels as Functions of the Potential Well Depth. 5. Elec­
tron in a Constant Homogeneous Magnetic Field.
CONTENTS

§ 13. Motion of an Electron in the Field of a Nucleus


1. Solution of the Radial Equations for the Coulomb
Field. 2. Wave Functions for the Continuous Spectrum.
3. Isotopic Level Shift. 4. General Investigation of the
Effect of Finite Nuclear Size.
§ 14. Electron Scattering
1. Spinor Scattering Amplitude. 2. Expression for the
Cross Section in Terms of Phases. 3. Polarization and
Azimuthal Asymmetry. 4. Scattering by a Coulomb
Field. 5. Small Angle Scattering.
§ 15. Nonrelativistic Approximation
1. Transition to the Pauli Equation. 2. Second Ap­
proximation. 3. Application of the Dirac Equation to
Nucleons.

CH A PTER III

QUANTIZED ELECTROMAGNETIC AND


ELECTRON-POSITRON FIELDS
§ 16. Quantization of the Electromagnetic Field
1. Four-Dimensional Form of the Field Equations.
2. Variational Principle. Energy-Momentum Tensor of
the Electromagnetic Field. 3. Expansion of the Potentials
into Plane Waves. 4. Quantization of the Electromag­
netic Field. 5. Use of the Indefinite Metric.
§ 17. Commutators of the Electromagnetic Field
1. Commutation Relations for the Potentials and the Field
Components. 2. Chronological and Normal Products of
Components of the Potential. 3. Singular Functions As­
sociated with the Operators □ and (□ — m 2).
§ 18. Quantization of the Electron-Positron Field .
1. Variational Principle for the Dirac Equation. Energy-
Momentum Tensor of the Electron-Positron Field.
2. Quantization Rules for the Electron-Positron Field.
§ 19. Anticom mutators of the Electron-Positron Field.
Chronological and Normal Products of Field Compo­
nents. Current Density
1. Commutation Relations for Field Components.
2. Chronological and Normal Products of Operators of
the Electron-Positron Field. 3. Electric Current Density.
xii CONTENTS

§ 20. G eneral P ro p e rtie s of W ave Fields . 214


1. Wave Functions of a Field and the Lorentz Group.
2. Irreducible Finite-Dimensional Representations of the
Lorentz Group. 3. Energy-Momentum Tensor and An­
gular Momentum Tensor. 4. Current Density Vector.
5. Relativistically Invariant Field Equations. 6. Wave
Equations for Particles of Spin Zero and Unity.
§ 21. Q uan tization of Fields. C onnection betw een Spin an d
Statistics .237
1. Nondefiniteness of the Charge in the Case of Integral
Spin and of the Energy in the Case of Half-Integral Spin.
2. Quantization of Fields for Integral and Half-Integral
Spin. Pauli’s Theorem. 3. Inversion of Coordinates and
Time Reversal.

CH A PTER IV

FUNDAMENTAL EQUATIONS OF QUANTUM


ELECTRODYNAMICS
§ 22. In te ra c tin g E lectrom agnetic an d E lectro n -P o sitro n
Fields . . . . 253
1. System of Equations for Interacting Fields. 2. La-
grangian. Energy-Momentum Tensor. 3. Field Equations
in Poisson Bracket Form. 4. Invariance Properties of the
Equations of Quantum Electrodynamics.
§23. E quations of Q u an tu m E lectrodynam ics in th e I n te r ­
action P ictu re. In v a ria n t P e rtu rb a tio n T h eo ry 268
1. Heisenberg and Schrodinger Pictures. Interaction Pic­
ture. 2. Transition to the Interaction Picture in Quan­
tum Electrodynamics. 3. Charge Conjugation Operator.
4. Perturbation Theory.
§24. T he S catterin g M atrix 290
1. The Scattering Problem and the Definition of the
Scattering Matrix. 2. Matrix Elements of Field Opera­
tors. 3. Representation of the Scattering Matrix as a Sum
of Normal Products. 4. General Relation between T-
and iV-Orderings. 5. Symmetry of the Scattering Matrix
under Time Reversal.
CONTENTS

§ 25. Graphical R epresentation of the Elements of the Scat­


tering Matrix. The Scattering Matrix in Momentum
Space
1. Graphical Representation of Normal Products. 2. Var­
ious Interaction Processes between Fields. 3. Transition
to Momentum Space. 4. Closed Electron Loops with an
Odd Number of Vertices. 5. Rules for Writing Down
Matrix Elements.
§ 26. Probabilities of Various Processes .
1. General Formula for the Probability. 2. Effective
Cross Section. 3. Summation and Averaging over Polari­
zation States of Electrons and Photons. 4. Probabilities
of Processes Involving Polarized Particles. 5. Probabili­
ties of Processes in the Presence of an External Field.
6. Feynman’s Notation.

vch a pter

INTERACTION OF ELECTRONS WITH PHOTONS


§ 27. Emission and Absorption of a Photon
1. General Expression for the Matrix Element. 2. Elec­
tric Multipole Radiation. 3. Magnetic Multipole Radia­
tion. 4. Selection Rules. 5. Angular Distribution and
Polarization of the Radiation.

§ 28. Scattering of a Photon by a Free Electron


1. Scattering Matrix Element. 2. Application of Conser­
vation Laws. 3. Differential Cross Section for Unpolar­
ized Particles. 4. Angular Distribution and Total Cross
Section. 5. Distribution of Recoil Electrons. 6. Scatter­
ing of Polarized Photons. 7. Scattering of Photons by
Polarized Electrons.

§ 29. Brem sstrahlung . . .


1. Perturbation Theory for an Electron Wave Function in
the Continuum. Incoming and Outgoing Waves. 2. Effec­
tive Cross Section for Bremsstrahlung. 3. Angular Dis­
tribution of the Radiation in a Coulomb Field. 4. Polari­
zation of the Radiation. 5. Spectrum of the Radiation.
6. Screening. 7. Radiative Energy Losses. 8. Exact
Theory of Bremsstrahlung in the Nonrelativistic Domain.
9. Exact Theory of Bremsstrahlung in the Extreme Rela-
xiv CONTENTS

tivistic Domain. 10. Radiation Emitted in Electron-


Electron and Electron-Positron Collisions.

§ 30. Em ission of P hotons of Long W avelength 413


1. “The Infrared Catastrophe.” 2. Investigation of the
Divergence in the Low Frequency Domain by Means of
the Scattering Matrix. 3. Relation between the Photon
“Mass” and the Minimum Frequency.

§31. P hotoeffect . . . . 429


1. Photoeffect in the Nonrelativistic Domain. 2. Photo­
effect in the Relativistic Domain.

§32. P ro d u ctio n of E lectro n -P o sitro n P a irs 438


1. Production of an Electron-Positron Pair by a Photon
in the Field of a Nucleus. 2. Exact Theory of Pair Pro­
duction by a Photon in the Field of a Nucleus in the
Nonrelativistic and Extreme Relativistic Cases. 3. Pair
Production by Two Photons. 4. Pair Production in a
Photon-Electron Collision. 5. Pair Production in a Col­
lision of Two Fast Charged Particles.

§33. A n n ih ilation of E lectron-P ositro n P a irs in to P h o to n s 457


1. Annihilation of a Pair into Two Photons. 2. Polariza­
tion Effects in the Two-Photon Annihilation of a Pair.
3. Annihilation of a Pair into One Photon. 4. Positro-
nium Decay. 5. Three-Photon Decay of Orthopositro-
nium. 6. Multiple Photon Production Accompanying the
Annihilation of a Pair.

§34. T he M ethod of E q u iv alen t P h o to n s 473


1. The Number of Equivalent Photons. 2. Brems-
strahlung from a Fast Electron in the Field of a Nucleus.
3. Radiation Emitted in an Electron-Electron Collision.
4. Pair Production by a Photon in the Field of a Nucleus.
5. Pair Production in a Collision of Two Fast Particles.

§35. S catterin g of a P h o to n by a B ound E lectro n , Em ission


of Two P hotons . . . 484
1. The Dispersion Formula. 2. Resonance Scattering.
3. Compton Scattering by Bound Electrons. 4. Emission
of Two Photons. The Metastable 2 State of the Hy­
drogen Atom.
CONTENTS XV

CH A PTER VI

RETARDED INTERACTION BETWEEN TWO


CHARGES
§36. Electron-Electron and Positron-Electron Scattering 499
1. Electron-Electron Scattering. 2. Positron-Electron
Scattering. 3. Scattering of Polarized Electrons and Posi­
trons. 4. Annihilation of an Electron-Positron Pair into
a /i-Meson Pair.
§37. Retarded Potentials . . . 509
1. Interaction Function for Two Charges. 2. General
Form of the Matrix Element. 3. Retarded Potentials and
Transition Currents.
§38. Interaction Energy of Two Electrons to Terms of
Order v2/ c 2 517
1. The Breit Formula. 2. Schrodinger Equation for a
Two-Electron System. 3. Interaction between an Elec­
tron and a Positron. 4. Exchange Interaction between an
Electron and a Positron.
§39. Positronium . . . 527
1. Hamiltonian Operator and the Unperturbed Equation.
2. Perturbation Operator. 3. Fine Structure. 4. Zeeman
Effect.
§40. Interna] Conversion of Gamma-Rays 537
1. Expansion of Retarded Potentials in Spherical Waves.
2. Conversion Coefficient. 3. Conversion in the E-Shell.
4. Effect of Finite Nuclear Size. 5. Effect of Electron
Shells on Radiation from the Nucleus.
§41. Conversion Accompanied by Pair Production. Excita­
tion o f Nuclei by Electrons 554
1. Conversion of Magnetic Multipole Radiation. 2. Con­
version of Electric Multipole Radiation. 3. Excitation of
Nuclei by Electrons. 4. Monoenergetic Positrons.
§42. Coulomb (M onopole) Transitions 565
1. Reduction to the Static Interaction. 2. Conversion and
Nuclear Excitation in the Case of an EO-Transition.
xvi CONTENTS

CH A PTER VII
INVESTIGATION OF THE SCATTERING MATRIX
§ 43. Properties of Exact Solutions of the Equations of
Quantum Electrodynamics. Propagators 571
1. Stationary States of a System of Interacting Fields.
2. Propagators and Their Spectral Representation.
3. Connection between Propagators and the Scattering
Matrix. Integral Equations for Propagators. 4. Electro­
magnetic Mass of the Electron.

§ 44. Structure of the Scattering Matrix . 593


1. Self-Energy Parts of Diagrams. 2. Vertex Parts of
Diagrams. 3. Renormalization of Electron Mass.

§ 45. Renorm alization of Electron Charge 605


1. Physical Charge of the Electron. 2. Renormalization
of Propagators and Vertex Parts. 3. Three-Photon
Vertex Parts. 4. Renormalization of Matrix Elements.
5. Formulation of Perturbation Theory as an Expansion
of Powers of ec.

§ 46. Divergences in the Scattering Matrix and their


Removal . . . .619
1. Divergences in Irreducible Diagrams. 2. Introduction
of a Cut-Off Momentum. 3. Convergence of Regularized
Expressions for Irreducible Vertex Parts and Self-Energy
Parts. 4. Convergence of Regularized Quantities in the
Case of Reducible Diagrams.

§ 47. Evaluation of Self-Energy and Vertex Parts . 631


1. Evaluation of Integrals over Four-Dimensional Re­
gions. 2. Second Order Electron Self-Energy Part.
3. Second Order Photon Self-Energy Part. 4. Third
Order Vertex Part in the Case of External Electron Lines.
5. Third Order Vertex Part in the Case of One External
Electron Line.

§ 48. Functional Properties of Green’s Functions. Limits of


Applicability of Quantum Electrodynamics 657
1. Expansion Parameters of Perturbation Theory. 2. Zero
Order Approximation in the Expansion in Powers of e2.
3. Integral Equations for the Zero Order Approximation.
CONTENTS

4. The Renormalization Group. 5. Derivation of Asymp­


totic Expressions for the Green’s Functions with the Aid
of Differential Equations of the Renormalization Group.
6. The Problem of Closure of Quantum Electrodynamics.
§ 49. Generalized Green’s Functions
1. Green’s Functions in the Presence of External Fields.
2. Green’s Function for Two Electrons. Equation for
Bound States of the Electron-Positron System. 3. Equa­
tions for Green’s Functions in Terms of Variational De­
rivatives. 4. Expressions for Green’s Functions in Terms
of Functional Integrals.

C H A PT ER V III

RADIATION CORRECTIONS TO ELECTRO­


MAGNETIC PROCESSES
§ 50. Effective Potential Energy of the Electron. Radiation
Corrections to the Electron Magnetic Moment and to
Coulomb’s Law
1. Energy of Interaction of the Electron with the Elec­
tromagnetic Field Taking into Account Corrections of
Order «. 2. Radiation Corrections to the Electron Mag­
netic Moment. 3. Radiation Corrections to Coulomb’s
Law.

§ 51. Radiation Corrections to Electron Scattering


1. Electron Scattering by the Coulomb Field of a Nucleus
in the Second Born Approximation. 2. Differential Cross
Section for the Scattering of an Electron by the Coulomb
Field of a Nucleus taking into Account Radiation Correc­
tions of Order a ■ 3. Elimination of the Photon “Mass”
from the Scattering Cross Section. 4. Removal of the
Infrared Divergence for an Arbitrary Scattering Process.
5. Scattering of High Energy Electrons by an External
Field. 6. Radiation Corrections to Electron-Electron and
Electron-Positron Scattering.

§ 52. Radiation Corrections to Photon-Electron Scattering,


to Pair Creation and Annihilation, and to Brems-
strahlung
1. Radiation Corrections to the Compton Effect. 2. Lim­
iting Cases of Low and High Energies. 3. Radiation
xviii CONTENTS

Corrections to Two-Photon Pair Annihilations. 4. Radia­


tion Corrections to Bremsstrahlung. 5. Radiation Correc­
tions to Photon Production and Single Photon Annihila­
tion of Pairs.
§ 53. R ad iatio n C orrections to A tom ic Levels .751
1. Radiation Shift of Atomic Levels. 2. Radiation Shift
of the Levels of ^-Mesohydrogen. 3. Natural Line
Widths. 4. Photon Scattering near Resonance.
§ 54. P h o to n -P h o to n S catterin g an d th e L a g ra n g ia n fo r th e
E lectrom agnetic Field . 764
1. Photon-Photon Scattering Tensor of the Fourth Rank.
2. Photon-Photon Scattering. 3. Connection between the
Photon-Photon Scattering Cross Section and the Radiation
Corrections to the Lagrangian of the Electromagnetic
Field. 4. Exact Expressions for the Lagrangian of the
Electromagnetic Field.
§ 55. P h o to n S catterin g by th e C oulom b Field o f a N ucleus 792
1. General Expression for the Cross Section for Photon
Scattering by a Constant Electromagnetic Field. 2. Rela­
tion between the Forward Scattering Amplitude for a
Photon and Pair-Production by a Photon in the Field of
a Nucleus. 3. Momentum Distribution of Recoil Nuclei
Accompanying Pair Production by a Photon in the Field
of a Nucleus. 4. Angular Distribution of Recoil Nuclei
and Total Cross Section for Pair Production by a Photon
in the Coulomb Field of a Nucleus. 5. Small Angle Co­
herent Scattering of Photons by the Field of a Nucleus.

C H A PT ER ix
ELECTRODYNAMICS OF PARTICLES OF
SPIN ZERO
§ 56. Field E quations fo r S calar P a rticle s 819
1. First Order Equations. 2. Quantization of the Free
Scalar Field. 3. Commutators of the Field. Vacuum Ex­
pectation Values of Products of Field Components.
§57. T he S catterin g M atrix in S calar E lectrodynam ics 827
1. The Interaction Picture. 2. Rules for Calculating Ele­
ments of the Scattering Matrix. 3. Divergences of the
Scattering Matrix.
CONTENTS xix

§ 58. Scattering of Scalar Particles . 835


1. Scattering of Scalar Particles by the Coulomb Field of
a Nucleus. 2. Scattering of a Charged Scalar Particle by
a Scalar Particle.
§ 59. Scattering of a Photon by a Scalar Particle. Brems-
strahlung Photons from a Scalar Particle 838
1. Scattering of a Photon by a Scalar Particle. 2. Brems-
strahlung from Scalar Particles.
§ 60. Production and A nnihilation of Pairs of Scalar
Particles . . . 842
1. Production of Pairs of Scalar Particles by a Photon in
the Coulomb Field of a Nucleus. 2. Production of a Pair
of Scalar Particles by Two Photons. 3. Two-Photon An­
nihilation of a Pair of Scalar Particles. 4. Annihilation
of Pairs of Scalar Particles into Electron-Positron Pairs
and the Inverse Process.
§ 61. Polarization of the Vacuum in the Case of Charged
Scalar Particles 847
1. Vacuum Polarization Tensor for Scalar Particles.
2. Correction to Coulomb’s Law. 3. Photon-Photon Scat­
tering. Radiation Corrections to the Lagrangian of the
Electromagnetic Field.

Concluding Remarks 852

R eferences 855

Subject Index 863


CHAPTER I

Quantum Mechanics of the Photon

§ 1. The Photon Wave Function


1.1. Introduction
The corpuscular properties of light were historically the first fun­
damental fact which served as a basis for the development of quantum
theory. The Planck-Einstein relation between the energy w of a quantum
of light, the photon, and the frequency co of the electromagnetic field
corresponding to it is expressed as w = tuo and is historically the first
relation containing the quantum constant ti.
The systematic quantum mechanics of the atom was developed be­
fore that of the photon and this situation has a profound physical reason.
Atomic particles, the electrons and the nuclei, have rest masses different
from zero. For them there exists a range of energies, small compared
to their rest energy, in which relativistic effects may be neglected. Since
the rest mass of the photon is zero, there exists for it no nonrelativistic
energy region, and the quantum mechanics of the photon must neces­
sarily be relativistic from the outset.
As the counterpart of a particle, quantum mechanics introduces
a field of one or more wave functions which determine the probability
distribution and the expectation values of the various physical quanti­
ties related to the partticle. The wave functions satisfy a certain system
of differential equations which determine the nature of the motion of
the particle.
In order to transpose this description into the relativistic domain
we must, first of all, take into account the requirements of the principle
of relativity. They amount to the requirement that the field equations
must be invariant under Lorentz transformations. However, this by
itself is not sufficient for a unique determination of the form of the
field equations which would give expression to the individual properties
of a given particle. In the case of photons the setting up of the equations
is facilitated by the existence of a classical analog, viz., of the equations of
[1]
2 QUANTUM ELECTRODYNAMICS

the classical electromagnetic field. It is natural to use Maxwell’s equations


as the quantum mechanical equations of motion for the photon and
it follows that the wave properties of the photon will coincide with
the properties of the electromagnetic field. As will be shown later, this,
together with the relation w = ha>, suffices for the formulation of
a theory of photons and of their interaction with charged particles.
Our first problem is the study of photons in the absence of electric
charges. Although properties of particles become apparent only as a result
of their interaction with other particles, such an investigation is necessary
as a preparatory stage for the study of interactions.

1.2. The Photon Wave Function in k Space


The electromagnetic field is described by the electric field vector
E and the magnetic field vector H, which in the absence of charges
satisfy Maxwell’s equations for a vacuum:
dH
curl E =
IT ’
di v H = 0,
( 1. 1)
dE
curl H =
~dT’
div E 0.
(Here and subsequently we make use of the system of units in
which the velocity of light is equal to unity, c = 1.)
In accordance with the foregoing we shall interpret the vectors
E and H as quantities describing the quantum mechanical photon state.
In order to give such a “corpuscular” interpretation to the system of
equations (1.1) we compare this system with the Schrodinger equa­
tion in ordinary quantum mechanics. This may be done conveniently
if we first subject equations (l.l) to a Fourier transformation with
respect to the space coordinates r, i.e., if we go over into k space. By
writing E and H in the form1
E= f Ekeikrdk,
, (1-2)
H= Hkeikr dk,
1 Following the authors’ usage, the scalar product of two three-dimensional
vectors will be denoted by boldface letters standing next to each other without an
QUANTUM MECHANICS OF PHOTON 3

we can easily see that in virtue of (1.1) the Fourier components Ek and
Hk satisfy the following system of equations:

Hk = - i [ k , E k\,
kH k = 0,
(1.3)
Ek = i[ k ,H k],
kEh - 0,
where dots denote differentation with respect to time. (For the sake
of brevity we shall omit here and subsequently the argument t in the
functions Ek = E {k ,t) and Hk = H (k, t).) To this system we must
also add the condition that the fields be real:
E -k = E*,
(1.4)
H - k = H *.

Instead of the two vectors Eh and Hk we can use the vectors Ek and
Ek by eliminating Hk with the aid of (1.3):

(1-5)

It is convenient to eliminate the necessity of taking into aecount


the requirement of reality. In order to do this we introduce a transfor­
mation which automatically guarantees that the relations (1.4) will
be satisfied:
Ek = N (k )(fk+ f* k),
Ek = —ikN (k) (fk—f * k).
where N(k) is a certain normalizing factor, which, as we shall see later,
it is convenient to choose equal to

( 1 '7 )

The substitution (1.6) means that instead of the two functions Ek and
Ek, which in virtue of (1.4) are actually defined in a half-space, we intro­
duce the single function f k, which is independently defined over the
whole &-space.
intervening dot, and either with or without parentheses: thus kr or (kr ), but not
k r. Similarly, the vector product will be denoted by the square brackets [A, B] with
or without a comma, but not by A x B .
4 QUANTUM ELECTRODYNAMICS

If we make use of (1.6), the expansion (1.2) will assume the following
form:
£ '= g + ( y * , H = < q

(y= f N{k)fkeikrdk, ^ ^

$ = /* ]< * < * •

We can easily obtain the equation satisfied by f k. By eliminating


Hk from (1.3) we obtain
d2
\-k2\E k = Q ,
dt2
which may be rewritten in the following form:

This second-order equation may be transformed into a first-order


equation for f k. Indeed, by utilizing the relation

- ik^Ek = -2 iN (k ) k f k,

which follows from (1.6), we obtain

!~ d f= k J *■ ° - 9)
The substitution of expressions (1.5) and (1.6) into (1.3) yields
* fk = 0. (1.10)
Equations (1.9) and (1.10) may be combined into one equation. On
multiplying (1.10) by k jk and on adding it to (1.9), we obtain

(1. 11)

where
QUANTUM MECHANICS OF PHOTON 5

It follows from (1.11) that

- ( * / ; > = o.

Therefore (1.10) will always hold if it holds at some initial time.


Equation (1.11) together with the initial condition (1.10) are equiv­
alent to the system of Maxwell’s equations.
Equation (1.11) has the form of a Schrodinger equation in which
w is the Hamiltonian operator. (Here and subsequently we shall make
use of the system of units in which ft = 1.) The eigenvalues of this opera­
tor are equal to k. This would amount to nothing other than the quantum
relation between the energy and the frequency of the photon, if we
could identify the Hamiltonian operator formally introduced in the
foregoing with the physical operator for the photon energy. We shall
show later that such an identification can be justified, and that the
function f k can be interpreted as the photon wave function in A>space
in the usual quantum mechanical sense of this word. We shall also
be able to define in A>space operators for other physical quantities
referring to the photon, for example, the operators for the momentum,
the angular momentum, etc.

1.3. Energy
We shall show that the operator w which we have just introduced
can indeed be interpreted as the operator for the photon energy. In
order to do this we introduce the expression for the energy w of the
electromagnetic field corresponding to a given photon as

w = \ J (E2+ H 2)dr. (1.12)

(We make use here of Heaviside units for E and H.) This expression
represents a space integral of quantities quadratic in the field vectors.
On the other hand, space integrals of expressions quadratic in the wave
function are interpreted in quantum mechanics as expectation values
of the corresponding physical quantities. Since we are considering
the field corresponding to a single photon, it would appear that the
natural generalization is the interpretation of expression (1.12) as the
expectation value of the photon energy.
6 QUANTUM ELECTRODYNAMICS

We show that w may be represented in the form

w = J f ^ w f kdk, (U 3 )
where w is determined by formula (1.11).
In order to do this we substitute expansion (1.2) into (1.12):

w = i f {EkEv + H kHv} e"k+k)rdkdk'dr.

After carrying out the integration over r with the aid of the re­
lation
J el<k+k')rd r = (2tOs<3(*+*'),
where <5(k-\-k') is a three-dimensional (5-function, and on utilizing (1.5),
we obtain

« = 4*» J k £ _ , + - i

We express and Ek in terms of f k in accordance with (1.6). This


yields
w = 16n* f N * f* fk dk.

If we choose for N(k) the value (1.7), we obtain relation (1.13).


We consider a monochromatic solution of equation (1.11):
f k ^ f ( k , t ) = f 0( k ) e ^ , (1.14)
where co is the eigenvalue of the photon energy operator w. It is ob­
vious that f Q(k) differs from zero only when k = co, and the expression
for the energy (1.13) assumes the form

w = <a j f * f k dk.
We demand that this equation be identical with the quantum relation
vv = co and then obtain the condition for the normalization of the
photon wave function
/ / ? / . < * = 1- 0.15)

Thus, if the condition for the normalization of the wave function


(1.15) is satisfied, the energy of the electromagnetic field becomes
identical with the photon energy.
QUANTUM MECHANICS OF PHOTON 7

1.4. Normalization o f the Photon Wave Function


In practice it is more convenient to deal with a Fourier series rather
than with a Fourier integral, i.e., with a discrete rather than with a con­
tinuous spectrum. For this we must introduce a certain normalizing
volume Q and assume that the whole field is confined within it. Certain
boundary conditions, which select from all the solutions of Maxwell’s
equations a discrete spectrum of solutions that satisfy these boundary
conditions, can be specified over the boundary of the region Q. It is
simplest to choose the volume in the form of a cube of side a and to
postulate that the boundary conditions amount to the conditions of
triple periodicity of the fields E and H with the period a along each
one of the cartesian axes. If these conditions are satisfied, then those
fields are selected for which the components k i of the propagation
vector satisfy the conditions

where wf are integers. Thus, A>space is divided into cells of volume


(2ji )3
(1.16)
Q
One allowed value of k corresponds to each such cell. Therefore, the
number of values of k lying within a certain range dk is equal to
dk
(1.16')
A
Integration over &-space is replaced by summation over these cells;
i.e.,
f F(k)dk = A - £ F(k),
J k

where the summation is carried out over all the allowed discrete
values of k.
The expansion (1-8) assumes the form

<&= A 2 ,N { k ) f ke»"- (1-17)


k
i.e., the coefficients in the expansion of (5 in a Fourier series are
the quantities A -N -f k.
8 QUANTUM ELECTRODYNAMICS

The normalization condition (1.15) can now be written in the follow­


ing form:
= (I-'*)
k a
If the volume Q chosen is sufficiently large, then the relation (1.16)
will depend neither on the shape of the volume Q, nor on the nature
of the boundary conditions. In other words, relation (1.16) is a uni­
versal asymptotic relation valid for ka^> 1, where a is a quantity charac­
terizing the linear dimensions of the region.
Later we shall show that in the solution of any physical problem,
the quantities A and Q appear only in the intermediate calculations
and disappear from the final result.
The transition from the Fourier integral to the Fourier series may
be carried out only partially, with respect to only one of the coordinates.
We assume, for example, that the boundary conditions are such that
A-space is divided into spherical shells of thickness <5. Then, on making
use of spherical polar coordinates, we obtain

f F(k)dk = f F (k)k2dkdo -> d £ k 2f F(k)do.


k
The expansion (1.8) in this case takes the form

6 = <5£ k 2N(k) f f keikrdo, (1.19)

while the normalization condition takes the form

= ( 1-20)

The division of A>space into spherical shells can be carried out


by choosing the volume Q in the form of a sphere of large radius R and
postulating that the field vanishes over the surface of the sphere. It
can be shown that in this case

= ( 1. 21)

The number of different possible values of k in the range dk is evidently


equal to
( 1. 21 ')
QUANTUM MECHANICS OF PHOTON 9

If we assume that the two-dimensional space of solid angles is also


divided into cells 50, then the expansion (1.19) is transformed into
a three-dimensional series. In this case formulas (1.17) and (1.19) become
identical if we postulate that in the latter formula the cell A is chosen
equal to A — k? <5<50.

§ 2. Photon States of Definite Momentum

2.1 Photon Momentum Operator


As is well known, the momentum of the electromagnetic field p is
given by the following expression:

p = J [EH] dr. (2.1)


We shall show that when the normalization condition for the photon
wave function (1.15) is satisfied, the vector p can be interpreted as
the expectation value of the photon momentum.
In order to do this we express p in terms of f k. Substitution of ex­
pansion (1.2) into (2.1) yields

p = J [Ek Hk] e'{h+k')r dk dk' dr

= (2n)3f [EkH _k] d k = j ~ - ( E kE^k)dk.

On expressing Ek and Ek in terms of f k , in accordance with (1.6), we


obtain
p = Qjif J

On replacing k by — k in the integrand, we can easily see that the inte­


gral of the last two terms vanishes, while the integrals of the first two
terms are identical. By using the definition (1.7) of N(k) we obtain
finally
J
P = fLV kadk, (2.2)

where the subscript a = 1, 2, 3 denotes the components of the vector f k


and summation over a is implied.
It follows from (2.2) that we are justified in identifying the operator
indicating multiplication by the propagation vector p = k with the
10 QUANTUM ELECTRODYNAMICS

photon momentum operator, and /r-space with momentum space.


The quantity f * f k may be interpreted as the probability density of the
photon having momentum k, while relation (2.2) may be interpreted
as the usual quantum mechanical expression for the expectation value.
Thus, the normalization condition (1.15) acquires a simple and natural
meaning.

2.2. Impossibility of Introducing a Photon Wave Function in the Coor­


dinate Representation
It might appear that by subjecting the function f k to the inverse
Fourier transformation

it should be possible to obtain the photon wave function F in the coor­


dinate representation (121); cf. also (148).1 Because of the normalization
condition (1.15) for f k, the function F will also be normalized in the
usual manner:
J F* F dr — 1.

However, the quantity F*F will not have the meaning of the prob­
ability density of finding the photon at a given point of space. Indeed,
the presence of a photon can be established only as a result of its in­
teraction with charges. This interaction is determined by the values of
the electromagnetic field vectors E and H at the given point, but the
latter are not determined by the value of the wave function F at that point
but by its values over all space. Indeed, the Fourier components of
the field vectors expressed in terms of f k contain in accordance with
(1.6) and (1.7) the factor ] / k . Formally this can be written in the form
4 j ------ 4 t-------
(S = y —A F , where A is the Laplacian operator; but }/ —A is in
fact an integral operator, and therefore the relationship between E
and F is not a local one, but an integral one; in other words, F(r, t)
is determined not by the value of the field E(r, t) at the same point,
but by the field distribution over a certain region whose effective di­
mensions are in order of magnitude equal to a wavelength. As a re-

1 References to the literature appear in parentheses. The complete list of refer­


ences is given at the end of the volume.
QUANTUM MECHANICS OF PHOTON 11

suit of this, the localization of a photon in a region smaller in order


of magnitude than a wavelength has no meaning, and the concept
of probability density for the localization of a photon does not exist.
This result is closely associated with the behavior of particle den­
sity under Lorentz transformations, since it is impossible to construct
a bilinear expression in the electromagnetic field vectors which would
form a four-vector satisfying an equation of continuity.
We note that in practice it is often sufficient (instead of the equation
of continuity for the probability density) to utilize the equation of
continuity for the energy density

iji+ d iv j = 0 , (2.3)
ot
where = \ (E 2-{-H2) is the energy density, and s = [EH] is the energy
flux density. However, these quantities do not form a four-vector.

2.3. Plane Waves


We return to the photon wave function in momentum space f k
= f { k , t ) . Equation (1.9) defines its time dependence
f ( k , 0 = fo(k) e~ikt.

The time-independent function / 0 (A:) is restricted only by the transversal-


ity condition (1.10). By utilizing the method described in subsection
1.4 we shall treat A>space as discrete and shall consider the photon
state for which f 0 (A) differs from zero only at point p. This is a state
of definite momentum, and the wave function corresponding to it
will be an eigenfunction of the momentum operator p. By taking into
account the normalization condition (1.18), we can write the eigen­
function of the momentum operator in the form

f( k , t) = 6k , p g—iOit (2.4)
j/zl
where cu = \p\ and e is the photon polarization vector whose absolute
value is equal to unity and which is perpendicular to p: \e\2 = 1, ep = 0.
(The phase factor i is introduced into (2.4) for convenience.)
For a given p two linearly independent vectors e are possible. We
take the z-axis directed along p. Then for these two vectors we can
12 QUANTUM ELECTRODYNAMICS

choose the unit vectors Xi ancl X2 corresponding to the x-axis and


the y-axis:
Xl*=l. XlV= X U = Q >
X2 y = l , X2X= X2 Z= 0, (2.5)

XiX2= 0 .

If e — (p = 1,2), then the photon is linearly polarized.


We can also choose the following two mutually perpendicular unit
vectors x( and X2 :

y' - 1 xL = o,
Xu f 2 ’ Xl* / 2 ’
f _ n (2-6)
X2z
f 2 ’ f 2 ’
x(*x; = o.
In this case if e = x^> the photons will be circularly polarized.1
Thus, states of definite momentum are twofold-degenerate. In order
to specify a state uniquely we have to specify its type of polarization.
Therefore, the function f ( k , t) should be provided with two subscripts
(quantum numbers) p and p (p = 1, 2). The quantities p x ,p y ,pz, and
p form a complete set of quantum numbers for the photon. (Of course,
the energy is thereby also determined, since co = p.) The system of func­
tions f pp(k, t) is a complete orthonormal system in terms of which
an arbitrary transverse function f ( k , t ) can be expanded:

f ( k , t ) = JT 1ap/lf p/i(k, <) = Z ene~ikt’


p, p p * (2.7)

/ f f p p dk = - i ] / A (fe*)em .

By making use of (2.4) it is possible in accordance with (1.16) and


(1.17) to represent the electric and magnetic fields corresponding to

1 We shall refer to xl as right circular and to X2 as left circular polar­


ization. (In optics the converse definition is usually adopted).
QUANTUM MECHANICS OF PHOTON 13

a photon state of definite momentum p a n d . polarization p, in the


following form:

In quantum mechanics a plane wave is frequently normalized in


such a way as to correspond to unit flux. Similarly, we can normalize
the photon field to unit energy flux density. The field corresponding
to such normalization has the following form:

<S - -— e e ^ P '- ^ , £ = ^ (v j . (2.9)

Indeed, in the expression for the energy flux density s = [EH] we


may omit the terms [(£,£] and [(i*.<}*], which vanish when averaged
over time, and we can take
* = [6 £*]+[(£*&]■ (2-10)
On substituting (2.9) into this expression we obtain s = pjp.
The expansion (2.7) corresponds to the expansion of an arbitrary
electric field satisfying Maxwell’s equations (1.1) into plane polarized
waves:
(y = V1Q fv
P
where the coefficients ap/l are determined in accordance with (2.7),
while (v have the form (2.8). Similar expansions hold also for the
magnetic field.
2.4. Polarization Density Matrix for the Photon
We have just given a complete quantum mechanical description
of a photon state. In accordance with (2.4) the photon state for a given
momentum is characterized by the polarization unit vector e. On ex­
panding e in terms of the unit vectors (2.5) or (2.6),
e = e 1Xx+e2xs , (2.11)
we can say that the state of polarization is determined by the pair of
complex numbers el and e2; the quantities |e]|2and |c2|2 represent prob-
14 QUANTUM ELECTRODYNAMICS

abilities of a definite (linear or circular) polarization of the photon,


determined by the unit vectors Xi and / 2- Since e1 and e2 are related
by the normalization condition
kil2+ |e2|2= 1
and since, moreover, the common phase factor of e is arbitrary, the
polarization is determined by two real parameters and (2.11) can be
written in the following form:
e = Xi cosa+X 2 sina-e^. (2.12)
If Xi ar|dX 2 are defined in accordance with (2.5), then /? = 0 denotes
linear polarization at an angle a to the x-axis, /? = ±7r/2 and a = n/4
denote circular polarization, and arbitrary a and /? correspond to ellip­
tical polarization.
A more general case is that of an incomplete quantum mechan­
ical description, when a definite wave function cannot be ascribed to
the photon. For example, if the photon has been previously scattered by
an electron, then there exists only the wave function for the combined
photon-electron system whose expansion in terms of the free photon
wave functions contains the electron wave functions. We shall confine
ourselves to the case when the photon has a definite momentum, i.e.,
when there exists a wave function (2.4), but the polarization state
cannot be specified definitely, since the coefficients in (2.11) depend
on the parameters characterizing the other system.
Such a photon state is referred to as a state of partial polarization.1
It can be described by means of a density matrix (c., for example,
(119, §12), (120, §5))
Qvv = ef>e* (ti ’ v = 1>2), (2-13)
where the bar denotes averaging over appropriate parameters.
From the definition (2.13) it follows that £ is a Hermitian matrix
= 9*. ■
From the normalization condition for e it follows that
Sp Q = 5ll"i“522 = 1 •

1 Compare the description of partially polarized light in classical electrodynam­


ics, for example, in (118), §50.
QUANTUM MECHANICS OF PHOTON 15

Thus, the density matrix depends on three real parameters. It can


be written in the following form:

_ 1 / l + f . fi
6 ¥ Ui+if2 i - f , r
or

e = }(i + l v , ) , (2.14)
1=1

where

are four matrices forming a complete system of linearly independent


two-dimensional matrices. (We note that the matrices r, are identical
with the Pauli matrices cf. §9.)
By using (2.14) it is possible to express ^ in terms of g, viz.,
^=Sp(Q-Tj). (2.15)
The quantities called the Stokes parameters, can be directly de­
termined experimentally. For example, let the unit vectors Xi and
X2 in (2.11) correspond to the linear polarizations (2.5). Then £3 will
be determined by the value of the probability of polarization along the
x-axis:
(?n = 2 0 + £ 3) •
In order to determine it is necessary to obtain the probability of
polarization along an axis making an angle of 45° with the x-axis.
Indeed, the transformation to such an axis corresponds to the unit
vector transformation
Xl+Xa x ;-x z
Xi = X2 =
l/2 ’ l/2
By writing (2.11) in the form
e = ei X l+ e2X2
and by defining g' in accordance with (2.13), Q,ltv= , we ob­
tain for the probability of linear polarization corresponding to %' the
following expression:
(?n = 4 1 + 6*2 2 + 0 1 2 + 0 2 1 ) = 4 O+ f i ) -
16 QUANTUM ELECTRODYNAMICS

Finally, in order to obtain £2 it is necessary to consider circular


polarization. Let Xi and Xz be the unit vectors corresponding to the
circular polarizations (2.6); i.e.,
X1 + X 2 X1 - X 2
Xi = X2 =
]/2 |/2
The probability of polarization corresponding to the unit vector Xi'
will be expressed by analogy with the foregoing in the following form:
t?n — ^[f?ll4~<?22~b* (@12 £?2l)] — 2 (^~b £2) •
We see that the determination of the polarization of a beam of
photons requires the measurement of two linear polarizations (at an
angle of 45° to each other) and of one circular polarization. We note
that if we had taken the initial matrix (2.13), as determined by the unit
vectors (2.6), then pn = -t(l+ £ 3) would have given us the probability
of circular polarization, while = |(1 + £ J) an^ £>n = K l + ^ 2) would
have given us the probabilities of the linear polarizations.
Since the probabilities must be positive and smaller than unity,
it follows from the expressions for gn , £>X1, and that |£;] ^ 1. On
noting that
Det e = K W 2) = R l “R P - t e R j ( R ? ) > 0
it may be concluded that

+ + I-
As an example we consider some limiting cases. When = 0,
(? = }■ In Ibis case the probability of any arbitrary polarization is
equal to -2-. Such a state is said to be completely unpolarized.
When i 2 = 1 we can write g, in the following form:
i 3 — cos2a, = sin2a cos/3, £2 = sin2a sin/S.
In the case of these values of the parameters the matrix q is identical
with the matrix obtained on the basis of expression (2.12). Therefore, it
describes a state of complete polarization.
In the case of partial polarization the density matrix can be regarded
as a linear combination of two density matrices corresponding to a com­
pletely unpolarized and a completely polarized state. Indeed, let
£ ; = FVj,
QUANTUM MECHANICS OF PHOTON 17

where

Then expression (2.14) can be written in the form

e= jd - P ) + ^ + I ^ ! j-
i
The coefficient P is referred to as the degree of polarization of the
photon.

§ 3. Angular Momentum. Photon Spin

3.1. Angular Momentum Operator


We now express the angular momentum M of the electromagnetic
field corresponding to one photon in terms of the wave function for
this photon f k. We identify this angular momentum with the expecta­
tion value of the angular momentum of the photon in this state f k. The
angular momentum of the field is determined, as is well known, in the
following manner:
M = j\r[E H ]\d r. (3.1)

By substituting into this formula the Fourier expansions for the fields,
we obtain
M = f \r[EkHk,]]ei(k+k')rdkdk'dr.

We first carry out the integration over r :

J reKk+k’)r J el(k+k')r dr = —i{ ln) zWk'd{kJr k ' ) ,

where V*. corresponds to differentiation with respect to the variables


k ' . We next carry out the integration over k'\

f d k ' \ V , . d ( k + k ' ) [ E , H ^ = - f d k '5 ( k + k ') \\,.[ E u H A -


On replacing H in the above equation by the expression (1.5) connecting
Hk with Ek, we obtain, after some simple rearrangements,

4
i(k'Ek) curlfc.
k'2
18 QUANTUM ELECTRODYNAMICS

Thus, we obtain

M = a-!)5 f rf*rf*'<3(* + A'){T*'Vf

- [ A Eh] + (k'Et) curl4. | t j .

After carrying out the integration over k!, and on noting that the last
term vanishes in the course of this integration since kEk = 0, we obtain

M = (2n f J j
d k ^ k V ( j ^ Eic) f-fcj ~ 5

where V denotes differentiation with respect to k; the superscript c on


Eh indicates that in carrying out the differentiation this quantity should
be regarded constant. Finally, we express Ek and Ek in terms of f k in
accordance with (1.6). This gives

M = j f d k { { W ( / ! ‘> / ? - / 3 ‘>£t + f 3 ' > f * - / Z ‘>f.k)]

The terms that contain different arguments, i.e., k and —k, vanish
on integration. For example, the integral J [f_kf k]dk on replacement
of the variable of integration k by - k changes into J [fkf _ k]dk, i.e.,
into an expression of opposite sign. The integral J [kV(f£c)f _ k)]dk on
replacement of k by - k changes into J [kV(fic^ f k)]dk, whereas inte­
gration by parts leads to the same expression but of opposite sign.
Also, on noting that the integrals of terms which depend on the argu­
ment - k do not change when k is replaced by -k, and, on integrating
by parts,
/ [ « ( / / ■ ' / , • ) ] * = - f [kV ( / * “ /„)] dk,

we finally obtain the following expression for the angular momentum:

m = / dk [W(A»i“>A )]-i [/*/„]}


or, in terms of its components,

= p * / / { - i [ t V l , / r 'V J . (3.2)
where eapy is the unit antisymmetric tensor of the third rank.
QUANTUM MECHANICS OF PHOTON 19

3.2. Photon Spin Operator


We introduce the vector operator s defined by the result of the
operation of its components on the components of some arbitrary
vector / according to the formula

(Sa /) f l= —(S{if)a= —ieapyfy (3.3)


On noting that the vector product of two vectors f and g is expressed
in terms of s in the following manner:
[gf]a = igpsjp, (3.4)
we write the angular momentum defined by formula (3.2) in the form

M = J f * { - i [kV]+s}fadk. (3.5)

We see that the expression for the expectation value of the angular
momentum of the photon (3.5) has the structure of quantum mechan­
ical expectation values, and that for the angular momentum opera­
tor for the photon we must take the operator

M = - i [*V]+s. (3.6)
We now show that the operator (3.6) is identical with the operator
for the infinitesimal rotation of the vector field multiplied by i. In
order to do this we consider an infinitesimal rotation of &-space about
the origin (189). As a result of such a rotation the radius vector of
each point in &-space is changed by the increment
8k = [dak],
where (5a = v<5a is the vector characterizing the infinitesimal rotation,
8a is the angle of rotation, and v is the unit vector along the axis of
rotation. This infinitesimal rotation of A>space transforms the vector
field f( k ) into the field f '( k ) related to f ( k ) by the following expression:

f'(k+ 8 k ) = m + [ 8 * f( k ) ] .

As a result of this the field at the point k is altered by the amount


8f(k) = / ' ( k ) - f ( k ) .
Since
f ( k + 8 k ) = f '( k ) + ( 8 k V ) f '( k ) = f'(k)+(8oi[kV])f(k)
20 QUANTUM ELECTRODYNAMICS

[in the last term we may neglect the difference between f ' ( k ) and /(£)],
we obtain
<5f(k) = - (Sa[kV]f(k)+ [daf(k)]. (3.7)

For a spherically symmetric field which can be represented in the


form f ( k ) = g(k)k, evidently d f = 0.
The second term in (3.7) can be transformed by utilizing the opera­
tor s introduced in (3.3):
[da/] = i(das)f.

On the other hand, the infinitesimal rotation operator J is defined by


the relation
<3/= (<5aJ)/= <5a(vJ)/
On comparing this expression with (3.7) we see that
J = —[ArV] —is - —/M,
i.e., that up to the factor i, M is identical with the infinitesimal rotation
operator.
By using expression (3.6) we can easily see that the components
of the angular momentum operator satisfy the following commutation
relations:
M j M j, —M VM X = z'Mx,
MaM 2~ M 2Ma = 0.
These relations hold for any quantum mechanical system and are a con­
sequence of the general relation between the angular momentum opera­
tor and the infinitesimal rotation operator.
From these relations it follows, as is well known, that the eigen­
values of the operator M 2 are equal to
M * = j ( J + 1),
where 2y'+l is a positive integer; later we shall see that for a photon
j is also an integer.
The eigenvalues of the operator are equal to
M2= M
The operator M commutes with the energy. Therefore photon states
with definite values of w, A/2, and M z are possible (quantum numbers
co,j, M). We now proceed to obtain the wave functions for these states.
QUANTUM MECHANICS OF PHOTON 21

3.3 Photon Spin Wave Functions


Formula (3.6) shows that the angular momentum operator for
the photon consists of two terms. The first term is identical with the
usual quantum mechanical operator L for the orbital angular momentum
in the momentum representation:
L =-/[*V J. (3.8)
The second term s may be called the operator for the spin angular mo­
mentum. However, the separation of the angular momentum of the
photon into an orbital and a spin part has restricted physical meaning.
First, the usual definition of spin as the angular momentum of a particle
at rest is inapplicable to the photon since the rest mass of the photon
is equal to zero. Second, as we shall see later, states with definite values
of orbital and spin angular momenta do not in the general case satisfy
the condition of transversality. Therefore, only certain superpositions
of such states have physical meaning. Nevertheless, from the formal
point of view the representation of angular momentum in the form
of a sum of two terms is very useful. It enables us to construct wave
functions for photon states having definite values of angular momentum
from simpler orbital angular momentum and spin eigenfunctions.
The vector subscript a of the photon wave function f a(k, t) can be
regarded as an independent variable which assumes three values: a
= x, y, z. In accordance with this we introduce the notation
f a(k, t) = f ( k , a, t).
The function f ( k , a , t ) is a scalar in the generalized momentum and
spin space defined by the variables kx, ky, kz, a. The different compo­
nents of the vector / are now values of the scalar f at different points
of the spin subspace. The operator L acts only on the variables k, while
the operator s in accordance with its definition (3.3) acts only on the
variable a. Therefore, the operators L and s commute.
The eigenfunctions of the operators L2 and Lz corresponding to
the eigenvalues L2 = /(/+ l), Lz = m will be denoted by Ylm. They
are functions of the variable n = k /k and satisfy the equations
V Y lm= l ( l + l ) Y lm,
L J i m = mYlm.
As is well known, Ylm represent spherical harmonics. We give the
22 QUANTUM ELECTRODYNAMICS

explicit expression for the spherical harmonics with the choice of


the phase factor which will be used in all subsequent discussions
(41), (119) (the factor —1 is associated only with positive odd
values of m)
( —iy (/-j-m )! ^ im<p 1 dl~m
sin2i#
¥) = -~Yl\ (/—tri) ! sinm# cf(cos#);~m
(3.9')
They are normalized by the condition

(3.9)

We now obtain the eigenfunctions of the spin operator. We denote


by x sp the eigenfunction of the operators s2 and s3, which corresponds
to the eigenvalues s2 = j(5+1) and sz = [x of these operators.
The argument of the function x sfl is the spin variable a. Therefore,
this function can also be represented in the form of the vector x str
If we write x in the form of the column vector,

H 9 ’
then by utilizing (3.3) we can easily obtain an explicit representation
for the operators sa and s2 in the form of the following matrices:
0 0 °\ / 0 0
0 0
o

sX 0
G/3
II

A
0 i 0 00 / ’
0/ \—
(3.10)
0 —i 12 0 0
°\
i 0
o

ss 2 0
II

0 .
0 0 0/ \o 0

It follows from (3.10) that the quantity s may assume only the single
value s = 1. In other words, the photon spin is equal to unity. Therefore,
in the future we shall sometimes omit the subscript s on the function %sll
and denote it simply by ^ .
The magnitude // of the spin component may assume the three values

(x — 0, r t 1.
QUANTUM MECHANICS OF PHOTON 23

The functions x? satisfy the following equations:


s2^ = 2 ^ ,
= .ax?-
By utilizing (3.10) we can easily obtain the solutions of these equations:

Xo = (3.11)
l ) ; "o'
The functions (3.11) are orthogonal to one another since they are
eigenfunctions of the Hermitian operator s^ corresponding to different
eigenvalues. Their normalization has been chosen so that
2 1Z*(«)
a
or, in vector form,
X„X* = <V- (3-12)
The unit vectors x M provide three basic unit vectors along which
any arbitrary vector / may be decomposed:
i
/ = /"X ,- 2 (3.1.3)
We shall refer to f M as the contravariant components of the vector
/ along these basic unit vectors. By utilizing the expressions (3.11) for
XMwe can easily establish the relationship between _/> and the cartesian
components of the vector / :
f°= A ,
/*■ = =f -p = c /; T '/ ,) - (3-I4)

Along with the contravariant components we can also introduce


the covariant components of the vector f M, which are defined by the
condition that the scalar product of two vectors / and g.

f g = L g z + H f x + i f y) f e x - t e „ ) + £ (/,-* /„ ) (g*+igv)>
should have the following form:
i i
f g = 2 g % = 2 f Mg»-
//=—1 ^=—1
24 Q U A N T U M E L E C T R O D Y N A M IC S

This will hold, as may be easily seen, if


/„= (- (3-15)
i.e.
/o = f z \ f ±i = "F ( y i ± */!,)•

/= 1 /,* ;•
n=- i
Since the spin operator commutes with the momentum operator,
it is possible to speak of states of definite momentum p and spin com­
ponent fi. The components of the polarization vector in (2.4) may be
chosen in such a way that
e= ^
The two possible polarizations correspond to only two values of the
component of the spin angular momentum /x. The third value is exclud­
ed by the condition of transversality. If the z-axis is directed along p,
then the transversality condition excludes the state Xo- The two vectors
Xi and x£> corresponding to circular polarization (2.6), are equivalent,
respectively, to Xi and X-i- Thus, the value of the spin component
H = 1 corresponds to right circular polarization, while p = —1 corre­
sponds to left circular polarization.

§ 4. Photon States of Definite Angular Momentum and Parity

4.1. Eigenfunctions o f the Photon Angular Momentum Operator


We now proceed to determine the form of the eigenfunctions of
the operators M 2 and M2. We denote them by Ym and call them vector
spherical harmonics or spherical vectors. As we have done previously,
we can introduce the spin variables a. Then

( Y 1M 0 ))a = Y jm 0 , a )•
These functions satisfy the equations
M zr m (n, a) = j 0 '+ 1 ) Ym (n, a).

Instead of solving these equations directly, we can utilize for the


determination of Ym the quantum mechanical rules for the addition
Q U A N T U M M E C H A N IC S O F P H O T O N 25

of angular momenta (9). Indeed, the problem reduces to the problem


(well known in quantum mechanics) of the construction of the wave
function of a system consisting of two noninteracting subsystems.
In our case such subsystems are the orbital and the spin degrees of
freedom of the photon associated respectively with the variables k
and a.
Since the spin angular momentum is given by s = 1, then according
to the rules of addition of angular momenta the total angular momentum
of the photon may assume a given value j if the orbital angular momen­
tum / is equal to
= j , j ± 1 for y # 0,
-1 for j — 0

Thus, in the general case there exist three different wave functions YjM
corresponding to three orbital states. We denote them by Ym .
The wave function Yjm is given, as is well known, by a linear com­
bination of products of orbital and spin functions,

Y jUi ( « , < * ) = H (4.3)


m+ti=M
for which the angular momentum components m and /j, are related
by the addition rule = M. The coefficients are in the
general case determined by the properties of the rotation group (cf.,
for example, (119) and (41); cf. also § 8).
For a given / and M the coefficients Cf%lfI form a matrix of three
columns (J = I, / ± 1) and three rows (^ = 0, ±1). The rows and columns
of this matrix satisfy the orthogonality and normalization condition

IciX C !X = *„■’ (4-4)

2) c>"lrc>»lr-= V- (4-5)
The elements of the matrix are found in Table 1.
Formula (4.3) can also be rewritten in vector form:
YjiM(n) = ^ m- p (n) X/i • (4-6)

A comparison of this formula with (3.13) shows that the contravariant


components of the vector spherical harmonic are equal to
(Yjmy = Ylt . (4.7)
26 Q U A N T U M E L E C T R O D Y N A M IC S

T able t

j I 1-1
/1 \

/ ( / —M )(/—M + l) / ( /+ M + 1 ) (/—M ) / (l+ M ) (/+ M + 1 )


-1
\ ( 2 / + 1 ) (2 /-h 2 ) \ 2 /(/+ l) V 2 /(2 /+ 1 )

/(/+ M + 1 )(/-M + I) M /(7 + m )77= m )


0
\ (2 /+ 1 M /+ 1 ) ]//(/+ !) \ 1(2l + l )

/ { I + M X I + M + 1) / ( / + M ) ( / - M + 1) / ( l - M ) ( l - M + 1)
1
\ (2 /+ 1 ) (2 /+ 2 ) \ 2 / ( / - f 1) \ 2 /(2 /+ 1 )

The covariant and the cartesian components are determined from


this formula by means of formulas (3.15) and (3.14).
The vector spherical harmonics Ym form an orthogonal system
of functions (a difference in the values of any one of the subscripts on
two functions indicates that the two functions belong to different eigen­
values of the self-conjugate operators M 2, Mz, or L2). Taking into account
the normalization condition on the coefficients (4.4), we can write

f = ’ (4-8)
where Y*M denotes a vector whose cartesian components are com­
plex-conjugates of the cartesian components of the vector Yj m . We
note that (Y*MY is not identical with (YjlMY*, but is equal to
(Y*lMY ( - i y ( Y jlM) - ^ = (Ynu)*.
It follows from this that
Y
-**jiM — Y n-M\( — lY+'+i+ju
— -*

4.2. Longitudinal and Transverse Vector Spherical Harmonics


We have obtained a system of eigenfunctions of the operators of
the square of the angular momentum of the photon M 2 and of its com­
ponent Mz. The state of the photon with definite values of j and M is
described by a wave function which in the general case is a linear com­
bination of three vector spherical harmonics:
Q U A N T U M M E C H A N IC S O F P H O T O N 27

The coefficients of this linear combination are not independent since


the photon wave function must satisfy the transversality condition
(1.10):
f}Mn = °-
Therefore there are not three, but two, different photon states with
given quantum numbers j, M. We denote the corresponding wave func­
tions by f jMX, where X may assume two values: X = 1, 0.
In order to obtain the explicit form of f jM, we note that from the
three vectors Yjm , which are linearly independent at each point of
&-space, we can construct three linear combinations Y fy (X assumes
the three values: X = 0, ffil), which are mutually perpendicular, with
one of these vectors, to which we assign the value X = — 1, being lon­
gitudinal, i.e., directed along the radius vector k, while the other two
vectors corresponding to the values X = 0, 1 are transverse. We now
obtain these combinations.
We utilize for this the well-known formula for the expansion in
spherical harmonics of the product n^Ym , where rf are the components
of the unit vector: T
n° = cos#, n*1 = =f=— sin#eTi'p
]/2
(# and cp are the polar and the azimuthal angles in the spherical polar
system of coordinates.) In our notation this expansion [cf., for example,
(16), (41)] has the following form:

nYjM = ~j//'2jX-l ^ M~ ~ \ / 2/+T 1+1‘ M'

Formula (4.9) can also be easily obtained directly, by expanding the


vector nYm in terms of vector spherical harmonics; the expansion
coefficients are equal to / nYm Y*Mdo. Consequently, expression (4.9)
represents that linear combination of vector spherical harmonics which
forms a longitudinal vector. Therefore, we define T $ for X = —1 by
Y W = nYm . (4.10)
We now consider the scalar product of the unit vector n with the
vector spherical harmonic Yjm
nYjm = ^ tin(iYj, M-(i"
28 Q U A N T U M E L E C T R O D Y N A M IC S

It vanishes, as may be seen by utilizing expression (4.9) and formulas


(4.5). Thus, Yjm is one of the desired transverse vector spherical har­
monics. We assign to it the value A = 0:
r 'S = r „ . (4.ii)
Finally, we define the second transverse vector spherical harmonic
Y $ by the relation
it / » = [»y/S5>]. (4.12)

By utilizing the expansion of nYm , it is possible to express Y $ in


terms of the vector spherical harmonics YjlM:

r% = j/w V w y— -(4'13)
On solving equations (4.9) and (4.13) with respect to Yjtj_ 1M and
YUj+1M we obtain

-— =■ ( j /y + l Y $ l + \ / j Y f c » ) ,
(4.14)

= -J ^ J W (l/J Y&' ~ ^ Y‘“ 1>)'

It is also possible to express the vector spherical harmonics Y}$


and Y $ in terms of the derivatives of the spherical harmonic YjM. In
order to do this we have to use the well-known formula for the differ­
entiation of spherical harmonics, which may be written in the following
form:

= , _ _ - [j ]/j+ 1 ^;j+ i.ii/+ 0’+ l ) ]/y Yi.j~i.Ai)-

On comparing this formula with (4.13) we see that


k
yd) —
x jM — v*JV (4.15)
l/z O + O
and, in accordance with (4.12),
y(0) _ __ YiM ___________ 1__ ~
jM 9 (4.16)
m I ® ) 7 x /+ fT ’
where L is the operator (3.8).
Q U A N T U M M E C H A N IC S O F P H O T O N 29

We shall also write down the expressions for the components of


the vector spherical harmonics Y}$ along the unit vectors of spherical
polar coordinates. On utilizing formulas (4.10), (4.15), and (4.16) we
obtain
(y < -m _ y
1M )n 1 jM ’

0,
( 57t,)» = T O )» = o,
i SYm (4.17)
T O )* = +KY}2X= -
] / /0 '+ i ) d& ’
1 dY.1M
V7(y+1) sin?? dtp

It may be seen from the definitions (4.10), (4.11), and (4.12) that
the vector spherical harmonics Y}$ are normalized in the same way
as Y nM. The functions Y $ , like the functions Y jlM, are mutually orthog­
onal for different values of j and M. Since, moreover, for different
A they are mutually perpendicular at all points, we have

/ ri»*r<.”.<h = s,r dKU.du ..

4.3. Parity o f Photon States


Thus, having taken into account the requirement of transversality,
we can represent the photon wave function of a state with a definite
value of the angular momentum in the form
f m = °iYja+<Iars>,
where the coefficients a1 and a0 are now arbitrary. This means that
the states f jM are twofold degenerate. We can remove this degeneracy
by requiring that each state must also have definite parity, i.e., that
it should be an eigenstate of the inversion operator I. The latter is de­
fined for the vector field f{k) in the following manner:
Im = -/(-* )•
(The minus sign is introduced since under inversion any direction goes
over into the opposite one I n = —#i.) Since I 2 = 1 the operator I has
the two eigenvalues ± 1 .
30 Q U A N T U M E L E C T R O D Y N A M IC S

The operator I commutes with the angular momentum operator


M and, therefore, we can set ourselves the problem of finding those
states for which the angular momentum and parity have definite values
simultaneously. These states can be easily obtained if we use the well-
known property of spherical harmonics
= ( - l) 'I 'm W ,
from which it follows in virtue of (4.7) that

For transverse vector spherical harmonics the following relations hold:


lY® = ( - i y + '- * Y % (4.18)
This can be seen from their definitions (4.11) and (4.13). Thus, for given
j and M two states differing in parity are possible. We denote the wave
functions of these states by f jMX(k = 0,1).
The states corresponding to the value A = 1 are said to be states
of electric type, while the states with A = 0 are said to be states of mag­
netic type. These names are associated with the fact that, as we shall
see later (cf. § 27), the emission of a photon in the corresponding states
is determined by the electric or the magnetic moment of the system
of charges.
The vector spherical harmonics Y}$ determine only the angular
dependence of the photon wave function f jMX, which is also a function
of the time and of the absolute value of the propagation vector k
f , m = “( k - O i'S S W -
According to (1.9) a(k, t) = a(k) exp ( ~ ikt), where a(k) is an arbi­
trary function which is restricted only by the normalization condition,
oo
J \a(k)\zk * d k = l .
0
If we make use of the discrete spectrum of k, then in accordance with
( 1. 20),

]?\a(k)\*k* = ^ . (4.19)
k

If, simultaneously with the angular momentum and the parity,


the photon energy also has a definite value, then a(k) differs from
zero only for k = a>, i.e., a(k) = (//co ]/< 5 )^ (the choice of the phase
Q U A N T U M M E C H A N IC S O F P H O T O N 31

factor is, of course, arbitrary). Thus, the photon state may be uniquely
specified by giving four quantum numbers corresponding to the values
of the energy co, the angular momentum j, the component of the an­
gular momentum M and parity X. The normalized photon wave function
then has the form
= (4.20)
coyd
We note that for X = 1 it is not possible to ascribe a definite value
of the orbital angular momentum to the photon state, since according
to (4.13) the vector spherical harmonic Y fy is a* linear combination
of the vector spherical harmonics Ym with different values of I. Here
we have evidence of the fact that, strictly speaking, it is not possible
to separate the total angular momentum of the photon into orbital and
spin parts, as we have already stated in the preceding section.
If j = 0, then according to the rules of addition of angular momenta
(4.2) there exists only one vector spherical harmonic F0io which is
identical with the longitudinal vector spherical harmonic F0(0-1) = nY00.
From this it follows that there are no transverse vector spherical har­
monics for j = 0. This result has a simple meaning. The state with
angular momentum zero represents a spherically symmetric state,
but a spherically symmetric vector field can only be longitudinal.
Thus, a photon cannot exist in a state of angular momentum zero.

4.4. Expansion in Spherical Waves


The wave function of a photon in an arbitrary state may be expanded
in terms of the functions (4.20):

Since the system (4.20) is orthonormal the expansion coefficients will


be determined in the following manner:
(4.21)

Formula (4.21) can be used, in particular, for expanding states of definite


momentum p and polarization p in terms of states of definite angular
momentum. In such a case
(4.22)
32 Q U A N T U M E L E C T R O D Y N A M IC S

where <50 is the element of solid angle defined in subsection 1.4. With
the aid of (2.7) we can easily obtain the expansion of the wave function
(4.20) of a state of definite angular momentum in terms of states of
definite momenta. This yields
= («:$«)*■
The probability that a photon which has a definite angular momentum
and parity is moving in a given direction /» and has a given polarization
H is evidently equal to | a ^ 1MX' |2.
The probability of the direction of motion lying within a certain
solid angle has more physical meaning. In order to determine this prob­
ability we must multiply |a |2 by the number of different states lying
within do, which is equal to do/80. If we sum this probability over the
polarizations, then in accordance with (4.22) we obtain:
do
x 2 K T ” !2 = IY$ (*) 1! * - A *) * . (“.23)
0 V-
where x == cos §.
We note that since the two transverse vector spherical harmonics
are related by expression (4.12), expression (4.23) does not depend
on A; i.e., the angular distribution of the photon is determined only
by its angular momentum, and not by its parity.
The function F(x) can be represented in the form of an expansion
in terms of Legendre polynomials

A *) =
a * A .M -
n=0
The coefficients a.Zn are given in Table 2. Explicit expressions for the
functions F(x) are given in Table 3.
T able 2

j 1 2 3
\M\ 0 1 0 1 2 0 1 2 3
flan
<*0 1 1 1 1 1 1 1 1 1
-1 1/2 5/7 5/14 -5/7 1 3/4 0 -5/4
-12/7 8/7 -2/7 3/11 1/22 —7/22 3/22
«6 -25/11 75/44 -15/22 5/44
Q U A N T U M M E C H A N IC S O F P H O T O N 33

4.5. Expressions for the Electric and Magnetic Fields


We now obtain expressions for the electric and magnetic fields
corresponding to a photon state of definite energy, angular momentum,
and parity. As before, we use the notation

£ = (> ;+ (* * ; H =§+§*.

In accordance with (1.8), (1.19), (4.20), and (4.12) we have

/~i~j \ 3 /2 _

In order to evaluate the integrals occurring in these expressions we


use the well-known expansion of a plane wave in terms of spherical
harmonics:

=2 ft(*')y,i if] Y,m . (4.24)

where
g, (kr) = (2i i j ^ i‘ /; +1//2- r- (4.25)
ykr
Vi +I/2 are Bessel functions).
From the asymptotic expansion of the Bessel functions it follows
that for kr !> 1
71
sin I k r—l —
(4.26)
g f k r )& 4m'
kr
We note that the function gt satisfies the normalization condition

/ * , • ( * » * * = k\ (4.27)

Indeed, in the last integral the important contributions come from


large values of r. Therefore, we can utilize expression (4.26), and this
leads to (4.27).
34 Q U A N T U M E L E C T R O D Y N A M IC S

H K
* oinHo
os os
CN *
O H os +
a>
X
CN
Os Tf- * +
so in + ><
CN
<N +
H Cl OS +
a> +
H VO
r«x1
+ + Os w
*# K
r- Os H H
H © H m CN cn cn
so w H ©
so
CN
in
©
r-
in
CN r^ *
in J i
<N CN * H
O © CN © K
K in
*
OS CN
K
oo so
«n
H CN
CN
£
in
in
h 8 I S
hf
r*-
cn CN cn
CN ©
CN

CN
1 + © CN + cn
+

H
OS K -H
x cn + H
S O r- I H —< CN
OS N
Jn +
X
Os
•^T X
cn 7I **
Os to
+
The Functions F(x), (x = cos $)

in I X
H I 2 K
t J*
s
os so I H
rf tXJ- Os
H
so
X
00 * in in
*n r- in CN
+ •n cn
CN
Os
IK oo
2 +
so + cn
so
^ hin° +
Its
T able 3

K +
»n *
cn »n
©
+ I
in H
<N H OS cn
so
CN
4 + CN IK
in oo
+ cn cn
so in so
© in
CN

K
*
l cn

<n I00K
(
H in Iso
K

»n Iso
K

+
CJ
H
I rt
cn Iso
K

CN
QUANTUM MECHANICS OF PHOTON 35

By utilizing explicit expressions for the components of the vector


spherical harmonics we obtain

/ (I) e“r* = ft<k r > (7 ).

f Y’" (x)e“"do=ft<kr>r<"(7 )’
I YSl ( I ) (✓ J W * ) F ,)t, „ ( l j

+ 1 / j + 1 gi-i (kr) Yu y_!_ M

From this it follows that for states of electric type (A = 1),

® - ‘V ^ Y W ^ h em(kr)r
+ ] / ( 4 . 2 8 )

&= - ] /? [ £ )
while for states of magnetic type (A = 0),

* = - / * t e l ’V 2 / + r f t - (ir) (4-29)

We note that formulas (4.28) reduce to (4.29) if we make the substi­


tution @ -*■ —i<Q, which corresponds to the invarianceof
Maxwell’s equations under the transformation H E, E -*• —H .
Since the vector spherical harmonic Y $ = YjjM is transverse, the
magnetic field in states of electric type and the electric field in states
of magnetic type are both transverse, i.e., we have
(rtyx-x = (r®)x=o = 0-
36 QUANTUM ELECTRODYNAMICS

But the electric field in states of electric type and the magnetic field in
states of magnetic type are expressed in terms of the vector spherical
harmonics Yj j+l M, which are not transverse. By utilizing (4.14), we
can represent (y and Ap as follows:

e,,-., = ' ] / ! ( £ W JU + V V a 1'

+ Ue,H+U+ O ft-.) v s I e-'“ (4.30)


^P(a=o) = •
The first term in (4.30) gives a radial component of the electric
field. At large distances, for kr > 1 (the “wave zone”), the radial
component vanishes. Indeed, for A r > l , (kr) = gl+1 (kr), in accord­
ance with (4.26). Therefore the following asymptotic expressions hold
for the fields:

i-s-7 sin ( k r - ^ i j - k ) )
e W> = - * • $ ( ! - « = i>+1- y - n- 1 • ( 4 . 31 )

An arbitrary electromagnetic field may be represented in the form


of an expansion in terms of fields corresponding'to photon states of
definite angular momentum and parity:

wj MX

where awjMX are determined by formula (4.21). In particular, the coef­


ficients (4.22) define the expansion of a plane-polarized wave in terms
of spherical waves

or

« '* ' = 2 y m [ j ) \ *,(*<■) V m ( f ) • (4.32)

§ 5. Scattering of Photons by a System of Charges

5.1. Incoming and Outgoing Waves


So far we have been studying states of the free photon, i.e., such
states as occur only in the absence of electric charges. However,
QUANTUM MECHANICS OF PHOTON 37

this is already sufficient to draw some conclusions concerning the be­


havior of a photon interacting with charges.
We turn to formula (4.31), which gives the asymptotic form of
those solutions of Maxwell’s equations (1.1) which correspond to
a photon of definite energy, angular momentum, and parity. These
solutions may be interpreted as a sum of waves coming in to a central
point and going out from the central point. Indeed, they may be rep­
resented as follows:
(y = (yin + (vout )

& = fyln-hS?oat,

= AYjU(n)------------,

p-lk(r-t) (5-1)
G0Bt = 2>r«>00----------
r ,

£ out = [n ,
where n = r/r, while A and B denote normalizing factors, such that
B = ( - \ y + 1~xA. (5.2)
We note that each of the fields (vln, t*pln and (^out, ,f)out separately
satisfy Maxwell’s equations (1.1). A similar decomposition into incom­
ing and outgoing waves can be carried out likewise in the exact expres­
sions (4.28) and (4.29). It corresponds to representing the Bessel functions
occurring in the radial functions gt (kr) (4.25) in the form of a sum
of Hankel functions H (1\ k r ) and H (1){kr):

However, the terms obtained in this manner will not individually satisfy
Maxwell’s equations (1.1) at r = 0, since the radial functions will
in this case have singularities at the origin. The requirement that the
solution remain finite at r = 0 imposes on the amplitudes of the incoming
and the outgoing waves the relation (5.2).
We now consider the case when there exists a system of charges within
a certain bounded region r < r0 in the neighborhood of the origin.
For the sake of brevity we shall refer to it as the scatterer. Outside
the scatterer, equations (1.1) evidently hold and, therefore, the asymptot-
38 QUANTUM ELECTRODYNAMICS

ic solutions (5.1) are valid. However, since at r = 0 equations (1.1)


do not hold, the amplitudes of the incoming and the outgoing waves
will no longer be related by expression (5.2). The relation between
A and B will in this case be determined by the properties of the scatterer.
In the general case, the outgoing wave may correspond to a photon
whose energy and angular momentum differ from the energy and the
angular momentum characterizing the incoming wave. Therefore we
write the relation between the amplitudes in the following form:
*„ao= I
co'i'Af'A'
The matrix S occurring in this expression is called the scattering matrix.
We note that this relationship does not cover all the processes
which may occur because of the presence of the scatterer. For example,
as a result of the interaction two or more photons may be emitted
or a photon may be absorbed. Such processes cannot be investigated
within the framework of the quantum mechanics of a single photon.
Later (in Chapter IV) we shall study in greater detail a more general
scattering matrix which includes all possible processes.
Here we shall confine ourselves to the case when the scatterer,
first, has spherical symmetry, and, second, scatters photons elastically.
In this case conservation of angular momentum, of parity, and of
energy holds and, therefore, the scattering matrix must be diagonal
and must not depend on M. Its elements can be represented in the
following form:
{oyjMX\S\co'j'M'X') = djr8mm,8u .SJX(a>) (-1 )'+ * -* ;
i.e.,
(5.3)
where the factor (—1),+1-A has been introduced in order that in the
absence of the scatterer SjX(ay) should reduce to unity, and (5.3) should
reduce to (5.2).
The property of elastic scattering means also that the energy flux
of the outgoing wave is in absolute magnitude equal to the energy
flux of the incoming wave. On the basis of the definition of energy
flux density and of formula (5.1) it can be easily shown that the flux
densities in the incoming and the outgoing waves are respectively given by
sln = - 2 |z f |2|F |2/i; s°ut = 2 |5 |2|T |2n.
QUANTUM MECHANICS OF PHOTON 39

Therefore from the equality sln = .yout it follows that \A\2 = \B\2; i.e.,
I^aO)! = 1 or
SjX(co) = e2i6^ w\ (5.4)
where djX is real.
We can eliminate the restriction imposed on the properties of the
scatterer if we alter somewhat the formulation of the problem, and,
in particular, regard all processes other than elastic scattering as “absorp­
tion.” In this we shall include both absorption proper, and also scat­
tering processes involving a change in frequency or in angular mo­
mentum, or the conversion of a photon into two photons, and all other
incoherent processes. We shall take all these absorption processes into
account in a summary fashion as a reduction in the intensity of the
outgoing wave in comparison with that of the incoming wave. In this case
we can retain formula (5.3) if instead of (5.4) we take

(5.5)
where
0 < aj} < 1.

5.2. Effective Scattering Cross Section


We can now construct a formal theory for photon scattering analo­
gous to the theory of scattering of particles by a central field in quantum
mechanics.
We represent the photon field at large distances from the scatterer
in the form of a plane wave and an outgoing spherical wave

j/2 6 = eeikz+F{n)—^~ > (5.6)

where e is the polarization unit vector; the normalization corresponds


to unit energy flux density in the incident wave, s0= n0 (n0 is a unit
vector directed along the z-axis, while the factor exp(—ikt) has been
omitted for the sake of brevity). The energy flux through the area
r2do in the outgoing wave is equal to ds= |F |2 do. Therefore, in accord­
ance with its definition the differential cross section for elastic scat­
tering is given by
do = | F(n) |2do. (5.7)
40 QUANTUM ELECTRODYNAMICS

We now show that F can be expressed in terms of the elements of


the scattering matrix In order to do this we write the general asymp­
totic expression for a field of some given frequency k in the form
of a series of incoming and outgoing waves. In accordance with (5.1)
and (5.3) this has the form

e = 2 1 Yi‘S t K - 1y ^ e - K ’+ S , ^ * ' } .(5.8)


]MI
We now choose the coefficients ajMX in such a way as to pick out the
incident wave in (5.8). In order to do this we make use of the plane-
wave expansion (4.32). By utilizing (4.14) and the asymptotic repre­
sentation of the radial function (4.26) we obtain

= 7 7 2 (c r S ’> . ) ) YU! (") { (-1 ) ' « - V « r+ < * ') . (5.9)


mi
We see that if we set

then (5.8) assumes the form (5.6), with

A *) = (5.io)
mi
If we expand the polarization unit vector of the incident wave e in
terms of the unit vectors Xi and x 2 corresponding to the x- and j-axes:

e = CrXi + ^aXa.
and if we expand the vector amplitude of the scattered wave in terms
of the unit vectors x# and x P belonging to spherical polar coordinates
$ and <p:
F = FiXv+FzXv,
then by utilizing the explicit expressions for the vector spherical harmon­
ics (4.3), (4.11), (4.13), and (4.17), we can represent (5.10) in the
following form:

Fx
= R ex 9

Fz *a
QUANTUM M E C H A N IC S O F P H O T O N 41

where R is the following two-dimensional matrix:


R = ^CD+ jR(0 )j

RW = - i (A = 0 , 1),

fi?P;_!(C0 S$)]
dcosft I
dPj_1(cos§)
d cos#
and Pj is a Legendre polynomial. The matrix i?(1) determines the scat­
tering of waves of electric type, while R w determines the scattering
of waves of magnetic type. If the polarization density matrix of the
incident wave is given, then the differential cross section can be
expressed in the form da/do = Sp R qR +. The density matrix for the
scattered wave q is related to q by the expression q' = R qR +/ Sp R qR +.
The total elastic scattering cross section is given by

(5.11)

We now determine the effective absorption cross section oa. It


is evidently equal to the difference between the energy fluxes of incom­
ing and outgoing waves:
«. = 2 0 -W )W ‘
iMX
where ajMX are the coefficients in formula (5.8); i.e.,

(5.12)

5.3. The Optical Theorem


There exists a remarkable relationship between the total effective
cross section, i.e., the cross section both for scattering and for absorption,
and the forward scattering amplitude for elastic scattering.
42 QUANTUM ELECTRODYNAMICS

We consider the total flux through a sphere of large radius r of


the energy associated with the field (5.6). In the absence of absorption,
the total flux is equal to zero. The existence of absorption means that
there is a net flux equal to oa entering the sphere (if, as is the case in
(5.6), the field is normalized to unit flux density in the plane wave).
Thus, we have
(5.12')

The flux density vector s occurring in this expression which corresponds


to the field (5.6) is given by

s= + -^-~ {{eF*)e~ikiT~z)-{-{eF)eik(-T 2)}

+ — ^ {(,eF)eik(-T~z)-\-{eF*)e~i k .

It is obvious that the first term in this expression will make no contribu­
tion to the integral (5.12'). The second term will give the total elastic
scattering cross section. The integrals of the remaining terms can be
considerably simplified if we first integrate them by parts. We consider,
for example, the integral

o o

On further integration by parts the last integral over # will give


a term of the order of 1/r, and therefore we can neglect it in comparison
with the first term, which does not contain r in the denominator. Further,
the oscillating term containing exp(2ikr) can also be neglected. In
order to justify this it is sufficient to average over a small frequency
interval Akr-~> I jr:
QUANTUM MECHANICS OF PHOTON 43

On dealing with the other integrals in a similar manner we obtain

- f (n0n)(eF(ri))e^-^r*do = - - ^ ^ ( 0 ) ,

r J (non) (eF*(n))e~ik^ do = r j (eF*(n))e-ik^ do =~ eF*( 0).

On substituting all these expressions into (5.12) we obtain

2 tz
at = as+ aa = -rj^e(F(0)—F*(0))
or

at = — Im (^(0)). (5.13)

This relationship is referred to as the optical theorem.


We see that the total cross section is proportional to the imagi­
nary part of the forward scattering amplitude for elastic scattering
(or, more accurately, to the amplitude of elastic scattering without
a change in polarization).
The essence of the optical theorem may be interpreted in simple
terms as follows. The total cross section is a measure of the weaken­
ing of the incident wave. On the other hand, this weakening may be
regarded as the result of the interference of the primary wave with
the wave scattered forward.
We note that the optical theorem may be applied not only to photons,
but to any particles.

5.4. Dispersion Relations


The scattering amplitude F has definite analytic properties which
enable us to set up integral relations between its real and imaginary
parts analogous to the relations obtained by Kramers and Kronig for
the dielectric constant.
For the sake of simplicity we restrict ourselves to the case of small-
angle scattering. We can then neglect the change in the polarization
and take F to be a scalar quantity.
We investigate the amplitude F a s a function of the frequency co for
a given change in the momentum of the photon as a result of scattering
44 QUANTUM ELECTRODYNAMICS

q — k ' —k (q = k-&, where $ is the scattering angle), F= F(co,q).


By utilizing the Cauchy formula we can write F(co, q) in the following
form:

f
1 F(o)', q)
F(oi,q) dco', (5.14)
2ni co CO

where the integration is carried out in the plane of the complex variable
cjo' along a contour surrounding the point co' = co, and F(co', q) is a func­
tion of the complex variable co' for a given parameter q. If the function
F(co',q) has no singularities in the upper half-plane, then the contour
of integration in the integral (5.14) may be chosen in the form shown
in Fig. 1.

Since the integral along the semicircle going around the point co' — co
is equal to half of the total integral (5.14), we have
CO

F(co,q)
711 J CO —CO Til J CO —CO

where the principal value is taken of the first integral along the real
axis, while the second integral is taken along a semicircle of infinitely
large radius R -*■ oo. If this latter integral tends to zero as R co ,
we have
OO

F{<o,q) = X i I^ - ^ d a > ' . (5.15)


711 J CO —CO
—co

We note that as a result of the fact that the field is real F(—co)
- F*(co) .
QUANTUM M E C H A N IC S O F P H O T O N 45

We shall not investigate in detail here the analytic properties


of the function F(a), q), as can in principle be done on the basis
of an investigation of the properties of the series in terms of which
the scattering matrix S is expressed (cf. Chapters IV, VII). We shall
merely introduce certain simple considerations based on very general
assumptions with respect to the properties of the propagation of
electromagnetic waves (94). In the presence of a scatterer the field E must
satisfy a certain homogeneous equation. It may be expressed in the
form of a certain integral relationship (the Huygens-Kirchhoff principle)

E(r,t) = J K ( r , t; r't')E(r',t')dsdt' , (5.16)

where K is a function simply related to the Green’s function for the


equation of the field, and the integration is carried out over t' within
the light cone (t—t')2 > (r—r')2, and over an arbitrary closed surface
surrounding the point r. This surface may be chosen in the form of
an infinite plane perpendicular to the incident wave and situated in
the region where the incident wave is not distorted by the presence
of the scatterer. Then E{r' , t') ~ exp(/Ar'—icot') . We choose the point
r in the asymptotic region; i.e., we take£(r, t) = (F/r) exp [i(kr—tut)]-
In this expression we have

— F = J K ( r , t ; r', t)eia,(t- t')~Hkr- hr')dsdt'.

We introduce the following notation: t —t' = r; r = p + r ' ; then


r = Q+ (r'k'Ik) (r > r'), and

± F = f K(p, x ; r, t)e{^ T~e)- {<ir'dt'ds. (5.17)


r>Q
If q is held constant, then it follows from (5.17) that in the upper half
of the a> plane F has no singularities regarded as a function of co,
since the coefficient of u> in the exponential is positive. From this it
also follows that F falls off exponentially along a semicircle of large
radius in the upper half-plane.
Thus, the amplitude F(a>,q) regarded as a function of co for fixed
values of q possesses properties which guarantee that the dispersion
relation (5.15) holds.
46 QUANTUM ELECTRODYNAMICS

Formula (5.15) implies the following relations between the real and
the imaginary parts of the scattering amplitude:
00

—00 (5.18)

We now evaluate the difference F{o, q)—F(0, q). From (5.15) we obtain
00

— CO

For F(0, q) = 0 we obtain the following relations:


CO

OO
(5.19)

—oo

§ 6. The Photon Field Potentials

6.1. Transverse, Longitudinal, and Scalar Potentials


In the investigation of the interaction between photons and electric
charges we shall later need expressions for the potentials of the electro­
magnetic field corresponding to a photon in some definite state. There­
fore, by utilizing the expressions for the electromagnetic field of the
photon obtained in the preceding sections we shall now determine the
vector potential A and the scalar potential A 0 corresponding to the
photon. As is well known, A and A0 are related to the fields E and
H by the following expressions:

( 6. 1)
H = curl A .
QUANTUM MECHANICS OF PHOTON 47

If we expand A and A 0 into Fourier integrals

A = f A heikrdk, A 0 = J A0keikrdk,

then equations (6.1) will assume the form

Ek = —Ak—ikA 0fe,
Hk= i[kAh].
Since Ek and Hk are transverse vectors, it is convenient to decompose
Ak likewise into transverse and longitudinal parts:
Ak = A ^ + n A ^ , (6.3)
where

n = ~ 5 kAip = 0 .
k k
From (6.2) and (6.3) and the condition of transversality of the field
Ekk = 0 we obtain the following relation between A 0k and A[ll):

ikA0k+ A l "> = 0. (6.4)


By eliminating the magnetic field with the aid of expression (1.5)
we find the following simple relations between AlL), A ^ and Ek, Ek:
1 .
A lL)- ^ E k,
(6.5)
= ~ E k.
By utilizing these relations and also (1.6) we can express A {kl) and AJ!-0
in terms of the photon wave function in momentum space f k:

Ail)
—I
=(fk-f~k)>
)/ 2(2ji):
(6.6)

( f k+f*k).
4 “= - ] / ^
where <x>— | k | .
The quantities A 0h and A[^ are related only by expression (6.4),
while otherwise they are arbitrary. This gives expression to the property of
gauge invariance according to which the fields E and H remain unal-
48 Q U A N T U M E L E C T R O D Y N A M IC S

tered if an arbitrary function dtj/dt is added to the scalar potential,


while simultaneously —V^ is added to the vector potential. If, in addi­
tion, we impose on the potentials the Lorentz condition

di^ + ^ r = 0’

then the following relation will hold between the quantities A0k and
^ ll):
ikA{kM+Aok= 0. (6.7)

Moreover, the fact that A and A 0 are real imposes the following con­
ditions:
a
SiQk— a *
— ^O.-ky• ,4(11) — _ J ( I D *
A k — n -k •

In analogy to (6.6) we can introduce the complex function f Qk de­


fined in the following manner:

| / 2(27r)3a)

A ok — (/ofc+/o*-fc) •

Then in accordance with (6.4) and (6.7) we obtain

AW= (fok~^~
y/'2(p,) ’CO

A0,\)
-V 2(2tc)3
(fok-fo*-k)'

On substituting (6.2) and (6.3) into the equation Ek = i[kHk] and


on expressing A0k and in terms of f 0k, we find that f 0k, like f k,
satisfies the first-order equation
QUANTUM MECHANICS OF PHOTON 49

With the aid of the functions f k and f 0k the potentials of the field
may be written in the following form:

"4o = i'/o+.cY*,
A = S (+ 9 (* ,
i r l
s-i0 = - - -----=- — = f 0kelkrd k, (6.8)
] / 2 (2 n f j j/co ^ J

51 — ~ —----- — f —^=rhkeikrdk,
]/2 ( 2 7if J j/oj
where
hk = (6-9)
6.2. Longitudinally Polarized “Photon”
The set of quantities hk, f 0k, may be utilized in place of the quanti-
tie s/t as the photon wave function. The expectation values of the energy,
the momentum, and the angular momentum of the photon are related
to hk and f 0k by the following expressions:

w = f o ( h t h k- f ; j , k)dk,

p =

M = / ( / . * , f^lUUIk,
where M is the total angular momentum operator (3.6), while L is the
orbital angular momentum operator (3.8). These quantities do not
depend on the choice of f 0k. Therefore, the wave function (ft,., f 0k) does
not satisfy the usual normalization condition.
We shall obtain the potentials of the electromagnetic field corre­
sponding to a photon state of definite momentum or angular momentum,
if we demand that the wave function (hk, f 0k) must be an eigenfunction
of the corresponding operators. Since we know the wave functions
f k it remains for us only to determine the functions f 0k.
We can write the eigenfunctions f 0(k, t) of the momentum operator
in a form similar to (2.4):
50 QUANTUM ELECTRODYNAMICS

where C is an arbitrary constant. For states of definite momentum the


function hk in accordance with (6.9), (6.10), and (2.4) has the form

h(k, t) = ee —i k t .Jkp 5
( 6. 11)
V*
where e may contain an arbitrary longitudinal vector
e = e±-\- Cn
(eL is the transverse unit vector).
It is sometimes useful to give an interpretation of the existence of
a longitudinal component of the photon wave function by introducing
an additional fictitious photon state with longitudinal polarization.
Then, for a given momentum one can speak of three polarizations
defined by the vectors 2, 3). Two of these are transverse,
and are determined by formulas (2.5) or (2.6), while for the third one
we have = Cn. Since in this case the photon wave function also
has a scalar component equal to the longitudinal component, the energy
and the momentum of the photon in this fictitious state are both equal
to zero.
The eigenfunctions f 0k of the angular momentum operator, which
in the case of a scalar is identical with the orbital angular momentum
operator, are given by spherical harmonics, which we shall write in
a form analogous to (4.20):

fomSk, t) = Ym {ri) e ^ 6 k(a. (6.12)

We note that the parity of the function (6.12) is uniquely determined


by its angular momentum and is equal to I = (—1)L It is identical with
the parity of the states of electric type. The vector hk for states of electric
type is in accordance with (6.9), (6.12), and (4.20) equal to

K k , t) = (6-13)

For states of magnetic type 1 = (—iy +1, and therefore the wave
function has no scalar part:
fok = 0, hk = f k . (6.14)
QUANTUM MECHANICS OF PHOTON 51

On the basis of (6.13) we can say that a state of electric type appears
to be a superposition of a transverse photon state (A = + 1) and a
“longitudinal photon” state (A = —1).
We also note the following important fact. The free electromagnetic
field may, in general, be described without making use of potentials.
In such a case no fictitious states corresponding to photons with lon­
gitudinal polarization arise, since the fields E and H have the property
of transversality, and this property of the fields is relativistically invar­
iant. However, for the investigation of the interaction between photons
and electrons the introduction of potentials is necessary, since the poten­
tials of the electromagnetic field occur in that part of the Lagrangian
or the Hamiltonian which determines the interaction. At the same
time the necessity arises of investigating “longitudinal” (and also “ sca­
lar”) photons, since it is not possible to make the potentials transverse
in a relativistically invariant manner. These questions will be discussed
in detail in Chapters III and IV.

6.3. Potentials for Plane and Spherical Waves


With the aid of the wave functions (hk, f 0k) it is possible by utilizing
formulas (6.8) to obtain expressions for the potentials of plane waves
describing photon states of definite momentum and polarization, and
of spherical waves describing photon states of definite angular momen­
tum and parity. We write the potentials in the form
A = 21+31*, A0 = c70+ + 0*.
It can be easily seen that the quantities 21 and sA0 have the following forms:
In the case of plane waves:

21 = ——= (e + Cn) ei(Pr- mt\


} / '2 Q
a>
(6.15)

a) — \p\;
In the case of spherical waves of magnetic type:

(6.16)
.<+o = 0 ;
52 QUANTUM ELECTRODYNAMICS

In the case of spherical waves of electric type:

- C[ ] / ^ i gi+i(pr)Yu + l ' „ Q (6.17)

By making different choices for the arbitrary constant C which


determines the amount of admixture of the “longitudinal” state, we
can make expression (6.17) assume different forms. Thus, for C = 0
(86) differs from G only by the factor l//co. For C = (_//(j+l))1/2
the expression for $( reduces to a single term and contains only the func­
tion gj-i, which is advantageous for a number of applications (cf.
§ 27 and § 40):

(6.18)

§ 7. System of Photons

7.1. Wave Function for a System of Two Photons


So far we have been considering a single photon and the electro­
magnetic field corresponding to it. However, we can generalize the
method just presented for introducing the photon wave function in
momentum space, and we can investigate also an arbitrary number
of photons in a manner analogous to the way a system of particles
is treated in quantum mechanics.
We investigate in greater detail a system of two photons. The argu­
ments of the wave function of such a system are the momenta k l} k 2,
QUANTUM MECHANICS OF PHOTON 53

and the spin variables al5 a2 for the two particles, and also the time t.
We denote this wave function by

i, oq; k 2, a2; /).


The square of the absolute value of the wave function determines the
probability that the first photon has the momentum k x and is polarized
along the ax axis, while the second photon has the momentum k 2 and
is polarized along the a2 axis.
Instead of representing this wave function as a scalar in the space
defined by k x , k 2;ax , a2 we can regard it as a tensor function in the
space of the momenta kx, k 2:
f ( k t , a 1; k 2, a 2; t ) = f aiai (kx, k 2;t) (7.1)

in a manner analogous to the way in which we represented the wave


function for a single photon in §3 either as a scalar f ( k , a, t) or as a
vector f a(k, t).
In first approximation the photons may be regarded as non-interact­
ing particles. (In fact, owing to the interaction between photons and
electrons, a weak interaction between photons appears, cf. § 54.) In
such a case the energy of the two-photon system is equal to the sum
of the energies of the photons, and their wave function satisfies the
equation
w/ ,

where w = w ^ J + w ^ ) , while w(Aq) is the Hamiltonian for a photon


defined by formula (1.1 T).
In addition to this equation the wave function must also satisfy
two other conditions. The first of these (the initial condition imposed
on the equation for / ) requires that the polarization of both photons
should be transverse. It may be formulated in the form of the following
expressions:
(k1)atf aiat= 0 ,
W X ^ 0-
The second condition represents a symmetry condition and follows
from the identity of the photons. Photons obey Bose-Einstein statistics,
and their wave function must be symmetric with respect to an inter-
54 QUANTUM ELECTRODYNAMICS

change of the particles. Therefore (omitting for the sake of brevity


the argument t), we have
f i K , a x, k 2, a2) = f { k 2, a 2, k u , (7.3)
or, in tensor form,
f aiaXk l ’ k 2)=fa*aXk 2’ k l)-
Instead of the variables kx, k 2 we can introduce the total momentum
of the system (momentum of the center of inertia) K :
K = k 1-\-k2
and the relative momentum 2k:

In terms of these variables the wave function of the two-photon system


may be written in the following form:
/(* i, <h, kz> a2) = <p(K)f(k ; a l5 a 2) .
In virtue of the conditions of transversality (7.2) and of symmetry (7.3)
the function f ( k ; ax, a2) = f aia,(k) satisfies the following relations:

faiak ai = fa^K = 0 > (7’4)


/ ( * ; <*1 , a2) = / ( - k ; a2, ax). (7.5)
If we consider states of definite total momentum K, then we can
always (with the exception of the case when k x and k 2 are parallel) trans­
form to a coordinate system in which
K — 0, k 1= k ; k 2 = —k .
The function f ( k , a lf a2) represents the two-photon wave function in
this coordinate system.
We note that two photons of total momentum equal to zero are
experimentally observed in the case of the decay of a neutral system
at rest (neutral 7r-meson, positronium).

7.2. Even and Odd States o f Two Photons


We now set ourselves the problem of finding the wave functions
of the system of two photons of definite angular momentum and parity
(113). We take the momentum of the center of inertia K equal to zero.
QUANTUM MECHANICS OF PHOTON 55

The angular momentum operator for the system M is the sum of the
operators for the orbital and the spin angular momenta:
M = L + s,
where L is the orbital angular momentum operator for relative motion
L = —/[&Vfc], and s is the sum of the spin operators for the two photons
S = Sj + Sj .
As in the case of a single photon, the eigenfunctions of the opera­
tors L2 and L, are given by the spherical harmonics Ylm(n).
The eigenvalues of the square of the spin angular momentum are
equal to s2 = 1), where in accordance with the rules for the addition
of angular momenta the quantum number s may assume the values
s = 0, 1,2. Since for a given value of s the eigenvalue of the operator
for the component of the spin assumes the values // = —s, —5 + 1 ,..., s,
there exists for a two-photon system a total of nine different spin wave
functions which we denote by xs/J(alt a2). They are bilinear com­
binations of the spin wave functions of the two photons x^iiai)
and ^ 2(a2).
These combinations may be written in the form of six symmetric
functions XpifaJXrti^+Xpifadx&fai) and three antisymmetric func­
tions ^ i ( a i ) ^ 2(a2) —^/il(a2) / /<2(ai)- Since states with different values of the
quantum number [m for a given quantum number s must have the same
symmetry, we must identify the three antisymmetric functions with
the functions xSfl corresponding to the value 1, while the six sym­
metric functions must be ascribed to the values s = 0, 2. Of these,
the four functions for which ^ + ^ 2 7^ 0 must correspond to the value
s = 2, while the remaining two functions
X o ( a i ) X o ( a z) and X i ( « i) % -i (“ 2 ) + X - i ( a i ) X i ( < h )

represent linear combinations of the wave functions Xzo and £00.


We note that the division of spin states into symmetric and anti­
symmetric ones has a real meaning. As regards the classification of
states according to the value of spin, it has no deep physical meaning,
since, as will be seen later, the condition of transversality requires the
superposition of functions with s = 0 and s = 2.
The wave function for a system of two photons with a definite value
of the angular momentum j and of its component M along the z-axis
al5 a2) represents, in accordance with the rules for the addition
56 QUANTUM ELECTRODYNAMICS

of angular momenta, a linear combination of products of orbital and


spin functions Yim(klk)xs/t(al , az).
States of definite angular momentum can be classified with respect
to their parity. The inversion operator I acts only on the orbital part
of the wave function:
I / 0l a, ( * ) = / 0l » , ( - * ) »
i.e.,
IA k , a 1, a J = f ( —k , a l , a J . (7.6)
Since

k
the parity of the state is determined by the quantity I = (— 1)'. On
the other hand, the replacement of & by —k is equivalent to an inter­
change of the momenta of the two photons. From the condition of
symmetry of the wave function (7.5) it follows that Ylm(n) %s/l (cq, a3) —
Ylm(—n) %S//(a2, cq). We thus see that there exists a connection between
the parity of the state and the symmetry of the spin function:
XSlXai ’ az) = ( - l ) ' ^ ( a 2 , cq) (7.7)
i.e., even states (/ = 2n, where n is an integer) are symmetric with
respect to the spin variables, while odd states {I = 2n-\-\) are anti­
symmetric. In other words, the tensor f aiai is symmetric in the case
of even states and antisymmetric in the case of odd states. In the former
case it has six components, while in the latter case it has three compo­
nents, which is in agreement with the number of symmetric and anti­
symmetric spin functions.

7.3. Classification of the States of Two Photons of Definite Angular


Momentum
The state of a single photon of given energy is completely determined
by its angular momentum and parity. To each set of the quantum
numbers co, j, M, I there corresponds one photon state (with the exception
that the state for j = 0 is missing). In the case of two photons the situation
is, generally speaking, different. Let us count the number of states
corresponding to a two-photon system when the angular momentum
and the parity are specified. We shall denote by JVj+> the number of
even states of given j and M, and by 7V<-> the number of odd states.
QUANTUM MECHANICS OF PHOTON 57

We begin by considering the odd states. As is well known, an anti­


symmetric tensor of the second rank may be written in the following
form:
(7.8)
where ea^ y is the unit antisymmetric tensor of the third rank, while
Fy is some vector. Therefore, the odd states may be described by a vector
wave function F(k). The properties of the vector F differ significantly
from the properties of the single photon vector wave function./^. Indeed,
the condition of transversality (7.4) applied to the antisymmetric tensor
(7.8) yields
eapvnpFy = 0, t.e. , [l»F] = 0.
Thus, the vector F is longitudinal in contrast to the transverse vector f k.
According to the results of § 4 there exist three linearly independent
vector functions Y fy corresponding to given values of the quantum
numbers j and M, with two of these being transverse (A = 0, 1), and
one of them being longitudinal (A = —1). From this it follows that
for a given j and M the vector F has the following form:
F = IV >(/i).
In accordance with (4.10) this function is odd only when j is even.
Thus we have
for j = 2n,
A<-> = 1 I for j — 2 n + l .
(7.9)

We note that since the tensor wave function for a system of two
photons f aiai is bilinear in the components of the photon vector functions
f x and / 2, then it follows from the antisymmetry property of the tensor
f that the vectors f x and / 2 have no components along the same axis.
This means that the polarizations of two photons in an odd state are
mutually perpendicular. The neutral jr-meson and positronium in the
ground state (cf. 33) can serve as examples of odd systems decaying
into two photons.
We now proceed to investigate the even states. As we have seen
earlier, they correspond to the values of s = 0, 2. In accordance with
the rules of addition of angular momenta for given values of j and M
there are six different wave functions corresponding to different values
of /:
I — ydr 1j , / i 2, _/ for s — 2
58 QUANTUM ELECTRODYNAMICS

and
/= j for s = 0.
Since / = (—1)', then of all these functions we must retain only those
for which I — In. The number of these functions is equal to four when
j = 2n and is equal to two when j = 2 n + l .
In order to obtain N f f l it is also necessary to take into account
the transversality condition (7.4). In order to do this we consider any
one of the tensor wave functions f aiai corresponding to given values
of j and M. We construct the vector Ga = which represents the
longitudinal component of the wave function. Since the vector n is
spherically symmetric, the vector G, like f aia., is an eigenfunction of
the angular momentum operator with the same eigenvalues j and M.
But for given values of j and M there exist three linearly independent
vector spherical harmonics Y $ . Therefore, of the six tensor functions
/ aia>, only three can have longitudinal components different from
zero, while the remaining three satisfy the condition of transversality.
The parity of the former is identical with the parity of the vector spherical
harmonics corresponding to them. Therefore, the number of even,
linearly independent tensor functions which have a longitudinal compo­
nent different from zero is equal to two when j = 2n and is equal to
one when j = 2n- \ . On subtracting this number from the total number
of even functions we obtain the number of even states which satisfy
the condition of transversality:
N )M) = 2 > j — 2n, (7.10)
N jP = U ./= 2 h + 1 .
The cases of j = 0 and j = 1 require separate consideration. When
j = 0 the rules for the addition of angular momenta allow only two
states (both even),
1=0, s — 0 ; 1=2, s= 2 .
On the other hand, when j = 0 there exists only one vector spherical
harmonic (even). This means that of the two tensors f aiQt only
one satisfies the transversality condition. Thus, there exists only one
even state of the system with angular momentum equal to zero. In this
case, as may be easily checked, the wave function of the system satisfying
the conditions of symmetry and transversality is given by the function
= const («ainaa — <5aiai) • (7.11)
QUANTUM MECHANICS OF PHOTON 59

It may be concluded from (7.11) that the polarizations of the photons


in this state are parallel. Indeed, let the z-axis be directed along k, then
we have f xy = 0.
When 7 = 1 the rules for the addition of angular momenta allow
only one even state: / = 2; s = 2. But the corresponding tensor function
has a longitudinal component, since of the three vector spherical harmon­
ics only Yfjtf is even. This means that there are no even states
satisfying the transversality condition when j = 1. In accordance with
(7.9) when j = 1 there are also no odd states. Thus, a system of two
photons, generally speaking, cannot be in a state of angular momentum
equal to unity, while a system having an angular momentum equal
to unity cannot decay into two photons.
T able 4

Angular momentum, No. of even states. No. of odd states,


j M<+)•
WjM xr(—
™ jM)

0 1 i
1
2n 2 i
2n + l 1 —

In the accompanying table a summary is given of the classification


of the states of a two-photon system.

7.4. Wave Function for a System of an Arbitrary Number of Photons


A system of an arbitrary number N of photons can be described by
the wave function f ( k x, cq ; k 2, a2; k 3, a3; ...; kN, aN) which depends
on the photon momenta k t and the variables at which determine the
photon polarizations (each variable at assumes three values: at = 1,2,3).
The wave function must be symmetric with respect to interchanges
of particles; i.e.,
f (... ; k t, a( ; ... ; k j , cq , ...) = f { . .. , kj, a.^, , k t, o.l , ...),
and, moreover, it must satisfy the transversality conditions
k t, a l...)(ki)ai = 0 ( / = 1 ,2 ,
ai
where the summation is taken over the three values of the variable
60 QUANTUM ELECTRODYNAMICS

ctj (the number of transversality conditions is equal to the number


of photons).
If the state of an individual photon is characterized by the set of
quantum numbers r (for example, p , f i , or <x>,j,M,X), then the state
of the system will be completely determined by specifying the “ occu­
pation numbers” Nr, i.e., by indicating for each individual state r the
number of photons Nr occupying this state. The wave function of
the system with a specified distribution of occupation numbers can
be expressed in terms of the individual photon wave functions f T{k, a ) :

•/aTj-jJ JVr2‘ ” (^1 » > ••• > , Q -i, . . . )

where N = £ 7Vr , and the summation is taken over all the permu­
r.

tations of the subscripts r. The function f N is normalized by the condition

An arbitrary state of a system of N photons can be represented in


the form of a superposition of states with definite distributions of
occupation numbers:

(7.12)
When photons interact with charges the number of photons can
change, since photons can be emitted and absorbed. Therefore, in
the general case, we have to consider states with an indefinite number
of particles. Such states are described by a set of wave functions f N

«i/i(*i, a)
azfz(ki, cq, k2, a2)
(7.13)

aArfhr(ki, cq; ••• ; kN, aA)


QUANTUM MECHANICS OF PHOTON 61

where the square of the absolute value of aN determines the probability


that N photons are present ( £ \ a N\2 = 1).
Such a column is equivalent to a certain functional with the aid
of which we can describe a system containing an indefinite number
of photons. This functional is determined in the following manner
(64). Let QN(<p) = —— j f NQ x ... f *)?>(£x) ... <p{£N)d£x ... d£N, where

<p(£n) is an arbitrary function of the variables kn and an which are for


brevity denoted by (integration over f n also includes summation
oo
over an). Then, if we specify the functional D (<p) = aNQN{y), this
N = 0
is equivalent to specifying the column (7.13). Indeed,

6NQ(y)
i ••• h ) = V m
<p= 0

where d/d<p denotes a variational derivative.1


As an example ,we consider the case of N photons all having the
same momentum and polarization |. In this case aN = 1, a, = 0 (z't^ N),
and therefore

f i ( , 0 = 7 k ’’" ® ’
As is well known in quantum mechanics, in order to describe a sys­
tem of identical particles obeying Bose-Einstein statistics we can go
over to a representation in which the role of the wave function is played
by the quantities <J> in (7.12), while the independent variables are the
occupation numbers Nr. This method of description by means of a wave
function in occupation number space is called the method of second
quantization. The transition to this representation is carried out by

1 The variational derivative may either be obtained from the expression for
the variation of the functional

10 = / • w
where 8<p is the variation of the function <p, or it may be defined as

= lim
tj—
yo ‘n
62 QUANTUM ELECTRODYNAMICS

means of introducing operators for the creation and annihilation of


a particle in the state r (cf., for example, (119); (27)). These operators,
which we denote by c+ and cr, are defined by the following expressions:
C . . . ) = ) / W ...).

In Chapter III we shall introduce these operators and study their


properties in detail from another point of view.

§ 8. L- Vectors and Spherical Harmonics

8.1. Irreducible Tensors


In this section we shall investigate the basic properties of quantities
which transform under spatial rotations in accordance with the irreduc­
ible representations of the rotation group. Such quantities are, for
example, the wave functions of a particle with a given value of angular
momentum, and with different values of its components along the
z-axis. In future we shall encounter other quantities of a similar
type. For brevity we shall call them L-vectors. An L-vector has 2L + 1
components. (The transformation matrix for an L-vector corresponding
to a rotation of the coordinate system will therefore have 2L+1 rows
and columns.) Since (2 L + l)is an integer, L may be either integral
or half-integral.
We note that if L is an integer, then the L-vector is equivalent to
an ordinary three-dimensional tensor. In the general case, a tensor
of the Lth rank has 3L-components, whose transformations are re­
ducible. However, by means of symmetrization, antisymmetrization
and simplification it is always possible to select sets of a smaller number of
components which transform under rotations independently of the
others. Thus, for example, from the nine components of the second
rank tensor Tik it is possible to pick out its trace (scalar), and its anti­
symmetric part, consisting of three quantities which transform like
the components of a vector (a tensor of the first rank), after which
there remain five independent components T°k of a traceless symmetric
tensor:
n ^ i ( . T , k+ T k, ) - i T „ 6 , t .
Linear combinations of the components T°ik form an L-vector (L = 2).
QUANTUM MECHANICS OF PHOTON 63

It may be easily shown that the number of independent components


of an L-th rank tensor in n-dimensional space T?kl ... (i, k, /, ...
= 1,2, ..., n), which is symmetric with respect to any pair of subscripts
(Tiki ••• = Tku •••, T?m ••• = Tm •••), and which vanishes on contraction
with respect to any two subscripts (7?,, ... = T?kk ... = 0), is given by
( n + L - 1)! ( n + L - 3)!
(n—1)!L! (n—1)!(L—2)1
In three-dimensional space N = 2L+1; when n = 2 , N = 2 ; when
n = 4 , N = ( L + 1)2.
We denote the components of the L-vector F by FM(M = —L,
—L + l , ..., L ). The set of basis vectors, i.e., the set of unit vectors
in the (2L+l)-dimensional L-vector space, can be so selected that the
invariant bilinear form in the components of the L-vectors F and G
(with the same values of L), i.e., their scalar product, will have the
following form:
L
(F ,G )= £ ( - 1 Y F mG~m. (8.1)
M=-L

For L = \ the quantities FM are identical with the components


of a spinor, while for L = 1 they are identical with the contravariant
components of a vector defined in §3 in terms of the basis vectors y_M;
i.e.,
F o = Fz> F ±i = T - ± - (.Fx=fiFu). (8.2)

In the general case, we shall refer to FM as the contravariant compo­


nents of the L-vector. If we also define the covariant components Fm
by means of the following formula:
FM = { - \ ) mF_.m, (8.2')
then the scalar product of two L-vectors may be written in the
following form:
(F, G) = £ F mGm .
M
8.2. The Algebra o f L-Vectors
The following problem is a basic one in the algebra of L-vectors:
to construct an L-vector H M from products of the components of the
/-vector F m and of the /'-vector Gm'. This problem is solved in the theory
64 QUANTUM ELECTRODYNAMICS

of the representations of the rotation group; here we quote only the


results.
By way of a preliminary remark we note that if £ = / = / ' = 1,
then the problem reduces to the determination of the vector product
of two three-dimensional vectors, which, as is well known, is solved
as follows:
Hi = eUk Fj Gk ,
where i,j, k denote the cartesian components of the vectors, while
eljh is the unit antisymmetric tensor of the third rank.
Similarly, in the general case, we have
HM = C " m.FmGm., (8.3)
where the coefficients Chvfm. form a “third-rank tensor” whose compo­
nents, just as in the case of the tensor em , do not depend on the choice
of the system of coordinates. Since the values of C%m. depend on L,
/, /', they are usually represented in greater detail as C\\m. = Cf™rm,.
In (8.3) summations are implied over m and m' respectively from —/
to I and from —/' to
Formula (8.3) expresses the quantum mechanical rule for the addition
of angular momenta / + / ' = L. The coefficients are called
the vector addition coefficients or the Clebsch-Gordan coefficients.
We shall not quote here the general expressions for C\™Vm.. For
the case /' = 1 they are given in § 4, while for the case /' = \ they will
be given in § 11. We note that differ from zero only when the
following rules for the addition of angular momenta are satisfied:
M = m+m', | / - / ' | < L < /+ /'.
Formula (8.1) is a special case of (8.3) for I = V and L = 0. The
formula for ordinary vector multiplication corresponds to the special
case of (8.3) for / = /' = L = 1. If H = [FG], then
H „ = - i ] / 2 C Z m, F mGm„
The coefficients Cf^rm. have the following orthogonality properties:
QUANTUM MECHANICS OF PHOTON 65

From (8.3) and (8.4) the following formula can be obtained which
expresses the product of the components of the /- and /'-vectors F
and G in terms of the components of the L-vector H (the Clebsch-
Gordan series)

(8.5)
L
If /' is the smallest of the three quantities L, I, /' then for given /', I,
and m the quantities form a square (2/'+l)-row ed matrix
C, whose rows and columns are orthogonal and normalized in accordance
with (8.4). (L can assume values from /—/' to /-\-V, while m' can assume
values from —/' to /'.)
We note that the number of conditions (8.4) is equal to 2(2/'+1),
while the number of elements of the matrix C is equal to (2 /'+ l)2.
Therefore, the number of independent elements Nr is equal to
Nr = (2 /'+ l)2—2 (2 /'+ l).
If l' = \ , then N r = 0 ; i.e., the coefficients are completely
determined by the properties of orthonormality.
We note the connection between the Pauli matrices a** (<+ are related
to cfx, av, az by expressions (8.2)) and the vector addition coefficients
(a, /? are spinor indices)

The coefficients also have a number of symmetry properties


with respect to interchanges of subscripts or superscripts. These symmetry
properties may be most easily formulated if we introduce the coefficients

( 1 2 3 ), which are defined in the following manner [(119) § 97;


\m1 m2 m3J
also E. P. Wigner quoted by (44)]:

C fm r m ' ~ (2 T + l)1/2(—1)'"''+M ( 8. 6)

It is obvious that m1+ m 2+ m 3 = 0; /x+ /2+ / 3 = 0. Let I m( m


trij) m
mkk fJ
\ i
Ik k h\ .
be the coefficient obtained from I I by a permutation of
6 6 QUANTUM ELECTRODYNAMICS

columns. Then we have


4 4 4
' + = , / '‘ (8.7')
\ W1 m 2 m 3
where e = 1 if the permutation is even, and e = (— if the
permutation is odd.
Further, we have
I 4 4 4 \ -- f, _jVi
„s , , , , , / 4
+fi-Ha I 4 4
(8.7")
\mx m2 m3/ \— I
The orthogonality conditions (8.4) may now be rewritten in the
following form:
'4 4 4\ 4 4
(2/3+ l ) I '1 TR$T7l$
7” l m 2 w 3/ m 2 "4,
(8.8)
'4 4
4 \ ( '■ 4 4' _ c c
2 (M .+ 1 )1 m i m 2 ™3/ \ m l m 2 ^7Thm\^mtmi
^3
(A summation is implied whenever a lower index is repeated.)
We also note that when /3 = 0 we have

C £ .w = ( „ _ m, “j = ( 2 / + l ) - 1'aM , . , ( - l ) ' - " . (8.9)

It is convenient to make use of the vector addition coefficients


in the symmetric form (8.5) for carrying out the summation over the
components of the ^-vectors (the m, indices). We state the basic summa­
tion formulas (161), (25):
Iji 4 4 4 j 2 4
l + f a + fs + Tni+tfla + tfls
(-iy
\/A ^ 2 ^ 3/ — fi2 mA
h1 i
4 4 jz \ \ j i J 2 7 s) / 7 i 72 73 1
x m1 —m2 ^ 3 / 14 4 4 J '/^i 7^2 t ^ /
( 8. 10)
/A A y .U A 4 a \ 2( _ „+I,M.+„,(
//3; \mx m2 - / / 3/ X
^ 17*i 72 73 ) / 4 72 4 \ / 7*i 4 /;
‘ 3

14 4 4J v»i ^ " v W -m.


w 2 —■>
7i 72 )

(
7 3
?, which do not depend on the m( indices, are
4 4 4J
QUANTUM MECHANICS OF PHOTON 67

called Racah coefficients. Sometimes the following notation is also used


for them:

t 1 1 }= ( ~ l )h+ll+h+h w ( j i U k k \ AA>-

Relations (8.10) may be illustrated by the following prescription


for the vector addition of angular momenta. Suppose that we have
to add the three angular momenta A + A + ^ i = h- This addition may
be carried out in two stages in two different ways (Fig. 2):
1) A A 72 = y*3> hi \
A~bA ~
2 ) 7*2+ h ~ h i h+h = h-\ ( ' }
Formulas (8.10) establish the relations between the coefficients in these
two methods of addition.

Evidently the Racah coefficients differ from zero only if the relations
(8 . 1 1 ) hold, i.e., if the lines corresponding to the values of 7), lt appearing
in that equation can form a triangle.
The Racah coefficients possess the following symmetry and orthog­
onality properties:
|y'l 72 A j _ (A jpjy\ __ \h^ Jy\
U h h\ = U hg “ U jp lvV
where (a, /?, y) is any permutation of the numbers (1,2,3);
(A 7 2 AW A A A1 ( 8. 12')
£ (2 7 3 + I ) (2 /3 + I ) = dUA
u \ h h k) U h h)
68 QUANTUM ELECTRODYNAMICS

If one of the six indices is taken equal to zero, then the Racah coeffi­
cient has the following simple form:

In later sections expressions are given for the Racah coefficients


for the cases when one of the subscripts is equal to \ or 1 .

8.3. Spherical Harmonics


The most common example of an /-vector is a spherical harmonic.
The covariant components of the unit direction vector nm {m = 0,
± 1 ) coincide up to a normalizing constant with the first order spherical
harmonics Ylm. By utilizing the multiplication rule (8.3) it is possible
to construct from products of nm successive I -vectors with / = 2, 3, ...,
etc. They are identical (up to a normalizing constant) with the spherical
harmonics Y*. Formula (8.3) expresses the recurrence properties of
spherical harmonics.
With the aid of formula (8.5) it is possible to obtain the expansion
of the product of spherical harmonics YIm(n) Yrm-(n) in terms of the
spherical harmonics YLM(n). This expansion must obviously have
the form

I
where the coefficients q1/ do not depend on m, m', M. In order to de­
termine these coefficients we multiply both sides of the preceding equa­
tion by On utilizing the first of formulas (8.4) we obtain

Further, by choosing n along the z-axis ($ = 0), we obtain in virtue

Since Y*m{n) = ( - l ) m7, _m(n), we have

(8.13)
Q U A N T U M M E C H A N IC S O F P H O T O N 69

We also give the expansion of the product of two vector spherical


harmonics,
y*M 0 0 YrvM.(/I) = ( - 1 ) "+ 1 [(21+1) (21'+ 1) ( / + 1) (2j'+ 1)]1/2
2

x (4 ^1 1/2 ^- +2L+
- ............
1 ) 1/2| ^ | CforoCfy'i-M YZ;Xn>),

(8.13')
= (2 ^ ) (-
1 2 1 ) - +1

)
X£ c fS?oc fw-MW(jj,jj] 2n, l)P2n(cos&).
71= 0

On the basis of the properties of spherical harmonics regarded as


/-vectors we can easily determine the character of the expansion of
particle wave functions in terms of them.
The wave function y of a particle of spin 0 is a scalar. Therefore
the expansion of the wave function of a state of definite angular mo­
mentum / in terms of spherical harmonics has the following form:

W = I a mY:m(ri) (8.14)
m
where the coefficients am form a contravariant /-vector which determines
the orientation of the angular momentum.
It may be seen from (8.14) that the spherical harmonics Ylm(n)
can be interpreted as “basis vectors” in /-vector space.
If a particle has spin s, then its wave function xp can be repre­
sented in the form

v> = v =2- s V xM(«)

where ^ ( a ) are the eigenfunctions of the spin operator (a are the


spin variables). Thus y> is characterized by the contravariant j-vector
ijf. On the other hand, for states of total angular momentum Ly)
must have the form
= 2 a MYLM(n,d)
M
where YLM(n, a) are the eigenfunctions of the total angular momen­
tum operator. On the basis of (8.3) they can be obtained in the form
YLM(n, a) = Cf“s/1Ylm(n)xM(a)
70 QUANTUM ELECTRODYNAMICS
<M Ol

+ <1

sx
+ +
k'»n ISCN
+x
7 +
Racah Coefficients

i x
— *-T
^-( + i X
++ <4 77
9
4+
C
N s
1 /—N + *< <N
+
+t
s
7 ^ —
a '/W '- 9
4
1|fS <
Nc? -•s <N
+ l'-> X
+ 1+
k +
II 1-J

+ ;+ + + 1-1
s >-C -J S I
-1

+
s
II II
QUANTUM MECHANICS OF PHOTON 71

^ 2)

+
C4 + x»
1+

(M

+' +
+CJ
— /-—\

<1 + + xC
I +
J 1 +J2 - L f ( L + / i + / , + 2 ) ( L + / i - / 2 + l ) ( / i + / , - L + l ) ( L

+ k(N
<N H. < b
x->|rl
fS ^
+ + + X X'J s
+ +
N
x> 5 +
+ " f |7 Xj xT
s—✓(N X?' + N
sLb /—s
+
+ , ^ + Xj a
+ ? 7 x>C4 I
s + I +
<N XT
I ts x^ <
X? +
'w '
+ fe
+ ^ xr + + Xj 1-^

+ +
Xj
? + +n ICJ
+ > x -,
x» ^ "t
XH- + + r5
l<N. x? |< s ^
IN k y

xC
+ S + +
+ Xj xT
I +
xj xq
I XII)
l_ I
+ + +
N
+

+ +
xC X, X? X,

I II II I I I
72 QUANTUM ELECTRODYNAMICS

Thus
v = H
M aMY&(n)
where
Y&{n) =
The functions Yfi(n) (basis vectors in L-vector space) can be called
generalized spherical harmonics . 1 In particular, for s = 1 they coin­
cide with spherical vectors (cf. § 4), and for s = 1/2 they coincide
with spherical spinors (cf. § 1 1 ).
Equation (8.15) defines the transformation to a dual tensor. Just
as in vector algebra an antisymmetric tensor of the second rank A u
can be introduced which is dual to the vector A k,
Ai) eiHc^k>
so in the algebra of L-vectors the tensor Gfa ((L-j)-vector) can be defined
which is dual to the /-vector Gm,
qh _ rM g

The tensor G£ can be represented by a rectangular matrix of 2L+1


rows and 2s-\-1 columns. It depends on all three indices L, I, and s.

1 These wave functions can also be obtained directly by solving the equations
for the eigenvalues of the infinitesimal rotation operator. In the general case these
wave functions can be expressed in terms of Jacobi polynomials; cf. (75).
C HA P T E R II

Relativistic Quantum Mechanics of the Electron

§ 9. The Dirac Equation

9.1. Spinors. Pauli Matrices


In nonrelativistic quantum mechanics (cf., for example, (119) § 54)
an electron state is described by a two-component wave function—the

, where ± 1 / 2 ) are the components of the spinor.


When the coordinate system is rotated through the infinitesimal
angle H = v§, where v is the unit vector along the axis of rotation, the
spinor components transform in the following manner:

(9.1)

Here cp'11 are the spinor components in the new coordinate system (a
summation is carried out over the index p), are the elements of the
matrix
a (v) = o x Vx + O u Vy + Og Vg

and a is the set of the three Pauli matrices

In accordance with the general relation between the infinitesimal


rotation operator and angular momentum the matrix |ct is the operator
for the intrinsic angular momentum of the electron—its spin—while
the transformation (9 . 1 ) is an unambiguous consequence of the basic
physical fact that the electron spin is equal to
The Pauli matrices possess the commutation properties of the com­
ponents of the angular momentum operator
oxay—oyox = 2ioz, cty oz —oz <7y = 2iox, ozox—oxoz = 2icy.
[73]
74 QUANTUM ELECTRODYNAMICS

Moreover, they also satisfy the following relations:


cfiakJt~akai = 2(5
From these properties it also follows that
(9.2)
where eikl is the unit antisymmetric tensor of the third rank.
On integrating (9.1) over the angle # we obtain the transformation
law for spinors corresponding to a rotation through a finite angle 6
about the v axis:
9 o = exp V•

Since (o,(,,))2n = 1 , (o-(*’ ) ) 2 n + 1 = a(v){n is an integer), then on expanding

exp ‘n a scries, we obtain

/ 0 , . . 0 ,A ,
cp = I cos y + j s m — aw I <p .

From this formula it follows that when 6 = 2ti, <p = —<p'. But rotation
through an angle 2 n is physically equivalent to the identity transfor­
mation. We thus see that the definition of a spinor contains an ambi­
guity, since the sign of the spinor remains undetermined. Therefore,
physical quantities cannot involve spinors linearly.
Spinors, like tensors, may differ in the way in which they transform
under inversion. The eigenvalues of the inversion operator I are usually
determined by the fact that a double inversion is equivalent to the iden­
tity transformation; i.e., I 2 = 1. Therefore, two cases are possible when
the operator I acts on the spinor <p:
<p' = l(p = ±(p. (9.3)
However, in view of the double sign in the definition of spinors another
definition of the operator for the inversion of a spinor is possible. Indeed,
a double inversion may be defined not as the identity transformation,
but as the transformation corresponding to a rotation through the
angle 2 n\ i.e., instead of the operator I we can introduce the operator
Ils which has the property If 95 = —cp, and, consequently, the law of
transformation of spinors under inversion can be formulated in the
following manner:
9 / = l x<
p = ±i<p. (9.4)
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 75

(If in (9.3) or (9.4) the upper sign holds, then we adopt the convention
of calling y a polar spinor, while in the case of the lower sign we call
it a pseudospinor.)
The alternative definitions (9.3) and (9.4) of the inversion operators
are equally valid.

9.2. Dirac Equations. Dirac Matrices


The relativistically invariant equations which in the case of the
electron play the same role that Maxwell’s equations do in the case
of the photon, were obtained by Dirac (47). They consist of the follow­
ing system of homogeneous first-order differential equations in the two
spinor functions y and %:

i-jf=™ y+<*vx>
(9.5)

^ 7 = -w *+ap9>.

where p = —/V is the momentum operator and m is the electron mass.


We note that Maxwell’s equations can also be written in a form
analogous to (9.5) if we introduce the following formal substitution:
y -> E, y - ^ i H , a -+ s, where s is the photon spin operator defined
by formula (3.3), and if we set m = 0.
Later, in Chapter III, we shall obtain the Dirac equations starting
from a general variational principle. But here we shall investigate specific
properties of the system of equations (9.5).
The set of two spinors y and % can be represented as a single four-
component quantity:

which is called a bispinor.


In order to be able to write equations (9.5) in the form of a single
equation in y, we introduce the four-dimensional matrices (the Dirac
matrices)
(9.6)
76 QUANTUM ELECTRODYNAMICS

where each element of the matrices a and fi denotes the corresponding


two-dimensional matrix. With the aid of the matrices a and /? it is pos­
sible to write (9.5) in the following form:

lw = (a p + ^ m)^- (9
Equation (9.7) has the form of the Schrodinger equation,
• u

in which the Hamiltonian operator is given by the following expression:


H = ap+ySm. (9.7')
From the definition of the matrices a and /3 it follows that
ai ak+ akai = 2$ik’ a/?+/3a = 0. (9.8)
We also introduce the following auxiliary four-dimensional matrices:

The matrices Z have the same properties as the matrices a. Moreover,


as may be easily seen, the following relations hold:
^ -s? = o , £Z -Z £ = o, fc+eP = o,
ea = i, « = e2 ,
[a, a ]= 2/Z, |a, Z] = 2/a,
ai^k = Z iak = Q8ik+ ieikiar
We note that the unit matrix, the three matrices £)(_/ = 1 ,2 ,3 ), the
three matrices q, which commute with them, where = q, q2= ifig,
and q3 = /3, and their nine products QjZh (a total of sixteen matrices)
form a complete system of linearly independent four-dimensional matrices.

9.3. Unitary Transformations o f Bispinors


The components of the bispinor wave function xp can be subjected
to the transformation xp’ = Uxp, where U is a certain unitary four-rowed
matrix. This transformation leaves the form of the Dirac equation (9.5)
unaltered if, in accordance with the general rules, the Dirac matrices
are also subjected to the following transformation:
a,i = U a iU+; 0' = UfiU+. (9.6')
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 77

The commutation relations and the other properties of these matrices


are not altered as a result.
We give examples of such transformations.
If
1
U= 0 +e),
w
then the Dirac matrices transform as follows:
P = Q, e' = P, S'=E,
a' = e 'S ' = j8S,
while the transformed wave function y>' = Uys has the following form:

where

£ = - j ^(<p + x ), v = -j=(<p-x)-
|/2 j/2
Tlie components o f,the spinor £ are usually denoted by f A(2 = 1, 2),
while the components of the spinor rj are denoted by ^(A = i, 2 ).
We note that under Lorentz transformations (cf. subsection 9.6)
the spinors £ and t) transform independently of each other, while under
the transformation of inversion (9.3) they go over into each other:
U = V>
As a second example we take U in the following form (130):

U{M>=
y 2

If we designate the quantities transformed by means of this matrix


by the index M, we obtain
<4 M) = - a x; a <M>= —a2; £ xl M>=
a (M) = p . pm = v = - Q, Z X.

The matrices at(M) are real, while and 27/M) are purely imaginary,
and the Dirac equation contains only real coefficients. It has the follow­
ing form:
78 QUANTUM ELECTRODYNAMICS

1
The choice of the matrix U in the form t / (M) = (/3+Oy) is referred
7^
to as the Majorana representation.

9.4. The Necessity for Four-Component Electron Wave Functions


We now investigate the reason for the doubling of the number of
components of the electron wave function in relativistic quantum mechan­
ics in comparison with nonrelativistic quantum mechanics.
In order to do this it is convenient to go over with the aid of the
unitary transformation U = ——(/3+g) discussed in subsection 9.3,
j/2
from the matrices a, /? to the matrices o', /?' defined by formula (9.6').
The spinors £ and rj forming the bispinor y> now satisfy the follow­
ing equations:
.d£ _ ap£+m ?j,
1 dt
(9.5')
—op ??+m£

We see that due to the presence in the Dirac equation of the matrix
the first equation (9.5') for the spinor £ now contains the components
of the spinor rj, while the second equation for rj now contains the com­
ponents of the spinor £. If the mass of the particle is equal to zero, then
there is no such mixing of components of the two spinors, and equa­
tions (9.5') assume the form

/-—- = ± a p £ . (9.10)

These equations, like the original Dirac equations, are relativistically


invariant.
Thus, for a particle of zero rest mass we can obtain a relativistically
invariant equation containing a two-component, rather than a four-
component, wave function (209). However, it can be easily seen that
this equation will not be invariant under the transformation of space
inversion.
Indeed, we investigate the transition from one coordinate system
to another coordinate system moving with respect to the former with
an infinitesimal velocity v. Since the four matrices 1, ax, cry, form
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 79

a complete system of linearly independent two-dimensional matrices,


the transformation which contains v linearly and reduces to the identity
transformation for v = 0 must necessarily be of the following form:

where A is a constant. But an expression of this type is not invariant


under an inversion, since v££* is a polar and is an axial vector.
This may also be seen if we compare the foregoing expression with
formula (9.1.) Both transformations are of the same form, but in (9.1)
3 is a pseudovector, while v is a polar vector.
Thus, the two-component wave function satisfying equation (9.10)
may be used to describe a particle of zero mass only if we drop the
requirement of the invariance of the equations under space inversion.
Such a situation exists in the case of the neutrino which is described
by equations (9.10); here the — sign corresponds to the neutrino, while
the + sign corresponds to the antineutrino (114), (123), (166a).
In the case of a particle of finite rest mass the requirement of rela­
tivistic invariance leads to the conclusion that the wave function must
have four-components. Moreover, as will be shown later, the equations
determining this function, i.e., the Dirac equations, will also be invariant
under the transformation of space inversion. This invariance is due
to the fact that the bispinor y> comprises the spinor <p and the pseudo­
spinor %, which have different transformation properties under inversion.

9.5. Symmetric Form of the Dirac Equation. Equation o f Continuity


The Dirac equation may be put into a more symmetric form if we
introduce the matrices y^i/J- — 1,2, 3, 4), which are defined as follows:
y} = —ifiaj O '= 1 ,2 ,3 ) ,
(9.11)
y* = P-
The matrices y satisfy the following commutation relations:

and are Hermitian, like all the matrices introduced earlier:


v i = vv y ^ = y*
(y+ denotes the Hermitian conjugate matrix, while y ^ y j denotes
the transposed matrix).
80 QUANTUM ELECTRODYNAMICS

We multiply both sides of the Dirac equation (9.7) on the left by 1/3.
On utilizing the definitions (9.10) we obtain
( n Pi+YV-™ )V’ = °>
3 3
where p , = —i^r— = — and y is the set of the matrices y lf y 2, y3-
dxt ot
If we introduce the abbreviated notation for the “scalar product” yvp v

P = 2 Pv7v = YP+y 4 P4>


V

then the Dirac equation assumes the following form:


(ip+m)y» = 0 . (9.12)
The operator ip which appears in (9.12) is not self-conjugate. There­
fore, the complex-conjugate wave function if* satisfies an equation
which does not coincide with (9.12). In place of if* it is convenient to
introduce the function
if = if* ft; if* - ifft,
which satisfies the equation
y>(—/’p+ m ) = 0 (9.13)
(it is understood here that the differential operators pa operate on the
function if which appears on their left).
We now prove the validity of formula (9.13). Since p* = /V = —p
and p* = p4, it follows from (9.12) that
(—■
iy* P4 + 'Y *P + m)V’* = 0.
Further, on noting that y*if* = if*ya, and on multiplying the last equa­
tion on the right by = y 4 we obtain equation (9.13) after making
use of (9.8).
We now multiply equation (9.13) on the right by —if, multiply equa­
tion (9.12) on the left by if, and add them. This yields
W n ip ^ + ip ^ y jp = o
or
^ = ° , (9.M)

where the following notation has been introduced:


(9.15)
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 81

Relation (9.14) may be interpreted as an equation of continuity,


and s^ may be interpreted as a four-dimensional flux density vector.
In three-dimensional form equation (9.14) has the form
dsn
- a r + d ,V J = 0>
where
= v>*v>>
s = iyjyip = ip*ay).
Just as in nonrelativistic quantum mechanics, the quantity s0
= y>*ip represents the probability density of the electron being localized
at a given point in space. The equation of continuity allows us to nor­
malize ip so that j ip*ipdr — 1 .

9.6. Invariance o f the Dirac Equation


We shall now prove the relativistic invariance of the Dirac equations,
i.e., their invariance under spatial rotations, proper Lorentz transfor­
mations, and the transformation of inversion.
Let us begin with the transformation of space inversion and define
this transformation in the case of the bispinor ip(r, t), in accordance
with formulas (9.3) and (9.4) in the following manner:
y>(r, t) -►ip'(r, t) = lip (r, t) = t), ^
xp{r, t) ->• y'(r, t) =1 y»(r, t) = r)*ip(-r, t)yt,
where rjp has one of the four values, + /, —i, + 1 , —1 .
It can be easily shown that this transformation leaves the Dirac
equation unaltered. Indeed, in equation

( - y * Y t ~ i y i F ~ im) v’*r ’ ') = °


we replace r by -r\

{ ~ y , §i + ' Y i () = °-

and substitute into this expression in place of y>(—r, t) the function


r]-l y l ip'{r, t) in accordance with (9.16):
82 QUANTUM ELECTRODYNAMICS

By utilizing (9.11') we can move the matrix y 4 immediately preceding


xp' to the left. Then, on multiplying the resulting equation on the left
by y 4 we obtain the original Dirac equation which now contains xp'
in place of xp.
Corresponding to each of the four values of rjp we have four types
of bispinors which are usually denoted by A(r)p = i), B(rjp = —i),
C(r]p = 1) and D{rjp = —1). The bispinors A and B (and also C and D)
can be used to describe two kinds of particles obeying the same Dirac
equation but differing in the way in which they transform under inver­
sion; it is commonly said that these particles differ by the sign of their
“intrinsic parity” .
We now investigate the transformations under rotation. It follows
from (9.1) that under a spatial rotation through an infinitesimal angle
ft the transformation of the bispinor xp has the form

¥> = | l + —ftsjy /.

By utilizing the general relation between the infinitesimal rotation


operator and the angular momentum operator we can conclude that
the matrix 2 represents the operator for the intrinsic angular momentum
(spin) of the electron.
The following transformation corresponds to a finite rotation through
the angle 6 about the v axis

— O vE / o q \
xp = e 2 xp' = I cos — + iv 2 sin— \xp'.

In accordance with the arguments presented in subsection 9.4 we


assume that the following transformation of the wave function corre­
sponds to the infinitesimal Lorentz transformation
xp = (l+£5iia)y>',
where 5u is the infinitesimal relative velocity of the two coordinate sys­
tems. In this case the following transformation corresponds to a finite
Lorentz transformation defined by the finite velocity u:

xp= elrnaxp' - cosh ■


—+ not sinh—) xpr,
■)
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 83

where

<p — arc tanh u, n U.


u
Both formulas for the infinitesimal transformations may be amal­
gamated into one,
v = ( l + ^ y ay/,)v', (9.17)
where # a(3 is the four-dimensional antisymmetric tensor for the infin­
itesimal rotation

&u = 2 eukdk; ^ 4/ = y ( f , j , k = 1 ,2 ,3 ).

We shall now demonstrate the invariance of the Dirac equation


under the transformation (9.17).
On substituting (9.17) into equation (9.12) and on multiplying it
on the left by (l + 2 ^ „ y vyu) we obtain
(1 y„yM) (y,. v >.+ 2 tfa/}yap'P- im) (l + \'9 ceyayeW = o,
where p' are the momentum components in the new coordinate system
P a = Pa + 2 ^ p ' .
On neglecting terms quadratic in $a/, we can rewrite this equation in
the following form:

|y aP a - f w - - | l- ^ ( y ay^+y^yn) + i ^ ( y ^ y ap + p y oy(3)-f2i7a/9yap ^ ' = 0 .

On noting that §ap = —§fta and

YffYa P' = p'y/>ya + 2y/jP a-2yaPp»


we can easily verify that terms linear in # a(9 vanish, and the transformed
equation becomes identical with (9.12).
We can also obtain the transformation law for yj. We take the
expression complex conjugate to (9.17),

V* = V*' (l+iKpYpya)’
and multiply it on the right by y4. Since = — -&t) (i,j = 1, 2, 3) and
= —# 4j, we obtain
V>= V'O—\®af>YaYp)- (9.18)
84 QUANTUM ELECTRODYNAMICS

9.7. Bilinear Combinations o f the Components o f the Wave Function


In relativistic quantum mechanics, just as in nonrelativistic quan­
tum mechanics, considerable significance is attached to bilinear combi­
nations of ip and y>*. Since the wave function y> has four-components,
sixteen such combinations can be constructed. From them quantities
may be formed which behave under Lorentz transformations as a scalar
S (one-component), a vector V (four-components), an antisymmetric
tensor of the second rank T (six-components), an axial vector A (four-
components), and a pseudoscalar P (one-component). We shall show
that these quantities have the following form:
S = - y>y>,
VM=
t »v =2V>( (9 -i9)
^ = w ^YbV*
P = y y sy),
where
Ys = YiYiYaYi = ~Q- (9-20)
In order to do this we utilize the transformation formulas for and
V>. First of all we consider S. It follows from (9.17) and (9.18) that
y > V = y ( \ - \ & aliyayji) { \+ \& aliyayf>)xp'.
This equation contains no terms linear in $a(3, i.e., y y = yi'y)', and,
consequently, S is an invariant. Under inversions S also remains unal­
tered.
In a similar manner it can be easily shown that

VYfi> = w 'Y f.f+ ^ 0 ^ ( 7 ^ aYp-YaY^Y


or by utilizing the properties of the matrices y ,
Vft = F /<'+12 i? ptv V"v •

Under inversion WY^W = WYiYftYiW' or Vj = - V ] , Vi = V ', i.e.,


transforms like a vector. From this it can be seen that the flux density
vector introduced in (9.15) is indeed a four-vector.
In a similar manner it can be easily shown that T transforms
like a tensor.
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N *5

lit order to establish the transformation properties of the quantities


Ap and P it is sufficient to utilize the property of the matrix yt which
follows directly from its definition,
y*y„ = -y„y»-
From this it follows that the matrix y5 commutes with the matrix y yv
which appears in the transformation under a rotation (9.17), (9.18),
and anticommutes with the matrix y t , which defines the transforma­
tion of the inversion operator in accordance with (9.16). Therefore,
the extra factor ys in P and A in comparison with S and V does not
change the character of the transformation of these quantities under
rotations, but changes the sign of the transformation under inversion.
Thus, P is a pseudoscalar, and A is an axial vector.
Since the components of S, V, T, A, P exhaust all the sixteen bilinear
combinations which may be formed from the components of y and y,
it is clear that there exist no other independent quantities. In particular,
the quantity y y a yazya3 ... which transforms like a tensor of the
nth rank may be reduced to quantities of the types already investigated.
For example, the tensor T'^ = lyy^iy^yy—y^/^W is dual with respect
to the tensor T ^ , T^v = - \ e aPllvTar
We also give expressions for the components of V, A, and T in
three-dimensional form:
V = (V, V0) = - iy * ( a , l)y,
A = (A ,A g)= y*(Z, g)y,
T = ( H , E ) = iy * ( f1 X J * ) y .
Here V and A are the spatial, and V9 and A e are the time-like compo­
nents of four-dimensional vectors, E is a three-dimensional polar vector
and H is an axial one, which together are equivalent to the tensor Tpv
(T12 = H 3 , T ti = iE j).
It follows from (9.19) that any quantity which transforms like
W^Pv (we shall call a spinor ° f the second rank) can be expressed in
terms of the matrices ya in the following manner:
RMy= (S -l+ P y 6+V',ya+ A aysya+ Tapyayp)/lv, (9.2l)
where S is a scalar, P is pseudoscalar, F4s a vector, A is an axial vector,
and T is an antisymmetric tensor of the second rank.
86 QUANTUM ELECTRODYNAMICS

§ 10. Electron and Positron States. States of Definite Momentum and


Polarization

10.1. Solutions with Positive and Negative Frequencies


The general solution of the Dirac equation can be represented in
the form of a Fourier integral: y>= f ipkeikr dk. The Fourier compo­
nent ipk is the electron wave function in momentum space. In accord­
ance with (9.7) it satisfies the equations

‘ dg f = ( a k + P m ) rP k -

If we write xpk in the form


<Pk\
V»* =
Xkl
we have
i 3- - = m(Pk+°kxk,
ot

1— = — mXk+^k(Pk-

We seek the solution of this system of equations in the form


Vk = Vo(k)e~i(ut
or
( <Pk\ _ /P o(*)\
0—i(vt

Then the equation for y;k is


(a k-\~f}m—td)xpk = 0,

or (<o-m)(Pk-akXk = 0,
-ok(pk+(co+m)xk = 0.
From the condition that this system of equations should have nontrivial
solutions
| co—m —ok
! —ok co+m
we obtain the frequency ro = ^ e , where e = )/Ar2 + m2. The last formu­
la expresses the well-known relativistic relation between energy and
momentum.
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 87

We see that there exist two kinds of solutions of the Dirac equation:
solutions with positive and negative frequencies. The general solution
of the Dirac equation has the following form:
ip =
where
ip(+) = J ip\+'>(k)elikr-*t) dk,

y>i~) = j k)e~i(kr~et) d k .
Such decomposition into positive and negative frequencies is relativ-
istically invariant, since the sign of the frequency cannot change under
Lorentz transformations. This can be seen, from the fact that the
lowest positive frequency is equal to m, while the highest negative
frequency is equal to — m; i.e., the domains of positive and negative
frequencies are separated by the finite interval 2m, while Lorentz
transformations contain a continuous parameter.
Of the two spinors y k and xk forming a bispinor, one is an arbi­
trary function of A:, while the other can be expressed in terms of the
first. For example, by specifying (p<k+) and Xk'] we can obtain and
xi+)‘
ok
e+m <Pi+)■
.
( 10. 1)
ok
vP = e+m XiT>
Solutions with positive and negative frequencies belong to different
eigenvalues of the self-conjugate operator H = ap-f/?m. Therefore

K + )> J r ) = <Pk+)t(pj r )+%fc+ )* * Jr) = 0


or
j ipt'+)*y>(- -)dr = 0 .
These relations can be easily verified with the aid of (10.1).
The existence of two types of solutions is of fundamental significance
for the theory of the electron. We examine the special solution of the
Dirac equation in the form of a monochromatic wave
V = Vo( 0
QUANTUM ELECTRODYNAMICS

The function f 9(r) is an eigenfunction of the Hamiltonian operator

H -- ap+iSm, Uip9 = a>y0 = ± e y 9.

In quantum mechanics the eigenvalues of the Hamiltonian operator


are interpreted as values of the particle energy. However, in relativ­
istic quantum mechanics of the electron it is impossible to retain
this interpretation. Indeed, the existence of wave functions with neg­
ative frequencies would imply the existence of electron states of negative
energy, and the absence of a lowest energy state. This means that in
interacting with other particles an electron could give up unlimited
amounts of energy by going to ever lower energy states. The physical
absurdity of such a conclusion requires a change in the fundamental
assumption of quantum mechanics which relates the possible values
of physical quantities to solutions of corresponding wave equations.
The altered assumptions must ascribe positive energies to states
with negative frequencies. However, two different states will then
correspond to a given value of the energy states with different signs
of the frequency. Therefore, the altered assumptions must give a physical
interpretation to this twofold degeneracy. We assume that the Dirac
equation applies to electrons of both signs of charge (the electron and
the positron). In this way a possibility arises of interpreting the sign
of the frequency as characterizing the charge state of the electron.
Corresponding to a positive frequency the electron must, for example,
have the charge e, while corresponding to a negative frequency it must
have the charge -e. We shall see later that such an assumption provides
the possibility of constructing a theory of electrons and positrons and
of their interaction with the electromagnetic field.

10.2. The Charge Conjugation Transformation


To facilitate the formulation of changes introduced into quantum
mechanics as a result of the foregoing requirements, charge conjugation,
a transformation of great significance, is introduced. We saw in § 9
that the wave function ip and the function ip — ip*ft obey different
transformation laws in the transition from one coordinate system
to another one and satisfy different equations. We shall show that it
is possible to construct linear combinations of the components of ip
which will satisfy the same equations as the components of and will
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 89
transform in the same way as the components of rp. We denote these
combinations by y c and set
VG= C y ,
where C is a four-rowed matrix.
The function rpc is said to be charge-cmjugate with respect to rp.
We now investigate the conditions that must be satisfied by the
matrix C in order that the function f c should satisfy the same equation
as rp; i.e.,
( /> + m ) v > c = 0, p m= - i — .

In order to do this we use the Dirac equation for rp

(~-ipeya+ m)v--= 0.
Oh substituting into this expression in place of rp, rp= C " Y , we obtain

( - /> a>V-l-m)C ~1¥’c = 0 -


If the matrix C satisfies the conditions —yaC - 1 - C- 1ya then this
equation may be rewritten in the form C~1(ipJrm)rpc = Q, and after
being multiplied on the left by C it may be reduced to the Dirac equation.
Thus, the matrix C, which is called the charge conjugation matrix,
must satisfy the following conditions:
C y a = - y ac (10.2)
in order that rp and rp° should satisfy the same equations.
It may be easily shown that if these conditions are satisfied, the
function will transform under Lorentz transformations in the same
way as the function rp. Indeed, in accordance with (9.18), under infin­
itesimal four-dimensional rotations the function rpc transforms in
the following manner:
r f = C y '(l- 4 0 B/Jyaf y ) = C ( l - ^ ( y ay/}))'V ,
where the superscript T denotes transposed matrices.
We require that the function rpc should transform in the same way
as rp, i.e., that the following relation should hold:
v8 = ( i + \ ^ y ayPW ,
where
rpG>= C y ' .
90 QUANTUM ELECTRODYNAMICS

In order to achieve this the matrix C must satisfy the following con­
ditions:
C(yayp)T = - y a7 pC, a ^ \3
or
Ca = —aC.
But these conditions follow directly from (10.2).
We now investigate how yjc transforms under space inversion.
We assume that the transformed functions y c> and yj' are related
by the same expression which connects the functions ip° and yj; i.e.,
y>c>= Cy>'.
The functions y) and y>transform in accordance with (9.16) in the follow­
ing manner:
¥ = %y*v>
v = v tw *
where % = + / , ± 1. We set
yC> = rjCy^O

where rfp, like r/*, may have any one of the four values ± 1, ± 1 . On sub­
stituting this expression into the equation relating yj°' to yj', we obtain

v° = — — y i C ^ 4.
—v P
Further, on utilizing (10.2) we obtain
n* C y>,
c=
VP
and, consequently, rfp = —rj*. Thus, if r]p = ± 1 , then rfv = q=l, while
if Vp = ± i, then also rjcp = ± 1.
In other words, if under inversion the spinor transforms in accord­
ance with (9.4), then the charge conjugate spinor transforms in the
same way; but if the spinor transforms in accordance with (9.3), then
the transformation of the charge-conjugate spinor differs in sign from
the transformation of the original spinor.
We shall now show that a matrix C with such properties actually
does exist. In order to do this we utilize the Majorana representation
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 91

(cf. subsection 9.3). Since the matrices ajM ) are real, while the matrix
/9(M) is imaginary, the conditions ( 1 0 .2 ) will be satisfied if we set

C<M> = = - v (10.3)

It may be easily seen that in the Majorana representation the charge-


conjugate function coincides with the complex conjugate function y/M>c
= yjW>\
We obtain the transformation law for the matrix C corresponding
to a unitary transformation of ip. Let ip' = U ip, ipc' = Uyj°. On intro­
ducing the relations ipc' = C'ip' and ipc = Cip and on comparing these
formulas we obtain
c = ucu,
(10.4)
c= U+C'U*.

Since in the Majorana representation the matrix C has the following


properties, as may be easily seen from formula (10.3):

C = -C ,
C+C=l, (10.5)
C*C = - 1 ,

then by utilizing (10.4) it may be easily shown that these properties


do not depend on the choice of the representation.
On going over from the Majorana representation to the usual one,
i.e., on setting

C '= C (M), U = — (v fiS ),

we obtain in accordance with (10.4)

C = ay. (10.6)

We exhibit the different relations between ip and i f:

ip° - Cyi, i f = C+ip,


(10.7)
ip=Cipc, ip = C+ipc.
92 QUANTUM ELECTRODYNAMICS

We also give the expressions for the bilinear combinations (9.19)


m terms of the charge-conjugate functions

S=

VM= ^ y ^ >

Tp, = —^ ( y ^ y . - y ^ X f ^ 0 &8)

a? = - y ^ i y ^ y ^ w 0,
f = —v cysve-

10.3. The Positron Wave Function


At this point let us consider the electron wave function in momentum
space. Let
ip = j ip{k)eikzd k ,

■f = J ijKk) e~ikz d k ,

where
kx = k r—<x)t, o) = i \ / k 2-\-m2, ip(k) =ip{k, co).

On substituting this expansion into (9.12) and (9.13), we obtain:

(ik-\-m)ip(k) = 0 , ip(k) (ik+m) = 0 .

On multiplying the last equation on the left by C and on moving


Cf{k) to the right, as was done earlier, we obtain

(—ik-\-m)if{k) = 0 .

We see that ipc(k) satisfies the same equation as —k). Since


the replacement of k by — k includes the replacement of <y by —<w,
we see that f c(k) is a solution with positive frequency if f{k) is
a solution with negative frequency.
We introduce the following notation:

Vie\ k , e) = yr(k, e),

^ (p)(*< e) = y>c(—k, —e), (e = H )


RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 93

and correspondingly

f le) = J f( k , e)ei(kr~Et)dk,

yi(_) = § y (k , —e)ei(kr+e°dfc.

If we take for the particle wave function corresponding to negative


frequencies not y (~ \ but y>(p), then its energy will be always positive.
We shall call y (p) the positron wave function.
It should be emphasized that in defining the positron wave function
by means of these relations we go outside the framework of transform­
ations admissible in quantum mechanics, since we subject a part of
the solutions of the Dirac equation to a nonlinear transformation (the
transition to a complex-conjugate quantity is a nonlinear operation).
As a result of this, the general solution of the Dirac equation is now
y = and does not have the meaning of a wave function.
On the other hand, the superposition of states of opposite sign of the
charge ^ (e)+ y (p) cannot give a general solution of the Dirac equation
since solutions with positive frequencies do not form a complete system
of functions. This circumstance is, of course, immaterial as long as we
are investigating states of a free electron with a definite sign of the
charge. In this case the interpretation of the states with negative fre­
quencies as states of the opposite sign of charge is of a purely formal
nature, as long as the interaction with the electromagnetic field is not
taken into account. Later we shall see that the use of the transformation
yM _ c^<-) enables us to construct a general theory for a system of
electrons, taking into account their interaction with the electromagnetic
field, with the states y le) and y (p) actually corresponding to the charges
e and ~ e .
In concluding this subsection we also investigate the question of
the relative intrinsic parity of the electron and the positron. In order
to do this we must determine the transformation properties under space
inversion of the wave functions of the electron and the positron at rest.
On setting / = 0 in the Dirac equation we can easily show that
these functions satisfy the following equations y4^ (e) = y {e) and yiy)(p)
= y {p). As has been shown earlier, the functions y M and y {p) transform
under inversion into y {eY = r]pyi y (e) and y {pY = rjpyt y {p). On utilizing
the equations for y ie) and y ip), we obtain y leYy ipY = rjprjly(e)y ip). But
94 QUANTUM ELECTRODYNAMICS

in accordance with ( 1 0 .2 ") we have yfv = —rj*, and therefore ip<e)'ipipy


- —iplehplp\
This relation means that the electron and the positron have intrinsic
parities of opposite sign ( 1 1 ).

10.4. Plane Waves


The electron states of definite frequency may be described by Hyi
= anp, where H is the Hamiltonian operator defined by formula {9.1').
States corresponding to a definite value of the frequency are degener­
ate and in consequence we require that the wave functions of these
states simultaneously be eigenfunctions of other operators which com­
mute with each other and with the Hamiltonian operator.
We begin by examining states of definite momentum and denote
the wave functions of such states by ipp. They satisfy the equation
prpp = pip . For the sake of definiteness we take the frequency to be
positive so that a> = £ = + | / p 2 + m 2.
Then we have
I = ^-P pipr— iet (10.9)
- p a ’

where up is a constant bispinor (we shall also call it the spinor ampli­
tude), which satisfies the equation (ap-j-fim—£)up = 0 , and Q is the
normalization volume. We assume that up satisfies the normalization
condition u*up = 1. We also have: f \ip\2 dr = 1.
a
We write u in the form of a column consisting of two spinors vv
and vv':
«p = ( j W , (10.9')

where N is a normalization constant. The spinors vv and iv' are related


to each other by equations (10.1). The spinor vv may be taken to be
arbitrary, with

w' = - ^ - v v . (1 0 .1 0 )
e+m

We assume that vv is a unit spinor, i.e., that it satisfies the condition


w *w = 1. Then the constant N has the value N = ]/(e+m )/2e.
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 95

10.5. Polarization o f a Plane Wave


For a given value of the momentum two different electron states
are possible corresponding to two linearly independent spinors w. We
refer to these two states as states of different polarization.
The two linearly independent spinors w may be chosen, for example,
in such a way that they are eigenfunctions of the operator \ a z. We
denote these eigenfunctions by % \

where X is the spin variable which assumes the two values X =


From the form of the matrix oz it follows that p, = ± 1 and xfX) = d
We note that if w = xMthen vv' is not an eigenfunction of the opera­
tor az. Indeed, in accordance with (10.10) we have

However, if we choose the z-axis in such a way that it is directed along


the vector p , then a p = poz and w' will also be an eigenfunction of
the operator oz:

These results are a consequence of the fact that the electron spin
operator does not commute with the Hamiltonian operator
H = ap+fim, while the operator pE does. Therefore, the eigenfunctions
of the operator H, in general, are not eigenfunctions of the operator
£ z, but can be eigenfunctions of the operator pE:

( 10. 11)

These amplitudes describe states with a definite value of the component


of the spin along the direction of motion. We denote the plane waves
corresponding to them by
96 QUANTUM ELECTRODYNAMICS

We shall refer to states for which the bispinor up is defined by for­


mula (10.11) as states of definite polarization. The bispinors upft corre­
sponding to different values ofp are mutually orthogonal u*pupp. = bptl..
In a similar manner we can discuss states of definite momentum
p and of negative frequency a> = —e. We shall write the wave functions
of such states in the form = (o / ) / Q)el^r+Ut, where vp is a unit
bispinor satisfying (a.p-\-^m-\-e)vp = 0 .
For a given p there exist two linearly independent amplitudes vp
which may be chosen to be the eigenfunctions of the operator Up :
^ 0 ^ = \p\[lVpp, p = ± $ .
These functions are mutually orthogonal, i.e., o*pv . =
We compare the spinor amplitudes u and o .
The first of these amplitudes satisfies
(ip + m)upfl= 0 ( 1 0 .1 2 )
and the second satisfies
(ip—m )v-pll= 0 , ( 1 0 . 1 2 ')
where p — ( p, ie) . Thus, up and u_p may be regarded as eigenfunctions
of the matrix ip, corresponding to the eigenvalues p p w . Since ip is a
four-rowed matrix, the four linearly independent functions u and
u_pp form a complete system of its eigenfunctions. (We note that
= /?»,-).
On introducing u = uPm and u = « m, we rewrite the equa­
tions for upp and v in the following form:
iP uMA = ~ A um A = ^m , (10.13)
We now prove that

u* A * ) = ~ <W • (10.14)
a

The functions u/tA which are eigenfunctions of the non-Hermitian


operator ip are not orthogonal in the usual sense of this word, i.e.,
u*au/i'a- 0. However, it follows from (10.13) that upAup.A- = 0 for
A ^ A '.
Similarly, it follows from (10.11) that u ^ u ^ . = 0, p ^ p . We now
obtain uM upA. By replacing m in (10.10) by A, we obtain uu
= N 2(w*w—w'*w') = — . Thus, the proof of (10.14) is complete.
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 97

In view of the completeness of the system of functions ujlA the follow­


ing relation also holds

£ = 5 aa'- (10‘15)

10.6. Polarization Density Matrix for the Electron


The spinor amplitude of the plane wave up may in the general case
be expanded in terms of the amplitudes u corresponding to definite
components of the spin along the direction of the momentum:
+ 1/2

up = Z ""V - (10.16)
,<= - 1/2
The expansion coefficients w^ in this expression are the components
of the spinor amplitude of the electron wave function in the rest system.
Indeed, in the rest system in accordance with formulas (10.9)-(10.12)
the amplitudes have the following form:

^ 1/2 \
pt> Xp P= o, (10.17)

and, consequently,
w\ w1!* \
w=
or w-'l2}'
Just as in the case of the photon (cf. subsection 2.4), we can describe
the states of an electron of a given momentum by means of the polari­
zation density matrix
Q„, = w t'w 'f (10.18)
where the bar over the right hand side denotes averaging over the param­
eters which characterize the system of which the electron is a part.
The diagonal elements of the matrix q are equal to the probabilities
that the component of electron spin along the z-axis (the direction of p)
is equal to or — The normalization condition for q has the form
Sp £ = 1.
Just as in the case of the photon, the matrix q may be written in
the form
6— 2 (1 + ^°)* (10.19)
98 QUANTUM ELECTRODYNAMICS

where £ < 1. The parameter £z = £ >p/\p\ determines the probability


of polarization along the z-axis and pn = - |( 1 + Q . The parameters
£x and £y appear in an analogous manner in the expressions for £u in
the system of coordinates which has been rotated by a certain angle
with respect to the original one. Indeed, when the coordinate system
has been rotated about the y-axis by the angle n j l the spinor w is multi­
plied by the unitary matrix exp(£ijrcry) = {\+ ia y)l\^2 and in accord­
ance with (10.18) the matrix q goes over into {?'= £(l + hry)g(l —iay),
from which we obtain ^ = | ( 1 +C)-
Similarly, a rotation about the x-axis through the angle —nj2 corre­
sponds to the transformation of the density matrix q" = +
from which it follows that = -|-(1—£y). We note that the expression
for and q[[ cannot be interpreted as the probability that the com­
ponent of the electron spin along the x- or the >>-axis has the value
However, we can say that is the probability that the component of
the electron spin along the x-axis is equal to \ in the rest system of the
electron (and similarly in the case of
When £ = 0 we have a completely unpolarized state, and when
£ = 1 we have a state of complete polarization. If we introduce the
notation £ = P£0, where | £0| = 1 , then the matrix q may be written
as a linear combination of two matrices corresponding to the unpolar­
ized and the completely polarized states
1 —p p
<? = - 2 - + y ( l + S o ").

The vector £ £ represents the average value of the electron spin in


the rest system. Indeed, in the rest system the average value of E is
equal to E = a = w*aw. On utilizing (10.18) and (10.19) we obtain
a = Sp go = %. (10.20)
The density matrix (10.18) is a matrix in the space of the eigenfunc­
tions upfl of the spin operator. We can also investigate the density
matrix in the “coordinate” representation, i.e., in the space of the bispi­
nor indices
y x a = upWUp(<?) where A, a = 1, 2 , 3, 4. (10.21)
In contrast to the two-rowed matrix (10.18) the matrix (10.21) is a four-
rowed one. However, its values for different A and a are not independent
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 99

if we are discussing, as we did earlier, states with a definite sign of the


frequency. Therefore the matrix (10.21) may be characterized by the
same three parameters as the matrix (10.18).
It is convenient to write the matrix G? in a form in which it is not
necessary to take explicitly into account the relationship between the
different components up(X). In order to do this we write, up = rj(p)u,
where

V(P) = - = £ (10.22)

and u is a certain bispinor, whose components are no longer related


by the Dirac equations. The validity of this equation can be easily veri­
fied if one notes that (ip+m)r](p) = 0 , in virtue of which up satisfies
the Dirac equation for arbitrary u.
Similarly, the amplitude of the state of negative frequency v_p can
be written in the form v_p = rj(—p)u.
The matrix rj(p) is called a projection operator. It can be easily shown
that the following relations hold

(v (p ) ) 2 = y(p), y= v +Vi = v> (10.23)


up = urj(p), rj{p) u/lA = uppSmA. (10.24)
On substituting (10.22) into (10.21) we obtain the following expres­
sion for J>:
= n ) M < a)u(P)(v(p)Y^Pa- (io.25)
We first consider the unpolarized state. We show that in this case
m m —ip
9 = V*- (10.26)

In order to do this we substitute the expansion (10.16) into (10.21),


and this leads to the following relation between 9 and g:
+ 1/2

2 (>0-27)
n, >>=—1/2
Since in the unpolarized state gpv — therefore

9 ;u ,= (*)*#>)■
100 Q U A N T U M E L E C T R O D Y N A M IC S

By utilizing the projection operator we can, in accordance with (10.22)


and (10.24), rewrite this expression in the following form:

where u A are defined by formula (10.13). In virtue of the orthogonality


property (10.15) we obtain

" ka (10.28)

2 » PA * W M = { - 2 ^ r . ) •
fi \ /Aff
which leads to the expression (10.26) for 9)-
On comparing (10.26) and (10.25) we see that in the case of an unpo­
larized state, u(a)~u(^) = mdaji/2e holds.
In the general case of partial polarization the quantity n(a)n(/9),
which in accordance with {10.25) defines the density matrix, must be
of the following form:

(io.29)

where R is a certain four-rowed matrix (a spinor of the second rank).


For the construction of the matrix R we start from the fact that
the density matrix 5>, and consequently also R, depends on three param­
eters. In the electron rest system these three parameters form the
axial vector X,.
Therefore, the state of polarization may be described in a relativ-
istically invariant manner by two equivalent methods: either with
the aid of the antisymmetric tensor t , or with the aid of the axial
vector ahl, which is related to t/IV by t/iv = J - e llvlepx ae, where eflvXe is
the unit antisymmetric tensor of the fourth rank and px is the momen­
tum four-vector.
The tensor tMv must satisfy t.tilpv = 0, from which it follows that
in the rest system only three of its components differ from zero. The
tensor t/lv is equivalent to the two three-dimensional vectors s and d
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 101

(tu = eijksk \ tj4 = — idj ,i,j,Jc = 1, 2, 3), which may be expressed in


terms of the vector ^ in the following manner:

d = j«n> = C(||); ^ > = — Ca ) , (10.30)


m
where the symbols (||) and (_]_) denote longitudinal and transverse com­
ponents of vectors with respect to p.
Similarly, the vector a^ must satisfy aflp/j= 0 . Its components
a}, j = 1, 2, 3 and as = — iat are related to the vector £ by the following
equations:
fl<!l> = _if<il>; fl<!> = £<!>; (10.30')
m e
In accordance with the general expression (9.21) the matrix R may
be represented by either R = — ■^itjkyj yk or R = —iayb. On substitut­
ing this expression into (10.29) and (10.25), and on noting that
r}(p)R = Rr)(p) , we obtain two identical expressions for J>:

? = ^ 1 - ' 2 I (« -» > )? . (10.31)

or (135)
'? = (m—Jp)y4. (10.31')

From these expressions it follows that

t — SdT’v W - W
(10.32)

im
Thus, if the density matrix (7> is known, the state of polarization may
be obtained by determining or au from formulas (10.32), and then
obtaining the vector £ with the aid of formulas (10.30), (10.30').
It is easy to obtain results similar to (10.28) and (10.31) also for
states with negative frequency and of momentum p .‘
(m+ip),a
(10.33)
2e
102 QUANTUM ELECTRODYNAMICS

rj>{
(10.34)
= - ^ ( i - ^ y 5) ( m +ip)y4»

where
9L> =
We note that the density matrix for the positron ^>j^> = u ff uft*
is related to 9)(“ ) by (?(p)y 4 = —(C+j7)(“ )y 4 C)T. If 3>(_) is of the
form (10.34), then j?(p) coincides with expression (10.31).

10.7 Averaging over Polarization States


In future we shall often encounter the following problem. We are
given the expression M = aupfi = J? a(?>)up/t(X), where upp is the spi­
nor amplitude of a state of definite momentum and polarization (and
of positive frequency), while a is a particular bispinor. We are required
to find the sum of the squares of the absolute values of |M |2 over the
different polarization states of the electron:
2 \M\2 = Y a(X)a*(o) £ u ^ u ^ i p ) .
H Xo (J.

This summation can be easily carried out with the aid of formula
(10.28). On utilizing it, and also the definitions a = a*yi , (a)* = y 4a,
we obtain

Y \ a uP,\* = ~ ^ a ( m —ip)a. (10.35)

Let a be of the form a = Rup.y, where R is a four-rowed matrix,


while up.v is the amplitude of the state of momentum p' and polarization
v, and consider the problem of evaluating Y \up'vRuPfi\2■ On utiliz-
V-v
ing once more the formula (10.28), and also the equality a = up v R,
where
R = yi R+yi , (10.36)
we obtain

Y \ u p'VRupM\2 = — Sp {m—ip )R (m ~ ip')R . (10.37)


RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 103

Similar results may also be easily obtained for states of negative


frequency:

2 I ^ - p J 2 = — — a(m + ip)a, (10.38)


fi

E |R_P'v^-p^l 2 = 4— 7 Sp (m+ ip) R(m-\-ip') R, (10.39)

U \UpvR v - p J 2 = — -S p (m + i;)/? (m -if)i (10.40)

The matrix appearing after the trace symbol Sp in formulas (10.37)-


(10.40) represents a sum of terms each of which is a product of several
matrices y . For the evaluation of the trace we can utilize the fact that
the quantities Sp YVlYv2 Yvn form a four-dimensional tensor of the
nth rank. The last assertion follows from the fact that in accordance
with the proof given in § 9 the expression wyViyVl---yVny is a tensor
of the nth rank. Therefore the quantities ^ uA yv yr ... yVn uAft
Ap 1 2

= Spy ■■■Vvn wiU also be tensors if the amplitudes are normalized in


an invariant manner, i.e., uu = 1 .
Since the matrices y p have the same form in an arbitrary coordinate
system, the form of the tensor Spy ...y Vn also does not depend on
the choice of the coordinate system. The only tensor which has this
property is dafs. Therefore the desired tensor can be constructed from
V
From this it follows immediately that if n is an odd number, then
we have
sp(y,iy,l ••• y Vn) = 0 (n = 2 k + \ ) .
If n is an even number, then we have
Sp(yVl ••• y„n) = 2 aPdikdim
p
where i , k , l , m , is some combination of indices vx, v2, v3, ... and ap
are numerical coefficients. The sum is taken over all possible combi­
nations of pairs of numbers ik, Im, ..., with the number of terms in
the sum, i.e., the number of such combinations, evidently being equal
to
104 QUANTUM ELECTRODYNAMICS

For the determination of the coefficients ap it is sufficient to set


i = k , l = m , etc. On utilizing (9.11) and on noting that Sp 1 = 4,
we obtain
iS p (y ...y ) = (n = 2k). (10.41)
P

The signs of the individual terms in the sum may be found in the
following manner (215), (153), cf. also (37). We let each matrix y v
correspond to a point on the circumference of a circle, and we arrange
these points in the same order in which the matrices yv occur. We then
join these points in pairs by straight lines. Then to each straight line
joining the points i and k there corresponds a factor dik, while to each
way of joining points (or, in other words, to each decomposition of
the subscripts i, k, /, m, ... into pairs i—k, l—m , ...) there corresponds
a term in the expansion of the trace (—l)pdifcd m ..., where P is the
number of points of intersection of the straight lines. Thus, we have
« , = ( - ! )P-
We give the values of Sp y1 ... y„vn for n = 2, 4, 6 :
i Spy,y* = Sik,
i Sp y i y i c Y i y m = ^im^kl ^U ^km’

i SP Y i7k7i7m Y T 7s = $ik <5/ m + d is <5W S mT + d i k d l s d mr + 8 i m &k l d rs

+ $kT dim + $il ^kr ^ms + m $ks $It + ^ ir ^km $ls ~ ^is m & Is

We note that the expression for the matrix y ... yv can be simplified
if among the subscripts there are two identical subscripts vt = vk = a
over which summation is carried out. By utilizing equation (9.11) we
can rearrange the product in such a way that the two matrices with
identical subscripts turn out to be side by side, and after this we can
carry out the summation: yaya = 4. For example, if two matrices are
separated by one, two, or three factors, the following relations hold
(it is understood that summation is carried out over the subscript
a from 1 to 4):
yayvya = - 2y ,>
YaYrYvYa = (10.43)
yayxylly ,y a = - 2y ,Y mYx-
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 105

§ 11. Electron States of Definite Angular Momentum and Parity

11.1. Orbital and Spin Functions. Spherical Spinors


We now investigate electron states characterized by definite values
of energy and of angular momentum.
The angular momentum operator M is a sum of the orbital angular
momentum operator L = [r, p] and the spin angular momentum oper­
ator ^Z ; M = L + | Z . We shall denote the eigenfunctions of the oper­
ators of the square of the angular momentum M 2 and of its compo­
nent M 2 corresponding to the eigenvalues M 2 = j { j + \ ) and M z = M by

They satisfy the following equations:

= MxPiM•

Each of the spinors <pjM and satisfies these same equations if we


replace in them the four-rowed matrix Z by the two-rowed matrix
a, i.e., if we set M = L + i a .
Just as in Chapter I, we do not solve these equations directly, but
in order to obtain the wave functions we use the quantum mechanical
rules for the addition of angular momenta, by regarding the orbital
and the spin degrees of freedom of the electron as two subsystems.
We already know the eigenfunctions of the operators for the orbital
and the spin angular momenta. The eigenfunctions of the orbital angular
momentum are the spherical harmonics T,m(r/r), where m and / ( / + 1 )
are the eigenvalues of the operators L, and L2, while the eigenfunctions
of the spin angular momentum are the functions xM(^) (cf. sub­
section 10.5).
In accordance with the rules for the addition of angular momenta,
a given value of the total angular momentum j can be obtained for
two values of angular momentum of the orbital subsystem 1 = j ± \ .
We use the notation <pjlM to denote the spinor function <pjM which corre­
sponds to the same values of j and M, but to different values of /. We
106 Q U A N T U M E L E C T R O D Y N A M IC S

can represent it as a linear combination of products of orbital functions


Ylm and spin functions xft:

C
PjlM ~ a (r) ^ Ciml/2M^Im(fl) X,i > n
(j.m
where a(r) is a radial function which we shall define later.
This equation can be regarded as an expansion of the spinor rp in
terms of the orthonormal spinors (pi?) = which are
the basic unit spinors analogous to the basic unit vectors ^ in § 3.
We call qf the contravariant spinor components. Since
we have <p(X) = <px, i.e., the contravariant components coincide with
the components defined in subsection 9.1.
In addition to the contravariant components we can also utilize
covariant spinor components defined as cpfl= (—1 . In virtue
of the formulas for the transformation of a spinor under rotation (9.1)
this definition guarantees the invariance of the scalar product ^<pn/i
of the two spinors cp1 and q>n under three-dimensional rotations.
The expression for <pjlM can also be written in the following form:
<Phm = a{r)Qjm {n), (11.1)
where Qjm (n) is a spinor whose contravariant components are given
by
(Q,,*(»)? = (H . 2 )
We refer to the quantity QJlM as a spinor spherical harmonic or a
spherical spinor.
The coefficients appearing in (11.2) differ from the corre­
sponding coefficients in Chapter I by the value s = They are normal­
ized in accordance with (8.4) and have the values in Table 6 .
T a b le 6

p 1 l+i I-l
1

1 / l + M+ % /l-M + i
2 X 2/+1 X 21+ 1
1 / / - M + i /l+M +i
2 X 2/ + 1 X 21+1
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 107

The spherical spinors form an orthonormal system of functions.


For different values of j, M, or I the functions are mutually orthogonal,
since they are eigenfunctions of self-conjugate operators belonging
to different eigenvalues. In virtue of (8.4) they are also normalized:
d o = <5;T(5ir< W , where do is an element of solid angle
in the direction n.

11.2. Wave Function o f a State of Definite Angular Momentum


In order to determine the radial dependence of <pj[M, i.e., the function
a(r), we make use of the Fourier transformation:

<p(r) = J <p(k)eikrdk.

States of definite angular momentum correspond to wave functions


<p(k) in momentum space which are eigenfunctions of the operators
for the square of the angular momentum and for one of its components.
The angular momentum operator has the same structure in the momen­
tum representation as in the coordinate representation, i.e., M = L + - |o ,
where a has the same meaning as before, while L = —i[AVk]. There­
fore, we can directly write cpjlM(k) = a(k)Qm (v), where v = k j k .
Since the electron energy has a definite value, this fixes the absolute
value of its momentum p = (e2 —w2)1/2. We, therefore, assume that
a(k) differs from zero only when k = p. On substituting <pm (k) into
the Fourier integral we obtain <pnM(r)= C J eipvrQjlM(y)dov, where C
is a constant.
We then make use of the expansion of exp(fpvr) in terms of
spherical harmonics (cf. (4.24)). This yields

<Ph m = C 2 S r ( P r ) Y rm '(» ) f
I'm'
On substituting into this equation the expression for the spherical
spinors ( 1 1 .2 ) we obtain

m'
YI'm' 0 0 J ^ iv ( v) ^ ilM^y) d ° v =

and, consequently, <pjlM = Cgfpr)QnM{n), where gt is the function


defined by formula (4.25). On comparing this expression with (11.1)
we see that a(r) = Cgfipr).
108 Q U A N T U M E L E C T R O D Y N A M IC S

We now obtain the second spinor %HM which appears in the electron
wave function y>jlM. On making use of the fact that in momentum space
the spinors %{k) and 99 (A) are related by the simple expression ( 1 0 . 1 )
we obtain
k A
Xhm ( * ) = a (* ) (• * ) . v = j '

On the other hand, Xhm, like is an eigenfunction of the operators


M 2 and Mz belonging to the same values of j and M. This means that
its angular dependence must be determined by a spinor spherical
harmonic. From this we can conclude that the product (ov)QjlM(y)
is a spherical spinor. Since in accordance with (11.2) QjlM contains the
spherical harmonic Ylm, then (cf. (4.9)) the product v f lm(v) contains
the spherical harmonics Yrm, where / '= / + 1 . Of these values only
the one value of /', viz.,
,
/ + 1
l= j-h
l'= 2 j- l =
/-I, I= j+ b
is compatible with the given value of j in accordance with the rules
for the addition of angular momenta.
Therefore avQjlM( y ) = cQjrM(y), where c is a constant. In order
to find this constant we choose the direction along the z-axis and make
use of the definition ( 1 1 .2 ) and the explicit expression for the coefficient
C{o 1/2/1 and for the spherical harmonic Yl0 (0). This yields
(v) = where l ' = 2 j —l. (11.3)
From this we obtain in the coordinate representation, in a manner
analogous to the transition from <pjlM(k) to <pjlM(r), the following expres­
sion for XjiJrY-

Xhm = - C - £ — g r (j,r) Q iru (n ) |/' = 2 j-l; n= j J '

The coefficient C is determined from the normalization condition


f\VjiM\2d r = 1. The integration is here taken over the normalization
volume which may be chosen in the form of a sphere of radius R (cf.
subsection 1.4). Since R may be taken to be arbitrarily large, the integra­
tion can be easily carried out if we make use of the asymptotic expression
for the functions g{ (4.26).
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 109

We give the expressions for the normalized" wave functions:

_ 1 / (e+m)p2d (r\
(2T ifi* \ Ye g'^ Q >m ( 7 )
( 11.4)
( e—m)p2d
XjlM
J> )^ ]/ g r(p r)G ,rJ ~ \*

where d = n/R , /' = 2j—l.


We note that the wave functions normalized per unit energy interval,
i.e., in accordance with the condition f y>*m y>cjmdr = d(s—e’), differ
from the functions (11.4) only by the replacement of 5 by e/p.
In momentum space the normalized wave functions have the fol­
lowing form:

tPiiMify (2 te) 3/2 ] / 2 e(e—m) 5 1,M\kJ


(11.5)
1
Xj l M( ^ )
( 2t t )3/2 " j / 2 E(eJrm)d Qjl'M1 ’
11.3. Parity o f a State
We note that <pjlM and %jm belong to different eigenvalues of the
operator L2: /(/-(-1) in the former case and /'(/'+ 1 ) in the latter case.
This means that the electron wave function y)jlM is not an eigenfunc­
tion of the orbital angular momentum operator. It can also be easily
verified directly that the operator L does not commute with the Hamil­
tonian operator H. We therefore conclude that the decomposition of
the total angular momentum into orbital and spin parts has a restricted
physical meaning in the case of the electron, just as in the case of the
photon. However, this decomposition becomes fully valid in the non-
relativistic approximation. As may be seen from (10.1) the ratio of
X to <p tends to zero as k -+ 0. This means that for low values of the
energy we can use a wave function consisting of only two components
(the remaining two components are small). In such a case I acquires
the meaning of orbital angular momentum. However, in the general
case the index / on the wave function serves only to designate two differ­
ent states of the electron with the same values of j and M.
110 QUANTUM ELECTRODYNAMICS

We can uniquely associate the quantum number I with the parity


of the electron state. In accordance with the law for the inversion of
spinors the inversion operator has the following form:
l = V,Plr 01.6)
where Ir is the operator for the reversal of sign of the space coordinates
I rV(r) = V( —»■)■
We apply the operator (11.6) to the wave function ipjlM

\X)IM/ \ XjlMI
Since <pjlM contains the spherical harmonic while XjiM contains
the spherical harmonic Yrm, and l rYlm= (—1)'Ylm, l rYVm = (—1)''
Yrm= ( - l ) ' +1 3'i.BI, we have
= / = ( - ! ) '. 01-7)
Thus, ipjlM is an eigenfunction of the inversion operator corresponding
to the eigenvalue 1 = (—1)'. We shall refer to the quantity I as the
parity of the state. Thus the quantum number I which for a given value
of j is uniquely related to I defines parity, rather than orbital angular
momentum, which does not have a rigorous meaning. Two states char­
acterized by different values of I for a given j differ in parity.

11.4 Expansion in Spherical Waves


We have constructed two different complete systems of electron
wave functions ipPtt and y nM. An arbitrary solution of the Dirac equa­
tion may be expanded in terms of either one of these systems: y)
— ^ ,anWn'> where n denotes the set of quantum numbers pn or jlM
n
(and also e). If ip and yjn are normalized, then we have an = f yip*dr. In
particular, the wave function of a state of definite momentum and
polarization can be expanded in terms of wave functions of states of
definite angular momentum and parity:

Vp = ] L a UM ty j l ’ Wp/i — A a \ PlM WjlM ^


HM jlM
or the converse expansion can be performed

W m = Z a%M)VPI>■
pp
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 111

It is evident that
uU M \up/j. ) •

These coefficients can be conveniently calculated with the aid of


wave functions in momentum space. On utilizing expressions (11.5)
and (10.9) and on going over from Fourier series to Fourier integrals
(cf. subsection 1.4) we obtain

where vv is the spinor amplitude of the plane wave defined in accordance


with (10.9'), while (50 is an element of solid angle. In the case of states
of definite component of spin along the direction of the momentum,
i.e., when w = xfl we denote by a(0 .
The quantity defines the angular distribution of the

electrons in a state of definite angular momentum. If we sum this quan­
tity over the polarizations, we obtain

°o p
For a given value of j this quantity does not depend on /, since

\Q,iu \2 = = l ^ w l 2-
The quantity \Qjm \2 can be evaluated by the method presented in
§ 8 ; this yields

\Q H m \2 = - ^ Z! a 2nP 2 n ( C 0 S & ) ,
n= 0

where

< ■ »»= (- 1 )2”- 1- " 2" ^ ) ’ 0 1-9)

W are the Racah coefficients. Explicit expressions for the coefficients


aZn are given below in Table 7.
By utilizing (11.8) we can easily obtain the expansion of a plane
wave in terms of spherical waves. If the amplitude of the plane wave
112 QUANTUM ELECTRODYNAMICS

is normalized (uu* = 1), while the polarization is determined by the


normalized spinor w (w*w = 1 ), then we have

ueipr = - l ^ r ' E r ______ > ( 1 1 .10)


* UM r j/'^ T ^ ^ )^ " )/
where

« = -> v = ^> /' = 2j—l.


r p

T able 7

3 5 7
2 2 2

1 3 1 3 5 1 3 5 7
a 2n |M ! 2 2 2 2 2 o 2 2 o

aa i i i i i i i l i

a 2 i - l 8 /7 2 /7 -1 0 /7 2 5 /2 1 195/28 -5 /3 -5 /2 1
ai 6 /7 -9 /7 3/7 8 1 /7 7 2 8 8 /7 7 6 3 /7 7 -1 1 7 /7 7
a* 2 5 /3 3 -5 /1 1 -5 /3 3 2 5 /3 3

For pr > 1 we can utilize the asymptotic expression for the function
g, (4.26). Then we have
m-\-e n . I In
~ 2 f~ Qm («)sin IPr ~
uelPr™ An— ^ t e M(v)H-)|
” r jlM I
- , y . e~ m n t \ ■/

(1 1 .1 1 )

12. Electron in an External Field

12.1. The Dirac Equation with an External Field


In considering the one-body theory exclusively, only a limited number
of problems can be formulated. These are problems in which the number
of particles does not change, and the interaction can be introduced by
means of the concept of an external field.
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 113

The equation of motion for an electron in a given external field


may be obtained in the same way as in nonrelativistic quantum me­
chanics. Let A be the four-potential of the external electromagnetic
field (A is the vector and Ag = —iA4 is the scalar potential). We shall
obtain the desired equation if in the Dirac equation we replace the
four-momentum operator p by p —eA, where e is the electron charge.
Evidently an equation of this type is relativistically invariant, since
the transformation properties of the quantities p and A are the same.
Thus, the Dirac equation, taking the external field into account,
has the form
(ip—ieA-\-m)y> = 0, (12.1)
where

We confine ourselves to the case when the external field is inde­


pendent of the time. Then stationary solutions such as
V(r, t ) = y 0( r ) e - imt

exist, where y>g(r) is an eigenfunction of the Hamiltonian operator, i.e.,


Hy>0 = coy)g, H = a.p-\-fim-\-eA0~ e ( tA . (12.2)
The general solution of (12.1) can be represented as a superposition
of wave functions of different frequencies co.
From the point of view of the values of the frequencies a signifi­
cant difference from the free electron case now arises. As we have
seen in § 1 0 , in the absence of an external field the frequencies form
a continuous spectrum with a gap in the range from -m to m. In the
presence of an external field, in addition to the continuous spectrum,
a discrete spectrum can also occur (as we shall see later), i.e., bound
states can occur in the interval - m < co < m .
We have interpreted the solutions with values of co > m as wave
functions of electron states, and solutions with co < —m, transformed
in accordance with (10.7), as wave functions of positron states.
Now the problem arises of the interpretation of the discrete spectrum.
We shall confine ourselves here to the practically important case
when the discrete spectrum is situated close to one of the boundaries
of the continuous spectrum. In this case we can, as before, separate
the frequencies into positive and negative ones, and interpret them
in accordance with the rules given in § 10. If the wave function ip =
114 QUANTUM ELECTRODYNAMICS

belongs, to negative frequencies, it is necessary to go over to the


charge-conjugate wave function tp(p) = Cy)(~K We derive the equation
satisfied by y>(p). On carrying out the same operations as in the case of
the transition from equation__
(9.12)
A
to A(9.13) and on noting that
.
A* = A,
A* = —Aif we obtain y>(—ip+ieA-j-m) = 0. On multiplying this
equation on the left by C and on utilizing (10.5), we finally obtain
(;ip-j-ieA-f-m)y(p> = 0. (12.3)
We see that equation (12.3) differs from (12.1) by a reversal of the
sign of the charge. Thus, instead of solutions of positive and negative
frequencies we can discuss solutions of only positive frequencies cor­
responding to the two signs of the charge. This is in accord with the
basic idea for the interpretation of negative frequencies presented in
§ 10.

12.2. Separation of Variables in a Central Field


We consider the problem of the motion of an electron in an electro­
static central field
A = 0, A0 = A0(r).
The Hamiltonian operator (12.2) has in this case the form
H - ap+ /3m + F(r),
where V (r)= eA0(r).
In view of the spherical symmetry of the field the angular momentum
operator and the inversion operator both commute with the Hamilton­
ian operator. Therefore, there exist states of definite energy, angular
momentum, and parity. As in § 11 we denote the corresponding wave
functions by
Vhm =

The eigenfunctions of the angular momentum and the inversion


operators were obtained in § 11. Therefore, we know the angular depend­
ence of the wave function, and we can write <pjlM and %jlM in the follow­
ing form:
<Pnm='g(r)Qn,M(n\
Xm= (12'4)
where V = 2j —l, n = r/r and QjlM is the spherical spinor defined by
formula (11.2). Our problem thus reduces to the determination of the
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 115

radial functions #(/•) and f(r). In order to do this we substitute (12.4)


into the equation for the eigenfunctions of the energy operator Hy
= ey), i.e., into the equations
(e—m —V)(p =apx, (e-Jr m —V )x — ap<p. (12.5)
We rearrange the right-hand sides of equations (12.5). On noting
that ap 9? = dgjdripn) Qjm-\-igap&jlM, and on making use of the fact
that (ari)QjlM= —QjVM we obtain

-(o p )^ jm = (®p) = (P«+ *[P”\o)Q jVU = - i l j + —La) QjVM,

where L = [rp] is the orbital angular momentum operator. The function


Q)VM is an eigenfunction of the operators L2, M 2 and ( |a ) 2 which corre­
sponds to the eigenvalues/'(/'+ 1 ), y(y+ l) and 3/4. Therefore we have
LoQjrM= (M 2 - L 2 - ( | a ) 2) ^ = 0 - ( y + l ) - / '( / '+ l ) - f } OjVM.
On introducing the following notation:
—(^+ 1 )> j= l+ b
( 12. 6)
« = T O '+ i) = I, j = l-b
and on noting that
—( /+ 2 ), j — 1+b
J U + i ) - / '( / '+ ! ) - !
^—1 , j = l~ b
we obtain
j t 'M — (1

dg x+ 1
a P9? -- Q jl'M
dr r
In an analogous manner we obtain
Id f * - l,\
a p X = J j

On substituting these expressions into (12.5) and on cancelling the


angular functions appearing on both sides of these equations, we obtain
the desired equations for the radial functions

— +(1 +x) — - ( £ - h m - V ) f = 0,
dr r
116 QUANTUM ELECTRODYNAMICS

or
^— + - ( r g ) - ( e + m - V ) { r f) = 0,
dr r
(12.7)
^ p - - - ( r f ) + ( e - m - V ) { r g ) = 0.
dr r
The functions rg and r f must in addition satisfy two boundary con­
ditions which guarantee the possibility of normalizing the wave func­
tion: they must remain finite as r -> 0 and r -* oo.
When V — 0 the solutions of the system (12.7) are given by the
radial functions found in subsection 1 1 . 2
i g = cg,(pr),

/= / , = 2y- ' '

12.3. Asymptotic Behavior o f the Radial Functions


We now obtain the asymptotic behavior of the functions g and /.
If we assume that the potential tends to zero sufficiently rapidly for
large values of r (V -*■ 0 as r -> oo) we obtain from (12.7) for r -*■ oo
d{rg)
—j f - —( s + m ) r f= 0 ,

d(rf)
- f r - + (e—m )rg = 0 .

The general solution of these equations is of the form


r g = Cie-^+qe*7),
( 12. 8)
rf = c l - '
\ \
|/ ^mLZ±
+e
e- »+q W
\
e
m+e
where A = (m2 —e2) 1/2 and C and q are constants. The constant C is
determined from the normalization condition for the wave function,
while q is obtained from the condition that the solutions of equations
(12.7) remain finite for r = 0. In particular, this quantity depends
on the energy q = q(e).
The nature of the behavior of the functions (12.8) is essentially dif­
ferent in the two cases e > m and e < m. If £ > m the coefficient in
the exponential is imaginary, and the condition at infinity is satisfied
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 117

independently of the value of e. The asymptotic expression of the radial


functions ( 1 2 .8 ) can then be given in the following form:

irg = i ' c V l i — s i n |p r - — + d\,


' ______ ' ' (12.9)
] / — - ! sin ( / > r - ■ £ + « ) ,

where p = — iX = ) /e2—m2 while Cl is a constant determined by the nor­


malization condition. The phase S depends on the nature of the external
field, and also on the energy and the momentum of the electron. In
the absence of an external field (5 = 0 , which may be easily seen on
comparing (12.9) with the asymptotic expression for g, (4.26).
If £ < m, then the quantity X is real, and therefore the second terms
in (12.8) increase exponentially. In this case the condition at infinity
may be satisfied only for such values of e for which q(e) = 0. The roots
of this equation determine the allowed energy values. Thus, as was
asserted earlier, we have a continuous spectrum for e > m, and a discrete
spectrum for e < m. The latter may be completely absent if q{e) has no
roots.

12.4. Behavior o f Energy Levels as Functions o f the Potential Well


Depth
We investigate the frequency spectrum in the special case of the
spherically symmetric potential well V = — V0, r < r 0, V = 0 , r > r0,
where VQ is a constant. On this simple example we shall demonstrate
certain general features of the behavior of the discrete spectrum.
On substituting these values of V into the system of equations (12.7),
and on eliminating r f from it, we obtain the following second-order
equation for the function W = rg:

-------- -{ ^ r J -W + [(e+ V oy - m * ] W = 0 , r < r 0,

d2W x (x + l)
---------- Ji---- w + (e 2- m 2) W = 0, r > r0.
The general solution of these equations has the following form:

W = ) / r { A2J\ x+1I2\(P r) + -®2-^U+ l/2|(^ r)}> r ^ r0>


118 QUANTUM ELECTRODYNAMICS

where / is a Bessel function, and N is a Neumann function


0 = |/(e + K0)2- m r , /?' = j/e 2—m2 .
The condition that IF should remain finite at r — 0 requires that = 0.
Instead of the condition that the function should remain finite as r -> oo
we introduce the condition W{R)— 0 where R > r0. This corresponds
to placing the system in a spherical box with impenetrable walls of
radius R. Such a change in the boundary condition has no fundamental
significance. However, practically it is convenient because it makes
the whole spectrum discrete with a very small distance of the order
of l/R between the frequencies in the range |co| > m. Due to the discrete
nature of the spectrum we can follow the changes in the position of
each level as V0 is varied.
From the boundary conditions it follows that

B = _ A J W l /21 (£'*)
2
The requirement of continuity of W(r) at r = r0 finally leads to the fol­
lowing formula:

— J \x+ l/2 \ (P R)Nlx+ml(P r0)]/|x+1^2| (/5r), r < r0,


JV(r) = A ]/ r
J\x + ll2 \(fir o) [ ^ | x + l / 2 | ( ^ -^) ^lx+l/2| ( /5 r )
- 4 +i/2|(/?'^)^Ix+i/2|(/5 >)], r > r0,
where A is a normalization constant.
The energy spectrum is determined from the following condition:
I WA I W1
W , r = r 0- o W
\ " / r = r 0+o
where Wx — rf. On expressing W1 in terms of W with the aid of equations
(12.7), and on utilizing asymptotic expressions for the functions con­
taining R In their arguments, we obtain for e < m

J',Aa)
4i(«)

where v = a — (3r0, a’ — /3’r0 and H is a Hankel function.


The roots of this equation may be found graphically in the general case.
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 119

We discuss the case / = 0, x = —1 , v = Then we have

and the last equation takes on the following form:

m
This equation can be easily solved in the limiting cases of a wide (mr0 > 1)
and a narrow (mr0 < 1) well. In the former case (mr0 > 1)

1 — a cot a =
1+- m
From this it may be seen that the first bound state e = m appears when
a = ts/2, i.e., for V0 = n ^ / Z m r just as in the corresponding nonrela-
tivistic problem. The energy of this level attains the value e = 0 for
n*
V0 = m 1 +
2m2r%
and the value e = —m for
1 TV
^o = 2 1 + -r
4 m2r,l ) m ■
In the latter case (mr0 < 1) we have
e= m for V0 = — —m,
ro

e= 0 for V0 — ,
r0
s = —m for V0 = ---4-m.
ro
We see that in both these limiting cases the width of the range of V0
within which the level shifts from m to —m is equal in order of magnitude
to 2m. This also holds in the general case.
Figure 3 shows schematically the dependence of the frequency spec­
trum on the potential depth V0 for a given well radius r0 (170). We see
that for V0 < F0(1) there are no bound states (|co| < m). The frequency
spectrum consists, just as in the case when there is no external field,
120 QUANTUM ELECTRODYNAMICS

of the two regions a>> m and co < —m to which we shall refer respectively
as the upper and the lower continuum. For V9 > F0(1) the lowest level
of the upper continuum attains a value less than m, i. e., there appears
one bound state; for V9 = V£2> a second bound state appears, etc.
The value of the frequency for each bound state diminishes continuously
as V0 increases, and even becomes negative for V9 > V 9. Nevertheless,
we may include these states among the electron states, since such a state

can be returned to the upper continuum by an adiabatic variation of


the external field. A difficulty arises when at a value V0 = V0c the level
crosses the boundary co = —m and merges with the lower continuum
which represents the set of positron states. The problem of the behavior
of an electron in a potential whose depth exceeds V0c cannot be solved
within the framework of the quantum mechanics of a single particle.
We shall return to this question later (cf. Subsection 18.2).

12.5. Electron in a Constant Homogeneous Magnetic Field


We now discuss the motion of an electron in a constant homogeneous
magnetic field H. If it is directed along the z-axis, then A0 = Ax
= Az = 0, Ay = Hx. We write the Dirac equation in the form
{a(p—eA)+j3m}y = eip or (e—m)rp = o(p—eA)x and (e+m)%
= a ( p —eA)<p. By eliminating £ from these equations we obtain a sec­
ond order equation for <p:
(p2-j-e2H 2x2—eH(ax+2xj7y)}^ = (e2—m2)<p. ( 12. 10)
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 121

The solution of this equation may be sought in the form cp = eHpyy+p*z)f(x )


where f ( x ) satisfies the equation

{— + (eH x-P y)2- eHa^ f = 0 2 -™ 2 - P * ) /


or
( - ( d * /d ? ) + ? - a z) f = af.
where

eH
The spinor function / may be so chosen that it is an eigenfunction
of the operator az: azf = pf, where p = ± 1. This means that /=
l0
for p = 1 and / = j for p = — 1. For f^(p = ± 1) we obtain the
following equation = —(a + p )f , the solution of which
is given by
f ^ = C e - ^ H n(i), (1 2 .1 1 )
where //„(£) is an Hermite polynomial, C is a normalizing constant and
n = (a-\-p —1/2) = 0, 1, 2 ,... From this we obtain for the electron
energy e2 = m2jrp2 z JreH(2n—p Jr \).
The function % may be expressed in terms of cp as follows:

^ L p M e H ) v 4 J L L. ( 12. 12)

We give the expressions for the normalized wave functions xp^:


I (e+ni) vin) \
exp i(pyy + p zz) 0

V+i = (n)
(Ly L z ) 1/ 2 [2(e2 + ws)]1/2
\ —i(eH2n)ll2v{n~1) /
(12.13)
o/ \
_ exp i(pyy + p zz) (e+ w )u(n)
W~x (Ly Lz)ll2[2(e2+me)YI2 i[2eH(n-\-i)]ll2v{n+1)
\ -pzv / (n )

where L„ and L, are normalization lengths along the y-and z-axes, and

u —e [ 7ri/4 2»/*(„ !)i/2 J


122 QUANTUM ELECTRODYNAMICS

§ 13. Motion of an Electron in the Field of a Nucleus

13.1. Solution o f the Radial Equations for the Coulomb Field


The most important application of the Dirac equation is the investi­
gation of the states of an electron in the field of a nucleus. The latter
is not a strictly central field. The deviation from spherical symmetry
is associated with the fact that in the general case nuclei possess electric
quadrupole, magnetic dipole and also higher multipole moments. If we
neglect these effects, which give rise to the hyperfine structure of electron
levels, and assume that the nuclear field is spherically symmetrical,
then we can discuss electron states with definite values of energy, angular
momentum and parity. In this case the wave function is of the form
(12.4) and the problem reduces to the solution of the equations for
the radial functions (12.7).
At large distances the field of the nucleus is a Coulomb field, i.e.,
V = — Za/r where r > r0 and where —Ze is the nuclear charge, a
= e2/4n, and r0 is the nuclear radius. We shall at first discuss this problem
in the approximation when the finite nuclear dimensions may be neglect­
ed, and we shall assume that the above expression for V holds over
all space.
In accordance with the nature of the asymptotic behavior of the
radial wave functions (12.8) we seek the solution of equations (12.7)
in the following form:

Substitution of these expressions leads to X = \ m 2—e2 and the


following equations for Fl and F2:

dF}
do
(13.1)
dF2
dq
where q= 2Xr.
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 123

The solution of the system of equations (13.1) for small values of


q is of the form , and F2 = a2qv, and a2 are related by
the following equations:

= 0 ,
4 + z “ -x
m
x —Za = o,
T
from which it follows that y = \/x2—Z 2a2. The sign of the square root
must be chosen positive in order to satisfy the condition that the wave
function should remain finite at q = 0 .
If we eliminate one of the functions, for example Fls from equations
(13.1), we obtain a second order equation for q~vF2 of the following form :
d2

where
b= 2 y+l, y —Za
A'
As is well known, the solution of the last equation which is finite at
q = 0 is given by the confluent hypergeometric function

m v r ( a + n) gn
F(a, b;e) =
v ' n^= 0 r ( b + rij nT
F{a)
( r is the gamma-function). Thus we have

F2 = CQYF \ y — Za~j-> 2 y + l ; e j > (13.2)

where c is a constant. By using the recurrence formula for the hypergeo­


metric functions,
q 4 - F ( ci, b ; q) = a[F(a+1, b; o)—F(a, b ; q)],
dQ
we can obtain F1 from (13.1):
124 QUANTUM ELECTRODYNAMICS

In the case e < m these solutions, as we have seen earlier, have


meaning only for discrete values of e. In order to find them we use the
asymptotic expression for the hypergeometric function
m r(b) h
F(a, b; q) e~ ina _\-L0a~beQ e > i, (13.3)
r(b-a) *
from which the following conditions for the absence of exponentially
increasing terms in Fx and F2 can be obtained:
1 „ 1
-= 0 , = 0. (13.3')
/ 'I y —Za~
Z ar f r Z aT
If (ZafX)m—x 9 ^ 0, which always holds for x < 0 , then the second
condition coincides with the first one. Since the poles of the y function
are given by negative integers and zero, we have y —Zae/X= —nr,
where nT is a non-negative integer. From this it follows that

| / l - i [ Z 2«V (7 i o 2] '
If x = Zam/A, then y = Zae/A and nT= 0. At the same time the second
condition (13.3') is not satisfied. Thus,
0 , 1 , 2 , ..., x < 0,
, 2 ,...,1 x > 0.
Formula (13.4) gives the fine structure of the levels of a hydrogen-like
atom. The lowest level corresponds to the values nT= 0, %= —1 and
is equal to e = m(l —Z 2a2y>2. For Za < 1 the expansion of the expres­
sion (13.4) in terms of this parameter to terms of order (Za ) 4 yields
the following expression for (e—m)/m:

where « = «r+ \x\.


The quantum number n coincides with the principal quantum num­
ber of the nonrelativistic theory. The first term in (13.4') yields Balmer’s
formula. Formula (13.4') can also be written in the form

where ae = Ijma is the Bohr radius.


R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 125

A characteristic feature of the fine structure formula is the fact


that it contains only the absolute value of x. Therefore, all the energy
levels, with the exception of the lowest one, are twofold degenerate.
In particular, the two levels 2j 1 /2 («r = l , x = — 1 , / = 0, j = |,
n = 2 ) and 2 plj2 (nr = 1 , x = 1 , / = 1 , j = n — 2 ) coincide.
Later (cf. § 53) we shall see that formula (13.4) based on the use
of the concept of a given external field is not exact. Corrections to this
formula remove the degeneracy with respect to x. The order of magni­
tude of these corrections for smalj values of Z is equal to (Za)4 a. From
this it may be seen that, for example, the use of formula (13.4) for hydro­
gen, attempting an accuracy which exceeds the accuracy of the expansion
(13.4'), is meaningless.
We now give the final expressions for the radial wave functions of
the discrete spectrum:

(13.5)

/ r ( 2 y + n r+ 1 )
1 / " I -(elm ) j 2 Z _ \ 313
r ( 2 y + l ) \ / n T\ X 4N{N—x) \N a 0J
X e
)!
(13.5')

where

N = ]/n2—2nT( \x \—y), a° = — «r+ \x\.

y = ,V - Z % * ) . — = 137 *
a = 4n

They are normalized in accordance with the condition J ( / 2 + g 2)r2</r = 1 .


126 QUANTUM ELECTRODYNAMICS

13.2. Wave Functions for the Continuous Spectrum


We now proceed to investigate the continuous spectrum (e > m)\
in this case I is a purely imaginary quantity, X = ip = i(e2—m2)112.
It is now convenient to write the formulas for Fx and F2 in the fol­
lowing form:
Fx = ceiS(y+ iv)F (y+ \+ iv, 2y + l ; lipr) (2pr)y,
F2 = ce~^(y—iv) F(y+ iv, 2y + 1 ; lipr) (Ip r y ,
where
x —iZa[m/p]
—y + iZa[e/p] ’
v is the electron velocity (f is real).
The hypergeometric functions appearing in (13.6) are related by
a simple expression which can be obtained by utilizing the integral
representation for the hypergeometric function:

m 4 d+ 2 )
F(a, b ; x ) = ( l + z ^ - H l —z) b-a-l, dz.
r ( a ) r ( b - d ) 2 b- 1
-i
If x is an imaginary quantity and 2R eo+ l = b, then
F(a, b; x) = exF *(a+ 1, b; x).
Upon utilizing this equation and the expressions for Fx and F2, we
obtain the radial functions in the following form:

irg = c ]./1 +(e/m) (2pry{e~ipr+ii(y-\-iv)F(yJr 1+/V, 2 y + 1 ; lipr)


+ eipr- it ( y —iv)F*(y-ir l+ iv, 2 y + l ; 2ipr)),
______ 03.7)
r f — c )/(e/m) —1 {2pry{e~,pr+^ { y Jr iv)F(y-\-\Jr ivy 2y + l; lipr)
-~eipT- i^(y—iv)F*{yJr \ Jriv, 2 y + l ; lipr)).

Now we obtain the asymptotic expressions for the functions (13.7).


To do this we utilize formula (13.3). Since now o = lipr is a purely
imaginary quantity, the nature of the falling off of the hypergeometric
function is determined not by an exponential function, but by the
power-series factor in (13.3). In our case we have Rea = y -|-l,b
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 127

= 2 y + l and the first term falls off faster than the second one, so
that we can retain only the second term. Thus we have
r ( 2y+l) 1
F (y + l + zV, 2 y + l ; 2 ipr) -( 2 ipr)iveiiPT
/" (y + l + zV) (2 ipry
A 2 y + 1 ) 1 — i | 2 p r - y + vl n 2 p r j

(y + zV) / ’(y + iv) (2 pry Pr > 1 .

On substituting this expression into (13.7) we obtain

A 2 y + 1)
irg = c 1 / H — sin [pr—~ + <5-f-v ln 2 pr
\ m \r{y+iv)\ )•
_____ ’ (13.8)

r/= c ] / i - - 1 p w ‘)T e" T’ r " ‘ si^ [ p r - ^ + i + r 1*%>') •


pr > 1 ,
where
<5 = £—a rg /" ( y + z y ) - y ( y —/—1 ),

v= — , y = i / ( y + | ) 2 - z 2 a^ r= 2 j-i.

Formula (13.8) agrees with the general asymptotic expression for


the radial functions (12.9). In this case the phase turns out not to be
constant, but to depend logarithmically on r, which is a characteristic
feature of the Coulomb field associated with its slow falling off with
distance.
For the determination of the constant c we shall make use of the
normalizing condition in the form
R
j (l?l* + l/l* )r> rfr= 1.
0

Since the integrand does not fall off as r - y o o , while the radius of
the normalizing sphere R is arbitrarily large, we can substitute into this
integral the asymptotic expressions for the radial functions (13.8);
the integral in this case turns out to be an elementary one, and we obtain
71

., e * V\r(y+iv)\ /~m (13.8')


C ' r ( 2y+l) X SR'
128 QUANTUM ELECTRODYNAMICS

The phase factor il in this expression has been chosen so that at Z = 0


the wave functions should coincide with the free electron wave functions
in the form (11.4).
If the wave functions are normalized per unit energy interval, then
R should be replaced by Tipje in (13.8').

13.3. Isotopic Level Shift


The results obtained so far have been based on the use of Coulomb’s
law for all values of r. In those cases when the value of the electron
wave function in the neighborhood of the nucleus is important, it is
necessary to take into account the fact that the expression for the
potential V is valid only for r > re (r„ is the nuclear radius).
An example of an effect due to the finite nuclear size is the isotopic
level shift in which a difference becomes noticeable between the electron
energy levels in the field of nuclei having the same charge, but different
radii (34), (159). We shall show that the isotopic shift may be directly
expressed in terms of the wave function at r = r%(186).
Let e denote the value of the electron energy in the Coulomb field,
and e-\-s1 the shifted value of the energy in the field of a nucleus of
finite size. We denote the radial wave functions in the former case
by / e and g%, and in the latter case by / and g. We discuss the case
x = — l. The functions / 0 and g 0 in this case satisfy equations (12.7):

-jr (rg«)~ —(rg9) = ( e + m ~ V ) (r/e) ,

Yr (rfo) + — (rf%
) = ( m + V - e ) (rg0),

where V = —Za/r. For r > r9 the functions and g satisfy the equations

^ ( r s ) - y ( r g ) = ( e + m - V ) r f+ e 1rf,

^ 0 / ) + j ( r f ) = (m + V - e ) (rg) —e, rg.

From the equations for / 0, g9 and f g it is simple to obtain the relation


(d/dr) (gf0r2—g0f r 2) = exr2(g9g + f0f ) . We integrate it with respect to r
between the limits r 0 and oo. On assuming that the functions /„, g0 differ
from f , g only within a small region near the nucleus, and that this
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 129

region makes a small contribution to the normalization integral, we


can approximate that
o o oo

/ r2(g0g + f0f ) d r ^ f ( g \+ fl) r 2d r = 1 .


ro 0

Thus we have
£i = r\ [go 0 o)/<>o)- g Oo) f 0 Oo)] • (13.9)
The isotopic shift is equal to the difference in the values of expression
(13.9) for two different values of the nuclear radius r9 corresponding
to two isotopes.

13.4 General Investigation of the Effect of Finite Nuclear Size


The fact that the nucleus is of finite size may also play a significant
role for the states of the continuous spectrum in those cases when
the effect is due to the behavior of the electron wave function in the neigh­
borhood of the nucleus. Such cases occur, for example, in /? decay (181).
In order to estimate the distortion of the wave function we return
to the equations for the radial functions (12.7) in the general case of
an arbitrary central field. The nature of the behavior of the wave functions
for small r may be determined in the following manner (166). We
construct the formal solution of each of the two equations (12.7) by
treating the last terms as inhomogeneous ones. This yields

rg = c1r~x-\-r~xj xx(e-\-m—V )f{x)dx,


(13.10)
r f = c2rx—rxJ x~x(e—m — V)g(x)dx.
o
From the boundary condition at r — 0 we obtain cx = 0 where x > 0,
c2 — 0 where x < 0 , and therefore for small r,
rg = q r 1,
r (x < 0 ),
rf = - q r ‘ MJ x 2,x,(e—m — V)dx
o
(13.11)
rf = c2r*,
(x > 0 ).
rg = c2r~xJ x 2x(e-\-m — V)dx
0
130 QUANTUM ELECTRODYNAMICS

We see that for x < 0 the dependence of the function g on r is deter­


mined by the magnitude of x, i.e., by the electron angular momentum,
and in the first approximation is independent of the field. But the function
/ is essentially determined by the field, and, therefore, it is affected
by the difference between the actual field and the Coulomb field. If
we can assume that for small r V > m (this holds for not very high
energies and for nuclei with not too small a value of Z), then we have
r
r f = c1/ " |x| f x2|x| V(x)dx where x < 0. A similar situation occurs when
o
x > 0 with the functions / and g interchanging their roles.
In order that the expansion (13.7) be meaningful it is necessary
that the integrals obtained by substituting (13.11) into (13.10) should
exist. This will be true for all x if lim rV = 0. Thus, this expansion is
r —>0
not applicable to the pure Coulomb field. However, the field of a nucleus
of finite size remains finite (or vanishes) at r = 0 , and therefore we
can utilize expansion (13.11).
We shall not give in detail the solutions of the equations for the
radial functions taking finite nuclear size into account. We shall merely
point out the main features of this problem.
First of all, it should be noted that the function V(r) is not exactly
defined for r < r0, and the nuclear boundary itself is to a certain extent
arbitrary. We shall obtain a qualitatively correct picture if we assume
that the charge density is constant within a sphere of radius r0. In such
a case we have
Za I 3
V for r O 0-

Another limiting case is the field of a charge distributed over the surface
of a sphere of radius r0. In this case
V — —Zajr0— V2 for r < r0 .
The fields V1 and V2 lead to results for the distortion of the wave function
near r0 for r > r0 which differ little from each other. The essential
feature here is not the exact shape of the potential for r < r 0, but the
fact that it does not increase in accordance with Coulomb’s law as r -*■ 0.
The solutions of the equations for the radial functions in the case
of the field V2 (constant potential) coincide with the solutions for the
free electron if e is replaced by e+ V 2. In practice it is sufficient to restrict
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 131

ourselves to the first terms in the expansions of/ and g in terms of r.


For this purpose we can use formula (13.11) both in the case of the
field V1, and the field V2. This determines the solution in the range
r < r0 up to a constant factor.
The expression for the potential V = —Za/r is valid in the region
r > r0. Therefore solutions exist in the form (13.1), (13.2). However,
since now the point r = 0 no longer lies within the region where
the solutions of equations (12.7) hold, the corresponding boundary
condition no longer applies. Therefore, in formula (13.2) we must take
into account the double sign of y. We shall, as before, utilize the notation
y = + ] /x 2—Z 2a?, then F2 = F2{y)-\-qF2{—y), where F2(y) is given
by formula (13.2), and q is a constant.
Similarly F1= F1(y)Jr qF1(—y) , where expression(13.2) holds for Fi(y).
Thus, the functions/ and g depend on the two constants c and q with
the former appearing as a common factor. Therefore, their ratio contains
only the constant q which may be determined from the condition of
continuity at r = r0 and may be expressed as (//g ) r = r o + 0 = (//g )r=ro_0-
The finite size of nuclei alters in a radical way the question of the
existence of bound states for large values of nuclear charge. According
to (13.4) when Za = 1 the value of the lowest electron energy level
emin(x = — l , n r = 0 ) becomes equal to zero.
For Za > 1 formula (13.4) leads to imaginary values, and this
indicates that corresponding states do not exist. However, this conclusion
is valid only for the Coulomb field and is inapplicable in the case of
the actual field of the nucleus. Taking finite nuclear dimensions into
account leads to the possibility of bound states occurring also for Za > 1
(156). As Z increases, the value of emln diminishes and, just as in the
case of the example of the potential well discussed in § 1 2 , for a certain
value of Z = Zc it attains the value -2m. The value of Z amounts to
Z c = 200 for r0 = 1.2 • 10~ 12 cm, and Zc = 175 for r0 = 8 • 10“ 13 cm.

§ 14. Electron Scattering

14.1 Spinor Scattering Amplitude


The problem of the scattering of particles by an external field may
be solved in relativistic quantum mechanics by the same methods as
in nonrelativistic quantum mechanics (cf., for example, (119) and (139)).
132 QUANTUM ELECTRODYNAMICS

In order to obtain the differential scattering cross section in the


general case of an arbitrary stationary external field, it is necessary to
find the asymptotic solution of the Dirac equation which represents
a superposition of a plane and an outgoing spherical wave:
eipT
ip ph ueipz + G(n) — •

Here u is a constant bispinor which defines the polarization of the incident


wave (the z-axis coincides with the direction of the primary beam);
G is the amplitude of the scattered wave and, like u, is a bispinor. It
depends on the direction of scattering n = r/r. The differential scattering
cross section da is equal to
da _ r2nj(r) G*anG
do j0 u*azu
where j(r) and j 9 are the flux densities of the scattered and the incident
waves. The amplitude G(n) is equivalent to the amplitude of a plane
wave propagating in the direction n. Therefore da/do = G*G/u*u or
if we represent u and G in the form

then we have da/do = |F | 2/|vv|2. Thus, in the case of the scattering


problem it is sufficient to investigate only a part of the wave function
, viz., the spinor 9 ?. If we write the asymptotic expression
for <p in the form
eipr
cp = weipz + F----- (14.1)

and if the spinor w is normalized (h>*vv = 1), then we have

d a = \ F \ 2do. (14.2)

In order to obtain the solution of the Dirac equation in the form


(14.1) we use the asymptotic expressions for the wave functions found
in § 12. Special solutions corresponding to definite energy, angular
momentum and parity, are determined by formulas (12.4) and (12.9).
The general solution corresponding to the definite energy e = (p2-{-m2) 1/2
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 133

may evidently be written in the form of a superposition of these so­


lutions:
^ s in O + ( 5 ;,- /W 2 ) r
a ]lM ^ j l M \ H) n= (14.3)
r - EilM r

In order to bring expression (14.3) into the form (14.1) we compare


it with the asymptotic expansion of a plane wave. In accordance with
( 1 1 . 1 1 ) we have

Wt
P jlM

where n0 = p/p. It may be easily seen that expressions (14.3) and (14.1)
will coincide if ajlM = (4n/p)ile*>1 . The amplitude of the
scattered wave turns out to be equal to

F= ~ (M.4)
lP ilM

We also give expressions, useful in a number of applications, for


the exact solutions of the Dirac equations which have the asymptotic
form & ue‘Pr-\-G(H)eipT/ r . On replacing in (14.3) the asymptotic
value of the radial functions by the exact one it may be easily shown
that

(14.5)
P m

where # i e = pjp, n = r/r, V = 2 j —l, gn and f n are radial functions


satisfying equations (12.7) and normalized in such a way that in the
asymptotic expressions (12.9) the coefficient cx is equal to unity.
In an analogous manner we can construct the solutions

(In)
= t £ * " ( f i S - W " 1) w ■ <1 4 5 >

which asymptotically go over into a superposition of a plane and an


incoming spherical wave
g-ipr
y)(ln) an uelpr+ G(*)------ •
134 QUANTUM ELECTRODYNAMICS

Solutions (14.5) and (14.5') for different values of p form two complete
orthogonal systems of functions:

in terms of which the general solution of the Dirac equation may be


expanded. In the absence of an external field this expansion goes over
into the expansion in terms of plane waves.

14.2. Expression for the Cross Section in Terms o f Phases


The relation between the amplitudes F and w can be conveniently
represented in the form
F = R(ri)w. (14.6)

Here R(n) is a two-dimensional matrix which must obviously have


the structure i?(/i) = , where a is a scalar, and a 2 is an
axial vector. But the only axial vector which can be introduced in
a scattering problem is the vector [n0«]. Therefore, R may be written in
the form
R = a(f))—ib(p)xa = a(fF)—ib(&) (oy cos<p—oy sin 99), (14.6')

where t is a unit vector perpendicular to the scattering plane

= [»0 »]
T sin # ’
# and 9) are spherical polar coordinates of the vector n in the system
in which the vector n0 is taken as the polar axis. The functions a{d)
and b(f) may be obtained by comparing (14.6) with (14.4). On utilizing
the explicit form of the spherical spinors Qjm and the matrices cri we
obtain

(14.7)
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 135

where
^ — -F (j~h%), ( j — l ih i) ,
Pt and P ;(1) are Legendre polynomials:

*>i = Y, o,

dPt
rp i<i) —
— sin# d cos# ’

We note that the function b (#) contains the factor sin #. Therefore,
the matrix R does not contain the spin matrices a when # = 0.
We also introduce another form in which the matrix R can be
expressed:
R — A -\-B{an) (<rn0). (14.6")

Comparing with (14.6') we can easily see that


a = A-\-Bcos$,
b - i?sin#.

The functions A and B have the following expansions

a - ' a
p d cos# £ f+Pi+i (cos # )+ /,-# ,_ ! (COS#)

(14.7')
p d cos# ^ ( f n ~ f ' ) P i ( cos#)>
d

where

$i± = y = / ± y |'

We now determine the total scattering cross section. According


to (14.2) and (14.4) we have

a= f \F\l do = 1 ^
J P )l M
136 Q U A N T U M E L E C T R O D Y N A M IC S

Since YlM(na) = \ / (21+1/4jt)dm , we have

Q fiM w = £ MtoC&x Ylm(na) = w (AOCffi*


Am

But in accordance with the table in subsection 11.1 we have

Therefore, 4 7 r ^ | |2 = |x| and the cross section assumes the


following form:

(14.8)

If for all values of I the quantities 8X coincide for x = —(/+1) and


x = /, then formula (14.8) goes over into the nonrelativistic formula
for the scattering cross section

£ (2 /+ l)sin 2 <5,.

14.3. Polarization and Azimuthal Asymmetry


An essential feature of electron scattering is its azimuthal asymmetry,
i.e., the dependence of the differential scattering cross section on the
angle y. According to (14.6) the angle y is contained in the matrix R
which determines the relationship between the amplitudes of the incident
and the scattered waves, with the coefficients of cos y and sin y con­
taining electron spin operators. Therefore, the azimuthal scattering
asymmetry is closely connected with the nature of the polarization
of the incident wave.
The differential scattering cross section can be easily related to
the state of polarization of the incident wave if we use the polarization
density matrix q defined in accordance with (10.18). It follows from
(14.2) and (14.6) that

^ = s Pf > ™ . (14.9)
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 137

On substituting into this equation the general expression (10.19)


£ = i 0 + ?*)> where ££ is the average value of the spin in the rest
system of the electron, and expression (14.6') for R, we obtain

= Q(P) (l + 4 (0£ t ) = 0(0) [1 + 4 (0 ) (£„cosp—C^sinp)], (14.10)

where

<m = w + w , m = - (w -n )

It may be seen from (14.10) that the scattering system is an “analyz­


er” of the polarization. A measurement of the azimuthal asymmetry
enables us to determine the component of £ perpendicular to the scat­
tering plane. On choosing for (p the values 0 and n or +tt/2, we obtain
from (14.10)

a ,A\r = daS^_i ° 1 ~ da(P>n) } = da(®, —(7i12 ))- d a ( &, + 0 + 2 ))_


V 0 ) + da(d, n) ’ 1 da{d, - (jr/2 >)+ da{$, + ( + 2 )j'

The longitudinal component of the polarization Cz cannot be determined


by single scattering.
If 4(0) — 0, then the scattering does not have any azimuthal asym­
metry. This occurs, in particular, in the nonrelativistic approximation
when scattering by a central field gives rise to the same phases dn for
a given value of / and different values of j. In such a case in accordance
with (14.7) 6 = 0 and, consequently, A = 0.
Moreover, there is no azimuthal asymmetry in the Born approxi­
mation. Indeed, in this approximation exp(2/6x) —1 ^ 2idx, a(&) and
6(0) are real, and according to (14.11) 4 = 0.
We note that A -> 0 when 0 -> 0, since 4(0), like 6(0), contains
the factor sin 0 .
Because the matrix R depends on a, the state of polarization of
the electron is altered by scattering. The polarization density matrix
q' of the scattered electron can be defined as = FaF^*/\F\i where
Fa are the components of the spinor F. We obtain from (14.6)

R qR+ (14.12)
Q Sp qR +R
138 QUANTUM ELECTRODYNAMICS

Therefore, the polarization vector of the scattered electron is given by


_ Sp£>/?+oi?
(14.13)
^ " S p qR +R
If the incident electron is unpolarized, then it follows from this that

r = « t , t = (14.14)
sin#
i.e., the scattered electron is polarized in the direction perpendicular
to the scattering plane, while its degree of polarization is equal to A.
Thus, the scattering system acts as a “polarizer” .
On utilizing (14.14) and (14.10) we conclude that secondary
scattering of unpolarized electrons will exhibit azimuthal asymmetry.
The differential cross section for secondary scattering is given by

— - = Q (V)(l+A(ff)A@ ’)'ZT'), (14.15)

where cos#' = (nit'), x — [/i/i']/sin#', it' is the unit vector in the


direction of the secondary scattering.
In the general case, if initial polarization is present we have
_ (\a\2—\b\2)'E>-{-i(ab* —a*b)x-lr (ab*-lra*b)[T'E>]-\-2\b\2('^z)T
** \a\2+ \b\2+i(ab*-a*b)% x ‘

14.4. Scattering by a Coulomb Field


The foregoing general theory of scattering by a central field is also
applicable to the case of the Coulomb field. However, since (cf. § 13)
the phase in the asymptotic expression for the wave function depends
logarithmically on r, the solution of the Dirac equation which describes
scattering should be sought in a form which takes into account the
distortion of the plane wave at infinity, viz., in the form (cf. 13.8)
ip r — i f l n 2p r
y = WeiPZ- l>’lnp(T-Z)jr P _ _______ _

Formula (14.4) remains valid and, in accordance with (13.8), we have

= f-a rg /X y + j V ) - ~ ( y - / - l ) , v = y = ]/x 2 —Z 2a2. (14.16)


2 V
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 139

The series (14.7) for the functions a(ft) and b(ft) in which 6X is defined
by formula (14.16) cannot be summed analytically. For low values
of Z it is possible to expand these functions in terms of the parameter
Za. In this case the scattering problem can be solved by perturbation
theory. This solution will be obtained in §§ 26, 51. Here we shall
only state the results (138), (133).
In the lowest order approximation (to terms of order (Za)2) we have

and J w = ° - (,4 -17>


This formula differs from Rutherford’s formula by the last factor in
(14.17). In the next approximation (to terms of order (Za)3) an azimuthal
asymmetry appears, and also a difference between the scattering of
electrons and positrons:
Zct \ ^ 1
(
/ ^ ) sinHW 2 ) ^
2 sin2 W 2) + 7tZa sin(tf/2 ) ( l-s in (tf/ 2 ))],
~ V 2

(14.18)
Ar%\— 1 7 v ( \ ~ v 2)112 sin3(ft/2) , 1

( ) Z“ l —v2 sin2 (#/ 2 ) ' cos(0 / 2 ) ln sin(tf/2 ) (14-19)


(in the case of positron scattering Za should be taken as negative).
In the general case of arbitrary Z the determination of the cross
section requires numerical calculation. Figure 4 shows some results
obtained in this manner (50). The curves show the ratio of the differ­
ential scattering cross section for unpolarized electrons to the value
obtained from Rutherford’s formula daR, i.e.,
_____ 2 0 ___
d<yR [Za/lpv sin2 (#/2 ) ) 2
as a function of the scattering angle ft for different values of the energy
and of Z. Figure 5 shows similar results in the case of positrons (Z < 0).
Figure 6 gives the curves of the dependence of A on the energy for given
ft and Z in the case of electrons (179), while Fig. 7 gives the same infor­
mation for positrons (8 ). We note that the signs of A shown on the
graphs are different for electrons and positrons.
At very high electron energies it is necessary to take finite nuclear
size into account. The change in the phases due to it may be taken into
account by the method presented in § 12. However, we shall not repro­
duce the corresponding results here (164), (61), (2).
140 QUANTUM ELECTRODYNAMICS

ad

(a)

Fig. 4
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 141

dd

(a)

Fig. 5.
142 QUANTUM ELECTRODYNAMICS

14.5. Small Angle Scattering


As has already been noted earlier (cf. (14.5)), the solution of the
Dirac equation for an electron in a Coulomb field which has the asymp­
totic form of a plane and an outgoing spherical wave

W\ eiPT
(^ = ue»*+G(n)— ,

0.1 0.3 0.5 0.7 0.9 0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
^kin.Mev
Fig. 6. Fig. 7.

can be expressed in the form of the following series:

x{(y+/v)e~i(pr~7')fr(y-f-l + /v, 2 y + l; 2/pr)+ compl. conj.) (14.20)

and of an analogous series for %, neither of which can be summed analyt­


ically.
However, it is possible to obtain an approximate solution ipn) having
the following properties: if yia) is expanded in a series in terms of
spherical waves jx), then the terms of this series will differ
i
significantly from the terms of the exact series y) = £ yi, only for low
p i
values of /. This approximate solution is useful for the investigation
of those processes in which the principal role is played by large values
of /, for example, the case of small angle scattering.
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 143

The function ip{1) can be obtained in the following manner (69), (22).
We transform the Dirac equation

e-----------/3m+z'aV \ip = 0
r

into a second-order equation. In order to do this we apply to the equa­


tion given above the operator (e + Z a /r+ ^ m —zaV):

If in this equation we omit the right-hand side, then we obtain an equa­


tion which coincides with the nonrelativistic Schrodinger equation for
a Coulomb field. It is well known that its solution having the required
asymptotic form is given by xp = cueipzF, where F is the confluent
hypergeometric function F = F ( iv , I; ipr—ipr), v = Z a / v , v is the
electron velocity, u is a unit bispinor, c is a constant.
On this basis it is useful to seek the solution of equation (14.21)
in the form tp = ceipz(\-\-Q)uF, where Q is a particular operator. In
this form it is possible to take partially into account the first term on
the right-hand side of (14.21). The second term may be neglected since
it contains 1 /r2, and therefore it may be expected that for large values
of / and for small angles small values of r are not important.
It may be shown that the function

(14.22)

actually possesses the required properties. In order to prove this the


following expansion should be used (43):

eipzF = ^ (2/4- 1)

Gt = F(l-1-1—zV, 2/+2; —2ipr),

the function y>{1) should be expanded by means of it, and then this expan­
sion should be compared with the series (14.20). It then turns out (22)
that for (Za ) 2 / / 2 < 1 the terms of the two series coincide.
144 Q U A N T U M E L E C T R O D Y N A M IC S

eipr
From the asymptotic form ^ (1) ~ ueivz+f{ri)u—— it follows that

/("> = 2^ s l ° W 2 )(1+ ^ (a " ~ a ' )) '


For small angles the second term may be neglected and we have
f~ 2Z a/pv-& 2. In this case the cross section coincides with formula
(14.17) for & < 1.

§ 15. Nonrelativistic Approximation

15.1. Transition to the Pauli Equation


If the electron energy differs but little from the rest energy, i.e.,
|e—m \ < m, then, as we have seen earlier, one of the two spinors com­
posing the electron wave function is much smaller than the other: x ■< <P-
This enables us to obtain an approximate equation which contains
only the spinor <p by means of a formal expansion of the wave functions
in a series in powers of 1 jc, where c is the velocity of light.
We start with the Dirac equation for an electron in an external
field which we write in the form

c a |p — - +/5wc 2 +E/I 0 jYJ.

Since in nonrelativistic theory the energy is taken to mean the difference


between the total energy and the rest energy it is convenient to trans­
form the wave function by introducing instead of y the function rp'\
xp — xp'e~imcit.
On substituting this expression into the Dirac equation

i^jt~ = j c a |p —~ / l J + (p — l)mc2+ e A ^ y '

and by representing as usual y>’ in the form y)’ = we obtain the


following system of equations:
d<p I e \

05.1)
i~ fc= ca \ V - ^ A j <p+ eA0x - 2 m c 2x
R E L A T IV IS T IC Q U A N T U M M E C H A N IC S O F E L E C T R O N 145

(here we have omitted primes on the spinors <p and this will not lead
to any confusion since in future we shall use only the transformed func­
tion ip').
In order to carry out the expansion in powers of 1 /c, we assume that
X is of the order of magnitude of q>/c (this is justified by the result of
subsequent computations). Then in order to obtain the lowest-order
approximation it is possible to neglect in the second equation all terms
containing % with the exception of the last one the coefficient of which
contains c2. We thus obtain
1

X= (15.2)
2 me
The substitution of this expression into the first equation yields
dep
dt <P-

By utilizing the properties of the matrices ot. it can be easily shown


that

c
where H = curl A is the magnetic field. By utilizing this transformation
it is possible to write the equation to first approximation in the form

, d9 T_T (15.3)

Equation (15.3) is known as the Pauli equation. It differs from the


nonrelativistic Schrodinger equation by the presence of the Hamiltonian
H in the last term, which has the form of the potential energy of a mag­
netic dipole in an external field. Thus, in first approximation the elec­
tron behaves as a nonrelativistic particle having in addition to its charge
a magnetic moment
a. (15.4)

15.2. Second Approximation


We now obtain the second-order approximation by continuing
the expansion to terms of order 1/c2. At first we confine ourselves to
he case when only an external electric field is present (A = 0 ) .
146 Q U A N T U M E L E C T R O D Y N A M IC S

On substituting the first-order expression (15.2) for % into the


terms of the second of equations (15.1) which were neglected earlier,
and on solving the resulting equation with respect to % we obtain

^ ai K 4 (15'5)
On substituting this value of %into the first of equations (15.1) we obtain
the equation for <p\

;( 1 + 4^ ) 4t “

On multiplying the last equation by ( 1 —p 2 /4ra2c2) and on retaining


terms of order not higher than 1 /c 2 we rewrite it in the following form:

,J| = HV- H' = ( 1- 4 ^ ) ( i 4 + e'4») + 4 ^ (,’pM«('’p)' (15'6)


Although this equation does have the form of a Schrodinger equa­
tion, nevertheless it cannot be directly regarded as a wave equation,
since the function does not have the exact meaning of an electron wave
function, and the operator H' cannot be regarded as a Hamiltonian
operator. Indeed, in the first place the integral f <p*<p dr is not conserved
in time, since according to the equation of continuity only the integral
f (cp*<p-\-%*%)dr is conserved. Therefore the function <p cannot be nor­
malized. Secondly, H' is not a self-conjugate operator, since it contains
the term (e/4m2c2)p2T 0. Both these deficiencies can be eliminated if
we carry out a transformation of the wave function by introducing
in place of <p the wave function 0 :
0 = Ocp, (15.7)
and by choosing O in such a way that 0 can be normalized.
After such a transformation equation (15.6) assumes the form
80
i-T T = H 0 , (15.8)

where H = O H 'O -1.


We obtain the operator O by requiring that for terms of order 1/c2
the equation
J (<P*9J+X* fydr = J 0 * 0 dr
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 147

should hold. To achieve this it is sufficient to substitute into the above


equation the first-order expression (15.2) for which yields with the
required degree of accuracy

J <p*[l-\-(jp2/4m2c2))(p d r = J 0 * 0 d r .

From this it follows that the operator O may be chosen in the form

0 = (15.9)
8 m2c2
Since the operator O differs from unity only by a term of order
1/c2, the transformation (15.7) does not affect the first-order equation
in which <p is a normalized wave function to terms of order 1 /c.
We now find the Hamiltonian operator in the second-order approx­
imation. On substituting expressions (15.9) and (15.6) for the operators
O and H' into (15.8) and retaining terms of order 1/c2, we obtain

H- -\-eA0-
2m 8 m3 c2 8 m2 c' ,(p M „ - A p 2)- 4m2 c2 P2^o

(op)/t0(<*p).
4m2 c2

The last two terms in this expression contain operators which are
not self-conjugate. However, it can be easily shown that they cancel.
Indeed, (op) T0(ap) = A 0 p2—o (A0 p—pT0) op = A0 p2—i (aW40) (op)
= / l 0p2+ z F p - [ £ p ] a , where E = —VA0 is the electric field. On noting
that p2A 0—A 0p2 = —A A 0-\-2iEp, we finally obtain the following expres­
sion for the Hamiltonian operator:

H= -eAc o[£p]- A A (15.10)


2m Sm2c2 4m2c2 8 m2 c2 °’

which is Hermitian.
The last three terms in (15.10) are corrections of order 1/c2. The
first of them takes into account the dependence of the mass on the
velocity, while the second one can be interpreted as the energy of inter­
action between the moving magnetic dipole and the electric field. Since
the equation A A 0 = — o holds for the electrostatic field where q is the
charge density giving rise to the external field, the last term differs
148 QUANTUM ELECTRODYNAMICS

from zero only at those points where these charges are present. Tn partic­
ular, for the Coulomb field of the charge Ze we have
A A 0 - —Zet5(r).
By utilizing the Hamiltonian (15.10) it is possible to solve by pertur­
bation theory methods the problem of the fine structure of the levels
of a hydrogen-like atom, taking for the perturbation the last three terms
in (15.10). This leads to formula (13.4') obtained earlier by means of
expanding the exact solution (13.4) of the Dirac equation.
In the presence of an external magnetic field the Hamiltonian differs
£
from (15.10) by the replacement of p by p ----- A and of p 2 by
c

al p— == (p — — — —gH, and has the form

e e
H= aH+ AA0
8m 3c2 + e A ° 2me 8 m2 c2

(i5u)
The transition to the approximate equation for the two-component
function 0 can also be accomplished (6 6 ) by means of the unitary trans­
formation tp' = Uxp, H' = U H U -1, which leads to the system of equa-
• dty' , , (W'X
tions ! - - = H yj = H'l ,1, which breaks up into two pairs of inde­
pendent equations in <p' and %.
There exists no matrix U which has precisely the properties indicated
earlier, but it can be obtained approximately in the nonrelativistic
domain in the form of an expansion in powers of I/me2. In order to
do this we write the Hamiltonian H in the form
Hu H 12
H
lH2i H 22
where Hj/Care two-rowed matrices, and we break up H into the diagonal
. H ..
u 0 \. / o Hl.
and the nondiagonal H_ = I I terms
o h J \H 21 O)
H = H++ H _ ;
H+ = f^mc2A-eA0‘, H_ = ec(p—(e/c)A)a.
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 149

The problem consists of making the matrix H(_ equal to zero after
the unitary transformation. This may be achieved approximately by
choosing U = exp(/3H_/2mc2). On expanding U into a series it can
be easily seen that in the new Hamiltonian H' the nondiagonal part
contains only terms of order 1/mc and higher. Further, on carrying
out the transformation H" = U 'H 'U ' - 1 with the aid of the matrix
U '= exp {—(/?/2mc2)H^_} we obtain H " of order 1/(me)2, etc.

15.3. Application of the Dirac Equation to Nucleons


The Dirac equation describes any free particle of spin However,
its application to the behavior of a particle in an external electromagnetic
field is limited to electrons or to other particles (p mesons) for whieh
the electromagnetic interactions are the strongest ones.
In view of the restricted nature of the concept of an external field
the Dirac equation (12.1) is not exaet even for an electron. Thus, for
example, as has already been noted earlier, the formula for the energy
levels of the hydrogen atom is correct only to terms of order a4. If the
interaction of the electron with its own electromagnetic field is taken
into account (radiation corrections), then, as we shall see later (Chapters
VII and VIII), this will lead to correction terms in the equation describ­
ing the behavior of the electron in an external field. Thus, in equation
(15.11) the coefficient of the term containing aH will be altered, and
this can be interpreted as a change in the magnetic moment, and also
the coefficient of A A 0 will be altered. The use of equations (12.1) and
(15.11) is justified by the fact that these corrections are small.
The interaction of nucleons with meson fields is of the greatest
importance. Therefore, an investigation of the behavior of a proton
in an electromagnetic field cannot be carried out in even a partially
satisfactory manner. As a result of its strong interaction with the meson
field, which is associated with charged particles, the motion of a proton
in an electromagnetic field does not coincide with the motion of a par­
ticle which is characterized only by its charge. Equation (12.1) is not
applicable to the proton; the exact equation must take into account
the additional interaction with the electromagnetic field through the
meson fields.
In particular, as is shown by experiment, equation (15.3) of the
nonrelativistic approximation turns out to be inapplicable to the proton.
150 QUANTUM ELECTRODYNAMICS

For vjc < 1 the proton has a magnetic moment which is not equal
to the nuclear magneton /u0 = ehj2Mc (M is the proton mass), as would
follow from (15.4), but which has the value fxp = 2.7/x0. The neutron,
which has no charge, has a magnetic moment given by jxn = —1.9//0.
Finally we can take into account the additional “anomalous”
magnetic moment of the nucleons by introducing into the Dirac equa­
tion an additional interaction in the form —fx'oH where /x' is the “anom­
alous” part of the magnetic moment equal to fx' = /xp—[x0 in the case
of the proton and (x = in the case of the neutron.
The corresponding relativistically invariant equation containing
the electromagnetic field tensor Faff must be of the form

(*'(p —^ A) + * y y ay/Ji > + Mc2)v , = ° . (15.12)


or
= Hy> = \ c a U - ^ A ] - / . i 'P ( Z H - i a E ) + p M A y . (15.12')

We investigate with the aid of this equation the behavior of a neutral


particle which has a magnetic moment (e = 0, /x' ^ 0 ).
We now go over to the nonrelativistic approximation. On setting,
as we have done previously, -tp = we obtain
S r - "

[ i 4 t + f l , o H ) (p = c a ( p ~ ^ r E ) x
and

[ i ~ —fx'oH-sr 2 M c ^ x = c a |p + — -E^(p.

From here we obtain in first approximation %= (1 /2M c)a(p+(z//jc)E)cp.


The substitution of this expression into the first equation yields

• dcP = Ho?.
z-=— u
dt Y
(15.13)

H=
2^ d iv £ + 2 ^ ’£l-[£ ^ > + W

In accordance with (15.13) a neutral particle which has a magnetic


moment interacts with a static electric field (65). In particular, if this
RELATIVISTIC QUANTUM MECHANICS OF ELECTRON 151

equation is applicable to the neutron then the latter must interact with
the Coulomb field of the electron. On neglecting the quadratic term
in (15.13) we obtain for the energy of this interaction
t

F = ~ 2 J f c d]vE = ■

This interaction is equivalent to a potential well of finite width. For


example, for a well of width r0 = e2/4nmc2, the depth of the well turns
out to be approximately equal to 4 x 103 ev. This value apparently
agrees with the available data on the interaction between neutrons
and electrons. We note that the Hamiltonian (15.11) also contains the
energy of interaction between a magnetic moment and an electric field
(the term — (ju0/4Mc)d\vE), but with a coefficient which is only half
as big as the one in the analogous expression for the interaction of
the anomalous magnetic moment.
However, it should be emphasized that the conclusion about the
existence of the interaction (15.14) on the basis of equation (15.12)
is illusory. In actual fact there can exist no equation of the type (15.12),
nor one similar to it, which does not take into account the interaction
of a nucleon with its own meson fields. We have already noted that
even for an electron the radiation corrections distort the coefficients
of aH and div E in its Hamiltonian. It is a fortiori impossible to establish
a relation between these coefficients for a nucleon in the presence of
strong meson interactions. Such a relation can be obtained only on
the basis of a theory of meson-nucleon interactions, which must also
give a quantitative expression for the neutron magnetic moment. Since
at the present time such a theory does not exist, the behavior of nucleons
in external electromagnetic fields may be approximately described
by equations of the type (15.11), (15,13) with empirical coefficients.
In variable fields these coefficients may depend on the field frequency.
CHAPTER lit

Quantized Electromagnetic and Electron-Positron Fields

§ 16. Quantization of the Electromagnetic Field

16.1. Four-Dimensional Form of the Field Equations


In Chapter I, starting from the classical Maxwell’s equations we
obtained photon wave functions in momentum space, and investigated
quantum photon states. However, for the general formulation of the
theory of the electromagnetic field and of its interaction with electrons,
and also for the investigation of specific phenomena to which this inter­
action gives rise, it is more convenient to utilize another method, to
the presentation of which we now turn. It consists of treating the electro­
magnetic field as a certain generalized dynamical system, and of regard­
ing the quantities defining the field as generalized coordinates of this
system. The quantum nature of the system is taken into aceount in
this treatment by replacing the classical quantities by the corresponding
operators. These operators act on the wave function of the system
which depends on the “occupation numbers,” i.e., on the numbers
of particles—photons—present in the different quantum states. This meth­
od of describing the electromagnetic field as a quantum mechanical
system, utilizing the wave function of the system in occupation number
space, is referred to as the method o f second quantization.1
We note that the transition from the wave function in configuration
or in momentum space to the wave function in occupation number
space is a canonical transformation. Therefore, the two approaches
to the quantum theory of electromagnetic processes from the “corpus­
cular” (photons and their wave functions) and the “field” (field and
its quantum mechanical study) points of view are equivalent.
Thus, we shall treat the four-potential (A, iA0) and the electromag­
netic field vectors E and H not as ordinary quantities (c-numbers),
but as operators (^-numbers). In the same way that Newton’s mechanical
1 This method, as applied to radiation theory, was first developed by Dirac (46).
[153]
154 QUANTUM ELECTRODYNAMICS

equations which are satisfied by the momentum and coordinate opera­


tors are retained in nonrelativistic quantum mechanics, so in quantum
electrodynamics we retain Maxwell’s equations which are satisfied
by the field operators. But in quantum mechanics the equations of
motion acquire a concrete meaning only after the establishment of
commutation relations between the components of the momenta and
the coordinates; therefore in quantum electrodynamics it is also neces­
sary, first of all, to establish the commutation relations between the
dynamical quantities characterizing the electromagnetic field. These
relations are referred to as the field quantization rules.
Since the field quantization rules must be invariant under Lorentz
transformations, we shall, before formulating these rules, give the invar­
iant four-dimensional form of the equations of the electromagnetic
field (cf., for example, (118)).
As is well known, the electric and the magnetic fields E and H together
form the antisymmetric electromagnetic field tensor F ^ whose compo­
nents are equal to Fik = em H,, Fi k = i E k, where eikl is the unit
antisymmetric tensor (/, k, I — 1,2, 3), or

0 , H z, ~ H y , ~ iE x
0 , Hx, — lEy
fiv
Hy, ~ H X, o, ~ iE z
iEx, iEv, iEz, 0

With the aid of the tensor F the first and the second pairs of Maxwell’s
equations can be written in the following form:
dF.VQ dF.ev
= 0,
dxe dx„ dxv
(16.1)
dF. _
0 .
dx..
Here all the subscripts assume the values 1, 2, 3, 4 and the repeated
subscript v is summed over from v = 1 to v = 4; x 4 = ix0 = it.
The tensor components can be expressed in terms of the four po­
tential of the field A^(A, iA0)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 155

These relations are equivalent to the first pair of Maxwell’s equations


(16.1).
The four-potential A)t is not uniquely determined by the field F ,
since the addition to of the gradient of any scalar does not alter
the field. In other words, Maxwell’s equations are invariant under
the transformation AM-*■ A ^ = A flJr(dr]/dxfi), where rj is an arbitrary
scalar. Such a transformation is referred to as a gradient or a gauge
transformation.
All physical relations containing the electromagnetic field must
obviously be invariant under gauge transformations.
Since the potential A^ is not uniquely defined, a subsidiary condition
(the Lorentz condition)
dA^
= 0 (16.2)
dx,<
may be imposed upon it.
By utilizing equations (16.1) and the Lorentz condition it can be
easily shown that the components of the potential satisfy the wave
equation

U A ^ 1 ^ A» = ° ’ (16.3)

These equations, together with the subsidiary condition (16.2), are


equivalent to Maxwell’s equation.
Condition (16.2) does not completely determine the gauge of the
components of the potential A . It restricts gauge transformations to
those functions rj which satisfy D?y = 0 .
We note that in order to satisfy condition (16.2) it is sufficient to
require that at some initial instant of time t = 0 the quantity x =
dA /dx: and its time derivative dx/dt should vanish. Then the function
X will also vanish at all later times (208).

16.2. Variational Principle. Energy-Momentum Tensor of the Electro­


magnetic Field
The second pair of Maxwell’s equations (16.1) can be obtained
from a variational principle if we assume that a definite Lagrangian JQ
corresponds to the electromagnetic field regarded as a generalized dynam­
ical system. In such an approach the field potentials at different points
156 QUANTUM ELECTRODYNAMICS

A (x) play the role of generalized coordinates of the system, while


dA/1/dxv are the generalized velocities. The integral of the Lagrangian
with respect to the time 1 = J 12 clt is known as the action. It can be
represented in the form of a four-dimensional integral / = / L d Ax , where
L is known as the Lagrangian density, and the integration is carried
out over the invariant four-volume (both now and later we denote the
element of volume dx1dx2dx3dx0 = drdt by dix).
The field equations can be obtained with the aid of the variational
principle <5/ = 0 , analogous to the variational principles of classical
mechanics. In carrying out this variation the generalized coordinates,
i.e., A^(x), are varied, and it is assumed that the variation of A^(x)
vanishes over the boundaries of the region of integration.
We derive the field equations by assuming that, in general, L depends
both on A M, and on 8 A jd x v. The variation of the action is obviously
equal to
dL 8L 8A„
61 = dAx
dA„ ^ d id A jd x J 8xv
r l 8L 8 dL
6A,,dA
f* x,
J \ 8Afi dxv d(8A J8xr)
and, therefore, the field equations, which are the Lagrangian equations
of the system, have the form
8 8L
-v-j—= 0, where /u = 1 ,2 , 3, 4. (16.4)
dxv d(8AJ8x J
In order that the field equations should be invariant under Lorentz
transformations the action must be invariant, and, consequently, the
Lagrangian density L must also be invariant.
On the other hand, the equations of the electromagnetic field must
possess the property of gauge invariance, and, therefore, the Lagrangian
density cannot contain the components of the four-potential directly,
but only the components of the field tensor F . But from the field
components we can construct only two independent invariants: E 2—H 2
and (EH)2, so that the Lagrangian density can depend only on the two
invariants, L = L[E2—H 2 and (EH)2).
If, in addition, we require that the superposition principle should
hold, i.e., that the field equations should be linear, then the Lagrangian
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 157

density can only be a quadratic function in the field components. In


other words, up to a numerical factor the Lagrangian L must coincide
with E2—H 2. We set
L = k(E2—H 2) — (16.5)
which corresponds to the use of Heaviside units.
In accordance with (16.4) the field equations for this Lagrangian
assumes the form

dxv didAJdx^)
and since

<5L= = - I f I (3 f
dx„ ^ dx.. ’
i.e.,
___ — ____ = F
didAJdx,)

we finally obtain the second pair of Maxwell’s equations dF)lJdxv — 0.


We use the Lagrangian equations (16.4) to obtain the four-dimen­
sional form of the conservation laws for the energy and the momentum
of the electromagnetic field. We consider the derivative dL/dxv;
dL _ dL d(dAg/dxg) dL dAg
8xv ~ d(dAJdxg) dxv dAg dxr '
On substituting into this expression the value of dL/dAg from (16.4)
we obtain
8L _ 8Ag d 8L d dAe dL
l x v ~ ~ dx^ dxg d{dAJdxg) + dxv dxg d(dAg/dxg)
d ld A p dL \
d x ^ \d x v d{dA Jdx^J
whence it follows that

d I f S ______— ___ ^ | = 0 ,
5*..
> \\ d(dA
- eJdxJ dxv
or, on introducing the notation
dL dAr
(16.6)
Tflv d(dAJdxv) ~ d x ’
158 QUANTUM ELECTRODYNAMICS

we obtain finally

= o where u = 1, 2, 3, 4. (16-7)

The quantities T form a four-dimensional tensor which is called the


energy-momentum tensor of the electromagnetic field.
The laws of conservation of energy and momentum follow from
equation (16.7). Indeed, we consider the four-vector

Tr,dr, (16.8)

where the integration is taken over the whole volume of the field.
We evaluate d P jd t in accordance with (16.7)
dT.Mi
dr.
dt " J dxt J dx<

The three-dimensional integral appearing in this expression reduces


to a surface integral which can be taken equal to zero. Therefore, P
is conserved:
Pp = const. (16-9)

It represents the energy-momentum four-vector for the field; its spatial


components represent the momentum of the field, and its time-like
component divided by i is equal to the energy of the field.
On utilizing the formula for L it is possible to express the tensor
in terms of F^:
T,v = (16.10)
or
Tik = \{E*+H*)dik~ E iEk- H iHk, Tk i = - i s k (i , k = 1 ,2 ,3 ),
Tu — —w.
where w = \ { E 2-\-HT) is the energy density of the field, and s = [E, H]
is the energy flux density of the field.
The wave equations (16.3) for the potential can also be obtained
from a variational principle, but in this case the Lagrangian density
must be taken not in the form (16.5), but in the form

-IF ^-U d A Jd x^,


E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 159

or

L_ (16.11)
2 dxv dxv

These two expressions differ by the quantity

_ i JL A
2 \ ' dxM * dxu I

which has the form of a four-dimensional divergence, and therefore


lead to the same field equations. In order to obtain Maxwell’s equations
the subsidiary condition (16.2) must now be added to the wave equations.
On taking the Lagrangian in the form (16.11) we obtain the expression
for the energy-momentum tensor

r _ r ( ____8L____ dAa d A jd A , J_ dA^dA.


(16.12)
^ d (d A jd x ^ d x v dxv dxM 2 dxe dxe ■

In the future we shall make use of the wave equations (16.3) and
the subsidiary condition (16.2) as our field equations.

16.3. Expansion o f the Potentials into Plane Waves


The general solution of the wave equation can be represented in
the form of a superposition of plane waves with arbitrary amplitudes.
We write it in the form

^O )
_L y _ L , i x) (cMeikl+c+ye-ikx). (16.13)
ki

Here Q is the normalizing volume, k x = kr — cot is the scalar product


of the four-vector x(r, it) and the four-dimensional propagation
vector k(k, ia>) which satisfies the condition k 2 = k 2—a>2 = 0 , ckx
are the wave amplitudes, and e{2) are the polarization unit vectors.
For a given k four linearly independent solutions are possible differing
by their polarizations. We choose the vectors ew as the basis vectors
for the coordinate system e ^ = SXfi. If we direct the x3-axis along k
and the other two space axes x 1 and x 2 at right angles to k and to each
other, then 2 = 1 , 2 will correspond to transverse, and 2 = 3 to longi-
160 QUANTUM ELECTRODYNAMICS

tudinal polarization; 2 = 4 corresponds to the so-called scalar polari-


zation. The following relations hold:

II
o
e '% = co, 2=3, (16.14)
ico, 2=4,
where
pWpW)
e e — — uu
X ,. and = 6 ^. (16.14')
From the relation k 2 = A2 —co2 = 0 it follows that the frequency
co can be of either sign. This fact is taken into account in (16.13) by
the existence of two types of terms, containing exp (ikx) and exp (-ikx)
respectively, in each of which we must take co = |A|.
We note that the sign of the frequency is relativistically invariant,
as can be verified directly with the aid of Lorentz transformations.
Therefore, the decomposition of A (x) into terms with positive and
negative frequencies is also invariant.
Since the spatial components of the potential A^(x) ( j = 1,2,3)
are real, while the time-like component A t (x) = iA0 (x) is a purely
imaginary one, the amplitudes ckX of waves with positive and negative
frequencies must satisfy the following relations = ckj, c^ 4 = —cki,
where c* denotes a quantity which is complex-conjugate with respect
to c. We also use the notation cki = ick0, c^ 4 = ic%0.
As we have seen earlier, the potentials A (x) must in addition to
(16.3) also satisfy the subsidiary condition (16.2). On substituting
the expansion (16.13) into (16.2) we obtain, after utilizing (16.14),
dA, k
dx» m l
|/ 0 t i
* ■{ckx<P*-c&er«*)
j/2co

= y g J J j/Y + ^ 4) e - ikl] = 0 ,

from which it follows that


(16.15)
By utilizing the expansion (16.13) it is possible to write the field tensor
in the form
F = — V — ==■ (<*,«“ ' - ‘ tie-*-)
U J t i V2 CO
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 161

The electric and the magnetic fields are equal to

The energy-momentum tensor defined according to (16.12) can be


expressed in terms of the plane wave amplitudes in the following manner:

The energy and the momentum of the field, which are related to
T^v by expressions (16.8), are equal to

(16.16)

fc;A=l,2
We see that the electric and magnetic fields, and also the energy and
momentum of the field, do not depend on the longitudinal or scalar
waves. This is connected with the existence of conditions (16.15). If
we had not used these conditions, we would have obtained for H and P
the same expressions (16.16) in which, however, X would assume the
values 1, 2, 3, 4. The terms corresponding to the polarizations A= 3, 4,
cancel because of the conditions (16.15).

16.4. Quantization o f the Electromagnetic Field


The expressions for the energy and the momentum of the electro­
magnetic field enable us to establish the quantization rules for it. These
rules must lead to the correct corpuscular picture, which consists of
the energy and the momentum of the field being sums of the energies
or the momenta of the individual particles—the photons.
In order to arrive at such a picture it is necessary to regard the ampli­
tudes ckX of the plane waves in the expansions for the potentials (16.13)
not as ordinary numbers, but as operators satisfying definite commu­
tation relations.
162 QUANTUM ELECTRODYNAMICS

We assume that these relations are

[CfcA> Cfc'rl = ^kk'^XX'^ tCfcA> Cfc'A'l [ < * , Cfc'A'] 0, (1 6 ] 7 )


A, A '= 1,2, 3,4,
or, in abbreviated form, [cA, qt] = 3XX,, [cA, cA<] = 0 , [cjf, c£| = 0 ,
where cA= ctA and [a, b] = ab —ba. Moreover, in accordance with the
fact that A and A0 are real we must assume that the operators cAand
c^ with A = 1, 2, 3, and also the operators c0 and cjf are Hermitian
conjugate.
We shall show that from the quantization rules (16.17) it follows
that the eigenvalues of the operator NA= NfcA= cfc x with A = 1, 2, 3
are equal to Nx = 0, 1, 2, ...
First of all, it is evident that the eigenvalues N x are not negative.
Indeed, from the equation c fc x0 = Nx0 , where 0 is the eigenfunc­
tion of the operator NAbelonging to the eigenvalue Nx it follows that

¥ _ _ (ca^> c^ )
* (0 ,0 ) (0 ,0 ) ^ ’

where ( 0 X,0?) denotes the scalar product of the wave functions 0 X


and 0 2.
We consider the function <p = cx0 on the assumption that Nx > 0.
It differs from zero, since c f ( p = c x cx0 = N x0 ^ O , and it satisfies

N x<p = c+cAcA0 = (cAc + - l ) c A0 = (N A—\)<p.

Therefore, cp is the eigenfunction of the operator NA belonging to the


eigenvalue jVa-1. Similarly, we can show that Nx-2, NA-3 , ... are also
eigenvalues of the operator NA. But such a series must, evidently, break
off since the eigenvalues N x cannot be negative. This is only possible
in the case that Nx is either a positive integer or zero.
It can be easily shown that all integers are eigenvalues of the opera­
tor Na. This follows from the fact that the function %— c f 0 which
satisfies the equation

n a% = c t c xc + 0 = c£(\+c+cx) 0 = (Na+ 1)x,

is the eigenfunction of the operator NA belonging to the eigenvalue


iVA+ l .
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 163

We note that the foregoing problem is equivalent to the problem


of finding the eigenvalues of the energy of an oscillator. Indeed, if
we set

ca = (Pa+ . ca+ = (Pa - i m d ,


y 2co y2co

then q; and pAwill satisfy the commutation relations for the momentum
and coordinate operators [q^, p j = i, while the operator Hx = a)
(c^cA+ ^ ) will assume the form of the energy operator for the oscillator
HA= i(p^+cu 2qi).
We now obtain the matrix elements of the operators cA and c£.
We denote the normalized eigenfunctions of the operator N, = c^cA
by 0 Vr Then from the foregoing it follows that cx0 N} — const • 0 Nx_1
and c f 0 x x = const - 0 Nx+x. These relations mean that the only
nonvanishing matrix elements of the operators cA and are
(CaW^+a and ( c ^ ) ^ ,W;i (Nx = 0, 1, 2, ...), Up to a phase factor
exp (fa), which may taken equal to unity, they are equal to

(Ca) ^ . ^ + 1 = ’ (CaOat;i, = A = 1’ 2, 3, (16.18)


since
•^A= (Ct C?)n x ,N x ~ (C^ ) w ; . ^ - i ( Ca)v^-1,Ata = I (C;t 12 = II2-

On utilizing these relations we obtain

‘A , = V ^ nx- i > ct®NX= V N x + I ^ a + i • (16-19)


The quantization rule (16.17) for X — 4 can be rewritten in the form
[c0+ ,c 0] = l , (16.20)

whence it follows that the eigenvalues N 4 of the operator c0qf are equal
to 0, 1, ... The nonvanishing matrix elements of c0 and cjf are obviously
equal to
(co)w,..v«+i = (co)^,ut- i = (16.21)
We now show that the quantization rules (16.17) actually lead to
the correct corpuscular picture of the field. In order to do this we obtain
the operators for the energy H and the momentum P of the field, which
164 QUANTUM ELECTRODYNAMICS

in accordance with (16.16) can be expressed in terms of the operators


cA and c^:
H = i ^ w ( c ; c++c+ca),
k, X
(16.22)
k. X

where the summation is taken over k, and over all the polarizations
A including the longitudinal and the scalar ones.
From these expressions it follows that the states for which the
operators NA= c ^ c A have definite values are states of the field with
a definite value of the momentum. This is connected with the fact that the
amplitudes cA and c^ are the coefficients in the expansion of the po­
tential in terms of plane waves, i.e., in terms of photon states of definite
momentum.
We note that the angular momentum of the field in these states
does not have a definite value, since the angular momentum operator does
not commute with P.
We shall obtain the states of the field having definite values of
angular momentum if we expand the field potentials not in terms
of plane waves, but in terms of the spherical waves AV)jMX which
describe states of definite angular momentum (cf. formulas (6.16),
(6.17)). This expansion can be written in the form
A/t a (A / ) u CWJMX ( A a t 5
co3j, M, X

where the expansion amplitudes cwjMX and c+/MA regarded as operators


satisfy the following commutation relations:

\-Cw j M X ’ Cto'j'M'X'] ^coco' ^ j j ' ^ M M ' ^ X X ' ’

\-Cu>iMX ’ = t Ci f M X ’ Co>'j’M'X'1 = 0 -

The validity of these relations can be verified if we use the expansion


of plane waves in terms of spherical waves (cf. subsection 4 .4 ).
The operators for the energy and for the component of the angular
momentum of the field are expressed in terms of the amplitudes cm)m
and ctmx as
^ 2 S °->( CaijMXCu)jMX~^~Cm ]M X (:uijMx') ’
0), j, M, X2

2 S ^ ( Cw)MXCmi MX' ^ CtojMXCmjMx) •


ui, i , M, X
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 165

From this it can be seen that the states of the field in which the
operators = c+i;jJl have definite values (they are equal
to 0 , 1 , ...), are states of definite energy, angular momentum, and com­
ponent of angular momentum.
The operators for the energy, momentum, and angular momentum
of the field contain all the polarizations including the longitudinal
and scalar ones (2 = 3,4). We have seen earlier that the subsidiary
condition (16.2) imposed on the potentials guarantees that in classical
electrodynamics all physical quantities are independent of the longi­
tudinal and scalar oscillations. This independence must, of course,
also exist in quantum electrodynamics, but the subsidiary condition
in its ordinary classical form is incompatible with the conditions of
quantization. Indeed, in accordance with (16.15) we have cfc3 = —icfc4,
while in the quantization rules the operators cfc3 and cki are independent.
The subsidiary condition for the quantized field can be formulated
in the form of a condition imposed not on the operators cfc3 and cfc4
but on the wave function 0 (59). In particular, we shall regard as ad­
missible only those wave functions which yield the result zero when
the part of the operator <3AJ d x M which contains positive frequencies,
i.e., terms of the type exp (ikx) is applied to them. On denoting this
part of the operator dAJdx^ by (SAJdx^)+ we write the subsidiary
condition as

( - ^ 4 * =0> <‘« 3 )
or
( 2 \f (O(C*3+ ' 0 ^ 0 = °,
k

whence it follows that the function 0 must satisfy for all k the conditions
(c* 3 + fcfc4) 0 = O , (16.24)
and, consequently, also the conditions
^ * (ci -3+ /cii) = 0. (16.24')
In other words, the function 0 , in addition to satisfying condition
(16.23), also satisfies the following condition:

(16.25)
166 QUANTUM ELECTRODYNAMICS

where (8A Jdx ) is that part of the operator dAJdx^ which contains
negative frequencies, i.e., terms of the type exp ( —ikx).
On taking the scalar products of (16.23) on the left with 0*, and
of (16.25) on the right with 0 , and on adding these two expressions, we
obtain
= 0, (16.26)
dx.

where <(L> = (0, L 0 ) denotes the expectation value of the operator


L in the state 0.
We see that in quantum electrodynamics the quantity which vanishes
is not dA Jdx , this requirement being impossible to satisfy, but the
expectation value of the operator dAJdx^ in the state 0. As a result of
this, in quantum electrodynamics the equations for the field tensor
instead of being of the form (16.1) are

( f . ^ L . 4>) = 0. (16.27)

It follows from (16.24) and (16.24') that

( 0 , (c^, /c+) (ck3+ icki) 0 ) 4- ( 0 , (C++/C+) (cfc3 icki)0) = 0,


*

(^» (C*3Cfc3“fiCfc4Ck4)^) = <\Ck3 Ckl)' Ck4^)= (16.28)


However, it does not follow from this that N3 = A 4 = 0. On the contrary,
as we shall see later, there exist no states of the field with iV3 = 7V4 = 0 .
We now determine the eigenvalues of the operator for the energy
of the field (16.22). On noting that c^c^ = c^cA+ l , and on utilizing
(16.28), we find that the eigenvalues of the energy of the field are equal
to
£= I o W a + i). (16.29)
k;A=1 ,2

We see that if we disregard the infinite sum \ to, the energy


k;X=1 ,2
of the electromagnetic field is indeed expressed in the form of the sum
of the energies of the individual particle—the photons. The integer
NklL= Nx represents the number of photons with propagation vector k,
polarization X (A = 1,2) and energy co.
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 167

The energy of the field is determined only by the transverse vibra­


tions (A = 1, 2). The osciliations with X = 3, 4 and the corresponding
numbers of longitudinal and scalar photons Nk3 and Nkl do not appear
in the expression for the energy.
In a similar manner we can obtain the eigenvalues of the momentum
and of the component of the angular momentum of the field, which
turn out to be given by P = J] k(Nkx+ \ ) and M t = £
where Najm are the eigenvalues of the operator
We see that if we disregard the sums £ and £ , then
h, X a), /, M ‘X
the momentum, or the component of the angular momentum of the
field, is equal to the sum of the momenta or of the components of
the angular momenta of the individual photons.
It follows from (16.29) that the energy of the field has the lowest
value when all the numbers Nkx(X = 1,2) are zero, i.e., when no
photons are present. This lowest energy state of the field is named
the electromagnetic field vacuum. It is a characteristic feature that
in accordance with (16.29) the value of the energy of the vacuum is
different from zero. To each degree of freedom of the field corresponding
to given values of k and X there corresponds in the vacuum state the
zero point energy equal to ^ co.
Taken by itself, the presence of an additive constant in the expression
for the energy of the field is, of course, not essential, provided we are
dealing only with the eigenvalues of the corresponding operators or
with their expectation values. However, it would be incorrect to assume
on this basis that the zero point energy can be simply neglected. We
shall see later that zero point oscillations appear in the investigation
of the interaction of the electromagnetic field with electrons and lead
to a number of observable effects (these effects will be discussed in
Chapter VIII).
In concluding this subsection we note that the operators cA and
Cj” have a simple physical meaning. If the field contains Nx photons
of polarization X and &N denotes the corresponding wave function,
then it follows from (16.19) and (16.21) that

CA * =
(16.19')
co<ZV. = V ^ 4 + 1 n> = A ^ , - 1 •
168 QUANTUM ELECTRODYNAMICS

Thus, the operators cx and c f with 2 = 1, 2, 3 represent annihilation


and creation operators for a photon of polarization X. For 2 = 4 the
roles of c0 and are interchanged: c0 represents a creation operator,
and cj~ represents an annihilation operator for a scalar photon.

16.5. Use o f the Indefinite Metric


We have already seen that the vacuum of the electromagnetic field
represents a state of the field in which no transverse photons are present,
i.e., the numbers Nx ( 2 = 1,2) are equal to zero. However, insofar
as the longitudinal and the scalar “photons” are concerned, their numbers
N 3 and N i differ from zero. Let us investigate that part of the wave
function of the field which describes longitudinal and scalar oscillations.
This function, which we denote by 0 3 4, can be represented in the form

^3 . 4 = 2 C C ^ 3 > -^4 )^AT„ N, »


N,. V,
where the function 0 NliNl describes states of the field with definite
numbers of longitudinal and of scalar “photons”, equal respectively
to N3 and Ni . On substituting this expression for <P3 4 into (16.24)
and on utilizing (16.19'), we obtain
2j c(N3, N±) {)/N 30 Na_lt Nt )/iV4+ l = 0 -
N„Nt

On making in the first term the substitution JV3- 1 -» N 3, and in the


second term the substitution AT4+ 1 -» N iy we rewrite this relation
in the form
2 c(7Y3 + l,A r 4) |/A ^ + T 0 N3Ni
W,>0 ,
- I c(A r,,N 1- i ) | / i v > , , l. , , . = o,

from which it follows that the quantities c(N3, N 4) satisfy the difference
equation
c(iV3+ 1, AQ j/A ^+T = c(N3, N , - 1) |/7V;,
where JV3 = 0 , 1 , ..., JV4 = 1 , 2 , ...,
and the boundary condition c(N3, 0) = 0; N 3= 1, 2, ...
On introducing in place of c(N3, Af4) the new unknown 6(7Y3 ,A^4)
so that
c(N3, AQ = j/(N J/N 3!) b(N3, N ,) ,
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F IE L D S 169

we obtain b(N3+ 1, NA) = b ( N 3, NA—l) = b(N3— 1, NA- 2 ) = . . . ,


whence it follows that b (N3, NA) depends only on the difference Ni —N3.
By utilizing also the boundary condition we finally obtain the desired
function c(N3, 7V4):
0 , na < n 3,
c(N3, Na) _______ (16.30)
|/(AT4! / A 3! ) / ( iV 4- A Q , N A ^ N 3 ,

where f ( N ) is an arbitrary function of the non-negative integer N.


We see that various states of the field of the longitudinal and the
scalar oscillations are possible, but that there exists no state in which
the numbers of the longitudinal and the scalar “photons” are equal to
zero.
It follows from (16.30) that the wave function cannot be normalized,
since I |c(AT3 ,Ar4) | 2 = oo, and consequently, the usual definition of
N 2. N t
expectation values in the state 0 3 4 has, strictly speaking, no meaning.
The longitudinal and the scalar “ photons” do not appear in the
expressions for the energy of the field, and have no physical meaning
in the case of the free electromagnetic field. Their appearance in the
theory leads only to a formal difficulty, which can be eliminated by
altering the quantization rule for the scalar potential, and by using
a definition of the expectation value of an operator which differs from
the customary one. According to the usual definition the expectation
value of any operator L in the state F is equal to

<L> = (F, L!F) = f F * L F d q ,

where F* is the function complex-conjugate to F and dq is the prod­


uct of the differentials of the variables of which IF is a function; more­
over, it is assumed that F satisfies the normalization condition

(F, F ) = J F* Fdq = 1 .

However, it is possible to adopt a different definition of the expecta­


tion value of an operator, and of the norm of a wave function (150):

<L> = (F , LF ) = J F+LFdq, ( F , F ) = J F+Fdq = ± 1 , (16.31)


where F+ = F*r\ and r\ is a certain Hermitian operator.
170 QUANTUM ELECTRODYNAMICS

With such a definition of the expectation value the norm of the


wave function may be negative. In this case we speak of the indefinite
metric in Hilbert space.
Obviously, the expectation value of any operator which corresponds
to some physical quantity must be real.
We now show that the expectation value of L will be real if the
operator L satisfies the condition
L=L, (16.32)
where L = y]-1 L+y) and L+ is the operator which is Hermitian conjugate
to the operator L. For the sake of simplicity, we examine the case of
discrete variables. In this case we have

i.k, I

Since, by definition, r\ is a Hermitian operator, then the quantity which


is complex-conjugate to <L>, can be represented in the following form:

<L>* = 2 ! P , * ( L U ? , = 2 'P fn , = <L>.


i,k, I i.k, I

If condition (16.32) is satisfied, then <L>* = <L>, i.e., <L> is real.


We refer to the operator L = y)“ 1 L+y) as the operator conjugate
to L, and to an operator satisfying condition (16.32) as a self-conjugate
operator.
When the indefinite metric is used in Hilbert space, with the norm
of a vector defined not by J W*Wdq, but by J xF*y]Wdq, the self­
conjugate operators in the sense of (16.32) play the same role as
Hermitian operators do in the usual scheme, which defines the norm
as J W * W d q = 1 .
The general definition of an expectation value in the form (16.31)
based on the indefinite metric can be utilized for carrying out the quanti­
zation of the electromagnetic field. In this way we eliminate the difficulty
noted earlier which is associated with the introduction of longitudinal and
scalar “photons”, and which consists of the impossibility of employing
the usual definition of expectation values in the state 0 34 which de­
scribes longitudinal and scalar oscillations.
The method of quantization utilizing the indefinite metric is devel­
oped in the following manner (80):
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 171

We start with the expansion of the potential (x) into plane waves
in the form

a /W = 4 = y ~ 4 = ( ckxe>kx+ c kxe - iki) e^ , where j = 1 ,2 , 3,


V Q k.X y 2 a )

(16.33)
A 0(x) = — y —L= (cu eikx+ ckxe~ikI)
V & k.x V 2 o >

and we assume cw = and cu = cA to be operators conjugate in the


sense of (16.32) and satisfying the following quantization rules:
[c>> c j = 1, [c0 ,c 0] = —1. (16.34)
The Hermitian operator rj which appears in (16.32) is defined by
the conditions
Cj = c+, y = 1 ,2 ,3 ,
co = C<M
i.e.,
Cj~i] — iQCy~, j 1 ,2 ,3 ,
(16.35)
Ccf*l= *]co"-
We note that the expansion (16.33) and the quantization rules
(16.34) differ from (16.13) and (16.17) formally only by the replace­
ment of c+ bye. Therefore, in the future, we shall, for the sake of uniform­
ity, utilize the notation (16.13), keeping in mind that all the formulas
will remain valid if we replace cH‘ by c.
It follows from (16.34) and (16.35) that the eigenvalues of the oper­
ators Ny = C)Cj = c f c j ( j = 1, 2, 3) are equal to Nf = 0 , 1 ,2 , ... ,
while the eigenvalues of the operators c0 c0 = —c<j~c0 are equal to N0
= 0 , -1 ,-2 , ; TVj and N 2 represent the numbers of transverse photons,
while N a and N0 represent the numbers of longitudinal and of scalar
“photons.”
As can be easily shown, the operator rj commutes with CyCy and
c0 c0. Therefore, we may assume that the operator r) is diagonal in
the representation in which CyCy and c0 c0 are diagonal. It follows from
(16.35) that the diagonal elements of Y] can be represented in the form

(N1, N 2, N a, N 0 |Y) | Nx, N2, N 3, N0) = yInJInJInJIn*’


172 QUANTUM ELECTRODYNAMICS

where rjN satisfy the conditions

Vs’j+i = Vxj’ J = ^! 2 , 3,
^w0+i = Vn, •
These conditions show that the matrix elements of the operator rj have
the form
(A^ , N,2 ! Kz A* <16-36)
We denote by 0 N the eigenfunction which describes the state which
contains N x photons of polarization A (A = 1, 2, 3, 0). It follows from
(16.31) and (16.36) that these functions must be normalized as follows:

(P n}>0 n’) = <5^.np 7 = 1 ,2 ,3 ,


(16.37)

By utilizing (16.35) and (16.19') we can show that the result of


operating with the operators cA and c; on 0 N and 0 Na is given by
the following formulas:

— V'Nj&N)—is j — 1» 2, 3,
Ci = ] / \ 0 Nj+x,
(16.38)
CO^V0 = ]/^O^A'0-l>
C0 ^V„ = l/A o+ 1 0KO+1 •

We see that if the quantization is carried out in accordance with


relations (16.34), then both c} (j = 1,2,3), and c0 will be photon
annihilation operators, while c; and c0 will be photon creation operators
in contrast to the quantization based on the quantization rules (16.17)
in which the operator c0 was the creation operator for the scalar “pho­
ton.”
We now determine the form of the allowable wave functions 0
which describe the states of the electromagnetic field. The subsidiary
condition (16.23) retains its form and we now have (dAfJdxf) ± 0 — 0.
In place of (16.25) we must write 0+{dA /dx )_ = 0. It follows from
(16.23) that

(c3 CO) 0 — 0 . (16.39)


ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 173

We seek 0 in the form

Ni, N2, Ns, Ns

where 0 NiiNliNaiNll is the normalized wave function which describes


the state of the field containing Nx and N 2 transverse, jV3 longitudinal
and N0 scalar photons, while b(N1, N2, Na, N0) are numerical coeffi­
cients. Conditions (16.38) together with (16.37) lead to the relations
y ' N M N l t N 2tN 3- l , N0) + }/^>(tV 1 ,A^2, tV3, AT0 - 1 ) = 0
(the quantities b with negative arguments are equal to zero), whence
it follows that these states of the free electromagnetic field are possible:

0 j = 0 N i . N t , 1.0 0 ,15 (16.40)

02= 0N l N z. 2 . 0 } / 2 0 N^ N i l l + 0 Nj.Nz,0,2 ,

or in general 0 n = —~ ( c 3 —co)7!0 o, where n is the total number of


| / n\
longitudinal and scalar “photons.”
We see that in contrast to quantization based on the quantization
rules (16.17), in the case of quantization based on conditions (16.34)
and the indefinite metric, a state of the field is possible in which
no longitudinal or scalar “photons” are present (the state 0 O).
On forming an arbitrary superposition of states (16.40) we obtain
the general state of the electromagnetic field 0 = 0)o-\-a10 l Jr a20 2Jr •••,
where alf a2, ... are arbitrary numbers. A definite choice of these num­
bers corresponds to a definite choice of the gauge for the potentials
with the condition (16.39) being taken into account.
The function 0 is normalizable. Indeed, by utilizing (16.37) and
(16.40) it can be easily shown that for all i and j, with the exception
i = j = 0, we have (0 j5 0 ;) = 0. Therefore,
(0 ,0 ) = ( 0 O, 0 O) , (16.40')
independently of the values of a1}az, ...
Thus, the norm of 0 contains only that state in which no longitudinal
or scalar “photons” are present.
174 QUANTUM ELECTRODYNAMICS

If we neglect the zero point energy, then the operator for the energy
of the field in terms of the variables cAand cA has the form

H = ^ co(cj Cj CgCg-f- c3 c3 c0c0) .


k

From this, on making use of (16.40'), we can easily conclude that the
states 0 1} &%, ... make no contribution to the expectation value of
the energy of the field, which, as should have been expected, is deter­
mined only by the transverse vibrations. We can therefore completely
leave the states &1} ... o,ut of consideration, and assume that in
the free electromagnetic field only transverse photons are present, while
the numbers of longitudinal and scalar “photons” are equal to zero:
n 3= n 0= 0 .
In this case vacuum may be defined as the state which satisfies the con­
ditions
ca^ 0 , 0 . o.o = 0, 2 . = 1,2,3,0. (16.41)

This definition is convenient because it is symmetric with respect to


the transverse and the longitudinal degrees of freedom of the electro­
magnetic field, but we must keep in mind that it corresponds to a cer­
tain particular choice of the gauge for the potentials.

§ 17. Commutators of the Electromagnetic Field

17.1. Commutation Relations for the Potentials and the Field Compo­
nents
If we know the quantization rules for cfcAand c^A, we can determine
the commutation relations which must be satisfied by the operators
for the potentials and for the field components.
We first of all evaluate the commutator for the potentials
[A^(;c), A„(x:')]. For this we use the expansion (16.13):

X')
K . (*) ’ A*(*')] = £ 7 = 7 ^ few >c*xl eHkz+k'

+ [Cju. cfcw]ei{kx~k'I‘) + [c+,


+ [ c i,c + r ] 7 < ^ ' ) } ew e'f>.
E L E C T R O M A G N E T IC A N D E L E C T R O N -P O S IT R O N F IE L D S 175

In accordance with (16.17) and (16.14') this expression may be rewrit­


ten in the form

[ \ ( x ) , A„(x')] = —
k
On going over from the summation over k to integration over A:-space
in accordance with the formula

% m = £ r S m * -

we finally obtain

[A„(x), A„(x')] = ~ i D { x —x')d^v, (17.1)

where

® W = (2 <17-3

It may be easily seen that the function D(x) is invariant under Lo-
rentz transformations. Indeed, expression (17.2) for D(x) may be writ­
ten as

i W = (2^)5 J (17.2')

where
tj(k) = k0/\k0\ and d4k = dkdk0.

But £(k) is an invariant, since (17.2), because of the presence of the


(5-function, contains only propagation vectors which satisfy the condi­
tion k 2 = 0 , while, as has been noted previously, the sign of the fre­
quency is in this case relativistically invariant.
The invariance of the function D(x) shows that the quantization
rules (17.1) are invariant under Lorentz transformations.
We note some properties of the D-function.
It may be seen from (17.2) that D(x) is an even function of the coor­
dinates, and an odd function of the time
D(x) — D(r, t) = D(—r, t) = —D(r, —t).
176 QUANTUM ELECTRODYNAMICS

On carrying out the integration over the angle between the vectors
k and r in (17.2) we obtain

1
D(x) = ,i£o(r-0 ioi(r-t) gi£o(r+0 g-ia>(r+0)

o
1

4nr {<5(r—0 —<5(r+ 0 }»


i.e.,
D (x) = - ^ ( S (r2- / 2), (17.3)

where £(x) = t/\t\.


Expression (17.3) shows that the function D(x) vanishes everywhere,
except at points lying on the light cone x2 = r2—t2 = 0 ; at these points
D(x) has a <5-like singularity.
From the integral representation (17.2) it follows that D(x) is the
singular solution of □ D{x) — 0, which satisfies

D(r, 0) - 0, — D(x ) | t . 0 = d(r). (17.4)

On differentiating the commutation relations (17.1) with respect


to t', and on setting t' = t in the result, we obtain

0 . -qj = i&u,<5(r—r ) . 07.1')

On differentiating (17.1) with respect to x and x ‘ and on utilizing


the definition of the field tensor, we can easily obtain the commutation
relations for the field components
d2 82
I F „ W , V W = 1! ^ dx/t ox]. dxfldx'v,
d2
—8.-.
"" 8xvdX;,j- +
(17.5)

[E,W . E*0Ol = [H ,M . H ,M ] =
dXj dx'k
d2
[Ey(x), Hfc(x')] = iejkl r D{x—x')\ j, k , l = 1, 2, 3,
dtdx't
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 177

where em is the unit antisymmetric tensor of the third rank. We see


that the field components may fail to commute with each other only
in the case when the corresponding space-time points lie on the light
cone (x—x ' ) 2 = 0 .
As is well known, the commutation relations admit the following
physical interpretation. If the operators L and M which correspond
to certain physical quantities L and M do not commute, but satisfy
the commutation relations [L, M] = G, G ^ 0, then it can be asserted
that there exist no states of the system with well defined values of the
quantities L and M, i.e., with well defined eigenvalues of the operators
L and M. The states of the system can be characterized only by ranges
of values of L and M. The intervals AL and AM, within which these
values lie must satisfy the uncertainty relation AL ■A M ^ |(7|.
Since the commutation relations for fields contain the singular
function D(x), we cannot obtain definite physical results directly from
them. Such results can be obtained, if we average the fields over certain
space-time regions. It can be shown that such averaged values of the
field components commute with each other if the time-like or the space­
like intervals over which averaging is carried out coincide. In particular,
the values of the fields, averaged over the same space-time region, com­
mute. However, if the regions over which averaging is carried out can
be connected by light signals even partially, then the averaged fields
do not commute with each other (31), (87).
In place of c; and c^ we introduce the new variables NAand <px:
(17.6)
Here N; = c f c x is the operator for the number of photons; the operator
<px may be interpreted as the phase operator.
We obtain the commutation relations between the operators NA
and (px. On substituting (17.6) into (16.17) we obtain
(17.6')

This relation will be satisfied if NAand (px satisfy the condition


= ~L (17.7)
Indeed, the following relation can be obtained from (17.7):

[Na, <p;n] = - i n y T 1,
178 QUANTUM ELECTRODYNAMICS

which immediately yields (17.6') if we note that

°° 1
el* = £ - j ( i < p ) n.
71^-0
The commutation relations (17.7) yield /1NA/l<p; ^ 1. The physical
meaning of this inequality consists of the fact that the electromagnetic
field cannot be simultaneously characterized by definite numbers of
photons of different kinds, and also by definite phase relations.
If the difference in phase between two waves has a well defined
value, then only the total number of photons associated with these
two waves can be specified. But the manner in which the photons are
distributed between the two waves remains completely undetermined.
The classical electromagnetic field is characterized by large expecta­
tion values of the numbers of photons. In this case, even for very small
valiles of Acpx, the expectation value of the number of photons (Nf) is
considerably greater than ANX, i.e., the states of the field are practically
characterized both by definite phase relations and by definite intensities
(photon numbers).

17.2. Chronological and Normal Products of Components of the Potential


An important role will be played by products of the operators for
the potentials arranged in a definite order, in the future investigation
of the interaction between the electromagnetic field and electrons.
These ordered products bear the names of chronological (time-ordered)
and normal products.
In the chronological product of the operators A/((x) and Av(x'),
which we denote by T[A/1(x ) A v(x ')), the factors are arranged in chrono­
logical order, viz.: the operator corresponding to the smaller value
of time is placed on the right and the operator corresponding to the
greater value of time is placed on the left. In other words,
t>(.
r(A„(.x)A,.(.v')) (17.8)
I A,(y)A„(x), t < ('.
We note that this definition is relativistically invariant. Indeed, if x
and x' are separated by a time-like interval, then this assertion is obvious,
since in this case the concepts of a later or of an earlier time have abso­
lute significance. However, if x and x' are separated by a space-like
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 179

interval and t t can change sign under Lorentz transformations,


then the operators A /t(x) and A„(x') commute, and, therefore, can be
arranged in all coordinate systems in the order which is required for
the chronological product.
By utilizing the expansion for the potentials (16.13) into plane waves
we can represent the chronological product T(A/t(x)Av(x')) in the fol­
lowing form:

r ( A t(x)Av(x')) = I _ v _ L - " {CJUC*Xg i ( f c i - l - f c 'i ') + CAACA',1' g i ( f c i - f c 'x ')

+ ^ c k-re'i(kl- k'I') k'i') }£«>««> t > t'. (17.8')


On interchanging x and x', we obtain the chronological product for
t < t'.
We now define the normal product of the operators A^(x) and
A„(x'). In this product, which we denote by N{Alt(x)Av(x')), the photon
creation operators c^Aare placed to the left of the annihilation opera­
tors ckX. If we represent the operator A/((x) as a sum of two terms with
positive (A^+)(x)) and negative (A(~\x)) frequencies, then we have
N( A , » A v(x')) = AJ+>(x)Ai+>(x')+AJr)(x)A ^(x')
+ A<~\x) A<+>(*')+A<->(x') A<+>(x)
1
= J_ V — (c C
I AA A A
, g i ( k x +k ' x ' ) _ |_ q +
' A'A
q
AA
gi(kx-k'x')
2 Q k.Vk'.X-
kXCk-Z‘
A remarkable relation exists between the chronological and the
normal products—their difference contains no photon creation or
annihilation operators, and is an ordinary function (c-number). Indeed,
on utilizing the expansion of the potentials into plane waves and the
quantization rules (16.17) we obtain for this difference the following
expression:

r |A ,(< )A ,M )-» (* ,(I)A ,M | = A 2 ’7!re'‘l'"rM"l“ '1V


k

which is valid both for t > (’, and for t < t'.
We call the difference between the chronological and the normal
products of two operators the pairing or the contraction of the two
operators.
180 Q U A N T U M EL E C T R O D Y N A M IC S

By making the transition in the preceding expression for the pairing


from the sum to an integral over £-space we obtain
^ A ^ A ^ - N l A ^ A X x ’)) = = D*(x-x')d^v ,
(17.10)
where the pairing of the operators A ^ x ) and Av(x) is denoted by
A“(^)A“(x), and the function Dc(x) is

(17.11)

We now show that the pairing of the operators A^(x) and Av(x)
defines the expectation value of the chronological product of these
operators in the vacuum state.
In order to obtain the vacuum expectation value of T(A^(x) Av(x')
we must know in accordance with (17.8') the vacuum expectation values
of products of the operators ckX and eft- Since in the vacuum state
the numbers of transverse photons are equal to zero, we have
<c2icM > 0 •— 0, (q jC ^ ),)— 1, A— 1 ,2 ,
where <L) 0 denotes the vacuum expectation value of the operator L.
It is obvious that for these polarizations the following relations also
hold:

(17.12)

As regards the longitudinal and the scalar polarizations, if that gauge


is chosen for the potentials which corresponds to the definition of the
vacuum (16.41), the numbers of longitudinal and scalar photons in
the vacuum state will also be equal to zero. However, here we shall
not utilize this gauge for the potentials and, in contrast to subsection
16.5, we shall not even specialize the definition for the expectation
value in the state <Z>34, but shall assume that the vacuum expectation
value of the operator c^ 3 ck,3 has been defined in some way:
< c £ |C w > o = ( f c c (3>)«Z (fe )d M . (17.12')
where %(k) is, generally speaking, some arbitrary function of k. More­
over, we shall assume that for all A the following relations will hold:

<(CUCft'V/'0 = ( CUCfc'A'\ = 0 -
E L E C T R O M A G N E T I C A N D E L E C T R O N - P O S I T R O N F IE L D S 181

From relation (16.28), which we can use since it is based only on the
subsidiary condition (16.23), and also from formula (17.12') it follows
that
^cfe4cfe4^o = ( cfc3cfc3/’o= .
Further, it follows from (16.24), (16.24'), and (16.14) that
<c*3c*4>o = <ct4c*3>o = »<Cjkc*3>o = i(ke™Yx(k) = (kew ) (ke^)x(k).
All these expressions, and also expression (17.12), valid for X = 1 , 2,
can be written in the form
<c^cfc.r >o = (kew) ( k ' e ^ ) X(k)dkk,, <cMc* - r > 0 = 0 ,
<^c+r > 0 = {6x, + { k e ^ ) {k'e^)x{k)} dkk,, <c&c+ A, > 0 = 0, (17.13)
A;A' = 1 ,2 , 3, 4.

On utilizing these formulas we obtain, in accordance with (17.8'),

<r(A ,W A .(x'))>, = - £
k, A, A'

+ -ir V —■(kew ) ( k e u'))cosk(x—x')e^)eY">x(k),


Q A M
or, since = dpv, we have

<r(A„M a .m ) > , = dr —
k

+ -q - ^ cos fc(x-* ')* (* )•


k

Here the first sum is equal to Dc(x—x'), while the second sum can
evidently be written in the form (d2/dx/idxv)q)(x—x'), where

<p(x) = - - ^ — cos k xxik ).


k

Therefore, we finally obtain the following expression for < r(A /i(x:)Al,(x:'))>0,
which is valid both for t > t', and also for t <

^ ( i - x ) = ( r | A , ( x ) A , M ) > 0= Dc( x - x ' ) d /n,+ dx-d- <p{x-x).


(17.14)
182 Q U A N T U M EL E C T R O D Y N A M IC S

The second term containing the arbitrary function cp(x) is due to the
arbitrary gauge for the potentials. If we use the gauge in which N 3 =
Nt = 0 then this term will vanish. It is clear that no physical quantity
can depend on the arbitrary function cp(x). This is the meaning of the
property of gauge invariance.
On repeating the calculations just given it can be easily shown that
the vacuum expectation value of the normal products of the opera­
tors A m(x) and Av(x') is equal to

<^( A ^ W A ,^ ))), = (17.15)

We expand the function Dc(x) into the four-dimensional Fourier


integral:

D°(x) = -J Dc(k)e*kxdik, (17.16)

in which all four components of k are independent and d*k = dkdk0.


In order to do this we use the following integral representation of the
function (l/a>) exp(—/a>|r|):

(17.17)
c
where k 2 = k2—k\ = co2—k\, and the integration is carried out in the
complex plane of k 0 along the contour C shown in Fig. 8 . In order to
show the validity of formula (17.17) we note that for t > 0 the contour
of integration can be completed by the semicircle of infinite radius
situated in the lower half-plane. Then the integral will be determined
by its residue at the pole k 0 = co and will be equal to the left hand side
of (17.17). However, if t < 0, then the contour of integration must
be completed by the semicircle in the upper half-plane. In this case
the integral will be determined by the residue at the point k 0 = —co
and will again be equal to the left hand side of formula (17.17).
Instead of integrating in (17.17) along the contour C which passes
above the pole k 0 = co and below the pole k 0 = —co, we can integrate
along the real axis of k 0, by displacing the former pole into the lower,
and the latter pole into the upper half-plane (Fig. 9). For this we must
replace the denominator k 2 of the integrand by k2—irj, where rj is
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 183

an infinitesimal positive number, since in this case the poles will be


given by the points k 0 = ±a>(\—(ir]/2a)2)) situated as shown in Fig. 9.
On substituting formula (17.17) into (17.11) and on comparing
it with (17.16) we obtain the Fourier component Dc(k):
1
Dc(k) (17.18)
i(k2—i0) ’
where iO denotes i\rj\ for \rj\ ->• 0. This expression can also be written
in the form
Dc(k) = nd+{ - k 2), (17.18')
where the singular function <5+(x) is defined by the integral
OO
<5+00 = — (e^+^dZ= —— L -, M - 0 . (17.18)
71 J 71 X " p ZU
0

*0 kQ
-c a Cs ^ - co +,•o
+ LU
^ “fO

Fig. 8. Fig. 9.

We note that <5+(x) = d(x)+(i/n)P(\/x). This equation, which follows


from (17.19), means that integrals containing the (5+-function are to
be evaluated according to the following rule:
+ oo _ +~
f f ( x ) S +(x)dx = J / W - L - * = / ( 0 ) + - L p j ' f W - ~ , (17.19')
-C O C -O O

where the contour of integration C coincides with the real axis, but
passes above the pole x = 0, while P denotes the principal value of
the integral.
On substituting (17.18') into (17.16) we obtain
OO

Dc(x) = ——-- f dikeikx f e~^(k'~llr,l>d£


J %J J

OO
x' +i'M
1
d£e e-ik'tsd*k'.
(2 ^
0
184 QUANTUM ELECTRODYNAMICS

where k' = k —(x/2£). On noting that

f e ± i ^ da==

and that, therefore,

f e-*™ d*k' = (ti2//£2),


we obtain

0 0

i.e.,

(17.20)

If we know Dr(k), we can easily obtain the Fourier component


Dc^v(k) of the function D^ipc) defined by formula (17.14). Since differ­
entiation of any function with respect to xMcorresponds to the multi­
plication of its Fourier component by ik , we can write DcMV(k) in the
form (115)

(17.21)

where <7,(k2) is an arbitrary function of the invariant k 2 corresponding


to the arbitrary function <p(x). On setting d ^ k 2) = 1 in this expression
we obtain DcMV(k) = Dc(k)dflv. When dt(k2) = 0 the function Dc/iv(k)
will have the property of “ transversality” :

W /.= °, dt( k * ) = 0 ,
or in the coordinate representation

o.

It has been pointed out previously that physical quantities do not


depend on the arbitrary function <p(x). Therefore we can say that the
property of gauge invariance of physical quantities means that they are
independent of the quantity d ^ k 2).
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 185

We note, that since k2Dc(k) = —i, the function Dc(x) satisfies the
equation
□ Z)c(x) = id(x). (17.22)
Thus, Dc(x) is the Green’s function of the wave equation.

17.3. Singular Functions Associated with the Operators □ and (□ —m2)


We have discussed the two singular functions D(x) and Dc(x) asso­
ciated with the electromagnetic field. The function Dc(x) can be ex­
pressed in the form of a four-dimensional integral

<m 3 >
where the integration over k 0 is carried out over the contour C0 shown
in Fig. 10.

Fig. 10.
The function —iD(x) can also be represented in the form of a four­
dimensional integral with the same integrand as in Dc(x), viz.,
1 _ eikI
D(x) -dik, (17.24)
~k*
186 QUANTUM ELECTRODYNAMICS

where the integration over k0 is carried out along the contour C0 shown
in Fig. 10. Indeed, in carrying out the integration over k 0 with the aid
of the residue theorem we obtain
i r r eifcr+(a,f giki twt i 1 r sin cot ,,
D(- » = ~ x w f dk --------- 2 ^ r \ = '* •
We obtain other singular solutions of the wave equation associated
with the electromagnetic field if we investigate the integral
. 1 r dik
F(x) = / e‘ 1 ’
£
in which the integration with respect to k 0 is carried out over the contour
J2 which coincides with one of the contours C+, C_, Cu CR, CA, shown
in Fig. 10. Moreover, if J2 surrounds one or both poles of the integrand,
then F (x) will be a solution of the homogeneous wave equation, while
if the contour extends to infinity we shall obtain a Green’s function
of the wave equation, i.e., a solution of the inhomogeneous wave equation
with a ^-function on the right-hand side.
We first of all discuss the singular functions D± (jc) for which J2
has the form of the contours Cj_ or C_:

r eikx
(2 jt ) 4 J
J k2
d*k. (17.25)

Integration with respect to k0 yields


1 dk
D,.(x) =
2 i{2 7 if / co
(17.25')
1 dk
D- i x) = 2 i (2
2 jc) 3i e ~ ■-*
From this it follows that with k 0 = co\k\,
D _ ( x ) = - D h( - x ) . (17.25")
It is clear that
D(x) = D+(x)+D_ (jc) . (17.26)
We now show that the functions D ±(x) define the expectation
values of products of the components of the potential in the vacuum
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 187

state. By making use of the expansion of A^(jc) into plane waves, and
relations (17.13) we obtain

<A„(x)A,(*')>, = A y
kjlv 0)

~\~~2 o —'Xft) (kew ) (ke^X)) {eik(-x- x')Jt-e-ik<x- x">} e^e^'K


Z ft. A. A' M

On noting that = dMV, and on going over from summation to


integration over k, we obtain

<AM(x) A„(x')>o = - i D +( x - x ' ) d tiv+ — — (p ( x - x ') ,

<A„(aO A „ (*)>0 = i D ^ { x - x ' ) d llv+ - ^ - j - - ( p { x - x ' ) .

From this it follows that the vacuum expectation value of the anti­
commutator (A ^ x ), A„(x')} = A ^ A / x O + A / x ^ A ^ x ) is equal to

({A„(*), Av(x')})o = D f a —x ^ d ^ + l dx ( p ( x - x '\ (17.27)

where
A W = - i { D +( x ) - D ^ ( x ) } . (17.28)

Evidently, the function D1(x) can be represented in the form of


the integral

D^ = W r J P ‘k- <17-2 8 '>

where the contour C1 is shown in Fig. 10. Integration over k0 with


the aid of the residue theorem yields

- ( k > f d k { ^ e‘'‘r+‘" ‘ + = ( k ’ f e“‘rC^ r dk-


(17.29)
This expression can be rewritten in the form

A W = (2 ^ / ^ d Q c ^ k . (17.29')
188 QUANTUM ELECTRODYNAMICS

On carrying out in (17.29') the integration over the angle between


k and r we obtain
CO

D^ x) — ^ {eim{r+t) — e-ia)(r+t)-heia)(r- t)—e - ia>(r~n}dco


2 1

(2 jrp r 2- ? 2’
i.e.,
2 1
Di(x) = (17.30)
(2 ^ x^-

We see that the function D ^ x ) , like the function D(x) , is a singular


solution of the wave equation, but in contrast to the latter vanishes for
no finite values of x2.
On comparing the integral representations (17.23), (17.24) and
(17.28') for the functions Dc(x), D{x), and D^x) we can express Dc(x)
in terms of Dx(x) and D(x):
D%x) = i { D y{x)-i£{x)D{x)}, (17.31)
where
t> 0 ,
t < 0 .
We also give the expression for Dc{x) in terms of the functions
D ± (x) :

D%x) = ± { d ( x ) D +( x ) - d ( - x ) D _ ( x ) } , (17.31')

where d(x) = | ( l + ^(x)).


We now consider the function
1 r p i kx

( 1 7 -3 2 )
CR
where the integration over k 0is carried out along the contour C^fFig. 10).
When t > 0 , the contour of integration can be completed by a semi­
circle in the lower half-plane, and we then obtain the function D (x).
However, if t < 0 , then the contour of integration can be completed
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 189

in the upper half-plane, where the integrand has no poles, and we obtain
zero as the result. Therefore,
D(x), t > 0 ,
Dr (x ) = (17.32')
, t <o. 0

On applying the d’Alembertain operator to (17.32) and on making


CR coincide with the fc0-axis we obtain

□ D*(x) = - — j- je***d*k = -<5(jc). (17.33)

Similarly, it may be easily shown that the function

<17-34>
CA

where the integration with respect to k0 is carried out along the contour
CA (Fig. 10), reduces to
( 0, t> 0,
= { -/)(* ), <<0. (17'34,)
and satisfies the equation
D D a (x) = - d ( x ) . (17.33')
It follows from (17.32') and (17.33) that the solution of the inho­
mogeneous wave equation I34(x) = —j(x), which vanishes for t = —oo,
can be written in the form
A(x) = f DR(x—x')j(x')d4x'; (17.35)
i.e., the function DR(x) is the Green’s function of the wave equation
which leads to retarded solutions. Similarly, we can easily show that
D a (x) leads to advanced solutions.
In the future study of the electron-positron field we shall need
a number of singular functions which bear the same relation to the
operator □ —m2 that the functions which we have just discussed bear
to the d’Alembertian operator. These functions are defined by the
four-dimensional Fourier integrals
1 f 1 C eipx
A(x) = Ay(x) d*p,
(2ny J p 2+ m 2 P ’ (27i)Ai c,J p 2-\-m2
C0
(17.36)
1 f elpx 1 r eipx
A%x) =
(2 ri)Hj p 2+ m 2 ^ P ’
Zl±(x) (2jt)4J p 2+ m 2 <Pp,
190 QUANTUM ELECTRODYNAMICS

where the integration with respect to p0 is carried out along the con­
tours C0, C l t C, shown in Fig. 10. On applying to these expansions
the operator □ —m2, and on making the contours coincide with the
/>0-axis, we find that the functions A(x), A ^x), A c(x) satisfy the follow­
ing equations:
(□ —m2)A (x) = 0,

(□ - m ^ A ^ x ) = 0, (17.36')

(□ —m2)A c(x) = i6(x).


On carrying out in (17.36) the integration with respect to p0 with
the aid of the residue theorem we obtain representations of the functions
A(x), A x{x), A c(x) in the form of three-dimensional Fourier integrals.
For example,

= - w /" {
(17.37)
where e = ]/p2-\-m2 .
In a similar manner we obtain
. .. 1 C . coset ,
( 1 7 ' 3 8 )

<17-39)
These formulas differ from the corresponding formulas for the functions
D(x), Dt (x), Dc(x) by the fact that e appears in them in place of a>.
On comparing the expansions just obtained we can easily show that

■4%x) = - {At (x)-i£(x)A(x)}. (17.39')

It is evident that all these functions are invariant under Lorentz


transformations, and therefore depend only on x2 for x 2 > 0 (forx 2 < 0
these functions depend on two quantities: x 2 and f(x)). From this it
follows, in particular, that since the function A{x) vanishes for t = 0 ,
then it vanishes everywhere outside the light cone when jc2 = r2—t2 > 0 .
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 191

By utilizing (17.37) and (17.38) we can easily show that the functions
A(x) and A 1(x) may be written in the form

A(,x) = -( - )3 f eipI^(p)d(p2+ m 2)dip,


(17.40)
= - J 2 / eivxd(p2+ m 2)dip.

Thus the Fourier components of the functions A(x) and A x(x) are
given by
A(p) = 2ni£(p)5(p2-\-m2),
, (17.40')
Ai (p) = 2nd{p2+ m 2).
The Fourier component of the function A c(x) can be obtained
in the same way as the component of the function Dc(x) if we displace
the + £ pole into the lower and the —s pole into the upper half-plane,
and then integrate in (17.36) along the real axis. Such a displacement
is evidently equivalent to a replacement of p2 by p2—i\rj \ where \rj \ 0 .
From this it follows that the Fourier component of the function A c(x)
is equal to

By utilizing formula (17.19) it is possible to obtain in a manner


analogous to the way this was done for the function Dc(x) the following
integral representations for the functions A(x), A x(x), Ac{x)\

^ i( * ) = 2 ~ z j sin{x2z — ^ j d z , (17.42)
0

0
*•
We now show that the functions A c(x),A(x), and A t (x) can be
expressed explicitly in terms of Bessel functions. We start from the
integral representation (17.36) of the function A c{x). Since this function
192 QUANTUM ELECTRODYNAMICS

is relativistically invariant and, consequently, depends only upon


x2 = r2—t2, it is sufficient to discuss the two cases when t = 0 and
jt2 > 0 and when r 2 = 0 and x 2 < 0. In the first case on carrying out the
integration over the angle between p and r, and then on setting
p = msinhy, e = m c o s h y we bring the integral (17.39) into
CO
wi r
A c(x) ~-g—2 t _ I eirmsinh ysinhy d y .

Further, on utilizing the following representation for the Hankel


function (204),

H<2)(r) = ^- J e-fcosmijy (17.43)
— 00

and setting r -*■ —ir, y -» y —ijt/2 in this expression, we obtain

J
00

H!>2)(—ir) = ~ eiTSlDh!/dy.
—CO
A comparison of this formula with the one given earlier for A c(x)
shows that

But

and therefore
. . im2 H[2){—imr) im2 H[2)(—im ] / x 2)
A c(x) = -o------ — ------ = ------ — ------T = - ~, X2 > 0 .
67i —imr Sn —im ]/x 2
When x -> 0 the term (i/4ji)d(—x 2) arising from the differentiation
of the function H^2) must be added to the right hand side of this expres­
sion. Indeed, when \z\ << 1 we have
= J0(z)—iN0(z) & -(2i/ji) ln(2/yz).
and since

—7tid(z),
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F I E L D S 193

the additional term

will appear in A c(x) when x 2 -> 0. Thus, when x 2 > 0 we have

4 -M - ^
v 2 ~ i m j/x 2 1

If x 2 < 0, then we can take r = 0 in (17.39); on setting


p = m sinh z, e = m cosh z, we obtain

A e-im|t|oosh»2sjnJ12z

This integral can be reduced with the aid of (17.43) to

A % x ) = w { ‘^ w ®(m |<l)+m 2/,“ (",|<l)}’


and since

it can be easily shown that we again obtain formula (17.44) which is,
consequently, valid for arbitrary x2.
If we have an expression for Zlc(x), we can find with the aid of
formula (17.39') the functions Zl(x) and ^ i(x ). On separating in (17.44)
the real and the imaginary parts we obtain

-iw A = 2 ^ 4n _ im y x 1
m2 Kx(m ]/ x2)
x2 > 0 , (17.45)
2 ^ 2 m [/ x 2
A W = ■
rri2 (m j / - x 2)
x2 < 0 .
4^ 2 m ( / —x 2
When x 2 -> 0 these formulas assume the form

(17.45')
1 m2l
----- 2 + m 2ln
AW = 4 ^ 1 x2 ”T f
where y = 1.781.
194 QUANTUM ELECTRODYNAMICS

We also give expressions for the functions A ±(x), A R{x), and


A a {x), which are analogous to the functions D±(x), DR(x), DA(x)\
1 r eipx
(2 n j i j p 2+ m 2 d *P ’

1 C eipx
j T + iS r * * (17'46)
CR

1 r elpx
(2^J 7 + ^ d' p -
CA
where the contours of integration C±, CR, CA over p0 are shown in
Fig. 10.
It can be easily shown that

-M * ) = T - a & y f e*“’ ~ r = i ( J W ± (>7-47>


From (17.47) it can be seen that similarly to (17.25”) we have
A +(x) = - A _ ( - x ) . (17.47')
The function Ac{x) can be expressed in terms of the functions
A, (x) as follows:
A c(x) = --{O(jc)^l + (x )-0 (-x )2 l_ (x )} . (17.48)

By repeating the calculations given earlier with reference to the


function DR(x) we can show that
A (x), t > 0,
A a(x) (17.49)
0 , t < 0
and also that
C-3—m2) A R(x) = —d (x). (17.49')
In conclusion we note that all the functions considered above
can be expressed in terms of one analytic function. In order to do
this we consider the function A +(x). From (17.36) it can be easily
seen that the integral defining it is still meaningful if x is replaced
by z = x — ir], where rj is a four-vector for which r f < 0, rj0 > 0. This
means that A + is a function of z 2 analytic in the plane of this vari­
able with a cut along the negative real semi-axis ( —oo, 0 ) (81a, 2 1 1 a).
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 195

Moreover, the region x 2 > 0 corresponds in the z 2 plane to the


positive real semi-axis, while the regions x 2 < 0 ,x 0 ^ 0 correspond,
respectively, to the upper and lower edges of the cut (z2 = x 2 + / 0 ).
Indeed, from the preceding formulas an explicit expression for A +
follows:

. A M = F (A = — i K + ( m / z % f 7 \ (17.50)

The remaining functions can be expressed in terms of A +, if rela­


tions (17.47), (17.47'), (17.48) are utilized. In particular, we obtain
iAc(x) = F(x 2 + ;0), (17.51)

- ~ Z l( x ) = F(x 2 + /0 ) - F ( x 2 —z0). (17.52)


]*o]

§ 18. Quantization of the Electron-Positron Field

18.1. Variational Principle for the Dirac Equation. Energy-Momentum


Tensor o f the Electron-Positron Field
In Chapter II we studied the properties of a single electron. In
order to study an arbitrary system of noninteracting electrons and
positrons we can, just as in the case of a system of photons, go over
from the description by means of the wave funetion in configuration
space to the description by means of the wave function in particle
number space. In the case of a system of photons such a transition
is equivalent to the quantization of the electromagnetic field. In the
case of a system of electrons it is equivalent to the quantization of
the electron-positron field, i.e., of the field of the wave functions ip (x)
and ip (x) satisfying the Dirac equations. In this case the Dirac equations,
like Maxwell’s equations, play the role of field equations, and not of
equations for an individual particle, while the quantization of the
electron-positron field means that the wave functions ip(x) and y(x)
are regarded as operators satisfying certain commutation relations
and operating on the state vector, or on the wave function of the system
in particle number space. As regards the dependence of these opera­
tors on the coordinates and the time, it is determined by the Dirac
equations.
196 QUANTUM ELECTRODYNAMICS

In order to emphasize the field nature of the Dirac equations and


to be able to treat them similarly to the equations of the electromag­
netic field, we show, first of all, that the Dirac equations can be
obtained from a variational principle: S f L d i x — 0, if for the Lagrangian
density we take

L= — C18-1)

and in varying L regard the functions ip and ip as independent.


On treating ip, ip and dip/dx^, dip/dx as generalized coordinates
and velocities of a certain dynamical system—the electron-positron
field—we obtain from the variational principle the following Lagrangian
equations:
dL dL
tep d[dipjdx/t] dn- = 0 .
- v
(18.1')
d dL___ dL
— = 0, A = 1 , 2 , 3, 4.
dx^ d[dipJdxM] dV>.
On substituting into these expressions
__5L_ = _ 1 _ dL
d[dipjdx/t] ?2 V
TV,
Mlv,n,,
vh "d[dyjdxj -- j f y X V ’, ’
(18.2)
dL_ 1 dipfI dL 1 I x dip
2 dx,
or, in abbreviated form,
__ dL________ 1 _ dL
W /^ J = 2 r y >1’ W /^ ,]
dL 1 dw _ dL dip
mip,
2 Yv dx
we obtain the Dirac equations

(y" 3 ^ + y = °, = ( 18-3)
If we know the Lagrangian density, then we can, just as in the case
of the electromagnetic field, define the energy-momentum tensor of
the electron-positron field.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 197

We begin with expression (16.6) for the energy-momentum tensor


of the electromagnetic field. As can be seen from the derivation of
this expression, it can be applied directly to define the energy-mo­
mentum tensor of the electron-positron field, if in it we replace the
components of the potential A by the components of the Dirac functions
tpfl and yj which we regard as generalized coordinates of the electron-
positron field. Thus, we arrive at the following definition of the energy-
momentum tensor of the electron-positron field:
s l __ <tyx_ dL dtpA
TltP =
d[dyjjdxv\ dxM ^ d[dyjdxv] d x j '
On substituting into this expression formulas (18.2), and on noting
that the functions tp and tp satisfying the Dirac equation make the
Lagrangian density vanish, we finally obtain
dtp
(18.4)

From this we see that the energy density and the momentum density
of the field are respectively given by
1 dtp dtp*
w = ~ T i4 = —
2 / dt dt
(18.4')
dtp dtp*
tp
2i V dr dr
The energy-momentum tensor (18.4) in virtue of the Dirac equations
satisfies the condition (d/dxv) T/lv = 0, and this, as we already know,
implies conservation of the four-dimensional vector

- i f T/tidr. (18.5)
This vector can be interpreted as the four-dimensional energy-mo­
mentum vector of the electron-positron field.
The total energy and momentum of the field are obviously equal
to integrals of w and p over all space:
198 QUANTUM ELECTRODYNAMICS

We note that the tensor T ^ is not symmetric, but that we can con­
struct the symmetric tensor 0 ^ = \ { T tlv+ T v^), which, like the tensor
T , satisfies the equation of continuity (dOMJ d x v) = 0.

18.2. Quantization Rules for the Electron-Positron Field


We now proceed to establish the quantization rules for the electron-
positron field. As has been noted earlier, we must, just as in the case
of the electromagnetic field, regard the components of the field y> as
operators which operate in particle number space, and which obey
definite quantization rules. However, these rules must differ from
the rules for the quantization of the electromagnetic field (16.17).
Indeed, from the latter it follows that the number of particles in a def­
inite state may be arbitrary, while the Dirac equation describes particles
which obey Fermi-Dirac statistics, and therefore their number in any
given state may be equal either to unity, or to zero (the Pauli exclusion
principle).
In order to quantize the electron-positron field, we start with the
general series expansion of y) in terms of the orthonormal system of
functions y){T+) and y)iT~) which are solutions of the Dirac equations with
positive and negative frequencies respectively. These may be solutions
which correspond either to free motion, or to motion in an arbitrary,
but sufficiently weak, stationary field. In a strong field it is impossible
to carry out the separation into positive and negative frequencies
(cf. § 12). The wave functions with positive frequencies correspond
to electron states, while the wave functions with negative frequencies
correspond to positron states. We write the expansion of rp in the form

V>= £ ary>‘+ , +J£b+y>‘~),


; _ ' (i8.7)
V = X, K v \ +) + ’>
7 T

where yjlT±) are normalized in accordance with

f v tr’,>V><
r?'>md r = fir r ‘#w'> V , rf = ± >
and we assume that the expansion amplitudes ar, a+, br, b+, are not
ordinary numbers, but operators, just as we did in the case of the
expansion amplitudes ckx in (16.13) for the potential of the electro­
magnetic field.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 199

We assume that ar and a+, and also br and b+ are Hermitian con­
jugate operators which satisfy the following commutation relations:
{ar ,a+} = ^rr., {br,b^} = (5rr,,
ar-} = {a+, a+) = 0 , {br, br,} = {b+ b+) = 0 , (18.8)
{ar, M = {ar, b+} - {a+ br,} = {a+ b+} = 0 ,
where {A, B } = A B + B A is the anticommutator of the operators
A and B. These relations, which are the quantization rules for the
electron-positron field, lead, as we shall see later, to the correct cor­
puscular picture of the field.
By making use of the quantization rules we determine the form
of the operators ar, a+, br, b+. We first of all find the eigenvalues
of the operators
n<+ )= a + a r, and n<_ ) = b + b r.
On squaring n<.+) and n<~>, and on noting that a2 = b 2 = 0, we
obtain
n ‘+>2= a+ara+ar = a+ar( l - a ra+) = a+ar,
n<-)2= b+brb+br = b+br( l - b rb + )= b+br,
i.e., n < + ) 2 = n<+), n < - )2 = n(r_), whence it follows that the operators
n(r+) and n jr' have only two different eigenvalues equal to unity and
zero.
We make use of the representation in which the matrices n(r+) and
n jr’ are diagonal; then the matrices ar, a+, br, b+ can be taken to
be two-rowed ones, and we can write
n, /=
k + ) ) ii = E k ) i f c ( a r)fci = E I f o ) * ! I 2 1 0 , i= 2
*= 1 *= 1

(it is assumed that the eigenvalue h<+) = 1 corresponds to the subscript


/ = 1 , while the eigenvalue n<+ ) = 0 corresponds to the subscript i = 2 ).
Since a 2 = a + 2 = 0, i.e.,
2

( a r)(fc = E ( a r ) i l ( a r) j k = 0 >
1=1

(a+2).* = l ( ar+)«(a^ * = °.
1=1
200 QUANTUM ELECTRODYNAMICS

we finally obtain
ar+ v t at>
(18.9)
0, 0
Y+ --
1, 0

where r]r is a certain matrix which does not operate on the variables
of the rth electron state and which satisfies the condition r]+rjT= 1 .
By proceeding in a similar manner we can show that the matrices
br and b+ are of the form
br = £ rar , b j = £+a+, (18.9')
where £r is a certain matrix which does not operate on the variables
of the rth positron state, and which satisfies the condition £+£r = 1 .
In order to obtain the form of the matrices r}T and £r we arrange
all the electron and positron states in a definite (arbitrary, but fixed
once and for all) order, and we label these states uniquely by the subscript
r. The matrices rjTand £Tmust be chosen in such a way that the following
anticommutation relations hold:
{ar, ar,} = {br„, br^} = {ar , br„} = 0 .
For this it is sufficient to assume that the matrices £r and r]T (we shall
in future denote both kinds of matrices by £r) operate only on the variables
of those states which precede the state r, i.e.,
t r = t r(l, 2, 1),
and which have the form of the product
Cr = t(l) t(2) ... £(r—1), (18.10)
where £ (n) operates only on the variables of the state n.
We now consider the anticommutator
K > ar+i} = £ Ta.T£ T + 1 a T+ x + £ T + 1 a.r + 1 £ r a T = 0.

Since the matrices ar and a r + 1 refer to different states, i.e., operate


in different spaces, they commute with each other. Therefore, the
anticommutator {ar, ar+1} can be rewritten, on taking (18.10) into
account, in the form

aT£(r)-\-£(r)aT= 0 . (18.11)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 201

Thus, the matrix £(r) anticommutes with ar; similarly it can be easily
shown that £(r) also anticommutes with a+. From (18.11) it follows
that
[£(r),a+ ar] = 0.
Therefore, the matrix £ (/') may be assumed to be diagonal in the rep­
resentation in which the matrix a+ar is diagonal, and we may write

K k (CWlii + (COOU(O21 = 0 ,
i.e.,
(C00)n = -(tO O U ,
where the subscripts 1 and 2 correspond to the eigenvalues of nf
equal to 1 and 0. This relation, together with the relation £+£r = 1
shows that £ 0 0 has the following form:

£ 0 0 = (—i)d^ dr= l - 2 d + d r,
where dr denotes both the operators ar, and the operators br. By making
use of (18.10) we finally obtain

£r = n ( l - 2 d+dr,) = T i (1—2n<±>), (1 8 .1 2 )
t' --1
where n<-> = d£dr. We now show that this quantity represents the
number of particles (electrons or positrons) in the state r.
Thus, £r is equal to +1 or —1 depending on whether the number
of occupied electron and positron states preceding the state r turns
out to be even or odd.
We now show that the quantization rules (18.8) lead to the correct
corpuscular picture. In order to do this we obtain the energy of the
electron-positron field. This quantity, like the field components, is
an operator defined by means of formulas (18.6) and (18.7) in which
the operators ar ,a + ,b r ,b £ obey the quantization rules (18.8).
By making use of the orthonormal properties of the system of
functions we can easily show that the field energy is given
by
H= £ e< + ' a + a T- £ s ^ b r b + ,
T T

where e<+ ) = w<+ ), = —co<-> and w* are the frequencies of the


corresponding solutions of the Dirac equation. By introducing into
202 QUANTUM ELECTRODYNAMICS

these expressions the quantities n‘+ ) = a+ar and nj.-) = b+br, we


rewrite H in the form
H = ^ £'+>n '+ > + i; £<-> n<-> - £ £<->. (18.13)

This expression is not positive definite. In other words, the energy


of the electron-positron field, both in classical theory, and also after
the quantization rules (18.8) have been imposed on the quantities ar
and br, can assume both positive and negative values.
However, we can define the vacuum state of the electron-positron
field in such a way that the energy of the field with respect to the vacuum
state is always positive definite. This state is defined as the state of
lowest energy, in which all the numbers n(~ T+) and n(r“ are equal to zero:
<a+ar>0 = 0 , <b+b r > 0 = 0 ,
(18.14)
<ara+) 0 = l , <brb+>0 = l ,
where <L) 0 denotes the vacuum expectation value of the operator L.
According to (18.13) the energy of the vacuum state is equal to
E0 = —^ e^~), where the summation is taken over all the individual
r
states of the electron with negative frequencies. Therefore, the energy
of the quantized field with respect to the vacuum state is given by
E = H - E 0= Ze'+ W + '+ Sei-'rt-'. (18.15)

We see that the numbers n(r+) and n(r_), which can have the values
0 and 1 , appear in the expression for the energy of the field in the form

of factors multiplying the energies of the electron and the positron.


Therefore, these numbers should be interpreted as the numbers of
electrons and positrons having respectively the energies e(r+) and £(r_).
Thus, the quantization rules do indeed lead to the correct corpuscular
picture, since the energy of the field is equal to the sum of the energies
of the individual particles—the electrons and the positrons.
We note that in accordance with (18.15) the energy of the electron-
positron field attains its minimum value (equal to zero) for n(r+)
= ni-'>= 0 only in the case when £(r+) > 0, e*-* > 0. These inequalities
hold both for free electrons, and for electrons in a weak external field.
In a strong external field, as we have already seen in § 12, when the depth
of the well exceeds a certain value V'0>a negative energy e(+) < 0 can
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 203

appear. In this case the energy of the field will be a minimum, and
equal to £(+) not when all the n(r+) vanish, but under the condition
that the state of energy e(+) is occupied, i.e., = 1. In this state
the field has the charge e. When V0 > Voc the energy e(+) becomes
less than —m, and a state of the field with negative energy and of zero
charge is possible, in which in addition to an electron of energy e(+) < 0
there also exists a positron of energy e(_) > 0. Moreover, the positron,
for which the external field is repulsive, can escape to infinity, while
the electron charge will cancel a part of the charge which gives rise
to the external field, and will therefore reduce it. These considerations
lead to the conclusion that the external field cannot be too strong
(the depth of the potential well cannot exceed Poe in the example of
§ 12). The complete solution of this problem requires a more detailed
investigation of the interaction between the electron and the source
of the external field, taking into account the energy and the charge
states of the latter. But this already goes beyond the limits of applica­
bility of the very concept of an external field.
We now show that the matrices ar(br) and a+(b+) have a simple
physical meaning, and represent annihilation and creation operators
for an electron (positron) in the state r. Indeed, if we denote by &n<.+)
T

the wave function (or the state vector), which describes n(T+) electrons
in the state r, we can write

(18.16)

In a similar manner we can show that bT and b+ are annihilation


and creation operators for a positron in the state r.
We write the nonvanishing matrix elements of the particle annihila­
tion and creation operators as
(a r) n (r+)- l. n (r+) — £ r j/ w < + ) ,

(a J)n<r « >+ l.n<r+> = fr |/n</r+ T ,


(18.17)
—1, j / nSr \
( b ^ ) nH + l, n (r > £r ] / f l(r * + 1•

If we write the operators tp and tp in the form


y>= a + b +, y> = A + + B , (18.18)
204 QUANTUM ELECTRODYNAMICS

where A and B+ comprise the terms in the expansion of the operator


ip which respectively contain positive and negative frequencies, while
A+ and B play a similar role in the case of the operator ip, then in accord­
ance with (18.17) we can conclude that A and B are electron and pos­
itron annihilation operators, while A+ and B+ are electron and positron
creation operators. In other words, ip is an electron annihilation opera­
tor and a positron creation operator, while ip is an electron creation
operator and a positron annihilation operator.
We return to the expansion (18.7) and examine in greater detail
the very important case when the external field is absent. In this case
we can start with the expansion of ip into plane waves, describing states
of definite momentum p and polarization r:

(18.19)
777T Z j {a-t(p)ur{p)e-ipx-\-br(p)vT( - p ) e ^ 1},
,2

where the spinor amplitudes u(p) and v{—p) satisfy the equations
(ip+m)u(p) = 0 ,
(—ip+ m )v(—p) = 0
and the orthogonality and normalization conditions
4

E K (p )Uq(p)* = dTT',
e=i
4
(18.19')
2 vT
Q( - p ) v re ( - p ) * = 5rr,.

We note that without loss of generality we can assume that


the solution wr( -p )e x p (—ipx) is charge-conjugate with respect to
uT(p) exp (ipx) , i.e.,
v { - p ) = Cu(p), v(—p) = C+u(p).
The energy and the momentum of the free quantized electron-pos­
itron field with respect to vacuum can be represented as

p \ r = 1,2
(18.20)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 205

where and n(pf are the numbers of electrons and positrons of


momentum p and polarization r.
We note in concluding this section that the quantization of the
electron-positron field in accordance with the rules (18.8) is in accord
with the quantum mechanical description of a system of electrons by
means of a wave function which is antisymmetric with respect to the
coordinates and spins of each pair of electrons.

§ 19. Anticommutators of the Electron-Positron Field. Chronological


and Normal Products of Field Components. Current Density

19.1. Commutation Relations for Field Components


We now establish the commutation relations satisfied by the field
operators ip and ip. In order to do this we use the quantization rules
(18.8) for the operators ar,a+, br,b+.
From (18.8) it follows, first of all, that
{Va(*)> Vp(x ')} = 0, {wa(x) ^ p ( x ') } = 0. (19.1)
i.e., the operators tpa and ipp anticommute, and so do the operators ipa
and xpp.
Further, we determine the anticommutator {ipa(x), rpp(x')}. On
making use of the expansion of ip into plane waves, we obtain

{Va(x),V’p(x ')} = JT 1 <(pW j}{p)eip(x~I')


p'. T= 1 , 2

+ £ v*(-P')vT
p(—P')e~iPiT~X)- (19‘r )
p ; r = 1, 2

We evaluate the following sums which appear in the preceding expressions :

-iS + (x -x ')= - y ul(p)ul(p)eip(x- I'\


i> : r = 1 , 2 (19.2)
~ i S - p( x - x ' ) = — £ ‘vra ( ~ P ) ^ ( —P)e' tpix~I')-
p; r = 1 ,2

On noting that in accordance with (10.28) and (10.33)

£ < ( p ) “fi(p) = —
r = 1 .2

^ v ra( - p ) v } ( ~ p ) = — — (ip+m)ap.
206 QUANTUM ELECTRODYNAMICS

we can write the functions S ^ ( x ) in the form

-iS+p(x)

— iS~p(x) = - - i p i:
p

—m
Q 2e
dx.
a0 P

On making the transition from summation to integration over p and


on utilizing the definition of the functions A +(x) we obtain finally

-----m jzl±(x) = ~(-pJr im)A±(x), (19.3)

where p = —iy^d/dx^]. In accordance with (17.47) the functions S ±(x)


can also be written in the form
S&(x)= - i ( i p - m)afl(A ( x ) ± i A x(x )) . (19.3')
On utilizing these formulas, and also relation (19.1'), we obtain the
following relation for the anticommutator {y>a(x) , y>p(x')}:
{%(*)> YfoCO} = ~ i S afi( x - x ' ) , (19.4)
where

S„,(x) = S & M + S ^ x ) = mj M x ) (19-5)

and A(x) is defined by formula (17.37).


We note that from (19.4) the following relation is immediately obtained:
{Va(r » 0 , V > p ( r \ 0 } = d a a <5(r— r ' ) , (1 9 .6 )
in which both operators ipa and ipfi refer to the same instant of time.
The functions S^(x) have a simple physical meaning: they are
vacuum expectation values of the products of the field components
Va(x)y)p(x') and y>p(x’)y>a(x). Indeed, by utilizing the expansions (18.19)
and formulas (18.14), we can easily show that
<V’a(X)V',i(X')'>0 = - i S a d i x - x ) ,
<Vp(x')y’a(x)')0 = - i S - p ( x —x ’).
From these expressions and from (17.47) it follows that
<fya(*)> V/»(*')]> o = S t y i x —x ’), (19.8)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 207

where
d
—m I 4i(x). (19.9)
aft
19.2. Chronological and Normal Products of Operators of the Elec­
tron-Positron Field
We now define the chronological and the normal products of the
operators of the electron-positron field, analogous to the corresponding
products of the operators of the electromagnetic field, which are need­
ed for our subsequent development.
If and <p2(x2) are any two components of the field operators
y> and yj, then their chronological product denoted by T(<p1(x1)(f>2(x2)),
is defined as follows:
Pl(*lVs(*2)> ^y
T(<Pi(xi)<Pz(xj) (19.10)
9>2(*2Vi(*i)> t1 < t2■
We note that in contrast to the chronological produet of operators
of the electromagnetic field the chronological produet of operators
of the electron-positron field changes sign when the operators <px and
cp2 are commuted. This is in accordance with the difference in the com­
mutation rules for the electromagnetic and the electron-positron fields.
The chronological product of an arbitrary number of field compo­
nents cpfxx), (p2{x2), ..., (pn(xn) is defined by the formula
T(<p1(x1)(p2(x2) ...) = d p f p ^ x j t p ^ x j ... <pin(xtn), (19.11)
where the operators q>t (xt^, <Pi2(Xi) ... are arranged in chronological
order, i.e., so that tt > t^ > ... tln and 6P is equal to + 1 or —1
depending on whether the permutation (1, 2, ...)-» Oi, i2, ..., /„) is
even or odd.
The chronological product of field operators defined in this manner
is relativistically invariant. Indeed, if the interval between the points
is time-like then this assertion is obvious, since in this case the inequality
t1 > t2 in the K system leads to the inequality t[ > t2 in the other
system K'. We therefore consider the case when the interval between
xx and x2 is space-like, with the inequality tr > t2 holding in the K coor­
dinate system, and the inequality t[ < t2 holding in the K' coordinate
system. Then in the K system the T-product of the operators ^ (jq )
and <p2(x2) can be written in the form
T(<Pi(x{) <p2(x^)) = (piix^fpfx^), > /2.
208 Q U A N T U M E L E C T R O D Y N A M IC S

It would appear that on the basis of this definition we should write


the T-product in the K' system in the following form :
T (rf i 01) 0 01)) = 0 (0 ) 0 01), t'l < t'2,
where (p[ and <p2 are the values of the field operators in the K' system.
At first sight this expression contradicts the definition of the T-product,
since if in the K 1 system t[ <t'2, then in accordance with (19.10) the
T-product in the K' system should be written in the form
T (<P'i 01) 9>101)) = —<p'201) <p'i 01), <!<*!•
However, actually the two expressions are identical, since for a space-like
interval the operators <p[ and cp2 anticommute.
We consider also the normal product of field operators in which
the particle annihilation operators are placed to the right of the particle
creation operators. If yj = A + B +, y = A++B , where A and B are
respectively the electron and positron annihilation operators, while
A+ and B+ are the creation operators for these particles, then the nor­
mal product of the operators y>1, denoted by N{y)l rp2 ...) is
defined by the distributive law N(y)1y)2(A-\-B+)ip3 ...) = N(y)1if2Ay)3...)
+N(VW 2 B+y>z ...) and by the rule, according to which
N{Z xZ 2 ... Z n) = d pZ i Z i2... Z in, (19.12)
where each of the operators Z x, Z 2, ... is either a particle creation
operator (A+, B+) or a particle annihilation operator (A, B), with
these operators arranged in the right hand side of (19.12) in such a way
that the creation operators are placed to the left of the annihilation
operators and Sp is equal to + 1 or —1 depending on whether the per­
mutation (1, 2 ,...,« ) - > (fi, /2, ..., in) is even or odd.
We give examples of A-products:

N(y(x)y)(y)) = A[(A(x) + B +(x)) (A+(y)+B(y))]


= -A+OOA(*) + A(a-)B(t)
+ B+(x)A+(y) + B+(x)B(;;),
N(yj(y)yj(x)) = A+(y) A(x)+B(y) A(x) (19.13)
+ A+(y)B^(x)-B+(x)B(y),
N(y(x)y>(y)) = y>(x)y>(y),
N[v(x)y(yj) = y>(x)y(y).
E L E C T R O M A G N E T IC A N D E L E C T R O N -P O S IT R O N FIE L D S 209

From this and from the field quantization rules (18.8) it follows that
N (yj(x) y(y)) + N (yj(y) ip(x)) = 0.
We shall show that, just as in the case of the electromagnetic field,
the difference between the chronological and the normal products
of the operators of the electron-positron field, which we shall call the
pairing of these operators, does not contain any particle annihilation
or creation operators and represents a c-number.
We consider the pairing of the operators and ipp(x'). By utiliz­
ing the expansions (18.19) and the quantization rules (18.8), we obtain
— iS+p(x— x'), t' < t,
T (wa(*) WpO')) -N(tpa(x) Wp(*')) (19.14)
iS-p(x-x'), t' > t,
where the functions S ^ ( x ) are defined by formulas (19.2). In accord­
ance with (19.3) and (17.39') this expression can be written in the form

T (Wa(X)^p(X' ) ) - N (y)a(X)V’p(X')) == Sap(x-X'), (19.14')


where

SS,M = - ( y „ 4 - - " < ) * & ■ <19-15)

and A c{x) is defined by formulas (17.36) and (17.39).


By proceeding in a similar manner we can show that the pairing
of the operators y a and and also of the operators xpa and ipis vanishes.
Thus, the following relations hold:
v U xWp(y) = s ap(x - y ) ’
wap (y)K(x) = - s ap(x-y), 16
yZ(x)yp(y) = o,
Wa(x)yap(y) = o,
where <p1<p% denotes the pairing of the operators cp± and y 2 (in place
of the superscript a we shall also use other letters of the Latin alphabet).
Since the expectation value of the normal product of any field opera­
tors (p1 and (p2 is obviously equal to zero in the vacuum state <Ar(991^ 2))o
— o, then the expectation value in this state of their chronological prod­
uct coincides with the pairing of the operators <p1 and <p2

,( ^(9,l992)/>0 — (Pl (P2 ■


210 Q U A N T U M E L E C T R O D Y N A M IC S

In particular, we have
<7’(va(JC)V/jO;))>o= S ^ ( x —y). (19.17)
We expand the function S c(x) into the Fourier integral
S cap(x) = 1/(27iy [ S^p(p)eipzdip.
By utilizing the expression (19.15) for the function S^(x) and (17.36),
we obtain
. (ip—ni) p
Cc
^a/3(P) = 1 y + r r ^ i O (19.18)

This formula will play a particularly important role in subsequent


developments.
Since —(ip—m) ((p-fm) = p 2jrm 2, the function S ^(p ) can be rep­
resented in the following symbolic form:

S c(p) = ——— , m -> m — iO.(19.18')


ip
In this expression we do not explicitly write in the denominator the
imaginary quantity —i\rj\, \r}\ -*■ 0, but we regard the mass m as complex
with an infinitesimal negative imaginary part.
From formulas (19.15) and (17.36') it follows that S c(x) satisfies
the following equation:

(ip-}-m)5c(A) = — /<5(.y), p = — (19.19)

Thus, iSc(x) is the Green’s function for the Dirac equation.


We also note that the function S c(x) can be expressed in terms of
S +(x) and S~(x) or in terms of S(x) and S il}(x) in analogy wtih Zlc(x):
Sc(x) = - i ( d ( x ) S + ( x ) - 0 ( - x ) S - ( x ) ) = l ( S (1)(x)-i£(x)S(x)}. (19.20)
In evaluating the pairings of operators we have expanded ip and
yi in terms of plane waves. However, we could have started with the
general expansion (18.7) of the operator y) in terms of the complete
system of functions y){T+) and which represent the electron eigen­
functions in some stationary external field. Such an expansion has
to be used if the external field cannot be regarded as a perturbation,
but must be taken into account exactly in the electron wave functions.
On repeating the calculations leading to (19.14')'it can be easily shown
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 211

that in this case the pairing of the operators ipa(x) and y>p(x') is defined
by the following general formula:

t' < t ,
S eaf}(x, X') (19.21)
— ’Z Wra)i . x ) ^ { x ' ) , t ' > t .
T

We note that the function iS{.e)(x, x') is the Green’s function for
the Dirac equation for an electron in an external field.

19.3. Electric Current Density


In § 9 we have introduced the electron current density vector s (x).
On multiplying this vector by the electron charge e we obtain the four­
dimensional electric current density vector for a single electron:
j„ = ieVY^P- (19.22)
If in this expression we interpret ip and ip as the operators of the elec­
tron-positron field, then the vector j can be interpreted as the electric
current density associated with the electron-positron field.
We note that the electric current density vector is related by a simple
expression to the Lagrangian of the electron-positron field:
dL dL
J„ = ~ ' e dlSv/dx j ’’1 (19.23)
didvjdxj
Indeed, on substituting into this expression the expressions for the
derivatives dL/d(dipJdxM) and dL/d(dy>Jdx/t) from (18.2), we obtain
(19.22).
Formula (19.22) shows that the charge density q — —iji and the three-
dimensional current density j are defined by the following expressions:
o = eipy^ip = eip+ip, j = ieipyip = eip^aip. (19.24)
As may be easily seen with the aid of the Dirac equations (18.3) they
satisfy the equation of continuity

( 12. 22 ' )
dx (1 jJ/J„ = o.

The total charge Q associated with the field is given by

Q = J Qdr = e j ip+ipdr. (19.25)


212 QUANTUM ELECTRODYNAMICS

On substituting into this expression in place of yj and y> the expansions


(18.7) and on utilizing the quantization rules (18.8) and also the orthonor­
mal properties of the functions y)\.+) and y)ir~) we can easily show that
g = e (^;a+ar+ 2 ’ brb + )= e ( 2 ’< +) + 2 ’( l - « r “ )) ) > where «'+) and
T r T r
are the numbers of electrons and of positrons in the corresponding
states. On setting n[+) = = 0 we obtain the charge in the vacuum
state Q0 = e ^ 1, where the summation is taken over states of nega-
r
tive frequencies.
This quantity is, evidently, infinite. Therefore, just as in the case
of our discussion of the energy of the electron-positron field only the
charge of the field with respect to vacuum has a meaning. This quantity,
equal to
q = Q - Q o = e E n tr+)- e E » {r -) > (19.26)
r t

represents the sum of the charges of the individual particles—the elec­


trons and the positrons.
The foregoing expressions for the energy and for the charge of the
vacuum can be interpreted in the following manner: The vacuum repre­
sents a state in which all the negative energy levels of an individual
electron are filled. This infinite “negative sea” of electron is not directly
observable in itself, but under the action of various external fields an elec­
tron can make a transition from a state of negative energy to a state of
positive energy (this will obviously require an energy not less than 2m),
giving rise to a “hole” in the infinite negative sea of electrons, which
will behave as a particle of positive energy, with its charge opposite
in sign, but equal in absolute value, to the electron charge. Such a “ hole”
in the infinite negative electron distribution can be interpreted as a pos­
itron, while the creation of the “hole,” i.e., the transition of an elec­
tron from a state of negative energy to a state of positive energy, can
be interpreted as the creation of an electron-positron pair.
In the theory of the quantized electron-positron field there exists
a possibility of defining the current density in a somewhat different
manner, viz., in such a way that the charge in the vacuum state is equal
to zero. Indeed, with the aid of the operators y>a(x)y)p(x') and —y>p(x')
y>a(x) which differ only by a onumber, we can construct the four-vector

j,/*) = (yMU (Va(x) Vfiix) ~ Vfs(x) Va(X) ) - (19.27)


ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 213

which, like the vector (19.22), satisfies the equation of continuity (19.22').
Therefore, this vector can be regarded, with as much justification as
the vector (19.22), as the electric current density associated with the
electron-positron field. As can be easily shown, the total charge of
the field will in this case be given by

9 = —i
^
f T T
[b+, br] = e ^ n [ +)~ e£ n < r > .
T r

This expression does not contain the infinite charge of the vacuum,
and coincides with expression (19.26).
Formula (19.27) may be written in the following abbreviated form:

j^OO = y r M x^ = y > (19.27')

where yMis the matrix transposed with respect to yM. On introducing


the charge-conjugate field operators

Vc(x) = Cyj(x), y%x) = C' >(x)

and on noting that in accordance with (10.8) we have

= wc(x) r flV ( x)
we can put formula (19.27') in the form

.U*) = y (v>(*)Y„V(x) - V c(x)Y„n>c(x)j • (19.28)

This expression will remain unaltered if we replace the operators ip{x)


and \p(x) by the charge-conjugate operators and at the same time reverse
the sign of the charge e.
We now show that the current density can be written in the form
of a normal product:
j^(x) = ieN(rp(x) y^ tp(x)). (19.29)

In order to do this we substitute into (19.27) the expressions

Va (x) = Aa(x)+ B+(x) and ^ (x) = A+(x) + (x) ,


214 QUANTUM ELECTRODYNAMICS

where A(x) and B(x) stand for the terms in the expansion (18.19) con­
taining positive and negative frequencies, respectively:

jM(x) = y (

= - (rXp Ap ( X + Bt ( x ) + Ba(x) Ap (x)

+ Ba(x) B / (x) - Ap (x) A+(x) - Ap(x) Ba(x)~B+(x)A+(x)^B+(x) Ba(x)).


On utilizing the quantization rules (18.8) and formulas (10.28), (10.33),
we can easily show that
{ A JW , B,+(x)} = {A„(x), B /x)} = 0,

{A+(x), A /x )} = T y Ti'Jp)u'yp) -- l / X -2 ^ (,P


p :r = 1, 2 p

{Ba(x), B+(x)} = ~ £ K ( —P)vp ( - P ) = j j 5^ 2“ (‘P+m)ar


p\ r = l , 2 p

Therefore we have
■U*) = ie(yXf> ( (x) A P (x) f Ba(x) A(, (x) + A+(x) B^(x) - Bj(x ) Ba(x))

+ “( r X j s i Z X rp+m)- e - E
' P P I

The first term in this expression represents, in accordance with (19.13),


the normal product of the operators ^ a(x) and v^(x), while the second
vanishes, since ( y ^ ^ ^ O .

§ 20. General Properties of Wave Fields

20.1. Wave Functions of a Field and the Lorentz Group


In the preceding sections we have studied in detail the properties
of the electromagnetic and the electron-positron fields. The theory
of these fields can be generalized to the case of wave fields describing
particles of arbitrary spin. Such a general theory of wave fields may
be of importance in connection with the existence of mesons and of
other particles. On the other hand, a general theory of wave fields will
enable us to understand more clearly the uniqueness of the structure
of the theory of the electromagnetic and the electron-positron fields
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 215

within the framework of present physical ideas. Therefore, in this and


in the following sections we shall discuss certain general properties
of wave fields.
The field is defined at the point x 2, x 3 , x 4) by one or by several
functions which are called the wave functions of the field. In
principle, the number of such functions may be infinitely great. The
wave functions of the field satisfy certain differential equations—the
field equations—which must be invariant under Lorentz transforma­
tions.
Since arguments associated with relativistic invariance play an
important role in the theory of wave fields, we shall recall here the
main properties of Lorentz transformations.
We shall give the name of the Lorentz group to the set of transforma­
tions
O u ,r= 1 ,2 ,3 ,4 ) (20.1)
or, in abbreviated form, x' = lx, which leave invariant the quadratic
form s2 = —x* = —x'f, which have a determinant equal to unity, and
which do not reverse the direction of the time.
The Lorentz group is sometimes referred to as the proper Lorentz
group in contrast to the full Lorentz group which, in addition to the
elements contained in the proper Lorentz group, also contains elements
associated with the inversion of spatial coordinates. We can say that
the full Lorentz group represents the set of all the transformations
(20.1) which preserve the magnitude of the interval, which have a de­
terminant equal to ± 1 and which do not reverse the direction of the time.
The wave functions of the field y f x ) are defined in each coordi­
nate system, with the functions y f x ' ) in the system K' being expressed
linearly in terms of the functions ipfx) in the system K: y>[ (x')
= Si}y)} (x), x' = lx, or, in abbreviated form,
y ' ( x ' ) = S yj(x). (20.2)
The linearity of this transformation is associated with the equivalence
of inertial coordinate systems. We note that the arguments of the func­
tions xp[ and tpj are related by the Lorentz transformation.
The matrix S is defined by the Lorentz transformation which trans­
forms the coordinate system K into the coordinate system K
S = S (1). (20.2')
216 QUANTUM ELECTRODYNAMICS

Since the transformation from the system K to the system K \ fol­


lowed by a subsequent transformation from K' to a new system K",
is equivalent to a direct transformation from K to K" the following
relation holds:
S(l') S(l) = S (l'l), (20.3)

where T is the Lorentz transformation which transforms the coordinate


system K' into the coordinate system K". This equation shows that
the set of transformations S of the wave functions yi gives a representation
of the proper Lorentz group1 in wave function space. We denote
this space by R and the representation by DR (the number of dimensions
of the space R is equal to the number of the functions ip and may be
either finite or infinite).
If for the wave functions of the field we choose some linear com­
binations of the ip, y l = Ayyjj or, in abbreviated form, op = Aip, then
the new wave functions will transform like <p{ = A uy)j' = A ijS]kipk
= Ai}Sjk Akl 9 ?, or, in abbreviated form, y = ASA-1*??. Thus, by mak­
ing a new choice of the wave functions we go over from the represen­
tation of the Lorentz group 1 -*■ S to the representation I -> ASA-1.
Since both these representations are associated with the same field
they are equivalent from a physical point of view; therefore such rep­
resentations are said to be equivalent.
If the wave functions ipi can be divided into two classes: iplf yi%, ...,
ipk and ipk+1, y’k+2 , ..., in such a way that under Lorentz transforma­
tions the functions of each class go over into linear combinations of
functions of the same class, i.e.,

V>'i = £ S iM (i = 1, 2, ..., k),


i=i

Vi = £ S^ ' (,-= k + l >k + 2*>■■■>”)■


)= k+ 1

1 We recall that if G is any group, and if we set in correspondence with each


element gt of the group an operator Tff. in such a way that T0. Tfffc = Tg B then
the correspondence g -> Ts is said to be a representation o f the group G. The number
of dimensions of the space in which the operators Te operate is said to be the
dimension of the representation.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 217

then the representation DR is said to be reducible. It is clear that every


reducible representation can be made up of representations of lower
dimensions. If the latter are in turn reducible, they can be further broken
up into representations of still lower dimensions. By repeating this
process we finally arrive at the irreducible representations into which
the representation DR can be decomposed. Thus, in order to characterize
the representation DR it is sufficient to specify the irreducible repre­
sentations of which it is composed, and to specify the number of times
that any such representation occurs in DR.

20.2. Irreducible Finite-Dimensional Representations of the Lorentz Group


In order to obtain all the irreducible representations of the Lorentz
group it is sufficient to discuss the infinitesimal Lorentz transformations,
i.e., the transformations which are “infinitely close” to the identity
transformation:
K = x tl+etlvx v, | £„„ | < 1. (20.4)

Since the interval xzfl is an invariant, then in the first approximation


the matrix e must be e = —e , which is antisymmetric. Therefore,
the number of independent parameters defining an infinitesimal Lorentz
transformation is equal to six.
We denote by S(e/ziJ the matrix S which in some representation
of the Lorentz group DR corresponds to the transformation (20.4).
On expanding S(£/x„) into a series in powers of eflv, and on restricting
ourselves to the linear terms we obtain

l+ l + K*I,v> I „ ,= - V (20-4')
fl>V

where 1 is the unit matrix. The matrices I are called the infinitesimal
operators. They completely determine the representation of the group.
The derivation of the infinitesimal operators is based on the follow­
ing theorem: The commutator of two infinitesimal operators

W)iv’ ^/il’
is equal to the sum of infinitesimal operators multiplied by certain
coefficients c, which are the same for all the representations of the
group (185):
218 QUANTUM ELECTRODYNAMICS

The coefficients c can be obtained if at least one representation of


the group is known. By choosing the Lorentz group itself as a repre­
sentation of the Lorentz group we can show that the infinitesimal
operators of the Lorentz group satisfy the relations

[I„ . u = <20-5)
or
[IfcM ^km^-lm
[I 4 /C’ ^kmI 4n ^kn I 4 m> (20.5')
^4ml I km’

where k, I, m , n = 1,2, 3.
We note that I!m(l, m = 1, 2, 3) represent the infinitesimal operators
of the rotation group. For the wave field defined by the Dirac equation
the infinitesimal operators can be expressed in terms of the matrices
7 (cf. § 9):

iMv = \ < y , y - y vv d '


Relations (20.5) enable us to determine the form of the operators
Ipv. Indeed, we introduce the operators

I, = - j e , (I, m , n = l , 2, 3), (20.6)

where e/mn is the antisymmetric unit tensor of the third rank. Then
we have
Ixmn = l/e^ m n p Ix p
and the relations (20.5') assume the form
[Ip >I,J feprs Is >
[I^p? I4r] Is >
[ I ^ j) > I f ] f e p rs I 4 S •

From this it follows that the operators


Xp = i(Ip + /l4 P), Yp = h(lp- i l ip) (20.7)
satisfy the following commutation relations:
[Xp, XJ = iepst X t ,
[Yp, Y J = i e pftYt , (20.7')
[Xp, Ys] = 0.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 219

We see that the operators Xp commute with the operators Ys,


and that both the former and the latter satisfy the same commutation
relations. These relations coincide with the commutation relations
for the operators of the components of angular momentum. From
this it follows that in their properties the operators Xp and Ys do not
differ from the operators of the components of angular momentum.
In particular, it is possible to choose a representation in which the
operators X3 and X2 = X2+ X 2+ X 2, and also Y3 and Y2 = Y f+ Y 2+ Y 2
will be diagonal. We denote the eigenvalues of the operators X3 and
Y3 by mj and mk\ as is well known, they can assume the values

mi = —j, —_/+1, ...j; mk = —k, —k + 1, ... k, (20.8)

where j and k are positive integers or half-integers. The numbers j


and k determine the eigenvalues of the squares of the “angular mo­
menta” X2 and Y2, which are equal to j ( j + 1) and fc(k+l).
Thus, the infinitesimal operators of the Lorentz group are completely
determined by the two numbers, j and k, which can independently
take on integral or half-integral values.
As is well known, rotations of coordinates in ordinary three-dimen­
sional space correspond to certain linear transformations of the eigenfunc­
tions of the operators representing the square of the angular momentum
and one of its components. These transformations form a representation
of the rotation group. We can now say that the transformations of
the Lorentz group correspond to certain linear transformations of
eigenfunctions of the operators X3, X 2, Y3, Y 2, which form a representa­
tion of the Lorentz group. This representation, which turns out to
be irreducible, is denoted by r ~ (j, k) . The dimension of the representa­
tion r ~ (j, k) is equal to the number of eigenfunctions of the opera­
tors X3,X 2, Y3, Y2,; i.e., (2 /+ l)(2 fc + l).
We note that if we go over from the wave functions ip to the complex-
conjugate functions ip*, then the operators 1^, are replaced by the
operators I*„, if p i,v = 1,2,3, and by —I*,, if one of the subscripts
H, v = 4, while the operators Xp, Yp are replaced by the operators
Xp = —Y*, Yp = —X*. These operators differ from Yp and Xp only by
a unitary transformation. Therefore, if the functions ip transform
according to the representation r ^ ( j , k ) , then the functions ip* will
transform according to the representation T * ~ ( k , j ) .
220 QUANTUM ELECTRODYNAMICS

Let the functions ip\1) and ip\2) transform according to the representa­
tions t (1) ~ 0 '1, k t) and r (2) ~ ( j 2, k 2)\ then the functions ip™ip™ will
also transform according to some representation of the Lorentz group
which, however, will not be irreducible. It is called a direct product of
the representations r (1) and r (2) and is denoted by t (1 )X t (2). The
direct product of representations can be decomposed into irreducible
representations in accordance with

Oi, k i ) X Os. k 2) = E O'. k ) (20-9)


where j and k assume the values
j — \ j \ - j 2 \,171—721 + 1. • • • .7 i+ 7 2 .
k = | ky—k 2\, \k±—k 2 1+ 1 , ..., k 1-\-k2.
In particular, we have
(y,0)x(0, k) = (j, k ) . (20.9')
An irreducible representation of the Lorentz group r ~ (j, k) can
be regarded, generally speaking, as a reducible representation of
the rotation group. In order to decompose-the representation (j, k)
into irreducible representations of the rotation group, we consider the
operators Ip representing the infinitesimal operators of spatial rotations.
According to (20.7) they are related to the operators X p and Yp by
Ip = Xp+ Y p. Since the operators Xp and Yj commute, the eigenvalues
m of the operator I 3 are equal to the sum of the eigenvalues and
mk of the operators X 3 and Y3, while the eigenfunctions %m of the
operator I3 are equal to the product of the eigenfunctions and cpk mk
of the operators X 3 and Y 3
Xm= w = m j + m fc. (20.10)

From this by utilizing (20.8) it can be easily shown that m may assume
the values
m = —I, —/+ 1 , ..., /,
where
l = \ j —k \, \j—k\ + \ , . . . , j + k .
Spatial rotations correspond to linear transformations of th e 2 y + l
functions 99^ . These transformations give an irreducible representation
of the rotation group of dimension 2y'+l, which we denote by Dr
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 221

In a similar manner the functions <pk transform under rotations


in accordance with the representation Dk.
If one of the numbers k or j is equal to zero the representation
r ~ (j, k) regarded as a representation of the rotation group reduces to
the representation or Dk:
(j, 0) = Di and (0, k) = Dk.
It follows from (20.10) that the representation r ~ ( j , k) regarded as
a representation of the rotation group is equal to the direct product
of the representations and Dk which decomposes into representa­
tions D, with l = \ j —k \ , \ j —k\ + \ , . . . , j + k ,
O', k) = D j X D k = Dt where / = \j—k\, . . . , j + k . (20.11)
The number of these representations is equal to 2 /-fl, if j < k, and
to 2/c+ l, if k ^ j.
We can say that the representation r<—-(j , k ) regarded as a rep­
resentation of the rotation group is equivalent to representations
according to which the eigenfunctions of one of the components of
the angular momentum of a system consisting of two particles with
angular momenta j and k transform under rotations.
We must differentiate between single-valued and double-valued
representations of the Lorentz group. In the case of the single-valued
representations a spatial rotation through the angle 2ji corresponds
to the identity operator 1, while in the case of the double-valued rep­
resentations it corresponds to the operator —1. It can be shown that
the representation r ~ (j, k) will be single-valued if the numbers 2j
and 2k have the same parity, and double-valued if the numbers 2j
and 2k are of opposite parity. The single-valued representations are
also called tensor representations, while the double-valued representa­
tions are called spinor representations. Quantities which transform
according to tensor representations are called tensors, while quantities
transforming according to spinor representations are called spinors.
We give several examples. The representation r -—•(0, 0) corresponds
to a scalar. A vector transforms in accordance with the representation
x ~ (i, i). A symmetric tensor of the second rank with a trace equal
to zero transforms according to the representation r ~ ( l , l ) (the
vanishing of the sum of the diagonal elements of the tensor is necessary
because this sum is itself an invariant, and if it differs from zero then the
222 QUANTUM ELECTRODYNAMICS

tensor does not transform according to an irreducible transformation).


The antisymmetric tensor of the second rank transforms according
to the representations r ~ (1,0) and 7 ^ ( 0 , 1). The Dirac bispinor
which was investigated in subsection 9.3 transforms according
to the representations r ~ ( | , 0 ) and r ~ ( 0 , |) .
Quantities that transform according to the irreducible representa­
tions of the Lorentz group r ^ (j, k) can be divided into four classes:
(1) the class +1 with integral yand k, (2) the class —1 with half-integral
y and k, (3) the class +£ with integral y and half-integral k, and (4)
the class — e with half-integral j and integral k.
If a quantity ip which transforms according to the irreducible rep­
resentation D r can be represented in the form of a sum of quantities
belonging to the same class, then we say that the quantity ip also belongs
to this class.
It can be easily shown that tensors with an even number of indices
belong to the class + 1, while those with an odd number of indices belong
to the class —1; in particular, a vector belongs to the class —1.
Multiplication of quantities belonging to different classes is carried
out in accordance with Table 8.

T able 8

-r 1 - l + e — E

+ 1 + 1 - l + e —E
- 1 - 1 + 1 —E -he
( 20. 12)
+ e —£ -M

—E —£ -he - 1 4-1

If a quantity y> that transforms according to the representation


r ~ (y, k) belongs to the class +1 or to the class —1, then the complex-
conjugate quantity that transforms according to the representation
r * ~ ( k , j ) belongs to the same class as ip. However, if ip belongs to
the class + e (or —e), then ip* belongs to the class —e (or +e).

20.3 Energy-Momentum Tensor and Angular Momentum Tensor


Just as in the case of the electromagnetic and the electron-positron
fields we assume that the wave field, regarded as a generalized dynamical
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 223

system (176), can be made to correspond to a Lagrangian represented


by the volume integral of a Lagrangian density L(y, dip/dxj. The
latter depends on the wave functions of the field ip and on their first
derivatives with respect to the coordinates and the time, but does not
explicitly depend on x.
The integral of the Lagrangian with respect to time is called action.
It can be represented in the form of an integral of L over the four­
dimensional volume Q:
J = J L(ip, ip_J<Px, (20.13)
Q
where ipt/i = dip/dx^.
We assume that the principle of least action holds, according to
which the action must be an extremum for actual motion, i.e., for those ip
which satisfy the field equations. Since these equations must be relativis-
tically invariant, the function L must also be relativistically invariant.
The total variation of the action resulting from varying both the
wave functions of the field and the boundaries of the region of inte­
gration is given by

" = / ( % * ' + £ - * ' ■ ' ) d‘x + / "*<■**•


a 1/2 x
where the second integral is taken over the hypersurface 27 bounding Q,
while dx represent the variation of the boundary coordinates. On
integrating by parts the second term in the volume integral we can
rewrite dJ in the form

f ( w - ^ ) Svd,* + f ( ^ Sv+L*xh - (2 0 ' 14)

We consider first of all the variation of the action in the case of


fixed boundaries and assume that the variation of the wave functions
at the boundary is equal to zero. In this case the second integral vanishes,
and by equating to zero the variation of the action we obtain the
Lagrangian equations for the wave functions of the field yii :
d dL dL _
(20.15)
dxM didxpjdx^ dy>t
The actual “motion” of the field must satisfy these equations.
224 QUANTUM ELECTRODYNAMICS

We now consider the variation of the action in the case of actual


motion, by assuming that the region of integration Q together with
the wave field is subjected as a whole to an infinitesimal translation
or to an infinitesimal rotation. The action will obviously be unaltered
in this case, i.e., the variation of the action dJ will be equal to zero.
In the case of an infinitesimal translation, i.e., of the infinitesimal
transformation of coordinates
^ -*• K =
where e is an infinitesimal four-vector, the wave functions are subjected
to the infinitesimal transformation
v>(xj -►y'( x M) =
whence it follows that
dip = ip’—ip = —£„¥».„■
On substituting this value for the variation of the wave function
together with Sx = e into the surface integral (20.14) we obtain
6J = f e ^ d a ^ 0, (20.16)

where

T T> i -
T , v= L6 i p = Ld,„.— „S L TAd L (20.17)
dy)' ^ dxM d(dipjdxv) ‘
This quantity is a four-dimensional tensor of the second rank to which
we shall give the name of the energy-momentum tensor.
It follows from (20.16) that
/ T ^ d a v = 0. (20.16')
v

On assuming that the volume Q is bounded by two hypersurfaces oriented


orthogonal to the time axis, i.e., dox = doz = do.s = 0, and doi = —idr,
where dr is an element of three-dimensional volume, we obtain from
(20.16') the conservation laws
— i j Tflid r = P/i= const, (20.18)
where the integration is carried out over the total volume of the field,
and the contributions from the surfaces at infinity are assumed to
vanish.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 225

Formula (20.18) shows that P^ is a four-vector. It is called the energy-


momentum four-vector of the field. The spatial components of P
define the momentum of the field P, while the time component divided
by i is equal to the energy of the field H.
The energy density of the field is evidently equal to w = —T44.
This expression can be obtained by analogy with classical mechanics
if we use the well-known expression for the energy of the system

dL
where L is the Lagrangian and p , = - — is the momentum canon-
dqt
ically conjugate to the coordinate qv In field theory we must replace
qt by ipt and pt by dL/dy^ and in this way we obtain the following
expression for the energy density:
w = y)(dL/dip^)—L = —Tu .
We note that the energy density, as a rule, is not positive definite,
and that the condition —Tu > 0 holds only for certain special fields.
By applying Gauss’ theorem to the surface integral (20.16') we
obtain
d T J d x v = 0. (20.19)
Thus, the four-dimensional divergence of the energy-momentum tensor
vanishes. By utilizing this relation it is possible to prove the conser­
vation of the vector P ^ as was shown in § 16.
We note that if we add to a quantity of the form dxp^Jdx^ where
rp is any third rank tensor antisymmetric in the indices v and 2, then
the four-dimensional divergence of the tensor
T f = TMV+(d/dxf) y ) ^
will vanish just as in the case of the tensor T)(„, where
(5 r; /dxv) = o.

Therefore the replacement of the tensor T)lv by the tensor T f will not
lead to any change in the total energy or the total momentum of the
field. In particular, ipMV? can be chosen in such a way that the tensor
TL will be symmetric.
226 QUANTUM ELECTRODYNAMICS

We now consider the variation of the action resulting from an


infinitesimal rotation which corresponds to the infinitesimal Lorentz
transformation given by
K=
where e is an infinitesimal antisymmetric tensor. The wave functions
in this case transform in accordance with (20.4'),
¥>'(*„) = V>(xll- e vltxv)+%ellvIltvy>(xl) ,
where I are the infinitesimal operators of the Lorentz group; from
this it follows that
dy) = y X x J - y t x J =
On substituting this expression for the variation of the wave function
together with the expression dx = sflvx v into the surface integral (20.14)
we obtain

= - 2 = 0, (20.20)

where

T r0~Xr TKS~ (20.21)

Thus, we have

/ M ed°e = 0,
whence it follows that

(20 .22)

The third rank tensor M fn, Q is called the angular momentum tensor.
On assuming that the hypersurface in (20.20) is oriented orthogonal
to the time axis we obtain the conservation laws

—i f M fu,tidr — M ^ = const. (20.23)


From (20.23) it follows that the quantities MMV form a second rank
tensor. This tensor is the angular momentum tensor for the whole
field. We shall show later that the terra x ^ T ^ —x ^ ^ in (20.21) for
^ hv.q can be interpreted as an “orbital” angular momentum, while
the term —(dL/dyg) can be interpreted as a spin angular momentum.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 227

We note that if we symmetrize the energy-momentum tensor of


the field, i.e., if we find such a tensor xnX = for which the
tensor

+ (20.24)

will be symmetric, then 0^ = 6v/l, then with the aid of the tensor 0
we can define the angular momentum tensor as
M,lv.e = (20.25)
In virtue of the equality 0 = 0 the divergence of this tensor van­
ishes and
dM^ . J dxe = ° y
and therefore the tensor M defined by formula (20.25) leads to
the correct value for the angular momentum of the whole field.
We shall refer to the tensor 6^ as the symmetric energy-momentum
tensor in contrast to the tensor T to which we shall refer as the canonical
energy-momentum tensor.
We now show how % can be obtained. On substituting (20.24)
into (20.25) and on utilizing the definition of M we obtain

x nd%ve?Jdxx - x v dxfiJ d x A= - — I Mvy.>+(d/dxA)(xflxve~ x vXMJ

(to the right-hand side of this equation we have added the nonessential
term (d/dxx) (x/lx re!L—xvxligi), whose divergence with respect to the
subscript q vanishes). From this it follows that

Xvnsi /* figv

and, consequently, that


1 dL
jCvQfl 2

where av^ is for the time being an arbitrary tensor satisfying ami = a ^ v.
We choose this tensor in such a way that the tensor xrgt, will be antisym­
metric with respect to the subscripts q and p. It can be easily seen that
228 QUANTUM ELECTRODYNAMICS

20.4. Current Density Vector


The wave functions of the field may be real or complex. In the
latter case the Lagrangian contains both the functions ipt and the
complex-conjugate functions ip* which must be treated as independent
of ipi . Therefore, in the case of complex fields the energy-momentum
tensor and the angular momentum tensor should be written
dL dL
T
X fiv
= Ld — ip
^ (IV T , ( i -V\V ~dip*v (20.26)

and
dL dL
* M * - (20.27)
dw.e ^ dip%
We assume that in the case of complex fields the Lagrangian density
is invariant under the transformation (149)
ip —> yj' = ipeia, ip* —>■ip*' = ip*e~ia, (20.28)
where a is an arbitrary real constant.
We consider the variation of the action for actual motion by assuming
that the infinitesimal transformation (20.28) dip = idaip, dip* = —idaip*
has been performed. Obviously the variation of the action in this case
vanishes, and in accordance with (20.14) we obtain
dL dL dL
- dip- - dip* I do = i da ip*)doM.
dV.u dip* f ( ? b W m
From this it follows that
d dL dL J
0.
dxL.\°V.n cV.v I
In other words, if we define the vector
dL
dL
jii -ip- -ip* (20.29)
dV.fi
where e is a constant, then its divergence will vanish.
Just as in the case of the electron-positron field we interpret j (j, iq)
as the electric current density four-vector associated with the field;
its spatial components j define the spatial current density, while its
time component q is the charge density. Later we shall show that e
represents the charge of the particles associated with the field.
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F IE L D S 229

In the case of real fields the current density vector vanishes. There­
fore, real fields which do not admit the transformation (20.28) describe
neutral particles. Charged particles are always described by complex
fields.
20.5. Relativistically Invariant Field Equations
Free fields obey the superposition principle and, therefore, satisfy
linear differential equations which in virtue of the homogeneity of space
and time must have constant coefficients.
Without loss of generality we can assume that the differential equa­
tions of the free field are homogeneous equations of the first order,
since equations of higher order can be reduced to equations of the first
order if we include the derivatives of ip among the wave functions of
the field. Equations of this type can be obtained from a variational
principle with a Lagrangian density quadratic in ip:

L= (20.30)

where A and B are certain matrices and % is a real constant different


from zero.1 Since L is a real quantity, the matrices A fi and B must
obviously satisfy the following conditions:
Afc= - A + (k = 1 , 2 , 3 ) , A4= A j - , B=B+.
On assuming that B has an inverse matrix (singularity of the matrix
B corresponds to a zero rest mass of the particle) we can introduce
the matrices / \ = B-1 A x and write L in the following form

( 2 0 ' 3 0 - )

where ip = ip*B. By varying^ and ip' independently we obtain the Lagran­


gian equations _

r„ J T + X'f = °> - = »• (20-31>


or
(rx j - = o. - f - ( r x + ^ = 0•

where the subscripts i and k label the wave functions. These equations
are obviously generalizations of the Dirac equations (73).
1 The case x = 0 corresponds to particles of zero mass.
230 QUANTUM ELECTRODYNAMICS

We now investigate the conditions. for the relativistic invariance


of the Lagrangian density (20.30). Under the Lorentz transformation
x' — lx (i.e., x^ = / x„) the wave functions undergo the transforma­
tion xp' (x ) = Sip(x) and the Lagrangian density assumes the following
form:

where A(d/dx) = A (d/dx^). In order that this expression should coincide


with expression (20.30) it is obviously necessary that certain conditions
should be satisfied: S+Al S = A, S+BS = B. On substituting A = Br
into the first condition we obtain r = \ s r s ~ 1 or F /J = Thus,
in order that L should be relativistically invariant the following condi­
tions must be satisfied:
^ = and S+BS=B. (20.32)
It is obviously sufficient that these conditions should be satisfied
in the case of the infinitesimal Lorentz transformation / = d -fs
£ltv = — In such a case in accordance with (20.4) S = Iv
and the conditions (20.32) assume the form
ir M>1J = r v—8pyl\ > where 1, 2, 3,4,
(20.33)
I^";B-)-BIw = 0, I^fcB—BI4fe = 0, where k, I = 1, 2, 3.
In particular, from this it follows that
A = [A, u , (20.34)
I + B - B I fc= 0 , k = 1, 2, 3.
The following expressions for the energy-momentum tensor, the
angular momentum tensor and the current density vector correspond
to the Lagrangian density (20.30'):

(20.35)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 231

where
I/... = B-1I+vB.
The components M MVtg for /n,v = 1 , 2 , 3 can also be written in
the form
M kl.e = X k T le ~ X , T ke + + k , I , VH = 1 , 2 , 3.

(20.35')
We now proceed to determine some general properties of equations
(20.31) . We note, first of all, that the same equation (20.31) can be
satisfied by wave functions describing different fields, for example,
a scalar and a pseudoscalar field. Therefore the field is characterized
not only by the matrices 7^, but also by the representation DR, accord­
ing to which the wave functions ip transform.
The derivatives dip/dx evidently transform according to the repre­
sentation which is the direct product of the representation DR and
the representation ( |, ^) according to which the four-vector x transforms.
On the other hand, in accordance with equation (20.31) the quantities
r dipldx transform according to the same representation as ip, i.e.,
in accordance with the representation DR. In other words, the repre­
sentation D r is contained in the representation Dn x ( M ) . From
this it follows that if an irreducible representation (/, k) of the Lorentz
group is contained in the representation DR, then DR also contains
at least one of the four irreducible representations
—-|), ( j — , £ + v ), ( j - \ - \ , k —J), (7+ \,k-\-\').
Thus, D r cannot be an irreducible representation, except in the case
x = 0. For example, in the case of the Dirac equation (with * =£ 0)
Dr is decomposed into two irreducible representations (|-, 0) and (0, -F).
If we decompose the representation DR into irreducible representa­
tions r ~ (/, k), we can assert that the sum j + k will be either integral
or half-integral in all the representations.
We now determine the masses of the particles described by equation
(20.31) . In order to do this we consider the solutions in the form of
plane waves in the rest system of the particle ip(x) = e~Uat, where
£0 is the rest energy of the particle and y><0) is a constant; the substitution
of ip(x) into (20.31) yields
£0T,4y <0)—^Y'(0) = 0, (20.36)
232 QUANTUM ELECTRODYNAMICS

whence it follows that e0 = («/Af), where Af is a nonvanishing eigenvalue


of the matrix jT4. This formula shows that the possible values for the
mass of the particles described by equation (20.31) are given by
K
(20.37)

where A, can be taken equal to any one of the nonvanishing eigenvalues


of the matrix r t .
Thus, in principle equation (20.31) describes particles of different
masses. In order that it should describe particles of only one value of
the mass it is necessary that the matrix r i should have only one non­
vanishing eigenvalue.
The matrix I), commutes with the infinitesimal operators Iifc(r, k =
1, 2, 3) . Therefore one of these operators, for example, I12 = z'I3 and
r 4 have common eigenfunctions. We denote them by <px. m. where m
is an eigenvalue of I 3. As we already know, this quantity may assume
the values mi = —/i;—/j+ 1 , ..., where l{ determines the square
of the “total angular momentum” I2 = lf + I I + I j. Thus, the eigen­
values Aj are, generally speaking, degenerate with respect to the “ mag­
netic” quantum numbers mt, with the degree of degeneracy of the
eigenvalue Xt being equal to 2/t+ l .
As has been pointed out previously, spatial rotations correspond
to linear transformations of the eigenfunctions of the operator I3, i.e.,
of the 2/j + l functions cpM (with different values of mt and the
same value of At). These transformations form irreducible representa­
tions of the rotation group We can therefore assert that particles
of mass = h/Aj have spin lv Thus, the spin is determined by the degree
of degeneracy of the nonvanishing eigenvalues of the matrix _T4, with
each value of the mass in principle corresponding to its own value of
the spin. We can also say this differently, viz., that the spin is deter­
mined by the number of independent solutions of the system of equa­
tions (20.36) corresponding to a definite value of e0, i.e., to the definite
eigenvalue Xl. The number of these solutions is evidently equal to 2/j-f-l.
(For example, in the case of the Dirac equation the number of inde­
pendent solutions is equal to two in accordance with the fact that the
electron spin is equal to |.)
Since the representation r (j, k), regarded as a representation
of the rotation group, decomposes into representations Dt with / lying
E L E C T R O M A G N E T I C A N D E L E C T R O N -P O S I T R O N F IE L D S 233

between \j—k\ and \j+k\, the spin will be integral or half-integral


depending on whether the number j + k is integral or half-integral.
We now show that the particle spin is determined by the term
—(dL/dyjJl^yj in the expression (20.21) for the angular momentum
tensor or by the term

in the expression (20.35). In order to do this we determine the energy


and the angular momentum of the field in the case when the field con­
tains only one particle at rest.
In accordance with (20.35) the energy of the field is given by

H = — j T44 dr — —1J dr(y)r4(dy)/dx4) —(dy}/dx4) T^p).


On substituting into this expression ip = ipw exp (—ze0 1), and on assum­
ing that the volume of the field is equal to unity, we obtain
H — e0Jy>r4ip dr = £0'tpmr 4ipw .
Since the field contains only one particle the function ip{0) must be nor­
malized in accordance with the condition
\ y {0)r 4ipw \ = 1.
We note that under this condition we obtain for the charge of the field
the value Q = ± e. Indeed, the charge of the field Q is given in accord­
ance with (20.35) by

Thus, as has been stated earlier, the constant e in the expression (20.29)
for the current density vector can be interpreted as the charge of the
particles.
We now obtain the component of the angular momentum of the
field along the x-axis. In accordance with (20.35') it can be written

since the first term in (20.35') vanishes for a particle at rest. But the matri­
ces and I3 commute, and therefore
234 QUANTUM ELECTRODYNAMICS

Since yj is a common eigenfunction of r i and I3, we have

A/1 2 = m f dr = m,

where m is an eigenvalue of I3, or the component of the spin along


the x3-axis.
We note that in addition to fields transforming according to the
finite-dimensional representations of the Lorentz group we cannot,
in principle, exclude the possibility of the existence of fields transform­
ing according to infinite-dimensional representations of the Lorentz
group. Particles described by such fields can exist in states with different
values of the mass, and, as a rule (73), the masses tend to zero if the
spin / -> oo.

20.6. Wave Equations for Particles of Spin Zero and Unity


The simplest example of relativistically invariant field equations
is given by

(2 a3 8 )

where <p is a scalar function of the coordinates and the time. Such an
equation in virtue of the scalar nature of <p describes particles of spin
zero; the mass of the particles is evidently equal to m.
We now show how this equation may be brought into the general
form (20.31). In order to do this we introduce in addition to cp the four--
vector rpM= (l/\/m) (d(p/dxfl). Then equation (20.38) is replaced by

" I ? + '”v“ = o - = o- (2a39)


where y>0= \/mcp. This system can be written in the general form (20.31)

(^ — + ^ = 0, (20.40)

where \p is a five-component wave function consisting of y>0 and xp


while 1 ,2 , 3,4) are four five-rowed matrices which, as can
be easily seen, satisfy relations (51), (104):
(20.41)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 235

In this case the representation DR consists of the representation (0,0)


according to which transforms the scalar ip0, and the representation
( 2'» i) according to which transforms the vector y> .
The wave equation (20.38) can be obtained from a variational prin­
ciple if for the Lagrangian density we take

dcp* d(p
L= - m2<f>*cp\ (20.42)
dx^ dxM
On taking

dwt difo
L— (20.43)
dx. ~9 x m

and on assuming that in carrying out the variation the functions


V o a r e independent, we obtain equations (20.39).
The Lagrangian (20.42) corresponds to the energy-momentum tensor
_ dcp* dcp dcp* dcp
(20.44)
dx^ dx~v + dxv d x ^
and to the current density vector
dcp*
j ti (20.45)
dx„ dx..
We note that a pseudoscalar field can be treated in a manner similar
to the scalar field. In such a case we simply replace the scalar cp by the
pseudoscalar cp Xa antisymmetric in any pair of subscripts.
We now consider the wave field describing particles of spin unity
(158). Such a field is characterized by the four-vector cp whose compo­
nents satisfy the equations
02
m2\cpv - 0 and 4^= 0. (20.46)
dx2 dx..
Because of the subsidiary condition dcpv/dxv = 0 imposed on cpv this
field describes only particles of spin 1. Indeed, we consider the free
field in the form of the plane wave <pti = cp^cxp(ipx). It follows from
(20.46) that p2-\-m2 = 0 and p^cp^ = 0. If we go over to the coordinate
system in which the particle is at rest, i.e., if we set p = 0, we obtain
<p£0) = 0. Thus, in this system the wave field is described by a three-
236 QUANTUM ELECTRODYNAMICS

vector. Under spatial rotations it transforms like a vector (according


to the representation DJ and consequently describes particles of spin 1.
If we introduce the antisymmetric tensor

then the equations (20.46) are replaced by the system

+mipM= 0, ftp, , d% fny>MV= 0, (20.47)


8x„ dx.. dx..
where ip = j / m<p . This system can, in turn, be written in the general
form (20.31) if we interpret ip as a ten-component wave function con­
sisting of the veetor and the antisymmetric tensor ip/iV

( ^ - a- + m ) v = ° . (20.48)

Here /3 are four ten-rowed matrices which, as may be easily shown,


satisfy the same relations (20.41) as the five-rowed matrices appearing
in the equations for a particle of spin zero.
Equations (20.47) can be obtained from a variational principle if
for the Lagrangian density we take

1 * ufyv ftPv\ 1 / dV>t dlP*


2 W^v\ d x v dx } 2\dxv dx V V + y V&Wv-wvtfVV
(20.49)
and in carrying out the variation regard the quantities iptl,ip*,iptlv,ip*v
as independent.
The Lagrangian density (20.49) corresponds to the energy-momentum
tensor
dw dip*
T/tv= L d MV- e
~ v%- - dx\ f . (20.50)
dx
and to the current density vector
U = ie (20.51)
Equations (20.40) for particles of spin zero and equations (20.48)
for particles of spin unity can be obtained with the aid of the following
Lagrangian density which is of the same form for both kinds of particles:
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 111

This Lagrangian density corresponds to the energy-momentum tensor

(20-53)
and to the current density vector
j M= (20.54)
where
V> = V * (2yffl— 1)

and the matrices /S are defined by relations (20.41).

§ 21. Quantization of Fields. Connection between Spin and Statistics

21.1. Nondefiniteness of the Charge in the Case o f Integral Spin and


of the Energy in the Case of Half-Integral Spin
We now show that in the case of a wave field describing particles
of integral spin it is impossible to introduce a positive definite charge
density, while in the case of a wave field describing particles of half­
integral spin it is impossible to introduce a positive definite energy
density (149).
We first of all consider fields describing particles of integral spin.
In this case if transforms according to the representation DR which
decomposes into representations r ~ ( y , k) with integral j f k . Therefore
ip contains functions only of classes + 1 and —1 which we respectively
denote by if{+1) and y>{~1). The basic field equation satisfied by if can
be written in terms of the Fourier components of i f in the form
( i r ilkM+x)y> = 0,
where k is the propagation vector (we assume that if ~exp(z‘fcx).
This equation can be decomposed into two equations:
t (r„)+, (/>. , *,y->+*?<-»» = o,
+ iCQ-i.+l*,vl+11+ xr'-1’ = o,
where (71 )±l are the elements of the matrix 7^ operating on the
functions if{±1). It can be easily shown that
C^)+i, +i = - i = 0-
238 QUANTUM ELECTRODYNAMICS

Indeed, since the propagation vector belongs to the class —1, then in
the case of nonvanishing ( r /x)+1.rl and the first and the
third terms in each equation would belong to different classes, but this
is impossible.
Thus, in the case of integral spin the equations have the form

(r M) + i . - i k ^ { 11 + ? V +1) = 0,
(r#4)_1>+1fc#1v'(+1) + *v'<~1) = o
We see that the role of the matrices r consists of relating functions
of different classes.
Equations (21.1) are invariant under the transformation
k^ -* — k^, ^ (+1) -»■ aip{+1) and ip{~l) -> —a y (-1),(21.2)
where a is an arbitrary constant. It can be easily shown that a may
be taken real. Indeed, the functions y (+1)* and ip<-u* transform in
the same way as the functions ^ <+1) and ip{~1); therefore from the point
of view of relativistic invariance conditions of reality for y <+1) and
^y<-1) can have the form
Y>(+d* = y>(+1) and ipi-v* —

In order that these relations should remain valid under the transforma­
tion (21.2) it is clear that a must be real. Without loss of generality
we can take a = 1 and rewrite the transformation (21.2) in the form
-> —kp ip{+1) -> v»<+», -> - y j 1- 11. (21.3)
We can construct different tensor quantities with the aid of the
functions yj(+1), yj*-1' and their derivatives. For subsequent development
we shall be interested in the structure of such tensors quadratic in ip.
If we go over to the Fourier representation, then the problem reduces
to the construction of tensors with the aid of yj<+1) and k , with these
tensors being required to be quadratic or bilinear forms in ip.
We start by constructing the even rank tensor T(2n). Since such
a tensor belongs to the class + 1, it can have only the structure

T<2n) ~ 2 kWW(+1)W{+l) + Z k w W ^ 1)W{- 1)Jr ' Z k 2N^ y (21.4)

where N are integers (we do not include terms containing i/j(+1)* and
ip*'-1''* since they also belong to the classes +1 and —1).
E L E C T R O M A G N E T IC A N D E L E C T R O N -P O S IT R O N F IE L D S 239

The odd rank tensor ^(2n+l) belongs to the class —1 and, conse­
quently, has the following structure:
r (2n+1) ~ ^ 2W"hV (+1)v;(+1)+ ^ 2JV+1vj(“ 1)v,<“ 1)+ ^ 2Wv,(+1)v,(-1)-
(21.5)
We now investigate the behavior of these tensors under the transfor­
mation (21.3). It is clear that this transformation does not alter T(2n),
but reverses the sign of T(2n+L).
In particular, the transformation (21.3) reverses the sign of the
current density vector j . In other words, to each solution of the field
equation there corresponds another solution with the opposite sign
of the current density components. Therefore, in the case of particles
of integral spin we cannot introduce a positive definite particle density
which transforms like the fourth component of a four-vector.
Thus, we have proved the nondefinite nature of the charge density,
and consequently of the total charge, in the case of particles of integral
spin.
We now consider fields describing particles of half-integral spin.
In this case j + k is a half-integer and the functions %p belong to the
classes + e and —e. We denote them respectively by xp{+e) and
In place of equations (21.1) we now obtain the following equations
(in terms of their Fourier components):
i'(^ 4)+« - ekflW(- E)+>iy>(+E) = 0 and i ( ^ - , 1+,fc(lV(+‘)+*V'l" t) = °>
( 21. 6)

where (71 )±£>_£ are the elements of the matrix operating on the
functions ip{±£). These equations, like equations (21.1), are invariant
under transformations of the form
k —►—k^, yj(+£) —> atp^£\ y)*--e) —> —ayj(-e) (21.3')
but in contrast to the case of integral spin, when a was real, we must
now regard a as a purely imaginary quantity. Indeed, from the point
of view of relativistic invariance the conditions of reality in the case
of half-integral spin can have the following form:
yj<+£>* _ yj(-e)* —(21.7)
while the following conditions are impossible:
yjH-®)* = y)t-+e\ e>* = ip^~£\
since ^<+£>* belongs to the class —e, while y><-£>* belongs to the class +e.
240 Q U A N T U M E L E C T R O D Y N A M IC S

On substituting into (21.7) the transformation (21.3') we obtain


a = —a*.
Without loss of generality we can assume that a = i, and rewrite
(21.3') in the form
-»• — k fl) ip(+c) -»• iipi+e), ip{' e) -> — iip{~e). (21.8)
We now investigate the structure of the tensors that can be con­
structed with the aid of ip{±e) and k^, if they are either quadratic or
bilinear in ip. By utilizing the multiplication table (20.12) we can easily
show that these tensors must have the structure

r (2n) ~ £ &2V +£V +£) + ^ & 2V ~ £V ~ £)+ ^ 2W+V +£V ~ £) ,


T ,( 2 n + 1 ) '— < ^ k 2 N + 1 i p l' + E)i p l' + !L)J\ - ^ ] k ‘2‘N J r l i p <'~ ’: ) i p {~ E)- \ - ^ l k 2 N i p (JrE)i p l' - e )

(we have not included terms with ip{+e)* and ip{~e)*, which belong to
the classes —e and +e).
We now investigate the behavior of these tensors under the transform­
ation (21.8). It is clear that this transformation does not alter T*n+1
and reverses the sign of T(in) in contrast to the transformation (21.3)
which does not alter the even rank tensors, and reverses the sign of
the odd rank tensors. We, therefore, conclude that in the case of particles
of half-integral spin it is impossible to introduce a positive definite
energy density, and consequently a positive definite total energy.
From the assertions which we have just proved it does not neces­
sarily follow that in the case of integral spin there always exists a pos­
itive definite energy density, and that in the case of half-integral spin
there always exists a positive definite charge density. On the contrary,
it can be shown that only in the case of spin | is it possible to introduce
a positive definite charge density, and only in the case of spin 0 and
1 is it possible to introduce a positive definite energy density (it is
assumed here that the matrix r o is brought into diagonal form).

21.2. Quantization o f Fields for Integral and Half-Integral Spin. Pauli's


Theorem
From the impossibility of introducing a positive definite particle
density in the case of integral spin, and a positive definite energy density
in the case of half-integral spin, there follows a very important physical
conclusion, which is that within the framework of the one body problem
E L E C T R O M A G N E T IC A N D E L E C T R O N -P O S IT R O N F IE L D S 241

it is impossible to formulate a theory based on relativistically invariant


field equations. In other words, a wave field must be associated not
with a single particle, but with a whole set of particles. Moreover,
the wave functions must be regarded not only as functions of the coordi­
nates and the time, but also as operators operating in particle number
space. We shall refer to them as the operators of the quantized field.
If we start by considering single particle states then, as is well known,
we must distinguish two possible cases, viz., when each such state may
contain an arbitrary number of particles, and when the number of
particles may not exceed unity. In the former case we say that the parti­
cles obey Bose-Einstein statistics, and in the latter case that they obey
Fermi-Dirac statistics, and we call the particles respectively bosons and
fermions.
In order that the field equations should describe either bosons
or fermions the operators of the quantized field must satisfy definite
commutation relations. By generalizing the results of §§ 16 and 18
we can say that the commutators of the quantized wave field leading
to bosons, and the anticommutators of the quantized wave field leading
to fermions, must be c-numbers.
We shall now prove the following fundamental theorem due to
Pauli: wave fields describing particles of integral spin must be quantized
in accordance with Bose-Einstein statistics, while wave fields describing
particles of half-integral spin must be quantized according to Fermi-
Dirac statistics.
We consider some physical quantity f(x) (for example, current
density or energy density) at the two points x 1(rl , i t 1) and x2(r2, i t 2)
separated by a space-like interval = (t1—tz)2—(rl —r2)2 < 0- Since
during the time \tY— t2\a. disturbance propagated with the speed of light
from the first point will have travelled a distance 111—t2| which is smaller
than ( ^ —r ^ 2, the second point cannot experience an effect due to
the first point. Therefore measurements at two points separated by
a space-like interval cannot influence each other, while the operators
corresponding to the quantities / ( x and /( x 2) must commute.
In the case of fermions the wave functions y> do not themselves
have a direct physical meaning, since spinors, in general, in view of
their transformation properties under rotations can be defined only
up to their sign (they change sign as a result of a rotation through
242 Q U A N T U M E L E C T R O D Y N A M IC S

an angle 2 n, cf. § 9). Therefore, in the case of fermions, only quantities


quadratic in ip have physical meaning, and it is these quantities that
must commute at points separated by space-like intervals. This commu­
tativity will exist if the corresponding commutators or anticommu­
tators of the field operators vanish.
A similar conclusion can also be drawn in the case of fields describing
bosons. However, in the case of bosons, in contrast to fermions, the
wave functions can have a direct physical meaning. In this case the
commutators of the boson field operators must vanish if the points
to which these operators refer are separated by a space-like interval.
We assume that the commutators or the anticommutators of the
fields yr(xx) and y f ( x 2) are c-numbers which vanish when ( ^ - r 2)2 > 0.
From the homogeneity of space-time it follows that the commuta­
tors [ipT,ip+]_ and the anticommutators fyr,y>+]_ can be functions
only of the difference of the coordinates x t ~ x 2, i.e.,

[y, O i ) >f t 0 2)1 ± = Grs( * i - * 2), (2 1 .1 0 )

where G>s(x) is a c-function. Our problem consists of showing that


in the case of integral spin this relation contains a commutator, while
in the case of half-integral spin it contains an anticommutator.
We assume that each operator ipT contains quantities belonging
to only one of the classes + 1, —1, +e, —e. In the case of integral
spin these will be quantities belonging to the classes + 1 and —1, while
in the case of half-integral spin they will be quantities belonging to
the classes + e and —e. Since complex conjugation leaves quantities
belonging to the classes + 1 and —1 in the original class, and transforms
quantities of the class + e into those of the class —e, and vice versa,
then by using the multiplication table (20.12) we can conclude that in
the case of integral spin the quantities ipTrp+ must belong to the class + 1 ,
while in the case of half-integral spin they must belong in the class —1.
From this we can conclude that independently of how the fields
must be quantized, i.e., independently of whether the commutator
or the anticommutator appears in (21.10), the quantities Grs(xr) must
be tensors of even rank in the case of integral spin, and tensors of odd
rank in the case of half-integral spin.
The structure of these tensors can be easily determined. Indeed,
we assume that the field tp describes particles of a single value of the
E L E C T R O M A G N E T IC A N D E L E C T R O N -P O S IT R O N F IE L D S 243

mass m. Then we have available only the one scalar function A(x) (cf.
§ 17), which vanishes for x2 > 0:

A {X) = ( 2 n f I eipr^ ~ dP ’ e = V ™ 2+ P 2-

If we differentiate A (x) once with respect to the coordinates x we obtain


a vector, if we differentiate it twice we obtain a second rank tensor,
etc. Therefore, in the general case the tensor Grs(x) has the structure
<j„(x) = F ‘j°(d/5x) A(x), where F<.">(d/dx) is a differential operator
containing derivatives with respect to x m u ltip lie d by constant coef­
ficients; n is the order of the highest derivative which coincides with
the rank of the tensor GTS(x).
We can, thus, rewrite expressions (21.10) for the case s = r in the
following form:
p(271) A (Xi—x2), 2 (j+ k) = IN,
rr \ d x 1
[ w ( * i) ,7 V (* 2)]
p(2n+l) | A (Xi—x2), 2( j+ k) = 27V+1,
dxx
( 21 . 11)

where the upper line refers to the case of integral, and the lower line
to the case of half-integral spin (N is an integer), and where, as can
be easily shown, (d/dx) and F(r^n ! 1) (d/dx) respectively contain
derivatives of only even or odd orders.
We now consider the expression
K(x, - x 2) — [rpT(*i), y+ (x2)}, + [y>T(x2), y j (xx)] +,
symmetric in x x and x 2. Since A(x) is an even function of the spatial
coordinates and an odd functions of the time
A (r, t) = A (—r, t) = —A(r,—t),
the symmetric quantity K(x1—x2) must contain an even number of
derivatives of A ( x ±—x 2) with respect to spatial coordinates and an
odd number of derivatives with respect to the time, i.e., an odd total
number of derivatives. This is possible, as can be seen from (21.11)
in the case of half-integral spin, and is impossible in the case of integral
spin if only FTT(d/dx) differs from zero. We therefore conclude that
in the case of integral spin K(x1—x 2) must vanish and
[ w ( * i) » ¥V+ (> 2)]± + [¥ Y (* 2 )> wt (* i)]± = °» 2 0 + fe ) = 2N. (2 1 .1 2 )
244 QUANTUM ELECTRODYNAMICS

But this equality evidently cannot be satisfied if it contains anticommu­


tators, since in this case the left hand side of the equaiton is essentially
positive for x x = x2.
Thus, we have proved that in the case of integral spin the commu­
tation relations for the operators of the quantized field must contain
commutators, i.e., they must be of the form

[W(*i)> Y ^fo)]- = A {xx- x 2) and 2 ( j+ k) = 2N. (21.13)

By utilizing the results of quantizing the electromagnetic field we can


conclude from this that the number of particles associated with the
field under consideration and existing in a definite state can be arbi­
trary. This means that particles of integral spin obey Bose-Einstein
statistics.
From the arguments just presented we cannot conclude that in
the case of half-integral spin the anticommutator of the operators
of the quantized field must be a, onumber, i.e., we cannot conclude
that particles of half-integral spin must obey Fermi-Dirac statistics.
Formally the commutator can be a c-number also in this case, i.e.,
formally we can carry out the quantization of the field in the case of
half-integral spin in accordance with Bose-Einstein statistics. However,
we must keep in mind that the energy of the field in the case of half-
integral spin is not positive definite, and since the energy of free particles
must be positive, it follows from this that quantization in the case of
half-integral spin must be carried out in accordance with Fermi-Dirac
statistics together with the requirement that all, or practically all, the
states of negative energy (we have in mind the single particle states)
must be occupied, or, in other words, we must from the outset introduce
the concept of the field vacuum in which all the single particle states
of negative energy are occupied.
Thus, in the case of half-integral spin the commutation relations
for the operators of the quantized field must contain anticommuta­
tors, i.e., they must be of the form

V t (*2)]+ = Frs ^ (*1 —* 2) and 2{j-\-k") = 2N-\-\.


(21.14)
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 245

21.3. Inversion of Coordinates and Time Reversal


The transformation of the propagation vector A: -> — k^, which
we utilized in a purely formal manner in proving the nondefinite nature
of the energy in the case of half-integral spin and of the charge in the
case of integral spin, acquires an easily visualizable meaning if we go
over from the momentum to the coordinate representation, viz., such
a transformation corresponds to an inversion of the coordinates and
time reversal
(21-15)
This transformation, together with the corresponding transformation
of the components of the wave function, is permissible in the case
of the equations for the free fields, but the theory as a whole (including
the commutation relations and the interaction between the fields)
turns out to be noninvariant under the transformation (21.15) alone.
Such invariance can be achieved, as will be shown later, if the transfor­
mation of the field operators associated with the inversion transforma­
tion xM-*■ — x^ is supplemented by a transition to the transposed
operators. This may be clearly seen on the example of the electromag­
netic field with which we begin our discussion.
We assume that the transformation (21.15) corresponds to the
transformation of the operators for the potentials of the electromag­
netic field
AM(x) ->■ A^(x') = —A^(x), where x' = —x. (21.16)
The equations for the free electromagnetic field

S ? A ,( * ) = 0 01.17)

are invariant under the transformation

while the commutation relations


[A^xj), A„(x2)] = - i d ^ D f a - x J (21.17')
are altered. Indeed, on replacing in expression (21.17') x by —x, and
on substituting into it A^(—x) = — A^(x) we obtain
[ a ; ( x O , a ; ( x 2) ] = - idMVD ( - x ^ x , ) ,
246 QUANTUM ELECTRODYNAMICS

and since we have D ( —x ) = — D(x), then


[ A ; ( ^ ) , A ; ( x 2)\ = iSfivD (x l - x 2).
However, it can be easily shown that if we go over from A^(x) to the
transposed operators A£(x) ^ A^T(x) (the index T denotes the trans­
posed operator), then the latter will satisfy not only equation (21.17),
but also the commutation relations (21.17'). Indeed, since (AB)T
= BTAT, then
[A£(*i), A?(x2)] = — ib ^D (xx—x2).

We thus see that the operators A^(,y) and A*(x) satisfy the same
equations of motion and the same commutation relations. From this
we can conclude that they are connected by a unitary transformation

A£(x) = V A ^(*)V -\
where V is a unitary operator V+ = V-1, operating in particle number
space. This relation can also be rewritten in the following form:
a ;( x) = u a jw u - 1, U = V*- i , (21.18)
where A'/t(x) = - A „ ( —x).
The form of the operator U can be easily obtained. On utilizing
the expansion of A;i(x) into plane waves (16.13), and on noting that
in accordance with (16.18) Cj£A= c;fA, we obtain AJ(x) = A (—x),
i.e., —A^(jc) = UA (x)U-1. Therefore the operator U satisfies the
relations
UcMU -* = - c feA. (21.19)

Since the operator U obviously commutes with the operators NfcA


= c+AcfcA we can regard it as diagonal in the representation in which
the NfcAare diagonal. Therefore it follows from (21.19) that the values
of U corresponding to two consecutive values of Nkx are related by the
expression (U)WfcA_L1= —(U)^ ; and from this we can conclude that
U = (—l)rw*L (We note that this operator coincides with the operator
of charge conjugation for the electromagnetic field, cf. § 23, formula
(23.29).)
We now investigate the manner in which the operators of the free
electron-positron field transform under an inversion of coordinates
and a reflection of the time.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 247

On writing this transformation in the form


Y>'(*)= Oy>(-x), (21.20)
where O is a matrix which we call the inversion operator, and on using
the Dirac equation

(yli^ — +rn^w(x) = 0,
we obtain

(— = °-
In order that this equation reduce to

the operator O must satisfy


y ^ ° + O y /t= 0 , /z = 1, 2, 3,4.
Since the operation of inversion applied twice in succession must not
alter the spinor, i.e., must reproduce it up to its sign, we must have
O2 = ± 1 . Finally, the matrix O must commute with all the transforma­
tions S of the wave function associated with the proper Lorentz
transformations, or, in other words, with all the infinitesimal operators
of the Lorentz group 1^ =
[o, y = o.
From these conditions it follows that O has the form
0 = £y5
where y5 = yiy2y3y 4 and f = ± 1 , ± i-
The inversion operator O can be represented in the form of the
product of the operator of spatial inversion and the operator of time
reversal which we respectively denote by Or and Ot\
0 = 0 r0t. ( 21 . 21)

We saw in § 9 that Or = I = fy4. Therefore we can say that


Ot = 71^2^3- (21.22)
We note that yj(x) = y>+(x)yi transforms under the inversion of
the coordinates (21.20) like
y>'(x) = xp\x) = xp(—x) O, ( 21 . 20 ')
248 QUANTUM ELECTRODYNAMICS

where
O = y i O+yi = f*y8-
We can easily show that
Q O =-l; Oy^ O = , (21.23)
and, therefore,
yj'(x)y'(x) = —y ( —*)v>(—*).
(21.23')
v ( x ) y fy ( * ) = v>(—aO^vvC—*)•
Thus, under the transformation x^ -> —x , yjy> changes sign, i.e., behaves
not as a scalar, but as a pseudoscalar, while y>y ip does not reverse sign,
i.e., behaves as a pseudovector.
It can be easily shown that the operators y/(x) and yi'(x) satisfy
the same commutation relations as the operators ip(x) and y>(x). Indeed,
on replacing x by —x, in the expression
d
M * i ) > ¥>(*2 ) } = i y y —m Zl (a-! * 2 )

and on substituting into it


ip (—x) = O“V M , y>(—x) = y>'(x) o -1,
we obtain

{O-hpXxj), y'(x9) O - 1 } = —Yn-fc- —m j d i —Xi + x J ,

whence we have

{V>’(Xi) ,V>'(x2) } = i O \ - y M - mj 0/1 (—xx+ x2) .

On utilizing (21.23) and the fact that the function zl (x) is odd, we obtain

{ y / ( X i ) , ? / / ( x , ) } = i\y/L ^ — mj/1 (xx— x 2) ,

which is what we set out to prove.


Thus, in contrast to the free electromagnetic field for which the
transformation of inversion leaves the field equations invariant, but
does not leave the commutation relations invariant, in the case of the
free electron-positron field the transformation of inversion leaves both
the field equations and the commutation relations invariant.
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 249

On the other hand, for the free electron-positron field the transposed
operators tpT(x) satisfy the same equations of motion and the same
commutation relations as the operators y(x). From this we can con­
clude that the following relations hold between the operators ip'{x),
ip(x) and ipT(x) of the free electron-positron field:
y>'(x) = LhyOOUf1 and y>'(x) = U o ^ O ^ lh r1, (21.24)
where U1? U 2 are unitary operators. However, these relations are not
equivalent, since the equations for interacting fields are not invariant
under the inversion transformation x„fj, -* —x„ji alone. We shall show
later (cf. subsection 22.4) that all the equations connecting the electro­
magnetic and the electron-positron fields remain invariant under the
transformation of inversion accompanied by a transition to the trans­
posed operators, i.e., under the transformation
\ ( x ) -►A£(x) = A^T(x) and y>(x) -> y R(x) = y>'T(x). (21.25)
This combined transformation we shall refer to as the transformation
of total inversion or of strong inversion. We emphasize that the trans­
formation of total inversion corresponds not only to a reversal of
the sign of the argument of the field operators, but is also associated
with an interchange of the initial and final states in the matrix elements
of the field operators. This interchange is brought about by the trans­
ition to the transposed operators.
Thus, of the two expressions (21.24) only the second expression
has a physical meaning. It can be written simultaneously for the opera­
tors A^(x) and ip(x) in the form
A'(*) = UAJ(x) U - 1 and y>'(x) = U^T(x)U _1, (21.26)
where U is a unitary operator operating in particle number space.
We now make an assumption which generalizes expressions (21.26),
viz., we assume that in the case of arbitrary wave fields y)(x) the follow­
ing relation holds:
y'(x) = Oy> ( - x ) = UyT(x) U " 1, (21.27)
where O is the transformation matrix for xp(x) associated with the inver­
sion of coordinates, while U is a unitary operator. Without loss of
generality we assume in accordance with (21.3) and (21.8) that the
effect of the matrix O reduces to multiplication by + 1, —1, + /, —/
250 QUANTUM ELECTRODYNAMICS

depending on which of the four classes: + 1, —1, +£, —e is the one


to which belongs the component of the field ip:
Oipi+1) = ip{+1), C)y(- 1) = —ip(- 1},
Oy(+E) = iipi+e),
We now show that assumption (21.27) is equivalent to the theorem
on the connection between spin and statistics; in other words we show
that if (21.27) holds, then the commutation relations for the field opera­
tors must contain commutators in the case of integral spin and anti­
commutators in the case of half-integral spin (176).
We start from the relations

[Vr(*i)» Vt(x 2)lr = F„ A ( * ! - x 2), (21.29)

where each of the operators ipr,ips belongs to one of the classes + 1,


—1, + e, —e and Fro, is a differential operator containing derivatives
multiplied by constant coefficients. Since [wOO, %+(*2)fi is a c-num-
ber, we can write

Frs( ) A ^Xl~ x ^ = u [w(*i)> ^ +(-^2)]


= ± UVi+T(x2) U - 1]±
= ± [(O ^ -x ^ r* ( O ^ - 4 ' l i
= ± 2 —*l), # ( - ^ 2)]i-
t ' s '

The bracket [ipr {—Xj), ip+(—x2)]J: appearing in this expression

is evidently equal to F rVj - - A - J A ( - x1+ x 2), and since A (— x)


= —A(x), we have

i v A - x i), vtf(-**)L = ^ ^ A ( Xl- x 2).


Therefore

K { A ) A ( x ' ~ x *>= 0 „ .O S .F ,...( _ 2 .) A ( x t - x , ) . (21.29’)

In the case of integral spin the quantity ipripf belongs to the class + 1,
if W and W, belong to the same class (+1 or - 1 ) , and to the class - 1 ,
ELECTROMAGNETIC AND ELECTRON-POSITRON FIELDS 251

if rpT and ips belong to different classes. In the former case the operator
/ 3 \
" | must contain derivatives of only even order, while in the latter
TS\ d Xl
case it must contain derivatives of only odd order. From this and from
(21.28) we can conclude that in the case of integral spin

2 ’° n, ° r , F, , . ( — )= f„ (2 L ), 2 o + * )= 2 * .

In a similar manner we can show that in the case of half-integral


spin

-F „ (A j, 20+ *)= 2*+ l.

Therefore, formula (21.29') can be rewritten as

+ F „ |~ 2(j+ k)= 2N ,
d
F rs A (xx—x 2)
dXy
l ± F" ( ^ J Zl(Xl-X2)’ 2 0 '+ * ) = 2 ^ + 1 -
From this we see that in the case of integral spin the lower sign must
be taken, which corresponds to the commutator in the commutation
relations (21.29), while in the case of half-integral spin we must take
the upper sign, which corresponds to the anticommutator in the commu­
tation relations (21.29).
Thus, we have shown that the theorem on the connection between
spin and statistics is a consequence of
ip'(x) = U yr (x)U-1.
The inverse theorem can also be proved, which consists of the
fact that the invariance of the theory of wave fields under the transfor­
mation of total inversion follows from Pauli’s theorem on the connec­
tion between spin and statistics (151), (129).
C H A P T E R IV

Fundamental Equations of Quantum Electrodynamics

§ 22. Interacting Electromagnetic and Electron-Positron Fields

22.1. System of Equations for Interacting Fields


In the preceding sections we began by assuming the independent
existence of the electromagnetic and the electron-positron fields, and
we studied their properties separately, making use of two inde­
pendent systems of equations—Maxwell’s and Dirac’s equations. In
actual fact, however, as a result of the interaction between them, the
electromagnetic and the electron-positron fields form a single dynamical
system, which must be described by a coupled system of equations.
In order to set up this system of equations we can utilize, firstly,
the Dirac equations for the electron wave function ip(x) in the presence
of the electromagnetic field A ^ x ) (cf. § 12)

(22 .1)

and, secondly, the classical equations for the potentials of the electro­
magnetic field in the presence of currents
( 22 . 2)
□ A '= -U ,
where j is the current density four-vector associated with the motion
of the electron. In this expression we must take for the electron current
density in the state tp the expression
J„= (22.3)

which formally coincides with the expression for the current density
in the case of the free motion of the electron. As can be easily shown,
this vector in virtue of the Dirac equations satisfies the equation of
continuity
(22.3')

[2 5 3 ]
254 QUANTUM ELECTRODYNAMICS

and can be used for determining the electromagnetic field to which


the electron in the state y gives rise.
The Dirac equation for a single electron situated in a given electro­
magnetic field, and the wave equation for the potential of the electro­
magnetic field produced by a given electron current enable us to solve
a number of problems, for example, scattering problems, electron energy
eigenvalue problems, etc. However, within the framework of the one-
body problem it is impossible to formulate a general theory of the inter­
action between electrons and the electromagnetic field. In order to
formulate such a theory, we must, as in the preceding chapter, utilize
the concept of quantized fields, and treat the potential of the electro­
magnetic field and the electron wave function not as ordinary quantities
(c-numbers), but as certain operators (^-numbers) operating in particle-
number space and satisfying, as functions of the coordinates and the
time, certain differential equations. We denote these operators by the
bold-face letters and 4h and we adopt equations (22.1) and (22.2),
in which 4* an^ A are regarded as field operators, as the fundamental
system of equations for the interacting fields. In doing this, as has been
already explained in § 19, it is convenient to alter somewhat the defini­
tion of the current density vector and to assume that it is given by

= y [4*> (22.4)

where 4*c = Ci]>, 4*c = are the charge-conjugate operators of the


electron-positron field. This vector, like the vector (22.3), satisfies the
equation of continuity, and can, therefore, be interpreted as the current-
density vector. We shall use this definition of the current density because
it does not lead to an infinite vacuum charge.
In addition to the system of differential equations for the field opera­
tors we must also formulate the commutation relations between these
operators. Since the time-dependence of the operators is determined
by the field equations, the commutation relations can be specified only
at some initial instant of time.
The introduction of definite initial commutation relations is equiv­
alent to prescribing them at identical instants of time t = t'. For such
commutation relations we adopt the expressions
[A^(x), Av(x ')W = 0,
FUNDAMENTAL EQUATIONS 255

M lM L l =0

{ 4*a (*)> *14 (*')}«-«■ = dapd(r-r'), (22.5)

[+„(*)> = °>
8 K ( x ') 1
dt \t=t’
In the case of free fields these relations are equivalent to the general
relations (17.1), (19.4) for arbitrary instants of time t and However,
in the case of interacting fields the general commutation relations for
arbitrary values of i ^ t' cannot be formulated in advance, since in
order to do this we would have to obtain the general solution of the
equations of motion of the coupled fields.
The state of the system of interacting fields is specified by a unit
vector in Hilbert space—the state vector 0°. In our method of descrip­
tion this vector does not vary with time, since the variation in the state
of the system is taken into account by the variation with time of the
field operators (such a method of describing the state of the system is
called in quantum mechanics the Heisenberg picture — cf. subsection 23.1).
However, not all state vectors are admissible: just as in the case
of the free electromagnetic field, the allowable states are restricted
by the subsidiary condition (16.26).
Thus, we adopt the following system of fundamental equations
of quantum electrodynamics for the field operators :
256 QUANTUM ELECTRODYNAMICS

r - x x ^ H ’

= idflvd ( r- r ') .
t = t'

{+„(*), 4»p(JC#)}*=*' = &ap6( r - r ')


[Am(x ), Av(x')]t=t- = 0,
{+«(*) = o, (22.6, cont.)
['I'aW- A^(x')]i=r = 0,
3A„(x) 5A,(x0j = 0;
dt ' dt
dM
A ( x^ ) l =0
dt' \t=t-
We note that from the third and the fifth equations of this system
we can easily obtain the second pair of Maxwell’s equations, which
now means that the expectation value of the operator F/iv + j
ox,.
in the state 0° is equal to zero:

(22.7)
r h r F- + i - r r °-
where F is the operator for the field tensor:
F,iv= (d/dx ) A v- ( d / d x v)A

22.2. Lagrangian. Energy-Momentum Tensor


The equations of motion of quantum electrodynamics can be obtained
from the variational principle 5 f Ldlx = 0, where L is the Lagrangian
density.
We first treat the quantities i A as c-numbers. In this case the
Dirac equation (22.1) for an electron in the electromagnetic field, and
the wave equation for the electromagnetic potential (22.2) can be obtained
from the variational principle if for the Lagrangian density we take
1 dAM dAfl
L = L0-{-Ley
2 oxv dxv

m\y)-\-iey)yvy)Av, (22.8)
7v d'x.
F U N D A M E N T A L E Q U A T IO N S 257

where L0 is the sum of the corresponding functions for the free fields
A> = Ly-\- Le:
(22.9)
2 dxv dxv
and
L = — ( 22. 10)
2V -mj w + ^ v \ y v- Q - ~ f n )y’>

while L ey is the part of the Lagrangian density which describes the


interaction between the fields:
L e y = iey>YvV>A v ( 22. 11)
Indeed, according to (16.4) the field equations have the following
form (ip, ip, A^ are treated as independent variables):
d dL dL — o
dxfi didAJdxJ dAv ~ ’
d dL dL d dL dL
dxfl d(dip]dxf) dip ~ ’ dxfi d{Snpfdx^ dip
On substituting into these expressions
dL dL dAv
dAv
dL
dip \-V (iey,A =
2 vy”

dL
^ (ie y .A -m ) - 1 (y, ( J L - +m L ^ - — 5 jVY„
dip
we obtain equations (22.1) and (22.2).
We note that the part of the Lagrangian density describing the inter­
action between the fields is equal to the scalar product of the potential
of the electromagnetic field A v and the current density vector j„:
Ley = A X - (22.11')
This vector can be defined with the aid of L in accordance with the
general formula (20.29)
. I dL dL \ . _
-A- ie[d(dip/dxfi) W V dXddpjdx^J
258 QUANTUM ELECTRODYNAMICS

We note that the function Ley is the only invariant in which the
electromagnetic field potentials appear linearly, and the electron func­
tions appear in bilinear form, and which does not contain any deriv­
atives of y and A .
In accordance with the general formula (20.17) the Lagran-
gian density (22.8) corresponds to the following energy-momentum
tensor:
dL dAg dL dy dL dy
d(dAJdxv) dx^ d(dy/dxv) dxM d(dy/dxv) dxfl

It represents a sum of expressions for the energy-momentum tensors


of the electromagnetic and the electron-positron fields.
In order to obtain with the aid of the variational principle the equa­
tions for the quantized fields we must take for the Lagrangian density
the operator

(22.13)

and in varying the action treat A^,4>,<J> (or A/4,4»c,4,c) as the independent
variables. On noting that

on carrying out the variation, and on neglecting the inessential terms


having the form of divergences, we obtain the expression
FUNDAMENTAL EQUATIONS 259

a = - i <5A„(D A „ + j „ ) + I ( n A „ + y

1
2

1
+ 2 y u 5 —ie A j+ m
dx

1
—m *
2

+ l ( M ^ + ieA ' r m) 'H + = 0 ’

from which the first three equations (22.6) follow. If in carrying out
the variation of L we treat i|F and 4>c as the independent variables
in place of i]> and *j> we obtain in place of the first two equations (22.6)
the equations for the charge-conjugate operators i|F and <pc.
We note that in the expression (22.13) for L the order of arrange­
ment of the field operators is important; if we treat these quantities
as c-numbers, expression (22.13) does not go over into (22.8).
In the theory of quantized fields, just as in c-number theory, the
energy-momentum tensor can be defined by

T dA^ - d dA" dA°


^ 2 \ 8xv 8x^ dx^ 8xv f,v 8xx 8xx

(2 1 1 4 )

which is equal to the sum of the energy-momentum tensors for the


free fields. It can be easily shown that this tensor is a Hermitian operator
and satisfies the continuity equation

-----T = 0.

From this follows the conservation of the energy-momentum four-vector


for the field
P, i f T 4 dr = const. (22.15)
260 QUANTUM ELECTRODYNAMICS

22.3. Field Equations in Poisson Bracket Form


We now show that the operators A^(x) and4*a(*) satisfy the follow­
ing relations:

-q— O) = i[A^(x), P J , (x) = i[v|>a(x), P„]. (22.16)

Since P„ = const., we can assume that in the integral defining P„ all


the operators are taken at the same instant of time as A^(x). We there­
fore obtain on utilizing (22.14) and (22.5) the following result:

*[A „(x),PJ = - i f

- m‘ 4 a ,m [a 'w ’~ k

A „M

We now prove the second of relations (22.16). It can be easily shown


that

= j f dr'^ a (x ),^ (x ,)}— ^ p(x')


~ W ^ ) { ^ ( 4 - ^ 4 / - * ') } - { + » > W * ') } +„(*')

+ -qZT W * ') { + a (4 +/»(*')} ) •


U A iu / t’ = t
F U N D A M E N T A L E Q U A T IO N S 261

From the commutation relations and the equations of motion it follows


that

and therefore

= J ({ W 4 + J M ji+ w )

The integral of the second term vanishes: for ^ = 1, 2 ,3 this is evident,


while for p = 4 it follows from the law of conservation of charge:

— J ip+(x)^(x)dr = — J d r = 0.

Therefore we finally obtain

W .W . / * '{ > ! < . « . < = 4 r ^ ■

We note that by utilizing (22.16) we can prove the validity of the


general relation

— F ( x ) = /[ F ( x ) ,P J , (22.17)

where F (x) is an arbitrary function of the field variables at the point x.


This relation enables us to determine the dependence of the matrix
elements of the operator F (x) on the coordinates. Indeed, on going
over to the representation in which the operators are diagonal, we
obtain, from (22.17),

= ilF IjiC -F f).

where (F (x))12 is the matrix element of the operator F (x) between the
states 1 and 2 of the system for which the energy-momentum vector
of the field is equal to P ^ and Pj^K From this it follows that

(F ( x ))12 = F12 e M' f* (22.18)


262 QUANTUM ELECTRODYNAMICS

where F12 no longer depends on the coordinates. This formula is a natu­


ral relativistic generalization of the well-known quantum mechanical
formula which determines the dependence of the matrix elements on
the time.
22.4. Invariance Properties of the Equations of Quantum Electrodynamics
We now show that the fundamental equations of quantum electro­
dynamics are invariant under Lorentz and gauge transformations, and
under the transformations of space inversion, charge conjugation and
strong inversion.
Invariance under Lorentz transformations is immediately evident
from the fact that the Lagrangian density is a scalar.
By a gauge transformation we mean a transformation of the follow­
ing form:

(22.19)
^ —>eievtp,

where rj is a scalar function satisfying


□ v\ = 0, (22.20)
but otherwise arbitrary.
The invariance of the equations of quantum electrodynamics under
these transformations follows from the fact that as a result of transfor­
mation (22.19) the Lagrangian density acquires an inessential addi­
tional term which can be written in the form
d / d2r] \d d2?] d I dr] d2r] \
\ ” dx»dxv / "~dxv dxl + dxv \ ten dx~dxv!
dr] d dh] . dr]
dxM dxv dx^dx, + ‘//‘ dx^
Here the first and third terms are of the form of four-dimensional
divergences and therefore may be neglected, while the second and
fourth terms vanish in virtue of equation (22.20). Finally, the last term,
which is equal to
(d/dxf (j]j\) ~~V (.djjdxf) ,
may be omitted, since dj jdx = 0.
FUNDAMENTAL EQUATIONS 263

We note that equation (22.20) must hold in order that the potentials
and A'M= A^—(drj/dx^) should satisfy □ = —j^, while the Dirac
equations retain their form for any function r) not even necessarily
satisfying equation (22.20).
We now consider the transformation of charge conjugation

4»(x) ->• «>c(x) = Cty(x),


(p(x) -»• c[>(x) = C F ^ x ), (22.21)
A„(x) -» A£(x) = —A^(x).

Under this transformation the sign of the current reverses in accordance


with (22.4): ->• j£ = —j^, and therefore the Lagrangian density remains
unchanged.
We note that the reversal of the sign of j agrees with the simple
obvious interpretation of charge conjugation as the reversal of the sign
of the charge.
The invariance of the field equations under the transformation of
charge conjugation is a mathematical formulation of the symmetry
of the theory with respect to both signs of the charge, i.e., with respect
to the electron and the positron.
Finally, we prove the invariance of the equations of quantum elec­
trodynamics under the transformation of strong inversion.
In the expression (22.13) for L we substitute x ->• —x:

L ( - x ) = L{ A „f(-x ),
T / . T / A \ + (( - x )W} = -
I
- -
1 5 A / X
f x—
- X )

— 2 " + ( - * ) [ ? / .( — -fx ~ ieA„ ( - x ) j + w j ( - x )

- 2 + ( ~ x ) [yj ( - + /eA„(-x)) - m ]$ (-* ).

Further, on substituting into this expression

<K~x) = cr^x*),
^ ( ^ ) = f W O ’ 1. ( 22 .22)

A ^ —* ) = —Ai«W,
264 Q U A N T U M E L E C T R O D Y N A M IC S

where O is the inversion operator introduced in subsection 21.3, and


on utilizing expressions (21.23)
0 0 = -1 , OyflO = y(l>
Ot Ot = —1, 0 Ty£0 T = r l ,
we obtain
T f , _ i 5a;(x) 3a;(x)
( X) 2 '"dx,

— ( —■ ; +/ eA; w ) (22.23)

- icA;(jr)| + m j <]/(*)•

We now go over from L to the transposed operator LT:

1 5A£(X) 5A£(*)
( X) 2 " " d x v" dxv

—j 4>*(*)[ v l (— + ieK ( x^ ~m] +*(*)

— ~ ■$*(*) ^ | — - - —ieA« (x)J + m j vj>R(x),

where in accordance with (21.25) we have


A«(X) = A 7(x); 4>*(x) = 4>'TW (22.24)
and the arrow over 5/Sjc denotes that the derivative is taken of the
expression standing to the left, and not to the right of d/dxft. Since
the Lagrangian density appears under the integral sign we may change
the “direction” of differentiation in LT if we simultaneously reverse
the sign of dldx , i.e.,
U ^ (x )^ (x )
LT{A,<(-*)=4»(-*)} 2 dxv dxv

j + * (* ) (■dl - + /cA« (*) j - m ] + » (x)

~ + “(x) | ) „ ( , ( -ic A ;,; U ) |+ m j + ' , (Aj.


F U N D A M E N T A L E Q U A T IO N S 265

A comparison of this expression with (22.13) shows that LT(—x) can


be obtained from L{A(x),i|>(x)} if the following substitution is made
in L(x):
+ (*)-<!>*(*) =<KT(x),
(x) -+ (x) = (x), (22.25)
A^(x) A « (x )= A J(x),
i.e.,
M a^ —x),4>(-x)})T = L{A£(x),<K(x)}.
From this it follows that the operators vJ>K(x),4>B(x) and A^(x) satisfy
the same equations as the operators*]>(x),t|>(x), A^(x). In other words,
the equations of quantum electrodynamics are invariant under the
transformation of total inversion.
It can be easily shown that the commutation relations (22.5) are
also invariant under the transformation of strong inversion.
The necessity of supplementing the transformation of inversion by
the transition to transposed operators is clearly seen using as an
example the equation
□ A /i(x) = - j ^ x ) .
The left-hand side of this equation reverses its sign under the trans­
formation of inversion, while in accordance with (21.3) the current
density does not change its sign:
□ a ;1(*) = &(*).
On the other hand,
= - m w -
Therefore, we have
□ A j?(x)=
where
j/?(*)= - # ( - * ) •
Thus, in order to preserve the invariance of the equation connecting
A with i , simultaneously with carrying out the transformation of
inversion we must also go over to the transposed operators.
Thus, the equations of quantum electrodynamics are invariant not
only under Lorentz transformations (which also include the trans­
formation of inversion in ordinary space), but also under transfor-
266 QUANTUM ELECTRODYNAMICS

mations of total inversion R and of charge conjugation C. They are


therefore also invariant under the product of these fundamental trans­
formations. In particular, the equations of quantum electrodynamics
are invariant under the transformation J = RC, which is called weak
inversion. In accordance with formulas (22.21) and (22.25) this trans­
formation has the following form:

A^(*) A£ 0 ) = AJ ( —*)>
<1*0) x))T, (22.26)

- ^ O ) = O M -* ) £ ) T»
where Q - 1 = OC = £y5C. In virtue of (10.2) this matrix satisfies the
conditions
Qy/iQ ^ 1 = y£ (// = 1,2, 3, 4, 5,);
QQ+ = l, QT= - & , (y5y/) T= - ^ y s y ^ - 1, y ^ = —
where
V i^W n V v-V rV ^-
It can be easily shown that in contrast to the transformation of
total inversion which reverses the sign of the current density
j/?(*)= - # ( - * )
the transformation of weak inversion does not reverse the sign of the
current density
# (* ) = £ ( - * ) •
This is associated with the that fact that the current density reverses
sign both under the transformation R and under the transformation C.
From this we can conclude that the transformation of weak inversion
is analogous to the space-time inversion in classical electrodynamics.
Indeed, since under the transformation xfl -» — x the velocity of
the particles is not changed, the current density also does not change
sign, and from this it follows in turn that under the transformation
-> —xtl the potential of the field should also remain unchanged,
since the Lagrangian for the particle must be invariant

L = —mc2] / \ —(v2/c2)-{-^-Av — e(p

(<p is the scalar potential, m is the particle mass).


FUNDAMENTAL EQUATIONS 267

From the form of the energy-momentum tensor (22.14) it follows


that both the transformation of total inversion and the transformation
of weak inversion leave TM „ invariant

T* (*) = T* ( - * ) and T£,(x) = ! £ ( - * ) .

Thus, the principal physical difference between the transformations


R and J consists of the fact that the former changes the current density
while the latter leaves it unchanged.
The operator J, as well as the operator O, can be represented in
the form of a product of the operators of space and time inversion which
we denote by P and T:
J == PT = RC. (22.27)
In this case the operator P is defined in the same way as in (21.21),
i.e.,
Py>(r, t) = O ry>(—r, t).

As regards the operator T, it is defined by formula (22.27). We note


that T ^ O , .
Since the square of each of these operators differs from the identity
operator at most by a phase factor, it follows from (22.27) that R = CPT,
T = RCP.
The product of the operators of space inversion and charge conju­
gation is called the operator of combined inversion (114),
K = PC.
The equations of quantum electrodynamics are invariant not only
under proper Lorentz transformations, but also under the transfor­
mation of space inversion P and charge conjugation C, and therefore
they are also invariant under the transformations T and K. However,
in the general case of arbitrary interacting fields we may require only
the invariance of the theory under the proper Lorentz group (without
the transformation of space inversion), and also under the transforma­
tion R.
In particular, it is well known that in the case of /3-decay parity is
not conserved, i.e., the theory is not invariant under the transformation
P (123). However, it may be assumed that in this case the invariance
under the operation of combined inversion is preserved-(114).
268 QUANTUM ELECTRODYNAMICS

§ 23. Equations of Quantum Electrodynamics in the Interaction Picture.


Invariant Perturbation Theory

23.1. Heisenberg and Schrodinger Pictures. Interaction Picture


It is well known that in quantum mechanics the state of a system
is described by a unit vector in Hilbert space—the state vector. In
this description the variation of a state with time consists of the
variation of the components of the state vector along the coordinate
axes. This variation can be described in two different ways (49). In
the first method of description the state vector is stationary while the
coordinate system rotates. Such a description is referred to as the
Heisenberg picture. In the second method of description the coordinate
system in Hilbert space is fixed, while the state vector rotates. Such
a description is referred to as the Schrodinger picture.
In the Heisenberg picture the operators corresponding to different
physical quantities vary with time, while in the Schrodinger picture
the operators do not depend on the time, but the state vector or the
wave function W (t) varies with time. This variation is determined by
the Schrodinger equation:

.dWit)
= H ^ (t) , (23.1)
dt

where H is the energy operator for the system.


The transition from the Heisenberg to the Schrodinger picture
is given by the canonical transformation

W ( t ) = e ~ i*t&°, (23.2)

where 0 0 is the constant wave function in the Heisenberg picture, while


F and F are operators corresponding to some physical quantity in
the Schrodinger and in the Heisenberg pictures. The determination
of the explicit form of this canonical transformation is equivalent
to a complete solution of the corresponding quantum mechanical
problem.
It follows from (23.2) that

F= H=H. (23.2')
FUNDAMENTAL EQUATIONS 269

On differentiating this relation with respect to time we obtain the


following equation:

- / • - |- F = [H ,F ], (23.3)

which determines the variation with time of the Heisenberg operators.


We denote by 0 ° the wave function of the stationary state of energy
En in the Heisenberg picture.
It satisfies the equation
H0»n = n n
The wave function ^ n(t) in the Schrodinger picture which corre­
sponds to 0°n has in accordance with (23.2) the form
Wn( f i = e - us»i€fi.
The matrix elements corresponding to some physical quantity
F are the same in both pictures:

(®S. F ® J.)= (T „, (23.4)


On substituting into this expression Wn(t) = e~xEnt0°n, we obtain
the dependence of the matrix elements on the time

(0°n, F0°n.) = (0°n, F0°n.)eilEn - En^ . (23.4')

In addition to the Heisenberg and the Schrodinger pictures another


intermediate picture is possible, viz., a description of the system in
which both the coordinate axes of Hilbert space and the state vector
rotate, with the coordinate axes rotating in a prescribed manner. Such
a picture is frequently used in quantum mechanics for the approximate
solution of problems by means of perturbation theory. It was intro­
duced into quantum electrodynamics in a relativistically invariant
manner by Tomonaga (197) and by Schwinger (172) and has received
the name of the interaction picture.
In order to explain the introduction of this picture we consider
v
a quantum mechanical system with the Hamiltonian H and decompose
v v
it into the two terms H0 and Hx which we call the free Hamiltonian
and the interaction Hamiltonian
V V V

H = H0+ H 1.
270 QUANTUM ELECTRODYNAMICS

We carry out the canonical transformation

&(t) = ei*I‘tlP(t), and p = emttpe-iH0t^ (23.5)

which, in contrast to the transformation (23.2'), involves not the whole


Hamiltonian of the system H but only a part of it, H0. We now determine
the variation with time of the function 0{t). On differentiating 0(t)
with respect to time and on utilizing (23.1) we obtain

i J L 0 = - H o0 + e '« .1(Ho+ H 1) e - ® ^ =
at
i.e.,

1-^0= ^$, (23.6)

where
Hl=

We see that the function &(t) satisfies the Schrodinger equation


with the Hamiltonian Hx. We refer to this function as the wave function
of the system in the interaction picture, and we regard the canonical
transformation (23.5) as the transition from the Schrodinger picture
to the new interaction picture.
We now investigate the way in which the operators vary with time
in the interaction picture. By differentiating the second relation (23.5)
with respect to time we obtain

F = H0F - F H 0,

and since in accordance with (23.5)

H0 = H0,
we have

- / — F = [H0, F]. (23.7)

In contrast to (23.3) these relations involve not the whole Hamiltonian


H, but only a part of it, H0.
F U N D A M E N T A L E Q U A T IO N S 271

Thus the fundamental equations of quantum mechanics in the


three pictures—the Heisenberg, the Schrodinger and the interaction
picture—have the form

. d&° n d¥
1-gj- = 0, — = [H, F] —Heisenberg picture,

dip v 3F
—i—Qj- = 0—Schrodinger picture, (23.8)

.30 5F
i - ^ - = Ht 0 , —i—^—= [H0, F]—interaction picture.

The functions 0°, 0 ( t ) , 0 ( t ) and the operators F, F, F in the three


pictures are related among themselves by the following canonical
transformations:

W(t) = e- im0°, 0(t) = em^W(t),


(23.8')
p* == ^ —iHt p* — e iH0t

23.2. Transition to the Interaction Picture in Quantum Electrodynamics


Quantum electrodynamics is concerned with the investigation
of the dynamical system involving both the electromagnetic and the
electron-positron fields. This system, like any other quantum mechani­
cal system, is characterized by a Hamiltonian function H which can be
written in the form H = H0+V , where H0 = H + H e is the sum of
the Hamiltonians of the free fields, while V is the Hamiltonian of the
interaction between the fields. Such a decomposition of the Hamiltonian
can be utilized for introducing the interaction picture into quantum
electrodynamics. In this picture the field operators satisfy equations
of the form (23.7), i.e., the equations for the Heisenberg operators
of the free field (16.3) and (18.3). Because of this the explicit form
of the field operators is known in the interaction picture, and the com­
mutation relations between them at any time are also known. This
makes the use of the interaction picture particularly convenient for
the solution of specific problems.
272 Q U A N T U M E L E C T R O D Y N A M IC S

We write the fundamental system of equations of quantum electro­


dynamics in the interaction picture

f v a r ; +m) v = 0> v { y* - £ r ~ m) = °'


8
□ A m= 0 , i j ( 0(t)= Y (t)0(t), (23.9)

[Am(x)A ,( x') ] = - ^ T > ( x - x'), {VaW ^ M } = - i S ^ ( x - x ') -


For the investigation of this system we must know the explicit
form of the operator V(/). We now show that
V (0 = — / (x) A m(x ) dr, (23.10)
where j is the current density vector written in terms of the operators
of the electron-positron field rp and xp in the interaction picture in accord­
ance with the general formula (22.4), and A^ is the operator for the
potential of the electromagnetic field also in the interaction picture.
To achieve this we go over to the interaction picture by starting from
the fundamental equations of quantum electrodynamics in. the Heisen­
berg picture, i.e., from equations (22.6) for the time-dependent field
operators. As has been pointed out already, this transition can be
regarded as a canonical transformation. If F and F are operators corre­
sponding to some physical quantity in the Heisenberg picture and in
the interaction picture, then such a transformation is of the form
F = S(0 FS+(0, (23,11)
where S(t) is a unitary operator
s + ( o s ( o = s ( / ) s + ( o = i,
which we shall refer to as the transformation operator. The transfor­
mation operator also relates the state vector 0 (t) in the interaction
picture to the state vector 0° in the Heisenberg picture
0(t)=S(t)0°. (23.12)
If we assume that at time t — — o o there is no interaction between
the fields, then we can impose the following condition on the operator
S (0 : S(—o o ) = l. it follows from (23.12) that
. d0 dS(t) 0 » = / 3S«
S+{t)0.
8t dt dt
F U N D A M E N T A L E Q U A T IO N S 273

By comparing this expression with the fourth equation (23.9), we see


that the operator V(r) is related to S(t) by

V ( / ) = / ^ W (23.13)

It can be easily shown that V (t) is a Hermitian operator. Indeed,


from the unitary nature of S(t) it follows that

^ ( s ( o s +W) = ^ s +W + s w ^ = o.

and therefore
d s (o ds+(t)
V(0 = i
dt
S+(/) = - / S ( 0
dt
i.e.,
V (0 = V+(0- (23.14)
In order to obtain the form of the operator V(t) we determine
the derivatives with respect to time of the field operators in the interaction
picture and compare them with the corresponding derivatives in the
Heisenberg picture. By differentiating (23.11) with respect to time
we obtain
3F ds dS+
S+SFS++SFS+S S - - S(t).
dt ~di dt ’

By utilizing (23.13) we can put this expression in the form

— = S ( 0 ^ - S + ( 0 + / [ F ,V ( /) ] . (23.15)

In a similar manner we can derive the following relationship'-

m = S (< )5 - S + (/)+ /| S ( / ) - ^ S+(0, V(0 j + / - |- [F, V(0]. (23.15')


a,
It follows from (23.15) and (23.15') that

^ ■ = S ( < ) ^ S + ( 0 + / [ v ’. V ( ( ) ] .
dt dt
(23.15")
32A,/
___ S( 0 ^ s +(0 + / [ S(0 ^ sm V (0] + i — [A„, V(0].
3/2
274 QUANTUM ELECTRODYNAMICS

With the aid of these formulas we evaluate the expressions


{y^d/dx^+m)y> and D A ^, which must vanish in virtue of (23.9).
We first of all obtain { y ^ d / d x ^ + m ) ^ :

dxk dt

= s(r) ( ^ ^ + " ,)^s+(0+y4[v,»v(0] = o, k = 1,2,3.

Since (yJd/dxJ + m)ty = ieAvp, we have

S ( o ( y ^ - ^ + ™ j 4 >s+(0 = icS(0A vpS+(0 = iekyi,

and we finally obtain


[V(0, W\ = iey iky>. (23.16)
The operators A^ and y) commute, and therefore it follows from
(23.16) that the operator V(r) must be linear in A . On the other hand,
the commutator of the operators A^ and Av taken at the same instant
of time vanishes. Therefore, we have
[A „ (* ),V (0 ] = 0. (23.17)
We now evaluate [JA^ by utilizing the. second formula (23.15"):

s ( o n A „ s + (/)-;[s (< )— -',- s + (o , v ( o l

- / ■ A [ a, . vo) ] _ ° .

Here the last term vanishes in accordance with (23.17), while the first
term is equal to —j since
S ( /)C A „ S + (/)= -S (r)j„S + (r) = —j„;
further, on noting that in accordance with (23.15) and (23.17)

% = S ( , ) ^ S +(,). (23.17')

we obtain finally
r sa i
FUNDAMENTAL EQUATIONS 275

We can conclude on the basis of this relation that the operator V(t)
must be linear in j;i. Thus, V(/) involves both operators A m and j linearly.
We now show that both relations (23.16) and (23.18) are satisfied
if we assume that
V (0 = —/ j„(*) \ (x) dr.
In order to do this we use the commutation relations for the field
operators in the interaction picture at the same instant of time, which
can be easily obtained from the corresponding commutation relations
in the Heisenberg picture with the aid of formulas (23.11) and (23.17'):
[A /x ), Av(x')]t,=t = 0,

\ \ ( x ) , ^ A v( x ' ) \ ^ = i d ^ d i r - r ) , (23.19)

{Va(*)» V>t = &aft &(r-r') ■


By utilizing the third of these equations we obtain

i[y, V(0] = —/ J A „(r', t)[y>(r, t),]v(r', t)]dr'


= e f A„(r', t)(yv)ap{y>(r, t), y>(r', t)}y>p(r', t)dr' = eyt Aip,
as is required by relation (23.16). Further, by utilizing the second rela­
tion (23.19) we obtain

- i f J > ', A,(r', <)] dr' = - j > ) ,

i.e., the relation (23.18).


We must still prove the equivalence of the commutation relations
(23.19) in the interaction picture involving different instants of time,
and the commutation relations in the Heisenberg picture for the same
time. In order to do this we note that the solution of the equation
C A ^ = 0 is uniquely determined by specifying A^(x) and dA^(x)/^/
over a certain hypersurface t = t0:

A > )= f ( d (x ^ x' )X P - - A ^ ) a"(a,---) *'• (23-2°)


We can easily show this, if we recall that

= d(r). (23.21)
£=0
27 6 QUANTUM ELECTRODYNAMICS

In a similar manner the solution of the equation (y J ^ j d x ^ + m ) ^ = 0


is uniquely determined by specifying ip (x) over the hypersurface t — t0:
Va( x ) = - i f S ap (x—x') ( y 4)p e y>e (p c ') dr', (23.22)
t ' = t 0

where

Formula (23.22) immediately follows from the fact that


S(r, 0) = zy4 <5(F). (23.22')
We now derive the commutation relations (23.9) from the commu­
tation relations (22.5). We evaluate the bracket [A^(*), A„ (*')]• By
utilizing (23.20) we obtain
dA„(x")
[A„(x), A,(*')] = J dr'
t " = t L J

- / [ A „ ( x ) ,A ,( * " ) ] ~ ! ^ 7 * " .
t " = t

In accordance with (23.19) this expression is equal to

[A„(x), A,(x')] = ')


J D(x'-*")fA,(x),iA^!l dr" = — id D(x—x'),
z"=t *-

which is what we set out to prove.


Further, we evaluate the bracket {xpa(jc), xpfr(x')}. On utilizing (23.19)
and (23.22) we obtain
f e w >v>t (*')} = —i { / S ^ i x - x ' Y y ^ ^ ^ x ^ d r " , y>+(*')}
r= t'
= —» / (Vhfe")»V#0 0} dr" = - i S ,oe(.v-x')(y4)tj, ,
t"=t’
from which it follows that
fe W > w t 0 0 } = —iSap(x—x') .
Thus, we have obtained the commutation relations in the interaction
picture from the commutation relations in the Heisenberg picture.
In concluding this subsection we note that, as can be seen from
a comparison of (23.10) with (22.11'), the operator V(t) can also be
represented in the form
V(<) = - / L „(x )* , (23.23)
F U N D A M E N T A L E Q U A T IO N S 277

where L ey(x) is the part of the Lagrangian density which describes the
interaction between the fields. This result can be obtained by means
of a direct derivation of the Hamiltonian from the Lagrangian. Relation
(23.23) holds in virtue of the fact that Ley contains only the “general­
ized coordinates” A , y>, and does not contain their time derivatives.
The transition to the interaction picture can also be carried out by
starting with the fundamental equation of quantum electrodynamics
in the Schrodinger picture

where W(t) is the time dependent state vector of the system of fields
and H is the time-independent Hamiltonian of the system. The latter
coincides with the Hamiltonian of the system of fields in the Heisenberg
picture H and is defined by the volume integral of the time component
T44 of the energy-momentum tensor

H= H = - J t u dr.

1 dAa dAa 1 3A, 3A„ 5* s«4>


T44 —
2 dr dr 2 dt dt Vi S*4 dxA

dxi dx4 ^

where Aa and ij* are the field operators in the Heisenberg picture. Since
the field equations (22.6) contain the first derivative with respect to
time of the operators ip and the second derivative with respect to time
of the operators A„, the independent variables in these expressions are 4>,
A a and dAjdt. Insofar as the quantities di|</5x4 (and also cty/cbt:4 and
dtyldx^) are concerned, they are not independent and must be eliminated
with the aid of the first two equations (22.6):

I f ^* =
278 QUANTUM ELECTRODYNAMICS

and analogous equations for the charge-conjugate operators <J>Cand tj>c.


As a result we obtain the following expression for T44:
1 / dAa dAa dAa dA0\
44 2 \ dr dr dt dt J

+ ■?{ - + + m) + + ,i' (y -|t

where
l£ _ —
j „=~2^y^ - ^ cy^c)-
On going over from the field operators F in the Heisenberg picture to
v
the operators in the Schrodinger picture F in accordance with the general
formula
F = e~lHtFelUt
we obtain the following expression for the Hamiltonian of the fields
in the Schrodinger picture
V V V

H = H o+H ,
where

= - J j „ ( r ) A f t ) dr.

The operator H0 is evidently the Hamiltonian for the system of


noninteracting fields, while H 4 is the operator for the energy of interac­
tion between the fields in the Schrodinger picture. Moreover, the field
operators in the Schrodinger picture satisfy the same commutation
FUNDAMENTAL EQUATIONS 279

relations as the field operators in the Heisenberg picture at the same


instant of time
{xp(r), yj(r')} = y 4d ( r - r ' ) ,
{v(»0. #0*')} = 0,
[A„(r), A r( r ' ) ] = i i „ i ( r - r ' ) ,

[A^r), A„(r')] = 0,

[ \ ( r ) , A v(r')] = 0.
We now go over to the interaction picture in accordance with the
general formulas (23.8” ). The state vector for the system of fields &(t)
defined by
0 ( t ) = cl5-e4»(0,
will then satisfy the equation

where
V(0 = e‘H„(H ie-» v = _ J dr

and the field operators A^(x), y a(x), ^ (x ) are defined in the interaction
picture by the relation
jp1-- glHoi —Wot

In virtue of the cannonical nature of this transformation, the commu­


tation relations between the field operators at the same instant of time
in the interaction picture coincide with the commutation relations in
the Schrodinger picture. Therefore, the equations of motion for the
v
fields in the interaction picture —i(dF/dt) = [H0, F] are, as can be
easily seen, of the same form as the equations of motion for the free
fields (23.9). Indeed, we have

- i^ = [H0, V] = - ^ [ ^ ( r ) , H 0]e- ^

_ giHn ty^ y g - i H
280 QUANTUM ELECTRODYNAMICS

Similarly we have

= where j = 1,2,3,

□M*) = o.
Since the field operators in the interaction picture satisfy the equa­
tions for the free fields, and the commutation relations between them
at the same instant of time have the same form as the commutation
relations between the free fields, it follows that the commutation rela­
tions between the field operators in the interaction picture coincide
with the commutation relations for the free fields at different instants
of time.

23.3. Charge Conjugation Operator


The charge conjugation transformation (22.21) corresponds to a trans­
formation of the state vector 0° of the system of fields, i.e., of the
system of electrons and photons. If we write this transformation in
the form
0 Oc = A 0 o, (23.24)
where A is the transformation operator which we shall call the charge
conjugation operator, then in accordance with the transformation law
for the operators, corresponding to the transformation of the wave
function, the following relations must hold:
rp -> y)c = AtpA-1,
(23.25)

or
AxpA x= Cip, -i_
AtpA-1 = C - Xy) and ylA^/l-1 = —A^.
FUNDAMENTAL EQUATIONS 281

We now demonstrate the manner in which the explicit form of the


charge conjugation operator A can be obtained (214). This can be done
in the interaction picture, since in this picture the explicit form of the
field operators is known.
We write the operator A in the form of a product of operators acting
respectively on the photon and the electron variables:
A = AvA et (23.26)
and we use the expressions for the field operators in the form of expan­
sions (16.13), (18.7). In this way we obtain the expressions

A yCX— C;.^ ) A yCt — A y,


A earA ? = br , A . i + A ? = b+, (23.27)
A e\ A ~1= ar, A ' b + A - 1 = a+,
relating the operators A and A e with the creation and annihilation
operators for photons, electrons and positrons.
We first of all determine the form of the operator A y. It can be written
in the form of the product
A y = X \ A yl,
X

where the operator A x acts only on the photon variables X = (k , X)


and satisfies the following relations:
A yX CX — CX A yX >

A vx 4 = - C t A y X,
(23.28)
A y X C X' = CX'A y X ’

A yxcr = Ct ’ Ayx j where X ^ X.

Since the operator A x commutes with the operator N x = c£cx, it


can be regarded as diagonal in the representation in which N x is diagonal.
From (23.28) it follows that the values of A yX corresponding to two
consecutive values of N; , are related by
(A yx) nx+X= ~~ (A yx)nx ■
From this we can conclude that
(23.29)
282 QUANTUM ELECTRODYNAMICS

Therefore, the operator A y has the form


(23.30)
where N = Z N X is the total number of photons.
We now obtain the form of the operator A e. Just as in the case of
the operator A we write it in the form of the product

,= rM „,
a
r
where the operator A„ refers to the electron variables r and satisfies
the relations
A „ z , A - ' = br, A,rbrA - ' = z , . (23.31)
In place of ar, br, a+ , b+ we introduce the operators ar, fiT, a+, /?+:

a, = — (a,+b,), o+ = ) ( a ) + b/j,
]/2 |/2

ft = -i =l (a -b,), ft+= ~
]/2
(at - b i ) ,
which, as can be easily verified, satisfy the same commutation relations
as the operators ar, br.
Expressions (23.31) defining the operator A er can be rewritten in
terms of aT and /9r in the form
A„a,A-} = a„ A J,A ~ i= ~ p ,. (23.31’)
We see that the relation between the operators A er and /?r is the same
as between the operators A x and cx. Therefore, on assuming that the
operator A er is diagonal in the representation in which is diagonal
we immediately obtain in analogy with (23.29)

A er= ( - l f ^
and, consequently,
ZPtPr
Ae= ( - l)r (23.32)
Since the eigenvalues of the operator are equal to unity or
zero, we have (—1)PtPt — 1—2/9+^r and the operator A e can be written
in the form
A = [] (l —2.PiPr) — /7 (1 —a+a,—b+br+a+br+b+ar). (23.33)
F U N D A M E N T A L E Q U A T IO N S 283

This last expression shows that the result of operating with A e


on the state vector 0 for the system reduces to the following: it transforms
each state occupied by an electron into a state occupied by a positron
(of the same momentum), and vice versa, and does not change the
unoccupied electron-positron states.
It can be easily shown that the operator A commutes with the energy,
momentum and angular momentum operators for the system, and
anticommutes with its total charge operator Q. Therefore, the states
of the system can be eigenstates of the operator A only in the case
when g = 0.
Since two consecutive applications of A are equivalent to the identity
transformation, i.e., A 2 = 1, the eigenvalues of A are equal to
A = ± \. (23.34)
Therefore, for given values of the energy and the momentum of a neutral
system we can distinguish two states corresponding to different values
of “charge parity” A = ±1. The state is said to be charge-even, if
A = + 1 , and charge-odd if A = —1.
For a neutral'system the law of conservation of charge parity holds.
Relation (23.30) shows that a system consisting of an even number
of photons is charge-even, while a system consisting of an odd number
of photons is charge-odd. Therefore, an arbitrary neutral system, if
it has a definite charge parity, can decay either only into an even or
only into an odd number of photons. For example, a 7i°-meson decays
into two photons, and therefore the 7r°-meson is charge-even.
We consider a system consisting of an electron and a positron.
The state vector for such a system can be represented in momentum
space in the form of a superposition of three states:
0= 2 A (P’ ai >ai) aX (P) K ( ~ P ) 0o,
Pi* Oi.Oi

where 0 Ois the vacuum state vector; p, and —p,(t2 are the momenta
and the spin variables of the two particles, whose total momentum
is assumed to be equal to zero.
The result of operating with A e on 0 is in accordance with (23.27)
A e0 = £ A(p, av o2) A ea+(p)Aj1-Aeb + ( - p ) A j 1A e0 o
p , Oi , Oa

= 2 A(p,<J1, o 2)b + (p )a + (-p )0 o,


P , o x,a%
284 Q U A N T U M E L E C T R O D Y N A M IC S

and, since a+ and b+ anticommute, we have

A e0 = - £ A (P>a1 »tf2) < (—P) GO


p , a i , o-i

= - 2 A ( —p , a a, a 1)si+(p)b+(—p ) 0 o.
p,OitOi

Thus, the result of operating with A e reduces to the reversal of the


sign of A and the exchange of the momenta and the spins of the two
particles:
A eA(p, crx, c2) = —A ( —p, <ra, ffj.

This expression can be rewritten as


A ey>(l>2)=-y>(2, 1), (23.35)
where ip is the wave function of the system, while 1 and 2 denote the
complete sets of variables of the two particles in momentum or config­
uration space.
It follows from (23.35) that the symmetric states of the electron-
positron system are charge-odd, while the antisymmetric states are
charge-even.
If I is the relative orbital angular momentum of the electron and
the positron, while s is the total spin of the system (s — 0,1), then
y l = ( - l ) '+ * . (23.36)
In particular, the electron-positron system in the state 1S(l = 0, s = 0)
can decay only into an even number of photons, while the same system
in the state 2S(l = 0, s = 1) can decay only into an odd number of
photons.
We note that not every neutral system can be characterized by
definite charge parity. For example, we consider the hydrogen atom.
We denote its wave function by ipH. Then AipK — ip^, where ip^ is the
wave function for an atom of “antihydrogen” consisting of an anti­
proton and a positron. The functions ipg and ipK are not connected by
any simple relation, and the eigenfunctions of the operator A are given
not by ipH and ip^, but by the combinations ip^-^ip^.

23.4. Perturbation Theory


In quantum electrodynamics the interaction operator in accordance
with (23.10) is proportional to the electron charge e which can be treated
FUNDAMENTAL EQUATIONS 285

as a small parameter. Therefore, the solution of the fundamental equation


of quantum electrodynamics in the interaction picture
d&(t)
(23.37)

can be sought by using the method of successive approximations in


the form of an expansion into a series in powers of the parameter e.
We write &(f) in the form

0 ( O = S ( M o W o), (23.37')

where 0 ( t o) is the value of 0 (f) at some initial instant of time t = f0,


while S(f, t0) is an unknown operator which obviously satisfies the
equation
V (f)S(b t0) (23.38)

and the initial condition


S(f„.fo)= 1. (23-38')
We note that if at / = - o o the wave function 0(f) coincides with
0°, 0 ( — o o ) = 0°, then the operator S(r, — o o ) goes over into the
transformation operator S(f) introduced earlier (cf. (23.12)).
We seek the solution of equation (23.38) in the form of a series
in powers of the charge e :
(X»

S(f,f0) = £ Sfc(f,f0), (23-39)


k= 0

where S ^ , t0) is proportional to ek. On substituting this expansion


into (23.38), and on equating terms of the same order, we obtain,
on taking into account the initial condition (23.38'):
t
s*(/, t0) = - i / v(OVi(^
( (23.40)
So(t> to) — 1•
On applying this formula k times we obtain
t U tk -1

Sk(t, t.) = ( - 0 * / * . / * » • • / * .V W V W ... V(I,). (23.41)


Co ^0 (o
286 QUANTUM ELECTRODYNAMICS

The A>fold integral appearing in this expression can be conveniently


represented in a somewhat different form. For the sake of simplicity
we first discuss the term S2(t, t0):
t tl
s 2(f, t0) = - f d t j ^ 2V(OV(^2). (23.41')

In the (tly t2) plane (tx is the abscissa, t2 is the ordinate of a point) the
region of integration in (23.41') will obviously be given by the triangle
situated below the bisector of the angle between the coordinate axes
(Fig. 11).

Fig. 11.

We now interchange the variables t1 and t2.


t t2

S2(f, 0 = —f dt2f ^ V ( t 2)V (M -

If in the (tlf t2) plane we, as before, regard t1 as the abscissa and t2
as the ordinate, the region of integration will now be given by the triangle
situated above the bisector of the angle between the coordinate axes.
The integrand will now differ from the previous integrand of expression
(23.41') by a change in the order of the factors. If the operators V(/1)
and V(t2) commuted, then both integrands would coincide and S2(t, t0)
would be given by half of the integral over the whole square. We can
do this also in the general case of noncommuting operators and
V(/2) if we introduce the chronological operator T(cf. subsection 17.2)
V(M V(t2), t2 <
r(vw vw ] = { (23.42)
V(r2) V(tl), tx < t2.
FUNDAMENTAL EQUATIONS 287

Then we have
t t

S 2( t , t 0) = - I f d t j d t i T f V O J V Q ) .
to t0

In a similar manner it can be shown that in the general case Sk(t, t0)
has the form (52)
t t t

'o) = / d t ^ d t , ... J dtk T{\{h) V(t2) ... VC/*)), (23.43)

where T is the chronological operator w'hich arranges the factors in


such a way that the time arguments of the operators V(f) decrease
from left towards right.
This solution can be interpreted as follow's. We write the formal
solution of equation (23.37) in the form

0 ( 0 = (exP ( - i f V ( t) d t ) j0 ( t o). (23.44)

This expression is a formally correct solution of the equation for 0(t)


if the values of the operator V (/) at different times commute. However,
if two operators A and B do not commute, then e x p (A + B )^ e x p
A-exp B. Therefore formula (23.44) cannot be used directly in the case
when the operators VOO, V (?2), ... do not commute. In order to
find out how to proceed in this case, we note that’ the solution of
equation (23.37) can always be written in the form
oo Tr‘
0(0= lim / 7 ( l - / f
V ( t) d t ) 0 ( t o),
T»“"T»+i "=0 ni+i
where r„ are times decreasing with increasing n and tending towards
r 0 = t > r x > t 2 > ... > = t0.
This expression shows that the function 0 ( t o) is first of all acted upon
by the operator V(f') at the time t = t0, and is only then acted upon
by the values of the same operator at later times. Therefore, expression
(23.44) can be used, but we must preserve the correct order of operating
with the values of the operator V(/) at different times: the values of
the operator referring to earlier times must be applied before those
288 QUANTUM ELECTRODYNAMICS

corresponding to later times. In order to express this we introduce


at the beginning of the right hand side of formula (23.44) the chrono­
logical operator T :

0 (0 = r(exp ( - 1 / V (0 <//')) 0 ( / o) (23.45)


^0
or
0 ( O = S ( M o) 0 ( r o),
where
t

S (t, Q = r ( e x p ( - / f V ( 0 dt'fj. (23.45')

If the exponential is expanded into a series, we obtain


t, i,

S 0 j t0) — v ( - 0* f d t , . . . f dt, r ( v ( 0 ... V (f„)). (23.46)


k[

Perturbation theory can also be used to obtain the solutions of the


fundamental equations of quantum electrodynamics in the Heisenberg
picture. On expanding the field operators A and into series in powers
of the electron charge
A ^ = A£0) + eA*41) + e2A*J2)- j - . .. ,

= + (23.47)

and on substituting these expansions into (22.6) we obtain, equating the


coefficients of like powers of the charge,

V ^Jx +mW(0) = °>


’ □ * ? . = <>.
, , . » (23.48)
[y^ +mr n+i)=l2 £ {A/”,)’ ^ <n' n')}’
n
□ A^n+1) = — ~ ^ [4><n,), where n = 0, 1 , 2 , . . .
FUNDAMENTAL EQUATIONS 289

These equations enable us to express 4,<1); A),11 in terms of <j>(0), ;


4»(2), A£2) in terms of4»(1), A),1*, etc., i.e., we can finally express all the
4>(n), A<tn) in terms of 4»(0), A4®*. Therefore, in order that the problem
of obtaining the field operators should be completely unique, we must
also specify the commutation relations which must be satisfied by
the operators <]><0) and A},0*. (Due to the existence of recurrence relations
it is not possible to specify general commutation relations between
the field operators, as we have seen earlier.) But <|»(0) and A)®5 are the
operators for the free fields, and the quantization rules satisfied by
them are known.
We now demonstrate the method of solving the recurrence equations
(23.48). We note, first of all, that the particular solution of the inho­
mogeneous Dirac equation

( y>‘Jx~ + m ) V ^ = (23.49)
where <p(x) is a spinor which vanishes at t = —o o , can be written in
the form of the following integral over four-dimensional space:

Va(*) = J S l a i x - x l y A x ^ d + x ’, (23.50)

where the Green’s function, or more accurately the Green’s matrix S R,


is defined by
(23.50')

Indeed, on making use of (23.50'), we obtain

( y, ‘ 1 (D — x'

but according to (17.49')


{ U - m * ) A R{ x )= - 6 { x ) ,
and therefore

Since the function A B{?c) vanishes for t < 0, the solution (23.50) vanishes
at t = — o o .
290 QUANTUM ELECTRODYNAMICS

In a similar manner the particular solution of the inhomogeneous


equation
□ AM( x ) = - a M(x), (23.51)
which vanishes at t = — oo, can be written as

A / x ) = j D r ( x - x ' ) afix^d^x’. (23.52)


Formulas (23.50) and (23.52) enable us to obtain in turn solutions
of equations (23.48) vanishing at t — — oo. It can be shown that these
solutions are of the following form (171), (172):
t t-i

4><n)(x) = i n J d i x 1 J d i x i ...
—oo — oo

... )'<***,,[ u™<x»),(... ...]],


—oc
t t,
(23.53)
A<f >(x) = i n J d i x 1 J d i x z . ..
—oo —oo

... J d4xn[u<°)(xn),[... A<°’( * )]-1 J ,


—oo

where
u « 'M = -p W A l" w ,

j r W = - j t $ “” W . y > ,0,W]-

§ 24. The Scattering Matrix

24.1. The Scattering Problem and the Definition of the Scattering


Matrix
In quantum electrodynamics, as in quantum mechanics, there are two
fundamental types of problems that have to be solved. Problems of
the first type consist of determining the energy levels of a system, while
problems of the second type are collision problems. We shall in future
concern ourselves with problems of the second type. They can be formu­
lated in the following way. Given the state of the electromagnetic and
the electron-positron fields at the time t — — oo, it is required to
F U N D A M E N T A L E Q U A T IO N S 291

obtain the state of the fields at t = oo. In the interaction picture which
we shall use these states are described by the wave functions 0 ( — oo)
and 0 (oo). In accordance with (23.45) they are related by
0(oo) = S 0 ( —oo), (24.1)
where the operator S is of the form

(24.2)
— OO

This operator which transforms the wave function of the initial state
0 ( — oo) into the wave function of the final state 0(oo) is called the
scattering matrix or the S-matrix.
In quantum mechanical collision theory it is usually assumed that
the initial and the final states of the system are free, i.e., it is assumed
that asymptotically at t = oo the interaction operator Y(t) vanishes
and V(=F oo) = 0. Such an assumption is justified if the collision does
not lead to the appearance or destruction of bound states, since, if
we exclude such cases, then at t = =p oo the particles are far from each
other, and therefore V (=f oo) = 0. Without taking into account the
appearance and destruction of bound states we assume that in quantum
electrodynamics we have in the initial and final states of the system
only free electrons and photons. In doing this we use here the term
“free electron” only in the sense that there is no interaction between
the electron and the quantized electromagnetic field; insofar as an
external electromagnetic field is concerned, it may be present, and
bound states due to the external field can be taken into account (cf.
the expansion of the operator for the electron-positron field in terms of
stationary states in the presence of an external field (18.7)).
We note that, strictly speaking, the electron cannot be regarded
as free even asymptotically at t = oo since the interaction of the
electron with its own electromagnetic field leads to a change in its mass.
We shall discuss this problem in greater detail in Chapter VII.
Since for V -» 0 the state vector in the interaction picture coincides
with the state vector in the Heisenberg picture, we can assume that
0(— OO) = 0°,
and rewrite (24.1) as
0 (oo) = S0°. (24.1')
292 Q U A N T U M E L E C T R O D Y N A M IC S

We characterize the different states of the free fields by the set


of state vectors 0°r, where r is the set of indices specifying the numbers
of electrons and of photons having different values of energy, momentum
and other physical quantities. If at the initial time t = —oo the system
was in the state 0 ° , then at t = oo it will be in the state
0 (oo) = S 0 ° .
Therefore, the probability amplitude for the transition of the system
from the state 0° at t = — oo into the state 0°f at t = oo will be equal
to
(00, 0 (00, S0?) = ( / | S | 0.
( 0 0 ) ) = (24.3)
We see that the probabilities of the different processes are determined
by the elements of the scattering matrix S connecting the corresponding
initial and final states.

24.2. Matrix Elements of Field Operators


In accordance with (23.10) and (23.46) the scattering matrix can
be represented in the form
S = r|e x p (—i f U(x) d4x)j (24.4)
or

S= S (n) = ^4xn^(U(^i) ••• U(x„)), (24.5)


n=0

where
U(x) = - L ey = —jM(x) A ft(x), (24.6)
and the integration with respect to each variable is taken over the
whole four-dimensional space.
Since both the chronological operator T and the interaction opera­
tor U(x) are invariant under Lorentz transformations, the scattering
matrix is relativistically invariant.
If the electron is acted upon by an external electromagnetic field
then we must take for the operators rp, y appearing in the expression
for the current density j the expansions (18.7) in terms of the eigenfunc­
tions of the Dirac equation for the electron in this field.
In many cases the external field is sufficiently weak to be taken
into account by means of perturbation theory. In order to do this
F U N D A M E N T A L E Q U A T IO N S 293

we must replace A/((x) in the expression for U(x) by the sum A ^x)
+A® (x) where A;i(x) is the potential of the quantized electromagnetic
field, while A*(x) is the potential of the external field which, in contrast
to A m(x), is a c-number. Insofar as the current density vector j is con­
cerned, it must be constructed with the aid of the expansions of xp
and xp in terms of free particle eigenfunctions.
An intermediate method of taking the external field into account
is also possible when one of its parts has to be taken into account
exactly by means of the Dirac equation, while its other part is taken
into account approximately by adding it to the expression for the quan­
tized potential. In this case the operators xp and xp are defined with the
aid of the expansions in terms of the eigenfunctions of the Dirac equation
containing only the first part of the external field.
We shall say that a scattering process, i.e., the process of interaction
between the fields, is an effect of the nth order if the element of the
S-matrix corresponding to this process is proportional to en. It is obvious
that all the nth order processes are described by the matrix S(n) which
is the nth term in the expansion of the S-matrix into a series in powers
of the electron charge.
Since the operators A^(x) and j^(x) commute, S(n) has the structure
S<~> = Jit,,,,... ( x 1; *2,...) T l A ^ x O A v(xz) ...)
where K ... (x1; x2, ...) contains a chronological product of the current
density operators. The latter satisfy the equation of continuity, and
therefore
(£>K„, .../<?*,„) = (3 K ,, .../a**) = ... = 0 .
From this we can easily conclude that the addition to T^A^-xq) A„(x2))
of the expression (5299(x1—x2))/(5x1/i dx2/), containing an arbitrary
function <p(x) (cf. formula (17.14)) does not alter the value of S(n), in
accordance with the gauge invariance of the theory.
We now proceed to establish the rules for evaluating the elements
of the scattering matrix connecting any given states.
The individual terms of the expansion (24.5) are integrals containing
products of operators xp(x), xp(x), A^(x). As has been shown in §§ 16,18,
xp(x) is an electron annihilation operator and a positron creation opera­
tor, xp(x) is an electron creation operator and a positron annihilation
operator, and A u(x) is a photon creation and annihilation operator.
294 QUANTUM ELECTRODYNAMICS

In order to obtain the matrix element of the operator ip(x) we use


the general expansion (18.7) of the operator y)(x) in terms of the eigen­
functions of the electron in an external field, and also formulas (18.17)
defining the matrix elements of the electron and positron creation and
annihilation operators. From these formulas it follows that the matrix
elements of the operators y)(x) and yi(x), corresponding to the anni­
hilation and creation of an electron in the state r, are given by
K j ( ° ° ) . V ( x ) & i t ( —° ° ) ) = (0+1 V ( * ) l!+ ) = ¥ r + )( x ) ,

(#i+(o°), y(x) 0 ot ( — °°)| ^ ( W t o l °+) = Wt,0W»


where ^ (r+)(x) is the normalized wave function of positive frequency
for an electron in the r-state, while ^ ^ ( z t 00) and &0+(± °°) are the
wave functions for the system of fields at the time t = ± oo (the subscript
1+ or 0+ denotes the number of electrons in the corresponding state).
The matrix elements for the creation or annihilation of a positron
in the state r are given by
Or Iv»(*) I°r) = ¥r\x),
(oT
- | v W | i r- ) = v (r- ,W , (
where yi(r ' ) is the normalized wave function of negative frequency for
an electron in the r-state.
If the states of the electron can be described by plane waves, then the
matrix elements for the annihilation and creation of an electron of
momentum p and polarization r will be given by

(°pr IVWI !pr) = ur{p) eipi,


V (24.9)
( IvKx) 10+) = ur(p) e - ^ ,
|- wb
where Q is the normalization volume.
Finally, the matrix elements for the creation and annihilation of
a positron of momentum p and polarization r are given by

(ijrlvtolO pJ = - j — v T( - p ) e ~ ipi,
(24.10)
t e l v W I ljr) = v \ —p)eipI.
F U N D A M E N T A L E Q U A T IO N S 295

These matrix elements have the form of normalized Dirac plane waves
of momentum —p corresponding to negative frequency.
In order to determine the matrix elements of the operator A ^x)
we use the expansion of A ft(x) in terms of plane waves and formulas
(16.38) defining the photon annihilation and creation matrix elements
and c+.
From these formulas it follows that the matrix elements of the
operator A^(x) corresponding to the annihilation and creation of
a photon of momentum k and polarization X are respectively given by
i z?(^) sylkx
(OkX\A/t( x ) \ l J
]/ 2coQ
(24.11)
l pW
(UaIa ^(x)[Ou ) Cfl Cip—ikx 5
j/~2coQ
where e(f is the four-dimensional unit polarization vector
g(A )g(A ) _ g(A )2_g(A )2 _ J

(ew is the spatial and e f \ = e[X)ji) is the time component of the vector
e f).
If for the potentials of the electromagnetic field we utilize not plane
but spherical waves, then in place of (24.11) we obtain

(OjATaI I 1jArt) (A jm a)^’


(24.12)
0 « J A /l(x)|0WJ) = (AiJltt)*,
where A jm are the normalized potentials of photon states of definite
momentum and parity (cf. § 4).

24.3. Representation of the Scattering Matrix as a Sum of Normal


Products
We return to the general expansion (24.5) of the scattering matrix.
Since the operators ip, ip, A are expressed as a sum of creation and
annihilation operators for the individual particles, each term in
the expansion (24.5) may be written as a sum of products of creation
and annihilation operators for electrons, positrons and photons in dif­
ferent states. We wish to determine the conditions under which such
products have nonvanishing matrix elements corresponding to some
process / - > / which is of interest to us. If, for example, in the state
296 QUANTUM ELECTRODYNAMICS

i there exists a single electron and no positrons or photons, while


in the state / there exists an electron and a photon, then, obviously, one
of the annihilation operators must “annihilate” an electron in the
state /, two creation operators must “create” an electron and a photon
in the state / , while all the other operators must break up into pairs
with the operators belonging to each pair “creating” and “annihilating”
the same particle.
Such virtual processes of creation and subsequent annihilation of
the same particle introduce considerable complications into the eval­
uation of the elements of the scattering matrix. We shall therefore
attempt to transform the scattering matrix in such a way that there
should be no need at all to discuss such virtual processes. Evidently
the problem is reduced to representing the scattering matrix in the
form of a sum of normal or ordered products of particle creation and
annihilation operators in which the creation operators appear to the
left of the annihilation operators. In the evaluation of matrix elements
of such products the annihilation operators will “annihilate” only
those particles which are present in the initial state, while the creation
operators will “create” only those particles which must occur in the
final state. Insofar as the virtual processes of particle creation and
annihilation are concerned they will clearly not appear in our discussion.
In order to write the scattering matrix in such a form we note that
in accordance with (19.29)
= ieN (y(x)v ^p(x))
and therefore
U(x) = —j^CxOA^Cx) = —ieN{y>(x) A(x')\p (x)j.
(24.13)

From this it follows that the nth term in the expansion of the S'-matrix
may be written in the following form:

S<"> = (■~ e ) n• dlXi ... J d'xnT{N(y(x l ) A(.vt ) ip(Xl))

••• % W A W ^ » ) | • (24.14)
The individual factors of the T’-product appearing in this formula
represent TV-products of field operators relating to the same instant
of time. We shall refer to a T’-product of this type as a mixed T’- product.
F U N D A M E N T A L E Q U A T IO N S 297

Thus, our problem consists of writing a mixed T-product in the


form of a sum of A-products. This may be accomplished with the aid
of the following two theorems (210).
I. A T-product of field operators is equal to the sum of their
A-products in which the operators are paired in all possible ways.
In other words, if Xi, x2, stand for the field operators ip, ip,
then the following relation holds:
T (XlXz — Xn) = dpXi1Xii - X i n
= N (Xl%2 ••• XJ +N( xZX$X3 ••• Xj +^ Xl X^ Xi Xi ••• Xn)
+ ••• +N(xiXlxXx\ ••• Xn)+ ••• + N ( x i x l ... Xn-iX%+ ••• (24.15)
Here the operators ... xin in the right hand side of the first
equation, which is a definition of the T-product, are arranged in chrono­
logical order, i.e., in such a way that the time increases from right
to left, and dp is the parity of the permutation of the electron-positron
operators required to achieve this.
The different letters serving as superscripts of operators in the
right hand side of the equation denote different pairings defined for
adjacent operators by means of
Xi7&= T(XiXk)-N(xtXk)-
Pairings between non-neighboring operators are defined in the
following manner: if the paired operators are photon operators, then
they may simply be written next to each other; however, if they are
electron-positron operators, then they may be placed next to each
other provided we first multiply the TV-product by the parity of the
permutation of the electron-positron operators required to achieve
this order. For example, if %1}X2 , are electron-positron operators, then
NiXiXiXsXtXsXg) = -(xixDixkiWiXiXe)
(since XiXl and XzXi are c-numbers, they have been taken outside the
A-product).
II. A mixed product of field operators is . equal to the sum of their
A-products of the form (24.15), but this sum does not contain pairings
between operators appearing in the same inner A-product. For example,
T(Xi N(x2X3Xd) = N (XiXzXzXi)
+N(xl X 2 XzXd + N(xlX2 Xz xd +N(XiX2X3Xi)- (24.15')
This sum does not contain the pairings xSXs> X%Xh X%X\-
298 QUANTUM ELECTRODYNAMICS

In order to prove theorem I we make the preliminary remark that


by a simultaneous identical rearrangement of factors within T- and
TV-products in (24.15) we do not violate this equation, so that without
loss of generality we may assume that the operators in (24.15) are
already arranged in chronological order.
We now arrange the factors in (24.15) in such a way that all the
creation operators stand to the left of the annihilation operators.1
We shall refer to such an arrangement of the operators as an TV-ordered
one. In order to achieve this we take the TV-unordered creation opera­
tor standing at the left and interchange it in turn with all the
annihilation operators standing to the left of this operator. In the
course of this additional terms with pairings between the operators
being interchanged appear in accordance with the following form ula:

%i%2 = T(x 1X2) = N(Xi%2)+XlXl = ± X2X1+X1XI

We perform the ordering operation also with respect to the other


unordered creation operators. As a result we express the initial T-product
in terms of a sum of TV-products. These TV-products may appear both
with positive and with negative signs, but if the factors within the TV-
products are rearranged in such a way that they again become T-ordered,
i.e., arranged in chronological order, then, obviously, all the TV-prod­
ucts will appear with positive signs and we shall obtain an expression
for the T-product in the form of a sum of TV-products of the form (24.15).
This sum will, obviously, not contain all possible pairings, but only
pairings between all pairs of TV-unordered operators; however, since
pairings between TV-ordered operators, which are at the same time
T-ordered operators, vanish, we can assume that the sum contains all
possible pairings. Thus we have proved the validity of theorem I.
Theorem II can be proved in a similar manner. In the course of the
proof it is only necessary to keep in mind that operators appearing
within the same TV-product need not be interchanged since such operators
are already TV-ordered; therefore there will be no pairings corresponding
to them.

1 For the sake of simplicity we assume that each of the operators Xi is either
a creation or an annihilation operator.
FUNDAMENTAL EQUATIONS 299

24.4. General Relation between T- and N-orderings


We now show that not only the individual terms in the expansion of
the scattering matrix, but also the matrix as a whole can be represented
in the form of an //-ordered operator.
First of all we consider the functional of the electromagnetic field
operators
CO

Fi A }= H f k ^v 2 „ (* 1 , X2, ..., xn)A (xJA (x2) ...


n=0 *' 2 71 12
••• Av„ (xn) d*Xi dix2 ... dixn, (24.16)

where K^\,li 2> , v71 are c-numbers, vk71= 1 ,2 , 3 ,4 , and the summation
is over n and vk. We show that (89)

TF{ A) = A(<?AF{A}), (24.17)


where
1
8^ D ' ( x - x ' ) (24.17')
2 d \ ( x ) <5Av(x')

In this expression the functional derivative with respect to the op­


erator A (jc) is defined as follows. If <5A^(.x)is the increment of the oper­
ator A (x) then the variation of the functional given by

(SFfA^x)} = F{A#1W+«5A#1W } - F { A #1W},

is related to the functional derivative [<5F{ A ^ * ) } ] / ^ ^ * ') ] by

aF{A „M } = (24' 18)

In order to prove formula (24.17) we first show that /1F{A} is obtained


from F{A} if in each term of the expansion (24.16) we form in all possible
ways single pairings between the operators A fi(x) and then sum the
expressions so obtained. In order to accomplish this it is sufficient to
verify
4l(A#l(*1)Av(*2)) = A“(*i )A“(x2),
300 QUANTUM ELECTRODYNAMICS

which follows immediately from (24.17') and (17.10):

A ( A ^ ) A„(a2))

- A a(xi) A a(x2)
= J 'J ' ^>c(x ~ x ) dix d*x ' <5A„(x) <5A„(a ')
y

S
(A^X j) davd(x2- x ' )
— 2 - f f D' (x- X')d ,X d ' X'sA,(x)
+ A ct(x2) d ^ d f a —x'))

= — f f Dc(x—x') (dXv d ^ d i x —Xj) <5(a'-a2)

+ (5a A v (H * ' — Xl ) <K* — * 2 )) < P x d * x ’


= ^ a (xi) A“(x2).
In a similar manner it is possible to show that the functional
(Ak/k\)F{A} is obtained from the functional ,F{A} if in each term of
expansion (24.16) we form in all possible ways k pairings between the
operators A and then sum the expressions so obtained.
From this it follows that theorem I stated above on the expansion
of a F-product in terms of TY-products can be written in the following
form applicable to the operators A^:

yy{A> =

On noting that _£"=0(2lk/£!) — A we obtain formula (24.17).


A formula of the same type also holds for electron-positron
operators, viz., if F{ip,ip} is a functional of the operators ip and \p
of the form

F { y),y } = £ f (x\ , x 2, . . . , x ni; A-;, a 2 , ..., A-;,) ip(ax) ip(a;) ...


n1.n, = 0
V ( X n ) V (x'n) ■d*x l ••• d i x n ,> (24.19)
where the summation is over and n2 and also over the spinor indices
on which G(ni,7lj) depends, then
TF{ip, y} = N(erF{ip, ip}), (24.20)
where
_d
27 = d4x d4a', (24.20')
dya(x) Safl (A * ') “*= a')
FUNDAMENTAL EQUATIONS 301

and the functional derivatives with respect to ip and ip are defined in


the following manner. If dip(x) and dip(x) are increments of the operators
ip(x) and ip(x), then the variation of the functional F{ip, ip} given by
dF{ip, ip} = F{ip+dip, ip+ dip}—F {ip, ip},
is related to the functional derivatives by

SF{V, y } = f - ^ S v S x ) d ‘x + f d ^ x ) - ^ d ‘x . (24.21)

In order to take into account correctly the change of sign resulting


from the interchange of electron-positron operators, the increments
of the operators ip and ip are subjected to the following conditions:
{dip, ip} = {dip, ip} = {dip, ip} = {dip, ip} = 0 (24.22)
or to the equivalent conditions

l w H =o’ l w H =o' (24.22')

{ ^ ) = = d^ x ~ x '>-

It follows from formulas (24.17) and (24.20) that if we have a general


functional F{A/i,ip,ip} of the operators A ^ i p , ^ , then
TF{Am, y, ip} = N(ea+£F{Alt, ip, ip}). (24.23)
We now apply these formulas to the transformation of the S-matrix.
Let the functional i^A ^, ip, ip} be of the form

, ip, ip} = exp (/ j j M(x) A m( x ) d4x ). (24.24)


In accordance with (24.4) the scattering matrix is related to F{A/i, ip, ip}
by
S = TF{Afi,ip,ip} = Tv TAF{A/i,ip,ip}, (24.24')

where the subscripts ip and A on T denote respectively the T-orderings


of the operators ip, ip and A . In accordance with the proofs just given,

Ta F{A/i, ip, ip} = NA(eA• exp (i f j / x ) A ^ x ) </4*)),


302 QUANTUM ELECTRODYNAMICS

where NA denotes the /^-ordering of the operators A^. On noting that

N a m T ( p ? ) exp(/ / ^ (x) d ix) = iN ^ x ) exp (' I A/‘W d 4x j j


it can be easily shown that

N a (a exp (/ j j^ x ) A^(x) d4xfj

= ~ N a ( ^ J d 4x ' J d 4x " D c { x ' - x " ) ) v { x ' ) j v ( x " ) exp(* / A„(*) ^4*)) •

Therefore we have
Ta F{ A, y>, xp) = exp(—| J d4x ’f d4x " j v(x') Dc(x'—x")jv(x")j

X ^ (e x p (z J j M(x) A M(x)d*xj^.

From this it follows that

S = Tv (exP ( - i f d 4x ' J d 4x" j ^ x ') Dc( x ' - x " ) j M(x'')j

X NA(exp i Jj„(x) A„(x) d4x )). (24.25)


In processes in which no photons participate the last factor in (24.25)
is equal to unity, and the scattering matrix assumes the form

S = T V exp( - i f d 4x ' f d4x ” j v(x') D f x ’- x " ) j„(x")). (24.26)

If the motion of the electrons can be regarded as given, i.e., if we


treat the current density j (x) as a given ^-function /^(x), then the scatter­
ing matrix can be written as
S = e x p ( - i J d 4x 'J d4x" J f x ’) Dc{x’- x " ) J v (x"))

X N exp |/ f j M( x ) A M(x)d4xj, (24.27)


where the only operators are the electromagnetic potentials A^(x).

24.5. Symmetry of the Scattering Matrix under Time Reversal


In proving the invariance of the equations of quantum electrody­
namics under the transformation of total inversion R we have shown
(cf. subsection 22.4) that

( M M - *)’ V(—*)})* = L{A«(x), y>fl(x)}.


FUNDAMENTAL EQUATIONS 303

From this it follows that

(j„ (-* ) V - * ) ) T = j * 0 ) A £(x).

On the other hand, R is a transformation which does not alter the equa­
tions or the commutation relations of quantum electrodynamics, and
can be regarded as a unitary transformation; therefore

j£ M A* O) = UTj„(*) A ^ x ) i f - 1,

where U is a unitary operator and, consequently,

A / - * ) = U - '( j,W A„(x))TU. (24.28)


We now use this relation to investigate the symmetry properties
of the elements of the scattering matrix under time reversal.
We write the elements of the scattering matrix in the form

S ^ f = \&f , W c x p [-z f V(t)dt


\ — oo

where
v(0 =
and make the substitution t -> — t. Since in accordance with (24.28)

v (-o = u-w^ou.
we have

(0 / ,r ( e x p [ - / J v ( O ^ ] )
\ ' — oo •

= U „ W ex p [-/ / v (-/)^ ])< f]


\ ' — oo '

0 n u - ^ f e x p f - / J VT(0 ^ ])u (P ,
— OO

where T is an antichronological operator which arranges the field opera­


tors from right to left in the order not of increasing, but of decreasing
time.
304 QUANTUM ELECTRODYNAMICS

The last expression can evidently be rewritten in the form


OO .
(
exp[—/ J VT(0 <*])Utf>,
■—OO

00 \
U 0 / 5 r ( e x p [ - i / VT(0*])ud> ,
—00

U 0 r, ( r e x p [ - z J v ( 0 ^ ] ) Tu ^ j -
' —OO '
Finally, on going over from the transposed scattering matrix to the
ordinary one, and in the course of this interchanging the initial and the
final states we obtain
/ 00 \
0 r ,:r(e x p [—/ J v (0 **])#«
—00

(U<ZC)*, W e x p [-z J V (/)^ ])(U 0 / )*|- (24.29)


—OO '
This is the relation which expresses the symmetry property of the
scattering matrix under time reversal. It can be formulated as follows:
the element of the scattering matrix between certain states and 0 f
is equal to the element of the scattering matrix between the time-reversed
states
0[ = (U0,)* and 0) = (U#,)*,
i.e.,
( 0 f , S 0 t) = (0'f , S01). (24.30)
We investigate the physical meaning of the time-reversed states.
In order to do this we consider a Hermitian operator F{<p(;c)} con­
structed from the components of the fields 99 (x) hand aving the form

F M * )} = 1 2 A

where each of the field operators 9oi belongs to one of the classes + 1 ,
—1 , +£, —e, while the summation is taken over all the permutations
P = I !’ 2’’’’ ) and, finally, the constants Ap are chosen in such a way
\ ••• /
F U N D A M E N T A L E Q U A T IO N S 305

that the summation appearing under the integral sign is symmetric


in the boson, and antisymmetric in the fermion operators, i.e., it does
not change sign on interchanging any pair of the former, and reverses
its sign on interchanging any pair of the latter operators.
We now investigate how F{9?(x)} behaves under the transformation
of inversion x -> —x. By utilizing the notation
<p'(x)= Oq>(—x),
where O is the matrix of the transformation of <j>(x) defined by formulas
(21.28) for the field components belonging to the classes ± 1 , ±e, we
obtain
F {<?'(*)} = J 2 Aptph(x)(p’tt(x) ... (p[n(x)dr.

But in accordance with (21.27)


<p’(x) = U - > T(x)U,
and therefore

F{?>'(*)} = u_1( j 2 Ap7\(x)(pf2(x) ... (pfn(x)dr) U

= U- 1 f [ 2 A p<Pin(x) ••• V i J ^ V i ^ Y d r l J ,
but since

2 A P<pin(x) ... <p,t (x)<pti(x) = ± 2 •••

where the + or the — sign is chosen in accordance with whether F con­


tains an even or an odd number of pairs of fermion operators, we have

Ffa'OO} = ± U -ifF{?(x)})TU. (24.31)


We now consider a state 0 in which F has the value A:
F {<p(x)}0 = X0,
and determine the result of operating with F{ 9?(x)} on the time-reversed
state vector (U0)*. Since F has been assumed to be a Hermitian oper­
ator we have
F { ? ( * ) } (U <£ )* = ( F * { ,p ( * ) } U 0 ) * = ( F t {99(*)}U <Z>)*

= (U U ^ F ^ W JU # )* .
306 Q U A N T U M E L E C T R O D Y N A M IC S

On substituting (24.31) into this expression we obtain


F { < p ( ; c ) } ( U < Z > ) * = ± ( U F { ? ' ( x )}<Z>)*.

We must distinguish two cases: when the integrand in F{<p(x)} is


a tensor quantity belonging to the class —1, and when it is a tensor
quantity belonging to the class + 1. We begin by examining the former
case. If, for the sake of simplicity, we assume that F ^ x ) } contains
the two operators <p?c and we obtain, in accordance with (21.28),
<p+e'(x)<pzc'(x) = (pic(x)<p^e(x),

F{<P'0)} = F{<K*)}>
and therefore
F{(p(x)} (U0)* = - (UFfo<jc)}tf>)* = -2 (U 0 )* . (24.32)
Thus, if F{ 9 >(x)} belongs to the class —1, and the state vector 0 is
an eigenfunction of F{<p(x)} belonging to the eigenvalue 2, then the.
time-reversed state vector (U<Z>)* will also be an eigenfunction of the
operator F{ 9 >(x)} belonging to the eigenvalue —2.
In particular, if in a state 0 the charge of the field is equal to q,
and the component of the spin is equal to fx, then in the time-reversed
state (U0)* the values of these two quantities will be —q and —[x.
In a similar manner it can be easily shown that if F{9 >(x)} is a tensor
quantity belonging to the class + 1 , and the state vector 0 is an eigen­
function of the operator F{9?(x)} belonging to the eigenvalue 2, then
the time-reversed state vector (U0)* will also be an eigenfunction of
the operator F{9 >(x)} belonging to the same eigenvalue 2.
In particular, if in state 0 the energy-momentum vector is equal to
p(p, ie), then this quantity will have the same value in the time-reversed
state (U0)*.
Thus, if
F{<Kx)}0 = 10,
then
2(U0)*, (F belongs to the class + 1 ),
F {<?(*)} (U4>)*
—2(U0)*, (F belongs to the class —1).
By utilizing these expressions we can write formula (24.30) which
expresses the symmetry property of the scattering matrix under time
reversal in the form
(■Pi’ »i“*Is IPf> ) = (Pf , ~ e { , — —eit —/q), (24.33)
F U N D A M E N T A L E Q U A T IO N S 307

where p t and p f are the four-momenta of the particles, et and ef are


the charges of the particles and pLf are the components of the particle
spins in the initial and final states.
Thus, if we interchange the momenta, the charges, and the spin
components of the particles in the initial and the final states, and at
the same time reverse the signs of all the charges and of all the spin
components, then the element of the scattering matrix will remain un­
changed.
Since the whole theory is invariant under the transformation of
charge conjugation, the elements of the scattering matrix are unaltered
when the signs of all the charges are reversed:
(A, ei>/fils lPf , ef >Hr) = (A* ~ ei ’ fJ‘i\S\Pf> - er ^ f)- (24.34)
By combining this relation with (24.33) we obtain
(.Pit A t Pi IS | p f , ef , [if) (p ft &ft Pf IS | A >£i t Pi) • (24.35)
Finally, we note that from the invariance of the equations of quantum
electrodynamics under the transformations P and T the following rela­
tions are obtained:
( A , £i ’ e i t P i \ S \ P f t en ef t P f) = ( - P i t . £ i , t i t P i | S | — P f , e f , Ef , p f ) ,
(24.36)
(A t £i j A t Pi IS | Pf £f Ef, [if) , , ( Pf , Ef, Ef, Pf | S | A , £j Ei p t)
, ,

where p is the momentum and e is the energy of the particle.

§ 25. Graphical Representation of the Elements of the Scattering Matrix.


The Scattering Matrix in Momentum Space

25.1. Graphical Representation of Normal Products


In the preceding section it was shown that the individual terms in
the expansion of the scattering matrix in powers of the electron charge
are given by integrals of mixed T-products, and that these T-products
can be expanded into a sum of normal products of the field operators.
We shall represent each such normal product, and consequently
every scattering process, graphically1, viz., we shall represent the four-

1 T he g r a p h ic a l r e p r e s e n ta tio n o f th e e le m e n ts o f th e s c a tte r in g m a tr ix w as

in tr o d u c e d by R . F e y n m a n (6 2 ).
308 Q U A N T U M E L E C T R O D Y N A M IC S

vectors, over which the integration in S(n) is carried out, by points in


a diagram, and we shall represent the operators by lines passing through
these points —the vertices of the diagram.
The representation of operators is carried out in the following manner.
We shall call those operators free which remain unpaired in products
of the type (24.15) and (24.15'); and we shall agree to represent a free
operator A^(x) by an (undirected) dotted line leading from the vertex
x beyond the boundaries of the diagram (and to “ infinity”), to repre­
sent a free operator y(x) by a solid line directed from the vertex x towards
“infinity”, and, finally, to represent a free operator ip(x) by a solid line
directed from “infinity” towards the point x.
Since y>(x) represents an electron annihilation or a positron creation
operator, the solid line leading from “infinity” to the vertex x can be
used to represent graphically either an electron which existed prior
to the scattering process, or a positron created during the scattering
process.
Similarly, since ^i(x) is a positron annihilation or an electron creation
operator, the solid line leading from the vertex x towards “infinity”
can be used to represent either an electron formed as a result of scatter­
ing, or a positron which existed prior to scattering.
Finally, since A^(x) is a photon annihilation or creation operator,
a dotted line joining the vertex x to “infinity” can be used to represent
a photon created or annihilated as a result of scattering.
We shall also use a dotted line to represent an external electromag­
netic field acting at the point x.
Pairings between operators are represented by internal lines of the
diagram connecting its vertices, viz., a pairing between the photon
operators A^(x) and A ^ j) is represented by an (undirected) dotted
line connecting the vertices x and y, while a pairing between the electron
operators y(x) and yj(y) is represented by a solid line joining the vertices
x and y and directed from y towards x.
Since pairings represent products of creation and annihilation opera­
tors averaged in a certain way, the internal lines of a diagram connecting
its vertices can be made the counterparts of virtual “ particles” created
and annihilated in the scattering process.
These “particles” are characterized, as we shall see later, by the
fact that for them there exists no definite relation between the time
F U N D A M E N T A L E Q U A T IO N S 309

and the spatial components of the momentum: the momentum of a


virtual electron, in contrast to that of a real electron, does not satisfy
the condition p2-\-m2 = 0, while the momentum of a virtual photon,
in contrast to that of a real photon, does not satisfy the relation k 2 = 0.
An internal electron line is directed from the point of creation of
a virtual electron towards the point of its annihilation (or from the
point of annihilation of a virtual positron towards the point of its crea­
tion).
Since in the expression for S(7i) each point x appears as the argument
of three operators—one photon and two electron operators— it fol­
lows that one photon and two electron lines pass through eaeh vertex
of the diagram.
D iagram s representing individual norm al products appearing in
the m atrix S(n) contain n vertices. We shall call them diagram s o f the
7 7 th order. They can be used to represent the effects o f the 7 7 th order
approxim ation o f perturbation theory.
We shall give several examples.

25.2. Various Interaction Processes between Fields


We begin with first order effects. In this case, obviously, there exists
only the one diagram shown in Fig. 12. It represents the scattering
of an electron or a positron by an external field, the creation or annihi­

F ig . 12.

lation of a photon by an electron (positron), and also the creation or


annihilation of an electron-positron pair. (On the right hand side of
the diagram we have symbolically shown the integrand of the matrix
element of S(1) omitting the symbol for the normal product.)
We now consider second order effects. In this case it is possible
to have only the six topologically different diagrams shown in Fig. 13.
310 QUANTUM ELECTRODYNAMICS

Beside the diagrams we have symbolically indicated the integrand of


the matrix element of S(2) (omitting the normal product symbol). The
lines connecting the different factors xp, xp, and A indicate the pairings be­
tween the operators.
Fig. 14 shows the 15 diagrams giving all possible third order effects.
Beside each diagram we have symbolically indicated the integrand
of the matrix element of S<3) and the pairings between the operators.

( x jjA x l) ) ( x p A x p ) ( i ) A x l) ) ( x j) A x p )
(4) (5)
o -

x p )(x p A ty ) m m * )

(6)

(ip A ip )(T p A ip )
v ' — r' -

Fig. 13.

We examine in greater detail diagram 8 of Fig. 14. If the external


lines in this diagram represent an electron and an external electromag­
netic field, then this diagram can be used to represent the following
process: an electron emits a virtual photon (vertex .vj, then the electron
undergoes scattering by the external field A e(xa) (the vertex *3), and,
finally, the electron absorbs the previously emitted virtual photon
(vertex Xj).
FUNDAMENTAL EQUATIONS 311

(4>A \p)(TpAip)(ipAip)
( 1 ) ( 9)
0

( 2) ( 10)
(fA \p ){T p A p )(T p A \p ) (V<A ip)(lpAtp){lpAlp)

( 3) (xj)k j>)(ipArp)(TpAip) A
( 11 )
(*) | (ipAip)(ipATp)(TpAip)

(12) .

( 5)

W (13)

( 6)
(ipA lp)(TpATp)(lpATp)

(14)
( 7) \N)A rp)(rpAip)(i(ATp)
\i>A}P)(^A\p)(-ipAxP)

) J
(ip A lp)(-pAip)(^Ai/>) A 0>A ip)(lpA i p j ^ A l p )
i i

Fig. 14.
312 QUANTUM ELECTRODYNAMICS

In the expansion of S(3) into normal products this process corresponds


to the six terms

— j r f d*x i f d*x *J d4x*S N (va(xi')^b(x i)v(x i)


X y (x,) A b(Xj) ip\x}) y d(xk) A (xk) ipa(xk)),
where the subscripts i,j, k, assume the values 1, 2, 3, with i # 7 , i 7 ^ k,
k 7 ^ j , and the summation is taken over all the permutations of these
subscripts. On omitting the integral sign, and on suppressing the argu­
ments of the operators, we can rewrite these terms in the form

A A A ___ A A A

(1) N(ipAyi)(ipAip)(ipAip), (4) N(ipAip) (ipAyj) {ipA y ) ,


___ a T "a a ___ A A A

(2) NjipAip^ {ipAip) (yAip), (5) N(ip£ip^ipAip)(ipAip),


A A '- A ___ A A A

(3) N(ipAip) (ipAip) (ipAip) , (6) N (ipAip) (ifAip) (ipAy),

and, in order to be explicit we have indicated both in these expressions


and in the diagram the pairings between electron operators by solid
lines with arrows directed from ip toward ip, and the pairings between
photon operators by dotted lines.
On noting that we can interchange the electron operators in an
TV-product, provided we multiply the N-product by the parity of the
permutation performed, and on utilizing the formulas which define
pairings between operators, we can easily show that all these six terms
are equal to one another.
We can say that the different normal products corresponding to
a definite nth order scattering process, which are analogous to those
just discussed, and which differ among themselves only by rearrange­
ments of subscripts specifying the points, are equivalent, since they
are equal to one another. The number of such equivalent normal prod­
ucts is evidently equal to z = n!/gw heregis the number of permutations
of subscripts of points which leave the form of the iV-product unaltered.
If the free operators describe different particle states, then g = 1
and z = n!. In this most important case we can in evaluating the
matrix element consider only one of the equivalent /^-products, and
disregard in S(7l) the factorial term 1jn\ which cancels against the number
of equivalent N-products.
F U N D A M E N T A L E Q U A T IO N S 313

The diagrams shown in Figs. 13 and 14 represent normal products


of field operators in general form, and describe a number of processes
simultaneously. These operators are sums of creation and annihilation
operators for particles in different states; therefore the normal product
corresponding to some specific physical process can also be represented in
the form of a sum of several terms which contain products of creation
and annihilation operators for particles taking part in the process
under consideration, and which differ from one another only by the
order of the arrangement of the operators. These terms can be repre­
sented by diagrams which are topologically equivalent, and which differ
from one another only by the order of the arrangement of the elec­
tron and photon lines.
For example, we consider the emission of a photon by an electron
in an external field. This is a second order process, and the normal
product corresponding to it in the integrand of S(2) is

JV(v(*0 A (* 0 v t3 v (* 2 ) A(*2)y(*2)) •
A

In place of A(x) we must substitute here

A (*) = Ae(x)+ A fc(x),


where Ae is the external field, and A k is the operator for the emission
of the photon k. The normal product can be decomposed into two
terms:

N[Vp{xl)Ae(xl) yAxSyj(x2) A k (x2) y (x2))

+N(y>(x i) Afc(x0 k 3 V (* 2) ^e(*a) y( * 2) ) ,

which correspond to the two diagrams shown in Fig. 15, and which
differ from one another only by the order of arrangement of the photon
lines representing A (e> and Afc.
If n photons take part in the process, then after the normal product
is decomposed into terms containing annihilation and creation oper­
ators for individual photons, we obtain n\ terms which correspond
to diagrams which differ from one another only by the order of
arrangement of the photon lines.
314 QUANTUM ELECTRODYNAMICS

Similarly, if several electrons and positrons take part in the process,


then the normal product can be represented in the form of a sum of
terms which contain the same electron and positron creation and anni­
hilation operators, and which differ only by the order of arrangement
of these operators, while the diagrams corresponding to these terms
differ from one another only by the order of arrangement of the electron
lines.

For example, the process of electron-electron scattering corresponds


to the two diagrams shown in Fig. 16; in these diagrams P i , p 2 and
p[, p2 denote the electron momenta before and after scattering, while k
and k' are the momenta of the virtual photons which are exchanged
by the two electrons.
In contrast to processes in which several photons take part and
for which the individual diagrams correspond to matrix elements having
the same sign, in the case of processes in which several electrons take
part, the individual diagrams correspond to matrix elements which
can have opposite signs. This is associated with the fact that the creation
and annihilation operators for electrons and positrons in different states
do not commute, in contrast to the commuting photon creation and
annihilation operators.
For example, the two diagrams of Fig. 16 correspond to matrix
elements of opposite signs. Indeed, the integrand of that part of the
S(2) matrix which describes electron-electron scattering is equal to

JV(v(*i )y}(x2) A ( x 2)rp(x2)].


F U N D A M E N T A L E Q U A T IO N S 315

On substituting into this expression in place of xp and xp,

xp(x)=a\xpp[(x)+a +>fpp>(x),
V O) = aPi xpPi (x)+ aPi xp^ (x) ,

where a+ and ap are the creation and annihilation operators for an


electron of momentum p, and xpv{x) is the wave function of an electron
of momentum p, we obtain four nonvanishing terms

The first two terms are two equivalent normal products which are
represented by diagram (a) of Fig. 16, while the second two terms
are two equivalent normal products which correspond to Fig. 16(b).
On interchanging in the third term the operators a and a+ it can
be easily shown that
a +/a
Pi Pi a p.
+.a Pi

and, therefore, as was stated previously, the matrix elements correspond­


ing to the two diagrams have opposite signs.
This result can be generalized in the following manner. Suppose
that n electrons take part in the process and that their momenta before
and after scattering are respectively equal to P i, p 2, ■■■ and p[,p 2, ...;
then the process will correspond to n\ diagrams which differ by the
designation of the electron lines after scattering. If in two such diagrams
the electron lines after scattering (which arc the continuation of the
electron lines before scattering) are denoted respectively by p[,p'2, ...
and pl^p'i,, •••, then the relative sign of the matrix elements corre­
sponding to the two diagrams will be determined by the parity of the
permutation
1 .2 ........n
316 QUANTUM ELECTRODYNAMICS

Thus, the matrix element corresponding to some physical process


can always be represented after the decomposition of the normal prod­
uct in the form
(25.1)

where the individual terms M}^lf differ from one another by the order
of the arrangement of the annihilation and creation operators for the
particles participating in the process; the diagrams representing these
terms are topologically equivalent, and differ only in the order of the
arrangement of the electron and photon lines.

25.3. Transition to Momentum Space


For the evaluation of the probabilities of different processes with
the aid of the scattering matrix, and also for a general investigation
of its singularities, it is convenient to go over to momentum space.
We now determine the form which the quantities have in mo­
mentum space. In order to do this, we express the pairings between
operators, and also the external potentials A iMe)(x) in the form of Fourier
integrals, and we substitute them, together with the expressions for the
matrix elements of the particle creation and annihilation operators,
into the normal product (in the expansion of the integrand of S(n))
which corresponds to
The pairings between the operators are given in accordance with
(17.21) and (19.18) by

(25.2)

\ ‘V __, n JaP
p 2-\-m2— i0

where dt{k2) is an arbitrary function of k2.


FUNDAMENTAL EQUATIONS 317

We write the Fourier expansion of the external potential A<(e)(;c)


treated as a onum ber in the form

Al e)(x) = A itle)(q)eiQXdiq ,A ^ ( q ) = J A ^ ( x ) e ~ ^ d * x . (25.3)

Finally, the matrix elements of the field operators are defined by


the following formulas. The matrix elements of the operators ip(x) and
ip(x) corresponding to the annihilation and creation of an electron
or a positron of momentum p and polarization r are respectively given
by (of. (24.9), (24.10))

(0/rlvW I !pr ) = VpP(x) = ur(p)eipx,

(lpr IV(*) I°pr ) = Wpt](*) = UT(p ) e~iPX>


I - , _ (25‘4)
(OprlvWI !pr) = Vpr’M = Vr( - p ) e ipx,

(^rlvW lO pr) = VpT'ix) = vr(—p')e-ipx,

where uTa(j>) and v ra(—p) are spinors satisfying the following normali­
zation conditions uTuT* = vTv r* = 1.
The matrix elements of the operator A (x) corresponding to the
annihilation and the creation of a photon of momentum k and
polarization X are respectively given by

(0fcJ A ^ ) | l w) = -— e ^ e 'teJ
(25.5)
( I\ ( x ) 10M) = e~ikx

(in all these formulas for the sake of simplicity we take the normali­
zation volume equal to unity).
On substituting these expressions into the normal product corre­
sponding to M /^ O n the integrand of the expression for S(n)) we first
of all carry out the integration over x lf x 2, ..., xn. On collecting the
factors exp (ipx,) with a definite x, (p denotes here the four-momenta
of the free particles, and also the variables of integration in the Fourier
expansions of the functions Ajf^x), Sc(x) and Dc(x)), we obtain
318 QUANTUM ELECTRODYNAMICS

exp ( i x ^ p ) , where the number of vectors in the sumJT/? is obviously


equal to three (the number of lines passing through the vertex x, of the
diagram). The integral of exp( i x ^ p ) is equal to (2tt4) d(£p), where
d(p) is a four-dimensional ^-function. Therefore, oh carrying out the
integration over x2, xn, we obtain the product of n four-dimen­
sional ^-functions.
In momentum space, to each line of the diagram there will correspond
a certain four-vector p. Since the functions Dc and S c associated with
the internal lines of the diagram depend on the difference in the coordi­
nates of the ends of these lines, then in the two (5-functions referring to
some internal line the vector p associated with this line will appear
with opposite signs. This enables us to interpret the vectors p corre­
sponding to the internal lines of the diagram as four-momenta of the
virtual “particles,” “created” at one end and “annihilated” at the
other end of the internal lines. In this case, obviously, there exists no
relation between the time and the spatial components of the four-
momentum of the virtual particle, as has been already mentioned.
Insofar as the external lines of the diagram are concerned, i.e.,
lines going outside the boundaries of the diagram, they obviously
correspond to four-momenta of the real particles taking part in the
process.
In order to obtain the final expression for M ,^ it is necessary to
arrange the matrices operating on the spinor indices in a definite order,
carry out the integration over the momenta of the virtual particles and
the variable q (arising from the expansion of the external potentials
into Fourier integrals), and sum the expression so obtained over the
polarizations of the virtual photons.
We first of all determine the order in which the spinor matrices must
be arranged. In order to do this we consider an example—the scattering
of an electron by an external field, shown in Fig. 17. The normal product
corresponding to this process (in the integrand of the expression for
S(3)) is given by
F U N D A M E N T A L E Q U A T IO N S 319

On interchanging y x and y^, and then y A and y a>. which will leave the
sign of F unaltered, and on utilizing (17.10) and (19.16), we rewrite
F in the form

F = N{yA(x2) ( y X , S cee(x2- x 3) A</>(x3) (yd)g„Scaa(x3- x r) { y ^ p Wp(x,))


X £ ^ (* i—* 2) = N(y(x2) y vS c(x2—x 3) A ie)(x3)
X S c(x3—x 1) y /ty (x 1))Dl,(x1- x i)
(the interchange of y a and yp leads to additional multiplication by —1).
This expression shows that the internal photon line corresponds to
the function D and to the two matrices yv and y^ which must be taken
to correspond to its ends (the vertices x 2 and Xj). In this case the ma­

trices operating on the spinor indices, i.e., the matrices y and S c, must
be arranged in the same order read from left to right as that in which
they occur if we move in the direction opposite to the electron line
(cf. Fig. 17).
Having determined the order in which the spinor matrices must
be arranged we can now write the following general expression for
in the form of an integral in momentum space:

dP( — l) n+t en(27i)iin- F)j d 4p 1dip 2 ... dip Fe J d^kxdik 2 ...

••• d*k r J <


P(h d*th ••• d*Qs

X iij 4 ip ) o { n ( u(Pf)v(-Pf)

i~r A {e)(q) j j ip—m X 1 /


X f ] Y* (25.6)
1 1 (2t i) 4 1 J p 2+ m 2 k 2 ) /fi
i '
320 QUANTUM ELECTRODYNAMICS

Here the integration is taken over the four-momenta of the virtual


particles, i.e., over the 4Fe variables Pi, p 2, ■■•■>PFe arising from factors
of the type S c, and over the 4Fp variables k ±, k2, . . . , k Fp arising from
factors of the type Dc, and also over the 4s variables qL, q 2, . . . , q s
arising from the expansion of the external potentials into Fourier
integrals (Fe and Fp are the numbers of the internal electron and photon
lines, i.e., the numbers of virtual electrons and photons, F = Fe+ F p, s is
the number of vertices at which external potentials exist, / is the number
of closed electron loops with an even number of electron lines, 8p is
the parity of the permutation of the electron operators, cf. below).
In addition to integration a summation is carried out in (25.6)
over the four values of the subscripts v which denote the different po­
larizations of the virtual photons, with each virtual photon correspond­
ing to its own subscript v, which assumes the values v — 1, 2, 3, 4.
The individual factors in (25.6) have the following meaning. The prod­
uct Y l ( u(Pl) v{ — p ^ [ei/]/(2coi]) is a product of the spinors u ( p ^ , v ( ~ p ^
and the matrices ei/j/2co,. describing electrons, positrons and photons
in the initial state i ( p { is the three-dimensional momentum of an
electron or a positron, is the unit polarization vector of a photon
of momentum k t, e = e^y^, 1); [](^(P/)V(~Pf)[^fl]/2oJ/]) denotes
a similar product for the final state /.
The factors i(ip—m/p2+ m 2) and yv(\/ik2) ( 8 ^+ id — l)[kfik j k 2]) y^
arise from the pairings of electron and photon operators with the
number of the former being equal to Fe, and of the latter being equal
to Fp.
Finally, the symbol O denotes a definite order in the arrangement
of the spinor matrices in (25.6), viz., the matrices y and S c operating
on the spinor indices must be arranged reading from right to left in
the same sequence in which they occur if we move in the direction of
an electron line of the diagram.
A special feature is present in the case when the diagram contains
closed electron loops with an even number of electron lines (in the
case of an odd number of electron lines the matrix element vanishes
in accordance with Furry’s theorem, cf. below). In this case to each
electron loop there corresponds in the expression M ^ f the negative
trace of the product of the matrices y and S c associated with this loop.
FUNDAMENTAL EQUATIONS 321

In order to show this we consider the part of the diagram containing


the closed electron loop shown in Fig. 18 (the squares A and B denote
certain complex diagrams whose explicit form is not needed for this
discussion). The factor associated with this loop in the general expression
for the normal product corresponding to Fig. 18 has the form

G = JvfyEfo) {y X ^ % xi)W.x2) ( y X $v ae(x2)).


On utilizing expressions (19.16) for pairings of operators we obtain
G = -(yJ a P Sf a ( X l - x J ( y X X e a i ^ - X l )
= x2)y fScf e - ^ ) | ,
which is what was asserted earlier.

A _____ _*ir B
Jjv

Fig. 18.

If the diagram contains / closed electron loops with an even number


of lines, then M \ ^ f acquires the factor (—1)' which appears in (25.6)
together with the factor (—1)ndp, where dp determines the relative
sign of M ^ f in the case when several electrons participate in the process.
As has been explained earlier, <5p represents the parity of the permu­
tation of the electron indices

where the numbers 1 ,2 , ... enumerate in the diagram the momenta


of the electrons before scattering, while the numbers j \ , j 2 do the same
after scattering.
As an example of the application of the general formula (25.6)
we give the expression for the matrix element which corresponds to
the diagram of Fig. 17:

^ (e)(?) i(Pi -k)—m (MX


M™f = p e * u A yv ---
(P k)2~\-m2 (2rr)4 (y»1—A:)2+ w ;
(25.7)
322 QUANTUM ELECTRODYNAMICS

Here ux and u2 are the spinor amplitudes of electrons of momenta


p x and /?2 before and after scattering, and q = p 2—p x. We have elimi­
nated the three four-dimensional c)-functions appearing in the general
expression for by integrating over q and over the momenta of the
virtual electrons, having replaced q by p 2—p x and the momenta of
the internal electron lines by p x—k and p2—k. For this reason the
integration in (25.7) is carried out only over the four-momentum of
the virtual photon k; moreover in (25.7) the summation is carried
out over the four values of the indices v, and this corresponds to summing
over the four polarization states of the virtual photon. The arbitrary
quantity d f k 2) has been taken equal to d , = 1.

25.4. Closed Electron Loops with an Odd Number o f Vertices


We now show that diagrams containing closed internal electron
loops consisting of an odd number of lines correspond to a total matrix
element which vanishes {Furry's theorem).
If a diagram Gx contains an internal closed electron loop I (Fig. 19),
then in addition to this diagram we must, obviously, also consider

F ig . 19.

another diagram G2 which differs from the former only by the direction
of circulation around the closed electron loop (loop II). The sum of
the matrix elements corresponding to these diagrams, M Ci and MCj,
determines the total matrix element for the process

M - M Gi-\-MGi.

In accordance with (25.6) the loops I and II correspond to the


following parts of the matrix elements M Gi and M Ct:
FUNDAMENTAL EQUATIONS 323

M, = j d ' p S p i y ^ S y p k k J y ^ S H p + k , l k,J ... S H p - k . p y , ^ , ) .

M „ = j d ‘p ' S p ( S ‘(p')

X r ^ V + k J ... S ^ p ' - k t - k 1)yIHS ' ( p ' - k 1)yrj ,


which appear in A/0i and MCa in the form of factors (TV is the number
of vertices in the loop).
On noting that
Sp A = Sp C“1AC,
where A and C are arbitrary matrices, we write M n in the form
M u = f d y Sp(C-hS'QOC
x C - ^ ^ C C - ^ ^ p ' + ^ C ... C_1 s c{p'—k1) C C-’y^C ).
We now choose for C the charge conjugation matrix. Then, in accord­
ance with (10.2) we have
C~lyf !i C = — yrjii,
and, consequently,
C-1S‘( P )C = S ' ( - p ) .
On utilizing these relations we obtain

M u = ( - \ r f d Y S p ( S ^ - p ' ) y , NS ^ - p ' - k N) ... S ' i - p ' + k j y ^ .


Further, on introducing the change of variable p ' = —p, and on noting
that for an arbitrary matrix A the relation Sp A = Sp A holds, we
obtain finally
r f d ‘p Sp(y ^ p + k j
X r ^ S H p + k i + k J ... S'(P k,,)ytltS‘(p)].
Since the total matrix element M contains as a factor the sum of
the expressions M, and M n , it follows from the last formula that in
the case of odd TV the matrix element M vanishes.
Thus, diagrams containing closed loops consisting of an odd number
of electron lines may be left out of consideration entirely.
We note that Furry’s theorem is a consequence of the law of conser­
vation of charge parity (cf. subsection 23.7). Indeed, a closed electron
loop with TVvertices, i.e., with N photon lines approaching it, corre­
sponds to a neutral system of real and virtual photons whose total num-
324 QUANTUM ELECTRODYNAMICS

ber is equal to N. Since for an individual photon A — —1, an odd value


of N would mean that the initial and final states of the system would
have opposite charge parity and this is impossible.

25.5. Rules for Writing Down Matrix Elements


In concluding this section we state the rules by means of which
we may directly write down the matrix element corresponding to any
diagram. These rules were formulated by Feynman (62) and consist
of the following.
1. To each external electron line there corresponds a spinor of
one of the following types: u f p ) , u T(p),vr(—p ) , v T(—p), where ur(p)
and uT{p) correspond to the annihilation or creation of an electron
of momentum p and polarization r, while vT(—p) and vr( ~ p ) corre­
spond to the annihilation or creation of a positron of momentum p
and polarization r (the four-momentum of the positron is equal to
the negative of the four-momentum which defines the spinor v corre­
sponding to negative frequency).
2. To each external photon line representing a photon there corre­
sponds the matrix e / ] / 2 o j , where oj is the frequency and e is the unit
polarization vector (e = e^y^, e = e2—e\ = 1).
To each external photon line representing an external electromagnetic
/*v

field there corresponds the matrix >4(e)(g)/(27i)4.


3. To each internal electron line of momentum p there corresponds
the matrix —i(l/ip+rri).
4. To each internal photon line of momentum k there corresponds

the factor y ( ^ + ( ( /,- l) ^ and to its ends there correspond


the matrices y and y v.
5. To each vertex of the diagram there corresponds a (5-function
containing the momenta of all the lines converging at this vertex, with
the momenta at the two ends of an internal line being taken of opposite
sign.
6. All the matrices operating on the spinor indices are arranged
reading from right towards left in the sequence in which they occur
if we move in the direction of the electron line.
7. If the diagram contains a closed electron loop with an even
number of electron lines, then in there occurs as a factor the
F U N D A M E N T A L E Q U A T IO N S 325

negative trace of the product of the matrices —ijip + m and y


which are associated with the individual lines of the loop and with
its vertices.
The matrix elements of diagrams containing closed electron loops
with an odd number of electron lines vanish.
8. The numerical factor which appears in in front of the
product of the spinors u, v, u, v and the matrices y, (—i/ip+m) is equal
to (—l)n+'(5J)(27i)4(n- J:'), where F is the total number of internal lines,
/ is the number of electron loops with an even number of electron lines,
and dp is the character of the permutation of the indices of the electron
momenta dp = ^ . j (1 ,2 , ... enumerate the initial, and j 1, j 2, ...
enumerate the final electron momenta).
9. Integration is carried out over the four-momenta of the internal
lines representing virtual particles, and over the variables q associated
with the external potentials, and summation is carried out over the
four-dimensional polarizations of the virtual photons.
As an illustration of Feynman’s rules we shall give the matrix elements
corresponding to the scattering of a photon by an electron (the Comp­
ton effect), and to the conversion of an electron-positron pair into
two photons (two-photon pair annihilation).
As has been explained earlier, each of these matrix elements is made
up of two terms which differ in the order of operation of the photon
operators. The diagrams representing these terms are shown in Fig.
20 (adjacent to the electron lines are shown the four-momenta and the
spinor amplitudes for the electrons and the positrons, and adjacent
to the photon lines are shown the photon four-momenta).
We see that the two processes are represented by topologically iden­
tical diagrams, and the difference consists only of the fact that the elec­
tron line of momentum p 2 represents in the case of the Compton effect
the emission of an electron of four-momentum p e = p2, and in the
case of pair annihilation the annihilation of a positron of four-momentum
pp = —p 2 (the corresponding spinor amplitudes are designated in the
diagrams by u2(pe) and v(—pp)).
In accordance with Feynman’s rules and the general formula (25.6),
the matrix elements corresponding to the scattering of a photon by an
electron, and to pair annihilation, have the following form:
326 QUANTUM ELECTRODYNAMICS

For the case o f photon-electron scattering

i{Pi+K)-m
= ie2{2n)4u2
]/2co2 ( P i + k ^ f m 2 ^'20h
iCPl- K ) - m e2 \ i{p^ _ p i + k t _ kiy
+ ]/2o>r ( P i - k J 2+ m2 \/2
(25.8)
For the case of pair annihilation
S ! S, = +
i(j> i— k i ) — m e1
= i e 2( 2 n ) 4v ( 2
\/ 2o>2 ( P i - k xT + m 2 j/2^
ex i(Pi—k ^ —m e2
ud(p2—pj.+kj.+kz),
]/2co^ ( P i - k 2)2+™2 j/2 w2
(25.9)
where ex and e2 are the polarizations of the photons of momenta k x
and k2; the subscripts I and II denote the two diagrams corresponding

Fig. 20.

to each process. We have eliminated one of the two (5-functions appear­


ing in accordance with (25.6) in Mlf!r by integrating over the momentum
of the internal electron line after replacing it by p 1-\-k1 and p x—k 2 in
the Compton effect diagrams, and by p 1—k 1 and p x—k 2 in the diagrams
for pair annihilation.
F U N D A M E N T A L E Q U A T IO N S 327

§ 26. Probabilities of Various Processes

26.1 General Formula for the Probability


Having defined in accordance with the rules of the preceding section
the matrix elements corresponding to some process i -> f we can obtain
the probability of the process P — [ S ^ l 2.
If only free electrons and photons participate in this process, then
the matrix element obviously has the following form

s i^r = M i f d( H p — E p ,),

where Pt— ^ p f) is the four-dimensional (5-function containing the


difference between the initial and final four-momenta of the particles.
Therefore P = \M tf\2d2(]?Pt — ^ Pf ) - We replace one of the (5-functions
in this expression by the integral

m = w ? f emdlx-

We take the region of integration here to be finite. On assuming that


the spatial volume, i.e., the volume to which the particle wave functions
are normalized, is equal to unity, and on denoting the interval of time
variation by T, we obtain the following expression for P:

1
P
(2^7 M u \ ' * [ 2 , P i - 2 , P , ) T-

Thus, the probability of the transition / -> / is proportional to the time.


Expressed per unit time it is given by

w = ^ M ‘M E p > - E p f (26»
In the processes under consideration the.initial and the final states
belong to the continuous spectrum, and therefore it is necessary to
evaluate the probability that the three-dimensional particle momenta
p f in the final state lie within the ranges dpf . In order to obtain this
differential probability, which we denote by dw, we must multiply W
by the product JJ {(2n)~3dpf), where dpf /(2n)3 is the number of states
328 QUANTUM ELECTRODYNAMICS

of a particle of definite polarization whose momentum p f lies within


the interval dpf (the normalization volume is taken to be equal to unity).
Thus we have

* = ^ w « ( 2 a - 2 >,) n ^ - (26- v

On carrying out the integration over one of the momenta p f we


eliminate from dw three spatial (5-functions:

dw = ( 2 ^ |M " 12'5 ( S e‘S er) n ' 0 y - <26-2')

Here f ] ' denotes the product [] from which one of the factors dpf l{2n)z
! /
has been omitted (e, and ef are the energies of the particles in the initial
and final states).
In order to eliminate from (26.2') the (5-function involving the energy,
we represent any one of the factors appearing in f ] in the form
f
dp f I( I n f = Qedef do,
where do is the element of solid angle containing p f , and Qe is the density
of particle states per unit energy and per unit solid angle. On carrying
out the integration over def we eliminate the ^-function containing
the energy and obtain

( 2 6 ' 3 )

where

Qf = 5 Q A 2 ei - 2 Er)dEf
and [ ] " denotes the product [ J 1 in which one of the factors dpJQjjif
f f
has been omitted.

26.2. Effective Cross Section


On dividing the differential probability dw by the flux density of
the colliding particles I we obtain the directly observed quantity—the
differential effective cross section for the process
_ dw
do (26.4)
I
FUNDAMENTAL EQUATIONS 329

The current density I is most naturally defined in the laboratory


coordinate system in which one of the two particles is at rest before
the collision:
/<°) = «|0)w(0)
where nx is the density, vx is the velocity of the moving particle, and
the superscript zero refers to quantities in the laboratory coordinate
system, v^0) = 0. We note that this definition is inapplicable to the
case of photon-photon scattering, when there exists no laboratory
coordinate system. However, in this case we can repeat the following
argument by using the center of mass system from the outset.
In the case of an arbitrarily moving scatterer the definition of current
density is, generally speaking, not unique. However, we can sensibly
define it in such a way that da will be invariant.
In order to do this we make use of the fact that the number of scat­
tering events occurring in the volume Q during the time T is invariant.
In the laboratory system of coordinates this is given by d a ^ n ^ v ^ n ^
Q w T {0), while in the arbitrary system it is given by dan^n^QT, where
da is the cross section in this system, while I = nxv is the flux density
with v being the quantity that must be suitably defined. Since Q T is
invariant, we have
£/a’<0)«{0)«20M 0) = danYn2v . (26.5)
On the other hand, the quantities (n1v 1,in^) and (n2v 2, in2) are four-
vectors. Therefore
n1n2( l —v 1v 2) = n[0)ni20).
We see that if we define v by
V = « 4 0)(1 — V i V 2)

the cross section will be invariant


da = da{0).
We note that in the special case, when v k and v 2 are colinear, v[0)
= (v1—v 2) / ( l —v 1v 2), and it follows from (26.5) that
v= \ v 21.
In the case of a collision involving a photon (v1 = 1), the expression
for v assumes the form
V = 1 — V 2 COS ■&,

where -& is the angle between and v 2.


330 QUANTUM ELECTRODYNAMICS

On utilizing the expression for z^0) in terms of invariant quantities

wi 0) = - { ( p i P i Y - m l m f 12/(P1 P2 ) ,
where plt p2 are the four-momenta and m1, m2 are the masses of the
colliding particles, we can eliminate from the expression for v any
reference to the laboratory system
v = ((PiP?y—m\m§m /£i£2- (26.5')
This formula is applicable to any case, including the case of two colliding
photons.
Formulas (26.2), (26.4) and (26.5) yield the following expression
for the effective differential cross section:

da - IM,if I2<5(/>! + p 2— EP f) /y — t 2 - T W - f- (26.6)


{Inf 2
1n 1 (2n)3 ’
where the subscripts 1,2 refer to the particles in the initial state, and
the subscript / refers to final states.
Formula (26.6) can be rewritten in such a form that da is expressed
only in terms of invariant quantities. In order to do this we explicitly
pick out from the matrix element M if those factors which are not invar­
iant. The following factors are of this kind: to each photon in the
initial and final states there corresponds the normalizing factor (2e)“1'2,
where e is the photon energy. To each electron there also corresponds
a noninvariant factor associated with the noninvariance of the ampli­
tude normalization u * u = u y i u = \ . It corresponds to the normali­
zation uu = m/s. Therefore, we can pick out from M if the factor (m/s)11-
or e~1/2 corresponding to each electron state. Thus, we can write M if
in the form

k
where the product is taken over all particles both in the initial and in
the final states. The quantity A is called the invariant scattering ampli­
tude.
On substituting this expression into (26.6) and on utilizing the fact
that

d3p f
= 2 dip f d(p2f-\-mlf )0(Ef)
-f
F U N D A M E N T A L E Q U A T IO N S 331

1 for zf > 0
where and d(zf) is relativistically invariant,
0 for ef < 0 ’
we obtain

da

(26.6')
Formula (26.6') gives the invariant form of the scattering cross section.
We note that the invariant form of the statistical weights of particles
utilized in (26.6') may turn out to be practically useful for integration.
As an example, we consider the case of two particles in the final state
with momenta k x and k 2. By introducing the new variables
q = k l —k 2, p = P!+Pi, Q = k xf k 2,
we obtain

dik 1 d4k 2 8(klJr mf) 8{k\-\-m%) 8(Q—p)

— ^ d4q 8{q2-\-k2-\- 2m\+2m%) 8 (qk+m\—m\

where W is the total energy and k 1 is the momentum of the particle


in the center of mass system, while do is the element of solid angle in
that system.

26.3. Summation and Averaging over Polarization States of Electrons


and Photons
In those cases when we are not interested in a definite polarization
state of the product particles (i.e., the orientation of electron spins
and the direction of photon polarization), expression (26.2) must be
summed over all possible particle polarizations in the final state. If
in the initial state particles are not polarized, then (26.2) must also
be averaged over the polarizations of the particles in the initial state.
We start by carrying out the summation, and by averaging over
the different orientations of the electron spin.
For the sake of simplicity we discuss the case when we have only
one electron in the initial and the final states. The quantity M if appearing
in the matrix element S\fl{ has in this case the form
M if = uf Qut,
332 QUANTUM ELECTRODYNAMICS

where and uf are the spinor amplitudes for the electron in the initial
and the final states, and Q is a certain matrix. We are interested in the
quantity
~I!\Mlf\2 = — £ Z ( ^ 7 i Q ud O f S ^ w , ) ,
Uf Z Uf Hi

where ^ and denote summations over two orientations of the


H
electron spin in the final and initial states. In accordance with subsection
10.7 we have

y ^ lMv!2 = "sTp1 - s P { 2 ( ^ r m) 2 ( '^ - " « ) } - (26.7)


z ° £ie f
where
Q = YiQJr7i-
If there is only one positron in both the initial and the final states
then the following expression holds:
■ M if = ViQvf ,
where vf = v{—p {) and vt = z>(—p t) are the spinor amplitudes corre­
sponding to the final and the initial positron states. In accordance with
subsection 10.7 we have

y E \ViQvf \2= l - 2 2, (vt7 iQ vf)(vj Q +7A^d


H-vi ft ff

= g^-£- SP {QOPf+m) Q(iPi+m)}. (26.8)

If one of the states is an electron state while the other is a positron


state, then we have
(uf Q v , in the case of pair creation,
M lf = \ J f .
1viQui in the case of pair annihilation.
Moreover, we have

y H \ u f Qvf \2 = t J — Sp{Q0pp+ m )Q (/y -m )} , (26.9)


Mf 0t eEp

is
Z Mi
l*\6«fl2 = 0 Ee Ep
(26.9')
where p e and pp are the momenta of the electron and the positron, and
ee and ep are the corresponding energies.
FUNDAMENTAL EQUATIONS 333

We note that formulas (26.8), (26.9) and (26.9') can be obtained


from formula (26.7): the latter goes over into (26.8) as a result of the
substitution p t -> —p f , p{ -> —pt. As a result of the substitution
Pf -*■ Pe> Pi - * —Pp(£f - > £ e > £ i - + — £ p) it g°es over into (26.9), and as
a result of the substitution p t -> pe, p{ -> —/^(e, -> ee, ef -> —ep) it
goes over into (26.9').
In a similar manner the summation over the polarizations can
also be carried out in more complicated cases when several electrons
participate in the process. In such a case it is possible to utilize formulas
(26.7) , (26.8), (26.9), (26.9'), by assuming that the matrix Q itself contains
spinor amplitudes. Summation over the polarizations corresponding
to them can be carried out by the same method by which formulas
(26.7) —(26.9) were obtained.
We now proceed to the summation over the photon polarizations.
For the sake of simplicity we discuss the case when only one electron
and one photon participate in the process. In this case the matrices
Q and Q are
Q = eG, Q=&,

where G does not contain the photon polarization e . Since


e = ey-!-/£v/4, e+ = e y —ie0y i eZ~ eo =
we have
A
e = y4c+y4 = —e.
Therefore
Q = - Ge
and
j £ \Mif\2 = - ^ - S p { e G ( i p - m ) G e ( i p f - m ) } . (26.10)
H>t1! {
If we are not interested in a definite state of photon polarization,
expression (26.10) must be summed or averaged over the two independent
directions of the vector e .
We denote by k the space part of the photon propagation vector
and choose a coordinate system with the z-axis directed along k. Then
for the two independent polarization vectors we can take the vectors
334 Q U A N T U M E L E C T R O D Y N A M IC S

# > (1 ,0 , 0 ,0 ) and # > (0, 1 ,0 , 0) and write the sum in which we are
interested in the form

y £ ^ \ M iAZ=
e Ul.Vf 1 r ’=1
= ^ ^ P { y }G(iP - m ) G y j (ipf —m)}, (26.11)
i f >=i
2
where denotes summation over the two polarization states of the
]=i
photon.
We now show that the summation in this formula may be carried
out over four values of j, i.e.,

£ l ^ v l -* )& .& ,-"< )}• (26-12)

where v = 1 ,2 , 3, 4. In order to do this, we use the fact that the equa­


tions of quantum electrodynamics are gauge-invariant. Therefore, if in
the matrix element containing the four-dimensional photon polarization
e we replace e by k we obtain zero. From this it follows that if in ex­
pression (26.10), which is quadratic in the matrix element, we replace
one of the vectors e by k , then this expression will vanish, i.e.,

Sp {{y-i +iydG{ipi-m ) G e { i p f - m ) } = 0,
Sp {eG(ip—m)G(y.3+ i y i) (ipf —m)} = 0,

while for the remaining vector e we can take an arbitrary vector. If


we take for this vector first (0 ,0 , 1,0), and then ( 0 ,0 ,0 , 1), we obtain

SP {(yz + i y d G { ip - m ) G y . A(ipf - m ) } = 0,
Sp {yi G{iPi-m)G{y.i + i y ^ ( i p f - m ) } = 0.

By multiplying the second equation by — i and adding it to the first


one we obtain
s P{yaG(ipi—m)Gy3(ipf —m) + y i G(ipi—m)G yi (ipf —m)} = 0.
Further, on adding this equation multiplied by — lj&Eief to (26.11),
we obtain formula (26.12).
F U N D A M E N T A L E Q U A T IO N S 335

In accordance with the general results of the theory this result shows
that no “ longitudinal” or “scalar” photons can be emitted.
A similar relation holds if several photons participate in the process.
If, for example, the process involves one electron and two photons with
polarizations e4 and e2ft, the matrices Q and Q can be written
Q = e2Ae1+ e1Be2 and Q = e4Ae2-\-e2Be4.
By repeating the preceding arguments, it can be easily shown that on
summing £ \Mif\2 over the different polarization states of the two

photons we shall obtain the expression:

y £ IMif I2 = 8eV S Sp {(%r)^ l s) + £is)£%r)) (i p - m )


«!>'! W./'/ * f T-S=1
X (e[*'Ae™+ e p B e p ) (ipf - m )} (26.13)
= j — S p & y p A y r + y ' B y J i i p - m ) (yvAyM+y^Byv)(ipf - m ) } .
°ei£f
The summation over the four photon polarizations can be carried
out with the aid of the following formulas (cf. subsection 10.7):
y vsyv = 4j, y va yv = - 2 a, y vabyv = 4ab, yvabcyv = - 2 cba, (26.14)
where 5 is a scalar and a,, bv, cv are four-vectors, ab = avbv. As a result
of this summation the number of matrices following the trace symbol
is evidently reduced by 2g where g is the number of photons.
In evaluating traces of matrices it is convenient to utilize
A A

£Sp A t A2 = A t A2, (26.15)


iS p A l A 2A3Ai = (Ai A 2) (A3A4) + ( A 1A4) (A 2Ai) - ( A l A3) (A2A4),
where A lf A2, A3, A4 are arbitrary four-vectors.
The trace of a product of an odd number of factors At is obviously
equal to zero.
We also note the following useful formula:
JSp (^ i+ ^ i) (A2-\-a2) (A3-\-a3) (T4+ a 4) = (A 1A2-j-a1a2) (A3Ai -\-a3ai)
Jr(A 1A i Jra1ai) (A2A3-\-a2a3) —(A4A3 a4a3) (A2A4 o2o4), (26.16)
where ai are numbers not involving the matrices y/t.
336 Q U A N T U M E L E C T R O D Y N A M IC S

26.4. Probabilities o f Processes Involving Polarized Particles


A complete solution of the collision problem involves obtaining
the probability of the process as a function of the polarization state
of the particles in the initial state, and determining the polarization
states of the particles in the final state.
For the sake of simplicity we consider, as we have done earlier,
the case when only a single electron is present both in the initial and
in the final states, but the methods which we employ can be easily gener­
alized to the case of an arbitrary number of electrons and positrons.
In the case under discussion the quantity M lf appearing in the matrix
element S ^ f has the form
M if = uf Qur
The square of the absolute value of M if averaged over the initial states
is given by

\ M tff = ( P j Q u d ( M i Q u f) = (uf Q ) a u ia u ifi ( Q u f)ii

or by
(26.17)
where f>{i) is the polarization density matrix for the initial state, defined
by formula (10.21). The matrix rJ )W can be expressed in terms of the
electron polarization vector in the initial state in accordance with for­
mulas (10.31), (10.31').
On summing (26.17) over the polarizations of the final state we
obtain

= (26.18)
'7 f
If the initial state is unpolarized, then in accordance with (10.26)

9 U) = — — 0?A - w ) y 4

and formula (26.18) reduces to (26.7).


For the determination of the polarization state of the electron in
the final state we obtain the corresponding density matrix.
The general expression for the density matrix corresponding to
transitions from the initial state i has the form
F U N D A M E N T A L E Q U A T IO N S 337

The elements of the matrix gf r , corresponding to a given value of


the electron momentum p, = pr and to different polarizations fxf ,
(tf = ± 4 g’ve rise to the polarization density matrix for the electron
in the final state
S\*f S\%. _ MifM;r
where p f — pr (26.19)

normalized in accordance with

spe,n = 2 C /=1-
^/=±Vj
On substituting into (26.19) M i ^ U f Q u ^ and on averaging over
the initial polarizations, we obtain, in analogy with (26.17), the follow­
ing expression for the polarization matrix Qf:

Qlf/fir = u'1i{pf ) rJ>wy i Qulir ( Pf) j (26.20)


where
N = - — Sp{Q 9^y,Q (ipf-m )}.

This matrix is a two-dimensional polarization matrix corresponding


to the definition (10.18). In order to obtain the four-dimensional polar­
ization matrix we utilize expression (10.27)

n p = 2 t l U r t ' O ’W ' i p , ) -
M/t*r
On substituting into this formula expression (26.20) in place of q^ , ,
and on utilizing the summation formulas (10.28) we finally obtain
1 (ipf - m ) Q 9 (l)yi Q{ipf - m ) y i
9 lr) - -
(26.21)
2z, Sp {ipf - m ) Q rJ>{i)yi Q
If the initial state is unpolarized, then we have
9 U) = 1
(ip-m )y4
4£i
and
1 (i p , - m ) Q ( i p - m ) Q { i p f - m ) y i
rj)W (26.22)
8eief Sp {ipf —m ) Q { ip —m)Q
338 QUANTUM ELECTRODYNAMICS

By utilizing this formula we can, in accordance with (10.32), obtain


the electron polarization in the final state

We now investigate the dependence of the probabilities of the various


processes on the photon polarization. If in the initial state we have
a polarized photon, then the averaged value of |M j;|2 is expressed in
terms of the polarization matrix for the photon in the initial state.
On writing M u in the form
M
l r , i f a *u ae c aU)>
where e(i) is the polarization unit vector for the photon in the initial
state, we obtain
\JQ * = (26.23)

where g(i) is the photon density matrix defined by formula (2.13).


In order to obtain the polarization density matrix for the photon
in the final state g(f) we write M if in the form
M
1YIif - e{f)*b
ca ua*
Then we have

el!> = k f = kr

where v determines the photon polarization in the final state (the photon
momenta kf and kr are the same in the states / and / ') .
Since elan = 6va, it follows that
bX
W (26.24)
X\b,.\2
If we know gin, we can obtain the photon polarization parameters
with the aid of formula (2.15).
The vectors a and b obviously depend on the polarization states
of the other particles over which appropriate averages and summations
must be taken.
The knowledge of the density matrix for the particle in the final
state enables us to determine the probability of the process which
results in the particle being left in a given polarization state.
FUNDAMENTAL EQUATIONS 339

Indeed, if we know the wave function of the final state 0 f , then


the probability of a certain state <&d is given by W id) =
Similarly, if a given state selected by the detector is characterized by
the density matrix Q(d), while the state of the particle is characterized
by the density matrix g(f), then we have

W™ = S p e<V*>. (26.25)
Since the density matrix can be written (both for electrons and for pho­
tons) in the form

e (/) = K 1+ f W and e(d) = i O + £ W >


we have
»™ = £(l + f'f*). (26.26)
Thus, if the cross section for a process summed over the polarizations
of the particle in the final state is equal to a, then the cross section
corresponding to the production of the particle with given polarization
parameters is given by
e ™ = o ( 1+ W ) . (26.27)

26.5. Probabilities of Processes in the Presence of an External Field


We now show how the probabilities of various processes are to
be determined in those cases when an external electromagnetic field
is present.
If a single electron participates in the process, then the matrix element
can be given in the first approximation in the form

S {ll}f = —euf (J e~iqiAe{x) d*xj u{,

where q = pf ~Pi- We first assume that the external potential does


not depend on the time, Ae( x ) = A e(r). Then we have

S}±Jf = —euf Ae(q)ui7ji 6 (ef —e,),


where

Ae(q) = / Ae(r)e~iqr dr

and e, and sf are the electron energies in the initial and the final states.
The probability of the process is equal to |5/i^|2. By'proceeding in
340 QUANTUM ELECTRODYNAMICS

the same manner as in the case of the derivation of (26.1) we obtain


the following expression for the transition probability per unit time:
W = 2K\uf eAe(q)ui\2d(sf —si). (26.28)
Formula (26.28) determines the probability of electron scattering
by the field Ae^(r) in the first Born approximation.
In order to obtain the differential scattering cross section da we
must divide W by the initial electron velocity andmultiply
it by dpf l(2ji)3—the number of final states for which the electron mo­
mentum Pj after scattering lies within the interval dp{\

da = ^ ~ I *
On noting that
dpf = p}dpf do = Pf£f dsf do,
where do is the element of solid angle containing pf , and on eliminating
the (5-function by means of integration over der, we obtain

(2«-29)

In particular, if A (e) = 0, A^e) = iA0, then we have

^ = ^ \u r e A M u ,P ^ - (26.30)

This expression is formally equivalent to the corresponding expression


for the scattering cross section in nonrelativistic quantum mechanics
in the Bom approximation.
If we are not interested in definite electron polarization states before
and after scattering, then formula (26.30) should be summed over
the electron spin orientations after scattering, and averaged over the
electron spin orientations before scattering. This may be done with
the aid of relation (26.7).
In the case when A (e) = 0, the matrix Q is of the form
Q = y ^ A 0{q)
and formula (26.7) yields

y ^ I |2= - ^ - - S p { y i (ipi- m ) y i (ipf - m } \ e A 0(q)\2. (26.31)


'V'V *£i
F U N D A M E N T A L E Q U A T IO N S 341

The trace of the product of the matrices y^ can be evaluated in accordance


with formula (26.15):

= i S p { y 4(ip(y —^ - - / ^ ( i ^ y —ey4—m)}
= 4SP { ( - 'A Y - £- m) (V/Y— ■e~ m))
= iS p{(piy)(^/ y )+ £ 2+ m 2} = p2 cos&+m2+e2

= 2e2 11—v2sin2y | *

where # is the scattering angle, e = et = ef , \p\ = \p{\ = \p, | , v = vt.


On substituting this expression into (26.31) and (26.30), we obtain
the following formula for the averaged scattering cross section:

P2 1—a2sin2 do
da ei{Pi~Pf)reA0 (r)dr (26.32)
v2 J2nf

This expression differs from the corresponding expression of nonrela-


tivistic quantum mechanics by the additional factor 1—vz sin2(#/2).
We examine in greater detail the scattering of electrons by the
Coulomb field of the nucleus. In this case we have
eA0(r) = —(Ze2/r),
where Ze is the nuclear charge (in Gaussian units) and

eA0(q) = - f— e*rdr= - • (26.33)


J r q2
Indeed, we consider the integral

J = f —e ^ ' - i ' d r ,
n J r
where the factor exp(—rjr) {ij > 0) is introduced in order to guarantee
the convergence of the integral for large r. On carrying out the inte­
gration over the angle between the vectors r and q we obtain
oo
In 4n
J J (e~ii)J+ri)r e(ic-vi^df
iq q2-\-rj2 '
o
By letting r) tend to zero we obtain formula (26.33).
342 QUANTUM ELECTRODYNAMICS

The substitution of (26.33) into (26.32) leads to the following expres­


sion for the cross section for the scattering of electrons by the nuclear
Coulomb field:
Ze2 & do
da = 1—v2sin2 (26.34)
2pv

26.6. Feynman's Notation


Throughout this book we use a Euclidean metric and four-dimen­
sional space in which x4 = it. The scalar product of two four-vectors
a and b is defined by
4

a b = E aA = ab+ ai bi-
/*=»
If we use the pseudo-Euclidean metric with a real time coordinate
t = x 0 = —z'x4, then we have
ab = ab—a0bQ.
Usually in this case the scalar product is defined with the opposite
sign. We denote by (ab) the quantity
(ab) = a0b0—ab — —ab and (a2) — —a2 = a%—a2.
In using the pseudo-Euclidean metric it is convenient to introduce
(first of references (62)) in place of the matrices y (p>= 1,2, 3,4) other
matrices which we denote by yft(ji — 0, 1, 2, 3) and which are defined
in the following manner:
7o = y* = f t y f = iy} = where j = 1, 2, 3.
The matrix y£ is Hermitian, while the matrices yf are anti-Hermitian
7o+ = yl and y f+ = - y f .

The commutation properties of the matrices y f are defined by the follow­


ing relation:
0, fi=£v,
rlrl+ rlrl 1, [X = v = 0,
—1, fi = v = 1, 2, 3.
In analogy with the quantity a — ya — y a -fy 4a4 we define
a = (yFa) = y Fa0- y Fa.
F U N D A M E N T A L E Q U A T IO N S 343

It can easily be shown that


.A V
ia = —a.
Application of the commutation relations for the matrices yF leads
to the following relations:

(■yFyF) — 4, ab-\-ba = 2 (ab),

(yFayF) = - 2 a , I Sp ab = (ab),

(yFabyF) = 4 (ab), \ Sp abed = (ab) (cd)+ (ad) (cb)—(ac) (bd),

(yFdbcyF) = - l e b a , y F— y£yFyS = yF.


The properties of the matrices y F under charge conjugation are the
same as in the case of the matrices y :
Cy F
M= - y 'C .
In this notation the Dirac equation assumes the form
v v
(p—eA—m) xp = 0,

y>(p + eA+m) = 0,
where

it being understood that in the last equation the operation of differenti­


ation acts on quantities standing to the left of it.
For a free electron we have
ip = ueikx,
v v
(k—m) u = 0 and u(k—m) — 0.
The current density vector has the form

= W mV and h = eN tv* y F
M •
The bilinear expressions S, V, T, A, P may be defined by formulas
(9.19) with y ^ replaced by y F and with
r l = vs = i y f y a y f y o -
344 QUANTUM ELECTRODYNAMICS

The projection operator t](p) given by (10.22) can now be written


in the form
p+m
v (p ) = 2m ’

while formulas (26.7)-(26.9) for summing over the polarizations and


the density matrix (10.34) can be written in the form

^ \ UfQui\2 = 7 7 --S p { 2 (A + '« ) Q(Pf+™)},

l2 = SP { Q ( K ~ m ) Q (P e+ ™ )} ,

9 = — (1 + a y 5) ( p + m ) y 4,

where Q = y ^ y ^ If Q = y£y$ ... yf, then we have Q = y \ ...


In this notation the Green’s functions for the photon and for the
electron assume the form

Dc(k)
(& y
I . p+m
s c(p) =
p —m 1 (p %
) —m2 '

The rules formulated in subsection 25.5 for writing down the matrix
elements can be easily put into Feynman’s notation by using the iden­
tities
A .V .A v 1 1
e = le, i p = —p ,
F W )'
CHAPTER V

Interaction of Electrons with Photons

§ 27. Emission and Absorption of a Photon

27.1. General Expression for the Matrix Element


At this point let us initiate a systematic investigation of various
specific processes which arise from the interaction between electrons
and the electromagnetic field. We use perturbation theory, and in this
chapter we carry out calculations only in the first nonvanishing approx­
imation. Corrections associated with higher order approximations
will be discussed in Chapter VIII.
We begin by considering first order processes. The first order scat­
tering matrix is of the form

(27.1)

If Ajpc) is the potential of the external electromagnetic field then the


matrix elements of S(1) determine the scattering of an electron by this
field in the first Born approximation, which we have discussed in the
preceding section.
But if A^(x) is the operator of the quantized electromagnetic field
then the matrix elements of S(1) will determine photon emission and
absorption, since the operator A^(x) appears in S(1) linearly, and the
nonvanishing matrix elements of A^ correspond to photon emission
and absorption.
However, it can be easily shown that the matrix elements of S(1)
vanish if the electron taking part in the process is free. Indeed, if the
electron and the photon are both in states of definite momenta, then
the matrix element corresponding to photon absorption contains as
a factor the four-dimensional (5-function <5(pi+fc—p2), where pL(j>i, i^i),
P2 (j?2 y ^ 2) are the four-momenta of the electron in the initial and
final states, and k(k, ioS) is the four-momentum of the photon. Therefore
[345]
346 QUANTUM ELECTRODYNAMICS

the matrix element will differ from zero only if the laws of conservation
of energy and momentum are satisfied:
P i + k = p 2. (27.2)
But these laws cannot be satisfied simultaneously since on squaring
(27.2) and noting that p\ = p\ = —m2, k 2 = 0, we obtain
p xk = p l k — co | / p \-\-m 2 = 0,
which is impossible.
On replacing in (27.2) k by - k we obtain the conservation laws
for the emission of a photon by a free electron, which also cannot be
satisfied simultaneously.
On replacing k by £ k t we can easily show that no process is possible
in which a free electron absorbs or emits any arbitrary number of
photons.
Finally, we note that from the conservation laws it also follows
that a single photon cannot give rise to a free electron-positron pair,
and a pair cannot give rise to a single photon. This can be seen directly
in the coordinate system in which the center of mass of the pair is at
rest: in this system the total momentum of the pair is equal to zero,
but their energy is different from zero (it is not less than 2m), while
in the case of a photon the energy vanishes whenever the momentum
vanishes.
In order for the conservation laws to be satisfied the participation
of a third body is necessary. Absorption or emission of a photon can
occur only as a result of a “triple collision” in which the interaction
of the electron with the “third” body plays an essential role.
In a number of important cases this interaction can be described
with the aid of the concept of an external field appearing in the Hamil­
tonian for the electron. The operator of the electron-positron field
can be expanded in terms of the eigenfunctions of this Hamiltonian
(cf. (18.7)), and the concept of the electron state will in such a case
automatically take into account the interaction of the electron with
other bodies. We shall say that such electron states are not free.
Such an approach enables us to study the processes of emission and
absorption of a photon by means of the first order scattering matrix,
since its elements between electron states that are not free are, generally
speaking, different from zero.
INTERACTION OF ELECTRONS WITH PHOTONS 347

We now proceed to determine the probability of emission and ab­


sorption of a photon by an electron in an external field.
We write down the general expression for the matrix element
which determines the emission of a photon.
If y x{x) and y 2(x) are the wave functions of the initial and the final
electron states and A ^ x ) is the potential of the electromagnetic field
corresponding to a definite state of the photon, then we have

s i ¥ f = ~ e f V>2(x) A*(x)xp1(x)d*x \/N k + 1 (27.3)


where N k is the number of photons of the kind in which we are interested
before emission has taken place. In the future we shall discuss only the
case N k = 0.
We assume that y ^ x ) and y z{x) describe stationary states of an
electron of energy e1 and e2. Then the matrix element corresponding
to the emission of a photon of frequency io can be written in the form
S $ f = - 2 7tiU^f d(£l- e 2-a>),
r- - (27.4)
J
U ^ f = - i e y 2(r)A*(r)y1(r)dr,

where y x(r),\p2(r), A(r) are functions only of r (without time dependent


factors).
In particular, if we are investigating the emission of a photon of
definite momentum k and polarization e, then the potential is given by
formula (6.15) and the quantity has the form

U ^ f = ----- * = f yZ(r)aee-lkry v(r)dr. (27.5)


y2co J

27.2. Electric Multipole Radiation


We now discuss the radiation in the case when the initial and the
final states of the electron are bound, and the wavelength X of the
emitted photon is large compared to the dimensions a of the region
to which the electron motion is confined. As we shall see later, the prob­
ability of emission in this case is related in a simple manner to the
electric or the magnetic multipole moment of the electron and therefore
the radiation from such a system is called multipole radiation.
We determine the probability of emission of a photon of definite
angular momentum L , definite component of angular moment M,
348 QUANTUM ELECTRODYNAMICS

and definite parity P = (—l)i +A+1. We recall that A = 1 corresponds


to states of electric type, while A = 0 corresponds to states of magnetic
type. Therefore, transitions involving the emission of such a photon
are called electric 2L-pole or EL-transitions when A = 1, and magnetic
2L-pole or ML-transitions when A = 0.
We begin by discussing the emission of a photon of electric type.
The potential corresponding to a photon in a state of electric type
is given by (6.17). The product of the components of this potential
and the matrices y is equal to (omitting the time dependent factor)

= 3- l / — +Cal-J>)+iCYl0 L„), (27.6)


where
L+ 1
gL+I O'") YL. L+1. gL-i(Mr)YL'L-i,M,
^ = V ^ i 1 / 2L+1

«(-i)
LM — - 1 / 32 L++E
U V 1 gL+l(°)r) YL V 2L+1 IK )

®lm = gL(^r)YLM, (27.6')


R is the radius of the normalizing sphere and C is an arbitrary constant.
A

This expression must be substituted in place of A ( r ) in formula (27.4).


Since the size of the system is assumed to be small compared to
the wavelength, the principal contribution to the integral U ^ f comes
from small r for which cor < 1. Therefore, we can restrict ourselves
to the first term in the expansion of the functions gL(cor) in powers
of cor:
(/cor)L
g l (a r)K i4 x (2 L + 1 )!, ■ (27.6")
where (2L+1)!! = 1-3-5 ... (2L+1).
Since the function gL(a>r) is proportional to (cor)L, we can keep
in (27.6') only those terms which contain g£(cor) with the lowest value
of L, i.e., gi _1(cor). As a result we obtain

aLM = j / 4m*, a>r < 1. (27.7)


Further, on taking into account that and &LM are related in
consequence of (4.10) by the expression

ULM — - - CO
V 0 LM’ (27.8)
INTERACTION OF ELECTRONS WITH PHOTONS 349

we can write Ut_^f in the form

C W tu- (27.9)
CO .)]
We now show that

u <~<= ~ eV ^ - Q E ^ § v n r ) r L Y i “ v ' (r)dr-


(27.9')
In order to do this we note that

/ v t M ( - "*v 0 ?„) Vi (<•) d r = f V>?(r) (<xp0*„ - 0 f „ a p) r , (r) d r

where p is the momentum operator and H is the Hamiltonian for an A

electron in the external field A {e)(x), H = ap-rfim—ief3A{e). Since


H is a self-conjugate operator and ipx(r) and ip2(r) are its eigenfunctions:
Hy>! = and H ^2 = e2 ^ 2 >
we have
/ V* ( - 0 *M) ip1 d r = - e o f xp$®*My>l d r, (27.10')

where co = el —e2. In virtue of this relation the terms containing the


arbitrary constant C drop out from expression (27.9), as might have
been expected from considerations of gauge invariance, and U ^ f
assumes the form

u '~ < - - -h i V i I ’* ^ d r■

On substituting into this equation expression (27.6') for &*M and expan­
sion (27.6''), we obtain formula (27.9').
We note that the condition %> a is equivalent to the condition
v ■< 1 where v is the ratio of the electron velocity to the velocity of
light. (This follows from the fact that in order of magnitude we have
v~^coa < 1.) Therefore, for the functions ^i(c) and -tp2(r) in formula
(27.9') we can utilize the nonrelativistic Pauli wave functions cp(cf. § 15).
350 Q U A N T U M E L E C T R O D Y N A M IC S

The matrix element U ^ f can be expressed in terms of the operator


for the electric multipole moment of the electron defined in the following
manner:
Q( i ) = l\ / — rLY,LM> (27.10)
2 1+ 1

where e/]/4n is the electron charge in Gaussian units. Indeed, on


introducing the matrix element of the electric multipole moment

( Q i u ) 2i = f < P * ( r ) QilMOO d r >


we can rewrite formula (27.9') in the form

£/‘- / = - j / ^ - ( ~ i ) L " j / l ( 2L+1) ~{2L-\)\\ ' ('27’1^

In accordance with the definition of the multipole moment operator


(27.10) we obtain in the case M = 0

o (1) — - f = - r LPL(costf),
V l o — (27.12)
)/ An
where PL{cosd) is a Legendre polynomial (the polar axis z coincides
with the axis of quantization along which the component of the angular
momentum M has been taken equal to zero). For L = 1 and L = 2
we obtain from the foregoing
e 1
Q10a) — " 7 ^ - z j Oa
V 2 0 >—
— (3z * - r 2),
]/ An 2
where z = r cos $. These formulas coincide with the usual definitions
of the dipole and quadrupole moments.
We now obtain the probability of photon emission. By utilizing
expression (27.4) for the matrix element and by proceeding in
the same manner as in deriving (26.3'), we obtain the following formula
for the transition probability per unit time:
W = 2w|t/l^/ |2«5(e1- e 2-o )).
The quantity W must be multiplied by the number of photon states
whose energies lie within the range dco, and it must then be integrated
over co. In accordance with (1.21) the number of photon states of a
given value of angular momentum L and of energy lying within the
INTERACTION OF ELECTRONS WITH PHOTONS 351

range da> is given by Qf = (Rda>/n). Therefore, after integrating the


quantity Wqf over co we obtain

• (27.13)
n
On substituting into this equation expression (27.11) in place of
we obtain the emission probability per unit time for a photon
of angular momentum L , component of angular momentum M, and
parity (—I)1".

WLM = £ (2 L + 1) f(2L—1)! !]2a)2l'+11(21m)2iI2 • (27.14)

In particular, for L = 1 we obtain the formula for the probability of


dipole radiation
0 = i " 3 l((2 ^ )2l|2. (27.15)

27.3. Magnetic Multipole Radiation


We now determine the probability of emission of a photon in a state
of magnetic type. In this case in accordance with (6.16) we have

"" 4 ^ ] / ^ Yai°M> (27.16)

where = £L(cor) .
On substituting this expression into (27.4) and on utilizing expression
(27.6") for gL((or) we obtain

y i-/ = - 4 ^ ] /

= ~ ey /i t 'g z +*i)T r J (27.17)


We have seen earlier that if the wavelength of the emitted photon
is large compared to the size of the region within which the electron
motion is confined, then the velocity of the electron is small compared
to the velocity of light. Therefore, in formula (27.17) the functions
xpi and y>2 can be expressed in terms of the nonrelativistic two-component
Pauli wave functions 99 (cf. § 15).
In order to introduce the functions (p into the integrand of (27.17)
we note that in accordance with (15.11) the term —eaA in the Hamil-
352 QUANTUM ELECTRODYNAMICS

tonian of the Dirac equation corresponds to the term —(e/2m) {(pA -\-Ap)
curl A} in the Hamiltonian of the Pauli equation (if we neglect
in it the term quadratic in A). Therefore, the matrix element of the
operator for the interaction energy eaA evaluated using the Dirac func­
tions must be equal to the matrix element of the interaction energy
(e/2m) {(pv4 + /Ip)+|x curl A} evaluated using the Pauli functions:

J ’v ie a A y 1dr = ^ <p%^^(pA + A p)+ p.cm \ A ^p l(ir, p. = ^ (27.18)

If div A = 0, then we have pA-}-Ap = 2Ap.


In the case of interest to us we have A = afy, with div a{fy = 0.
On noting that in accordance with (4.16)

v< o) = —I
1 LM — [rVYLM],
]/ L(L+X)
we can express in terms of the function <&LM= gL(cor) YlM\

Therefore, we have

2“ffip = }/ L ( r + 1) v 0 “ Ll
where L = [rp] is the operator for the electron orbital angular momen­
tum. Further, on taking into account the fact that in accordance with
(6.17) and (4.29)
curlfl^
v u l i UL M
= —UJULM'>
and on utilizing (27.7) and (27.8), we have

curl afy = V 0^.

On substituting these formulas into (27.17) and (27.18), we obtain

<Pidr.
m \/ L(L+\)
(27.19)
We now define the operator for the magnetic multipole moment
of the electron:
471 1
V{rLYLM)\ L+p. (27.20)
= /2L+1 m (L + 1), ]/ 471
INTERACTION OF ELECTRONS WITH PHOTONS 353

For the case M = 0 this definition leads to

Q « = V ( ^ ( cos ^ ) ) [ ^ L +(1] )?L .


If L = 1, then

in accordance with the usual definition of the magnetic dipole moment.


On introducing the matrix element of the magnetic multipole moment
(Qlm)2! = j <P*Qlm<Pi dr, we can write Ut^ f in the form

This expression differs from (27.11) only by replacing Qfy by


Therefore, by utilizing (27.4) we can directly write the expression for
the emission probability for a photon in a state of magnetic type of
angular momentum L, angular momentum component M, and parity
(—1)£+1:

LM L (2 L + \) [(2L—l)!!]2 " 2L+1i (Q1


l%)2i \2- (27.22)

We have derived formulas (27.14) and (27.22) for the probability


of multipole radiation only for the case of a single particle, but they
are also applicable to an arbitrary system of interacting particles, as
long as the velocities of the particles are small compared to the velocity
of light, and the wavelength of the emitted photon is large compared
to the size of the system, coa < 1 (a is the effective size of the region
within which the particles move).
In the case of a system of particles should be interpreted as
the multipole moment of the system which is equal to the sum of the
multipole moments of the individual particles.
In particular, formulas (27.14) and (27.22) can also be used to esti­
mate the probability of emission of radiation by nucleons moving within
a nucleus, but in this case we must take into account the anomalous
magnetic moments of the proton and the neutron. Moreover, the neutron
must be treated as a particle without charge, but possessing a magnetic
moment.
354 QUANTUM ELECTRODYNAMICS

We should keep in mind that while for electrons moving inside an


atom the following order of magnitude expression holds: c o a ^ v ^
(e2/47i) = (1/137), the quantities a, co, v are not related in any definite
manner in the case of nucleons inside a nucleus. Therefore, in order
that the formulas (27.14), (27.22) should be applicable to the determi­
nation of the probability of emission of radiation from nuclei, the two
relations ooa < 1 and v < 1 must be satisfied independently.
In order to obtain a rough estimate of the absolute value of the
emission probability we can assume that in order of magnitude the
matrix elements of the multipole moments are given by
(27.23)

Formulas (27.14) and (27.22) can be utilized for determining the


probability of emission of a photon of definite momentum and polari­
zation. For this we must make use of expansion (4.32) of a plane polar­
ized wave in terms of spherical waves

and substitute this expression into formula (27.5) for the matrix element
which determines the emission of a photon of momentum k and polar­
ization e. If the wavelength is large compared to the size of the radiating
system, then only one term in the expansion of e exp (ikr) will play
an essential role. Let this term be characterized by certain values of
L, M, X (generally speaking, if there are no special considerations for­
bidding this, we have L = 1 , 2 = 1 which corresponds to electric dipole
radiation). Then the probability of emission of a photon of momentum
k lying within the solid angle do, and of polarization e, will be given
by

(27.24)

If we integrate this expression over do and sum over the two independent
photon polarizations e we obtain the total probability of emission equal
to the quantity
In all the preceding formulas it was assumed that the number of
photons in the initial state (before emission occurs) is equal to zero.
INTERACTION OF ELECTRONS WITH PHOTONS 355

If there are some photons in the initial state, and their number is equal
to N, then the preceding formulas for the emission probabilities must
be multiplied by N + 1.

27.4. Selection Rules


In order that the matrix element of a particular multipole moment
should differ from zero, certain conditions arising from the laws of
conservation of angular momentum and of parity must be satisfied.
These conditions are called selection rules and consist of the following.
If j 1 and j 2 are the angular momenta of the electron, and ml and m2
are their components, in the initial and the final states, then the
following relations must be satisfied:
m1—m2= M and \ j \ - j 2 \ < L ^ j x+ j 2, (27.25)

where L is the angular momentum of the photon, and M is its component.


Moreover, the following condition must hold:

Pi = PZP, (27.25')
where Px and P2 are the parities of the initial and the final states of the
electron and P is the parity of the photon state given by P = (—l)i +1+A
( 2 = 0 corresponds to states of magnetic type, and 2 = 1 corresponds
to states of electric type).
If relations (27.25) and (27.25') are not satisfied, then the matrix
element of the corresponding multipole moment vanishes.
It can be easily shown that for given j x and j 2 the largest emission
probability occurs for a photon of angular momentum L = \jx—j 2\,
provided this is compatible with the parity selection rule (27.25').
Indeed, since the matrix elements Q{ll, contain (cor)L in the integrand,
with cor 1, the largest value of Q\ w i l l correspond to the smallest
possible L, i.e., u \ —j 2\.
In accordance with (27.14) and (27.22) the emission probability
contains the factor (coa)2L+1. Therefore, if the difference in the
angular momenta of the initial and the final states j x—j 2 is large,
then the emission probability may turn out to be very small, and
the time during which the radiating system remains in the excited
state may turn out to be large. Such long-lived excited states are
said to be metastable.
356 QUANTUM ELECTRODYNAMICS

Estimates of nuclear multipole moments made in accordance with


(27.23) lead to the correct order of magnitude for the lifetime of nuclear
metastable states, which for L = 3-5 can attain values of the order
of minutes, hours and longer. Nuclei existing in such states are called
isomers. Isomers behave like almost stable nuclei, and differ from nuclei
in the ground state in a number of their properties, for example in their
periods of /?-decay.

27.5. Angular Distribution and Polarization o f the Radiation


The state of a system (for example, of a nucleus) of a given energy
e1 and angular momentum j \ is degenerate with respect to the quantum
number specifying the component of angular momentum mx\
xp1 = Therefore, in the general case the state of a nucleus
mt
of given f is described by the polarization density matrix :
(w ll(? llm i) = Cm'Cm'*
where the bar denotes a certain average (cf. subsection 2.4).
It is convenient to introduce the concept of nuclear polarization
moments. The name of /th order polarization moment is given to the
/-vector defined in the following manner:
(27 .26)
where Cjj™j,m are vector addition coefficients (cf. §8); it is understood
that both here and later summation is carried out over repeated m-indices.
In particular, = S p ^ = 1. The polarization moment f >{1) is pro­
portional to the expectation value of the angular momentum operator J,
1
= J a’
Va Oi + 1)
9)(2) is proportional to the expectation value of the quadrupole mo­
ment tensor
______ 3JaJ , j i i j f 1)_____
aP l4 (7 i + lH 2 A ^ lH 2 ^ + 3 ) '
If the density matrix ^ is diagonal (wx[ | w^) = m-, then
the polarization moments have only the one component different
from zero, which we denote by $>,;
y t = 2 Q i( m ) C j$ l0.
m
INTERACTION OF ELECTRONS WITH PHOTONS 357

In particular,

,-p = . rp = 3w f-y1( A + l ) ______


1 VKh + ^ ’ 2 V ^ O i+ 1 )(2 A -1 )(2 A + 3 T ’
where f ( m j = £ ei(™i)/("b)-
mi

The density matrix (m jg |m[) can obviously be expressed in terms


of the polarization moments in the following manner:
2/1 i
(m, 1 i m|) = 2 ; s>™. (2 7 .2 7 )
1=0 4 /i-r1
We investigate the radiation from a nucleus whose initial state is
specified by the density matrix We denote by (j2m2L M \ S \ j 1m1) the
element of the scattering matrix corresponding to the transition of the
nucleus from the state j lf to the state j 2, m2 accompanied by the
emission of a photon whose angular momentum is defined by the quan­
tum numbers L, M. We assume that the matrix element (j2m2LM\S\jl m])
is normalized so that the probability of the transition -»j 2m2 is
equal to the square of its absolute value. On comparing this with (27.13)
we have
(y2w 2LA /1SIa/?*!) = j / 2 R UUm^ hmt.

A complete analysis of the emission process can be given by intro­


ducing the matrix
w = S^S+j
i.e.,
(m'2M '\ w \ m 2M)
= (j2m2LM' \ S\ j l m[)*. (27.28)
If we know the matrix w we can obtain, firstly, the transition probability
or the mean lifetime, r, of the excited state

-=Spw (27.29)
x
and, secondly, the density matrix for the final state
q2 = rw. (27.30)
We can easily obtain the dependence of the matrix element
(j2m2L M | S l y ' ! o n the quantum numbers ml5 m2, M. In accordance
358 QUANTUM ELECTRODYNAMICS

with (27.11), (27.21) it is proportional to the quantity S W ^ Q lu ^Ph^dr^


Since the wave function y>jm is a component of a y'-vector, while the
multipole moment Qfy is a component of an L-vector we have (cf. § 8)
{j2m2L M \ S \ j 1m1) = QC)\Z\lM, (27.31)
where Q is a quantity independent of m1, m2, and M. On substituting
(27.31) into (27.28), we obtain
(m'2M '\ w \ m 2M ) = \ Q \ H m 1\Ql\m[) Cj%LMC$ LM, . (27.32)
The substitution of (27.32) into (27.29) yields

— = \Q\2. (27.33)
T

Thus, the lifetime of the excited state is independent of its polari­


zation. On comparing (27.33) with (27.14) or (27.22) we see that
£ w<m
l= \ Q\Z an<3 does not depend on ml .
M
The density matrix q2 describes the state of the nucleus-photon
system. It is given by expressions (27.30), (27.32) in the representation
in which the photon state is described by definite values of angular
momentum and parity, and in accordance with (27.32) q2 does not
depend on the photon parity. It is convenient to go over to the represne-
tation in which the photon states are characterized by a definite momen­
tum k (i.e., for a given energy characterized by a definite direction of
motion n = k/k) and polarization a. On utilizing the expansion of
spherical waves in terms of plane waves (subsection 4.4) we obtain
(m2ria’\Q2\m2na) = (m ;M '|e2|m2M)(F<^(/i))(1(F ^ f(/i'))a/,
or in virtue of (27.30) and (27.32)
{m2ri a\Q2\m2na)
= (2 7 .3 4 )
We now determine the angular distribution of the radiation W(n).
It is expressed in terms of the matrix q2 in the following manner:
W(n) = £ (m2na\Q2\m2na)
m^a

or in virtue of (27.34) it is given by


W(n) = ( m ^ K ) C |; I ^ C f A M. F ^ ( « ) F ^ ( fl). (27.35)
INTERACTION OF ELECTRONS WITH PHOTONS 359

On writing ^ in the form (27.27), and on carrying out the summation


over m1( m[, and m2 in accordance with (8.10), we obtain
»n«) = (27.36)
I
where
(2/+ l)(2y1+ 1 )1/2
ai = W(LL,j1j 1;lj2) (27.36')
(2L+1)1/2
and W (LL ,jl j l \lj2) are the Racah coefficients.
Further, on making use of the expansion of a product of vector
spherical harmonics in terms of spherical harmonics (8.13') we obtain
= (27.37)
I
where

A = <*, j / ^ ^ { 2 L + \ ) C I U W { L L , LL \ 11).

If the matrix q± is diagonal, then (27.37) is an expansion in terms of


Legendre polynomials

» -(« )= (cos#). (27.38)

Since in the expansion of Yjfy? in terms of Ylm I are integers,


then only polarization moments of even order can affect the angular
distribution. The angular distribution is invariant under a replacement
of n by —n which is a consequence of the law of conservation of parity
in quantum electrodynamics. The even integer I must, further, in con­
sequence of (27.27) and (8.10) satisfy the two conditions (1/2) ^ j \ , and
W ) ^ L. Therefore, in particular, the radiation from a nucleus of
spin 0 or \ is always isotropic. If j\ = 1, 3/2 or L = 1, then

W(n) = — + ^ Y 2m(n).

We now consider the photon polarization. The polarization density


matrix Qaa. for a photon emitted in the direction n is given by

ft*.' = £ (m2na\e2\m2na).

On utilizing (27.34) and (27.36') we obtain

2 > c LM n.x n /\y1 uLM


)
Qna W(n) LM' Vm J m
360 QUANTUM ELECTRODYNAMICS

We now obtain the circular polarization of the photon (cf. subsec­


tion 2.4). In order to do this we take a and a' to be the Cartesian axes
in the plane perpendicular to n. On applying formula (2.15) we obtain

^ = K 'i W Y S W l * . (27.39)

Since —i[Yif y ( n ) Y l]%I*(ri)]n= — (n), then with the aid of


the expansion (8.13') we can write (27.39) in the form

(‘ = 1 (27.39')
where

n = 1 CL^ W ( L L , L - \ L - , H )

+ j/lC £ ^ 10r0W{LL, L + \ L \ l ' l ) \ .


If the density matrix q2 is diagonal, then

*■ = i k f ? <2 7 ' 3 9 " )

In particular, if only the first order polarization moment differs


from zero, then
(2yx+ 1)3/2(2L+1)1/2
£2 — W{LL, j 1j 1; 1h ) 9 x cos ft.
\ /L { L + \)
In the expansions (27.39) /' is an odd integer satisfying the inequality
V ^ 2j\ and / '^ 2 L + 1 . In contrast to the angular distribution the
circular polarization is determined by the polarization moments of
odd order. Therefore, only a measurement both of the angular distri­
bution and the circular polarization of the radiation completely deter­
mines the polarization state of the nucleus.
We now proceed to the investigation of the polarization of the final
state, i.e., of the state of the nucleus after emission has taken place.
It is determined by the density matrix

(mz 1£ 2 1mi) = f do (m2na\Qz\m2na) = (m2M\Q2\m2M ),

i.e., by taking the average of the density matrix (27.30) over the photon
states. On utilizing (27.32) or (27.34) we obtain
( m 2 | Qz I m 2) — (.m i I (?11 m 'l) C j ^ L M C ^ m ' LM (27.40)
INTERACTION OF ELECTRONS WITH PHOTONS 361

In accordance with (27.26) formula (27.40) gives the following relation


between the polarization moments 5 ^ of the initial and of the
final states:
9i!!, = (2 /i+ l)1'* (2/,-t-l)1” (27.41)
If the initial state is unpolarized, then the final state will evidently
also be unpolarized. However, in this case a correlation of polarization
occurs, i.e., for a given direction of propagation of the emitted photon
the nucleus becomes polarized after emitting the photon. Such a corre­
lation of polarization is defined by the density matrix

(m21Q2{n) |m'2) = (m2na\Q2\m2na).

On substituting into this equation expression (27.34) and on setting

we obtain, after utilizing the expansion (8.13),

(27.42)
/
where

«, = ] / W (LU LL; u-)


or

5’!!" = 0' w r y“ (”) ' (27-43)


If we choose the direction of motion n of the photon as the z-axis,
then
i = „ _ _ ^ ± 1___. (27.43')
111 1 1/M 2M T )

Since in (27.42) / is an even integer, polarization moments of only


even order occur with / ^ 2L and / ^ 2]2-
If the final state of the nucleus is also an excited one, then a subse­
quent emission of a photon of angular momentum L' may take place
with the nucleus making a transition from the state of angular momen­
tum j 2 to a state of angular momentum j 3. In this case the correlation
362 QUANTUM ELECTRODYNAMICS

of polarization leads to an angular correlation between the directions


of motion of the two photons emitted in cascade (83L
From (27.43') and (27.38) we obtain the following formula for
the angular correlation W(d):

W(B) = 7^ . ^ a ipi Pl {cos0), (27.44)


;
where 6 is the angle between the directions of motion of the two photons,
at is the coefficient in (27.42) and /S, is the coefficient in (27.37), in
which j 1 has been replaced by j 2, j 2 has been replaced by j 3 and L has
been replaced by L'.
We give the values for the coefficients in the expansion (27.44)
in some frequently occurring cases. We rewrite (27.44) in the form (71),
(127):
W ( V = Z A M cos0),

A 2n = (2 n + \)b 2n{L)b2n{L')t2n,
where
/ k (k + \) \ (2 L + l)\k \(L + k l2 )\
bk{L)
\ 2 L (L + \)l(2 L + k + \)\{kl2)\{L-kl2) \ ’

+ 1 for 7i = h i L, j 3 = j 2^ fL ,
(2j2—k) ! (2j2Jr k Jr 1)!
tk for _/] j 2 L ,j z -—j 2 L ,
(?j2)\(2 jt + iy:
(2J2)l (2/a+ l) !
*k for A — jxA-L, j z — y2+ L ',
(2j2—k)l(2j2+ k + i)!
2^2+3
^2 for 71 = 1,73 = 7 2 —^',
J2
2/ 2-1
for 7 i = 1,7 3 = 7 2 + ^ ' ,
72+ 1
( 2 / 2 - 1 ) (2/2 + 3)
*2 for 7i = 73 = 1•
72 ( 7 2 +1)

The phenomenon of angular correlation of y-rays is one of the


experimental methods for the determination of nuclear moments.
In order to be able to observe this phenomenon, it is necessary that
the transition probability from the intermediate state j 2 into the final
INTERACTION OF ELECTRONS WITH PHOTONS 363

state j 3 should be sufficiently great, otherwise the polarization of the


intermediate state will be destroyed by the interaction with the atomic
electron shells. It is. obvious that this condition reduces to the inequality
1/r > v, where l/r is the probability of the transition j 2 ->• j 3, while
v is the magnitude of the hyperfine splitting of the atomic levels.

§ 28. Scattering of a Photon by a Free Electron

28.1. Scattering Matrix Element


In § 27 we have shown that a free electron cannot emit or absorb
a photon, i.e., that first order processes are impossible for a free electron.
Therefore, the simplest processes arising from the interaction between
photons and free electrons are described by the second order scattering
matrix, which in accordance with (24.14) has the form
2 /*
s<2) = - - J T [7/(y> (xjA ( x j r p ( x ^ N (ip(x2)A ( x 2) y > ( x j ) \ d i x l d * x 2 .
Since the matrix S<2) contains the operator for the potential A^Cx)
bilinearly, its elements describe those processes in which the total
number of photon states is equal either to two or to zero. The latter
case refers to the interaction between electrons with no photons tak­
ing part, and will be discussed in Chapter VI.
The matrix elements of S(2) involving two photon states may be
divided into three kinds, corresponding to the total number of electront
in the initial and the final states. This number may be equal to four,
two, or zero. The matrix elements with four electron states are simply
products of first order matrix elements of S(1) (cf. Fig. 13, 1). The matrix
element which does not involve any electron states determines the
“proper mass” of the photon (cf. Fig. 13, 4) and will be discussed in § 44.
Matrix elements involving two electron states are of the greatest
physical interest. Processes involving two photon and two electron
states are described by that part of the second order scattering matrix
which is of the form
S(2) = e2N J y (x 2)A (x 2) S c(x2—x 1)A (x 1)y (x l)dix1dix 2, (28.1)

where the function Sc(x) is defined by formulas (19.16), (25.2). The


simplest process described by the matrix S(2) is the scattering of a
364 QUANTUM ELECTRODYNAMICS

photon by a free electron, the detailed investigation of which we now


undertake.
We denote by y>i(x) and the wave functions of the initial
and final electron states and by and A 2(x) the potentials of the
initial and final photon states. We obtain the element of the matrix
S<2) corresponding to photon-electron scattering if we replace in (28.1)
the operators y)(x2) and tpixj) by the functions yj2(x2) and ^ ( x j , and
the operator A(x2) A ^ ) by the function A*(x?) A ^ x ^ + A ^ x ^ A * (x ^ ):
£{!>,= e2f (x2) [A$ (x,) S c(x2- x 1)A1(xx)
+ ^1 (* 2) S c(x2—x 1) A* (x1)]y)1(x1)dlx 1dix 2. (28.2)

We choose the functions y T(x) and A r(x) (r = 1,2) in the form of


plane waves, normalized per unit volume:

w(*) = uTeiprx and Ar(x) =


\ 2 cor
where pT and kT are the four-momenta of the electron and the photon,
cor= \ kT\ is the photon frequency, ur and eT are unit (spinor and vector),
amplitudes.
k2 *1 k.

Pi Pix *Pi P\
Fig. 21.

On substituting these functions and the expression (25.2) for Sc(x)


into (28.2) we obtain the following result which is in accordance with
the general rules of subsection 25.5:
le“ i f i —m a . a i f 2—m „
S&, = “2|< p2-
2 / ia>, co, '2f \ + r t ei+ e'T l + ^ h e' r
(28.3)
where f x = p1+ k 1 = p2-\-k2 and = Py—k 2 — p2—k x. Figure 21
shows two diagrams corresponding to the two terms of the matrix
element S ^ f .
INTERACTION OF ELECTRONS WITH PHOTONS 365

28.2. Application o f Conservation Laws


The matrix element S ^ f differs from zero only when the laws of
conservation of energy and momentum are satisfied:
P i+ kx =p 2+ k 2. (28.4)
For given momenta p x and kx these laws enable us to obtain four of
the six components of the momenta of the final state; if, in addition,
we specify the direction of k 2 or of p 2, then we can completely determine
k 2 and p 2.
We now obtain the dependence of the frequency of the scattered
photon a>2 on the direction of its momentum. On squaring (28.4) p \+ k \
+ 2p xkx = p l+ k\-\-2p2k2 and on noting that
p\ = p\ = —m2 and k\ — k\ — 0,
we obtain p xkx = p2k2 or p xkx = pxk2-\-kxk2, whence it follows that

^ ( l — COS0J = a)2( l —v1cosff2) + (1—cost?), (28.5)


ei
where vx = \p1\/e1 is the initial electron velocity, ex is the initial energy,
&x and $ 2 are the angles between the momenta of the primary and the
scattered photons and the initial electron momentum, # is the angle
between kx and k2.
In the case of the scattering of a photon by an electron at rest (ax = 0,
= m) we obtain from (28.5) the well-known Compton formula

^ l+(co1/m )(l—cos??) (28.6)


The quantities f x and f 2 appearing in are the four-momenta
of the virtual electrons for which the usual relation between the energy
and the momentum does not hold. The difference between a virtual
electron state and a real one can be characterized by the quantity
/ r2+ m 2
where r — 1 ,2 , (28.7)

which vanishes in the case of a real state. It can be easily shown that
m2xx = 2p xk x = lp 2k 2 and m2x 2 = —2pxk 2 = —2p2kx. If the electron
was at rest before the collision, then
2cox , 2co2
x x — ------------- and x2 (28.7')
rn m
366 QUANTUM ELECTRODYNAMICS

28.3. Differential Cross Section for Unpolarized Particles


In accordance with the general rules of § 26 the differential cross
section for scattering in which the momenta of the final states of the
electron and of the photon lie respectively within the intervals dp2 and
dk2, is given by

da = -----1m2 |2 ^ (^ i + - ^ 2 - ^ 2) ^ + " i— e2 - " 2),


4co1<x>2 J{2n)1
(28.8)
where Q is the matrix appearing in the figure brackets of (28.3):

Q = — — e2 ( ; / ,- « ) « ! H------— e f i f f - m ) ^ , (28.9)
mi K2
and J is the flux density of the colliding particles. In accordance with
(26.5'):
J = v = k1p1/e1atl = m2x1l2e1a)l . (28.10)

The presence in (28.8) of (5-functions enables us to carry out the


integration over dp2 and da>2. In order to do this it is sufficient merely
to perform the substitution
<K/»i+£i—P2 —k 2) dp2 -*■ 1,
a>2do2
<5(ei + OJi - e2 - (0i) dk2
(dfda)2) ( c o 2 + e 2) j
where do2 is the element of solid angle containing the vector k 2. In
the last expression e2 is a function of co2, since the momentum p 2 is
related to k 2 by the conservation law
e2 = \/m 2 jr {p 1 + k 1 —f c 2)2
= ]/ £i + tUi + 2/7l a)1cos^1+ co‘2—2 pl co2cos ,&2—2(i)l (x>2 cos$.
From this it follows that
CO, ,—pl cos'&1— a>l cos& k 2{k2—p x—k^)
doj. ' (W2+ ^2) -- f + = 1+

Pi Piki m2xx
(28.11)
co2e2 £2(i>i 2e 2x>2
<

and therefore
INTERACTION OF ELECTRONS WITH PHOTONS 367

where r0 = e2/4nm is the classical electron “radius” (e/j/4n), is the


electron charge in Gaussian units.
If the initial electron state is unpolarized and we are interested in
the result of scattering without reference to the polarization of the
electron in the final state we must sum expression (28.12) over the
final values, and average it over the initial values of the electron spin
component. For the sake of brevity we denote this quantity, i.e.,
2 £ '^'^d o (Hi and p 2 are the values of the electron spin components

in the initial and the final states), as before, by da. In accordance with
(26.7) such an averaged value of the differential cross section is
equal to
Y^ (j-p- j __

do = ~T
Z tfi U Sp ( 2 0 'A - w ) 2 0 'A - w ) } do2, (28.13)
where
Q = yiQ +Yi-
On substituting into this expression the value of Q, and on noting
that for any four-vector q whose spatial components are real and whose
fourth component is imaginary, q = —q, we obtain
Q (iPi.-m ) Q(ip2~ m )

\rfl /v^ TYl /V2 I

x + e2(if2- m ) e ^ \ ( ip 2- m ) . (28.14)
m x2 J
If the primary photon is also unpolarized, we obtain the scattering
cross section, irrespective of the polarization of the scattered photon,
by summing (28.13) over the final polarization states, and by averaging
it over the initial polarization states of the photons. We denote this
quantity, i.e., \ E VuVtda, where vx and v2 are the photon polarizations,
also by da.
The evaluation of this averaged cross section is considerably simpli­
fied, since the summation may be carried out not over two, but over
four photon polarizations, including both the “longitudinal” and
the “scalar” polarizations (cf. subsection 26.3). In doing this we must
simply replace in expression (28.14) ex by yv, e2 by and sum over v
and X (v ,X — 1, 2, 3, 4).
368 QUANTUM ELECTRODYNAMICS

Thus, we obtain
da = r2 to! do2 SP (28.15)
°m2x?
where
m ifz—m
SpF
y' + y ' - m ^ n _

X ( j P i - m ) y , ifl m r i ( ‘P 2 - m ) ,

—m ifo—m
+Sp V + -—
m2xx v y v /wz—
x2 y*

X (;ip1~ m ) y x ^ m y„(iP2 - w ) [ ■ (28.15')


mzx2 )
It can be easily shown that the second term in this expression is
obtained from the first one by means of the substitution k x ~+ —k 2,
k 2 -+ —k x which also corresponds to the following substitution
/ i fir fz f r xi xir x 2 Ki ■Obviously, each of the terms is a func­
tion of only the two invariants xx and x2. Therefore, we can write SpF
in the following form:
SpF = P(xl5 x2) + P(x2, Xj),
(28.16)
P<*1, *2) = Ki^ir *2) + h2(xx, x2),
where
1 A A A

A(*i» X2) = —r r S p { y A(i/1- m ) y v(z/71-m )y ,,(i/1- m ) y A(i>2- m ) } ,


m

Kbii, >h) = —r — Sp {?„ 0 / 2 -


m xix2
A

m) yx(z>! - m) yv{ifx-m ) yx(ip2- m)}.


a A

On carrying out in A1(x1, x2) the summation over v and X with the
aid of formula (26.14), and on dropping the terms with an odd number
of matrices X, we obtain

K(xir * 2 )= —TT ™2A A + 4 m 4}.


nPx\ s P{fiPifiP 2 + ^ 2 ( f i P i + f i P 2 - f i A )
From here we can easily obtain with the aid of formulas (26.15) (26.16)
INTERACTION OF ELECTRONS WITH PHOTONS 369

In an analogous manner we can show that

h b i i. *2) = (4 —Xj —X2).


«1«2
Therefore
1
SPjF ~ U0 = A I——|- — j - 4 ^ + — . (28.16')
5 \«! «2/ \«1 *2/ l« 2 « l/

and we finally obtain the following expression for the averaged cross
section do:
da — _22T2 UnOjldo9
" U02 ’ (28.17)
m*x ^0L

This expression confirms that do is relativistically invariant. Indeed,


and U0 are invariants, and the quantity a>2do which transforms like
<odcodo = d3kl<jj = 25(k2)dik, is also an invariant.
If the electron was initially at rest, i.e. p 1 = 0, then on using
(28.6) and (28.7') we obtain U0 = (<u1/co2)+(co2/oj1) —sin2 and the
scattering cross section assumes the form (107), (193)

sin2# db. (28.18)

We note that the scattering cross section for the photon contains
the mass of the scatterer m in the denominator. Therefore as m -> 0 0
no scattering occurs. However, this conclusion is not exact, since it
was obtained only by investigating the second order scattering matrix.
It turns out that the fourth order scattering matrix already contains
elements corresponding to the scattering of a photon by an infinitely
heavy charge. The phenomenon of coherent scattering of a photon by a
nucleus corresponding to these matrix elements will be investigated in § 55.
28.4. Angular Distribution and Total Cross Section
By utilizing expression (28.6) v/hich determines as a function
of co1 and # we can express the differential scattering cross section
for unpolarized photons in terms of the frequency of the incident pho­
ton and the scattering angle for the photon
do —
rl 1 + c o s2# I " _________ (tu1/m)2( 1—cos#)2______
2 [l + (<u1/m )(l—cos#)]21_ (l+ c o s2# )[l+ (u j1/m )(l—cos#)]
(28.19)
370 QUANTUM ELECTRODYNAMICS

The angular distribution of the unpolarized photons given by this


formula is shown in Fig. 22 for different values of the parameter
y = cojm .
At low photon energies co1 < m formula (28.19) reduces to the
classic Thomson formula

da = (1 -fcos2#) do2, where cox -< m. (28.20)

If the energy of the primary photon is large compared to the


electron rest energy cox m then from (28.19) we can obtain simple
expressions for two limiting cases, when •&<!]/ml(o1 and when
$ >> \/mj(jol :

da ___ ..2r%do2 > where col >> m, & << (28.21)


X Oil

d o = -2 -
m do2
co, 1—cos# ’
where co1 >■ m, $ !>
V-
X 0)t
(28.21')

For -&= \ /mj(o1(co1 > m) both these formulas lead to the same
results. This shows that for cox >> m formula (22.21) gives a good ap­
proximation to the angular distribution over the angular interval
0^ ^ \/m/a)1,
while formula (28.21') gives a good approximation
over the angular interval •&^
INTERACTION OF ELECTRONS WITH PHOTONS 371

The differential scattering cross section can also be expressed with


the aid of (28.6) in terms of the energy a>2 of the scattered photon.
After integrating over the azimuthal angle we obtain
md(oz TtUj CO,
da = 7ir\ (28.22)
cof [ oj2 + CO,

where in accordance with (28.6) co2 varies over the range


co.
^ co2 cox,
l+2(co1/m)
Integration of formula (28.19) over the angle $ leads to the follow­
ing expression for the total scattering cross section:

a= 3 h+y 2y(i+y) - l n ( l + 2 y ) +
ln(l+ 2y) l+ 3 y
J a° [ V3 l+ 2 y 2y (l+ 2 y )2I
(28.23)
where a0 = (8n/3)r3 and y == (lojm ).
For coi < w, expansion of (28.23) in powers of y yields
cr = cr0( l —2y + ...), where y < 1, (28.24)
i.e., in the nonrelativistic domain the scattering cross section is almost
independent of the energy.

In the extreme relativistic domain cox > m the expression for the
cross section is greatly simplified:
3 m (28.25)
a where m.
8°°^ m
372 QUANTUM ELECTRODYNAMICS

In this region the cross section falls off almost inversely with the photon
energy.
Thus, at high energies the number of scattered photons is greatly
reduced, as a result of which the penetrating power of y-radiation
increases with increasing energy.
Fig. 23 shows the dependence of a/a0 on the energy of the primary
photon. The values of alo0 are given in table 9.

T able 9

r 0.05 0.1 0.2 0.33 0.5 1 2 3


tf/ft0 0.913 0.84 0.737 0.637 0.563 0.431 0.314 0.254
7 5 10 20 50 100 200 500 1000
ajo0 19.1 12.3 7.54 3.76 2.15 1.22 0.556 0.304

28.5. Distribution o f Recoil Electrons


As a result of its collision with the photon the electron which was
initially at rest acquires some energy. From the conservation laws
it can be easily shown that the energy acquired by it is given by
ft)?(1 —cos#)
E —m = (28.26)
m f - o o f l —cos#)
It varies between zero at # = 0 up to the maximum value £ max —m
= 2ft)J/(m+co1) at # = n.
The angle /? between the momentum of the scattered electron and
the momentum of the primary photon is related to the angle # by

c o s = ( l + y ) l / - v , /V --« v (28.27)
\ 2 + y ( y + 2 ) ( l —c o s # )
and varies from n i l at # = 0 to zero at # = n.
By utilizing this relation we can express do in terms of /?. Thus,
we obtain angular distribution of the recoil electrons
da = W (l + y)2co^ d°e
0 [ l + 2 y - f y 2 sin 2/?]2

2y2 cos4/? 2(1 + y ) 2 s in 2/? c o s2/?)


X 1+
[ l + 2 y + y 2 s in 2/ ? ] [ l + y ( y + 2 ) sin 2/?] [ l + y ( y + 2 ) s i n 2/?]2 j ’
(28.28)
INTERACTION OF ELECTRONS WITH PHOTONS 373

where doe is the element of solid angle containing the momentum of


the scattered electron.
In the extreme relativistic case y >> 1 this formula yields
•2 J
da = , where y > l and y/32 < l ;
T P ' + 1
(28.29)
4r$ cos (1 doe .
da = ------. , D , where y > 1 and y sin2/? > 1.
y2 sin4/?

28.6. Scattering o f Polarized Photons


We now investigate the scattering of polarized photons by unpo­
larized electrons. We describe the polarization state of the incident
photon by the density matrix (cf. subsection 2.4)
(28.30)
where are the Pauli matrices and £{j1)( j = 1 ,2 ,3 ) are the Stokes
parameters. By utilizing the general formulas of subsection 26.3 we
obtain in place of (28.12) and (28.13) the following expression for
the differential cross section:

d r(5 (1)) = } rm S p ( 2 ^ 1- m ) e<|j)^ 2-m ))rfo2 (28.31)


Z irl

where (cf. (28.9))-.

Qa, = m2x2 ------ e f )( ifi~ m ) e f)


*
and e ^ ) (a = 1, 2) are unit vectors in the plane perpendicular to
the vector k v We choose them in the following manner:
(1)_ [ k i,k 2] and P(i) _—
e2 [Anci1*];
61 | [Ari, A:231 I*i|
e f ] are unit vectors in the plane perpendicular to k 2(fj, = 1,2). As
has been explained in subsection 26.3, in expression (28.31) the summa­
tion over ^ may be extended over the values /j, = 1, 2, 3, 4, i.e., we
can replace e^2) by y^.
On evaluating the trace (over the spinor indices) in (28.31) we
obtain in the case of an electron at rest
da(%(1)) = d o o + i^ d o i, (28.32)
374 QUANTUM ELECTRODYNAMICS

where d o 0 is the scattering cross section for an unpolarized photon


determined by formula (28.18), and
1 2
d o x = ^ - r 20 — sin2# d o . (28.33)
2 oi\
The expression for do can also be written in the form

d a = \ rl[ ^ k ) Fdo' (2834)


where
F = - ^ + - ^ + ( £ < 1> -l)sin 2tf. (28.34')

We now obtain the polarization of the scattered photon. By apply­


ing the results of subsection 26.4 we obtain the expression for the
Stokes parameters for the scattered photon,

£(2) = (28 35)


Sp {Qai£ $ i - m) e§' QpXiPi - m)}
Here the summation over all the variables (a, /?, //, v) is carried out
over the values 1 and 2. On choosing the unit vectors e<,2)
(2) _ ftn ^ 2] e { 2) _
e2 —
1 i ^ i ^ 2r l*2|
we obtain after evaluating the trace in (28.35)

£{,2) = (sin2$ + ( l + cos2# ) ^ 1’) -y ,

£{2) = 2£[1) cos# — , (28.36)


r

£(2) _ ( Q>1 I 602 j £<1) C0S^


\ ^2 ^1 / F
From these formulas we can see, in particular, that the unpolarized
photon becomes partially polarized as a result of scattering. Indeed,
on setting f j1*= 0 in (28.36) we obtain
£<2>= £<2>= 0,
(28.36')
(ct»1/co2)+ (^ 2 /^ i)—sin2# ‘
INTERACTION OF ELECTRONS WITH PHOTONS 375

Since £^2) > 0 the photon is polarized in the direction e[2), i.e., per­
pendicular to the scattering plane. In the nonrelativistic approximation
we have £^2) = sin2# /(l+ c o s2#) and in the case of scattering through
90° the radiation is completely polarized.
We consider the case when the incident photon is linearly polarized
= o, (£|1))2+(£<1))2 = 1), and seek the differential scattering cross
section in the case when the scattered photon is also linearly polarized.
In order to do this we utilize formula (26.27) and expressions (28.18)
and (28.35) for da and £j2). On expressing the parameters £\2) and
^ in accordance with (2.15) in terms of the components of the photon
polarization vectors which we denote by e(1) and e(2), we obtain the
following expression for the differential cross section for the scattering
of polarized photons by an electron at rest:

(28.37)

where 6 is the angle between the polarizations of the primary and the
scattered photons, cos 6 = e{1) e{2).
The scattering cross section as a function of d reaches a maximum
when the polarization directions of the primary and the scattered
photons coincide.
We can obtain the total scattered radiation for given frequencies
and a>2 by adding expressions (28.37) for the two linearly independent
polarization states of the scattered photon. We choose two mutually
orthogonal directions e{2) {jx = 1, 2; e[2) e{2) = 0), corresponding to
these states in the following manner: one of them is perpendicular to
e {1): cos0(1) = e(1)e{2) = 0 and the other lies in the plane containing
k 2 and e(1), cos2 0(2) = (e(2)e{1))2 = 1—sin2 ■&cos2 y, where <p is the
angle between the planes (Aq, k 2) and (k2, e{1)).
If da{1) and da{2) are the scattering cross sections corresponding
to these polarizations, then the total scattering cross section will be
equal to da = da{1) + da{2). If the incident photons are unpolarized,
then the sum d a ^ + d a ^ averaged over the angle <p gives the scattering
cross section for unpolarized photons (28.18).
For o j 1 < < m
da{1) = 0, da{2) = /q(1 —sin2# cos2 <p)do2. (28.38)
376 QUANTUM ELECTRODYNAMICS

For > m we have to distinguish between the regions of small and


large scattering angles. If & < |/m /a ^ , then

do{1) = 0, dol2) = r%(l—sin2& cos2<p)do2, a>l > m,


(28.39)
If ft > ]/m/coj, then
1 rl m do2 m
doa) = do{2) = dcy — .
T - f « j tOj > w, # >>
2 4 1—cos# ft>i
(28.39')
We see that in the extreme relativistic domain the scattering for
small # takes place in the same way as in the nonrelativistic domain.
At high energies and large scattering angles the scattered photon is
unpolarized independently of the nature of polarization of the primary
photon.

28.7. Scattering o f Photons by Polarized Electrons


We now discuss the scattering of photons by a polarized electron at
rest (57), (126), (217). If the expectation value of the electron spin is
equal to then its polarization density matrix has, in accordance
with subsection 10.6, the following form:

9 ^ = l ( \ + y d ( l - i l My,). (28.40)

By utilizing (26.18) and (28.31) we obtain for the differential scatter­


ing cross section

,?<«) = 4 (— j (28.41)

On evaluating the trace we obtain

d o ( ^ , ^ ) = d o W » ) + r * ^ J f t'y u G , (28.42)

where

G = — -^ -(1 —c o s #)(& ! c o s # + & 2) (28.42')

and do(tj(1)) is the cross section for the scattering of a photon by an


unpolarized electron (28.32).
INTERACTION OF ELECTRONS WITH PHOTONS 377

We note that the second term in (28.42) contains only the component
which describes the circular polarization of the incident photon,
and the component of the polarization vector of the electron in the
scattering plane.
We give the expression for the total scattering cross section in the
case when the electron is polarized in the direction k x, while the photon
has only a circular polarization:

2 \ l+4co1+5a)* (l+«u1)ln (l+ 2 « j1) \ (28-43)


nr° \ w 1(l+ 2 a Jl)2 ''2 J f'" ' j’
where a0 is the scattering cross section for unpolarized photons.

Figure 24 shows the ratio a’/a0 as a function of the photon energy


(79).
If the incident photon is unpolarized, then the polarization of the
scattered photon can be obtained from

tm = SP
1 Sp{Q .r9 a'Q „ ( i h - ’n))
In this case the scattered photon has a linear polarization determined
by formula (28.36') and, in addition to that, the circular polarization

(28.45)

where G is defined by formula (28.42'), while F is defined by formula


(28.34') with = 0.
378 QUANTUM ELECTRODYNAMICS

Finally, we investigate the polarization of the recoil electrons. In


accordance with (26.21) and (10.32) we can find the parameter a}, and
then from formula (10.30') we can find the polarization vector ^ (2).
If the electron was unpolarized before the collision then

S(2)= #> — ’ (28.46)

where F is defined by formula (28.34'), and

G' = — —(1—cos??) (&! co stf+ £ 2)—( l+ c o s # ) C°-:! ~ ^ ( f c 1—£2) ■


m L m 2-r£ 2 J
(28.46')
We see that the electron is polarized only in the case when the photon
has circular polarization.

§ 29. Bremsstrahlung

29.1. Perturbation Theory for an Electron Wave Function in the Con­


tinuum. Incoming and Outgoing Waves
When an electron collides with a charge or a system of charges, in
addition to the electron being scattered photon emission can also take
place. Such a process is referred to as bremsstrahlung. If the electron
collides with a heavy particle (nucleus, atom) the effect of the latter
can be taken into account as the effect of an external field. In this case
bremsstrahlung is described by the matrix element (27.3) in which the
initial state refers to the continuous spectrum, and the final state refers
also to either the continuous or to the discrete spectrum.
Because of the complicated nature of the wave functions of an
electron in the field of a nucleus (cf. § 13) the matrix element (27.3)
can be evaluated only in the case of low energies, when the nonrelativistic
approximation may be used, and also in the limiting case of high energies
and small scattering angles for the electron when the wave functions of
subsection 14.5 may be used.
If the external field is such that it can be treated by perturbation
theory methods, then in order to determine the probability of brems­
strahlung it is sufficient to evaluate the corresponding element of the
second order scattering matrix S(2) (28.1). This matrix element will
INTERACTION OF ELECTRONS WITH PHOTONS 379

differ from the matrix element for photon-electron scattering (28.2)


only by the fact that the potential of the primary photon must be re­
placed by the potential of the external field Ajf>(x):

f = e2f y>2( x 2) [A*(x2) Sc(x2— x j A lei( x x)


+ A {e)(xJ Sc(x2—x^) A*{x^)]ip1{x^dix1dix 2. (29.1)
Here tp^x) and ip2(x) are the wave functions for the free electron
in the initial and the final states, A/l(x) is the potential of the emitted
photon and A jf^x) is the potential of the external field.
The criterion for the applicability of such treatment obviously
coincides with the criterion for the applicability of the Born approxi­
mation to the case of the Coulomb field (Ze2/z>) -< 1, where Ze is the
nuclear charge and v is the electron velocity. We first consider this rela­
tively simple case which has quite an extensive range of applicability.
Before evaluating the second order matrix element (29.1) we note
that a comparison of (29.1) with the first order matrix element enables
us to obtain the wave function of the electron in the external field in
the first approximation of perturbation theory. Indeed, formula (29.1)
must coincide with (27.3) if the wave functions in the latter take into
account scattering by the external field. To distinguish them from the
free electron wave functions y>x and ip2 we denote them by ip[e) and
y4e) and set
Wie) = Vi+Vi and v ie) = Vz+Vi- (29.2)
Substituting these expressions into (27.3), and taking into account the
fact that (27.3) vanishes if the free electron wave functions are substituted
into it, we obtain on comparing with (29.1)

y[e)(x) = y x(x) —e J Sc(x—x') A ^ i x ^ y j ^ x ' ) ^ ' ,


(29.3)
fy e>(x) = y 2(x) —e J yj2(x')A (e)(x')Sc(x'—x)d4x'.
These formulas have a straightforward physical meaning: in the
second terms the function ^ (x ) plays the role of the Green’s function,
and the product A (e)(x)y(x) plays the role of source density, while the
second terms themselves represent scattered waves obtained as a result
of the superposition of waves scattered by each four-dimensional volume
element d*x'.
380 QUANTUM ELECTRODYNAMICS

By proceeding in an analogous manner we can also obtain the


higher order approximations for the wave function of the electron in
an external field. For example, in second order we obtain

y)^'>{x)z= y)1(x)—e j S c(x—x') A le) (x')xp1(x,)dix r


+ e2f S c{ x - x ') A M{x')Sc{ x - x " ) A (e\ x " ) ^ 1{x")dix'dix",
(29.4)
wie)ix) = ef :<P2(x ')'4(e>(x')S c(x' —x )d ix'
+ e2J rp2 (x ") A (e)(x") S c(x " —x') A (e)(x') S c(x '—x) dix'dix".
The summation of the series obtained by perturbation theory should,
in principle, lead to the exact solutions of the corresponding Dirac
equations.
If we substitute into (29.3) in place of y;Jj2 (x) the plane waves y>li2
= u12 exp (iplt2x ), we obtain
A

v4e,(*) = ui e<PlI“ (2 ^ e\ f ~ Pl) Ul elfi d f ,


A (29.5)

where A {e)( f) is the Fourier component of the external potential defined


by formula (25.3).
In an electrostatic field
A (e)(f) = 0, A f> (f) = A {0e)( f) 2 n d (f0), / ( / , if0)

p r if — iy i
WU) = UieiPlI + _ _ 3_j — _ ^ 2 Atf >( f - Pl) y i uxelfxd f ,
(29.6)
« 2c - ^ + — - J u2yi A ^ ( p 2- f ) j , - ^ L e - ^ d f .

In the nonrelativistic approximation small values of | / | play an


important role in these integrals, and therefore if ^ —m and the func­
tion ip{e) assumes the following form which is well known from the
perturbation theory for the continuous spectrum:

W e)(r) = Wp+r A{o ]U r P l V jd f, (29.6')


( 2* ) ! / £P- £f
INTERACTION OF ELECTRONS WITH PHOTONS 381

where ipp — exp (ipr) and ep = p2/2m, and the integration is carried
out in accordance with the definition of the function S^x) by taking
into account the rule for going around the pole of the integrand (cf
§ 18).
It can be easily shown that the second terms in (29.6) which, as
has been pointed out already, represent scattered waves, have the
asymptotic form of outgoing spherical waves for large values of r.
Indeed, we examine the integral appearing in (29.6):
r e'fr
J = J 8 (P ’ f ) y z r p d f ,

where g ( p ,f ) includes all the factors other than those explicitly shown
in J. On integrating over the angle # between the vectors / and r we
obtain

g0eift >ifr£
_____ dg_
f dr­ ■ fd f + d£.
P -P ‘ - w P - P z d£
where £ = cos# and gn and g0 are the value's of g ( p , f ) for ■&= n
and 0.
The second term here is evidently of order 1/r2 and may, therefore,
be neglected. In the first term we can make the substitution / -» —/ ,
as a result of which the equation assumes the form

in
J= — where r -> oo.
r

In accordance with (17.18) the poles of the integrands in J are given


by the points p + /0 and —p —i0. Therefore, on completing the contour
of integration in the first integral by a semicircle in the lower half-plane,
and in the second integral by a semicircle in the upper half-plane, we
obtain J eipT, where r ->■ oo.
r
Thus, in virtue of the rule for going around the poles of the function
S c(p), both the function y)[e) and the function y>{2e) contain only outgo­
ing waves at v —> oo and contain no incoming waves. This assertion is
obviously valid not only for the first-order approximation of perturba­
tion theory, but also for all higher order approximations. Therefore, the
382 QUANTUM ELECTRODYNAMICS

exact wave functions for an electron in an external field, which belong


to the continuous spectrum, and which may be regarded as sums of
series obtained by perturbation theory, must also have the same prop­
erty, viz., the initial state function ^ (x ) and the complex-conjugate
function for the final state yi2(x) must have at r -»• oo the asymptotic
form of the sum of a plane wave and an outgoing spherical wave (187)

elPir + a—elp'r,
r (29.7)

From this it follows that the final state wave function itself must
at r ->• oo have the form of a sum of a plane wave and an incoming
spherical wave

Y>2 ~ ev»r+ - 2- e~ip2r. (29.8)

This fact is of the greatest importance for the evaluation of the


matrix elements using the exact wave functions for the electron in an
external field. The point is that the wave functions of the continuous
spectrum have a peculiar degeneracy. In order to define them it is not
sufficient to specify the electron momentum, but we must also specify
the nature of the asymptotic behavior at infinity, since for a given
momentum the wave function at infinity may have the form either of
the sum of a plane wave and an outgoing wave, or of the sum of a plane
wave and an incoming wave. We now see that from the rule for going
around the poles of the function S c it follows that the wave functions
of the initial and the final states must be chosen in such a way that at
infinity the initial state wave function must have the form of the sum of
a plane wave and an outgoing wave, while the final state wave function
must have the form of the sum of a plane wave and an incoming wave.
This rule can also be formulated in the following manner: the wave
functions of all the particles formed as the result of any process must
at r -* oo have the form of a sum of a plane wave and an incoming
wave, while the wave functions of particles which disappear (i.e., appear
only in the initial state) must a tr - + o o have the form of a sum of a plane
wave and an outgoing wave.
INTERACTION OF ELECTRONS WITH PHOTONS 383

The nature of this asymptotic behavior admits a simple physical


interpretation. If we expand a plane wave in terms of incoming and
outgoing waves, then in the initial state wave function y)t at r -*■ oo
only the incoming waves will have amplitudes which do not depend on
the nature of the scatterer, while the outgoing waves will be completely
determined by the scattering force field. This corresponds to the fact
that in the initial state the particles are incident on the scatterer. In
the final state wave functions at r —►oo only the outgoing waves
may have amplitudes which do not depend on the nature of the scatterer,
while the incoming waves must be determined by the scattering force
field.

29.2. Effective Cross Section for Bremsstrahlung


On substituting into (29.1) the electron wave functions and the
photon potentials in the form of plane waves (cf. (25.4), (25.5)), and
also the expression (25.2) for .Sc(.x), we obtain the following form for
the matrix element S Q f which determines the bremsstrahlung cross
section:
<T(2> * ifi—m ifi—m A

e ----- 2-----
A (e)(q )+ A {e)(q) e mx, (29.9)
|/2 Q) m2x1 m2x 2
where
f i = P z fk , fz — P i—k, (1 — P z f k —Pi>
m2x i = f\-\-m 2 — 2p2k ,
mPx2 = f l + m 2 — —2p±k ,
A (e)(q) = J A (e)(x)e~iQXd*x.

This expression can also be obtained with the aid of Feynman’s


rules (cf. subsection 25.5) if we first draw the diagrams corresponding
to the process of bremsstrahlung. These diagrams are shown in Fig. 25;
they differ from the diagrams corresponding to the Comptom effect,
only by the fact that corresponding to the primary photon k x we have
here a “pseudophoton” of four-momentum q, i.e., the Fourier q-
component of the external potential, while corresponding to the scat­
tered photon we have here the emitted photon of momentum k
and polarization e. Correspondingly, the expression for the bremsstrah­
lung matrix element (29.1) differs from the Compton effect matrix
384 QUANTUM ELECTRODYNAMICS

element by the substitutions k± -> q, k2 ^> k, —>A (e)(g) and


A A
e2 —■►e.
From the laws of conservation of energy and momentum it follows
that the four-momentum of the “pseudophoton” q(q, iq0) taken with
reversed sign is equal to the four-momentum p n given to the nucleus
as a result of the bremsstrahlung process: pn = —q = p i—p 2—k. Thus,
—q is the three-dimensional nuclear recoil momentum, while —q0 is
the energy imparted to the nucleus.

If the external field does not depend on the time and is described
by the static potential A (0e)(r), as is the case for a stationary nucleus, then
A (e)(q) = y i iA(0e\q)7jid(q0),
(29.9')
^oe)(?) = j A l0e)(r)e -i4*rdr.
In this case the polarization of the pseudophoton, in contrast to the
polarization of a real photon, is directed along the x4-axis.
In accordance with (26.33) A (0e)(q) is given for the nuclear Coulomb
field by A (0e)(q) = — (Ze/q2), where e is the electron charge in Heaviside
units.
We now proceed to the determination of the differential cross section
for bremsstrahlung in the Coulomb field of a nucleus. In accordance
with (26.3') it can be written in the form
^4 dn die
d<y = \ A e)(q)\2\u2Qux\2 ^ ■afe-C z-co), (29.10)

where J is the electron flux density in the initial state, whose magnitude
is equal to their velocity, and Q is the matrix appearing in the figure
brackets of formula (29.9),

2 = - - h ( M i.
INTERACTION OF ELECTRONS WITH PHOTONS 385

We first of all obtain the bremsstrahlung cross section averaged


over the spin directions of the incident electron and summed over the
spin directions of the scattered electron and over the polarizations of
the emitted photon. Such averaging and summation over the polariza­
tions of the particles reduces, according to (26.7), to the following
replacement in (29.10):
I Qu1 12 -+ Sp F, (29.11)
0£i £2
where
SP F = S p { Q p d P i - n O Q f i p t - m ) } ,
.A A
_ ifi—m //,—m
Qn yp m*xx y i m
m~x2 y ^
2 . .
(29.11')
7T \fx ~ m , ifz— m
Q» = y & ty * = m^xx V v + V nm^x
-za 2r-y *

The quantity Sp F is a function of xx, k2, q2, elf £,:


Sp F 7i(/<i, Kn, q , £j, £2) ”)” ^ 2(^15 ^2> 9^5 £i> £2)> (29.12)
where

7 i = S p |e ^ ( v 1- m ) y 4- ^ — -y(|( ^ 2-/w )J,

74 ( ^ 2 -™ )} •
1 m y~z )
If in the expression defining 7 2 we make the substitution px -*• /?2,
A “’’A* (7 -*■ — 9 and k -*■ —k, then ^ and x2, as well as f x and / 2,
will be interchanged so that xx -*• *2, Xx, fx / 2, h -»• / 1 , and
we shall obtain in the brackets following the symbol Sp in the expression
for 7 2 the same matrices as in the expression for 7 a, only arranged in
the inverse order. From this it follows that 7 2(a> x 2, <72> £i> £2) =
7 i(* 2 > 9 2>£2 >£i) and, therefore, it is sufficient to evaluate 7 a.

The evaluation of can be carried out simply: after summing over


the subscript p in accordance with (26.14) we must, by utilizing the
commutation rules for the matrices yv, arrange them in such a way that
the two matrices y 4 become adjacent. This may be accomplished with
the aid of the obvious formula y4b = —b*y4, where b* = b and b* =
—b4, which is valid for any real four-vector b (having a real spatial and
an imaginary fourth component). Since y\ = 1, after such a rearrange-
386 QUANTUM ELECTRODYNAMICS

ment of the matrices y4 there will remain in each term a product of not
more than four matrices y v for which the trace can be easily evaluated.
By proceeding in this manner we obtain

(29.12')

By interchanging in this expression y1 and y 2, and also el and e2, we


obtain (?2.
Consequently,
Z2er> dp2dk d(e1—e2—a))
4a>e1e2 J(2n)5 y\y\q^

(29.13)

In contrast to the Compton effeet the momentum of the scattered


electron p 2 and the photon momentum k are here independent. There­
fore the 6-function can be eliminated simply; on noting that dp2
= p2e2de2do2 and dk = (o2dcodo, where do2 and do are elements of solid
angles containing the vectors p2 and k, we have
d(e1—e2—oS)dp2dk -» oo2\p2\e2dodo2doo.
On substituting instead of the primary electron flux density J their
velocity J = = \p± \/el , we obtain the expression for the brems-
strahlung cross section averaged over the polarizations of the parti­
cles (21):
Z 2a3 | p21
O ) 2 IAI

(29.14)

where a = e2/4 n = 1/137.


INTERACTION OF ELECTRONS WITH PHOTONS 387

On taking into account the relations


m2xx = 2(p2k —e2co) = —2a>(e2—p 2cos#2),
m2x2 = —2(pxk —excd) = 2a>(ex—px cos'&1) ,
q2 = m2(x1+ x 2—2)-f2(e1£2—pxp 2).
Pi = e\—m2,
Pl = £\ - m 2,
where and ??2 are the angles between the vectors k and p x, and k
and p 2 we can rewrite do in the form
Z 2a3 p2 dcododo2 e»Pi l 2
do =
(2n)2 px coqi *2 J
- q 2^k, [k ,p 2- p xf (29.15)
or
Z 2a3 p2 da> dodo2 j p\ sin2??2
do (4e\—q2)
(2 tt)2 p x- oo q 4 {e2—p 2cos??2)2
p \ s i n 2$ ! p \ sin2??!+pf sin2??2
■ (4£
22- -q2)+ 2 o i2
(ex- p x cos&xy (£i~ P i cos^j) (e2—p 2cos??2)
p xp2 sin??! sin??2cos<j?
—2 (4ei£2—^2+2co2)[» (29.16)
(£x—p x c o s^ ) (e2—p 2cos??2)
where <p is the angle between the planes {kx, px) and (k 2 , p 2), p — \p\,
q = \q I-
The square of the momentum q2 imparted to the nucleus is related
to the angles ??1; ??2, and <p by the expression
q2 = p\-\-p\JrCo2—2pxoo cos??! + 2p2co cos??2
—2pxp2 (cos??i cos??2+sin??i sin??2 cos<p).
From (29.16) it follows that do vanishes when ??x = ??2 = 0.

29.3. Angular Distribution o f the Radiation in a Coulomb Field


Formula (29.16) determines the angular distribution of brems-
strahlung in the case of unpolarized primary electrons, averaged over the
polarizations of the photon and the scattered electron..
388 QUANTUM ELECTRODYNAMICS

In the general case this angular distribution of the radiation is of


a very complicated nature. Simplifications are possible in the two
limiting cases of low and high energies, which we shall now investigate.
Tn the limiting case of low primary electron energies p1 << m, which
correspond to the problem of the continuous X-ray spectrum, the
photon momentum is small compared to the electron momentum,
since co = ( p\ — pV)l2m < px. Therefore q2 = (px—p 2)2.
Further, on noting that in the nonrelativistic case m2x2 = —m2xx
= 2mco. e x = e 2 = m, and on retaining in the figure brackets of (29.15)
only the first term, we obtain
Z za3 dco p2 dodo2 [k, p x— /?2]2
where Pi < m. (29.17)
° 7 l2 CO P y ( / ? ! — p 2y CO2

Since p 2 differs little from px, the factor


(2m Zaf
da, do2 (29.18)
l/h -P a l4
appearing in the preceding expression is the Rutherford cross section
for elastic scattering of electrons. We can, therefore, say that in the
nonrelativistic case the bremsstrahlung cross section is equal to the
product of the cross section for elastic scattering of electrons das
and the probability of photon emission dwy;
da = dasdwy,
a k l 2 dco
dwy —, t>i —v 2\ ---- do (29.19)
CO J CO

This expression for the emission probability is in agreement with


classical radiation theory, viz., dwy is the ratio of the intensity of dipole
radiation at low frequencies to fico, i.e. the average number of photons
emitted per unit time determined by classical methods (in connection
with this cf. § 30).
The intensity of the radiation has a maximum in the direction per­
pendicular to the plane of the electron motion (p x, p 2).
We now discuss the angular distribution in the case of high energies,
when £1 ^>m, e2^> m. Due to the presence in the denominator of
(29.16) of the factors 1—w2cos$2 and 1—^ c o s ^ the bremsstrahlung
cross section has a sharp maximum in the neighborhood of the direction
INTERACTION OF ELECTRONS WITH PHOTONS 389

of the momentum of the incident electron, and the radiation is concen­


trated primarily within a narrow cone about this direction of angular
aperture of order of magnitude -&1 ~ m/ex. The same cone also contains
the momentum of the scattered electron. In the extreme relativistic
case the angular distribution of the radiation is approximately given
by the following formula (90), (190):
$1dd1
da (tfjJ = A [ l n ( l + t f 2— J + (29.20)
[0? + (w/ex)2]2
where > m and e2 >• m and where A and B are constants.

29.4. Polarization o f the Radiation


We now consider the nature of the polarization of the bremsstrahlung.
We take the electron in the initial state also to be polarized.
By utilizing the general formula (26.18) and the expression for the
bremsstrahlung matrix element (29.9), we obtain for the differential
cross section formulas (29.10), (29.11), in which

F = QfiPi—m) 0 - fays) QfiP a-™),


is defined by formula (29.IT), and the vector a is related to the
electron polarization vector in the initial state \ by expressions (10.30').
On evaluating SpF we can easily show that the terms containing
a vanish. Thus, the bremsstrahlung cross section is independent of
the electron polarization.
In accordance with (26.24) and (29.9) the photon density matrix
is of the form

eap = SP {Q attPi-m ) (1 - i h d Q piiPz-m )}, (29.21)

where
A

i f i —m i f i —m
Qa= m2xL m2x2

and ea{a — 1,2) are two mutually perpendicular unit vectors in the
plane perpendicular to the vector k. We chose them in the manner
390 QUANTUM ELECTRODYNAMICS

On performing the calculations in accordance with formula (29.21),


we obtain the following expressions for the Stokes parameters for the
photon fy (201):

f 3 = ~~?r~Tn
Kxq y I[^ (^£i —^2)(^ 2ei) 2—2 —x2 [ { Pz — P i ) » * ]2}) - 1 ,
t. 2(p2e^)\ 4e2 <72 ^ , 4£le2—q2f _ ^
£ l -- ' T j T T I “ \P2ez)~>T .. \Ple2J
Xi 94G [
2m
£2 — e ^ q t Q

^ lcop1a+E1ka <x>pxa—Exk a \
\ *1 *2 /

+ * + ^ a + a
?c, x, m' \ x, Xo

X \— {oipxa—Exka) + — (o>p2a—E2kd)U , (29.22)


\ x2 ^1 /1
where

(2 =
a>2^4

[ A , ( / 7 2 - / 7 i ) ] 2 ,

^ 1 ^ 2 ^7

= »i = A -
m | Pi |

It follows from (29.22) that independently of the nature of the


polarization of the incident electron the photon turns out to be linearly
polarized (£x and £3 do not depend on 1J). In the case of a polarized
electron the photon also turns out to have circular polarization. Thus,
the determination of the degree of circular polarization of bremsstrah-
lung can serve as a method for the measurement of the degree of electron
polarization. Circular polarization of photons manifests itself in scatter­
ing by polarized electrons (cf. subsection 28.8).
Of greater practical interest is the quantity £2—the degree of
circular polarization of the photon averaged over the final momenta
of the electron ^ but it is not possible to obtain simple expressions for
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 391

this quantity. Fig. 26 shows the dependence of £2 on the energy of the


photon emitted in the direction p x by an electron of kinetic energy

£1—m — 2 MeV (134). We see that near the upper limit of the spectrum
circular polarization becomes almost complete.

29.5 Spectrum of-the Radiation


On integrating (29.16) over the angles we obtain the differential
cross section dom for the radiation in the frequency interval o; to co+^co
independently of the directions of the momenta k and p:
oi
do,.. = o d -
e, —m
da) p 2 j4 p\+ pl Vie2 ■ V2 H VlV2_
1 ~ *Z 9 9 3" I 3
PiP\ Pi PI PIP2
[ _8 E1E2 to2
+L\ T - + ia(« !* l+ ^ I)
3 P 1 P2 p i n
m 2co2 I 8 ^ 2 + p\ ei e2~\~Pl , 2coe1e2
W i­ V2 + - (29.23)
^-PlPi rt Pi p Ip I
where
0 = rl Z 2a ,
L = ln r t+ P iP z-ei™ = 21n ei e2+PiP2—‘
P i- P iP i - ^ i ^ moo
£2^TP2
,h = l n h ± £ L = 2 \ n ei+Pl ?y2 = 21n- and Pi, 2 IPli.i
-P1 m m
392 QUANTUM ELECTRODYNAMICS

At low energies (px < m) this expression is simplified to


co 16 ^ dco m2 Pl+ Pt
a‘ d-T1 ~ 3 co p\ Pl- Pi

= i in l l/ 7 . t A i Z g j l , (29 24)
3 co 7\ co
where 7\ = p\l2m = e1 — m is the initial kinetic energy of the electron.
We see that in the nonrelativistic domain the radiation cross section
is approximately inversely proportional to the frequency. At the highest
possible frequency co = the cross section becomes equal to zero.
As co ->• 0 the intensity of radiation co d a m diverses logarithmically.
This divergence is related to the divergence of the Rutherford cross
section for small angles, and is characteristic of the pure Coulomb
field. As we shall see later, just as in the case of elastic scattering this
divergence can be removed by taking into account the screening of
the nuclear field by the atomic electrons.
In the extreme relativistic domain for ex > m, e2 >> m the radiation
cross section assumes the form

j i 00 .-r dco So I £i , e2 2\(. 2e,£2 1


d<xa = a a d — = 4 0 ----- M—+ — —T3 / \In—^
ex co e1 \e 2 mco 2

_2 I co 2£i 1—(co/ej)
co M 1 V ~ JV e1 - m w
(29.25)
We see that the probability of an electron radiating a definite fraction
of its energy, i.e., the probability of radiation for a given value of the
ratio co/e^ increases approximately linearly with the logarithm of the
ratio ejm .
The radiation intensity, i.e., the product of the frequency and the
cross section, codaw, diverges logarithmically as co -* 0 both in the
relativistic and in the nonrelativistic case.

29.6. Screening
In evaluating the bremsstrahlung cross section we assumed that
the nucleus gives to a pure Coulomb field. Such an assumption is cor­
rect for distances from the nucleus smaller than the radius of the A>shell
of the atom. At greater distances the field of the nucleus is partially
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 393

or completely screened by the field of the atomic electrons, and this


can lead to a decrease in the bremsstrahlung cross section. Therefore
in order to determine the extent to which our results are correct we
have to estimate the range of distances from the nucleus which is signif­
icant for the bremsstrahlung process.
From the form of the integral (26.33) defining the Fourier compo­
nent of the Coulomb potential it follows that the main contributions
to the integral come from distances of the order

9 = \Px~Pt—k \ ■ (29.26)

Indeed, large values of r are unimportant because of the rapidly oscillat­


ing function exp (—iqr), while small values are unimportant because
of the smallness of the corresponding volume dV = Anr2dr. Therefore,
we can assume that R0 determines the order of magnitude of the most
important impact parameters.
On the other hand, as can be seen from formula (29.16), the differ­
ential cross section for bremsstrahlung is particularly large for small
q which, in accordance with (29.26), correspond to large impact param­
eters. Obviously, the screening effect will play no role if the maximum
value of R0 corresponding to the minimum value of q is considerably
smaller than the effective size a of the atom /?0max. a- C>n the other
hand, when the inequality R0max > a holds, complete screening of the
nuclear Coulomb field must occur.
The minimum value of q is, obviously, equal to qmlu = P\—Pi—w,
where px, p 2, and a> are related by j/m 2jrp\ = \/m 2+ pl+ co.
For low energies and low frequencies px << m, a> << px, it follows
from the law of conservation of energy that
--------- ma>
Pz = y P i — 2mco (*f p 1-----— ,
Pi
and therefore
mco moi , ^ „
qmln = m ------ —co ^ ------- , where p1 < m, and a> < pl .
Pi Pi
In the relativistic domain we have
_________________ 9 “(j)
qmin = j / ef—m2— ]/ e\—m2 —--------, where et >■ m and e2 m.
2£j e2
394 QUANTUM ELECTRODYNAMICS

Thus, the maximum value of R0 is determined by

Pi Pi < m , o) < p1,


R,0 m ax
(29.27)
2e1e2
£x > m e2 > m,
m2oj

from which it follows that at sufficiently low frequencies R0max can


exceed the size of the atom. In other words, as co ->■ 0 complete screening
always occurs, and this leads to a considerable decrease in the brems-
strahlung cross section. In particular, when o j -> 0 the quantity co d a ro
vanishes, in contrast to the case of the pure Coulomb field for which
this quantity diverges logarithmically.
In the relativistic domain 7?0niax becomes large even at frequencies
of the order of the energy of the primary electron. Indeed, if co ~ el5
£2~ £ i , ex > m, then /?0max ~ (^/m 2).
If we use the Thomas-Fermi atomic model and take the effective
radius of the atom equal to a — a0Z ~1/3, where

0.529-lO"8 cm
me

is the radius of the hydrogen atom, then for an electron energy exceeding
£0 = 137 (m/Z1/3), the value of R0max will be greater than a. Therefore,
in the domain of high energies screening must always be taken into
account.
In order to take into account the screening of the nuclear field by
the external electrons we must in accordance with (29.10) find the
Fourier component of the total potential due both to the nuclear charge
and to the charge of the electrons. This potential satisfies A A (0e) = —q,
where o is the charge density q = —Zed(r) + en(r) and n(r) is the
electron density in the atom. On going over to Fourier components,
and on utilizing the notation (29.9'), we obtain —q2A [0e)(q) = e[Z—
F(^)], whence it follows that

(29.28)

where F(q) is the atomic form-factor F{q) = Jn(r) e~'ir dr. Thus, we
INTERACTION OF ELECTRONS WITH PHOTONS 395

see that in order to take screening into account we must replace Z


by the quantity Z —F(q). In the Thomas-Fermi model we have

F{q) = Zf(qZ -W ), (29.28')

where f( x ) is a certain universal function which is the same for all atoms,
and whose values may be obtained by means of numerical integration.

Fig. 27.

On introducing the atomic form-factor for the Thomas-Fermi


model we can represent the spectrum of the radiation in the relativistic
domain in the following form (21):

4 lnz]
~ e . e . [ ® , ( C ) - 4 1'> z ]}> <2929>
where
mo)
£ = 200a 100
^Omax e1e2Z 113

and 0 X(C) and 0 2(t) are the two functions of £ shown in Fig. 27. The
quantity £ which is proportional to the ratio a/R0max determines the
effect of screening. Small values of £ correspond to large screening and
large values of £ correspond to small screening. If £ = 0 we can speak
396 QUANTUM ELECTRODYNAMICS

of complete screening. In this case ^(O ) = 4 In 183, <Z>2(0) — 4 In 183—f


and the cross section (29.29) assumes the form
o) da)
= Z zar%
a) + £l — y £i £ 2 In (183Z~1/3) + - ~
-m

^ > 137 Z - 1/3, ey > 137 m Z -1/3 • (29.30)


mo)
Under these conditions the bremsstrahlung cross section for a given
value of the ratio a)/e1 does not depend on the primary electron energy
ev For large values of £ this conclusion becomes incorrect.
For £ >> 1 the screening effect plays no role. The small correction
due to it can be taken into account if we replace in the last bracketed
expression in formula (29.25) the term \ by -J-+c(£), where the values
of c(£) are given in Table 10.

u)/(cr m)
F ig . 28.

Figure 28 shows the dependence of the intensity of bremsstrahlung


on the ratio of the photon energy to the electron kinetic energy ro/(ex—w)
(21). The dotted curves do not take screening into account and are
valid for all the elements (the nuclear charge Ze appears only in the
quantity 0). The numbers beside the curves indicate the initial energies
£x—m in units of m. The solid curves take screening into account, and
refer to lead Z = 82 (with the exception of the solid curve corresponding
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 397

to {ex—m)lm = 0.125 which refers to aluminum). In the region of low


frequencies they approach the curve which refers to the case of complete
screening. The nonrelativistic curve for aluminum has been plotted
taking into account the exact wave functions for the electron in the
continuous spectrum (cf. subsection 29.8). We see that in the case of
light elements screening is not important in the nonrelativistic domain.

T able 10

C 2 2.5 3 4 5 6 8 10 15
c(£) 0.21 0.16 0.13 0.09 0.065 0.05 0.03 0.02 0.01

We note that the use of the Born approximation which we have


employed can lead to an error in the case of large values of Z not only
in the nonrelativistic domain. This is associated with the fact that the
parameter £ = Zez/v is not small in the domain of large values of the
energy and of Z, but is equal to Z j 137. As we shall see later, the deviation
from the Born approximation in the relativistic domain is noticeable
only at large values of q for which screening plays no role. Therefore,
corrections to the bremsstrahlung cross section due to the deviation
from the Born approximation and to the screening effect are simply
additive.

29.7. Radiative Energy Losses


If we know the bremsstrahlung cross section we can obtain the
average energy loss for an electron in passing through a medium.
This energy loss, evaluated per unit path length of the electron, is
obviously equal to
dej
= N o codco,
dx

where N is the number of atoms per unit volume of the medium. If we


define the radiation damping cross section 0 T as
Bi—m
(oo doj. (29.31)
398 Q U A N T U M E L E C T R O D Y N A M IC S

then the energy loss will be given as

(29.32)

By utilizing expression (29.23) for c w e can evaluate 0 r :


^ \ 12e2+4m 2 e1+ pl (8el +6pl)m'1 / e1+ p1\ 2 4
r \ 3elPl ~~ m 3ElPf \ m ) T
2m2 J 2 p 1(el + p1) \ \
(29.33)

where

0«/ ~y
On noting that
F (x )= [n2/6]+[(\nx2)/2 ]-F (\/x )
and that for small x, F(x) = x - ( x 2/4 ) + ( i3/9 ) - ( x 4/16)H----- , we can
obtain 0 T in the two limiting cases of low and high energies:

0 T= ——0 , where ^ < m,

I 2e 1 \- (29M)
0 r = 4 |l n — 1---- —10, where e, > m.
\ m 3/
We see that in the nonrelativistic domain the ratio of the radiated
energy to the initial electron energy does not depend on the initial
energy, while in the relativistic case 0 T increases logarithmically with
ex. However, this result holds only if the screening effect is neglected.
If we take screening into account, then in place of the second formula
(29.34) we obtain the constant cross section
0 T= 0(41n(183Z-1/3) + |) , where et > U l m Z ^ 2. (29.35)
The values of 0 T/ 0 for different energies are given in Table 11.
T able 11

(£ i ~m)lm 0 1 2 5 10 20 50 100 200 1000 CO

&r_ H 20 5 .3 3 5 .5 6 .5 9 .1 1 1 .2 1 2 .9 1 4 .6 1 5 .6 1 6 .4 1 7 .5 1 8 .3

0 Pb 5 .3 3 5 .5 6 .5 8 .7 5 1 0 .3 1 1 .4 1 2 .6 1 3 .6 1 3 .8 1 4 .5 1 5 .2
IN T E R A C T IO N O F E L E C T R O N S W IT H P H O T O N S 399

The energy dependence of the cross section &r taking screening


into account is shown in Fig. 29 on a logarithmic scale. The straight
line corresponds to the logarithmic increase in the cross section in ac­
cordance with the second formula (29.34) with screening not taken
into account.

0 .1 0.5 1 2 5 10 20 5 0 100 500 1000

(cr m)/m

F ig . 29.

The energy lost by the electron per unit path length due to its inelastic
collisions with atoms is determined by the following formula (21):

- f e ) ,= w z l n ^ ’ (2936)
where / is the average ionization potential which is approximately
equal to / — 13.5Z eV. The ratio of radiation losses to collision losses
is approximately given by (21):
1 £i Z (29 36')
^ ( d e jd x ) , ~ 1600 m
Here it is assumed that the logarithmic term in both formulas (29.34)
and (29.36) has approximately the same value. It follows from formula
(29.36') that losses due to collisions and to radiation become equal
at an energy given by e0 = 1600(m /Z). At higher energies the radiation
losses exceed the losses due to collisions (e0 is equal to 10 MeV for
lead, 55 MeV for copper, 200 MeV for air).
400 QUANTUM ELECTRODYNAMICS

29.8. Exact Theory o f Bremsstrahlung in the Nonrelativistic Domain


As we have already noted earlier, the matrix element (27.3) which
determines the bremsstrahlung in the Coulomb field of a nucleus can
be calculated exactly in the nonrelativistic and in the extreme relativistic
domains.
We shall, first of all, show how this calculation is carried out in the
nonrelativistic case (187).
The energy of interaction between the electron and the electro­
magnetic radiation field —a A corresponds in the nonrelativistic ap­
proximation (cf. § 15) to the operator —pA /m , where p is the electron
momentum operator. Therefore, the matrix element (27.3) assumes the
following form in the nonrelativistic domain:

S fl{ = —f
j/2 com J
V* ePe~ihr+imtip1 dr dt.

The exponential factor exp (ikr) appearing in this expression can evi­
dently be replaced by unity. Indeed, in the integral the important
values of r are of the order r ~ v r , where v is the electron velocity
and r is the collision time, which is given in order of magnitude by
t ~ 1 / c o ; therefore kr a > v —- v 1.
Thus,
S™ ,= - 2 n iU i^ ! d(e1- E 2~co), (29.37)
where

= - | 2oj J
f V>*W m
V, (<•) * (29.370

and y)i(r) and ip.fr) are the wave functions for the electron in the initial
and the final states with the time dependent factors omitted (e1 and
e2 are the corresponding energies).
Since the matrix element of the operator for the velocity v = p/m
is related to the matrix element of the radius-vector r by (v)12= i(o(r)12,
can be written in the form
icco r
u i^r = ------ ^ ip*(r)eripfr)dr.
\2(o J
Thus, the problem is reduced to the evaluation of the matrix element
of the coordinate using the continuous spectrum wave functions for
the Coulomb field.
INTERACTION OF ELECTRONS WITH PHOTONS 401

In order to determine completely the continuous spectrum wave


functions it is necessary to specify their asymptotic behavior at infinity,
which may be of two forms, since at infinity the function ip may have
the form either of the sum of a plane wave corresponding to a definite
momentum, and of an outgoing spherical wave, or the form of the sum
of a plane wave and an incoming spherical wave. As has been shown
in subsection 29.1, the wave function describing the initial state of the
electron must have an asymptotic behavior of the first type, while the
wave function of the final state must have an asymptotic behavior of
the second type.
In the case of a Coulomb field the wave functions with such asymp­
totic behavior have the following form in the nonrelativistic approxi­
mation :
V>i(r)= N ^ rF[i$lt 1, / ( / V — Pii*)],
(29.38)
VaW = N 2elP*rF [ — i f a , 1, — i ( p t r + p t r) ] ,

where F is the confluent hypergeometric function, p l and p2 are the


momenta of the electron in the initial and final states, 2 = <^Z/vl 2
(v12 are the electron velocities, a = e2/fic = 1/137), and N2 are
normalizing constants. At large values of r these functions are asymp­
totically given by
»t*ir . A(0) elklT >
y>i(r)ta N x

p— (n /2 )fi / fjff) \

where
r / r\\ fl 2(£ilnsln -§- -^ (l *'fl)
= 2k 1 sm2y ' A i+ ^ i)5

- 2(^2 In cos Y -^ (1
W ) = 0
2k, cos2—

If the constants and N 2 are taken equal to

Nl =
(29.38')
N 2 = c ^ V c i+ if a ) ,
402 QUANTUM ELECTRODYNAMICS

then the wave functions will have unit amplitudes at infinity. In the
future we shall use this normalization.
In accordance with the general rules of § 24 the differential cross
section for the emission of a photon of frequency co and polarization
e within the solid angle doy is given by
e2oj3dojp2
da = y>*rey>1dr doydo2’ (29.39)
2(2ti)4P i
where do2 is the element of solid angle containing the electron momentum
after scattering, p1>2^ |/h ,2|.
The differential bremsstrahlung cross section summed over the two
photon polarizations is given by
e2a>3dojp2 (k D Y \
da = D I2- (29.40)
IQji YP x k2 /
where
D = fy$(r)npx(r)dr. (29.40')

In order to evaluate D we use the following general formula (144):


r eiqr
J= I dre~*T— —F(i£1, 1, i p x r — i p x r ) F ( i £2, 1, i p 2r + i p 2 r )

a d — fiyX
l - / £ i . i'f2, (29.41)
a , a{y + 6)y
where
a = K?2+ * 2)» P = P2<i—&p2,
y = Piq +ftpi—*> <5 = P iP i-P iPz-P-
In order to evaluate D it is, obviously, sufficient to differentiate J with
respect to X and q and then to set X = 0, q = P x ~ p 2:

D=N^ - m ■ (29-42)
q = p i-p t

By utilizing this value of D we can obtain the. following expression


for the differential bremsstrahlung cross section integrated over the
angles (187):
16n2 Z 2a3 1 day ^d
da(i> F(x0) \\ (29.43)
T ~pf ( l - r ^ ^ M
INTERACTION OF ELECTRONS WITH PHOTONS 403

where F(x0) is the hypergeometric function


F(x0) = F(i , ; f 2, 1, x0)
and
_ _ 4£i ^ 2

We investigate the expressions to which this formula reduces for


fi < 1 and f 2 < 1.
By utilizing the relations (cf., for example, (119))

~r—F(i£1, /'£2, 1. x0)= — l + /'l2>2,*o)>

F(a, 6, c, z) = ( \ —z ) - aF \a , c—b, c , - — n

we have

— F ( ^ , /f 8, 1, *0) = - ^ 1^ ( l - x 0)-1- if-JF |l + ^ 1, 1-/£ > , 2 , - — -

If ^ < 1, f 8 < 1, then


F(i£i, i£z, 1, x0) ^ 1,

d F(i£lt i'f2, 1, *0) ^ 1. 2, x0—l



dx, l-x 0
Further, on noting that for 0 < z < 1

F(l, 1, 2, - r ) = - l n ( l + z ) .

and on setting z = x 0/(x0— 1), we obtain

d F(i£„ /f 8, 1, x0) ~ where ^ < 1, f 2 < 1.


dx o
Therefore, we have
d In (1 —x0)
F(x0)|2 = 2F (x „ )^ f
dxo1 2 Xn ’
i.e. ,
d \F(xQ)\* = 2 ^ ln 0 y —
dx o
404 QUANTUM ELECTRODYNAMICS

Thus, for £x < 1, f 2 < 1 the following formula holds:

16 Z 2a3_ 4gt2^ £2 ____ daj I px+ p2


do m (29.44)
3 p\ (e2n^ — 1) (1 —e~2n^) a) \Pi~P2

On replacing in this expression (2?i)2f 1| :2/[exp(27r|:1) —1] [1 —exp


(—2jrf,)] by unity we obtain formula (29.24) for the bremsstrahlung
cross section in the nonrelativistic domain in the Born approximation.
However, formula (29.44) has a broader region of validity than
formula (29.24), viz., it is valid under the single condition << 1,
and insofar as £2 is concerned this quantity may be arbitrary. Indeed,
if << 1, while f 2^ 1, then x 0 = —4 f1/ f 2, i.e., |x0| •< 1, and, therefore

F(i£i, »£8> 1* *o) ^ 1,


d
F(i£x, /f2, 1, x 0)
dx o

Further, on noting that for |x0| < 1 — In (1 —x0) —1, we obtain

4 ~ |F(x0) |2 ^ 2 f1f 2^11-^1— , where < 1, f 2 1.


wA XQ

On substituting this expression into (29.43) we obtain formula (29.44).

29.9. Exact Theory o f Bremsstrahlung in the Extreme Relativistic


Domain
We now show how the matrix element (27.3), which determines the
bremsstrahlung in the extreme relativistic domain, may be evaluated (22).
This calculation is based on using the relativistic wave functions
given in subsection 14.5

W = N elPr | l — ~ a v j w F ( j f , 1, ip r — ip r), (29.45)

where F is the same hypergeometric function which appears in the


nonrelativistic wave function (29.38), u is the normalized spinor amplitude
of the plane wave, a are the Dirac matrices, and N is a normalizing
factor.
INTERACTION OF ELECTRONS WITH PHOTONS 405

The basic property of the functions (29.45) is the fact that if they
are expanded in terms of spherical waves, then for any value of the
energy (not necessarily large) the terms of these expansions corresponding
to the values / > /0, where /0 = Z / 137, practically coincide with the
terms of the expansions of the exact solutions of the Dirac equation
for a Coulomb field. Thus, the functions (29.45) differ from the exact
wave functions for a Coulomb field only for values of / of the order
of / 0.
This property of the functions (29.45) makes them extremely con­
venient for the evaluation of matrix elements of those processes in which
small values of the angular momentum / are not important.
In particular, bremsstrahlung in the nuclear Coulomb field in the
domain of high energies is included among such processes (as well as
pair production by a photon in the field of a nucleus, cf. § 32).
Indeed, for the bremsstrahlung process the important impact param­
eters are of the order of, or greater than, and since the
effective values of / are of the order of magnitude of / ~ b/k where %
is the electron wave length, we have / ~ bp ~ (l/(e/a)m). If, therefore,
the electron energy e is considerably greater than m, then large values
of / will be the important ones and the approximation in which the
functions (29.45) are used as wave functions must yield good results.
Large values of / correspond to scattering through small angles,
which are prominent in bremsstrahlung in the relativistic domain. Thus,
the approximation under consideration must give particularly good
results in the domain of high energies and small scattering angles.
The matrix element determining bremsstrahlung (omitting the time
factors) can be written as

where and xp2 are the electron wave functions in the initial and final
states, for which we take the functions (29.45); here, as has been explained
in subsection 29.1, the wave function for the initial state must for r -+ oo
go over asymptotically into the sum of a plane wave and an outgoing
spherical wave, while the wave function for the final state must go
over into the sum of a plane v/ave and an incoming wave. The functions
yjj' and y>2 are assumed to be normalized to unit amplitude at infinity.
406 QUANTUM ELECTRODYNAMICS

In such a case Nx = e 2 7X1 — ifj) and N2 = e 2 7^(1 + /£2). Thus


we have

M S l,= - - ^ A W
y2u>

X J e -'P S [u*F? + -a* VF*u*


]

X (ea) e~ikr el^ r|^u17q — a VFi iq j dr

InieN xN *
[{u*a.eux)Ix
]/2 o)
+ (u* (ea) ( a /2) wj + (u* (<*73) (ae) w j], (29.46)
where
7*1 = FO'ix, 1, i p s - i p s ) ,
F 2 = F ( — i£ 2 , 1, — i p 2 r — i p 2r ),

J1 = J e 'v F f F x dr.

h = - ~ J eiqrF * ’V F1dr,

/ 3 = — J elqr(¥F*)Fx dr,

<1= P i - P i - k .
It might appear that in the high energy region, which we are inves­
tigating, there is no necessity to retain in functions of the form (29.45)
the second term, which is inversely proportional to e. However, this
is not so, because the factor u^eau^ arising from the first main terms
of (29.45) physically corresponds to the component of the electron
velocity parallel to the photon polarization, i.e., perpendicular to its
momentum, which in the relativistic domain is extremely small. For
this reason the factor w^eazq is of order of magnitude of ?n/e, while
at the same time the factor M2* (co)a«1 is of order of magnitude unity.
As a result, both terms of (29.45) lead to quantities of the same order
of magnitude in the matrix element.
INTERACTION OF ELECTRONS WITH PHOTONS 407

The evaluation of the integrals appearing in (29.46) can be carried


out with the aid of formula (29.41). Indeed,

' ' - - ( S L

h = A - ., (29.47)

h A=0•

The differential bremsstrahlung cross section summed over the


photon polarizations v is related to M ^ f by the expression
dp2dk
da = 2 * i ^ IM " f 5(ea—e2— (d).
w
We do not reproduce here the detailed calculations, but state only
the final result (22)
V2(x) + Z 2a2y 2 fV2(x)
da = daQ (29.48)
VH\)
where da0 is the bremsstrahlung cross section evaluated in the Born
approximation, and given by formula (29.16),
V(x) = F (—i£, if, 1, *)
f 2(l + f 2) 2 f 2(l + f 2)(22+ f 2)
= 1+
(2!)2 * + (3 !)2
sinh?rf
V{\) =

1 dV{x) _ , , 1+ f 2 (l + f 2)(22+ ^ ) v2 ,
W(x) =
f 2 dx + 1!2! 2! 3! X + ‘'
_ j_ _ o)2(e2— p 2 cosd 2)(E1—p 1 cosOj)
X~ y’ y~ C iW 2
and 0lt 62 are the angles between k and p x, p 2.
Over a considerable range of important angles the differential
bremsstrahlung cross section (29.48) practically coincides with da0.
This is associated with the fact that the scattered waves in the initial
and the final electron states in practice overlap very little.
408 QUANTUM ELECTRODYNAMICS

The bremsstrahlung cross section integrated over the angles has


the following form (145):

where da0(a>) is the cross section evaluated in the Born approximation


taking screening into account, and
CO

(29.49')

Thus, the bremsstrahlung cross section is always less than the value
obtained in the Born approximation. Insofar as the form of the brems­
strahlung spectrum is concerned, it is essentially unaltered when the
correction to the Bom approximation is taken into account.
We note that for the evaluation of the bremsstrahlung cross section
integrated over all the directions of the scattered electron we need
not necessarily use for the final state the electron wave functions which
have the asymptotic behavior (29.8); we can equally well use the functions
with the asymptotic behavior (29.7) since such a replacement merely
indicates a different choice of the complete system of functions describ­
ing the particle. This enables us to relate bremsstrahlung cross section
integrated over the angles to the cross section for pair production
(cf. subsection 32.2).

29.10. Radiation Emitted in Electron-Electron and Electron-Positron


Collisions
We now investigate the process of photon emission in electron-
electron collisions.
In contrast to the problem of bremsstrahlung in the collision of
an electron with a nucleus, we cannot in this case replace the effect
of one of the particles by an external field. Therefore, we have to
investigate the elements of the scattering matrix which connect one
photon state and four electron states (two initial and two final states).
Such elements are contained in the third order scattering matrix S(3).
The eight diagrams corresponding to them are shown in Fig. 30. In
these diagrams p l and p[ denote the momenta of the initial states,
p 2 and p'2 the momenta of the final states, k is the momentum of the
INTERACTION OF ELECTRONS WITH PHOTONS 409

emitted photon, f s are the momenta of the virtual electrons, and qs


are the momenta of the virtual photons (s = 1, 2, . . 8). The last
four diagrams (5, 6, 7, 8) differ from the first four by the replacement
of p 2 by p 2 and of p 2 by p 2. We shall refer to these diagrams as ex­
change diagrams (the terms corresponding to these diagrams appear
in the matrix element with opposite sign).
ik \k
i
p2 1 /; P2 f2 I Py Pi Py
* i* ^ 1*
j?i |9 2
-- 1 - J-r i ■,
P'l P\ P'l P
Jk
P’y
(D (2) (3)

I* !*
P2 0, P2 ! h Py P'l ^6 ! Py
«------- r<------- * 1^
|?e
Pi U p\- Pi L5 p\ A 1g Py
i
(4) !* (5) (6)

04 . Py
p'l
■*------- r*— M- P'l
!?0 ,
p, r, i? ' Pi !, '8 Py
* -------- r*-----------^ ---- i
1 i
\k (8) ik
<
1 (7)
Fig. 30.

In diagrams 1 and 3 scattering occurs first, followed by emission


of radiation, while in diagrams 2 and 4 conversely emission of radiation
occurs first, followed by scattering. In diagrams 1 and 2 the radiation
is emitted by the first electron, while in diagrams 3 and 4 it is emitted
by the second electron.
By moving along the electron lines of the diagram from left towards
right, and by replacing the lines and the vertices by the amplitudes
410 QUANTUM ELECTRODYNAMICS

and the matrices corresponding to them, we obtain in accordance with


Feynman’s rules the following expression for the matrix element:

In each term the first bracketed expression corresponds to the upper


electron line, and the second to the lower one. The values of f s and qs
are determined by the conservation laws at each vertex of the diagram:
*7l — #2 = P 2 P li $ 3 = #4 ~ P 2 P \t fl P 2 ^ >

f2 = P l~ k > f3 = P2+ k ’ f i = P'l — k -


The expression for the effective cross section corresponding to the
matrix element (29.50) turns out to be very awkward, since after sum­
mation over the polarizations of the electrons and of the photon a large
number of terms appears containing traces of products of six y ma­
trices.
We shall give here only the results referring to two limiting cases:
the nonrelativistic case | p 1 | < m, \ p[ | < m and the extreme relativ­
istic case | p x |
Since the dipole moment of a system of two electrons is equal to
zero, the radiation accompanying the collision of two nonrelativistic
electrons is in the first approximation quadrupole radiation. The
radiation cross section can therefore be evaluated not only by means
of (29.50), but also in accordance with formula (27.14). In doing this
we have to use the fact that in the nonrelativistic approximation the
system of two particles may be replaced by a single particle moving
in an external field, the Coulomb field describing the interaction between
INTERACTION OF ELECTRONS WITH PHOTONS 411

the two electrons. For the wave functions of the initial and the final
states we can take expression (29.6') in which y)q has the form

_ i _ ( ei<7r_j_ e-l
% ]/2
where r is the relative radius-vector of the particles, and q is the mo­
mentum of their relative motion (the plus sign corresponds to total
spin equal to zero, and the minus sign corresponds to spin unity).
We note that in § 27 in investigating quadrupole radiation we had
assumed that cor << 1, where the magnitude of r defined the distances
important for the problem. In the case of bound states r has the meaning
of atomic or nuclear dimensions, while in the present case of radiation
emitted in a collision r ~ vr, where v is the electron velocity and r is
the “collision time” which is of the order of magnitude of r ~ l / c o .
The expression for the differential cross section for the emission
of radiation has the following form in the nonrelativistic case (125), (58):

a r o | - ^ r ( 4 CPx ~ p D + 3 [ P 1 P 2 ? )

f (144 [/?! p2]6+ 2 6 4 ( p l - p l ) 2 [p ! p2]4+ 1 0 5 ( p \ - p i)* [ p ^ ] 2

+ l 2 ( p l - p i Y ) - - ^ 4— (36[p1p2]4+ 3 9 ( p ? - ^ ) 2[p1p2]2

dq
+ 12 {p \~ p l)\ (29.51)
Pi \(p \ - p 1)
where
q = P i—p ^ P — Pi+Pz-
The angular distribution of the radiation in the center of mass
system of the electrons is determined by its quadrupole nature, i.e.,
do ~ \ F2<°)I2.
Integration of the cross section (29.51) over the angles, which must
be carried out over a hemisphere, since the particles are identical, yields
3x2 1 2 (2 -x )4- 7 ( 2 - x ) 2x 2- 3 x i
do(a>) = T7 \2 +
15 17 (2 -x ) (2—x)3 \ / l —x
| / l —x
dx, (29.52)
x
412 QUANTUM ELECTRODYNAMICS

where x = co/ e 1 , co is the energy of the emitted photon given by


co = e1—e2 — and e 2 and e 2 are the initial and final
electron energies in the center of mass system.
The cross section for the radiation energy loss in electron-electron
collisions is given by the formula

8ar02. (29.53)
£i
It is of the same order of magnitude as the cross section for the radia­
tion energy loss in the field of a nucleus for Z = 1 (cf. (29.34)). From
this it follows that the cross section for the radiation energy loss in
the collision of an electron with atomic shell electrons is approximately
equal to 8a/-oZ where Z is the number of electrons. In order of magni­
tude this cross section is smaller by a factor Z than the cross section
for the radiation energy loss in the field of a nucleus, which is
proportional to Z 2.
In the extreme relativistic case the cross section for the emission
of a photon in the collision of a fast electron with an electron at rest
is determined by (72)

*(«,)= 2 ^ ^ 2 + fi±f! ( z l n ^ f i - l
E2 CO (\ 3 £l e2 / \ (om I

co-
- 1 .7 0 2.14 • (29.54)
PlP-2
In this case the cross section for the energy loss by radiation is
given by
2e1
0 = 4a/'ol In 1.23 . (29.55)
m

We see that the cross section for the emission of radiation coincides
up to a factor in the argument of the logarithmic term with the cross
section for the emission of radiation by an electron in the field of a nu­
cleus with Z = 1. Such a result can be easily understood on the basis
of the following considerations. As we have seen (cf. (29.27)) at high
energies the effective momentum transfer to the nucleus becomes very
small, and, therefore, the value of the nuclear mass becomes less and
less important.
INTERACTION OF ELECTRONS WITH PHOTONS 413

We now consider the radiation emitted in the collision of an electron


with a positron. In the nonrelativistic case this radiation, in contrast
to the radiation emitted in electron-electron or positron-positron
collisions, is of dipole nature.
If one of the particles is at rest before the collision, the radiation
cross section is equal to

(29.56)

In the c. m. system the radiation cross section coincides with the


radiation cross section in the field of a nucleus with Z = 1 in the non­
relativistic approximation.
The cross section for the emission of a photon in a collision between
relativistic electrons and positrons does not differ from the cross section
for the emission of radiation in a collision of two relativistic electrons.

§ 30. Emission of Photons of Long Wavelength

30.1. “The Infrared Catastrophe''


In the preceding section we have seen that the probability of photon
emission in the low energy range is inversely proportional to the frequen­
cy dw ~ (dco/cd), while the total radiation probability diverges loga­
rithmically as co -> 0.
This divergence in the region of low photon energies is referred
to as the “infrared catastrophe”. It is brought about by the unjustified
application of ordinary perturbation theory, based on the expansion
of the scattering matrix into a series in powers of e, to those processes
in which long wavelength photons participate. Indeed, it can be easily
shown that if the probability vv2 for the emission of one long wavelength
photon is proportional to e2 In (e/co) where e is an energy of the order
of magnitude of the electron energy, then the probability of emission
of two photons w2 will be proportional to (e2 In (e/co))2. Therefore, the
ratio of the probabilities is given in order magnitude by
e
where oj -+ 0 . (30.1)
0)
414 Q U A N T U M E L E C T R O D Y N A M IC S

It is this ratio, and not the quantity e2 as we have been assuming until
now, which provides the perturbation theory expansion parameter
applicable to the processes of interaction between the electron and
long wavelength photons. Since when co -+ 0, £ is not small compared
to unity, perturbation theory is, strictly speaking, inapplicable to such
cases.
The inapplicability of the usual perturbation theory is associated
with the fact that the number of photons emitted by an electron per
unit energy tends to infinity as co -*■ 0, while in perturbation theory
it is assumed that the emission of one photon is always more probable
than the emission of two or a greater number of photons.
In order to verify this we note that if the energy and the momen­
tum of the photon are considerably smaller than the kinetic energy
and the change in the momentum of the electron, and the photon
wavelength is considerably greater than the classical electron “ radius”,
we may regard the motion of the electron as given and use classical
electrodynamics. Assuming for the sake of simplicity that the electron
velocity is small compared to the velocity of light, we can start with
the following formula1 for the intensity of dipole radiation d£a
in the frequency interval dco (118)

d £ a, = ~ | < / J 2</co,

where dmis the Fourier component of the second derivative with respect
to time of the dipole moment
1 °°

— CO

If co —►0, then

(^to)u)->0 -- 2jj
{dl and d2 are the values of the dipole moment before and after emission).
In the case of interest to us d = ev and therefore

( < 0 ^ 0 = (e/2jr)(v2—i^),

1 B o th here and in subsequent fo r m u la s we e x p lic itly in tr o d u c e th e con stan ts


c and h.
INTERACTION OF ELECTRONS WITH PHOTONS 415

where v x and v 2 are the electron velocities before and after emission.
Thus we have

~ r ( v 2 - v i)2 do)• (30.2)

We see that the radiation intensity per unit frequency interval d£Jdoy
tends to a finite limit different from zero as a> ~» 0. From this it follows
that the average number of photons emitted which is equal to (\/hoy)
(dEjdoy) tends to infinity as co -> 0, as has been stated earlier.
Since the probability of transition of an electron from a state of
momentum p x to a state of momentum p 2 is always finite, the prob­
ability of the simultaneous emission of an infinite number of photons
of vanishingly low frequency is also finite, and different from zero.
Therefore, the probability of emission of a single photon or a finite
number of photons with co -» 0 is in fact equal to zero, and is not in­
finite as given by perturbation theory.
The quantity (l/hoy)(de(u/doy) is the average number of photons
of frequency co emitted by the electron in the frequency interval day.
We now obtain the probability of the electron emitting an arbitrary
number n of long wavelength photons whose frequency lies within
the interval oyx ^ co ^ co2.
On postulating, as before, that
hoy hoy
—A~ (30.3)
e cAp < 1’ me° - T < 1-

where e is the electron energy, Ap is the change in its momentum, %is


the photon wavelength and r0 is the classical electron “radius” we can
assume that the emission of photons does not affect the electron mo­
tion, i.e., we may regard the electron motion as given. Under these condi­
tions the processes of succesive emission of photons will be statistically
independent and, consequently, the probability of photon emission
will be given by the Poisson formula:

w(ri) = (h)n (30-4)

where n is the average number of emitted photons whose frequency


lies within the given integral cox ^ co ^ co2.
416 QUANTUM ELECTRODYNAMICS

If conditions (30.3) are satisfied the quantity n can be determined


by means of classical electrodynamics. Indeed, if dJ — Idcodo is the
classical radiation intensity in the frequency interval (co, co+dco) and
in the solid angle do, then we have

-J-dcodo. (30.4')
non
Q}= 0)i

We now demonstrate how / can be determined.


It is well known that dJ = c\Ha)\2R zdcodo, where Hm is the Fourier
component of frequency co of the magnetic field H at the point R; the
quantity H m is related to the Fourier component of the vector potential
Aw by the relation Hm= i[k A ^. By using the Lienard-Wiechert po­
tentials (118) we can easily show that the Fourier components of the
vector and the scalar potentials are given by
oo oo

A“ - I jic R J ^( 0 * dU 2n R J 6 dt’
— oo — OO

where r — r(t) is the equation of the electron trajectory and v(t) is


the electron velocity. On noting that k A O)—(l/c)axpa)= 0 , we obtain
d J = c\H (0\2R2dcodo = cR2k2Q A J 2-\(p m\2)d(odo
OO OO

-^ -r(r)r(0 -l J dtdt’k 2da)do. (30.5)


— o o — OO

If the frequency co satisfies the condition cor << 1, where r determines


the order of magnitude of the time during which the electron scattering
takes place, then in carrying out the integration in (30.5) we can assume
that r{t) pa v xt+ a , where —oo ^ t ^ 0 and r(t) an v 2t Jr b, where
O ^ t ^ o o , (vx and v 2 are the electron velocities before and after
emission, a and b are constants). On substituting these values of r(t)
into (30.5) we obtain
rfr- «■ / / « . _______
4tt2c3 (\ 1—(n v jc ) 1—(n v jc ) /

- (\1
- ( n v j c) ) l - ( n} v j c ) l \ dco do (30-6)

(n is a unit vector in the direction of emission).


INTERACTION OF ELECTRONS WITH PHOTONS 417

If v l and v 2 are considerably smaller than the velocity of light,


then formula (30.6) reduses to formula (30.2).
We now determine the probability of the scattering of an electron
by an external field accompanied by the emission of n long wavelength
photons. On assuming that the conditions (30.3) are satisfied, which
means that the radiation reaction on the electron motion is extremely
small, we can express the probability on the electron being scattered
and emitting n long wavelength photons in the form

dw = w(ri)dws, (30.7)

where dws is the probability for the elastic scattering of an electron


by an external field into the solid angle dos which was obtained in § 14.
For the frequency interval (o^, co2) we must obviously take the interval
containing zero frequency. The average number of photons emitted into
such an interval becomes infinite in accordance with (30.4'), and therefore
w(«) = 0. In other words, the probability of the electron being scattered
and emitted a finite number of long wavelength photons is equal to
zero. In particular, the probability of purely elastic electron scattering
CO
is equal to zero. Since, on the other hand, lim JF w(ri) = 1, it follows
. n —» o o n = 0
that
oo

£ dw = dws. (30.7')
n=0

Therefore, we can say that the probability dws for the elastic scattering
of an electron obtained in § 26, which does not take into account the
interaction between the electron and the radiation field, is the total
probability for electron scattering to occur, independently of the number
of long wavelength photons emitted by the electron. As we have shown,
this probability can be determined in the case of a sufficiently weak
external field by perturbation theory, in contrast to the probability of
scattering accompanied by the emission of a single photon or of any
finite number of photons of frequency co -» 0: according to perturbation
theory this probability is infinite, whereas in fact it is equal to
zero.
The average energy of the long wavelength photons emitted in the
frequency interval (pd,o)-\-doS) and solid angle do, with the electron
418 QUANTUM ELECTRODYNAMICS

being scattered into the solid angle dos, can in accordance with (30.4),
be given as

n= 0

Since in this case n = dJ/tioi, we have


de = dJ dw, . (30.8)

As should have been expected, this quantity is equal to the product of


the total probability dws of the electron being scattered into the solid
angle dos, and the classical expression dJ for the energy radiated in
this process. The usual perturbation theory leads to the same result in
the low frequency region. Thus, we see that the average radiated energy
and the total probability of electron scattering (independently of the
number of long wavelength photons emitted) can be determined by
means of the usual perturbation theory in the domains of both high
and low frequencies, although the probability of electron scattering
accompanied by the emission of a finite number of photons of o j -*• 0
is not given correctly by perturbation theory.
The probability of emission of photons whose energy is not vanish­
ingly small may be obtained by means of the usual perturbation theory,
i.e., by the'expansion of the matrix elements into a power series in the
electron charge, but the probability of emission of a single photon
determined in this manner should be interpreted not as the probability
of just that one photon being emitted, but as the probability of the
emission simultaneously with this photon of an arbitrary number of
long wavelength photons of frequency o j -* 0.

30.2. Investigation o f the Divergence in the Low Frequency Domain


by means o f the Scattering Matrix
We now show that the Poisson distribution for the probability of
emission of long wavelength photons can be obtained with the aid of
the scattering matrix.
We assume that the conditions (30.3) are fulfilled, which means
that the emission of radiation has only a small effect on the electron
motion, so that the latter may be regarded as given. Under these
conditions the current density can be regarded not as an operator, but
INTERACTION OF ELECTRONS WITH PHOTONS 419

as a c-number, and the expression (24.27) for the scattering matrix


can be used.
On expanding the current density into a Fourier integral

h (*) = ( 2 ^ ' / j ^ ky kI d*k » where J W = U ~k)

and on utilizing the expansion for the operators of the potentials of


the electromagnetic field into plane waves we can express the scattering
matrix S in the form

S = exp f j* {k )D \k )jv{k)d^k
2(2 nY

x N exp { p r / 2k,X v*(k)^ ~Ju(k)c tx W


y2co
where
i
Dc(k)

We now consider the integral appearing in the first exponential:

K = - 2 fj:(.k)D '(k)j,.(k)d ‘k = - - j

By utilizing the rule for going around the poles in the integrand of this
expression, i.e., by taking for the poles the points k0= ± \ / k * - i O ,
we obtain as a result of integrating over /c0
dk
K = n f j*(k)jr(k)
2m ’

where a> = \k\ and j v(k) = j v(k, (d).


Thus, the S-matrix can be written in the form

S = AN exp j — -Jj? — =- [y*WcfcA+A(/r)c+Je;vJ, (30.9)

where
dk
A = exp
( (2 71)4 T exp 2(o
We now evaluate the element of the S-matrix between the vacuum
state and the state which contains n photons of different kinds
420 QUANTUM ELECTRODYNAMICS

n = £ n k, where nk is the number of photons of momentum k . In


£
order to do this we investigate the «th term in the expansion of the
S-matrix

S'"’= A N \ ^ \ W W e “+^ )c“)"'.


and obtain the matrix element

M ™ ,= (n |S (n)|0 ) = A^n 0 . (30.10)

We use the formula

( B i+ B z + B .^ -------- 1- B r)n — ^ B*2 ... B / — --j


-r»/ = n 1 2 T
ni + n2+ ,“ + nr n
and in it set

Bt = r = k ,X , n'T= n'kX.
]/2 oj

We are interested in the final state containing nT= nkX photons of the
rth kind. Therefore, the contribution to introduces only one term
into this formula in which n[ = n1} n'2= n2, ..., n\ = nt = nkx, i.e.,

On noting that in accordance with (16.18) (nw!(cj^)n u |0) = ]/nu !,


we obtain finally

n kX
M}*\ = A (30.10')

We now obtain the probability of emission of n = £ n k photons


k
of different momenta. This probability, which we denote by Pn, is
evidently equal to the square of the absolute value of summed
INTERACTION OF ELECTRONS WITH PHOTONS 421

over A and k , with the sum £ nkX being equal to a given number
a=i
nk. Since j fieAfi= j x, we have

jP_
P„ = i nk
n k l + n k Z= n k «*2!

^ ril 1 (Ui(*)[, + Ij.(*)I,)"‘


a* 1 1 I nk\
The sum of the squares of the absolute values of the transverse
components of the current density \ji(k)\2-P\jz(k)\2 is equal in virtue of
the equation of continuity k j^(K) = 0 to the square of the absolute
value of the current density four-vector

ia (* )i!+ ia ( * ) i! = u ( * > r - l - i c - m != \j m ‘

Further, on going over in the expression for K from integration


to summation over k in accordance with the formula

we obtain

and therefore Pn can finally be written in the form


/U A)J!\"*
L/.w n n ' 2n)fl ' (nk)nk
Pn = [J CXP
k
2ojQ nk\
(30.11)

where
- _ \u w
k 2coQ ■
Thus, we have obtained the Poisson distribution for the probability
of emission of long wavelength photons. It is clear that nk is the average
number of photons of momentum k, since according to (30.11)

h nkPn = H -
nk
422 QUANTUM ELECTRODYNAMICS

It now remains to be shown that the resultant expression for the


average number of photons emitted

(30.12)

coincides with the expression for n which was found earlier by the
classical method (formula (30.4')).
In order to do this we determine j (k) by assuming that the current
density has the following components

j^ x ) = \evd(r—r(t)) and r(/))),

where r = r(f) defines the trajectory of the moving particle


oo oo

jfi) = » (/) e'(- ‘- krU,) dl, ie J


— CO — CO

From this it follows that

\ j li( k ) \2 = e 2 j J
—oo —oo

The average number of photons emitted of frequency lying in the


range o>1 ^ co ^ o>2 is, according to (30.12), given by
c o - co2

- f M W dk
J 2 (o ( 2 t i) 3
10 = COj

k2 ^ \ d o d m
(2(27if (0 J

(30.13)

This expression coincides with the classical formula (if we take into
account the fact that in (30.13) the charge is expressed in Heaviside
units).
INTERACTION OF ELECTRONS WITH PHOTONS 423

30.3. Relation between the Photon “Mass" and the Minimum Frequency
Since perturbation theory gives an incorrect value for the proba­
bilities of the different processes of interaction between the electron
and the long wavelength photons, we shall in the future, in using perturba­
tion theory, segregate the low frequency region, i.e., we shall assume that
the photon frequency exceeds a certain minimum value a>mln. This
quantity must later be eliminated by a separate investigation of the
interaction between an electron and the long wavelength photons
(cf., for example, the problem of the radiation corrections to electron
scattering, and the radiation shift of atomic levels in Chapter VIII).
However, in practice, it is more convenient to use a somewhat
different condition, which is equivalent to the condition a> > con)in,
viz., without imposing any restrictions on the photon frequency, to
assume that the photon has a certain very small mass Awhich is different
from zero. The introduction of this mass, which' is a relativistically
invariant quantity, in contrast to the noninvariant quantity comln,
simplifies the calculations considerably.
We now show how to obtain the relation between the photon “mass”
A and the minimum frequency wmin.
We consider the emission of a photon by an electron in some con­
stant external field Ajf^x). According to (29.9) the matrix element which
determines the emission of a photon of energy k 0, and with its polar­
ization vector directed along the ^-axis, is given by
y,< i( p 2+ k ' ) - m * i ( P i - k ) - m ytl
S<2> = ie2u, A 'e\q ) + AM{q)
|/2 F0 (P2+ k)2jrm 2 (Pi~k)2-\ m- y 2k0
(30.14)
where p t and p2 are the four-momenta of the electron before and after
scattering, and k is the photon four-momentum. Since we are interested
in the low frequency region we can neglect the quantity k in the numera­
tors. Further, on introducing the photon “mass” I by means of the
relation k2-fA2 = 0, we rewrite (31.14) in the form

where
pk = p k ~ e k 0, e = ]//>2+ m 2, k 0 — j//c2+A2.
424 QUANTUM ELECTRODYNAMICS

Since {ip i+ m )^ = 0, we have

M (e,(« ')('P i-w0 V /i = = 2iu2A 'e)(q)ulPl/i.


Similarly we have
u2yJ<iP2 -™ )A {eKq)ui = 2iu2A {e\q )u 1p 2/i.
Therefore
Pi u p (30.15)
S& f = (ut A {eKq)uM
V 2ko (X2j l ) —p 2k (X2/2)+ p1k

On assuming that e > A we may neglect X2, since p k contains the term
e \/k 2+X2 which is considerably larger than A2.
We now obtain the cross section for electron scattering accompanied
by the emission of a photon of energy not exceeding As. With respect
to the magnitude of Ae we assume that it is considerably smaller than
the electron energy e and considerably greater than the photon mass A:
A •< Ae << e.
The differential cross section for such scattering, summed and
averaged over the electron spin components in the initial and the
final states, is given by
2 n e4 B
da' = W2^(e)(?)Wll Gfdo,
v ~(2nf
where

B= Pip \ dk
(30.16)
Pik l K ’

do is the element of solid angle containing the electron momentum


p 2 after scattering, v is the initial electron velocity and

_ p\ dpi Pi £
1 (2n)2de (2tt)3
is the number of final states of the electron per unit energy, and per
unit solid angle.
In the case of the Coulomb field of a nucleus
Ze
A (e)(q )= ~ i y 4 2
\ q \
INTERACTION OF ELECTRONS WITH PHOTONS 425

and

£ ! M <e)(?)«il2= 2 |— - j | l —uzsin2y j = 87ra— | l —u2 sin2^

where $ is the scattering angle.


In order to establish the connection in which we are interested
between the photon mass X and the minimum frequency comln we evaluate
the quantity B appearing in (30.16) in two different ways: first by assum­
ing that X =£ 0 and that k 0 varies from k 0 = X to k 0 = As, and then
by assuming that X = 0 and that k 0 varies from k0 = comin to k0 = Ae.
By comparing the two expressions for B, which we shall in future denote
by Bx and Butmin, we obtain the relation between X and comln.
On noting that pt^A-m2 = 0 we obtain

|fc| r* k2d\k\ doy ( m2 m2


' J y ^ - 2 2 \ ( e 2 ? ^ + r 2- / 7 ^ ) 2 + (£l
1*1=0
-]/i^+X2—p 1k f

____ _J-PiP*________ ____


(ea ]//<2+ 2 2—p 2k){E1\/k 2+X2—pxk)
(30.17)
\k\=Ae
m m-
B" m i n = — f
I
\k\d\k\do
i I I I y
(e21k I—P*kf (ei I k I—Pikf
I =<um i n

i_________ 2 pifo________
+ (e21k I~ P i k)(ei I k I - P i k)

where doy is the element of solid angle containing the photon momen­
tum k.
We now use the identity

1 1 r ______dz ______
(e2\/'k2+X2--p 2k) (fxj / F + 2 2- P ik ) 2^ k 2+X2- p , k ) 2 ’

where
pz = A ( lA - z ) P i+ i( l- z ) p 2, + i ( l - z ) e 2.
426 QUANTUM ELECTRODYNAMICS

Then
\ k \ = At
k 2d \A:| doyj m2 nv
Bx= ~ \ /k 2+X2 ((«!]/k2+X2—p 1k)2 (e2)/k 2+X2—p 2k)2
l*l = o
dz
+P 1‘P‘
P2 I
(e* j / k2JrX2—p zk)2
-1

Since
do,, An
h (e\/k2+X2- p k ) 2 k 2{e2- p 2)+ X 2e2

we have
As
k2d\k\ \ m2
B = ~ 4™f 1+ Vi
| / k 2+X2 \ k 2{e2~ p 2+X2e2 k2(e2
2- p 2)+X2e2

+PiP, I (3018)

It can be easily shown that

k2d | k | 1 2zle e 1 £+\p\


I n - ------ --—- - ------In
. ( [k2{e2- p 2)+X2B2]\/k 2+X2 e2- P 2 a 2 1/71 e2—p 2 e ~ \ p \ ‘

Therefore we have

Bx = 4* { - [ 2 + 0 . , / , , - W / — ] In — + - [ r ^ | In £1+ 1Z7! 1
ei I/^l |

| e2 Jn £2+|P2l dz £*_ in £*+ l ^ ]}•


■(/»!/»*- f i l ^ j (30.19)
1^21 fi2 IP2 1 -1 £\ - p \ \Pz\ ez ~ \ p .

On introducing in place of z the new variable of integration £:


C= - = j/cos2($/2)+ z2sin2(#/2), we write Bx in the form
Vez
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 427

Bx = 4n J 2(20 coth 2 0 - 1 ) In — + i l n ^
I A V 1— V

------cosh 2 0 f In --------------—--------------1 n n 20)


•ysin($/2) J 1-vC ( i_ j;2C2 i/C2- c o s 2(tf/2) J (-30-20)
cos(5/2) ' v ' ’’

where the quantity 0 is related to q by the expression q2 = 4m2 sinh20 .


Since qi = 0, we have q2 = q2 = 4p2sin2 (&/2) and

sinh0 = sin(ff/2).
]/1 —v2
On setting A = 0 in (30.18) and on integrating over j A:[ from
l^l1= comin t0 \k\ = Ae we obtain Bm
<umin :
1 Ac

= ~ 4* (30.21)
-1 “ m in

Since
f 40
J e\ —pi m2 sinh 2$

and PiP2—Sie2 = —m2 cosh 20, we obtain finally


/is
Bm — %n{20 coth 2 0 —1) In----- . (30.22)
^ m i n

J n comparing (30.20) and (30.22) we obtain the desired relation between A


and comin:

2 ( 1 - 2 0 coth 20) In
A

i , \+ v (i - v 2) u<vf. r i+ < #
= — In ---------------;— 7qTt\ COsh ln 1 > 7 — — — •
« 1-® ®sin (0/2) J 1 VC(1 —v2C2)]/£2—cos2(ft 12)
cos(9/2) V
(30.23)
In the limiting case of low electron energies v < 1 this formula leads to
In 2a>mln = ln A + |. (30.24)
428 QUANTUM ELECTRODYNAMICS

On substituting (30.20) into (30.16) we obtain the following ex­


pression for the differential cross section for electron scattering in the
nuclear Coulomb field accompanied by the emission of a photon whose
energy does not exceed Ae(Ae<^ e ):

Za
da' = ) (1—v2) ( l —v2 sin2—
{2mv2sin2(0/2)
l+ o
X — {2(20 c o th 2 0 —l ) l n ^ ^ - + - l n
71 A o l-o

— •—7 r7 ^ -c o sh 2 0 C (v + ) (30.25)
wsin (0/2) J
where
L
l + < _________ dC_________
G (v,d) In (30.25')
- / & 1 - < ( 1 - C 2( 2) j/C2^ cos2(0/2)
cos—.

We now also evaluate the quantity Bx in the case when p1 = 0.


On setting e2= m c o s h 2 y , p 2 = m s\nh2y, and on noting that
e 2— \pz\2 = cosh2^ —z2 sinh2_y, we obtain

J
-1
m2 sinh2_y

Further, we have

.£± l n i l ± £ L = 2 , ‘i - \ n ‘t+ P * = 4 y ™ i h 2 y ,
Pi El Pi P2 e2 P’1
ez — cosh2_y—z sinh2^, \p2\ = (1—z) sinh_y cosh^
and

£z + 1/^1
In
ez-\P*\

dz cosh2^ —zsinh2^ e2y(co sh y —z si nh j>)


In
cosh2_y—z2 sinh2_y (1—z)sinh>>coch^ coshjy + z sinhj>
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 429

On setting z = tanh vjtanh. y, we obtain

dv coth^ —tanh. a
J= 2
J cosh2 7 tanhy — tanhv 0 —y)

cosh(y—?;) 8y
cosh2_y ta n h j f
J sinh(>>—v) ^ ^ sinh 2y h(2y).

where

h{y) = — j u coth u du.

The quantity Bx which is defined by (30.19) will finally be given for


Pi = o by

Bx = 4n |2 (2 t coth 2y— 1) —---J + 4y coth2.y[l —h(2y)]


I
(30.26)

§ 31. Photoeffect

31.1. Photoeffect in the Nonrelativistic Domain


If an atom has incident on it a photon of energy greater than the
ionization energy of the atom, then absorption of the photon can
occur accompanied by the liberation of an atomic electron. This phenom­
enon is called the photoeffect.
The matrix element for the photoeffect can be written as
S ^ f = —27ti U ^ f 8 (e0 + oj—e),
where

U ^ f = ------?L = -
y2co J
f ^f{r)yeeikrip{{ r ) d r = ---- -}L=rM
y 2 oj
(31.1)

yit(r) and y>f (r) are the electron wave functions in the initial and the
final states, e0 and e are the energies corresponding to these states,
and a>,k,e are the frequency, momentum, and polarization of the
photon.
430 QUANTUM ELECTRODYNAMICS

The cross section for the photoeffect is determined by

da = ^ I ^ o + °>-£) dp, (31.1')


''i-'V
where the summation is taken over the different spin orientations of
the electron in the initial and the final states; p is the momentum of
the ejected electron.
We restrict ourselves here to the calculation of the photoeffect
cross section only, for the A'-shell of the atom (formula (31.1') takes
into account two electrons in the Ai-shell).
We shall, first of all, consider the nonrelativistic case, when the
photon energy differs but little from the atomic ionization energy I.
On assuming that the photon wavelength A is considerably larger than
the atomic dimensions a, A > a, we can replace exp (ikr) in (31.1)
by unity. Further, on noting that in the nonrelativistic region the matrix
a corresponds to the electron velocity v, we obtain
M i ^ = (ev\+f = - k o ( e r \ ^ r (31.2)
Thus, the problem is reduced to the evaluation of the matrix element
of the component of the radius-vector of the electron along the photon
polarization vector e. We have encountered an analogous problem
in the theory of bremsstrahlung in the nonrelativistic domain. The
difference consists of the fact that in (29.37') a wave function belonging
to the continuous spectrum appears as the wave function for the initial
state, while in (31.1) it is replaced by the wave function for the AT-electron.
This function is of the form ^ (r) = Nie~VT, where i] = Z/aQ,
a0 = tr-jme1 and Nt = y Z^jua^.
Insofar as the wave function of the final state is concerned, we must
take for it a wave function for the electron in the Coulomb field of
the nucleus belonging to the continuous spectrum. Since an electron
is liberated in the final state, then in accordance with subsection 29.1
the function y>f(r) must at r -> oo have the form of a superposition
of a plane wave and an incoming spherical wave. The function having
this asymptotic behavior is of the form
—f
V>f (r) = e 2 r ( l + /f)F [—if, 1, —i(pr+pr)]eiPr,
where f = Ze2jtiv.
INTERACTION OF ELECTRONS WITH PHOTONS 431

The substitution of the functions ^ ( r ) and y>f (r) into (31.1) yields
M ^ f =i(oeM, (31.2')
where

M = e 2 r ( 1—i£)Nt j e <pr 1, i(pr-\-pr)]dr

— e2 (31.3)
On noting that

/= *' l(pr+pr)]T ) , . o
and on utilizing

J F [ i ( , I , i(pr-\ p r ) ] - -

l-i£
= 2 jIf c £ > ! . +
C l
^ y— + ^+ P (q -p) - iVPj ’
we obtain

3 X) (p*+ri*)* ( j - / p ) •
Since for a purely imaginary n = — i\n\

In- 1 __ q —2n a r cco t [n|


i« + l
we finally obtain

j — \6ni — -- e~2*Brccot 4 (31.3')


p5 (£2+ l ) 3
and

M ^ { = - \ 6 \ / n ( D e 2^ (31-4)
(£2+ l ) 3

On substituting this expression into (31.1') we obtain the differential


cross section for the photoeffect due to an unpolarized photon:

* = 2- ^ / - l y W i^ -e + ^ d p .
432 QUANTUM ELECTRODYNAMICS

where the summation is taken over the photon polarizations. It can


be carried out with the aid of

^ (pev)2 = P2 - ^ O * 2) = P2( l - c ° s2^)>

where # is the angle between p and k.


On eliminating the d-function by integration over the energy of the
emerging electron we obtain:
e2co g—4f arccotl
I2
d a = 24 — 1— ( l - c o s 2^)t/o, (31.5)
mH* P + l 1—e~2ni ( I - o j ) 2

where do is the element of solid angle into which the electron is emitted.
In order to obtain the cross section for the photoeffect in the case
of polarized photons we must carry out in this formula the substitution
sin2# -> cos20 = sin2# cos2<p, where 6 is the angle between the photon
polarization e and the momentum of the emerging electron p, while
99 is the angle between the (p , k) and (k , e) planes, and we must omit

the factor £ in front of Z v which corresponds to averaging over unpo­


larized photons.
Thus, the angular dependence of the photoeffect in the nonrelativ-
istic domain is determined by the factor cos2 d = sin2# cos2cp. We
see that the majority of the photoelectrons are emitted in the direction
of polarization of the incident photon.
On integrating (31.5) with respect to do, we obtain the total cross
section for the photoeffect in the A!-shell:
ne2 J3 g—4|arccot|
oK = 27 (31.6)
3m a>4 1—e ' 2-^
In the Born approximation, when f < 1, this formula assumes
the form
_ 26 e2 / I \7/2
(31.7)
* — T m / w"

On multiplying oK by the number of atoms n per cm3, we obtain


the absorption coefficient rK at the frequency co
ne2 nP e 4farccot£
xK noK = 27 (31.6')
3m oj4 1—e 2;tvt
INTERACTION OF ELECTRONS WITH PHOTONS 433

Figure 31 shows the dependence of \n(aK/a0) on the photon frequency.


We now consider the photoeffect in the L-shell. The photon ab­
sorption coefficient due to the electrons of the L-shell is given by the
following formulas (16). The absorption coefficient due to the two
2s-electrons (Lj-subshell) is given by
ne‘ n l \ / 2 \ g — arc cot i |
= 210 1+ 3- (31.8)
3m CO’ co 1- 0-2
— r
. > . \
\ s V
\ \
s. \ \ \
\ \
'-X.
\
^ ' ,
_____ Born approxim ation
\ v V N \ ____ _ E xact so lu tio n
\ ----
L x *
V s \ \
X
o x Experim ental p o in ts ~
\ \ \ N . X \ N
\ \
v V s X
h n vx ^ v \
\
XN
X N s
'<
s N ^P b
V \ «
»
>
\ \ _^ S r i ■’'n .

Ak
Cu L --
I.

0.001 0.002 0.005 0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10 20 50


f)Cl)/me7
Fig. 31.
and the absorption coefficient due to the six 2p-electrons (the Lu and L ul
subshells) is given by
= 211^ 1 ^ | 3.,_8 c2 £—4^ arccol £ % (31.9)
3w o r O) 1
where /, is the ionization potential for an L-electron.

31.2. Photoeffect in the Relativistic Domain


We now investigate the photoeffect in the relativistic domain, when
the photon energy is large compared to the energy of a /(-electron.
In this case we must use for the electron wave functions the solutions
of the Dirac equation corresponding to the nuclear Coulomb field.
Insofar as the function y>f (r) is concerned, we take for it the wave func­
tion introduced in subsection 14.5:
y /r ) = Nf u{p)e-iPr\yi - — y v \F [ it, 1, i(pr+pr)], (31.10)
434 QUANTUM ELECTRODYNAMICS

where
2jji$ \ l/2 aZ
**,= £=
1- e - * * ) ’ v
The electron wave functions describing discrete states in the nuclear
Coulomb field can, in accordance with (12.4), be written in the form
f iS»(r)Qnm(n) \
1—f x{r)Qn.m{n)j
where n — r/r, I' — 2j —l, OjVm = —{an)Qym. In the case of the X-shell
j= 1 = 0 , x — —1 and the function rpjlm assumes the form
I fr«(r)« \

where u is a constant spinor satisfying the normalization condition


u*u — IjAn.
In the case of light nuclei where aZ •< 1, we must in formulas
(13.5) and (13.5') defining the functions gx(r) and f j f ) set N = 1, n = 1,
ri = 0, y = ] /l —a2Z 2 & 1, and e m m (l+ ict2Z 2).
As a result of this we obtain g j / ) — 2rf!2e~VT, f J / ) = — 2r]3/2(aZ/2)e~T1T,
z
Tj = ---- »
1
0
.aZ e-nr
Vk(r) = Nt z—r—cos v
.a Z .
z-=—sin$elf Nt=
71ao
if the spin is directed along n, and
0
1
. aZ . . . .
z—— sm$e~1?l -nr

aZ
-i— -cos $

if the spin is oriented opposite to n.


INTERACTION OF ELECTRONS WITH PHOTONS 435

On introducing the matrices y t and the constant bispinor uQ with


components

we finally express the A’-electron wave function for aZ < 1 in the form
y k(r) = 7^(1 — where aZ < 1. (31.11)
We shall use this function for the evaluation of the matrix element
On substituting into (31.1) the expressions (31.10) and (31.11)
we obtain
M ^ f = NlN?u(p){(ey)JQ+ (ey)yi (yJ1) + (y J 2)y 4(ey)}u0, (31.12)
where
J0 = f eiik~p)r~7,TF[i£, 1, i(pr-\-pr)]dr,

J 4 = — ~ a Z J e i(k-p)r- r,r , 1,i(pr+pr)]dr,

/ 2= — y £j e Hk-P)r- nTVF[it;, 1, i{pr-\rpr)]dr.

The quantity £ Wt-+f\2 can be evaluated by means of formula (26.12):

j — Sp{A (m —ip0) A(m —ip)}, p0 = (0, ie0),(31.12')


4 SEo

where
A = (ey)J0F (e y )y i (J1y ) F ( J 2y )y i {ey),

A = y 4A +y 4 = - ( e y ) J f + ( J ? y ) y 4(ey)+ (ey)y4(J 2y).


436 QUANTUM ELECTRODYNAMICS

The evaluation of the trace of the matrix product in (31.12) yields


e—m ,

— - { J o { p J t ) - M P J t ) - 2JoieP) («/i*)-com pl. conj.

+ ^ { 2 ( e J J ( e J 2) J-i J 2. I compl. conj. (31.13)

The integrals J0, J l t J 2 can be evaluated with the aid of


dr -it
B = | eHk-P)r- 7irF[i£, 1, i(pr+pr)] — = 4n
r
where
a -- k 2+(r]—ip)2,
a+b = (k —p)2+v)2.
We can easily show that

SB f if i£\0
. . , , , 1 —■
•'•= - W = l7 M + 2 7 iT r'

(31.14)

,aZ m l 3 &—p d1 aZ m T //) x1


'■ = ' T 7 | ¥ - T W r - T T f [ f < * - ! » - / J *■

These formulas become considerably simplified if v ~ 1 and Za < 1.


In this case £ •< 1 and
ctZy ■y—cos^
Jn = B,
m ( l - ) / l ^ y " 2) 1— ycostf
/aZ I-!)2 k —p
J
\= 2m2 1—vcos'd'
B,
(31.14')
iaZ y \ —v 2
J2= — — 7= = = ( k —p)B,
2m2 l —| / i —^ 2
47 l 2 1—“Z,’2
1B 12 _
™4 \(1—j / l —y2)(l —VCOS'&)J
(■& is the angle between p and k).
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 437

On making use of these formulas we finally obtain the following


expression for the differential cross section for the photoeffect which
is valid for aZ < 1, w ~ 1, (169), (187):
2a6 Z 5 sin2# j (1 —l / f —w2) 2
m 2 A (1-?T co7#)"4\ (1 — v 2^ 2 V c° s #)

— —]/l —v2) ( 1 —wcos#) cos2 9?+4 cos2 <y| do, (31.15)


where
a = ^ o _ ^ )L = m
Ovc'
( l - | / l - » 2)5
*
' " ~ ro + rn
(v is the velocity of the emitted electrons, cp is the angle between the
( p , k ) and the ( k , e) planes).

The integrated cross section for the photoelfect in the AT-shell is


given by (169), (187)
.5
m
a = 4rra4 Z 5 rg(— 1 (y 2 —1 ) 3/2
a>

X
4 , y
(y—2) 1-------.= In 1 * H , a) > /, a Z < 1,
3 y
+1 2y|/y2—1 y — |/y 2—1
(31.16)
where
1 (o-\-m
rn =
l / l - vc m
Anm '
In the extreme relativistic case (o > m this formula assumes the form
YYl
a = Anr^a^Z5— , a> > m . (31.17)
ft)
For large values of Z, when these formulas are inapplicable, a numer­
ical calculation of the photoeffect cross section was made (93). The
following table lists the values of the photoeffect cross section a obtained
in this manner for a series of values of Z and ft).
T able 12
a)jm 0 .6 9 1 2 .2 5

A1 2 2 .3 8.1 1 .2 4 0 .3 5

Fe 1 7 .8 6 .5 1 .0 5 0 .3 0

Sn 1 2 .3 4 .5 0 .7 9 0 .2 4

Pb 7 .9 3 .2 0 .6 0 0 .1 9
438 QUANTUM ELECTRODYNAMICS

§ 32. Production of Electron-Positron Pairs

32.1. Production o f an Electron-Positron Pair by a Photon in the Field


of a Nucleus
We now proceed to investigate the production of electron-positron
pairs.
In order for such a process to be possible, obviously an energy is
required which is not less than 2m. However, a single photon having
sufficient energy cannot produce a pair, since according to subsection
27.1 the laws of conservation of energy and momentum cannot be
simultaneously satisfied. Therefore, for the production of a pair by
single photon the presence of an additional particle, for example a nu­
cleus, is required.
Electron-positron pairs can also be produced in a collision of two
charged particles or photons possessing sufficient energy. However,
such processes play a less important role than pair production by a pho-
^ton in the field of a nucleus. We shall, therefore, first of all consider
this latter process.
Pair production by a photon in the field of a nucleus corresponds
to the two diagrams shown in Fig. 32. In these diagrams p_ and p + denote

the four-momenta of the electron and the positron, while k denotes


the four-momentum of the photon. Topologically these diagrams are,
obviously, equivalent to the diagrams representing the process of
bremsstrahlung in the field of a nucleus (cf. Fig. 25).
The matrix element for the process of pair production can be written
in accordance with Feynman’s rules in the form
iec
sm f = u{P-) Ii > („) x i« > („) l A z U
(-/>+)>
l/2 C0 I fi+ m * W ' WV s+ "< 2
(32.1)
IN T E R A C T IO N O F E L E C T R O N S W IT H P H O T O N S 439
A

where Ajf^q) is the Fourier component of the nuclear field, u(p_)


and v ( —p +) are the electron and positron spinor amplitudes and
f i ~ P - ~ k = —P+ + <7, f 2 = — p + J r k = p_—q.
A comparison of the matrix element (32.1) with the matrix element
(29.10) for the case of bremsstrahlung shows that the former may be
obtained from the latter if the following substitution is carried out
in the latter: P i~ + —p +, p2 -+p_, k - + —k, «i(pa) -*• u(—/?+),
u 2 ( p 2) -*■ u ( P - ) (the replacement of k by —k corresponds to the fact

that in contrast to bremsstrahlung the photon is now absorbed). There­


fore, we do not need to calculate the cross section for pair production,
but can utilize the already available results for the bremsstrahlung
cross section. We must only keep in mind that in contrast to the brems­
strahlung process, for which the density of incident particles is equal
to the electron velocity, the density of the incident particles—photons—is
now equal to the velocity of light. Moreover, the number of final states
is altered and is now given by g f — d p _ d p + / ( 2 j i y .
On taking these changes into account we obtain the following
expression for the differential cross section for pair production by
a photon in the nuclear Coulomb field (21):

Z 2a? \p+\\P-\ de+


4 {
+
i

(2n)2 OJ3 ?4 m* { ' L ’ 1 J


+ q 21 [ ^ , / 7 _ + / 7 + ] 2 ] (32.2)
»2 J «1«2 i
where
m2X! = f i + m2 = —2p _k; m2x 2 = / | + m2 = —2p +k,
do+ and do_ are elements of solid angles containing the momenta p +
and p_ of the positron and the electron, e+ and e_ are the energies of
the two particles.
They are, obviously, related by e++e_ = to.
Formula (32.2) can also be rewritten in the form
Z 2a3 p +p_de+ sin0+ sXnQ_dQ+dQ_d(pJr
da =
2n or

J P\ sin20
si p?_ sin26_
X (4e2—q2)- (4e2 - q 2) +
p
1(£+ - JP++ cos 0+)‘ (e_—p_ cos0_)2
440 QUANTUM ELECTRODYNAMICS

2p+p_sin0+sind_cos<p+ g
q2—2a>2)
(e_—p_cosO_)(e+—p + cos0+) V +
^ p2+ sin20+ + p2_ sin2B_ \
(32.3)
(e_—p_cosO__)(e+—p + cos0+) )
where Q± are the angles between p ± and k, <p+ is the angle between the
(k, p +) and the (k , p_) planes and
?2 = { k - p _ - p +)2.
The angular distribution determined by this formula is of a fairly
complicated nature. It simplifies considerably only in the extreme rel­
ativistic domain when the principal role is played by small angles Q±.
In this case the electron and the positron are emitted predominantly
forward, i.e., within a narrow cone surrounding the direction of motion
of the photon; the effective angular aperture of this cone is given
in order of magnitude by 0± ^ m / o j .
On replacing in (32.3) sin 6± by 6± and cosfL by 1— we obtain
the following angular distribution for small angles:

da — 4 ^ - __£~£+ [a>2(n2-\-v2) 2e_e+(u2£2-\-v2r]2)


7i m 2 co2 q 4
+ 2(ei + £ + )uv&i cos 9o+]de+6_dd_ 0+ dd+ dy>+, (32.4)
where

u = P -0 - P+v+
V — + ,
m m -u-o 5 -v
Integration of (32.3) over the angles yields the differential cross
section for pair production with the positron energy lying in the range
£_i_ to £_j_“I- :

- sOW
da — l p A de+ \ _ ± —2e+e P l + P 2-
or I 3 " P \P 1

nr ___ z+z-
\p+\3 \p+\\p-
8 e+e_
f L (i?+f t + p \ p l ) -
3 Lp +||/>_|
m2to £+£- ~ P- > , e+e _ - p 2+ 2a>£+e_
9 9~ (32.5)
2 \P+\\P-1 \P- \ 3 \P+\3+ P+P-
INTERACTION OF ELECTRONS WITH PHOTONS 441

where

0 = /‘qZ2a , U = 2 In i i + J Z k ,
m

L = 21n-£-+£r - ± L ^ J i ^ l ± '" S .
moo
In the extreme relativistic domain when all the energies are consid­
erably greater than m(oo,e± > m) this expression simplifies consid­
erably:

d° = 4 $ M e l + e - + h e- ) { la2^ - j ) - (32-6)
Formulas (32.5), (32.6) do not take into account the screening of
the nuclear field. Therefore, in accordance with (29.27) they are valid
if (2e+ejmoo) < 137Z ~ ^ .
Screening can be taken into account in the same manner as in the
case of bremsstrahlung. In the extreme relativistic domain (co, > m)
we obtain for the cross section for pair production

do (£i + £y(<f,(C )-^lnZ


or
lOOmco
(A R )-jln z jj, f= (32.7)
+ y £- £- Z 1/3£ £+
where 0 X(C) and 0 2(O are the same functions which appear in (29.29).
In the case of complete screening when (2e+e_/mco) > 137Z-1/3,
formula (32.7) assumes the form

da= 1 + £ l + - £ ,£ _ )ln(183Z - i / 3 ) - .e+£ (32.8)


or
In order to obtain the total cross section for pair production we
must integrate (32.5) over the positron energy. However, it is impossible
to integrate this expression directly. Such an integration can be carried
out only in the extreme relativistic case by utilizing formula (32.6).
On integrating this formula we obtain the following expression for the
total cross section in the extreme relativistic case without taking screening
into account:
; * k/ 2-8l ,n ----
2(0 218 \ (32.9)
o= & where go m . >

\ 9 m " 2 7 /’
442 QUANTUM ELECTRODYNAMICS

On integrating (32.8) over e+ we obtain the total cross section for


pair production in the case of complete screening

(32.10)

In § 55 we shall establish the connection between the total cross


section for pair production by a photon and the cross section for
the forward scattering of a photon in the field of a nucleus. By utilizing
this relation we can show that the total cross section for pair production
without taking screening into account is determined by the following
formula (160):

= ■— 2 t [ - (6 4 + m n I ) E W ^r)
I

+ (.25 + A +42^ ) K( y / , - _ ! _ ) ] } , (32.11)

where
CO

r, . r . * + ]/* 2- y ___ dy___


n x - \ / x 2- y \ / y ( y - 1)

and F(x) and E(x) are elliptic integrals:


rt/2 nl2
F(x) = J (1 —x 2 sin29?)“^ d<p, E(x) = j (1 —x 2 sin2^)^ dcp.
o o
In the extreme relativistic domain this formula reduces to (32.9).
Figure 33 shows the energy distribution of the particles of the pair
determined by formulas (32.5) and (32.7). Along the horizontal axis we
have plotted the ratio of the kinetic energy of the positron e+—m to
the total kinetic energy of the pair co—2m, while along the vertical
axis we have plotted the quantity [(co—2m)/&] (do/de+). The different
curves refer to the different values of co/m which are shown next to the
curves. The curves corresponding to co/m = 6 and co/m — 10 are valid
for all Z since for these values of the photon energy screening can still
INTERACTION OF ELECTRONS WITH PHOTONS 443

be neglected. The other curves take screening into account, and refer
to lead and aluminum. The curve corresponding to co/m = oo refers
to complete screening.

Fig. 34.

For low photon energies the curves have only a'single flat maximum
corresponding to the electron and positron having the same energy.
444 QUANTUM ELECTRODYNAMICS

At large energies the curves have two identical symmetrically situated


maxima of which one corresponds to maximum positron energy and
minimum electron energy, while the other conversely corresponds to
minimum positron energy and maximum electron energy.
Figure 34 shows the dependence of the total cross section for
pair production on photon energy in the case of lead, copper, and
aluminum.

32.2 Exact Theory o f Pair Production by a Photon in the Field o f a


Nucleus in the Nonrelativistic and the Extreme Relativistic Cases
The foregoing formulas for the pair production cross sections are
valid in the Born approximation, when lj± = (Ze2jhv±) << 1 (v_ and
v + are the electron and positron velocities).
However, in the nonrelativistic (w± < 1) and the extreme relativistic
(1—v ± < 1) cases it is possible to evaluate the pair production cross
sections exactly by utilizing the exact electron and positron wave
functions in the Coulomb field in the former case, and the functions
introduced in subsection 14.5 in the latter case.
This calculation is carried out in a manner analogous to the evaluation
of the bremsstrahlung cross section.
The matrix element determining pair production can be written
in the form

S£f!r == i —^L= f y)?_eay)+eikr~imt dr d t = 2ji M^ (5(a>—e+—e_), (32.12)


y 2(o J

where ipj(r) and y>+(r) are the exact electron and positron wave functions
in the nuclear Coulomb field. Since both these particles are created,
the asymptotic behavior of the functions y>_(r) and ^ +(r) for large
values of r must be the same, viz., the functions y>__(r) and y>+(r) must
at r -> oo have the form of a sum of a plane wave and an incoming
spherical wave. This asymptotic behavior represents the difference
between the matrix elements for pair production and for brems­
strahlung: in the latter process only the wave function of the final
state has such asymptotic behavior, while the wave function for the
initial state has at r -> oo the form of a sum of a plane wave and
an outgoing spherical wave.
INTERACTION OF ELECTRONS WITH PHOTONS 445

Let us evaluate first of all the cross section for pair production
by a photon in the extreme relativistic case. In order to do this we
use as and ip+ the functions

= 7V_e1J>-r |^l —- — av]w .F(— 1, —ip ^r—ip_r),


2e_ J
(32.13)

W iV+ e -^ +r|^ l+ — a v j u f ( —if + , 1, ip+r + i p +r).

where u and v are the unit spinor amplitudes for the electron and the
positron, and N_ and N + are normalizing factors which we choose
--f.. --f+
equal to N _ = e 2 JT(l + /f_), N + = e 2 T ( l -J-/£+), corresponding
to unit amplitudes for the wave functions at infinity. (The subscripts
-|- and — both here and later denote quantities which refer respectively
to the positron and the electron.)
After substitution of formulas (32.13) into (32.12) the matrix element
Mj_ assumes the form
le*■
M± t T = N + N* 0i{k— p +— p_)r
[l + ^ a v j u
4 n y 2 (o +

XF(i£_, 1, ip_r-\-ip_r)ea 1+
[1 +27aV]
X v*F (—i£+, 1, ip+r+ ip +r)dr
■= C{u*eav Ix+u*(ea) (a/2)^ + w*(a/3) (e<x)v), (32.14)
where
11 = j e“trFxFzdr,

12 = — f eltirFxV F zdr,

h = 2~ ~ j e ‘q r F ^ F x d r ,
Fx F(ig_, 1, ip_r+ip_r),
Fj = F (—i£+, 1, ip+r+ ip +r),
q = k — p + — p_,

ie2 N +N*
C
2* j/2co
446 QUANTUM ELECTRODYNAMICS

The integrals Ix, I 2, / 3 can be evaluated with the aid of formulas (29.47).
The pair production cross section is related to the matrix element M ±
by the expression
j o 1 V i ^ p \ d p +p_e_do+do_
dot = 2 n ^ 2 (2*jif-- '
ft+M-.v

where ^ denotes summation over the components of the particle

spins.
The result of these calculations leads to the following formula for
the cross section for pair production by a photon:

na 2 1
da = — a —---- sin d_ sin Q,dQ,dQ_dcp
sinh na 2n m2 or + ^ ^
| V2(x) Tpi sin2 Q_(4e2+—q2) _ p \ sin2 B+(Ae2
_ —q2)
I 94 L (e_—p_ cos B_)2 (e+—p + cos6+)2
(4e_c+-\-q2—2a>2) 2p_p+ sin0_ sin0+ cos 99
+
(e_—p_cos 0 _) (e+—p+ cos d+)
2cu2(p isin 20 _ + p isin 20+) 1 a2[co2—(p_+ p+)2]2 W 2(x)
(e_—p_ cos d_) (e+—p+ cos 0+”)J [4eo2(e_—p_cos0_) (e+—p + cos0+)]2
I"pi sin2 6 J 4 e \ —q2) p \ sin20_h(4ei—q2)
L (e_—p _ co sQ_)2 (e+—p + cos0+)2
(4e+e_ + 9 2—2co2) 2p_p+ sin0_ sin 0 + COS99
(e_—p_ cos 0 _) (e+—p + cos Q+)
2 co2(p i sin2 0 _ + p i sin2 0 +)
—4co2 (e_e++ p _ p + cos 0 _ cos 6
(e+—p + cos 0 +) (e_—p_ cos 0 J
(32.15)
where the same notation has been used as in formula (32.3). The func­
tions V(x) and W(x) are defined by formulas (29.48'), a = Za, and

x = l _ _______ ga[fl>2 —(p+—p_)2]_______


4oj2(e+—p + cos 0 +) (e_—p_ cos 0 _)

In the particularly interesting case of small angles 0 _ ~ 0 +~ m /e


the cross section is of the form
INTERACTION OF ELECTRONS WITH PHOTONS 447

na V a 2 1 2 2
do = 8a - =^ Ld e , 6 , d d : d dd dtp
sinh^ta / 271 m- a)3 + + -i- - - ^
X{q~4V2(x) [a>2 (w2 + v2) £r\—2e_ £ + ( « 2 t?-\-v2r}2)
+ 2 (el-|-£^_) uv£rj cos 99] + a2W 2(x)£2rf [a>2(1 —(u2+ v2)rj£)
— 2e_e+(u2£2-\-v2r)2) — 2(e2_-\-e^)uvr]£ COS 99]}, ( 3 2 .1 6 )

where u = (p+d+/m), v = (p_B Jm ), f = ( 1 / 1 +w2) and = ( 1 / 1 + v 2).


The difference between the results given by formula (32.15) and by
formula (32.3) which is valid in the Born approximation is greater than
in the case of bremsstrahlung. This is associated with the fact that
the scattered waves for the two particles overlap in the case of pair
production to a greater extent than in the case of bremsstrahlung.
The pair production cross section integrated over the angles is of
the form
, _ Z 2a3 d s,
do = 2 — s----- ± |e + + e i + y £ ±£_J 21n— —1—2/(Z) , (32.17)
n r or m oi J
where /(Z ) is defined by formula (29.49').
If screening is taken into account the cross section is given by
, Z 2a3 de,
do — — 111 | k + t i ) [ ^ O O - y l n Z - 4 / ( Z ) j
m2 a>3

+ - s + t - U c r t - y lnZ -4/(Z )J}> (32.18)

where (y) and 0 2{y) are the same functions which appear in (29.30),
y — 100(ma)/£+£_)Z“1/3.
The total pair production cross section without taking screening
into account is equal to
(32.19)
9 137 L‘“ m 42
In the case of complete screening we have
28 Z 2
o— ^ ln (1 8 3 Z -3 /3) - — - / ( Z )j. (32.20)
9 137
We see that the correction due to the Coulomb field is always neg­
ative and is equal to
(32.20')
9 137
448 QUANTUM ELECTRODYNAMICS

In the case of lead the relative correction amounts to approximately


10% (it is assumed that ln(2ut/m) is of order of magnitude 3-5).
We now obtain the cross section for pair production by a photon
near threshold, p + << m. The wave functions are in this case de­
termined by formulas (32.13) in which the expressions in square brackets
are to be replaced by unity. Therefore the matrix element M± has
the form
M ± = Cu*eaivlx (32.14')

A simple calculation leads to the following expression for the energy


distribution of the positrons in the nonrelativistic case:
aZ 2 2______4ti% %
12ti r° (1 - e - ^ - ) (e8^ " — 1)
X [<u—2m + (aZ)2(a>—w)]

X [(e+—m) {co— 2m — (e+ —m)}]1!2 (32.21)

32.3. Pair Production by Two Photons


The process of pair production by two photons corresponds to the
two diagrams shown in Fig. 35. According to Feynman’s rules the matrix
element determining this process is equal to

X v{—p +){2n)i d(k-\-k'—p +~ /?_), (32.22)

K k

Fig. 35.
where
fi = -p+ H -k = p _ - k \ / 2= ~ p + + k ' = P --k,
m 2Xi = f l + m 2 = — 2p j k = — 2p_k',
m 2x 2 = f \ - \ - m 2 = — 2p +k ' = — 2p _ k ,
IN T E R A C T IO N O F E L E C T R O N S W IT H P H O T O N S 449

k and k' are the four-momenta of the photons, e and e are their po­
larizations, p_ and p + are the four-momenta of the electron and the
positron.
The differential cross section for pair production is related to
S S /b y
ei
da — uQd — d ( k + k ' - p+ - p_) § ( w + co'- e+ - e_ ),
4cuco'
(32.23)
where
ifx—m A A if2—m A,
Q ---- e + e — g----- e
nrx1 ml x2
(32.23')

and J is the photon flux density.


The matrix Q has the same structure as the analogous matrix in the
case of the Compton effect. On setting in the latter 'p1 = —p +, kx = k,
ex = e, p2 = p_, k 2= —k', and e2 = e', we obtain Q. Therefore, we
can use the results already available for the Compton effect, and write
the following expression for the quantity \uQv|2 summed over the spins
of the electron and the positron, and averaged over the photon
polarizations:

where
S p F = Sp {Q (ifo - m ) yiQ + ySP z-™ )} >
and the following substitution has been performed in Q : e - * y v and
C -*■y
?
.-
AI

It is convenient to carry out further calculations in the center of


mass system of the colliding photons, in which k' = k, p + = —p_,
a) = a>' = co0 and e+ = = co0.
On eliminating the (5-functions by integrations over dp_ and de+:
6 (k4rk' — p +—p_)dp_^> 1,
„ . p .e .d o ,
(5(tO~r ^ —£_j_ £—) ^P-\- ~^ ^’
450 QUANTUM ELECTRODYNAMICS

where do+ is the element of solid angle containing the direction of positron
motion, and on noting that the photon flux is given in the center of mass
system by 7 = 2c, we obtain the following expression for the cross
section for pair production by two photons:

- * ( ^ + vX2j
) A \«2
v + v Xll
) \ )' (3 2 *25)
On substituting into this expression
m2x1 = —2p +k = —2p+k-\-2e+on — 2a)0(co0— m2) cos 0),
m2x 2 = —2p__k= —2p_kJr2e_a>
— 2p+k-\-2e_co = 2a>0(cd0+ )/(o$—m2 cos 0),
where 0 is the angle between the vectors k and p +, we obtain
_ rl m2\fo)\—m2(2eo2—m2+(co2—m2)sin20
4 \ m2 cos20+a>osin20
2(a>o—m2)2sin40 |_ . n „
— 7-j—
(m2cos220fl, 2 • 2m2f271 sin 0 ^ -
+ ai2sin20)2J (32.26)

The total cross section for pair production by two photons is equal to

(32.27)

In accordance with the arguments of subsection 28.3 this cross section


is relativistically invariant. In an arbitrary coordinate system in which the
photon k' is moving in the direction opposite to the photon k the
relation obviously holds a>- to' = a>o(A||—it'). Therefore, in such systems
the cross section is defined by formula (32.26) in which we must
set ojq= j/coco'.
On introducing the quantity x — (1—m2/o>o)')1/2 we can rewrite
the total cross section in the form
l+ x
* = ^ ( l - * !) | ( 3 - * ‘) l n -\-2x(x2—2) (32.27')
l-x I
INTERACTION OF ELECTRONS WITH PHOTONS 451

32.4. Pair Production in a Photon-Electron Collision


For a pair to be produced in a collision of a photon with an electron
at rest the photon energy must exceed 4m. Indeed, from the conservation
laws written in four-dimensional form k + p = p '+ p '+ p " ', where k
and p are the four-momenta of the photon and the electron in the initial
state while p , p " , p " are the momenta of the particles (electrons and
positrons) in the final state, it follows that (k+p)2 — —m2—2ma>
= (p '+ p '+ p "')2- But the square of the four-momenta of three
particles of equal mass cannot be smaller in absolute value than 9m2;
therefore we have —m2—2ma> ^ —9m2, from which it follows that
co ^ 4m.
The process of pair production in a collision between a photon and
an electron differs from the process of pair production in a collision
of a photon with a nucleus by the fact that the effect of the electron cannot,
generally speaking, be replaced by an external Coulomb field, as we
have done in discussing pair production by a photon in the field of a nu­
cleus. In this sense the process is analogous to the process of emission
of radiation in a collision of two electrons, and like the latter process
it is described by elements of the third order scattering matrix S(3) connect­
ing one photon state and three electron states (more accurately speaking,
two electron states and one positron state). The corresponding eight

Fig. 36.

diagrams are similar to the diagrams of Fig. 30 representing the process


of emission of radiation in an electron-electron collision. Two of these
diagrams are shown in Fig. 36. The others can be obtained by interchang­
ing the momenta of the real and the virtual photons, and also the
momenta of the electrons in the final state p2 and p2 (pt is the electron
momentum in the initial state, p + is the positron momentum).
452 QUANTUM ELECTRODYNAMICS

In order to obtain the matrix element S ^ f corresponding to the


process under consideration it is, obviously, sufficient to perform the
substitutions p'2 -> —P+, k -> —k, and u' -> v (—p +) in the matrix
element (29.50) which determined the emission of radiation in a collision
of two electrons.
We give here only the results for the total pair production cross section.
Near the pair production threshold the cross section is equal to (140)

In the extreme relativistic case (200) we have

<t = ar2
0 — In— —10.8 . (32.29)
\ 9 m I
This expression differs from the cross section for pair production by
a photon in the field of a nucleus for Z = 1 only by the coefficient
in the argument of the logarithm.

32.5. Pair Production in a Collision o f Two Fast Charged Particles


We now consider pair production in a collision of two fast charged
particles, for example, nuclei of charges Z xe and Z2c. This process
corresponds to the diagram given in Fig. 37 in which the dotted lines
represent the external potentials associated with the two charges.

c(p--p) a(p++p)

Fig. 37.

The matrix element for this process is given by

— i1((2tr)4ez«(p_) j dip a(P-~P) i p - m a(p++p) v (—P+)>


°5i -(2)
>r
(2n)4 p ’L^r rrP (2k )4
where a{q) is the Fourier component of the total potential due to the
two charges.
INTERACTION OF ELECTRONS WITH PHOTONS 453

The probability of this process turns out to be independent of the


time. After summing over the electron and positron spin components
it can be written in the following form:
dp+ dp_
dw = £ I W
p+.p- " (2*)6

= 4 ( 2 ^ r Sp {f - f + i ? %p++p>«p+

X / diP'a<J’++P') pl l + ™2 a(P --P ') (32'30)

We now determine the field produced by the charges. The potential


(A, A0) produced by a particle moving with uniform velocity v satisfies
the equations
□ A = —ZevS(r—v t —r0),
□ A q=z —Zed(r—v t —r0).
On expanding A^ into the Fourier integral.

we obtain
, , 2izZe . , s
Afl(q) = — ^ - u ^ M e lQI°,
H
where u {[v l]/X —v2] ,[ i l \ / \ —v2]) is the four-velocity of the particle.
Therefore, the Fourier component of the total potential produced by
both particles will be equal to

an = A t A - A .‘■t* qi q^ (32.31)

where the indices 1 and 2 denote the two particles and r0=
We now note that
d[(p_-p)u™]d[(p+p+)u™] = 0,
d [(p _ -p )u {2)]d[(pJr p+)u i2)] = 0,
S [ ( p + p + ) u w ] 6 [ ( p — p _ ) u {1)] = 0,

S[(p' + p +) u i2)]S[(p_-p')u{2)] = 0.
454 QUANTUM ELECTRODYNAMICS

Indeed, since uq is an invariant, we can use the coordinate system in


which has only a fourth component, and in such a system
<5[(p_—P)u]d[{p+p+)u] = d (e_ — p 0) S ( e + +/?0) = <5(e_+e+)<5(e_—p 0) = 0,

where e± and p0 are the fourth components of the vectors p ± and p.


Therefore, the substitution of (32.31) into (32.30) yields

dP+dP_
dw = (27t)4 e+e_ P (32.32)
{2nY ’
where

Sp G = i Sp {[m< 0 7 - 70) w(2>+ D(2>( / / '- / ') w(1) ]


X ( i p + — m ) [ w ( 1 ) ( z 7 ' — / o ) m (2) + w ( 2 ) ( / 7 — 7 0) m (1)] ( i p _ — m ) } ,

f m P/1e~Hp+p+)r° 5[(p_—p)u^)] <5[(/?+ /?+ ) « (2)]


J P -- ({^p 2++ m2) ( p_ _ p)2(p + pJ 2
m 2) ( p _ — p ) 2 ( p + p + y

P = j d*P P „ e - i{ p — p ) r °6 \ { p _ - p ) u {2)]

{p2+ m 2) ( p _ - p ) 2 {p+ p+ f
<5[(/?+/>+)M(1)]

In order to obtain the differential cross section for pair production


we must integrate dw over 2nr0 dr0. We give here only the result for the
cross section integrated over the angles of emission of the electron and
the positron (117) on the assumption that m << e± << (m /j/l —v2),
where v is the relative velocity of the colliding particles:

g £+ + £?_ + -v-£_|_£_
e,e
da = in
m(e+ + e_)
m
X In ­ ode,“r de__. (32.33)
(e++e_) ] / l —v
The total cross section is equal to

(32.34)
" = i 7 ^ ( z ‘Z!a)V S( ln )7 f = !) 3'

This formula is valid if In (1 / ] / \ —v2) > 1.


In § 34 we shall obtain this formula by utilizing the method of
equivalent photons.
INTERACTION OF ELECTRONS WITH PHOTONS 455

In the foregoing calculation we have replaced the effect of the collid­


ing particles by an external field. Such a replacement is valid in the case
of very fast particles whose state of motion is practically unaltered as
a result of pair production. But in the general case the effect of the
colliding particles cannot be reduced to that of an external field, and the
process of pair production in a collision of two charged particles is shown
not by the diagram of Fig. 37, but by the two diagrams of Fig. 38.
In these diagrams the upper two solid lines correspond to the colliding
particles, while the lower ones to the electron and the positron of the
created pair.

7 --------- «--------- r*

(o.) (b)

Fig. 38.

If a pair is formed in a collision of two electrons then in addition to the


diagrams shown here we must also take into account the exchange
diagrams (cf. subsection 29.10).
Thus, the production of a pair in a collision of two charged particles
is, generally speaking, a fourth order process.
We shall not reproduce here the calculation of the cross section
for pair production in a collision of two electrons, but shall confine
ourselves to indicating the order of magnitude of the cross section.
Since this process is a fourth order process, the cross section for the
process amounts to approximately a = 1/137 of the cross section for the
emission of radiation in a collision of two electrons. If one of the
electrons was at rest before the collision, then from the conservation
laws it follows that the second electron must have an energy exceeding 7m.
In the case when one of the colliding particles is fast, so that its
state of motion is practically unaltered as a result of pair production,
the effect of this particle may be replaced by an external field. In this
case we may remove the upper line from the diagrams of Fig. 38, and
the diagrams representing the process of pair production will assume the
456 QUANTUM ELECTRODYNAMICS

form shown in Fig. 39 (the external photon lines in this case correspond
to an external field).
We see that in this case the pair production process is a third order
effect. Such a case occurs, for example, in the case of pair production
in the field of a nucleus. The effective cross section for pair production
by an electron in the field of a nucleus in the extreme relativistic case
£ > m ( e is the initial electron energy) is given in order of magnitude
by the formula (23a, 143)
a ^ rt Z Z - - ( \ n — ) , (32.35)
n \ mj
where Ze is the nuclear charge, r0is the classical electron radius. On set­
ting Z = 1 in this formula we obtain the order of magnitude of the cross
section for pair production in the collision of two relativistic electrons.

//
/ •*— /r*
/ /

(a) (b )

Fig. 39.

Formula (32.35) actually corresponds to formula (32.34) if we set


Z x = 1 in the latter, and assume that one of the particles is initially
at rest.
We note that in comparison with the bremsstrahlung cross section
the cross section (32.35) contains the coefficient ajAn, but also a higher
power of the logarithm (in formula (32.35) screening is not taken into
account).
The diagrams of Fig. 39 also represent the process of pair production
in a collision of two heavy particles moving at nonrelativistic velocities,
since the nonrelativistic system of two particles is equivalent to a single
particle situated in the external field. The pair production cross section
is in this case approximately equal to (88)
2
Z \ z \ IZ 1M 2—Z 2M 1\
a2rl (32.36)
m 2t 2 \ M1 )
INTERACTION OF ELECTRONS WITH PHOTONS 457

where M x and M 2 are the particle masses, T2 is the kinetic energy of the
particle of mass M 2, and it is assumed that the particle of mass M x is
initially at rest, and that T2 >> 2m.
In conclusion we note that every process of radioactive decay with
a sufficiently large release of energy may be accompanied by pair pro­
duction. This phenomenon can be regarded as pair production by an
external field with frequencies in excess of 2m. In the disintegration of
the nucleus into two parts the probability of pair production is determined
by the expression (136):

(32.37)

where Z ' = Z x (1 —Z 2A2/Z XA X), A x and A2 are the atomic mass numbers
of the fragments, Z xe, Z 2e are their charges, e and v are the energy and the
velocity of the smaller fragment, M is its mass, m is the electron mass.
The probability of pair production in /1-decay amounts to approxi­
mately 10“6—10“7 (196).

§ 33. Annihilation of Electron-Positron Pairs into Photons

33.1. Annihilation o f a Pair into Two Photons


The process of the production of electron-positron pairs by photons
has its counterpart in the process of a pair disappearing and giving rise
to photons. Such a process is known as pair annihilation. If the electron
and the positron are free then annihilation with emission of a single
photon is impossible since this process is not allowed by the laws of
conservation of energy and momentum. For the annihilation of a free
pair into photons at least two photons are needed, but if a positron
collides with a bound electron then single photon annihilation becomes
possible.
The most important process is the one in which the colliding free
positron and free electron are converted into two photons. Therefore
we start by investigating this process.
Two-photon pair annihilation corresponds to the two diagrams
shown in Fig. 40. They coincide with the diagrams representing pair
creation by two photons—a process which is the inverse of two-photon
annihilation.
458 Q U A N T U M E L E C T R O D Y N A M IC S

The matrix element which determined two-photon annihilation can be


expressed in accordance with the rules of subsection 25.5 in the form

ie“ \ 1~ i f i —m a, A, i f 2—m
C(2)
-+/ —

=-v{—p +) l e — ---- e + e — »------ey-
2 / ,coco tril xl m2x2
Xu(pJ)(27i)i d(p++ p _ —k —k'), (33.1)
where the same notation has been used as in (32.22).
The differential annihilation cross section is equal to
dkdk'
da 7 - 1^<2«l d(p+Jr p _ —k —k ') <5(e , + e _ —co— co'),
4 coco J{2 n f
(33.2)
where J is the flux density of the colliding positrons and electrons and

a ifx—m a, a, i f 2—m
Q = e — 5---- e + e — ----- e.
m2xy m2x 2

The structure of the matrix Q is obviously identical with the struc­


ture of the corresponding matrix for the Compton effect: on carrying
out in the latter the substitution px = p_, k x — —k', ex = e', p2 = —p +,
k2 = k, and e2 = e, we obtain Q. Therefore, we can utilize the
already available results of § 28 and write the following expression
for the quantity ^ ,\ v Q u \ 2 summed and averaged over the particle
polarizations:

It is convenient to carry out further calculations in the system asso­


ciated with the center of mass of the pair in which p + = —p_ = —p 0,
e+ = e_ = e0, k' = —k, co = co' and the flux density is equal to
J = 2(p0/e0) = 2v0.
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 459

On eliminating the (5-functions by integrating over dk' and da>:


& ( p + + P - - k - k ') d k ' -* 1,

d(e+-\-e_—c o - cd ' ) dk -> °J ,

where do is the element of solid angle containing the photon momentum


k, we obtain

(33.3)

where
m 2x i = — 2p + k = 2 p 0k -\-2 e 0a) = 2 e 0 (£ 0 + p 0 cos 0 ) ,
m2x2 = —2p_k ~ —2p0k-\-2e0(o = 2 e 0 ( e 0 — p 0 cosQ)
and 0 is the angle between p 0 and k.
Thus we have
m2 ^ p l+ p l^ o 2p\ sin4 Q
do = sin B dddcp
4poeo E q -— P qc o s 2 0 \e2—plcos2d)2
1 do
[1 + ^ s i n 220—®o(l—sin40)] (33.3')
4 vno°o
e; (1—^oCOS20)2
In order to obtain the total annihilation cross section we must
integrate do over 0 from 0 to n and over qp from 0 to n (these limits
of integration correspond to two indistinguishable particles—photons in
the final state). As a result of this we obtain
6= 71 <p=n

a= In 1+^0 2{v2- 2 ) \ . (33.4)


1 — Vn
0= 0 V = 0

Since the cross section is relativistically invariant, then in order to


transform to another system it is sufficient to express e 0 and p0 in terms
of the particle energies in that system.
We consider the especially interesting case when anihilation occurs
as the result of a collision of a positron with an electron at rest.
The positron energy in the rest system of the electron is, obviously,
equal to
0 = £p+ floA) _ _fo+jP?_ = 2e2 0—m2.
£+ “ “ i/3 -s ' "i2
460 QUANTUM ELECTRODYNAMICS

where v0 = p j e 0 is the velocity of the electron rest with system respect


to the center of mass system. From this we obtain

On substituting this value of v0 into formula (33.4) we obtain the cross


section for the annihilation of a pair into two photons in the electron
rest system (48), (193):

nr% j y2+ 4 y + l y+3 1


a= In (y+ / y 2—1) (33.5)
y+1 1 y2—l
where y — /m.
For low positron energies the cross section is inversely proportional
to the positron velocity:
nrl
a (33.6)

while the annihilation probability w is independent of the velocity:

w = Znv+a = Znnrl sec-1, (33.6')

where n is the number of atoms per unit volume. The lifetime of a slow
positron is equal to 1/ vv; in the case of lead 10-10 sec.
For large positron energies (e9_ > m)

a= ^ ( ln^ ± , ) . (33.7)

33.2 Polarization Effects in the Two-Photon Annihilation o f a Pair


In the preceding subsection we have given the expressions deter­
mining the cross section for the two-photon annihilation of a pair
averaged over the polarizations of the electrons and the photons. The
annihilation cross section for polarized particles, and the polarization
of the photons produced in this process can be investigated by the
general methods presented in §§ 2, 10, 26 and applied in §§ 28 and 29.
We confine ourselves here to stating some of the results.
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 461

1) Correlation o f photon polarizations. The differential anni­


hilation cross section for unpolarized particles with the production
of two plane-polarized photons has the form
rl m2 Jf 1
cos26) (exe2)2
8 v0el | 1 —z;2cos20 ^

do. (33.8)

where e0 and v 0 are the positron energy and velocity, and 6 is the angle
between the directions of the positron and photon momenta in the
center of mass system.
At low velocities it follows from (33.8) that

dou » = do(2-» = 0, do^-v = A do, (33.9)

where do{Uk) is the cross section corresponding to the production


of photons polarized in the direction et, ek, with ex lying in the plane
(pQ, k), while e2 is perpendicular to this plane.
We see that for v0 0 photons are produced which are polarized
perpendicular to each other. This result has a simple meaning. It follows
from the law of conservation of parity. In the low-velocity limit anni­
hilation occurs principally in an E state which is odd. But a system
of two photons is odd if their polarizations are mutually perpendic­
ular.
2) Annihilation of a longitudinally-polarized pair. If the polari­
zation vectors for the electron and ^ (+) for the positron have non­
vanishing components only in the direction of their relative motion,
the differential annihilation cross section has the form (147)
d o = d o 0(\+ V + <$>->/), (33.10)
where do0 is the annihilation cross section for an unpolarized pair
(cf. 33.3)) and
_ —cos40) — 1+ ^ —vl(\ —^ )sin 26
vl{\ —cos40 ) + 1—Vq+ vKI —f^ sin 20
(vQand 0 are, as before, the velocity and the angle between the momenta
of the positron and the photon in the center of mass system).
On integrating (33.10) over the angles we obtain
f f = f f 0(l-K<->S<+>n (33.11)
462 QUANTUM ELECTRODYNAMICS

where a0 is the total annihilation cross section for unpolarized particles


(cf. (33.5)) and
tp—
* I y

_ 3+^0 , l+ w 0 3+^2
a i~ ^ W vT ’

- — (3 ~ 3 ^+ 2 ^)+ (2 ^+ 3 ^-3 )ln |± ^.


v0 tv 0 1 v0
Fig. 41 shows the dependence of F on the energy in the laboratory system
of coordinates (i.e., in the system in which the electron is at rest).
If only one of the particles undergoing annihilation has a longitudinal
polarization equal to C, then da = da0; in this case the photons are
circularly polarized :

2a0—2z)g-^-sin20 cosd
£2= C --------------- 2 - - - --------------------------- . (33.12)
1+^o+^n—ir Um2 sin20 (l+ c o s20)-fz;§ sin20

c.MeV
A

On integrating the numerator and the denominator of this expression


separately with respect to d between the limits nj2 and n, we obtain the cir­
cular polarization x of that one of the two photons which has the greater
energy in the laboratory system in the case of annihilation of polarized
positrons in an unpolarized target. But if we integrate the numerator
and the denominator of (33.12) with respect to 6 between the limits 0
and n/2, we shall obtain the circular polarization of the photon, again
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 463

having the greater energy in the laboratory system, in the case when
annihilation of unpolarized positrons takes place in a polarized target.
The quantities x and y are shown in Fig. 42 as functions of the kinetic
energy of the positron in the laboratory system.
1.0
-
0.5
U S' :
-
^ 0
-
-0.5 N
: X
- 1.0 3f,M eV
10-2 10-1 1 10 102 103 104
Fig. 42.

33.3 Annihilation o f a Pair into One Photon


If a positron collides with an electron which is not free, both particles
may be converted, into a single photon (60), (92), (142), (18), (95).
Such a process is determined by the first order matrix element

Sl]lf = ----- f y)+eeihIip_dix, (33.13)


y2co J
where y>_ and xp+ are the electron and the positron wave functions.
If the electron is an atomic A'-electron, then we have
r — i e. t
v_ = y k = (33.14)

where y)k is the wave function of the A'-state, ek is the energy of the
Af-electron and a = %'2’lme2Z .
Insofar as the wave function of the positron is concerned, we must
take for it the exact wave function belonging to the continuous spectrum
in the Coulomb field of the nucleus. This function, as has been explained
in subsection 29.1, must at large r have the asymptotic form of the
sum of a plane wave and an outgoing spherical wave.
On comparing the matrix element (33.13) with the matrix element
for the photoeffect (31.1), it can be easily shown that the process of
one-photon pair annihilation in which we are interested may be regarded
464 QUANTUM ELECTRODYNAMICS

as the inverse photoeffect for the AT-electron accompanied by the emission


of a photon of energy
a) = m —I m2, (33.15)

where I is the ionization energy for the AT-electron, and p + is the positron
momentum.
In contrast to the photoeffect, the final state in this case has a neg­
ative energy and momentum p = —p +- Moreover, the density of final
states will now be given not by pedoJ{2ny, as in the case of the photo­
effect, but by to2do l(2n.y, where doe and doy are the elements of solid
angles containing the momenta of the electron and the photon; finally,
the density of the incident particles will now be equal to the positron
velocity v + instead of the velocity of light, as in the case of the photo­
effect. On taking these changes into account and on carrying out in
(31.16) the substitution y -> —e+/m, co -> e+-\-m, we obtain the following
formula for the cross section for single-photon annihilation of a posi­
tron in the AT-shell (normalized to two AT-electrons):

nr £+ -+P +
o = An:rS In
°(137)» P+(e+y m y m2 3 m 3 p+ m
(33.16)
This formula, like formula (31.16) for the cross section for the
photoeffect, is valid if Z a/v+ << 1.
In the nonrelativistic and the extreme relativistic cases formula
(33.15) assumes the form
4n 2 Z5 p +
where v + •< 1,
3 r°(137)4 m ’
(33.17)
Z5 m
a = Anr\ where e+
037y T ;

We see that the cross section for single-photon annihilation of a


positron is proportional at low energies to the positron velocity, in
contrast to the cross section for two-photon annihilation, which is
inversely proportional to v +. Therefore, at low positron energies single­
photon annihilation is much less probable than two-photon annihilation.
The ratio of the cross sections for single-photon and two-photon anni-
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 465

hilation has a maximum at e+ ~ lOw, and has the value of approxi­


mately 0.2 for Pb.
At low positron velocities and large values of Z formula (33.15)
ceases to be applicable.

33.4 Positronium Decay


The foregoing results referring to the annihilation of a pair into
two photons can be applied to the problem of the decay of positronium,
i.e., a bound system consisting of an electron and a positron.
Obviously, the ground state of positronium is an S-state. Since the
spins of the electron and the positron are each equal to we must
distinguish between the singlet S-state, in which the total spin of the
two particles is equal to zero, and the triplet state in which their total
spin is equal to unity. We denote these states respectively by 1S and 3S.
Positronium in the hS”state is called parapositronium, while positronium
in the 3S state is called orthopositronium.
It was shown in § 7 that a system consisting of two photons cannot
have states of angular momentum equal to unity. Therefore, orthopos­
itronium cannot decay into two photons. Moreover, since the 3S state
of positronium is charge-odd, we can assert that, generally speaking,
the decay of orthopositronium into any even number of photons is
impossible.
On the other hand, parapositronium is a charge-even system and,
therefore, it cannot decay into any arbitrary odd number of photons.
The main processes which determine the positronium lifetime are
two-photon annihilation in the case of parapositronium, and three-
photon annihilation in the case of orthopositronium.
The probability of positronium decay can be related to the prob­
ability of annihilation of a free pair. In order to do this we consider
the wave function for the S-state of positronium

y>(r)= ]/,1Jia3
—e a, (33.18)

where a is the positronium radius equal to twice the radius a0 of the


first Bohr orbit of hydrogen: a = 2a0 = (2h2/me2) (the doubling of
the hydrogen radius is associated with the fact that the reduced mass
466 QUANTUM ELECTRODYNAMICS

of positronium is equal to the electron mass divided by two). The Fourier


component of the wave function ip(r)

Vp = f yj(r) e~ipr cfr = (33.18')

obviously determines the probability amplitude for the electron in the


positronium having the momentum p, while the positron has the mo­
mentum —p. It follows from (33.18') that the wave function ip (r) in­
volves momenta primarily of order of magnitude 1/a which are
considerably smaller than m. Therefore, the decay of positronium can
be treated approximately as the annihilation of a free positron and
a free electron of momenta equal to zero, but of definite spin orienta­
tions.
We denote by and w,, the probabilities of decay of ortho and
parapositronium. Further, we denote by w the probability of decay
of a free pair of particles with zero momenta averaged over the orien­
tations of their spins, and evaluated on the assumption that the particle
density is equal to the particle density in positronium, i.e., |y(0)|2 = 1/na3
(and not l/Q, where Q is the normalizing volume, as is assumed in
evaluating the annihilation probability for a free pair).
Since the relative weights of states of spins 1 and 0 are respectively
equal to 3/4 and 1/4, we obviously have
w = fw j+ jivo. (33.19)
The quantity w can be written in the form
w = 'v2y+ ^ 3 y+ •••» (33.20)
where wny is the probability of the annihilation of a pair into n photons,
averaged over all spin orientations. Since parapositronium can decay
only into an even, and orthopositronium only into an odd number
of photons, it follows from (33.19) that
4 VV0 = W 2y + W 4y + ••• > 4 U ’l = U ’3 + H ’s + ... ,
i.e.,
w0^ 4 £ » V (33.21)
By utilizing formula (33.6) for the cross section for two-photon
pair annihilation for low positron velocities we obtain from the
preceding result:
>v0 = 4|y(0)|2(?;+ a-)I,+^ 0 = \a?m = 0.8 • 1010sec-1. (33.22)
INTERACTION OF ELECTRONS WITH PHOTONS 467

33.5 Three-Photon Decay o f Orthopositronium


We now proceed to the determination of the probability of three-
photon decay of orthopositronium (146).
Three-photon pair annihilation is a third order perturbation theory
process, it corresponds to the six diagrams shown in Fig. 43. The ma-

1*3 ]/r2 i/r3


*' i! !

//- A , “ " AP - \ r~P+ P -\ r-P + P -\

^3 1*1 *2 ]/r3

A ^ A . / ~ A A ~ * A
/- A . P -\ r~P+ P -\ r~P<- P~\
Fig. 43.

trix element for this process can, in accordance with the rules of sub­
section 25.5, be written in the form

eJ A i f i —m
- r , i f i —m i f i ~ m , if2—m,
v [e3— <?i+ei-m^x2i L 3 m*x 2 e.
ii
A

1 / 8 o>1 oj2 (o3 [


A A

f i — m , if3— m i f i — m , ifi — m. i f i —m , i f — m.
*3 + ^3 -T32
tsr

m2x 3 Cl m2x3 m ^~T~


4 e 'i “ m“«4
7„ a ------g 2 T f2 T /" c 3
m‘x5

if'- m , c
e3l m(2ji)4 &2—k3), (33.23)
m2^6

where ki and coi are the momenta and the frequencies of the photons,
f i = fh = P- *i> f \ = f \ — ^3 P+ >
f i —f\ = P- *2 > A ~ A = *i /*+ >
fs — = P— ^3’ ./l) f ^2 P+’
/ ' , 2 _ |_ ^ 2 2
fi+ m ' < = A = 1,2,3,4,5,6.
xn
m2 ’ ^ /rr
468 QUANTUM ELECTRODYNAMICS

In the region of vanishingly small particle velocities |p+| = \p_\ = 0,


f l = —/ 3, / 2 = —f X) / 3' = —f 2and the differential probability for
three-photon pair annihilation is in this case equal to

dw3 = --------------- 1 - ^ \ VQU\ ^{k1Jr k 2-\-k^)d{o)xJrto2-\-co2—2m),


7 6C01 t 0 2 0J3 ( 2 7l)aU

(33.24)
where 1/Q = | ip(0) |2 = 1/na2 and

e- * m2x2 | w 2x2 wj2x6


1j
A A A

A if2—m m L if3—m A A z/5—m A1


■'2 9
m2x. rr
mzx3 e>+<!3“ S S r ' , i

i/i —wi (a / / j - m if\—m


A A

+ ea 2 9 t i+ e i m.2..
^ 4 *2 !
3 m2« rl
= _ V a //„+ m 1a A |
^ \ /i" 1 m 2^ _ 2 " +1 ‘ /,+1 m 2x ^ x "“ Y

/W = 1,2,3.

(e0 ~ e3> e4 = el > f - \ = / ’ll f o = fz) •


The quantity \vQu\2 averaged over the spins of the electron and the
positron, and summed over the photon polarizations, can be written
in the form
1
^ ^vQuj2 SpF,
4 Zj 16m2
Hi. Hi Vl.V%.V3

where

JJj,z > - '”



“ m2xu ' u—1 -V»+i
h H p m Xu-J.

ifh -2 — m
y^+i y u'- i

i f s - i —m _ ) ifu.+m
+ y u--i m2xfi._1 ^/,+1J m2x„ y A ip + m ).

(p, p = 1, 2, 3).
INTERACTION OF ELECTRONS WITH PHOTONS 469

We shall not reproduce here the calculations of SpF, but shall


directly give the results for the differential probability for three-photon
pair annihilation:
a3 dk] dk2dk3
dw 0 *1«2)24 1 «2«3)2 (1 3
4CO] o>2to. { - - ( - + - n l « ) 3 }

X <5(/c1+/c2+A:3)(3(a;1+ a )2+a)3 —2m), (33.25)


where nt is a unit vector in the direction of the photon momentum k v
The total probability of three-photon pair annihilation is equal to
oo co n
— 1 3a3 r c r .
u ’o — —— ~ I d c o i I d(Oo I sin i / i 2 d o ] 2
6 J 2J 12 12 o j3
0 0 0
X (1 —cos#12)2 <5(oi1+ (u 2+a)3 —2nd), (33.26)
where i?12 is the angle between nx and n2 (here the factor 1/6 takes into
account identical states which differ by an interchange of photon mo­
menta). We now evaluate the integral appearing in (33.26)

CO, w 9
do>9 (1 —x)2d/x(5(ft>1+ a>24-a>3—2m),
* = J d 0 h ) (O n
0 o -l
where x = cos#12.
In eliminating the (5-function by integration over x we must keep
in mind that to3 is a function of x so that co3 = j/(of+cof + 2a>1<:o2x.
Therefore
i
J f(x)b[(o(x)+to1+ (o2 2m] dx - ’
-1 °
where x0 is the root of the equation o>3(x) = 2m —(o1— (o2:

x0 = i1-----------
2m
(w j+ w j-m j.
U)] co2
Further, on noting that (d(o3jdx) = ((^1oj2/a>3) and that the absolute
value of the quantity x0 cannot be less than unity we obtain
m rn
d<o2
K = J dco] j ' ((o1—o>2—m)2 (o\(o\ ‘
0 m — ttij
470 QUANTUM ELECTRODYNAMICS

On integrating over co2 we obtain the spectrum of the decay photons:


m
vv3y= j' Fiaj^do)1, (33.27)
0
where

" i) _ 2m jn
a0J \(2tm— aji)2 (2 m —aij)3 m
, 2m — (xix 2(m — c o j m . m — coj. \
+ ------------- 1-------- In -----------}.
co! col m }
The shape of this function is shown in Fig. 44.

(jj/m
Fig. 44.

On carrying out the integration over cot we obtain the total prob­
ability of three-photon pair annihilation

«;3yr = | 071
( f - 9 ) 4(7g. (33-27')

Further, by utilizing expressions (33.21) we obtain the probability for


the decay of orthopositronium
O *2
h1! = —— (ti2 — 9) a = 0.71 • 107 sec”1. (33.28)
y71 a0

33.6. Multiple Photon Production Accompanying the Annihilation


of a Pair
In the preceding subsections we have determined the cross section
for the annihilation of a pair into two photons. This is the smallest
number of photons which can be formed in the annihilation of a free
INTERACTION OF ELECTRONS WITH PHOTONS 471

electron-positron pair. The annihilation of a pair into a larger number


of photons, having an energy of the same order of magnitude, will
be considerably less probable than the annihilation of a pair into two
photons.
However, if in addition to two photons sharing almost the total
energy of the pair there is also emitted a number of long wavelength
photons whose total energy is small compared to the energy of the first
two photons, then the probability of such a process may be compa­
rable to the probability of two-photon annihilation.
We are here dealing with a process analogous to the radiation by
an electron of long wavelength photons for which, as has been shown
in § 30, the ratio of the probability of radiation of two photons to the
probability of radiation of a single photon is determined not by the
constanta but by the quantity f = a In (e/co), where w is the photon
energy, while e is the electron energy. Since for co < e this quantity
is not small, the probability of emission of a number of long wave­
length photons can be comparable with the probability of emission
of a single photon.
For the same reasons, in the case of pair annihilation the probability
of the annihilation of a pair into two hard photons and a number of
soft photons can be comparable with the probability of the fundamental
process—the annihilation of a pair into two hard photons.
We now give an estimate of the order of magnitude of the proba­
bility for such a multiple process of photon production aceompanying
pair annihilation (97).
As a preliminary remark we recall that in accordance with (30.15)
the matrix element for the emission by an electron of a long wave­
length photon is determined by

C(2> = 1) ( P ^ _ _ _ElL\
j/2co ^ '\p jc P ik)'
where 5 ' ^ is the matrix element for the fundamental process (the scat­
tering of an electron by an external field), />! and p2 are the initial and
the final momenta of the electron, k is the momentum and e is the
polarization of the photon.
On repeating the arguments leading to (30.15) we can easily show
that in the case of pair annihilation in which we are interested, the matrix
472 QUANTUM ELECTRODYNAMICS

element M n+2 for the annihilation of a pair into two hard and n soft
photons is related to the matrix element for the two-photon annihila­
tion M 2 by

e,P+ esP -
m 2. (33.29)
M n+2
p + q ,' p~q»

where qs, eos, es are the four-momentum, the energy, and the polar­
ization of the soft photon and p + and p_ are the positron and electron
momenta.
By utilizing this expression we can obtain the cross section for the
annihilation of a pair into « + 2 photons, which turns out to be given by

aC In —-
°n+ 2 (33.30)
— ° 2 '
n!

where cr2 is the cross section for two-photon annihilation, me2, mex
are the upper and the lower limits between which lie the energies of the
soft photon, and Cr*ss (2/te) (ln2y+ —1). (y+m is the positron energy
in the laboratory system. The factor l/n \ arises due to the indistinguish-
ability of the soft photons.)
The quantity ex is determined by the resolving power of the appa­
ratus, while the quantity e2 may be roughly estimated as e2 ~ y +. With
such an estimate we obtain

(33.30')

(r+ > 1 . 1 " - > >) •

In accordance with the results of § 30 this formula shows that the


emission of soft photons obeys the Poisson distribution. The average
number of soft photons emitted is evidently equal to

n = — In— (In 2y , —1). (33.31)


71 Ei ■r
INTERACTION OF ELECTRONS WITH PHOTONS 473

§ 34. The Method of Equivalent Photons

34.1. The Number o f Equivalent Photons


We have seen earlier that the matrix elements for bremsstrahlung
and for Compton scattering have a similar structure. However, a deeper
connection exists between these processes due to the fact that the electro­
magnetic field of a fast uniformly moving charged particle has prop­
erties which are very close to those of the field of a light wave.
We compare the two processes represented by the diagrams of Fig.
45. The diagram of Fig. 45a corresponds to the collision of particle of
momentum q (mass m) with a photon (momentum k), as a result of
which a set of particles of momentum Q is produced. The diagram
of Fig. 456 corresponds to the process of the scattering of a particle q
by another particle (momentum p, mass M), as a result of which the
same set of particles Q is produced while the particle acquires the
momentum p'. The second process is equivalent to the collision of the
particle q with a virtual photon. If the square of the photon momentum
k 2 is small, so that the difference between the virtual and the real photons
is small, the cross section for the process (6) can be expressed in terms
of the cross section for the process (a).

Fig. 45.
On denoting the invariant amplitudes for these processes respec­
tively by A r and A we can, in accordance with (26.6), write the
differential cross section which we denote by doT and da in the form
m2

M 2m2 dip'6(p'2-\-Mz)
d o — \A\ 6q (27iy
474 QUANTUM ELECTRODYNAMICS

where

N is the number of particles in the set Q, qa are the momenta, ma are


the masses of the particles produced. We now show that under certain
conditions the amplitude A can be expressed in terms of A r (180).
From the diagrams of Fig. 45 it follows that A r = e^J^ (0) and
A — jpJpik2). Here eMis the photon polarization vector, J^(k2) describes
the set q + k -> Q, with the same vector J appearing in the expressions
for both A t and A, but taken for different values of k 2 respectively equal
to 0 and (p—p')z. / describes the vertex for the process of emission of
a photon by the particle p (p -> p '+ k ); if the particle p is an electron,
then j fi = u(p')yu(p).
The quantities j and J^ satisfy the condition of transversality
j fik fl = 0 and J k = 0. We consider the process in the rest system of
the particle q(q = 0).
If the velocity of the particle p is large {\p\ > M) and p' differs but
little from p, then we have j fi = 2Zeu^, where Ze is the charge of the
particle, while u = p j M is the velocity four-vector

From the transversality conditions we can express the frequency of the


virtual photon co in terms of its momentum k: co = kv = k {]v. Also
k 2 = (1 —v2) <x>2-\-k\ and k L = k —k y.
We require that \k2\ be small, \k2\ -< in2. Then we have kx << m and
co < v 2).

We see that under these conditions the case is possible for which

k L •— - oj ] / l —v2^ c o , kL >> co(1 —v2) (34.1)

and which we now discuss. On expressing by means of the transversality


conditions J0 in terms of J, J0 = (l/co)./fc, we obtain

= — 7 = = z( - M i+ - V w0 - ® 2))-
coy 1—v2
INTERACTION OF ELECTRONS WITH PHOTONS 475

On taking (34.1) into account we can neglect the second term in this
expression. Moreover, since k L << k^, JL differs from Je, where e is
the unit vector perpendicular to k (ek = 0), by a quantity of the
order of magnitude ]/l —v2. Thus, for small k2 we have

= 2Ze, - ^ m e .
j / 1—v 2 CO

On substituting this into the expression for da and on comparing


it with the expression for daT, we obtain
4Z2e2k \M d*p'b{p'2+ M 2)
da (34.2)
] / l - v 2coki (2t)3

Thus, we express da in terms of da,. Formula (34.2) is valid if the con­


ditions (34.1) are satisfied.
We can express the quantities appearing in (34.2) solely in terms
of invariants characterizing the process. In order to do this we utilize
the relations
k2 = k\-\-{\ —v2)co2,

Q2 — {q-\-k)2 = k 2—m2—2com\ 1 _ E pq
y/ l - v 2 ~ M Mm ’

71
J ' dip'd(p’2jrm 2) dk2dQ2
4 j./ {pq)2—M 2m2

(the last relation can be easily obtained, for example, by going over
to the coordinate system in which p = 0) and obtain

* = O f rfe. (34.3)
71 \Q2-\-m2\ki

The conditions for the validity of this formula are


^ 2< m 2; \Q2\ < \ q p \; \Q2+ m 2\ > k 2.

We obtain a formula which is simple and convenient for applications


by choosing for our variables kL and co. We then have
dQ2dk2 = 4m kLdcodkL
476 QUANTUM ELECTRODYNAMICS

and
2Z2a dw k \ dk±
d a -d a T ^ w (34.4)

We denote by do’ the cross section for the process integrated over all
to and k ± (i.e., over the recoil momenta p'). It can be written as

where
(34.5)

We can interpret this formula in the following manner. In-order to


obtain the cross section of some process due to a fast particle acting
on a particle at rest (q = 0 in our coordinate system) we can replace
the fast particle by the spectrum of equivalent photons n(co)da>, and if we
know the cross section for the process due to photons, it is then sufficient
to multiply it by the number of equivalent photons in the frequency
interval dco and to integrate over the frequencies. This method is referred
to as the method of equivalent photons, or the Weizsacker-Williams
method (206), (212). The underlying idea was originally proposed and
applied by Fermi (59).
The equivalent photon spectrum can also be obtained from a classical
investigation of the electromagnetic field of a rapidly moving particle.
The four-potential of the field produced by a uniformly moving
particle is determined by
471
C A = ---- — evh{r—vt), D A 0 — —47ied(r—vt),

where v is the velocity of the particle (we employ both here and in
subsequent formulas of this subsection the usual CGS system of units).
On expanding the potentials into Fourier integrals we obtain the follow­
ing expressions for the fields (118):
{kv)v
( 3 4 .6 )
INTERACTION OF ELECTRONS WITH PHOTONS 477

where the x-axis is directed parallel to v .


On making the substitution kx -» £kx, ky ->£ky, kz £kz, where
f = 1 j / l ~ ( v 2/c2), we rewrite (34.6) as
ie£ [ k 1
E±(r, 0 = 2ji 2 .1 k2
ie f h
E\\(r, t) = (34.6')
2n2 „ k2

i/(r, 0 = — ,E 5
c
where
v (k v )
kx = k 2 ’
k« = k —k.
V

and Ej_ and E^ denote the perpendicular and the parallel components
of E (with respect to the direction of the velocity v) k x = |A|||.
We assume that £ > 1. In this case, obviously, IZTJ > |£ |||, i.e.,
the field remains practically transverse, just as in the case of a light
wave. The formula (34.6') determine the expansion of this field into plane
monochromatic waves of frequencies a>= kv1£ = k l c£ which are
propagated in the direction of v .
We now determine themumber of equivalent photons corresponding
to these waves. In order to do this we obtain the total flux of electromag­
netic energy in the direction of v :
OO

s -
— OO

where d f = dy dz = 2nb db is an element of area in the yz plane,


b = } / y 2-\-z2 (this quantity is known as the impact parameter). On
equating the total flux S to the energy of equivalent photons
OC

S = J hcon(oj)da>,
o
we obtain the number of photons n((o)da> in the frequency interval
to to co+c/co.
478 QUANTUM ELECTRODYNAMICS

On substituting (34.6') into the expression for S we obtain


cxj

= £ /* '■ / d,^ + E -)
oo
c Z 2e2£2 , ( - k yk'y- k j t ' z)
dy I dz dkdk
4n 4 or4
— OO
/
— OO
d‘ f k 2k '2

X +fcP +iz(-kz+ ~ (ki+kP


Z 2e2£ r k j+ k l Z 2e2f r k2
dk
2n 2 v J k4 2n2 ~J (k f+ k f) 2
On noting that dk — 2nkL dkL dkx = 2jikL dk±(d(o/c£) and on com­
paring this with the expression for S we obtain formula (34.5).
The upper limit of the integral in (34.5) can be determined from the
condition k± < m . Since n(co) depends on /rmax, logarithmically, we can
set ^max= r\m-> where r] is a quantity of the order of magnitude
unity. On carrying out the integration we obtain

n ( ro) d m = -- a Z 2— In^ . (34.7)


71 CO CO

Formulas (34.4), (34.7) are evidently valid for


m
oo < — (34.8)
)/1 — V2
]/l
The fact that the integral for n(co) diverges at large co has a simple physical
meaning: it is related to the divergence of the expression for the energy
flux S for small values of the impact parameter b.
The lower limit for b, and consequently, the upper limit for k L,
can be evaluated in the following manner. In quantum mechanics in
order to be able to speak of a definite value of the impact parameter
b it is obviously necessary that b should be considerably greater than
the transverse dimensions Ay of the wave packet associated with the
particle: b ^>Ay.
On the other hand, according to the uncertainty principle Ay > hlApy
where Apy is the uncertainty in the transverse component of the momen­
tum of the particle. The transverse component of the velocity of the
INTERACTION OF ELECTRONS WITH PHOTONS 479

particle vy must be sufficiently small so that the transverse displacement


of the particle during the collision time, which in order of magnitude
is equal to T ~ b / v ^ b / c , will be small in comparison with b, vy(b/c)<^ b,
i.e., the transverse component of the velocity must be small compared
to the velocity of light and, consequently, Apy <* me, where m is the
particle mass.
Thus, from the uncertainty relation it follows that Ay >> k/me.
Therefore, the impact parameter must be considerably larger than
h/mc.
In discussing the effect of the nuclear field on the electron we must
obviously interpret m as the mass of the lighter particle, i.e., of the
electron. We assume that in this case the minimum value of the impact
parameter is given by bmin = (1 /r])(h/mc), where in order of magnitude
rj is equal to unity. The quantity bmin corresponds to the maximum
value of k±, which is given by k max = rj(mcjh).
34.2. Bremsst rah lung from a Fast Electron in the Field of a Nucleus
By utilizing the method of equivalent photons we shall now obtain
the bremsstrahlung cross section for the collision of a fast electron with
a nucleus. In the coordinate system K*, in which the electron is at rest
before the collision, this process can be treated as the Compton scattering
of the electron by the spectrum of the equivalent photons of the nuclear
field.
The cross section for the scattering of the photon by the electron
is given in the coordinate system K* by formula (28.18)
CO„
+ —sin2#* do*
0), CO CO ,

where co* and co£ are the frequencies of the primary and the scattered
photons in the coordinate system K*, which are related by expression
(28.6) cor
to* =
1 + ( c o r / w ) (1 — C O S 0 * )

and d* is the scattering angle. The bremsstrahlung cross section in the


coordinate system K* is related to the scattering cross section by
wi max

dar*ttd« ) = f n{(ol)doildalMi{<ol,iol), (34.9)


w l m in
480 QUANTUM ELECTRODYNAMICS

where (o*min and <w*max are the minimum and the maximum frequencies
of the primary photon corresponding to a given frequency co% of the
scattered photon.
Since the scattering cross section is relativistically invariant, in order
to determine dcrrad in the coordinate system K we only need to express
the cross section, determined by formula (34.9), in terms of the frequency
co2 in the coordinate system K, which corresponds to the frequency cd$
in the coordinate system K*. The relation between the two frequencies
is given by the Doppler formula co2 = co*(l—v cos0*)f.
On introducing = a)*/m, £2 = ojjmt; and on noting that
2Z*a dtj rjl
n(cof)dco* y
n fi fi

cos 0* -

^catt = nrS-
nr% {' fs+i- f ,+ {.(i-w [i,(i-«' 2]} '
we rewrite dcrrnA in the form
2^2
d<7M (.m J = 2 r S Z * a d i, f

2( 1 - f 2)

( 3 4 '9 '}

where the upper and lower limits of integration over ^ correspond to


d* = 0 and n.
The principal contribution to the integral (34.9') obviously comes
from the values of near the lower limit. On carrying out the calcula­
tion with logarithmic accuracy, i.e., on assuming that In ( ^ f/^ ) is a large
quantity, and on taking it outside the integral sign, after setting it
equal to its value at ^ = f 2/2(l —f 2), we obtain

^ r „ , K ) = 4r02Z 2 a - l - 2| ^ + ' A _ l ] i n ^ i £ l f L , (34.10)


co 2 \ e2 ex 3/ mo)

where £x and £2 are the values of the electron energy in the initial and the
final states.
INTERACTION OF ELECTRONS WITH PHOTONS 481

Formula (34.10) is valid if £2 ;> m, e 2 > m. As should have been


expected, it coincides with formula (29.25).
34.3. Radiation Emitted in an Electron-Electron Collision
By utilizing the method of equivalent photons we shall now obtain
the radiation emitted in an electron-electron collision.
If a fast electron, which we shall refer to as electron / collides with
an electron at rest, which we shall refer to as electron r, then in order
to determine the cross section for the resultant bremsstrahlung by the
method of equivalent photons we must, in the first place, investigate
the scattering of the equivalent photons of the field of the fast electron
/ b y the stationary electron rand, secondly, on going over to the coordi­
nate system associated with electron / investigate the scattering by
electron / of the equivalent photons of the field due to electron r.
The second of these two processes does not differ from the process
discussed earlier of the scattering of the equivalent photons of the
nuclear field by an electron in its own rest system—the process to which
the bremsstrahlung of an electron in the nuclear field may be reduced.
We now show that the contribution made by the first process to the
cross section for bremsstrahlung accompanying an electron-electron
collision is considerably smaller than the contribution made by the
second process.
On assuming that the photon energies and o>2 before and after
scattering in the first process are considerably greater than m, we can
start from the following formula for the Compton scattering cross
section:
, , dto.
do scaU = nr\m —

(Here we have not included the term containing sin 0, which may be
neglected since in the relativistic domain scattering takes place primarily
into small angles O ^m jco .)
By utilizing expression (34.7) for the number of equivalent photons
we can represent the contribution of the first process to the cross
section for the emission of radiation in the form
1m a
d(»i , e CO O CO ,
do]aJ = 2r\amdoj2 (34.11)
f
o>i m in
3 ,n —
CO? CO,
— + —
O)2
482 QUANTUM ELECTRODYNAMICS

where e is the energy of electron / and colm in and e o lm a x are the limits
between which cox lies for a given frequency u>2. These limits can be
obtained from _ co2
a>l 1—(a>2/m) (1 —cos 6)
and are obviously equal to colmin — c o 2 and a,imax = °°- Although
the method is valid only for o j 1 < < e , nevertheless since the integrand
in (34.11) falls off rapidly with increasing col5 the upper limit in (34.11)
may be taken to be infinite.
The principal role in the integral (34.11) is played by the values
of cox near the lower limit. On carrying out the integration with logarith­
mic accuracy, just as in (34.9') we obtain
d(yl^(co2) = — r2
0am— \n-^-. (34.12)
3 CO2 CO2

On setting Z = 1 in formula (34.10) for the cross section for the


bremsstrahlung emitted by an electron, we obtain the contribution of
the second of the two processes described earlier to the cross section
for the radiation emitted in a collision of two electrons
d c 0 n E<y / £ i 2Ci £2
<Kad(W2) = 4r0a CO, (34.12')
3/ rruo2
Comparison of do\&i and d a shows that da\&A {mlaio) dcr"d < da\\A.
This relation is associated with the fact that in both integrals (34.9)
and (34.11) the principal contributions come from values of the varia­
bles near the lower limits, while the lower limit in (34.19), which is equal
to ma)2/2e, is considerably smaller than the lower limit in (34.11), which
is equal to a>2.

34.4. Pair Production by a Photon in the Field of a Nucleus


We now apply the method of equivalent photons to the evaluation
of the cross section for pair production. We start by determining the
cross section for pair production by a photon in the field of a nucleus.
We assume that a photon of energy to = £m, £ > 1 is incident
on a nucleus at rest in the coordinate system K. We go over to the
coordinate system K', in which the nucleus is moving towards the
photon with a velocity v equal to
INTERACTION OF ELECTRONS WITH PHOTONS 483

In this coordinate system the photon energy is equal to


1 —v
(0 = (0 pw m.
I ' 1— V 2

The field of the nucleus in the system K' can be represented in the form
of a superposition of pseudophotons, whose spectrum is determined
in accordance with (34.7) by

n ( f o 0) d < o 0 = — a Z In L — ) .
71 co0 \ w0}
The process of pair production by a photon in the field of a nucleus
can be treated as pair production in the collision of the incident photon
with the photons of this spectrum. In other words, the cross section
for pair production by a photon in the field of a nucleus is equal to

a f n (OJq) d(O0G y y , (34.13)


where ayy is the cross section for pair production by two photons given
by formula (32.27') into which in place of x we must substitute
m2 m
x=
O)'(O0 oj0 ‘

For the limits of integration in (34.13) we must take m and £m. On


introducing in place of a>0 the variable x we rewrite (34.13) in the form

l / ‘T
l+x!
a = 2r2a Z 2 J d x-x In [??£(1 —x2)]^(3—x4) In y - ~j + 2 x (x 2—2)^.

On assuming that In £ > 1, and on retaining only the main terms


(of order In £) we obtain after integration

< 7 = ^ a Z V > /£ , /~1. (34.14)

This expression coincides with formula (32.9).

34.5. Pair Production in a Collision o f Two Fast Particles


We shall now also investigate the process of pair production in a
collision of two fast charged particles. We assume that the relative
velocity of the particles v is sufficiently large, so that £ = (1/j/l —v2) > F
484 QUANTUM ELECTRODYNAMICS

In the coordinate system Kt , in which one of the particles A 1 is at


rest, the distribution of photons representing the field of the second
particle A 2 is determined by
day
nA(o)doj = —aZ\ In (r i- ^ -
71 \ co CO

where Z 2e is the charge of the particle A 2.


The process of pair production in the collision of particles A 1 and A 2
can be regarded as pair production in the collision of the photons
of this spectrum with the stationary particle Ax. The cross section
for pair production in the collision of a photon with a particle is, in
accordance with (34.14), given by

°V(°J) ^ — r\aZ\ In [rj' — J , rj' ~ 1,

where co is the energy of the pair and Z xe is the charge of the particle A 1.
Therefore, the cross section for pair production in the collision of par­
ticles A ± and A 2 in which we are interested can be represented in the
form
o = J ov(oS)n2(oS)doy. (34.15)
The maximum frequency of the photons of the field due to particle A 2
is equal in order of magnitude to fm. Therefore, as limits in this inte­
gral we must take 2m and tjm:

/ A aZ|ln dco
a — rlaZl In L '
co co
2m
28 r2
- 0 a2Z fZ .f(ln/O 3, /~1. (34.16)
Z / 71

This formula coincides with formula (32.34).

§ 35. Scattering of a Photon by a Bound Electron. Emission of Two


Photons

35.1. The Dispersion Formula


In § 28 we studied the scattering of photons by free electrons. We
now go over to the study of the scattering of photons by bound elec­
trons.
INTERACTION OF ELECTRONS WITH PHOTONS 485

For the determination of the scattering cross section in this case


we can, as we have done previously, utilize the part of the scattering
matrix S<2) defined by the general formula (28.1), but now in place
of the function Sc(x1—x 2) the expression for S<2) must contain the
Green’s function for the Dirac equation for the electron in the external
field e) (*!, x 2):
S i2) = e2f N lyjC xJA ixJSyK xi. x 2) A(x2)ip(x2)} dix 1d ix 2. (35.1)
The function S y ){x1, x 2) is defined in the following manner:

h > h,
Sl'KXl • Xi) =
n (35.2)

where
V>nl +)(x) = y>nl +)(r)e and ip ^ ix ) Wn \ r) e n
are the positive and negative frequency solutions of the corresponding
Dirac equation.
The electron Green’s function S^e)(x1,x2) can be written in a more
compact form if we note that
0icot a < 0,
dco o—iat a > 0, t > 0,
2ni a(l —/0)-)-a)
-ia t a < 0, t < 0.
By utilizing these formulas we can easily show that

Vn(r l)Wn(rd (35.3)


d(Dt*alt i-(2> ^
En( i - i o y ■CO

where the summation over n is taken over all states corresponding


both to positive and to negative frequencies.
From formula (35.3) and from the Dirac equation it follows that
S^e)(x1} x 2) satisfies

■ieAjf](x) +w>6'c(e)(x, x') = —id(x—x'), (35.4)


1
dx„
4

and since S c(x) satisfies


486 QUANTUM ELECTRODYNAMICS

the function S ci ei(x1, x 2) satisfies the integral equation


S ie)(x1, x 2) = S c(xl , x 2) — f Sc(x1, x 3) eA(x3)Sj.e)(x3, x 2) d % . (35.5)
We now apply these formulas to the determination of the cross
section for the scattering of a photon by bound electron.
We consider the scattering of a photon of frequency cot, propa­
gation vector k t, and polarization et by an electron in the state
y>i(x) = yii(r)exp(—iEit). As a result of scattering, a photon appears of
frequency cof , propagation vector kf , polarization ef , while the
electron goes over into the state y)f (x) — yf (r) exp(—iEf t).
The frequency of the scattered photon o)f may either coincide with
the frequency of the primary photon a> o r differ from it (in the latter
case we speak of Raman scattering). On noting that photon states
under consideration correspond to the potentials

A.ix) = At(r) e~i<u


|/2cof

Af (x) = Af (r) ei<ad =


y2ojf
we obtain in accordance with (35.1) the following expression for the
matrix element for this transition:

S& = d * x i j d ix2{y)f (r1) e ' lkfr'ef S c{e)(.x±, x 2)


2 ]/ a)ia)f
Xc,e*,r> t(r2) e ^ ' f 1 V";)
+ y)f (rl)elkirieiS ci e:)(x1, x 2)e,e^lkfr- ^ ( r ^ elt^Er m0
(35.6)
Af (X, ) A fa ) A f a ) A f (xz )

'i k -
S(c t^i ,x2)*V /"S^fx^Xj)

Fig. 46.

The two terms in this formula correspond to the diagrams shown


in Fig. 46. The difference between these diagrams and the diagrams
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 487

representing the scattering of a photon by a free electron consists of the


fact that in Fig. 46 the external electron lines represent bound states,
while the internal electron line corresponds to the function S ci e)(x1, x 2),
and not to —x2) .
On substituting into (35.6) expression (35.3) for S^e)(xlt x 2), we
write S ^ f in the form
CO OO oo

S,(2!f = ~ -- V f dt1 I dt2 f do (fre/ e-an ^ ) ( n \ e {e,,‘‘r\i)


, y (oiwf ^ J .1 J En-\-(o
— oo — oo — oo

y ' g :(l(£/ ' f e tt2(Ei - r a}i + ,’>)

'i/t/T
+ (f\e je kir'n)(n\ef e tk' r\i)
(35.7)
En+ co
where

(m\eelkr\n) = Jy)m(r)eeikry n(r) dr. (35.7')

On carrying out the integration over tx, t2 and co, we obtain

S& = — d (Et + iOi — Ef ~u)f), (35.8)


where
l ( f \ e f e~ik' r \n)(n\eteik'r\i)
1 <"|"V T [ En- E - o , t

> (35.8')
En —Ef+Oli J

and the summation is taken over all the states n corresponding both
to the positive and to the negative frequencies.
The most important application of this formula is to the scattering
of a photon by an atomic system.
We shall first of all consider the nonrelativistic case when the photon
energies are small compared to the electron rest energy (ot < m,
cof -< m. In this case we can assume that the electron energies both
in the initial and the final states E{and ^ d iffer little from m; \El—m\ < m
and \Ef —m\ < m. Moreover, we assume that in the sum (35.8') an
488 QUANTUM ELECTRODYNAMICS

important role is played only by those states n (“intermediate” states)


for which the energies are also nonrelativistic, i.e., | \En\—m \ •< m.
These assumptions enable us to simplify considerably the part
of the sum Ut^ f in (35.8') which refers to negative frequencies. First
of all, in each term of this part of the sum, which we denote by U\~\,
we can replace the denominator by 2m:

U \ ~ \ = ---- {(f\ef e 'k/r\«(+))(« ( >1e1e*,r |0


m (->
+ ( f\e ielkir\n)(n\ef e th' r \i)},

and after this we can transform U\~}f into a sum over all states, includ­
ing not only negative, but also positive frequencies. In order to do
this we must in the expression

(f\e f e ihfr\n{ >) = J y>f (r)ef e


make the substitution
m —H
Vn~](r)
2m v4 '(r).

where H is the Hamiltonian for the electron. In the approximation


in which we are interested we have

H >=» m yi and = i m y (J:|(r).


Therefore
m —H . . . . , w \ m —H . , w ,
~2m~'Pn ^ = Wn ~ 2 n T Vn ^ = 0

and
1 y 4 — iKsr
ikfr I | / -a ifc iri .\
e 1 \n I (n \ei e 1 \ i)
m y o>i(of n

%1__y ± j kir n j («| ef e ik/ i o | .


+ f 1 2
where the sum is taken over both signs of the frequency. On utilizing
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 489

the rule for matrix multiplication and the definition (3 5 .7 ') we can
rewrite U\~?f in the form

u u = - ^ = { l f y 4 1
------------------ .... £ 3 t *i)
m y 0 ) ^ , {\ e' ^ 2 - e

Finally, on noting that


y*e = - e y t ,
^ + ^ , = 2 ^
and that in our approximation we obtain

//(-) — 2na
u i~*r — = eter j V*(r)Vi(r)ellki k,)rdr. (3 5 .9 )
mj/cOjO)
We now consider the part of the sum U{^ s which refers to positive
frequencies. In the nonrelativistic approximation in the numerators
of the fractions appearing in this part of Ut^ f we can, in accordance
with subsection 15.1, replace a A by {\/m)Ap-\-^ curl A and ae by
(\lm)ep-\-iy,[k,e], where p, = (efi/2mc)a is the spin magnetic moment
of the electron. On carrying out this replacement, and on utilizing
formula (35.9) for £//“} we obtain the following expression for

u t.
4™ M yf (
p'
4 ?
),n
(p
'4 X
t
m lm n L +
(.P'^i)fn(P' Af)ni
(35.10)
E - E n- ( o f
where p = p+\e[aV] and the parentheses denote matrix elements
evaluated using nonrelativistic wave functions:
( L Xs- = / <P?Ecpt. dr.
We note that the matrix element of the form (35.10) can also be
obtained in a different manner, viz., if from the outset we utilize the
nonrelativistic operator for the energy of interaction of the electron
with the electromagnetic field (cf. (15.3)):

V = ------ p'A A 2, p = />+ y [oV].


m 2m
490 Q U A N T U M E L E C T R O D Y N A M IC S

The term quadratic in A in this operator determines the “direct transi­


tions” and in the first approximation of perturbation theory leads to
the last term in (35.10), while the linear term determines transitions
via the “intermediate” states, and in the second approximation of
perturbation theory leads to the sum in (35.10)
If the wavelength of the scattered photon is large compared to the
atomic dimensions, then we can take A t(r) and Af {r) to be constant,
since exp (iki f r) Qxp(iklfR) (R is the radius-vector of the center
of the atom). In this case U ^ f assumes the form

2jiae ikl kf)R M y T( p e ^ J p e X t 1, ,\


u t. m]/a7~oif \_ E —En+ iot ^ E - E n- o j f J 1 ! i{\ '
(35.11)
Here the last term differs from zero only in the case of coherent scat­
tering when coi = cof .
Formula (35.11) defines the dispersion law, i.e., the dependence
of the scattering ability of the atom on the frequency of the light. It can
be represented in a somewhat different form, viz., in such a way that
instead of the matrix elements of the momentum Ui^ ! will contain
matrix elements of the atomic dipole moment. In order to do this we
must utilize the relation
( p ) s S- = »n(Es- E s.)(r)ss,,

and also the commutation relations between the operators for the
momentum and the coordinate, whence it follows that
elef 6i f = i[(pet) (re,)-(re,) (/>«,)]„ = i ]T [(/«!,),„(rc/)nl- ( r c /)/fI(Jpift) J .
n

Further, on adding inside the figure bracket in (35.11) the zero term
0 = cof [(reJ (ref) —(ref) {rei)]fi = cof [(rc,)/n (ref)ni ~ (ref)fn( r ^ J ,
n
we finally obtain

— 2nae y (35.12)
n LE > ^ 0>i ^ (of J
The differentia] scattering cross section is related to tf. by
dkf
da = 2ti\ Ut^ f \2 (5(£i+coj—Ef ~cof)
(hi)3 '
I N T E R A C T I O N O F E L E C T R O N S W IT H P H O T O N S 491

On eliminating the 5-function by integration over «jf we obtain


2
da — {Qe_f) ni (ggt)/n(gg/)m co,(o3,do
E - E n+ toi + (35.13)
Et En (of
where Q = er is the atomic dipole moment.
If the scattering system contains not one, but several, electrons,
then we must interpret Q as the total dipole moment for the system.
Formula (35.13) holds if the photon wavelength %is large compared
to the atomic dimensions a, k a.

35.2. Resonance Scattering


The dispersion formula (35.13) contains a sum over all the excited
states of the atom. If the photon frequency is equal to the difference
between the energies of one of the excited states and the ground state
of the atom, i.e., if oii = En— £t, the scattering cross section becomes
infinite, which shows that the foregoing formula is not valid for
(oi = E n—Ei. This case is known as resonance.
We shall not develop here a rigorous theory of resonance, but
shall confine ourselves merely to explaining the physical reason for the
inapplicability of formula (35.13) near resonance, and we shall indicate
an approximate method for discussing the phenomenon of resonance.
A consistent discussion of resonance scattering utilizing the general
theory of radiation corrections will be given in subsection 53.4.
The reason for the inapplicability of formula (35.13) near resonance
lies in the fact that we have treated y>n(x) as stationary state wave func­
tions, whose time dependence is given by exp ( —iEnt).
But the excited states are only approximately stationary, since, be­
cause of the interaction with the electromagnetic field, there exists a finite
probability for the transition of the atom to the ground state (this
probability was found in § 27).
Stationary states can be described approximately (cf., for example,
(119)) as states of complex energy whose wave functions have the fol­
lowing time dependence:
Un - -1r n)(
Wn e
where r is a positive real quantity. The probability of finding the atom
in an excited state is proportional to |^ „ |2, and thus decays according
492 Q U A N T U M E L E C T R O D Y N A M IC S

to exp (—+ + ). This shows that Fn = w,, where wn is the emission prob­
ability determined by formulas (27.14), (27.22).
At frequencies close to the resonance frequency we may neglect
in formula (35.13) all the terms except for the resonance term and
replace in it En by En—i/2rn. Thus, we obtain in place of (35.12) the
following expression for the scattering matrix element:

______ H ( » * / ) / , ( r e t)s i

u ~ =
where the sum is taken over all the states of energy En. The correspond­
ing differential scattering section is given by

IT ( e ^ v ( e « , ) » r
5
do = cofijf d o f (35.14)
(£■„-£,-a>,)2+ ( l / 4 ) /^

In order to obtain the total scattering cross section we must integrate


expression (35.14) over the angles, average it over the polarizations
of the incident photon, and over the angular momentum components
of the initial state, and sum it over the polarizations of the scattered
photon, and over the angular momentum components of the final
states. Such a cross section is determined, in addition to £,, En and co,
only by the level width r n and by the values of the angular momentum.
In the case of coherent scattering the total cross section is equal to
27n+l r \ ________
(35.15)
2(2/,+1) CO2 ( £ n - E - c o y + ( r * J 4 p
where j n and j\ are the angular momenta of the system in the states n
and i and co~ co(~ a>r This formula preserves its form also in the
case when the state n differs in an arbitrary way from the state i with
respect to its angular momentum and parity. In this case r n = co„
must be interpreted as the emission probability for the corresponding
multipole radiation.

35.3. Compton Scattering by Bound Electrons


We now investigate the scattering of an energetic photon by a bound
electron on the assumption that the photon energy is considerably
greater than the electron binding energy.
IN T E R A C T IO N O F E L E C T R O N S W IT H P H O T O N S 493

We denote by 0 the state vector describing one bound electron


in the state yv = %{r) exp(—iEtt) of energy Et and polarization cft, and
one photon of energy ico(, momentum ki and polarization ei . Since
it has been assumed that co, > m we may neglect the electron binding
energy and approximately write 0 as a superposition of states xk{-p>a,
describing the same photon and a “free” electron of momentum p
and polarization a, whose energy, however, is equal not to j/ m2 -\-p2,
but to the energy of the bound electron Et : 0 = C a%k.. a.
pa ’ *’

The expansion coefficients C are obviously equal to

c P .a = (Xki:p.a &) = f Wi(r)ua(p)* e ~ i p r dr = (p)* ,

where yf* is the Fourier component of the bound state wave function
V>iO )
y>ap = J y i(r)e~lPrdr
and ua{p) exp(//jf) is the orthonormal system of solutions of the
Dirac equation for free electron satisfying the normalization conditions
“a(p)ua'(p)* = 5aa,\
We are interested in the transition to the state in which there exists
one photon of energy a>f , momentum k{ and polarization e{ , and a free
electron of momentum p f , energy ef and polarization ar The matrix
element for this transition is evidently equal to

(Xk/'Pf•<>{' ^p.aCp,o(Xk/;pfOf >SXki-.p.o)'

where Xkf\P/,af t^le state vector describing the state containing one
photon of momentum kf , polarization en and one free electron of energy
ef = ]/m2jt^Pj-, momentum pf and polarization ar
The element of the scattering matrix (Xkf;Pf.af > s Zfci:P)Cr) between
the free states £,;/>, ct and kf ;pr a. (but without the usual relation
between the electron energy in the initial state, and the momentum p)
can be obtained with the aid of formula (35.7), in which we interpret
the states y>n(x) as the free eleotron states = ua‘( p ) e ip'x, where
the values a ' — I, 2 correspond to the two polarization states of
different spin orientation and of positive frequency, and the values
a' — 3, 4 correspond to the two polarization states of negative fre­
quency.
494 Q U A N T U M E L E C T R O D Y N A M IC S

In the case under discussion the quantities ( f \ e le !fc‘r | n) have a


simple form, for example:
( f \ e ie~ll‘i r \n) = — i e , f u f (pf)*au° ( p ' ) e r(kr P ~Pf) dr
= — ;'(2jr)3(5(/ri+ / 7 '—/7/ )w°jr(/7/) * a e i w'7 ( p ) .

Summation over n in (35.8') reduces to integration over p ', and to


summation over ex'= 1 , 2 , 3,4. Integration over p' can be carried
out immediately due to the presence of a <5-function, and leads to the
formula

(Xkf •pf^r SXkt-.p.a) ----+ 5 (£,+£«>,—£,—CO,)


2 | ojto>f
*
V (ua/(Pr) o.ef ua'(ki+ p)ua'(ki+ p ) a e iua(p)
Zj Ea\ k iJrP)—(oi — Ei

uaf(pf )*aeiua'(p—kf )ua'(p—kl)ctef ua(p)\


E ° '( p ~ k f)-{-(of - E i }‘
On noting that

E ^i(p)Wa(P)K(PT = v>apl(p)>
we obtain

S{Vf = - ,e2(- ~ d(El+ e o - e f- (0 f)BaiV, (35.16)


2) (ot(of
where
i ( uaf(pf)*aefua‘(kf+ p f)ua'(kf+ p f)aeiy)ai(kf+ p
Baiaf — V
Z, I E"\kr \ p f ) - o ) - E l
o '—I ^

uaf(pl)*aef u”'(pf- k i)u"(pf - k t)a.ef yP'(krJr p f - k l) \


+ " E ' i p f - k ^ + 'O'-Et {'j
The transition probability per unit time is equal to

d\v = ---------- — O r ) | B aiar 12 dp, dkf


4rUjCo,
e4(27r)5 (o
—- (XEi-\-o)i—ef —(of) 1Baiaf\2 dpf do d<of
OX

On dividing dw by the flux density of the incident photons, which is


equal to their velocity, and on summing over final, and on averaging
IN T E R A C T IO N O F E L E C T R O N S W IT H P H O T O N S 495

over the initial polarizations of the particles, we obtain the averaged


scattering cross section
, e4(2rr)5 to, JL -------
d a = - ^ - - ,- - i - d ( E l+u}l- E f -cof) Y | Bnia/\2dpf do dojf , (35.17)
afar = 1
where the bar denotes averaging over all polarizations.
On eliminating the (5-function by integration over the angle between
k{ and k{, we obtain the following expression for the averaged cross
section expressed per unit energy of the scattered photon:

d° = ~ o j2
1 ( f E/ ^
Oj.Of =1 ! ^ I 2 d p ) d (° f
/ ( 3 5 ‘ 1r )

In order to obtain this formula it is necessary to transform from


the variables p f , o); to the variables p i = k1— kf—p, and <5w
= — E{ - \ - ( o f — Ei
— to; and to evaluate the Jacobian for this trans­
formation for <5w = 0.
Since the photon energy is assumed to be large in comparison with m,
the principal role w^ll be played by small scattering angles Therefore
2 _____
in the evaluation of \Baiaf\2 we can set d = 0.
avaf =i
On selecting for y)t the wave function of the ground state of the
electron in the Coulomb field of the nucleus (cf. § 13) and on expanding
da into a power series in Za, we obtain (38)

da = nrl + - ^ ) j 1+ ] (Z«)2+ ^ ( Z a ) 4+ 0(l)(Z a)5


o j f \ (Oj °>j-1 { 2 24

+ In--—(l -)-0(Z2a2))| dcof . (35.18)


n m ' )

35.4. Emission of Two Photons. The Metastahle 2si State of the Hy­
drogen Atom
The matrix S(2) defined by formula (35.1) can also be applied to
the problem of the simultaneous emission of two photons. The matrix
element corresponding to this process can be easily obtained from the
scattering matrix element S ^ ; defined by formulas (35.11) and (35.12),
if we replace in (35.12) the potentials At and Af by A* and A2, and the
496 Q U A N T U M E L E C T R O D Y N A M IC S

photon momenta kt and kf by A:, and k 2■We thus obtain in the dipole
approximation the following expression for the matrix element for the
emission of two photons of momenta k 1 and k 2:

U‘- ' = { e: - e = ^ + - e, ^ 4 - (35- '9)


The probability of emission of two photons of momenta k 1 and k 2
lying within the ranges dkx and dk2 is related to by the expression
dkx dk2
dw kl, k, — 2n\ Ut_yf\2 b{Ei—

On eliminating the ^-function by integration over co2, we obtain


1 (Qe2)h {Qel)Sj
dwk i, k z
(2ny E - E s- o h
2
(Qei)fs(Qe^si (35.20)
col col d(0i d°\ do2.
Et- E s- ( o 2
where Q is the atomic dipole moment.
On integrating (35.20) over the angles, and on summing over the photon
polarizations, we obtain the probability for the emission of two photons
of energies co1 and co2 = Ei—Ef —col

dwW(W \2\ (o^o^)3 dcox,


2 E - E s- o j 2 j j

(35.21)
where Q[]^ are the components of the dipole moment defined in § 27.
The probability for the emission of two photons of frequencies
cox and co2 is usually very small compared to the probability for the emis­
sion of a single photon of total frequency co = co1-j-co2. An exception
exists in the case when the emission of a single photon is not permitted
by the selection rules. This occurs, for example, if the angular momentum
of the emitting system (an atom or a nucleus) in the initial and in the
final states is equal to zero. In this case the probability of emission for a
single photon is exactly equal to zero, since the photon has no states of
angular momentum zero.
We note that states of angular momentum zero are possible in atomic
systems only in the case when they consist of an even number of electrons,
IN T E R A C T IO N O F E L E C T R O N S W IT H P H O T O N S 497

since the electron spin is equal to one-half. This conclusion also applies
to nuclei, whose angular momentum can be equal to zero only for an
even number of nucleons.
For a single electron transitions between states of orbital angular
momentum equal to zero are also strongly forbidden.
A typical example in which the emission of two photons is consider­
ably more probable than the emission of a single photon is the
2s1/2 excited state of hydrogen (n = 2 , I — 0). A transition from this
state to the 1j1/2 ground state (n = 1, / — 0) is possible with the emis­
sion of one magnetic-dipole photon (L — 1, A = 1). However, in the
nonrelativistic approximation the probability of such a transition given
by formula (27.22) vanishes in virtue of the orthogonality of the radial
wave functions with different principal quantum numbers. A non­
vanishing expression for the probability of this transition is obtained
only when exact relativistic wave functions are used, and turns out
to be of the order of 10~7 sec-1.
The transition 2sljz ->• H1/2 with the emission of two electric dipole
photons is considerably more probable. The probability of such a trans­
ition may be obtained with the aid of formula (35.21). In this case
we should interpret s both as the principal quantum number n, and
as the magnetic quantum number m.
The matrix elements of the dipole moment differ from zero if the
difference in the orbital angular momenta of the states, between which
it is evaluated, is equal to unity. Since the initial and the final states
are both j-states, it follows that the “intermediate” states must be
p-states with three values of the magnetic quantum number m = — 1,0,1,
while the principal quantum number in the intermediate state may
take on any arbitrary value starting with 2, i.e., n — 2 ,3 ....
The wave functions for the initial, final, and intermediate states
are or the form

y\(r) = —;= ^
v An
1
Vf {r) •^2o0") J
/An

%(r) = K i ( r) Yim> m = —1, 0, 1, n = 2, 3, ...,


498 Q U A N T U M E L E C T R O D Y N A M IC S

where Rn!(r) is the radial and Ylm is the angular part of the electron
wave function. Therefore, the matrix elements of the electric dipole
moment

can be written in the form

(2 im )/s = /— ■
=■'' Q l n ^ Mm and ((?&)„ = -4==r Q
ni bMm
\/\2 71 \ /\ 2 n

where

Q
u=fR
2o
(r)
Rnl{
r)rz
dr,
0

CO

Q
m=J R w ( r ) R ni ( r ) r z dr.
0

On substituting these quantities into (35.21) we obtain

J 8a2
— J! ( f o 1c o 2) 3 i/ a ) ! .
~ 274 71^= 02n0nl | E, —E„ — co. ■4" 7 . —£ n—(0i
2

(35.22)
The integrals Q2n and Qnl can be evaluated analytically, while sum­
mation over n can be carried out only numerically. The total probability
of the transition 2s1/2—l.y1/2 turns out to be approximately given by (35)
,nl = Er -Ef

vv = I dw,, „ % 7 sec-1
wl= 0
(more accurately it lies between 6.5 and 8.7 sec-1).
We note that the long lifetime of the hydrogen atom in the meta­
stable 2slj2 state was utilized for the experimental observation of the
radiation shift of the hydrogen levels (111).
CHAPTER VI

Retarded Interaction between Two Charges

§ 36. Electron-Electron and Positron-Electron Scattering

36.1. Electron-Electron Scattering


In this chapter we shall investigate processes of interaction between
two charges which do not lead to emission or absorption of photons.
We restrict ourselves here to the investigation of such processes
in the first nonvanishing approximation. Since the first order scattering
matrix S(1) contains only elements corresponding to emission or absorp­
tion of a photon, we must investigate the second order matrix S(2).

P'l P\ Pi P\
F ig . 47.

The simplest process of this type is a collision of two electrons.


It is described by the two diagrams shown in Fig. 47. In accordance with
the general rules formulated in § 25 we obtain the following expression
for the matrix element:
s}» = i e H i j i y M d i p . + p i - p z - p 'z ) ,
where
_ (u2>'/jwi) {ui7,tu\) _ iu2 . y Ov/,;Mi) (76 1)
~ ' (p2- p ly iP z-P vf '
Here p Y and p[ are the four-momenta of the initial states of the two
electrons, p 2 and p2 are the momenta of their final states; u1,u [,u 2
[499]
500 Q U A N T U M E L E C T R O D Y N A M IC S

and u2 are the normalized bispinor amplitudes corresponding to these


states. The normalization volume is taken equal to unity.
In accordance with the rules obtained in § 26 the expression for the
differential cross section has the form

da = ( 2^ ) 2 \ M \* d ( P i+ p 'i— P 2— />2 )<H£i+ % —^


e 2 — e'), (36.2)

where J is the particle flux density. In the center of mass system of


the two electrons (pi+p[ = Pz+p'i = 0),
J — 2v
where v is the velocity of the electrons.
On eliminating the (5-functions from (36.2) we obtain
e4 p2
da | M \ 2 do. (36.3)

In the case of unpolarized particles we must average expression


(36.3) over the polarizations of the initial state and sum it over the final
polarizations, i.e., we must replace \ M \2 by M \ 2, where £ denotes
summation over all the polarization states. This summation can be
carried out with the aid of the method described in § 10 which is not
significantly complicated by the fact that the matrix element (36.1)
contains four, rather than two, amplitudes.
In analogy with formula (10.40) we obtain
1
: ~ j 4s p ( m - 'P i ) y M - i h )
1 £2 ^2

O 2- P 1 )4 SP Yfl ^ ~ * ^ Vv (m ~ * ^ SP Vfl ^ ~ 7v ^ ~ *****

— Spy i‘(m ~ y" “ ' P 3 y n (m ~ y* ~ ^2)J•


In this expression we can replace everywhere yll = yi yflYn by y
Indeed, we are summing over fx (jx = 1,2, 3, 4), i.e., we are dealing
with terms of the type
3

^(■•■)y#l= H y/1(---)y/1+ }74O -O iv


/<= !
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 501

But 7^ = - Y ^ C“ = 1 ,2 ,3 ); y4 = y4 and, consequently,


y„i--)y„ = y „ ( - ) v
Moreover in order to simplify this expression we can utilize relations
(10.43), as a result of which we obtain a sum of terms each of which
contains traces of products of not more than four matrices y . Their
values were given in (10.42).
On carrying out the summation we obtain

da = — d o r Pl p + ^Pl P*y+2m2 (Pi Pi) ~ l np (Pi


2e2 \ (P2-P iY
( P i P i’ ) 2+ ( P i P2)2 + 2 m 2( P i p Q - 2 m 2 ( p 1p 2) ( P iP'i)2+ 2 m 2( p 1p'1)
{Pi — PiY (P 2 -P i )2(P2~P i )2
or
[2(e2/m2) —l]2 f 4
da = r2
4a4(e/m)6 bsin4# sin2#
([e2/m2]—l)2
■rfo, (36.4)
(2[e2/m2] - l ) 2 , ^ sin2# i
where w, e and # are the velocity, the energy and the scattering angle
of the electron in the center of mass system, rQ= ajm is the classical
electron radius.
In the nonrelativistic limit v <s^ 1, e/mt^ 1 and, as can be easily
seen, (36.4) goes over into the Rutherford formula with exchange taken
into account (139)
1 1 1
da = T -°— do. (36.5)
I6v4 Lsin4(#/2) ^ cos4(#/2) sin2(#/2) cos2(#/2)
We express v and e in (36.4) in terms of the velocity vx and the
energy £j of the incident electron in the coordinate system in which
one of the electrons is at rest before the collision (v[ = 0). From the
invariance of the square of the total momentum p2 of the system it
follows directly that
e
+1
m V m
Thus we have (137)

— + 1 3
m 4 U _ _ j - ( i + _ ± - ) do, (36.6)
da = 2r 4(el/m)2 \ sin2# /
° m)2 sin4# sin2#
502 Q U A N T U M E L E C T R O D Y N A M IC S

where # is the scattering angle given, as before, in the center of mass


system. It is related to the scattering angle d in the laboratory system
of coordinates by

cos # 2 — + 3 1sin2#
m / 2 + f e - ' N

In the nonrelativistic limit # = 20 and


1 1 1
da = 4 cos 0 do . (36.7)
( i) ' bsin4# cos4# cos2# sin2# ]
The energy lost by the electron as a result of its collision with
an electron at rest is uniquely determined by the scattering angle in
the center of mass system. We denote by A the ratio of the energy
transferred from one electron to the other one, which was initially at
rest, to the kinetic energy of the first electron: A = (e1—e2)/(e1—m)
= (e'2—m)/(e1—m). From the laws of conservation of energy and
momentum it follows thatZl = |(1 — cos#). On substituting this expres­
sion into (36.6), we obtain
27irl dA
da = A (I-A)
^ (x^-T) A 2(l — A j2

' ( Y A- 1) A2(l ~ A ? } ’ <3 6 -8 )

where x = ejm.
Formula (36.8) gives the energy distribution of the secondary elec­
trons arising as a result of the passage of a fast electron through matter.
For small values of A
1 dA
da -- 47ir20 (36.9)
vf~A*'

36.2. Positron-Electron Scattering


Formula (36.1) for the matrix element can also be applied to posi­
tron-electron collisions. In this case we must make the following sub­
stitution:
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 503

where p[p) and p (2p) are the initial and the final four-momenta of the
positron and v 1 and v 2 are the amplitudes of the corresponding wave
functions with negative frequencies. Thus we have

- m =
(.P2 - P 1)2 (Mp)+ A )2
The diagrams which describe the scattering of a positron by an
electron are shown in Fig. 48. The diagram of Fig. 486 describes the
virtual annihilation of an electron-positron pair with momenta p 1 and
p[p). This pair is converted into a virtual photon of momentum
k — p 1Jrp[p) which again creates a pair of momenta p2 and p[p).

F ig . 48.

The evaluation of ^ | M | 2 can be carried out in the same way as in


the case of two electrons. We give here the result of this calculation
(24):

da =
16(e/m )2
[(e/m )2

0
+ 2 1+ COS4
V id H i-
n r
( l + c o s 2$) — 1 (1+COS0)
] 0 M r
e2(t:2—m2) sin2

+ (—5 —1 (1+COS0 )2 do, (36.10)


m“
504 Q U A N T U M E L E C T R O D Y N A M IC S

where e and # are the energy and the scattering angle in the center of
mass system.
In the nonrelativistic limit (36.10) reduces to the first term of formula
(36.5), i.e., to the Rutherford formula without exchange. The exchange
(i.e., the annihilation) diagram (Fig. 48b) makes a small contribution
in the nonrelativistic limit.
For the energy distribution of the secondary electrons we obtain
do x
F(x; A), (36.11)
dA r°( x— I)2A 2
F(x;A)
A'
l+2(xr—1)(1—Zl)+(x—l)2 I —A
* (*+ 1) {[ ■)1
\-A + A > )\

h ^ - 3 h ! - [ 3 + 4 ( x - l ) ( l - / l ) H - ( x - l ) 2( l - / l ) 2)}'

where
r(p)
x: = and A = £2~ m
m e,(p) —m

(the electron is at rest before the collision). For A << 1 we obtain


formula (36.9).

36.3. Scattering o f Polarized Electrons and Positrons


We now investigate the scattering of polarized particles. The dif­
ference between these scattering cross sections and the formulas given
earlier for the case of unpolarized beams and unpolarized targets can
be utilized for measuring the degree of polarization of the particles.
We have seen (cf. §14) that in the case of scattering of electrons
by an external field there is no polarization effect in the Born approxi­
mation. Similarly, in the cases of electron-electron or positron-
electron scattering the unpolarized beam can become polarized only
in the case when the target is polarized, while the cross section will
differ from the formulas obtained previously only when both the beam
and the target are polarized.
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 505

We assume that the colliding electrons are characterized by the


polarizations £(1) and £(2). By employing the general methods for the
evaluation of cross sections in the case of polarized particles described
in § 26 and already utilized in Chapter V, we can obtain the expression
for the differential cross section

da = ^o-0| l + — (36. 12)

where da0 is the scattering cross section for unpolarized electrons deter­
mined by formula (36.4), while Tik is the tensor whose components
in the coordinate system with axes parallel to the vectors p lt [pip'^\
[/h[Pi/,i]] > have the values (108)
Txl = 4x(2;c—1)—(x—l)(;c+3)sin2??,
T22 = 4x-~ (x— l)(x + 3 )sin 2??,
T33 = 4(2x—1)—(x—l) 2 sin2??,
Ti2 = T2l = (x—l) ] /2{x-\-Y) sin 2??,
^31 = -^13 = T22= T32 = 0,
4x2(l + 3 cos2 i?)+(x—l) 2 (4+sin2??) sin2??
T sin2??
where, as in the preceding cases, & is the scattering angle in the center
of mass system, x = e1/m, ex is the electron energy in the system in
which the target is at rest.
In the nonrelativistic limit

^ = ( 3 6 ' 1 3 )

For ?? = ji/2 the cross section vanishes in the case of complete polar­
ization and when the two electrons have the same spin orientation.
In the extieme relativistic case

-2- [(8 - s i n 2 ??)Ci1 )Ci2)


0 I (3 + cos2 &y

+ sin2 ??(C<1 ) a 2 ) - a i)a 2))]}. (36-14)

We see from the last formula that if, for example, the colliding
electrons are completely polarized either parallel or antiparallel to the
506 QUANTUM ELECTRODYNAMICS

direction of the original beam, the cross section depends very strongly
on the relative orientation of their spins. For example, for ?? — jc/2
do w = i, q 2) = - l )
d a ( ^ = 1, Q2)= 1) “ '
Figure 49 shows the dependence of the coefficients Tik/r on x for
ft = 7i/2 in the case of electron-electron scattering.

V *

For the case of positron-electron scattering we write the cross section


also in the form (36.12), where £ (1) and 1^(2) are respectively the polar­
ization vectors for the positron and the electron, while da0 is the cross
section for unpolarized particles determined by formula (36.10). The
components of Ttk in the same coordinate system have the following
values:
7+ = ( x + l) 2(7 x + l) + ( x + l) (7x2—4x+5) cos ft
+ ( x —1) (x+ 3)2 cos2??+(x—l)2(x+ 3) cos3ft,
T22 = ( x + l) 2( x + 7 ) + (x + l) (x2+ 4 x —13) cos ft
—(x—1) (x + 3 )2 cos2??—(x—l)2(x+ 3) cos3ft,
T33 = ( * + l) (17+ 8x—x2)—(x2—1) (x—11) cosft
+ ( x —l)2 (x+3) cos2??+(x—l)3 cos3??,
T12 = T21 = 2 ]/2 (x + l) sin??[2(x+l)
+ ( x —1) (x+ 3) cos??+(x—l)2 cos2??],
^3i ^1 3 = T22 - r 32 = 0,

r = ( s - i H l - c o s f l ) 1(* + 1) (9x3
+ 4 (x 2—1) (3x—11)cos??+2(x—l)2 (3x2+ 1 2 x + 11) cos2ft
+ 4 ( x —l)3 cos3??+(x—l)4 cos4??].
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 507

In the nonrelativistic limit da coincides with da0, i.e., the cross


section does not depend on the polarization of the particles (this depend­
ence in the case of electron-electron scattering is associated only with
the exchange effect). In the extreme relativistic limit the dependence
of the cross section on the polarization differs only by its sign from the
corresponding dependence in the case of electron-electron scattering
(36.14)

+ sin^(C<1>a2)- a i)a 2))]J- (36.14')

Figure 50 shows the dependence of the coefficients Tih/r on x for


■&= n j l in the case of positron-electron scattering.

V'T'

If the incident electron beam is unpolarized (£(1) = 0) then in the


case of scattering by polarized electrons the incident electrons acquire
a polarization £ (1)' related linearly to the polarization of the target
electrons C(2)
?!1,'= (36.15)
We confine ourselves to stating the values of M ik in limiting cases.
In the nonrelativistic case we have
(1 —cos ■&)cos 0
Mik = _ 2(1 [-3 cos2) ? ) 1* '
In the case of positron-electron scattering in the nonrelativistic limit
M ik = 0.
508 Q U A N T U M E L E C T R O D Y N A M IC S

In the extreme relativistic case for electron-electron scattering the


coefficients Mik are equal to
(1 —cos$)
Mu = (2 + cos # + cos2'&),
(3+cos2$)2
(1—cos$)2
M22, — A/3 3 —
(3 + cos2 ■&)2’
M ik = 0

In the case of positrons the coefficients M ik are of the opposite sign.


Figure 51 shows the dependence of Mik on x for ■&= n j l in the case
of electron-electron scattering.

36.4. Annihilation o f an Electron-Positron Pair into a /.i-Meson Pair


As is well known, the ^-meson which has an electronic charge and
spin I does not participate in interactions stronger than the electro­
magnetic interactions, i.e., in nuclear interactions. Therefore, all the
electrodynamic formulas obtained for the case of electrons are also valid
for /^-mesons.
We. can also investigate the interaction between electrons and ju-
mesons. It will differ from the interaction between electrons by the absence
of the exchange terms, i.e., of the second diagrams in Figs. 47 or 48.
The diagram of Fig. 486 can be used to describe the peculiar process
of the annihilation of an electron-positron pair resulting in the creation
of a ^-meson pair.
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 509

The minimum energy which the positron must have in order to


create such a pair in a collision with an electron at rest is equal to
2/u
// = 4 - 1010 eV,
m
where // is the //-meson mass.
We give the expression for the total effective cross section for this
process (14)
£ V 12
a= 1 --) (36-16)

The maximum cross section (at e m 1.7 e0) is less by a factor of approx­
imately 20 than the cross section for two-photon annihilation at the
same energy.

§ 37. Retarded Potentials

37.1. Interaction Function for Two Charges


In the preceding section we have utilized the expression for the
matrix element in the momentum representation for the scattering
of one charge by another. We shall now investigate this matrix element
in the coordinate representation. This will enable us to investigate the
interaction between two charges in the presence of external fields, and
also to investigate the bound states of the electron-positron system
(positronium).
In accordance with the results of § 23 the general expression for the
second order scattering matrix S(2) has the form:

s<2)= (37.1)

where T is the chronological operator, j is the current operator,


j = ieNiyjy^y) (N is the symbol for the normal product), A is the
operator for the potential.
We denote the set of elements of this matrix between states contain­
ing no photons by eS(2). We shall obtain the expression for eS(2) from
(37.1) if we replace in (37.1) the operator T [A^ (x) A„ (_y)] by its expec­
tation value for the vacuum state of the electromagnetic field:
510 Q U A N T U M E L E C T R O D Y N A M IC S

< r[A (1W A„(j')])0. In accordance with (17.10) this expectation value
is equal to
<?’[A(1W A ,(j')]>0 = b ^ r r i x —y),

where the function Dc is defined by formulas (17.16), (17.18).


The operator consists of products of four electron (or
positron) creation or annihilation operators. In accordance with the
foregoing we shall be interested in those matrix elements of S(2) which
correspond to the presence of two particles both in the initial and in
the final states. In this case results different from zero can come only
from those terms of the operator j M(x)jM(_y) which contain two creation
operators and two annihilation operators, with all the operators refer­
ring to different individual electron states. Therefore, all these four
factors anticommute with each other, while their products in pairs
commute, i.e., j /J(x)jvOO = and the operator T may be omit­
ted from (37.1).
Thus, the scattering matrix for two charges is of the form

eS<2) = —y JJ *)j„(*) d*yd*x. (37.2)

We note the following fundamental fact in connection with the


structure of expression (37.2). In nonrelativistic quantum mechanics
interaction between particles is described by the potential energy of
interaction which is a function of the distance between the particles.
The operator for the potential energy U (in the occupation number
representation) has the form

U = —J Q(rl)U(rl , r 2)o(r2)dr1 dr.,,

where q = y)*ip is the operator for the particle density.


In electrodynamics there exists no potential energy of interaction
which depends on the coordinates of the interacting particles taken
at same instant of time, since the electrons interact not directly, but
via the electromagnetic field. The operator for the interaction energy
is replaced in the first approximation by the scattering matrix eS(2)
defined by formula (37.2) which plays an analogous role. The integrand
in (37.2) contains the coordinates (however, including also the time
coordinate) of only the electrons, while the electromagnetic field has
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 511

been eliminated from this formula. Therefore, the function e2Dc(y—x)


may be referred to as the interaction function for two charges.
We shall see later (cf. § 38) that in the nonrelativistic approximation
it does, in fact, lead to the expression for the energy of interaction
between two charges, i.e., to Coulomb’s law and to corrections to
this law.

37.2. General Form o f the Matrix Element


We construct the general expression for the matrix element of the
operator eS<2). We first consider the case when both particles are elec­
trons.
Let A and B denote the sets of quantum numbers of the initial
electron states, and C and D the sets of quantum numbers of the final
states. Then in the operator
j„ ( * ) O) = - e 2N[xp (x) ip(*)]N [ip(y) yMip(>>)]
nonzero values will come only from

— -^r }„(*)},&) -*■ac aAa t aB(vc O) wAW) (vd GOy j p BGO)


+ at aBa+aA(ipc (y) yMipA{y)) (ipD(x) y^ipB(x))
+a+ afl a+ a„ (vc (x) y ^ s (x)) [ipD(y) y^ipA(y))
+ ai a a a cf aB{ipc ( y) y/tipB(y)) (VpD(x) yMipA(x)), (37.3)
where a+ and a are the creation and annihilation operators for electrons
in corresponding states, while ipA are the corresponding solutions of the
Dirac equation.
By utilizing the fact that all the operators appearing in this expres­
sion anticommute we can arrange them in the form of the normal
product a^aBaA, whose matrix element is equal to unity.
We can then pick out the operator product in (37.3) as a common
factor, with the first two terms appearing with a plus sign, while the
second two appear with a minus sign. We note further that the first
term in (37.3) differs from the second only by the interchange of the
arguments (x) and 0>). On substitution into the integral (37.2) they
will yield identical results since the function Dc(x—y) does not change
when the sign of its argument is reversed. The third and the fourth
terms in (37.3) also differ from each other only by an interchange of
arguments.
Q U A N T U M E L E C T R O D Y N A M IC S
512

Thus, we obtain the following expression for the matrix element


eS,<2)/ :
eS £ \ = e2 j {^D{ y ) r ^ B( y ) D c{ y ~ x ) l p c { x ) y ^ A{x)

- y c { y ) y M y } DC{y - x ) VD{x)y^ A{x)} diy d*x - (37‘4)


In accordance with the general rules for graphical representation
of matrix elements the two terms of (37.4) correspond to the two dia­
grams shown in Fig. 52. They differ only by the interchange of the
indices of the final states. Therefore, the second term in (37.4) is said

F ig . 52.

to be the exchange term with respect to the first one, and the second
diagram of Fig. 52 is said to be the exchange diagram with respect
to the first one. The minus sign of the exchange term in (37.4) cor­
responds to the antisymmetry of the two-electron wave function in
configuration space.
We write the matrix element (37.4) in the form
eC<2) — C(2) _ C(2) H 7 51
------- 0 AB\CD J AB\DC>

where
cd = e
2fy
D(y
) (37.6)
while S {A% DC differs from S {A£. CD by the interchange of the indices
C and D.
We now proceed to investigate the interaction between two posi­
trons. In this case ip contains creation operators, while xp contains an­
nihilation operators. Therefore, if we denote (in contrast to the electron
case) the initial states by C and D, and the final states A and B, we
obtain a formula which differs from (37.3) by the replacement of the
electron operators aA, aB, a^ , a^ by the corresponding positron opera­
tors b^, bg , bc , bj,, and also by the fact that the wave functions xp now
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 513

denote the corresponding solutions of the Dirac equation with negative


frequencies. On going over from wave functions with negative frequen­
cies to positron wave functions, and on taking into account that
(cf. (10.8)) = v4p) we obtain the formula for the matrix
element which defines the interaction between two positrons

es \ l l f = e2f {V a 1(>’) y fl v,n ) (y) Dc( y —x) Wa ’ (*) y ^ c K x )

- y > ^ ){ y ) y ^ ){y)Dc{ y - x ) ^ ){ x ) y ^ % )(x) } d*y d*x.

Naturally, this formula differs from formulas (37.4) only by the relabel­
ling of the wave functions of the initial and the final states of the par­
ticles.
In the case when one of the particles is an electron, and the other
one is a positron, we shall use the following notation: A is the initial
electron state, C is the final electron state, B is the final positron state,
D is the initial positron state. In expression (37.3) we must at the same
time replace a£ by b^ and aB by b^.
On bringing once again the product of the operators a and b to
the form of the normal product a j a ^ b ^ b j = aJb^a^bD, we obtain

•S£>, = ~ (37.7)
The matrix element (37.7) differs from (37.5) by its sign, and this
corresponds to the electron and the positron having charges of opposite
sign. The exchange term in (37.7) (the second diagram of Fig. 52) now
corresponds to virtual annihilation and subsequent creation of an elec­
tron-positron pair.
We shall also apply the fundamental formula (37.2) to the interaction
between an electron and a proton. In this case j(x) can denote the
electron current, while j(y) denotes the proton current, or vice versa.
Let the states A and C refer to the electron, and B and D refer to the
proton. In the matrix elements of the exchange terms will
be absent, i.e., in place of (37.3) we shall obtain

jM(jtr) j #«(^> = 2e2a+ a i a BzA(y>c (x) yfly’A(xj)(y’D( y ) y fly’B(y))

and, consequently,

® = . ‘->
C(2)
AB: CD
(37.8)
514 Q U A N T U M E L E C T R O D Y N A M IC S

where S $ . CD is determined by formula (37.6), as before. The minus


sign corresponds to the opposite signs of the electron and proton charge.
The matrix element (37.8) is represented by the first diagram of Fig. 52.

37.3. Retarded Potentials and Transition Currents


The matrix element S)2J .CD can be written in a form which is very
convenient for applications, and which permits a simple physical inter­
pretation. In order to do this we shall explicitly write the time dependent
terms in the wave functions appearing in (37.6) tpA(x) = y’A(ri)e~i‘°A‘1
and similarly for tpB, ipc and ipD. Furthermore, we shall substitute
for Dc(y—x) expressions (17.16), (17.18):
g i f c ( r , —rt)-lu)(ti— tt)

jy(y-x) dk d(o.
(In) k 2^ 2
As a result of this we obtain

S A B :C D = - ( 2 ^ r /

X f eik (r,~ri) dk j* j eKw° - mz " m)t>dt2.


(37.9)
We first of all carry out the integration over t2:

= 2 n d ( c o D- o > B + co);

the integration over a> now reduces to the replacement of co by


(db d ^ oiB— o)D. We next carry out the integration over k . Integration
over the angles yields:
OO
e i f c ( r , —r 2 )
4n f s i n x \ r 1— r 2
/ — -dk
k 2—CO ri ~ r2\ •'0 *2---(°2
x dx.

Integration over x must be carried out taking into account the rule
for going around the poles of the integrand formulated in § 17, i.e.,
replacing k 2 by k z— i0 ; this yields

sinx | rx—r2 n g i|co |


/ x2—CO2
- x dx =
2
| r , —r , | _
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 515

Thus,
no

Dc( y - . x ) J' (>—icu(t\—U)+l\uj\\rl-r3\ df-Q' (37.10)


&7Z2
*2 —oo

Substituting this expression into (37.9), and introducing for the


sake of brevity the notation for the “transition currents” (here and
subsequently the vector indices on y, j , and y have been omitted,)

7 cm O) = ieWc M m W = 7 c a ( #,i ) c '" Ci‘'»


, x . (37.11)
j DB(y) = ie y jD ( y ) y y j B ( y ) = j D B ( r 2) e ,<ao B > ,
where

J ca^ i ) = ieyic(ri)yv>A(r i)»


Job (ft) = ! ^ D ( r 2) y y J B ( r 2) ,

we obtain the expression for the matrix element Sfy.CD


j r l« -i-rs |

; CD J JcA (**1> O j o B ^ Z i 0 ^"j ^*1 d f 2 dt (37.12)

or, on carrying out the integration over t,


j r !mBDI lri- r2l
■S’a Ucc = -9- ic x (h )-T -— —i— /DflW ^ 2 , (37.12')
Z, J ;' 1 »2i
with r/jCil = —c»DB, i.e., a>c —<oA = (>)B—(oD, which expresses the law of
conservation of energy. For the sake of definiteness we take coA > o>c ,
then (oB < o)D and o>Ac — —o)BD > 0. On taking this condition into
account we can rewrite (37.12) in the form

C(2)
‘~>A B \ C D ' I dt I j DB(r2, t)ACA(r2, t) dr2 = - e f y>D(y) ACA(y)yjB(y) d4y,
(37.13)
where
-ia>ACt + UoAC \r ! - r 2|

A
c A ( r -‘ > 0 — ^ J 7 cm (r i)
!r, —r0
d/T

or

(37.14)
Cj4V ’ ' J |r i - r 2|
516 Q U A N T U M E L E C T R O D Y N A M IC S

It is remarkable that formula (37.13) coincides with the expression


for the elements of the first order scattering matrix S(1) in the case of
the transition of an electron from state B to state D induced by an
external field whose potential is given by Aca. In these formulas A ca
is determined by the transition current j CA corresponding to the tran­
sition of the particle from state A to state C in exactly the same way
as in classical electrodynamics, i.e., by the formula for the retarded
potentials (37.14). Naturally, the differential equations for the poten­
tials of classical electrodynamics also hold:

S(ACA), _ Q

We note that expression (37.12) for Sjfy.CD is symmetric in the two


currents. In going over from this expression to (37.13) we have treated
the current j DB as “experiencing” the external field A ca , and the current
j CA as giving rise to this field. Here we have taken the current with the
positive frequency to be the one which gives rise to the field (we have
assumed that coAC > 0; the transition takes place from the higher energy
level to the lower one). If we regard the current with negative frequency
as generating the field, i.e., if we write (37.12) in the form

5’A'2>
B:CD — if Jca^dbW^X,
then in place of (37.14) we obtain ADB in the form of “ advanced”
potentials

We now go over from the scattering matrix to the matrix for the
effective perturbation energy for the system of two charges defined
in the following manner:
2ttiUi^f d((oA i coB coc coD)
or
^A
ABB:C
CDD — 2 7 l i U A B . C D d ( ( O A c ~ \ ~ l ,-)B D) •

Then we obtain from (37.12'):


1 r lr]-r2!
| JcA{ri)JDBir2)
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 517

Expression (37.15) can also be written in the form

U ab -.cd = J* J ob (^2) ^ca (^2) ^ * 2 5 (37.16)


where

The potentials A ca satisfy the differential equations of classical electro­


dynamics for monochromatic fields
AACa ~\-u>cAAc a = j CA, div A ca coca(A^)ca — 0. (37.17)
The results of this section provide a rigorous foundation for the
semiclassical method of treating the interaction between electrons
and the electromagnetic field (the so-called “correspondence principle”
(27)). In this method the quantum properties of the electromagnetic
field are not taken into account, and the equations of classical electro­
dynamics are used in which the currents are replaced by the transition
currents. It can also be shown that the energy flux in the classical
electromagnetic field corresponding to the potentials (37.14) is equal
to the probability of emission of a photon per unit time, with the electron
making the transition from state A to state C, multiplied by the photon
energy coAc.

§ 38. Interaction Energy of Two Electrons to Terms of Order v 2/ c 2

38.1. The Breit Formula


In the preceding section we have obtained an expression for the
matrix element for the effective interaction energy of two electrons.
It has the form
u ^ f = UAD:CD- U AB;DC, (38.1)
where A and B denote the initial, and C and D the final states of the
electrons. In accordance wnh (37.15) the matrix element UAB.CD can
be written in the form

U A B . C D = a f ? c * W ^ W T - 7 f el|" JC ' lri ^


J h r2i
(38.2)
where a = e2/47i. Here the operator oq operates on the function y A(ri)’
while the operator a 2 operates on the function yjB ( r 2) . We note the
518 Q U A N T U M E L E C T R O D Y N A M IC S

structure of expression (38.2). It contains an operator which depends


on the coordinates and the spin variables of both particles, (1—a ^ ) /
\fl —r2\, and the “ retardation factor” exp(i\coAC\ \r1—r2\). The presence
of this last factor, which depends explicitly on the initial and the final
energies of the system, does not allow us to introduce in a general
form the concept of the operator U for the interaction between two
electrons whose matrix element would be given by UAB.CD\

UAB-CD= J wi (n) v t 0 2) UyjJrJ rpB(r2) drxdr2. (38.3)


However, for small velocities (vie << 1, c is the velocity of light) such
an operator can be constructed. In order to obtain it we expand the
matrix element in powers of vie up to terms of order (v/c)2. As in
the case of the analogous expansion in § 15 we shall not use in this
seetion the system of units in which c = 1. This is convenient from
the point of view that the transition to the nonrelativistic approxima­
tion can be formally accomplished by expanding in powers of 1/c.
The expansion of the retardation factor up to terms of order 1/c2
inclusive is of the form

— \ mA ~ Mc \ | r i— r 2 i

e 1
---- I. 1(0
■ '•, —(or ((Oa -(Oc)2
(38.4)
Iri~ '2 —**2I C 2c2
In order of magnitude the matrix elements of the operators a are equal
to vjc. Indeed, y)*a.ip= while —’(p/c. Therefore, in terms
containing a ja 2 in (38.2) it is sufficient to retain only the first term
in the expansion (38.4). The remaining terms will have as factors not
1—a xa 2 but simply unity. The second term in (38.4) will vanish after
substitution into the integral in virtue of the orthogonality of the
functions ipA and xpc. The third term can be put into symmetric form
by utilizing the fact that coa —ojc — (od—(ob , and therefore
— (<*>A— " c ) 2k l — r 2 \ = ( f ' h - (Uc) '« h j) k l — 1*2I •
We can then eliminate the frequencies from the expression for
the matrix element by utilizing the Dirac equation H ^ f o ) = o.)Ay A(r1)
and similar relations for the other wave functions. Keeping in mind
that expression (38.4) is multiplied on the right by y)A{rl)ipB(r2), and
on the left by V*c (ri)y*(r2) and integrated over drx and dr2, we can
replace in it coA and coB by the operators Hi and H 2 to the right of
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 519

the factor \rx—r2\, and coc and a)D by the operators H t and H2 to
the left of \r1—r2\, i.e.,

(ft)A—wc)(wB— r2| -► ki~ »'2|H 1H2+ H 1H2| iq—1-2|

|#*i r21H, H2| r1 r21Hx = [Hls [H2, \r1—#*21]].

(The square brackets denote commutators of the corresponding quan­


tities.)
A

Since the Hamiltonian operator has the form H = cap-\-(imc2-\-eA{e),


where A (e) is the external field, it contains only a single term, viz.,
ap which does not commute with \rx—r2\. On evaluating the commu­
tator ^[[oq/q, a 2p2, | —r2|]], we find that in substituting (38.4) into
the integral (38.2) we can use the substitution:

« ia 2—(ttl«)(a 2«) (38.5)


r, —r.
where

n r_i — r 2
~r2

On substituting (38.5) and (38.4) into (38.2) we can show that UAB.CD
does indeed turn out to be the matrix element of the operator (33),
( 112), (20)

q «1q2-h(g1/i) (g2n)
(38.6)
'2 k i —ra|

for which relation (38.3) holds.


The first term in (38.6) is the Coulomb interaction between the
charges. The second term takes into account corrections arising because
of the motion of the charges.
Since this expression holds only up to terms of order (v/c)2, the
[(p\
electron wave functions tp = ^ I should also be taken to terms of
the same order, i.e., we must use expression (15.2)

CTP
X = 2me (p.
520 Q U A N T U M E L E C T R O D Y N A M IC S

On substituting this approximate expression for % and the operator


(38.6) into (38.3) we obtain

UAB,CD = a f <P*cO J <P%O 2) <pA0 i ) 9?b O 2) |


J Ir l r 2 I

+ ~ ^ T J {[aiPi?c(ri)]*[<fiPi<PA(ri)]?Z(r2)<
P£(r 2)

+ <P* O i) <Pa O i ) [® * P 2 <PdO 2)]* [ o 2 P 2 (P b O 2)]} y P ~ ~ T


Ir l '2 I

+ f {< P c O i) CT1(CT1P i) <Pa O i) + [® i P < P c O i) ] * ® i <Pa O i) }

c/f* dy
x O 2) CT2 (° 2 P 2) PPb 0 2 ) + [ ° 2 P 2 ? b ( 1' 2 ) ] * ^ b C ^ ) } , 1J
I*1 r2

---- 8 (CTl") (a i P i ) ^ ( r i ) + lffiP i? ’c (ri)]*(0 i #,) ^ ( ri)}


cfy dy
x { ^ ( r 2) (a2/i) (a2p2) 9?B(r2) + [a2p2<pB0 2)]*(a2/i)<pB(r2) } ~ — - .
I' 1 r2 I
(38.7)
We now introduce the approximate two-component electron wave
function 0 related to <p by expressions (15.7), (15.9):

<P 0. (38.8)
8m2c2
In doing this it is sufficient to replace <p by 0 in all the terms of
expression (38.7) with the exception of the first one, since they already
contain the factor 1/c2.
We now rewrite expression (38.7) in such a way that it assumes
the form (38.3). In other words, we find an operator U (e) which leads to

UABiCB = (38.9)
(As we have done previously, for the sake of brevity we again omit
here the spin indices of the wave functions.)
In order to bring (38.7) into the form (38.9) it is necessary to in­
tegrate by parts. Since the integrand contains high powers of the quantity
\rL—r21-1 we must exclude the origin (the point r1 = r2). The integral
over the surface surrounding the origin yields a finite quantity as
R E T A R D E D IN T E R A C T IO N B E T W E E N TW O C H A R G E S 521

k i-fa ! 0- Since the latter depends only on the values of the integrand
at /q = rz it can also be written in the form of a volume integral of
an expression containing <5(1*!—r2). We finally obtain the following
expression for the operator for the interaction energy of two electrons:

u<e) = 7 “ 7r^ 2 <5W - — -rT O P i K - kp2K + 2 kPi]®2

- 2 [rpa] o x) - ( y Pl p2+ — »■kPi ) p2

a I"ql q 2 _ 3 f o r ) f o r ) _ _8tc_
S(r) (38.10)
4/7 z2 c 2 [_ r 3 r5 3 CTl<J

where r = »q—r2.

The ^-functions contained in this expression have arisen in the manner


just explained. In this expression (since the region in the neighborhood
of r = 0 has already been taken into account) we need not pay any
attention to the seeming divergence of certain integrals arising, for
example, as a result of substituting the last square bracket of (38.10)
into (38.9). In order to avoid divergence we can first carry out the
integration over the angles.
Naturally, the presence of (5-functions in (38.10) does not indicate
a strong interaction. Terms involving ^-functions contain 1/c2 in their
coefficients and, therefore, from the point of view of the foregoing
expansion must be treated as small compared to the first term (the
Coulomb interaction).
We shall return to the interpretation of the individual terms of
the operator (38.10) in the next section when we apply it to a specific
problem. We note here a certain similarity between this operator and
the energy of the electron in an external field up to terms of order
(v/c)z (cf. (15.10), (15.11)). Thus, the second term in (38.10) is similar
to the last term in (15.10) if A 0 appearing therein is the Coulomb
potential A 0 = e/ 4 7 ir. This term is followed in (38.10) by terms of the
same type as the penultimate term in formula (15.10), if we interpret
E appearing therein as the Coulomb field E = er/Anr3. The last terms
in (38.10) correspond to the last term of formula (15.11) in which H
denotes the field of a magnetic dipole of magnetic moment ea/ 2 me.
522 Q U A N T U M E L E C T R O D Y N A M IC S

We note that the operator for the interaction energy of two electrons
(38.10) can be obtained from expression (36.1) for the matrix element
for electron-electron scattering. In the nonrelativistic approximation
(including terms of order (pjm)2) we can write, on the basis of (10.9')
and (10.10), the plane wave amplitude u in the form

u—

On substituting this expression, and also the approximate equation

j _ , J _ _i_ (wO
q2 4m2 m2(^2)2 (
q=p
'i—
Pi=P
z—P
z)
into the first term of (36.1) we obtain
e2
K y ^ ) («^/Mw2) — = w\* h’2* Ue{q)w1 u’2,

UM
(q
)= ^ — “4/-2
-+ 4 ^ -[?/>2]52+2 [w >2-2[

, (qpi) (qpz)__P1 P2 , (gCTi)(gq2)


m 2 (q2) 2 m 2 q2 4m4 4m2 j
This quantity is the momentum representation of the interaction energy
(38.10) Ue(r) = / ei(irUe(q) dq.

38.2. Schrodinger Equation for a Two-Electron System


The exchange matrix element in (38.1) UAB.DC can be brought into
the form (38.9) in exactly the same way as UAB.CD

UAS,DC = f ^O A rfl^M d ^d ri, (38.11)

where U (e) is the same operator (38.10).


If we introduce the antisymmetrized two-electron wave functions

r2i ^ 2)

= -$= [0A(ri,*i)4>B(ri, K ) - 0 b (Ti , (38.12)


R E T A R D E D IN T E R A C T IO N B E T W E E N TW O C H A R G E S 523

where and 22 are the spin variables of the particles, and analogous
functions for other pairs of states, then the matrix element (38.1) can
be written in the form
U ^ r = J 0*CDUe0 AB drx dr2. (38.13)
If we interpred A and B as arbitrary stationary states of the electron
situated in a given external field, then the wave functions 0 AB form
a complete system of antisymmetrized eigenfunctions of the operator
H i+ H 2, where Hx and H2 are, as before, the Hamiltonian operators
for a single electron.
Due to the existence of an interaction between electrons the state
described by the wave function <PAB is not stationary. If at / = - o o
the electrons were in the state 0 AB then at t = + o o they can turn out
to be in some other state 0 CD; the probability of such a transition is,
as we have seen, proportional to

I j \

We can formulate the problem of determining the stationary states of


a two-electron system taking their interaction into account (up to terms
of order v 2 /c2). Such states are superpositions of states (38.12), and
the wave functions 0 n corresponding to them are eigenfunctions of the
operator
H = H 1+ H 2+ U e. (38.14)
Thus we have obtained the Schrodinger equation for two electrons
in configuration space
H 0 n = E n0 n, (38.15)
where En is the energy of a stationary state. In addition to satisfying
equation (38.15) the wave functions 0 n must satisfy the requirement of
antisymmetry.

38.3. Interaction between an Electron and a Positron


We now proceed to the determination of the interaction energy
between an electron and a positron to terms of order v 2 lc2.
In accordance with (37.7) the matrix element is in- this case of
the form £/)_>,= —{UAB.CD— UAB.DC), where A and C, as before, denote
the initial and final states of the electron, while D and B denote the
initial and final states of the positron described by the wave functions
524 Q U A N T U M E L E C T R O D Y N A M IC S

ipB and yjD with negative frequencies. In place of these functions we


introduce the positron wave functions in accordance with subsection
10.3, yj{p) = Cy){~ \ and utilize the fact that V'ir’TVvV’ = Va’7/iWd’•
Then the matrix element UAB.CD assumes the form
'“vie ki—r,|
1—oqo^ VA(t‘i)VD)(l‘2)drl dr2.
U ab,cd = « J y’ * ( r i ) V JBP ) * ( r 2)

(38.16)
In (38.16), as in (38.2), we have arranged on the left the wave func­
tions of the final states of the particles, and on the right the wave func­
tions of the initial states, with all these functions corresponding to pos­
itive frequencies. Therefore, by proceeding in exactly the same way
in which (38.2) was transformed into (38.11), we can obtain from
(38.16) the formula

U Ab : c d = J ^ ( r J ^ , * ( r 2) U (e)^ ( » - i) < P i,p ) ( r 2 ) ^ r i t / r 2, ( 3 8 .1 7 )

where U (e) is the operator (38.10).

38.4. Exchange Interaction between an Electron and a Positron


After expansion in powers of 1/c2 the exchange matrix element
1 la -rj
1-oqcts °D i !r i-

U a B-. DC —
= aJ
'Va (Ti)w )(r2)dr1dr2,
(38.18)
assumes a form quite pifferent from the two-electron case. Since a>D
and coB are negative, the “retardation factor” now contains not the
difference, but the sum of the particle energies toA—<x>D = ea -{-e{J >),
including the rest energy 2me2. Thus, (oja —ojd)/ c is not a small quantity
of order 1/c, but a large one — of order c — and expansion (38.4) ceases
to be applicable.
In this case the transition to the low velocity approximation can be
easily carried out by replacing the energy in the exponential by the
rest energy ^ 2me2. We then obtain

1—«i«2
UAB: DC V>D )v>* (r2)IIn —r. I e,2i,nc|ri~r'1 ('■?.) dri dr2 -
- J
(38.19)
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 525

We note that in (38.19) the initial and the final states of the particles
enter only through the wave functions, i.e., this expression already has
a structure of the form of (38.3).
The retardation factor in (38.19) contains in the exponent the large
coefficient 2me and, therefore, oscillates with a wavelength (l/2mc)
(1/mc is the “Compton wavelength” for the electron). For this inte­
gration the important contributions come from the region 1^—r2\
~ l/m c . Since the electron and the positron wave functions (in accord­
ance with the assumptions that the velocities are small) do not vary
significantly over such a distance, we can segregate in (38.19) the integra­
tion over r = rx—r 2 by setting rx = r2 in the arguments of the wave
functions:

U AB, DC = a J \vD~)* ( r ^ V A(r i) y )*Ori) V B ' )OrO

/ „i2mcT
~ T ~ dr'
The last improper integral can be evaluated in the manner in which
expressions containing oscillating integrands are usually evaluated:

/
pilmCT f Jy 71
------- dr = lim e^rncr-xr — _
(me) 2
r n=0 J r
as a result of which the exchange matrix element assumes the form

UAB, DC= ~ n ~ ^ c2 j ^ y,A^ ) (V>^ ^ >(rJ)


- ( v P * Oh) *V a Oh)) (Wc (»’i)«Vb")Oh)) drx
or

U A B . DC = '/ Z~ m ^ c 2~ j ^ ’O h^cO hX 1 —« i “ a)
X S(r 1 —r 2)i/JA(r])yj(f >(r 2)dr 1 dr2. (38.20)
In (38.20) we express the wave functions with negative frequencies
in terms of the positron wave functions y>(-) = —/3Cy>ip)*;
= —^ (p)C/3; we then obtain

u AB, {(VS, C M W*cP O S T )


- ft f t * dr,. (38.21)
526 Q U A N T U M E L E C T R O D Y N A M IC S

The last expression already contains the coefficient 1/c2. Therefore,


in the wave functions

we can omit and to the desired approximation replace cp by the


two-component wave function 0 . Each four-rowed matrix
J\r2\
r
J \ r J *
where J \ , r 2, . . . are two-rowed matrices, can be replaced here
by jTx. Utilizing the explicit expression for the matrices C — ay (cf. § 10)
and a, we obtain
0,
c Pa„ = ayf i c -> - 1 ,
Cfiax = —afixC -> ioz ,
C(5az = — a*/?C -*• - i a x.
Using these expressions in (38.21), we obtain

UAB,DC — J ^ c ( rl ’ ^ 2) [ 1 + c rl^

+ ° i x a z A ® A ( r i> ^i)^bp)( U ^ 2)} d r l t


where X1 and X2 are spin variables (ax operates on Xx, while a , operates
on X2).
In order to put the exchange matrix elements into a form analogous
to (38.17) we bring the integrand in UAB. Dc into the form
0 * ( A 1 ) 0 ^ * ( X 2) T 0 a (2.1 ) 0 ^ ) ( L 1) .
It can be easily shown that the operator T must have the structure
T = c1 ~\-c2 o l o 2, where c1 and c2 are constants. Indeed, the matrix
element UAB.DC must be a scalar (in three-dimensional space). The
operator T, which is a general bilinear combination of four two-rowed
matrices (1, ax, ay, az), guarantees that this condition is satisfied.
By utilizing directly the explicit form of the matrices cr; we obtain
the following values for the coefficients and c2: c1 — if and c2= i.
Thus we have
U AB, l)c = J 0 2 ( r 1) 0 < j ‘> * ( r 2)U ^ 0 A ( r 1) 0 ^ ( r 2) d r 1 d r 2 ,
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 527

where (152), (12)

U " " h> = ~2 »*)<)(<•.-»•,). <38-22)


The matrix element for the transition Ui_^f can now be written
in the form
U ^ f = f 0*BV (p)0 ADdridr2, (38.23)
where the operator
U (p) = —U (e) + U (exch) (38.24)
is the complete operator for the interaction between an electron and
a positron and
0 A D ( r i > ^ 1 i r 2 > ^ 2) ~ 0 A ( r 1 > ^ l ) 0 {DP) ( r 2 > ^ 2) )

0cs(r 1 . * i ; r a, A2) = 0 c (r1, (r 2, A2).


The functions &AD (A and D are the quantum numbers of arbitrary
states of the electron and the positron in a given external field) form
a complete system of eigenfunctions of the operator H (e) + H (p), where
H (e) is the Hamiltonian for the electron, and H <p) is the Hamiltonian
for the positron (in the presence of an external field they differ by the
sign of their charges).
The wave function ( r ^ ) of the stationary state of the electron-
positron system satisfies the Schrodinger equation
[H"»(r1)+ H "» (r!) + U 'p»]$n = En0 n, (38.25)
and is not restricted by any symmetry requirements. The electron and
the positron appear in equation (38.25) as nonidentical particles.
The symmetry of the system with respect to the reversal of the sign
of the charge of both particles leads, as was shown in subsection 23.3,
to a definite charge parity. In particular, the symmetric states &n(x,y)
= 0 n(y> x) (* and y denote the set of the coordinates and the spin
variable) are charge-odd, while the antisymmetric states 0 n(x,y)
= —0 n( y , x ) are charge-even.

§ 39. Positronium
39.1. Hamiltonian Operator and the Unperturbed Equation
The Schrodinger equation (38.25) can be applied to the problem
of positronium, i.e., to the problem of the bound states of the electron-
positron system.
528 Q U A N T U M E L E C T R O D Y N A M IC S

The operator for the interaction energy between the particles is


determined by formulas (38.24), (38.22), (38.10), while the Hamiltonians
for the electron and the positron are determined by formula (15.10)
with A 0 = 0 (there is no external field).
The Hamiltonian operator in equation (38.25) consists of terms
of different orders of magnitude. We write it in the form H = H (0) + H (1),
where H <0) includes terms which do not contain 1/c2:

(39.1)
H "” = 2 ^ W + P |)

while H (1) is proportional to 1/c2.


The problem of the positronium energy levels can be solved by
perturbation theory. The operator H <0) can be treated as the unper­
turbed Hamiltonian, while H (1) can be regarded as the perturbation.
The unperturbed problem is exceedingly simple; it reduces to the
problem of the hydrogen atom in nonrelativistic quantum mechanics.
Indeed, we use the system of coordinates in which the center of mass
of positronium is at rest. Then we have p x = ~ P i = P, where p is
the operator for the momentum corresponding to the relative radius-
vector r = r^—r^. The unperturbed Schrodinger equation has the form

p2tf>= (39.2)

It coincides with the equation of motion for the electron in the hydrogen
atom, if we replace the electron mass by the reduced mass for two
electrons m/2. Therefore, the values of the positronium energy are
smaller by a factor two in absolute value compared to the correspond­
ing energies of the hydrogen atom, while the orbit radii are twice as
large.
As is well known, the unperturbed levels depend only on the prin­
cipal quantum number n and do not depend on the quantum numbers
j and /, which define the total and the orbital angular momenta.

39.2. Perturbation Operator


We now proceed to investigate the fine structure of the positronium
levels, i.e., their displacement and splitting due to the perturbing energy
H (1) (10).
RETARDED INTERACTION BETWEEN TWO CHARGES 529

We write H (1) in the form of a sum of five terms


H (1) = V1+ V 2+ V 3+ V 4+ V 5, (39.3)
where V1, . . . , V 5 are defined below by formulas (39.4)-(39.8). The
term Y 1 is a correction of order 1/c2 to the kinetic energy of the particles:
1 P4 (39.4)
V: =
8m3c2( P i + P a ) 4m3c2
The remaining terms are contained in the operator for the interaction
energy U (n). In V2we have combined all those terms from the operator
U (n) in (38.24), which contain no spin operators. The operator V2 describes
the interaction associated with the orbital motion of the particles1:

v’ = 4 ^ ) ,5w+ w { 7 p‘p2+^ ,'(rp‘)p}


This operator can also be written in the following form from which
it can be clearly seen that it is selfconjugate:
e2 [1
2—p2 - —4n h 2 (5(r) — — [rp]2j

We write V2 in still another form which is more convenient for further


calculations:
eti A, n 1 1 enx. \\ 2 1
V2 L2, (39.5)
m 2-5
= 71
me m 2 c2 r 3 2 c2 r
2 2 P
2 \ me
where ftL = [rp] is the operator for orbital angular momentum.
In V3 we have collected those terms which contain both the momen­
tum operator and the spin operator:

V 3 = 1^“ , ) -^r{[»'Pl]CTl - [ 1'P2]®2 + 2[''Pl]CT2 - 2 [»'P2]®l}-

This operator describes the spin-orbit interaction. It can be written


in the form
V3 = - ( — - ) - i- L s , where s = - - ( o 1+ o 2) (39.6)
3 2 \ mcf r 3 2
is the operator for the spin of the system.
1 In th e fo llo w in g d is c u s s io n w e u se th e o r d in a r y s y s te m o f u n its a n d in tr o d u c e

th e c o n sta n t h in to th e fo r m u la s ; e d en o tes th e c h a r g e in o r d in a r y G a u s s ia n u n its .


530 QUANTUM ELECTRODYNAMICS

The operator V4 describes the interaction between the spin magnetic


moments of electron and the positron:
12
v.= £
If we introduce the spin operator s , then V4 can be written in the form

eh x,xk
is-V*! 3) <5(r). (39.7)
V- = < W
Finally, we denote by V5 the exchange interaction operator (38.22):
Vr = U (exch). can ajso written as

s 2 d(r). (39.8)
= * (\ —
m c I) '
It is very essential that, as can be seen from (39.3)-(39.8), the pertur­
bation energy H (1) contains operators acting on the spin variables of
the particles only in the form of the total spin of the system s. This
means that the matrix of the perturbing energy is diagonal with respect
to the eigenvalues of the operator s2, which are given by s2 = ^(^+1).
From this is follows that positronium levels can be divided into singlet
or parapositronium levels ( 5 = 0), and triplet or orthopositronium
levels (s = 1).

39.3. Fine Structure


The operators V 1 and V2 are diagonal with respect to the orbital
quantum numbers I and do not depend on the spin variables. The
interaction described by these operators removes the degeneracy with
respect to /. In order to determine the correction to the energy level
we must evaluate the expectation values of these operators (V4 and V2)
in the unperturbed states. By utilizing the unperturbed wave equation
(39.2) we can easily obtain Vx:
e2
V ,= (E2jr 2Ee2 r~l + e4r~2) (39.9)
4m2 c2
The last term vanishes. Indeed,
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 531

Furthermore, we have
<5(>) = 0 2(O).
and
i 1 ,
)=
r
= JI 0* -V r 2 dr
dr = J
0
R - * - dr = ~ y * 2(0).

where £ is the radial wave function; since R(0) differs from zero only
when 1 = 0 , then R 2 (0) = 4 tt0 2(O). Consequently,

P = 0.

For the evaluation of the different expectation values we can use


the well-known formulas which hold for the hydrogen atom (taking
the reduced mass into account);

Jo_ 1
£ tf>2(0) = c5;0
4n2’ Rnaln2

(39.10)
2 a0 n2’ r 2 2 a\ ns (2 /+ 1) ’

4fl03«3/(M-l) (2 /+ 1 )’
where e0 = e2 /a0, a0 = h 2 jme 2 (the Bohr radius) and n is the principal
quantum number.
On utilizing these values we obtain

1
(39.11)
“ e° 8«3 \ 2/ • 1 8n 137
Similarly, in accordance with (39.5) and on utilizing (39.2) we obtain

eh f _ ( E ± ( e y r ) ] + 1 el. / ( / + , ) r ^
V = 3jt 02(0)-
me mc- / 2 \ me I-
We note that lr~ 3 must be taken equal to zero for 1 = 0 , in spite
of the fact that r 3 diverges for / = 0 as r -> 0. Indeed, in the derivation
of the expression for U <e) the terms for which the origin played an
important role in the matrix elements were treated separately. In the
532 QUANTUM ELECTRODYNAMICS

product lr~3 the factor / arises as a result of integration over the angles
which, as we have already noted in § 38, can be carried out first in the
case of singular terms.
On utilizing (39.10) we obtain

(39.12)
V2 <5,0 _8n3
L (_ J_
\ 2/+1

The operator V3 is also diagonal with respect to / and s. Its eigen­


values are equal to zero in singlet states, and depend on the value of
the total angular momentum j in triplet states. On utilizing the fact
that 21s = j ( j + l ) —l ( l + l ) —s (s + l), we may directly w rite-

3 2 ;(;+ i)-/(/+ i)-2 for s = 1, 1 7 ^ 0,


V3= 16 a £° «3/ ( /+ l) ( 2 / + l ) (39.13)
0 for s = 0, or / = 0.

In the operator V4 of (39.7) we treat the two terms separately.


The second of these is diagonal with respect to s and differs from zero
only for / = 0. It determines the expectation value of V4 for 1 = 0 :

v , = 4 ^ ( — ) “-L [ i,( i+ l) - 3 ] 4 > * ( 0 ) ,


3 \ me J 4
or
-1 for s=
1 (39.14)
for s=
3

The first term in (39.7) has matrix elements different from zero only
for 1 7 ^ 0, s = 1, both in the case of terms diagonal and nondiagonal
with respect to /. From the laws of conservation of total angular mo­
mentum, spin angular momentum and parity it can be easily seen that
the only nonvanishing nondiagonal elements of V4 are those which
connect states with the same values of j and M for s = 1 and / = j ± 1.
For the evaluation of the matrix elements we must write the posi-
tronium wave functions 0 )isM in the form
RETARDED INTERACTION BETWEEN TWO CHARGES 533

where cplm and xsu are the orbital and the spin functions, as a result
of which we obtain

( n v .io H c ; ^ c z v V '\r- ’ \o

/ „ , Xj xk
X I m — K^'ISjSfcl/*).
T**
The last two factors in this expression can be easily evaluated, since
the angular and the spin functions are of a simple form. After this we
can easily carry out the summation over m, m and /x and express
(/' |V4|/) in terms of only the radial matrix elements. We shall state
the result.
After the application of (39.10) the diagonal elements assume the
form
1 for j = /,
I
for j = l + 1,
(39.15)
( / >v - ^ = W + W / + i ) ^ j 2 /+ 3
/+ 1
for j = I—I
21—\
( l ^ = 0 , s = 1),
while the nondiagonal elements become

U + 11V41j - 1) = a% 0 + 1 | r 31j - 1). (39.16)

We shall not carry out the evaluation of (j~\-l|/'“3jy —1) here since
in the general form this does not lead to simple expressions, while in
all special cases referring to the discrete spectrum the calculations are
elementary. In particular, for j = 1 and n = 3 (39.16) vanishes, so that
the orbital quantum number remains a good quantum number in the
case of the first three principal quantum numbers.
Finally, the expectation values of V5 differ from zero only for I
= 0, s = 1:
0 for s = 0,
V5 a2 (39.17)
e cA o for s = 1.
4n*
534 QUANTUM ELECTRODYNAMICS

Thus, of all the terms making up the perturbation energy H (1)


only V4 contains nondiagonal elements belonging to the case / = y+1,
V = j — 1 and 5 = 1 . Therefore both singlet and triplet states for which
/ = j (parity (—1)0 can be classified by their orbital angular momentum.
We denote the shift of each such level with respect to its unperturbed
value by w(Zs+1/^) (the spectroscopic symbol of the term is indicated
in brackets). Then we have

w(00 — (39.18)
a £ 2n3 \ 2 / + 1 32n / ’
a2e
w(3j}) = w (0';)- Sn3j ( j + \ ) (2y+l) (39.19)

Terms of parity (—l)m are superpositions of unperturbed terms for


which / = y i 1. For these terms we have

w = -V± ,- —
~ V- - L i // l/yv + +
_l
vV.
V ( + 0 + l ,V 1! ; - l ) 2 ! (39.20)

where
v _ s V (> 0 -1 ),)

= w (10 '- i ) j - i )+ [-3 »+ 4 n , j £ j _ ,) ( \ ~ 2 jq rr)] •

a2£0
v + = v ( » o + i ) , ) = H . ( ‘ o -+ i)m )-
4 n ' ( , + l] \ 2 ! 2 j+ I

and (y’+ l IV4jy —1) is defined by formula (39.16).


These formulas enable us, in particular, to determine the energy
difference between the ground states of ortho- and parapositronium.
The dependence of the energy on s for / = 0 is contained only in formulas
(39.14) and (39.17) which define V4 and V5. From these formulas it
follows that the ground level of orthopositronium lies above the ground
level of parapositronium, with the energy difference equal to

A = *-e0a2| y + l j = 8.2-10-4eV. (39.21)

The first term in brackets corresponds here to the magnetic interaction


V4, and the second term to the exchange interaction V5.
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 535

39.4. Zeeman Effect


We investigate the Zeeman effect in positronium (10). The special
features of this effect are associated with the fact that the orbital magnetic
moment of positronium is equal to zero, while its spin magnetic moment
is not proportional to the spin angular momentum.
Indeed, since in positronium t a p j = [r2p2], the orbital magnetic
moment operator is given by
eh
Pl = 2 ^ ( [ r i P j + [ r 2 PJ ) = 0 .
The spin magnetic moment operator
eh
1* = 2mc (cti- o2)

does not commute with the total spin operator s = if o + a a ) . Therefore,


the eigenstates of the operators p. and s do not coincide in the general
case.
We consider the four eigenfunctions of the operators s2 and sz,
corresponding to definite values of the spin s and of its component s2:
Xu = a(l)a(2),

Zlo = — [a(l)0 (2 )+ 0 (l)a(2 )],


172
*oo=
^ [a (l)/? (2 )-/? (l)a (2 )],
172
where a and /? are the two spin functions for one particle corresponding
to the values o - = ± 1 , while the numbers in brackets denote: (1) an
electron, (2) a positron. With the aid of these functions we evaluate
the matrix elements for the operator of the perturbing energy for
a positron in an external magnetic field H ; V(H) = We can
easily show that all the matrix elements (ss2 1V(H) | s's'J vanish with the
exception of
(00 1^,110)= (10 W 00) = ~ H .

Thus, the matrix for the energy of interaction with the magnetic
field contains only elements which connect singlet states with triplet
states. But, as we have seen, the perturbing energy V which determines
536 QUANTUM ELECTRODYNAMICS

the fine structure of the positronium levels does not contain such ele­
ments. Therefore, in weak fields when {ehjmc)H < W0\, where
Wx and WQ are the values of the energy for the positronium triplet
and singlet levels, there is no effect proportional to the field.
In the opposite limiting case of a very strong field the interaction
between the spins may be neglected and the splitting of the levels will
be given by ehjmc)H corresponding to states described by the wave
functions a(l)/3(2) and a(2)/?(l).
For the ground state (and also for 5-states with n —2 and 3) we can
easily write the exact formula for splitting in a magnetic field. In
this case the wave functions for the stationary states are given by
superpositions of the functions %10, Xoo '•
%{i) = C P x i o + C P x n , *'= 1,2, (39.22)
where i = 1 corresponds to the state which reduces to the triplet state
in the absence of a magnetic field (C)1} = 1, C^1) = 0, H = 0), while
i — 2 corresponds to the state which reduces to the singlet state for
(Cj2) = 0, C£2) — \, H = 0). The coefficients C (i) satisfy the equations
pfi
------ HCP = 0,
me
(39.23)
eti
HC[i)Jr( E {i)— W0) Cq0 = 0,
me
where E (i) is the positronium energy in the field H.
It follows from (39.23) that
w< -w0 ± / ( w , - w af
E U) = (39.24)

The ratio |C |i)/C'(5i)|2 defines the relative weight of the triplet and
singlet states in a given stationary state of positronium in a magnetic field.
The magnetic field has an important effect on the probability of
positronium decay. We denote by iv0 and ivx the decay probabilities
per unit time for para- and orthopositronium. In the presence of a magnet­
ic field the probability of positronium decay in the state %{i) is given by
= IC l^W n+IC ^I2*,, (]C{1)|2+ |C ^ )|2 = 1). (39.25)
For the state %{1) (triplet state in the absence of the magnetic field)
even a small value of the coefficient C0 greatly increases the probability
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 537

of decay, since (cf. § 33) vv0 ;> Wj. In a weak magnetic field we have
in accordance with (39.23), (39.24)

C '1’ 2_ [{etijmc)Hf
(39.26)
CF ~
A measurement of the effect of the magnetic field on positronium
decay enables us, by utilizing formula (39.26), to determine experimentally
(45) the energy difference between the ground levels of ortho- and
parapositronium A = E10—E00. The experimentally found value of
A agrees with the theoretical value (39.21).

§ 40. Internal Conversion of Gamma-Rays

40.1. Expansion of Retarded Potentials in Spherical Waves


Because of its electromagnetic interaction with external electrons
a nucleus in an excited state may make a transition to a state of lower
energy giving up its excitation energy to the electrons in the atomic
shells, or producing an electron-positron pair. Processes which are not
accompanied by emission of radiation have the general name of internal
conversion of y rays. A similar phenomenon in atomic shells, i.e., an
interaction between two electrons as a result of which one electron
makes a transition to a state of lower energy, while the other makes
a transition into the continuum, is called the Auger effect.
The phenomenon of internal conversion in atomic shells may be
investigated on the basis of the retarded interaction between two charges
described in § 37, one of the two charges being an atomic electron,
while the other one is a proton in the nucleus.
We denote the electron wave functions in the initial and final states
by and y)2> and the proton wave functions correspondingly by
and W2. In accordance with (37.8) and (37.15) the matrix element
of the effective perturbing energy is given by
g i e u |r i - r ,| _
/
xl f f i ) y p 0i) - j~T V’ 2 (r2) y„ y>i(r2) drxdr2
\r 1 *2I
/ gfcoli-i-ral
Oi) w t O2) (1 —a i a2) - T ------ r ^ lO i) Yhfo) drx dr2,
\ri —' 2 1 (40.1)
538 Q U A N T U M E L E C T R O D Y N A M IC S

where co = E1—E2 = e2—e1, a = e2/4jc= 1/137, a x and a 2 are Dirac


matrices which operate respectively on the proton and the electron
spin variables, Ei2 are the proton energies, while e1 2 are the electron
energies.
The evaluation of these integrals is facilitated by the fact that the
dimensions of the regions within which the proton and the electron
motions take place are different. For the proton this region is of the
order of magnitude of the nuclear radius. But for the electron the anal­
ogous region is of considerably greater dimensions. Therefore, we can
assume that in the integrals of (40.1) the important contributions come
from a range of variation of the variables for which r1 < r 2. Under
this condition the following expansion holds1:

gl<n\ri —r , |
(40.2)
ri~r2 lm
where

<Pim(r) = gi(<or)Ylmy j j ,

®,nSr)= Gl(Mr) Ylm


(40.3)

VZ

G,(z) =
Vz
and Jn(z) are Bessel functions, H ^ { z ) are Hankel functions of the
first kind, and Ylm are spherical harmonics. We now show that the
following expansion also holds:
giailr,—
r,|
“ l « 2 (40.4)
\ri - r 2\ LIM

1 T h e e x p a n s io n (4 0 .2 ) is th e a d d itio n th e o r e m fo r c y lin d r ic a l fu n c tio n s a p p lie d


to th e fu n c tio n

2
Tfl/2 (w |ri —r2|) e i< u |r , - r j |
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 539

where
LI MSl(COf') YL l M ’
(40.5)
Bum ~ Gt(cor) YLlM,

a! and a 2 are arbitrary vectors which do not depend on the coordinates,


and YLlM are vector spherical harmonics defined in § 4. We note in this
connection that any vector function can be expanded in terms of a com­
plete system of vector spherical harmonics YLlM. Thus, the vector

regarded as a function of r2jr2 can be written in the form

(40.6)

In order to determine the expansion coefficients CL,M we multiply this


equation by Y*VM.{r2jr2) and integrate over all directions of the unit
vector r2/r2. By utilizing expansion (40.2) and also the orthogonality
property of vector spherical harmonics, and their relation to spherical
harmonics (cf. § 4) we obtain
CL!M (rv r2) = (a 1A f l„ (r2))G, (cor2).

Substituting this expression into the expansion (40.6), and multiplying


it by a 2 we obtain (40.4).
Instead of the set of vectors A LIM and BLlM (for a given L and
M , l — L ± 1, L) we can use the vectors a^ defined in (27.6), (27.16),
and similarly defined vectors B fy (the difference between B fy and afy
consists of the replacement of g, by G,). The vectors afy and B fy have the
property that, as was shown in §4, they arc asymptotically (for 1)
transverse ( f o r / = 0, 1) or longitudinal (for A= —1). It can be easily
shown that
L+ 1

X («1 ' ) ( ^ b l!m) = X (*2b [X


)m)-
l=L~ 1
On substituting these expansions into (40.1) we obtain
540 Q U A N T U M E L E C T R O D Y N A M IC S

where brackets denote matrix elements of corresponding quantities


evaluated using electron or proton states:

(.<Plu)n=

(^ L m ) £1 = J ’Pz'C’ I.u’pld? .

(“ S S V = lV i*
(In the future we shall sometimes omit the subscripts from these brackets.)
Expression (40.7) can be simplified if we use certain relations obtained
in § 27. In accordance with (27.8) and (27.10') we have (aa
= ((-pfM). Moreover, under the condition coR < 1, which we assume to
hold (R is the nuclear radius), according to (27.7)
a t-1) = i / _ L ad)
LM \ L + l LM'
In virtue of this, three of the four terms appearing in the figure brackets
of (40.7) can be combined in the manner
K i } ' ) - « ) («*& ) = (««&•) (* & ), (40.8)
where

= - |/ r - ] / ^ « * * .« .- <40'9>
In the future we shall also use the notation
(40-10)
We note that in spite of the fact that B ^ and 0 lM are related by an
expression analogous to (27.8):
» M I — ! _ XJ(f)
d L M ------------ v ^ i M >
CO

the matrix elements ( 0 LM)2i and (afi|_^1))2l cannot be expressed in terms


of each other in analogy with (27.10'), since the transformations which
we used in § 27 are inadmissible in this case because of the singularity
of the function <PLM. Indeed, due to the presence of a pole at the point
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 541

r = 0, in the course of integration by parts which is involved in the


transition from (27.8) to (27.10') there remains an integral over the
surface surrounding the pole which does not vanish as r -> 0. In other
words, the operator H is not self-conjugate with respect to the functions
0
^ L M •

Thus, the matrix element corresponding to internal conversion can


be written in the form
uo
Ut ■r — —a-4 tz 2 {(”“>&*)*1W .!)*!+ (“ " I D n U © » } ■ (40.11)
LM

40.2. Conversion Coefficient


The expression for the conversion matrix element (40.11) admits
a simple physical interpretation. Here (aaj^)21 and (aaj°^)21 are, apart
from numerical coefficients, the matrix elements for the emission of
a photon of angular momentum L and parity (—1)£ or (—l)-t+1 which
are proportional to the matrix element of the electric or the magnetic
2L -pole nuclear moment (cf. § 27). The operators B fy for which the
matrix elements are evaluated with respect to the electron states can be
interpreted as the perturbation energy due to the field associated with
the nuclear transition. The quantities B fy and 0 LM can be called the
potentials of the electric or the magnetic multipole. Indeed, the poten­
tials 0 LM and B\fy are, as can be easily shown, solutions of the wave
equation (37.17) having the form of outgoing spherical waves with
a singularity at the origin. This singularity corresponds to the presence
at the point r = 0 of a radiating electric or magnetic multipole of the
L-th order (2L-pole). The process of conversion can be interpreted as
the absorption of these waves by the electrons near the nucleus.
We note that the expression (Bfy) differs from the matrix element
of the operator A0LMX—a A Lm , which corresponds to the absorption
of a photon of definite angular momentum and parity only by the re­
placement of the regular function g, by the function Gt which has
a singularity at r = 0. It should be noted that the potentials of the
electric multipole (40.9) have the structure of (6.18) which corresponds
to a definite choice of the arbitrary constant in the general expression
for the photon potentials (6.17). An arbitrary choice of the constant
in formula (6.17) for ALMk, which is made possible by gauge invariance,
is not allowed in this case because of the singular nature of the potentials.
542 QUANTUM ELECTRODYNAMICS

Let the angular momentum of the nucleus in the initial state be


J1, and in the final state J2. Then the matrix element (oca<^)21 will
differ from zero if the condition \J1—J2\ ^ L ^ J 1-\-J2 is satisfied.
If ioR<< 1 then only the smallest possible value of L is important. For
a given value of L the matrix elements ( a a ^ ) 2i and ( a a ^ ) 21 cannot
both simultaneously differ from zero in virtue of the parity selection
rule. Cases are possible when both an electric and a magnetic multipole,
of orders differing by unity, take part in the process, for example
(aa(L]^)2l and (aaj.1^ M)21•However, we shall not investigate such processes
here. An estimate of such matrix elements (cf. § 27) leads to the conclusion
that, generally speaking, they are of different orders of magnitude.
Therefore, in the majority, of cases the process of conversion is deter­
mined by only a single term from the sum (40.11). The general expression
for the conversion probability
w = 2jr|i7w |2 S(EX— E 2—^ + £ 2)

can be simplified as a result of this.


In the case when the angular momentum of the nucleus changes1
by L
| J , ~ J 21= L

and parity changes by (—I)1, the conversion probability is determined


by the potential of the electric 2L-pole and is equal to

^Z1’ = 0)21 )2 1 12 I I2 <K £i~£2—£i+ e2) .

In the case when the angular momentum of the nucleus changes


by Land parity changes by (—l)i f l the conversion probability is deter­
mined by the potential of the magnetic 2L-pole, and is equal to

<■ = lU O n l* 5(z-,--B2- £l+ £ 2).

The conversion coefficient is defined as the ratio of the probability


of the conversion process to the probability of photon emission in the

1 If L = 0, and th e p a r ity ch an ges, th e n th e d ip o le tr a n s itio n (Z , = 1)

is th e m o s t p r o b a b le one. For th e ca se = J2 w i t h o u t a change o f p a r ity c f. § 4 2 .


R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 543

same nuclear transition. In accordance with the results of § 27 the prob­


ability of emission is given by

,U> _ aco I(“ 0^)2112.


L ra d
271
Therefore, the conversion coefficient in the case of an electric (2 = 1)
or a magnetic ( 2 = 0 ) 2i,-pole is given by
a
ft' = H E 1~ E 2- s 1+ez). (40.12)
^Zrad H
It is assumed here that the electron wave functions are normalized
so that f iff if i dr = f ip* ip2 dr = 1.
If the wave function ip2 (belonging to the continuous spectrum)
is normalized per unit energy, i.e., if we assume that / ip*ipe' dr = d(e—e'),
and at the same time retain the former normalization for ip1, the expres­
sion for the conversion coefficient assumes the form
ato 2
v ZBlmWi dr (40.13)
~4

40.3. Conversion in the K-shell


It is important that the conversion coefficient contains no
quantities which refer to nuclear properties. In order to evaluate it,
a knowledge of only the electron wave functions is sufficient. If in the
initial state the electron is in the ^f-shell, then its conversion coefficient
is equal to

K f

Here 2 denotes one of the two indices (1 or 0) which indicate whether


the transition is of electric or magnetic type; Z K denotes summation
over the two electrons of the /f-shell, and Z f denotes summation over
the possible final states, i.e., those allowed by the selection rules. In
order to simplify the calculations we can use the fact that for a closed
shell the conversion coefficient cannot depend on the magnetic quantum
number of the electron m. Indeed, the latter depends on the choice of
the z-axis, while a closed shell is spherically symmetrical. We can, there­
fore, determine assuming M = 0. Moreover, since the pertur-
544 QUANTUM ELECTRODYNAMICS

bation does not depend on the azimuth the conversion probability


for M = 0 cannot be different for two electrons which differ only
by the quantum number Therefore,

f
The final state must have the energy e2 = co—|% |, where | % | = —eK
is the binding energy of the electron in the X-shell (conversion can
take place if a> > |%|).
In order to determine the angular momenta of the final states we
can use the laws of conservation of angular momentum and of parity.
These laws show that the change in the angular momentum of the
nucleus will be exactly the same as in the case of emission of a photon
of angular momentum L and of parity (—I)1, or (—l)i+1, and that
the selection rules for the final state of the electron will be the same
as in the case of absorption of such a photon.
This can also be shown directly by a consideration of the angular
part of the integral in the matrix element (40.11) which coincides with
the angular part of the matrix element for photon absorption.
Since for the electron in the initial state (A'-shell) we have \
and lx = 0, then in the final state in the case of an electric multipole we
have /2 = L md j 2 = /2 ± f = L i J .
In the case of a magnetic multipole we have either /2= L + 1 ,
y2 = 12 ~ \ = L + \ or /2 = L —1 and j 2 = l2+ \ = L —\ . Thus

= f m w v ) l+ i.
(40.14)
= I ■<“ {I( * s ’)i+ , , I + L P+l M V , I*}.

(The two indices on the matrix elements indicate the quantum numbers
of the final state of the electron: the first is j 2, the second is /2.)
For the electron wave functions we use the general expression (12.4)
igMir)Qjln{ri) \
(40.15)

where
r
n
r l ' = V —h Q iVm = - ( o n ) Q jlm.
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 545

In particular, the electron wave function for the Ai-shell { j — \ ,


1 = 0 , x = —1) can be written in the form

W = ( ig*(r)u \, (40.16)
\ f K(r)(an)u)
1
where u is a constant spinor normalized as u*u = On utilizing the
471 '
definition (40.9) for the quantity B fy , we obtain

— ~^f ~ j / L-\-\ (40-17)


On substituting into this equation the expressions for the wave functions
(40.15) and (40.16) and for the potentials (40.3), (40.5), we obtain

( A o ) „ ,= j ( g t g K + f Z f J O ^ d r u l j . u (40.18)
where the following notation has been introduced:

u* = J Q*lmYl0do. (40.19)
(We note that for I = 0, j = \ and un coincides with n.)
The second term in (40.17) contains the quantity

+ ' i f l.Sv 1),_,,</,)) u

- i f » / * ( } *.,...»(<>») (40.20)
The integrals over the angles in this expression can be brought into
the form (40.19). Indeed, we expand the vector spherical harmonic
Yl L_ito into a transverse and a longitudinal component

YL,L. 1,0 i / i d - 1^ y<xo


i)

-
X 2L-
Since (an)a = n+ i [o n ], we have

(< jli)o Y L , = y ^ ± r i n Y t f ' + X

On expressing the longitudinal and the transverse vector spherical


harmonics in terms of the sphesical harmonic

iv> = - a » n 1.l) = n 0.1= L^ °


546 QUANTUM ELECTRODYNAMICS

(L = —/[f(7 ] is the operator for the orbital angular momentum),


we obtain

(0',)°y— = ( ] / ^ j - T z f c ) y“'
Similary

o(" ) =( f e + jwm) n °
-
Moreover, making use of the fact that the operator La is self-conju­
gate, we transform the integral
/ ^ * , m/ " ) L oyzo(*)“ do
in such a way that the operator La acts on the function QhUmi which
is its own eigenfunction corresponding to the eigenvalue j i { k Jr^)~~h
(4 + iH -
We then obtain the following values for the integrals over the angles
appearing in (40.20):

/ QLmAan)aYL,L-1.0 u do
_ ul,u j 0 for j2= L+ \,
]/L (2 L + l) \ 2 L + 1 for j 2= L —\ ,
(40.21)
/ Mdo
uj,u j 2L for j 2=
= ]/L(2 Z+T) [ - I for j 2 = L —\,
(h = L)-
On substituting (40.21) into (40.20) and (40.18) into (40.17) we
obtain finally the following expression for the matrix element which
determines the internal conversion coefficient for an electric multipole:

( T O .Z + I, L = (R y+ Ri-U Rd,

k —£+ 2 ’
(40.22)
(® z V )z -L Z “ z * - I .z “ ] / £ _ j _ L

h• _~ Lr y1,
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 547

where Rt denote the integrals


CO

Ri = f g t g KGLr2dr,
0

c o

* 2 = / f Z f KGLr*dr,
0
(40.23)
c o

* 3 = / f t g KGL- ^ d r ,

R i = j g t f K G L- i r 2 d r .

The quantity u*u appearing in (40.22) can be easily evaluated if


in the integral (40.19) we express the spinor spherical harmonic in terms
of the spherical harmonic in accordance with the formulas of § 4.
As we have already mentioned, the value of the quantum number
mx (the component of the angular momentum of the electron in the
initial state) can be taken equal to for the Ai-shell. Then m2 will also
be equal to In this case
1 /1\
u
\/4n \o]
and
1 \/L + l for Jz — L + i ,
u*h u — ]/47i (2L+1) (40.24)
-/L
jz = L ¥• for
On substituting (40.22) and (40.23) into (40.13) we obtain the following
expression for the internal conversion coefficient for an electric multipole
in the /f-shell:
aa) L
1-^1 \~RZ 2i'7?4| +
2>n I 2Z.+ 1
U 2L4-1 . 1 !2
+ \RX+ R 2— i- i?3 i — Rx, (40.25)
(2 L + 1 )(L + 1 )|“ 1 ' ' L \h = L-i, la=L

The matrix element which determines the conversion for a magnetic


multipole is of the form
(/*!?),„,= i j f i g ^ d r j
j (40.26)
548 QUANTUM ELECTRODYNAMICS

The angular integrals appearing in this expression can also be re­


duced to an integral of type (40.19). In order to do this we must express
Y[ i n terms of the spherical harmonic YL0 and utilize the fact that

Thus, we obtain

j QiumS™)°VtVu do = - J Y lou do

_ ~y20'2+ l) + ^(^2 + l) + f
\ L( L \ 1)

The second integral over the angles in (40.26) differs from the first
one only by its sign.
On introducing the notation for the radial integrals

^ 5 = J f * g KGLr*dr,
(40.27)
* e = / g * f KGLr2dr,

we can write (40.26) in the form


! _ (L for y2 — ■ £ ' + ' 2 >
iu *1'u (R5 +
L (L + 1 )\ - ( L + l ) for j 2= L —i ,

where u*ju is defined by formula (40.24).


The internal conversion coefficient for a magnetic multipole in the
Ai-shell is given by

^ 0> = "8+ 2L-{-T * L 1* 5+ R«'** ,2=i+1


+ ( L + 1) | R5+ R g\*i=L_h . (40.28)

The radial integrals (40.23) and (40.27) which appear in the expres­
sions for the conversion coefficients (40.25) and (40.28) can be simply
evaluated only in the limiting case of low nuclear charge Ze and of
high transition frequencies co. In such a case we can, on the one hand,
neglect the magnitude of the electron binding energy in the Ai-shell and,
on the other hand, neglect the effect of the nuclear charge on the
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 549

electron wave function in the continuum (final state). In this approxi­


mation we have
gK= 2 ( Z a m f 12, f K= 0,
e + m J,2+i(pr)
■= l / - 2 ,/7 (40.29)

"=W\V
l'i = 2y2—/2,
e-m
2
p = | / £2—m2,
Ji2+i(pr)
j/7
e = cu+m.
The substitution of (40.29) into (40.23) and (40.27) leads to the
following expressions for the radial integrals

Rx = iL(2nZam)312 - j ^ / 7i.»

f?2= 0 ,
I 2( e - m ) T
f?t1= iL 1(2nZam)3,‘2 ' 4 - i)
(0 V
*4=0,

i?5 = iL(2nZam)312

*6=0,
/ 2(e- m )
OJ
x
x

where

h = j JL+ i ( p r ) H ^ i ((or) rd r-
This integral is equal to

L ni \ a>I p2—or

On utilizing these formulas we obtain the following expressions


for the internal conversion coefficients:

(40.30)
ffi°>=2a(Za)»(l + ^ P " ; .
550 QUANTUM ELECTRODYNAMICS

If in these formulas we go over to the nonrelativistic approximation


(o) < m), we obtain
U-H

(40.31)

Formulas (40.30) and (40.31) give a quantitative idea of the nature


of the dependence of the conversion coefficients on the energy and
the angular momentum. The coefficients are large for small a>,
and for large L. In some cases ffff can considerably exceed unity.
However, quantitatively the free electron approximation gives very
low accuracy even for small Z and large o>. The magnitude of Rt can
be evaluated if we use the nonrelativistic approximation for the electron
wave functions (78). However, the accuracy of such results is not very
great, since even at low energies the electron wave functions differ
appreciably from the nonrelativistic ones at small values of r, which
turns out to play the most important role in the integrals Rt.
If we use exact electron wave functions Rt can be evaluated only
numerically (184).

40.4. Effect of Finite Nuclear Size


In a number of cases in evaluating the internal conversion coeffi­
cients we must take finite nuclear size into account. This introduces
two types of corrections into the evaluation of the matrix element which
determines the conversion probability. Firstly, in the region r2 < R
(R is the nuclear radius) expansion (40.2) is not valid. Therefore,
the exact expression for the matrix element for transitions of given
L will have the following form instead of (40.11):

Ui ^ f = ~ a j ^ ( aaL ))n J V’t O2) xpfrf dr2

eia>\ri-r,\
yjffr2) ) -------- drxdr2. (40.32)
I' 1 ' 2 1

(In the expressions obtained previously it was assumed that R 0.)


R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 551

Secondly, the electron wave functions near the nucleus may, as


we have seen in § 13, differ significantly from the electron wave functions
in the Coulomb field. As actual analysis shows, the main error is due
to extending the integral in the first term of (40.32) into the region
0 < r2 < R, and using Coulomb wave functions in this region. The
latter have a singularity at r = 0, and therefore the region 0 < r2 < R
in the case of large Z makes a contribution to the integral comparable
to that of the region R < r 2 < o o . But the true wave functions have
no singularities at r — 0 (rip -* 0 as r -* 0), and therefore, in fact,
the integral over the region r2 < R must be small compared to the
integral over the region r2 > R. Results good for practical purposes
can be obtained from the expression for the matrix element (183)

(40.33)
TZ> R

into which Coulomb wave functions can be substituted. Thus, the lower
limit r2 — R should be introduced into the integrals Taking finite
nuclear size into account leads to corrections which are inappreciable
for low values of Z, but which attain values of 30-40% at Z>—-80-90.

40.5. Effect of Electron Shells on Radiation from the Nucleus


In § 27 we have obtained expressions for the probability of emission
of a photon by a nucleus. The presence of electron shells introduces
corrections into these formulas. In addition to direct emission of
a photon by the nucleus, represented by the diagram of Fig. 53, there

F ig . 53.

also exists the possibility of the process of emission by an electron be­


longing to an atomic shell, due to its interaction with the nucleus.
Such a process is represented by the diagram of Fig. 54.
552 QUANTUM ELECTRODYNAMICS

Since the latter process is a third order process we can, as a rule,


neglect it in comparison with the process of direct emission which
is a first order process. However, this may turn out not to be so when
the conversion coefficient is very large (110). Indeed, the matrix element
for the process shown by the diagram of Fig. 54 amounts, roughly

speaking, to the product of the matrix element for the internal conversion
process and the matrix element for the emission of radiation. If the
conversion process is more probable than emission, then the effect
of the electron shells on the emission by the nucleus may also turn
out to be important.
We investigate in greater detail the process represented by the dia­
gram of Fig. 54, in which the thick lines represent the proton. As
we have done before we take W1(x) = (/) exp (JEYt) and
= !^2(i*) exp (iE2t) to denote respectively the proton wave functions
in the initial and the final states. We assume the initial and the final
states of the electron to be the same, and denote the electron wave
function by rp0(x) — y)0(r) exp (ie0t), and the vector potential of the
emitted photon by A(x) = A(r) exp (—mi). The matrix element for
this process (cf. Fig. 54) will be expressed as follows:

S \% = e3 j V/2(x1) y /y \ ( x 1)Dc(x1- x 2)^0(x2)

X S (ce) (x2, x 3) A* (x3) y 0(x3) di x 1 d*x2d*x3

+ e3 f if /2(x1)yfltf /1(x1)Dc(x1~ x 3)y}0(x2)A(x2) S ^ )(x2, x 3)

X y fly)0(x3)di x1d‘i x2di x3.


R E T A R D E D IN T E R A C T IO N B E T W E E N TW O C H A R G E S 553

On substituting into this equation the expression for Dc(x) in the form
(37.10)
CO

D
c(
x.
i-Xa) = — f d
(
Q"
— oo

and for S {ce) in the form (35.3)


OO

— — — dm’,
(O + £n

where rpn are the electron wave functions in the nuclear field, and n is
a set of appropriate quantum numbers, we obtain the expression for
S<3)
S \ % = - l n i d ( E 1~ E 2-co)Uif,

TT V*
— r/(conv) rr(rad) I ____ ^____

u ~ 4* n n U -—
« £nH
o +-CO E —ea—0)
where
/ i(Ei—Ei) ]ri—r
^2(»*i) yJP i(ri)—.-........ voC1*)
Ir l r 2\
represents the matrix element for the process of internal conversion
(40.3) in which the electron makes a transition from the initial state to
state n, and
d) = e f yj0(r)A(r)ipn(r)dr

represents the matrix element for the emission of radiation, in which


the electron makes a transition from the state n to the initial state.
In the case of states belonging to the continuous spectrum we nor­
malize ipn per unit energy, i.e.,

/ K ( r)Vnt(r) dr = 6(£n - £ n ,) d<K>


where n' are all the quantum numbers, except for the energy, i.e., the
quantum numbers for the orbital, spin and total angular momenta.
In this case Uif will be given in the following form:

— V 1 I rr(conv) r/(rad) I
de
u\V J n n \ £-£ 0 + CO
+ ' £ —£n— CO

1 1
+ % jy(conv) rad)
+ ■
£n- £ 0+co en—e0—a>
554 QUANTUM ELECTRODYNAMICS

Here the first sum refers to the continuous spectrum, and the second
to the discrete spectrum.
The principal role is played by the resonance region which lies in
the continuous spectrum near the energy value e = e0+a». Therefore
(198) we have
U \ f = in ? , U{nconv) U ( b £0 =
n'

Thus, the matrix element for the emission of radiation, taking


into account the effect of the electron shells, has the form
U\}' + U # ' = U (1) + i7l ^[/(eonvl^rad), (40.34)
n'

where U(1) is the matrix element for the emission of radiation by a nucleus
not surrounded by electrons. If the conversion coefficient /9 = ] {/(conv)/
U(1)|2, evaluated without taking the effect of the shells into account,
is large, then the second term in (40.34) may play an important role.

§ 41. Conversion Accompanied by Pair Production. Excitation of


Nuclei by Electrons

41.1 Conversion of Magnetic Multipole Radiation


If the nuclear excitation energy a> exceeds 2m, then in addition to
the phenomenon of internal conversion in the atomic shells creation
of electron-positron pairs may also occur. The coefficient of conversion
accompanied by pair creation can be determined from the general
formula (40.13) in which for rp2 we take an electron wave function, while
for y)1 we take a wave function of negative frequency corresponding
to a positron. The matrix element appearing in (40.13) can be reduced
to integrals over radial wave functions similarly to the way this was
done in the preceding section in the case of conversion in the A-shell.
We confine ourselves here to the investigation of only the one case
when the electron and the positron states may be regarded as free (15),
(178), (165). Such an approximation is applicable for small Z.
We choose the electron and the positron wave functions in the
form of plane waves
xp2 = ue?~r ,
(41.1)
y)± — ve >p+r
R E T A R D E D IN T E R A C T IO N B E T W E E N TW O C H A R G E S 555

where p_ is the electron momentum, p + is the positron momentum,


and u and v are unit bispinor amplitudes.
On substituting (41.1) into (40.12) we obtain the following expression
for the differential conversion coefficient for an electric or a magnetic
multipole accompanied by the creation of an electron in the momentum
interval dp_ and of a positron in the momentum interval dp+:

(41.2)

where q — p_-\-p+ and X denotes summation over the spin states


/<!■El
of the electron and the positron.
We begin our calculation with the case of a magnetic multipole.
In this case, in accordance with (40.10), we have
B ' & = -aY<°>G£(a>r).
Further, on utilizing the expansion of exp (iqr) in terms of spherical
harmonics we obtain
eriqrY ^ G Ldr = ( - \ Y Y ^ ( v ) I L,

where v =

h c ‘ w 6 ‘ w r! d r = 4 S H ' 2) ( « ) -
Thus, we have
a 1
dpj_dp_ ^ \ u * o. Y ^ ( ) \2 5 { o - > - £ _ - e+ ).
dPi0) = n 2 (or —q2)2co v v

H i, Hi
(41.3)
The summation over the spin states may be carried out in the usual
manner in accordance with § 26. It yields

v i-<*« F™.rj

+ {p- o ( p + y ! ° » ) + ( p - n s;* ) ( p + o i •
This expression can be simplified. Since p + = q ~ P - and qYj°jJ= 0
we have p_ Y l° J = —p+ Yj0^ . Further, we have p _ p + = p +q — p+- On
the other hand, from the law of conservation of energy
co = ] / ( p 2_ ± m 2) + |/(tf—/ O H
556 Q U A N T U M E L E C T R O D Y N A M IC S

we can obtain p +q = coe+—^(co2—q2). Therefore

We substitute this expression into (41.3) and go over from the variables
p p _ to the variables q, p +. We write the product of the differentials
of the momenta in the form
dp+dp_ = dp+dq = p \ d p +do+q2dq doQ,
and the argument of the 5-function appearing in (41.3) in the form

co— e+ = co—£+ — \/q2jrp\-\-m2~2qp+ cos# ,


where # is the angle between q and p. If we take the z-axis along q
hen do+ = d cos■&d<p+ and we can eliminate the 5-function:
£
5(co—£ , —e_)d(cos#) -> ——
p+q
Now the angle # is determined by the law of conservation of energy:
(co—£+)2 = q2+ e \ —2qp+ cos#. (41.4)
Formula (41.3) assumes the form
a Iy ( ° ) | 2
W i0) = n 2{co2—q2)2 I -* LM '

~ 2P+\ Y lm\2 sin de+ dq dcp+ doQ. (41.5)

We now integrate (41.5) over dcp+ and doq. On utilizing for sin2#
the value given by (41.4) we obtain the following expression for the
coefficient of conversion accompanied by creation of a pair with given
values of the positron energy £+ and of the magnitude of the total momen­
tum q\
2a q ^ f 1 p l+ p l co2 \
« ( £+, q) (41.6)
n co2L+1 \ 4 ^ 2 ( c o * - q 2j ^ ( c o 2- q 2)2\ + 9
On integrating this expression over q between the limits qmln to qmaK,
where
9 m i n = l/> 4 — P - 1 ’ = \P + + P -\ ,
RETARDED INTERACTION BETWEEN TWO CHARGES 557

we obtain the energy distribution of the converted positrons:

a I m2-\-p,p + £ ,£
TUX) r moj

- ^ l ( P + + P - y L- ( P + - P - y L]+ j [ ( e +e_+p+p_—m2)(p++ p _ y (L- 1)

- (e+e _ - p +p _ - m 2) (p+ - / 0 2(i"1)] - —[p\Jrp 2L ~ 2 { L — 1) m2]

L~1 w 2 ( Z ,- n - l) \

x£ -----~
n-----[(P++P~)2n- ( P +- P - T n) de+. (41.7)

We give in explicit form the corresponding expressions for the cases


of a magnetic dipole (L = 1), quadrupole (L = 2) and octopole (L — 3):

dP?>(e+) = ( p l + p l )In J ^ ± P ± P - + V - d ■
T 710)'* f f lf t l ^

^ 2 0)(e+) = —~5 {a)2(/*+ + .P -—2w2) ln ^ - + -P+/7~+6+£-


T 7zco° I ma>

+P+P-(co2—3 p l —3p2_)j d£+ ; (41.8)

d ^ ( e +) = — \o)i {p\-\-p2_ —4m2) \n~m2+P+P- + E^


^ nar 1 mco

+ y P+P- [w2(o»2+ 8m2) —2o)z(p2. +/?i) - 5 (p^_+ p i) 2]J<fe+.

The total conversion coefficient /3^0) will be obtained by integrating


(41.7) over the range from m to u>—m. The result of this integration
in the general case cannot be expressed in terms of elementary functions.
Later we shall give the results of numerical integration. However,
we shall first investigate the extreme relativistic case o> m when we
can approximately take p + m e +, p _ ^ e _ . Assuming, moreover that
(o>/m)2 > L we obtain the expression for the total conversion coeffi­
cient

;(o) = 3-____V _____4" + 5 1 (41.9)


1 in \ m 12 Z (2Z + 1) Z j 2 (2 n + l)( 2 n + 3 )
L n= l J
558 QUANTUM ELECTRODYNAMICS

In the opposite limiting case L > (o)/m)2 (for which also co > m)
we obtain the asymptotic expression for /3[0),

2+yJ, (41.10)
710)“ or
where y is Euler’s constant.
The distribution over the angle % between the momenta of the com­
ponents of the pair can be obtained on going over in (41.6) from the
variable q to %, and on integrating the resultant expression over the
energy e+. The relation between y and q follows from the definition
of q :
q2 = p 2++ p l —2p+p_ cos %.

As an example of the application of these formulas for the conversion


coefficients we now obtain the cross section for the capture of slow
neutrons by protons resulting in the formation of a deuteron, an electron
and a positron. On taking into account the fact that in such a process
a transition occurs from a 1S state to a 3S state, we can easily conclude
that the transition is of a magnetic dipole nature. Therefore, we have
cr= <JC(3[0), where a is the cross section for the process of interest to us,
while ac is the photocapture cross section, which in the case of thermal
neutrons is equal to ac 0.3 • 10“24 cm2. The binding energy of the deuter­
on is given by |e| = 2.15 MeV, i.e., co = 4.2m. Numerical integration
carried out for this value of co yields for /S[0> the value 3 x 10~4. Con­
sequently, we have a = 0.9TO”28 cm2.

41.2. Conversion of Electric Multipole Radiation


We now proceed to investigate the conversion of electric multipole
radiation. On substituting into (40.12)

B (1) —
^L M (pr) Yl „ - y ^ + 1 GL^ ( a r ) YL^ . M*
V IT T ° L

with the same wave functions (41.1) we obtain


(— \)L^(4Ti)2i ( q \ L 1
(* $ ) = v
co(q2 co2) \o) J j/ l +T _
CO

]/2L + 1 y y^ -L M u*a ^ |-
R E T A R D E D IN T E R A C T IO N B E T W E E N TW O C H A R G E S 559

Substitution of this expression into (40.12) and integration of (40.12)


over the angle between p + and q yields the expression for the differential
conversion coefficient
21,+1
__________
dPi11 = 7z 2 co* l + 1 ( co2 — q 2) 2 ( L + 1) I } /L Y l m u * v >

V 2L + 1 — Yl.l - i. e+e_de+dq doQdcp+, (41.11)

where £ denotes summation over the electron polarizations. On car­


rying out the summation, we obtain

£ I ■"I2 = L l Ylm\2(£+s -+P+P--™*)+(2L+ 1 ) ^ \ Y l,l- i ,m\2

X(s+e _ - p +p _ + m ^ - / L ( 2 L - i - l ) ^ ( Y LMYL*L_ liM+ Y * MYLiL_liM)


*/
X (e_p+-\-e+p_)-\-(2L-\-l')— [(YL L_l Mp_) (Y * L_l Mp +)
H

The products {YL L^ X Mp +) appearing in this expression can be easily


evaluated if we use the expansion of TZiI,-1iM into a longitudinal and a
transverse vector spherical harmonic. For the longitudinal vector we have
y lm1)P+ = P + Y l m co s & ,
y lm 1)P - = ( ? - P + c o s # ) 7 xm,
and for the transverse one
Y®p+ = - Y $ p _ ,
\ Y & P + \ * = \ Y & \ * f a m ' f i .

Substituting these expressions into (41.11) and integrating over d(p+


and doQ we finally obtain
2a (3L+l)cu2 1
dffl'(e+>9) 2Le+e
7i (L-\-1) 2 00“- q 2

2L(J02£ + £ _ \(
((L + 1)e+£_-\-L(jo2
K = ? ) 2“ J +

+ ( Z + l ) m!) ^ ^ + ^ ( 2 L £+e_ + (L + l)m ')(^ i ^ ] } * + rf?.


(41.12)
560 QUANTUM ELECTRODYNAMICS

M Cl Cl pi ei M Cl Cl Cl Cl
o 1o— o 1o-^ o O o o O O o o
o(N O
(N
VO VO vo <7\ r- vo «(N VO (N 00 00 r-
o (N
CO o 100 vo vo -H VO (N Os
(N CN —< i-H CO (N CN
d o o o o o o o O o o o
N (4 N Cl N r» Cl Cl Cl Cl CJ PS
o o o o 1o-H o o o o o o O
vo vo
CO ON
S r- O
CN r- (N On VO Os
os
*—1 1S—I 5
C l-H V vo
(N CN
r- VO CO 8
< Os
*—
d o d o o o o o o O o o'
Cl Cl ps rs PS PS CJ Cl Cl ei PS PS
o i-Ho © o © O O o 1-^
o o o ©
O O
Tj- OO (N o oo CN vo (N VO VO
1CO1 o O VO oN vo CN 8 vo
— Os r- vo CO < *—» oo Tf
d O © o o d o o O d O O
PS pj n « PJ PS to Cl Cl PS PS PS PS
o o © o o O o»—<o o o 1 o-H o
Values of total conversion coefficients

»—<
r- •*»* r-
Tj- r- 1
-H o r- oCO V) ovo VO <r» vo
VO CO Os ov
CO OV o1-H Tf r- oo
oo VO (N i—i Si? r- VY
© o o O O o o o o o o O o
SJ
ps ps ps ps n Cl PS PS PS PS *p
o © © © o oi-H **«. o o o o O o
T a b le 13

o
VO £ vo
VO VO <N CO oo
o vo 00
CN r-
r- r- Tf C
VO rN vo CO
VO CO N OCN vo 00 vo CO Os
d o o o o o o o o O o O o
Sj
PS
ps ps ps pj *^fc PS PS PS PS PS
© © © o o o o o o O o o
vo vo
ON OO
CO T oo t CO CO r- CO
VO o o 5 O
CO CN
j-
CN os oo r- vo
Ov vo CO CN
Os
CO
© O O o o o d O O o o ©
PS n Tf *p ws -£! PS PS PS PS 'p US
o o o o o o £ o o 1 o-^ o o O
Tj-
r- CO CO Os
vo Tj- CO i-^ < CO vo OS oo Tf
oo VO CN © co oo vo 00
(N O vo CO •—< O vo r-
© © o o O o o O o o O o
*p US irt r» PS PS «p «
o o o o o o
H o 1o—1 © o o
o1—
co CO
o1-H CO T}- CO Os OO CO V) CN r-*
r- O N vo
r- CN v> C
CN ON s <
(N C7O\ vo
CO
1-H f-
o o o O O o o O O O

3 |S 3 |S

J <N CO N- vo O (N CO ’t vo ©
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 561

By integrating over q we obtain the energy distribution of the con­


verted positrons. For 1 we obtain

d&L110 + ) = — ^ + r l c°2(/'-1) [£+ + £2-—2 ( ^ —1)W2] \n tn—^ +e- JrP+P-


nco^1 { T mco

+ W +T) - [ ^ E n y + m!] [0>++/>J * 11- "


CO2
—( p + — / 0 2(L-1)H- — [(£+£_+ p + p . - — m 2) ( p + + p _ ) 2{L~ 2)

- ( e +s _ - p +p _ - m 2) ( p +- p _ ) ^ L 2> ] - - [e2_ + £ !_ -2 (L -l)m 2]

x 2
71= 1
— L

U
" 11 [(p+ + p j ,n- ( p + - p j " ] } < f e +
'
a # o . (4 i.i3 )

For L = 1 (electric dipole) the differential conversion coefficient


is equal to

<W ’(£+) = - ^ { ( < $ + e i ) l n t £+ ir± :E±E^ + 2p+p_) de+. (41.14)


+ nor + mco ^ ^
We also give the expressions for the conversion coefficients for
electric quadrupole and octopole radiation which follow from (41.13):
a m 2+ E + E _ + p +p _
< W (*+) o)2(e\-\-e^_—2m2) In
nco5 mco

+ y p + p - ( £ + £ - — m 2) \ d £ + ;
(41.15)
m2+ £ +e _ + p H.p_
= A { co4(fi2+ £ i —4m2) In
mo' ( mco

—p +P - [2a»4—a)2m2—8m4—(9co2+4m 2)£+£_+4e2 c!e +

For the total internal conversion coefficient we give only the expres­
sions which are obtained in the limiting case co < m. In this case if
(co/m)2 > L we have
2a \ 2co 23 5L+1 Ly ! 4n+5 [
J n \ n m"“ T2 + 4 (L + 1)(2L+T) £ 2 (2 « + l)(2 « + 3 ) j '
(41.16)
562 QUANTUM ELECTRODYNAMICS

This formula holds for L ^ 2. For L = 1 we have

A1’- (4U7)
In the opposite limiting case L >> (oj/m)2 > 1 we obtain for
the same expression as for the conversion coefficient for a magnetic
multipole (41.10).
Table 13 gives the values of the coefficients for internal conversion
accompanied by pair production for the cases of electric and magnetic
multipoles obtained by numerical integration of expressions (41.7)
and (41.13).
41.3 Excitation of Nuclei by Electrons
The process which is the inverse of internal conversion of y rays
is the excitation of a nucleus by an electron. In this case the energy
of the final state of the electron is less than that of the initial state,
e2 < £ i, while the final energy of the nucleus is greater than the initial
energy E2 > El . By utilizing the results of § 40 we can write the fol­
lowing expression for the probability of excitation of the nucleus resulting
in its angular momentum being changed by L and its parity being
changed by (—1)-L+A+1, and with the electron making a transition from
the initial state (1) to the final state (2):
(41-18)
where is the probability of emission of radiation in the transition
of the nucleus from an excited (final) state into the initial state, while
is the “conversion coefficient” equal to
'2
W*B L U *W ldr (41.19)

Just as in the case of the investigation of conversion accompanied


by pair production we confine oursehes here to the free electron ap­
proximation (211), (194). We denote by p x and p 2 the values of the
electron momentum in the initial and final states, and by ux and u2
the amplitudes of the plane waves corresponding to them. In accordance
with (41.18) and (41.19) the differential cross section for the process
in which we are interested is equal to
aa> dp<i
dof) wL r a d
, U )
(41.20)
4^'i (2*Y '5(Si" £2
R E T A R D E D IN T E R A C T IO N B E T W E E N T W O C H A R G E S 563

where v 1 is the initial electron velocity


9 — Pi Pz

and Z t, denotes summation over the initial and final electron polari­
zations (the factor | in front of the sum corresponds to averaging over
the initial polarizations). On comparing formula (41.20) with formula
(41.2) for the coefficient of conversion accompanied by pair production
we can easily show that the last factor in these formulas, viz., the factor
Z
2 1| J ei{r'-P')ru * B ^ u xdr\

in (41.20) and the factor

£ | j eiip++p- )ru * B ^ vdr\

in (41.2) differ only by the following replacement of parameters: />_ -> p x,


—p + -*• p 2, —£4- -> e2, and e_ -> Therefore (41.20) can be written
in the following form:

da'» = | < ’, a«H£2- £ (41.21)


I V l

where b[X) is a function which can be determined from the expressions


for the differential conversion coefficient (41.3), (41.11) if we write
them in the form
= b(p_, p+)d(e++e_—«))dp+dp_.
On going over in (41.21) from the variables p 2 to the variables q
and on integrating over the angles we obtain

(41.22)
Z Pi

where
d^iL )ip+, q)
biL )(p+, 9) = de+dq
and d B ^ is defined by expressions (41.6) and (41.12).
In order to obtain the total cross section for nuclear excitation
a lLl) it is necessary to integrate (41.22) over q between the limits qmIn
and <7 max determined by the conservation laws. Since the total cross
564 QUANTUM ELECTRODYNAMICS

section cannot depend on the absolute direction of the initial electron


momentum, (41.22) can also be averaged over the directions of the
vector p + :
^max

^mln
This formula can be written in the form
(A)
rU) — VVLrad (41.23)
8np\
where

6» i (£) =
d
jS
1JI(£)
de
and dfi^ is determined by formulas (41.7) and (41.13) in which we must
replace p + and p_ by p x and p2.

41.4. Monoenevgetic Positrons


If there is an unoccupied state in one of the atomic shells, then
a pair can be created with the electron occupying this bound state,
and the positron having a sharply defined energy. We are here dealing
with the phenomenon of internal conversion involving the emission
of monoenergetic positrons. In order to evaluate the conversion coef­
ficient in this case we can make direct use of formula (40.13) where we
must interpret y2 as the electron wave function belonging to the discrete
spectrum of the shell, and rp1 as a negative frequency wave function
belonging to the continuous spectrum normalized per unit energy.
For the 7^-shell formulas (40.25) and (40.28) will hold in which the
continuous spectrum radial functions gx and f x must refer to negative
frequencies. In the approximation of small Z and of energies large
compared to the binding energy of the ^T-shell we obtain the expres­
sions for the conversion coefficients

2m [L+i
i9>0, = 2a(Za)3 T
CD
RETARDED INTERACTION BETWEEN TWO CHARGES 565

As an example we give the value of the ratio of the probability of


conversion accompanied by the emission of a monoenergetic positron
(with the electron being produced in the A'-shell) to the total probability
of conversion accompanied by pair production (182). Calculations
carried out with exact radial functions for = 1.4 MeV in the case
of an electric dipole give for this ratio the value 1/3. As a> decreases
this ratio increases.
The process which is inverse with respect to the production of
monoenergetic positrons is the following one: when a positron collides
with an electron belonging to the A'-shell of the atom this pair is absorbed
by the nucleus with no photons being emitted, and the nucleus becoming
excited. The probability of such a process can be written as
w = w yfi, (41.24)
where /3 is the coefficient of internal conversion accompanied by the
production of monoenergetic positrons, and w is the probability of
excitation of the nucleus by a photon of the corresponding energy.
Since the excitation energy acquired by the nucleus is sufficiently
great, then as a result of the absorption of the pair the nueleus ean
disintegrate. For example, the effective cross section for the fission
of uranium caused by such a process may be of the order of 10~31
cm2 (157).

§ 42. Coulomb (Monopole) Transitions

42.1. Reduction to the Static Interaction


The general formulas for the probability of conversion or excitation
obtained in § 40 become inapplicable if the angular momenta of the
initial and final states of the nucleus are both equal to zero. For L = 0
expression (40.11) vanishes in accordance with the fact that there exist
no photon states of angular momentum equal to zero. Nevertheless,
the matrix element of S(2) for a transition between two such states is
different from zero. Such transitions are called monopole or £0-transi-
tions.
As we have already pointed out in § 40, the integrals in (40.11)
extend over the region outside the nucleus, while the general expres­
sion for the matrix element also contains, in accordance with (40.32),
566 QUANTUM ELECTRODYNAMICS

an integral over the region occupied by the nucleus, which, generally


speaking, differs from zero for L = 0. Thus, the matrix element of the
effective interaction energy has the form for L = 0
e ia>\rj— r2\

U L/ = fj2 i ( r i ) J 2i(r2) --------- r dr1dr2,


Si
\ri - r 2\
where j 21 and J21 are the four-vectors of the electron and the nuclear
transition current density, and the integration is carried out over the
volume Q occupied by the nucleus. This matrix element can also be
written in the form (37.16)

^ i-/= - fA M ^ n ir J d r ! , (42.1)
a
where
gia>|ri—r2!
^21 (fl) dr2. (42.2)
4711f i —r2
The case L = 0 implies a spherically symmetric distribution of
nuclear transition current, whose spatial ( / 21) and time (72i) components
have the structure
fo) = / ( ''2> 2,
■^lW = s(r2),
where f(r) and g(r) are certain functions which are independent of
the angles. The potentials associated with these currents have an analo­
gous structure
^ 2i (r2) = a0(r2)r2 =
, (42.3)
^21 O'2) — tPoO2)'
On substituting (42.3) into (42.1) we obtain
+ J(./2i99+./2iV2) dr. (42.4)
Since in accordance with (42.3) the vector potential is the gradient
of a scalar function, it can be eliminated by means of a gauge trans­
formation.
We introduce the transformed potential <p0 = (p—iioX. On integrating
by parts the term containing V2 in (42.4) and on utilizing the equation
of continuity divy21 = —iw j^ , we obtain

u t-*f = fjn<P dr — e j y t W i dr- (42.5)


R E T A R D E D IN T E R A C T IO N B E T W E E N TW O C H A R G E S 567

In order to obtain the potential <p we utilize the fact that <p0 satisfies
the equation A(p0-\-(o2<p0 = —g(r), and therefore A<p = A{<pQ+ UoX)
— —g —a>29?0+ iioAX. Since A 21= —VA, it follows from the Lorentz
condition for the potentials that AX = — iio<p0 and therefore
A (p = —g = eW? ^ . (42.6)
Formulas (42.5) and (42.6) can also be written in the form

= -<■ f v fW f 1dr*. ■ <42-7)

We see that for L = 0 the transition is brought about by the electro­


static (Coulomb) interaction between the charges.
Formula (42.5) can be conveniently written in the form
Ut_>f= J J210 d r, (42.8)

where 0 is the electrostatic potential due to the electron density distri­


bution and satisfies the equation
A 0 = —ey*xpi- (42.9)
In this equation the right hand side may turn out not to have spherical
symmetry. However, in virtue of the spherical symmetry of the charge
distribution in (42.8) only the symmetric part of the potential will give
a nonvanishing result. Therefore, the right hand side of (42.9) can be
averaged over the angles.

42.2. Conversion and Nuclear Excitation in the Case of an EO-transition


We now obtain the probability of conversion and nuclear excitation
in the case of an £0-transition (68), (216), (194).
The electron wave function does not vary appreciably over the
volume occupied by the nucleus. Therefore, the right hand side of
(42.9) can be regarded constant, and we can start from the equation
1 d2(r0)
r dr2
where
2o = V^(0)y>i(0).
The general solution of this equation is given by 0 = c ^ c ^ r —egfr2!6).
The constant cl is unimportant, since the integral (42.8) vanishes for
568 QUANTUM ELECTRODYNAMICS

constant 0 in virtue of the orthogonality of the nuclear wave functions.


The constant c2 is equal to zero, this being a consequence of the condi­
tion that 0 is finite at r = 0. Therefore

v M = ~ e , Q „ (42.10)

where
Q. = — f Y iV i^ d r. (42.11)

The selection rules for EO-transitions follow from (42.11). They


consist of the statements that A J = 0, and that the parity remains the
same. Thus, EO-transitions can take place not only in the case when
the nuclear angular momenta in the initial and the final states are
both zero, but also in the case when Jr = J2^ 0 with the parity of the
two states being the same (40). In the latter case in addition to the
EO-transition M l-and E2-transitions are also possible, and they, in
contrast to the EO-transition, can take place not only by means of the
conversion process, but also by means of photon emission. Transitions
between states of J1 = J 2 = 0 , but of different parity, are possible
only by means of emission of two photons, or by means of still more
complicated processes.
The quantity Q0 is the “ zero moment” of the nucleus similar and
in order of magnitude equal to the quadrupole moment. Since for
L = 0 there is no photon emission, it is impossible to introduce a con­
version coefficient. In order of magnitude Q0 ~ E2e2, where R is the
nuclear radius.
The probability of conversion in the E-shell can be expressed as

w =2-2*te) | e . l ! l l M 0 ) N v ' ,( 0 ) l 2, (42.12)

where xpK{0) is the value of the E-electron wave function in the nucleus,
while y>e(0) is the value of the continuous spectrum electron wave func­
tion in the nucleus normalized per unit energy and evaluated for energy
e = c o + |e ff| and angular momentum j =
In the case of conversion accompanied by pair creation the dif­
ferential probability is equal to

«*»= ^ 12.1*Z le„l! «(o.-«_-*+).


Pl.p. ^ '
RETARDED INTERACTION BETWEEN TWO CHARGES 569

In the free electron approximation we have g0 = ufiv where u and v


are the amplitudes of the plane waves of positive and negative frequen­
cies; ^ denotes summation over the polarizations of the electron
and the positron.
On carrying out the summation we obtain
e+e _ + p +p _ —m2
4 ^— SpP (ip_— m )P(ip+ + m ) =
e_e+
Mi./'i
and, consequently,

dw = — ^ ^ \ Q 0\2\p-\\p+\de^do+do_(e^e+— m2-\-p-.p+). (42.13)

In a similar manner the differential cross section for nuclear excita­


tion is determined by the formula:

da = -jg- \Qo\2 j j ^ j d o 2(e1E2-\-m2-\-p1p 2). (42.14)

The total cross section for nuclear excitation is equal to


CHAPTER VII

Investigation of the Scattering Matrix

§ 43. Properties of Exact Solutions of the Equations of Quantum


Electrodynamics. Propagators

43.1. Stationary States of a System of Interacting Fields


In the two preceding chapters we have investigated various electro-
dynamic processes in the first nonvanishing approximation. For the
investigation of higher order approximations, it is necessary to use
subsequent terms in the expansion of the scattering matrix. In the lan­
guage of graphical representation of the scattering matrix this means
that for a given process we must consider more complicated diagrams
with the same external lines.
In investigating higher order approximations it is very useful to
make maximum use of those properties of the system of interacting
fields which can be formulated without reference to perturbation theory.
In this section we shall establish some of these properties.
The system of interacting fields (the electromagnetic and the electron-
positron fields) which we are studying can be characterized, like any
other quantum mechanical system, by a set of stationary states.
We denote the wave functions (the state vectors) of the stationary
states of this system by 0 Twhere r is the set of quantum numbers char­
acterizing this state. The fact that the state is stationary means, by
definition, that the wave function &r is an eigenfunction of the energy
operator. Since our system is a closed system, it follows from consid­
erations of invariance that 0 T is also an eigenfunction of the energy-
momentum operator

P f l r = PlP r - .( 4 3 ‘ 1 )

The four-vector p (p , ip0) is obviously contained in the set of quantum


numbers r.
[571]
572 QUANTUM ELECTRODYNAMICS

The lowest energy state of the system is called the vacuum state. We
denote it by 0 O and assume that in the vacuum state

P * = °-
This value is the only one which guarantees relativistic invariance of
the ground state of the system.
We assume that the electric charge of the vacuum is equal to zero.
For excited states the energy of the system must be positive, p 0 > 0.
In order for this property to be relativistically invariant obviously the
following relation must hold: p 2 < 0.
We consider a certain Heisenberg operator F(x) which is, in particu­
lar, a function of the coordinates and the time. It is defined by its matrix
elements
(r|F (* )|r') = (0?,F(x)<£°),
where &°T is a time-independent wave function in the Heisenberg picture.
We can easily obtain the dependence on the coordinates of the matrix
element (0|F(jc)|r) which connects the vacuum state with some station­
ary state r. In order to do this we utilize the result of subsection 22.3.
It follows from (22.18) that
( 0 |F W |r ) = /,e '" (43.2)
where f r is independent of *. Similarly
(r|F (x )|0 ) = / ; < r '" . (43.3)
If F(x) is a Hermitian operator, we have / ' = /* .
Theoperators of the electron-positron and theelectromagnetic
fields <]>(*) andA(xr) are of the greatest importance from our point
of view.
In accordance with (43.2) the matrix element of vj>(x) can be written as

(0 |'M * )lr ) = H r )u (p )e ipi, (43.4)


where f (r) is some invariant function of the quantum numbers r, while
u(p) is a normalized bispinor.
Since the operator 4>(x) (cf. § 18) has matrix elements corresponding
only to electron annihilation or to positron creation, and the charge
is an integral of the motion, then f (r) in formula (43.4) differs from zero
only if the state r has a single electron charge Q = e.
INVESTIGATION OF SCATTERING MATRIX 573

Moreover, for a given p the state r is fourfold degenerate corre­


sponding to the existence of four linearly independent bispinors. We can
choose them in the form of eigenfunctions of the matrices ip and Zp.
As was shown in § 10, the operator ip has the two eigenvalues ^ M ,
where M = ~\f—p2, corresponding to the different parities / =
For a given / the matrix Z f has the two eigenvalues 2\p\p (p =
corresponding to different polarizations.
Thus, for a state r for which (0 (jc) | r) differs from zero, the set
of quantum numbers includes p\ Q — e\ p = I = ± i . Moreover
the state may possess additional invariant quantum numbers. We
shall call these states single-charge states. It is known from experiment
that single-charge states cannot have an energy smaller than mc, where
mc is the electron mass. Therefore, we shall assume, although we cannot
give a rigorous proof of this assertion, that the four-momentum of
single-charge states satisfies the condition p 2 < — m\.
We identify the state p2 = —m2 with the state of a free electron.
Therefore we also assume that the parity in this state can have only
the one value I = i, and that the set of quantum numbers: p, Q = e,
I = i, p , = ± 1 is a complete set.
Single-charge states for which p2 < ~ m 2c can be interpreted as
combinations of one electron and a certain number of photons and
electron-positron pairs. Such a system must be characterized by addi­
tional quantum numbers and, in particular, may have arbitrary parity.
In analogy with (43.4) we can write the matrix element (r 14»(x) 10)
of the operator vpC*) *n the form
(r|4>(x)|0)= £'(r)u(p)e-iJ,x, (43.5)
where r stands for the same set of quantum numbers as in (43.4).
The matrix elements (0|4*(x)|r') and (r'|i|>(x)|0) have nonvanishing
values for Q = —e. They can be obtained from (43.4) and (43.5) with
the aid of the transformation of charge conjugation
(0 |$ W |r ')= C + (0 |+ (* )|r ).
(r '|+ W |0 ) = C ( r |$ M |0 ) ,

where r' differs from r by a change in the signs of the quantum num­
bers Q and p. Indeed, (43.6) follows from the relations (cf. subsection
23.3) 0?, = A0°r; &°0 = A0°o; A4»(x)A+= Cvp(x).
574 QUANTUM ELECTRODYNAMICS

We now proceed to investigate the matrix elements of the operator


of the electromagnetic field A(x). On the basis of (43.2) and (43.3) we
write them in the form
(01 ( x ) | r) = rj (r) aM(k) eikx, ^
(r IA/4(x) 10) = r j ' ^ a ^ k ) e~ikx.
Here a is a four-vector which we assume to be normalized just as in
the case of free fields (cf. subsection 16.3), i.e., = eJ(2k0Q)1/2, where
e is a unit vector, and Q is the normalization volume. The factor r](r)
is an invariant; it can depend on the invariants k 2 and k^a^, so
that r]{r) has different values in the case of transverse (A^a^ = 0) and
longitudinal polarizations. With respect to the transverse states the
operator A(x) is Hermit ian so that
V'(r) = ri*(r) (A ^#(= 0 ) . (43.8)
We note that for k 2^ 0 all the matrix elements of the operator A(x)
have ,the property of transversality. Indeed, from the equations of
motion □ A (1( x ) = —j/((x) and from relations (43.2), (43.3) it follows
that
*»(0|A(,W |r ) = ( 0 |j / . r ) |r ) ,
k*(r] A(,(x)|0) = (r|j„(x)|0).

Since the current operator j(x) satisfies the condition of continuity


[3L(x)/dxu\ = 0, we have A: (0| j.,(x)|r) = k (r\ j (x)|0) = 0, whence

^/£(°l k) = °>
(k2 ^ 0). (43.9)
M r IA„W |0) = 0
For k2 = 0 we identify^ the state r with a free photon. For k2 < 0
the state r can be interpreted as the combination of a certain number
of photons and electron-positron pairs. All the states r for which the
matrix element (43.7) is different from zero are uncharged ((9 = 0).
We now emphasize the principal points of difference between the
matrix elements of the operators and A which correspond to exact
solutions of the equations of quantum electrodynamics, and the matrix
elements for noninteracting fields discussed in Chapter III.
First of all, the exact solutions contain matrix elements of the
operators and A which connect the vacuum state with states for
INVESTIGATION OF SCATTERING MATRIX 575

which p2 < ~ m 2 and k2 < 0, while in the case of noninteracting


fields the matrix elements of and A connect the vacuum state only
with one-particle states, for which p2 = —m\ and k 2 = 0.
Secondly, for p2 = —m 2 and k 2 — 0 the matrix elements of v|> and
A differ from the matrix elements of ip and A by the factors f(r) and
rj(r). We write these matrix elements in the form
(0|<Mx)|r) = Z \ l2u(p)elpx, p 2 = - m2,
, , (43.10)
(0| A(x)|r) = Z ll2a(k)eikI, k 2 = 0,
where Z \ 12 and Z 1/2 stand for the values of £(r) and rj{r) for
p 2 = —m\ and k 2 = 0.

43.2. Propagators and Their Spectral Representation


In Chapter III we have defined for noninteracting fields the photon
propagator Dcfir and the electron propagator as the expectation values
of the chronological product of the field operators in the vacuum state.
These functions were the Green’s functions for the d’Alembert wave
equation and the Dirac equation.
We now consider similar vacuum expectation values for interacting
fields. They are also called Green's functions or propagators.
The photon propagator which we denote by G is defined as

GUix-y) = (o| 7’(A/i(x)Av(jO)|o). (43.11)


The electron propagator which we denote by Geap is defined in a similar
manner:
G*ap( x - y ) = (0|T(vi>a(x )^ (T ))|0 ). (43.12)
The functions Gv and Ge can obviously depend only on the difference
of coordinates x —y in view of homogeneity of space-time.
We shall show that the electron and the photon Green’s functions
Ge and Gv can be represented in the form of expansions in terms of
Green’s functions for free particles of different masses (99), (124), (76).
These expansions will enable us to investigate a number of general
properties of the propagators.
We begin with the photon Green’s function. Let x0 > >'0. Then
in accordance with (43.11) we have
G ^ f x - y ) = (0| A/jOA.OOIO), where *0 > y0.
576 QUANTUM ELECTRODYNAMICS

By utilizing the rule of matrix multiplication we write Gv in the form

G U X~ -y) = 2 (° IK t o | r ) (,r | A„(;/) 10),


r
where r is the complete set of quantum numbers characterizing the
stationary states. On substituting into this equation expressions (43.7)
for the matrix elements we obtain

Ov„ ,( .x - y ) = S v ( r ) v ' ( ' K W a , ( k ) ^ - > . (43.13)


T

We note that in accordance with (43.9) the longitudinal components


of the matrix elements of the operator A, i.e., those which do not
satisfy the relation a^k^ = 0, differ from zero only in the case k 2 = 0.
In Chapter III we have obtained an expression for the longitudinal
part D0^ of the photon propagator in the case of the free field (k2 = 0)

D ^{x-y)
x—y)
where % is an arbitrary function. The longitudinal part of the function
Gv will obviously have the same form. Therefore, expression (43.13)
can be rewritten in the form

= y »)i* y
k.s X X H° »
(43.13')
where the quantum number A defines the polarization state satisfying
the condition of transversality, while s is the set of the remaining quantum
numbers. Further, for the^sake of brevity we omit the arbitrary term
Fxltepdx,.
The four-fold summation over k in (43.13') can be carried out in
two stages: first the three-fold summation over k is carried out for
a fixed k0 = ]/k2-\-M2, and then summation over M is performed.
In this way we obtain
G v ( x - y ) = Z 2 1 a ^ { k ) a ^ { k ) e ik^
k. A(fc* = 0 )

+ E \ v ( ~ M \ s ) |2 £ a^{k )a[x\ k ) e lk^ \ (43.13'')


M.s fc.A(fc=-AT>)

(The index s is absent from the first sum, since single photon states
are characterized only by their momentum and polarization).
INVESTIGATION OF SCATTERING MATRIX 577

In a similar manner we obtain for x0 <r y 0 an expression which


differs from (43.13") by the replacement of k by—k.
On taking into account the fact that the summations over k and
over —k are identical, we can write for arbitrary times x0 and y0

G^v{x—y) = Z
u.x

+ E [?/(—M z, s) |2 £ ap( k)ai»(k)etkl^ y)- i / e ^ ' * - v']. (43.14)


Mt s k,X (k2= - / i 2)

We know the results of evaluating the sums over k and A for given
values of M which appear in expression (43.14). We have evaluated
them in § 17, and they are expressed in terms of the functions Dc
and A\M)
I a]?>(£)a?>(k)eikix^y)^ = A\M)fJV{ x - y ) ,
k.X

where A ciM) (x—y) is determined by the four-fold Fourier integral

and

= W + M 2Z~iO

(of. subsection 17.3). For M = 0 we have A cM(k) = Dc(k).


The final expression for the photon Green’s function can be conven­
iently represented in the form of the four-fold Fourier integral

& „ (* -? ) = (* „ - - f i ) d' k -

On introducing the notation


M+dM
2 2 \ V ( - M \ s ) \ * = ZQ{M*)dM\
M s

we can rewrite Gv(k) in the form


OO

G*(k) = z { D ‘(k')+ f e (M 2)AcM(k)dM2}. (43.15)


578 QUANTUM ELECTRODYNAMICS

or
00
1 r g(M2)d M 2 )
(43.15')
<p w = k2- i 0 +J
o
k 2+ M 2- i Q \
>
Thus, we have obtained a representation of the photon Green’s
function in the form of an expansion in terms of the Green’s functions
for free particles of different masses. The quantity q{M2) is essentially
positive and can be interpreted as the density of states of a given mass.
We see from (43.15) that Gy(k) has a pole at k 2 = 0, i.e., at a point
corresponding to a real photon.
The representation (43.15) implies that the function Gy(k) has
definite analytical properties. In order to establish these properties
we utilize the fact that the function A cM(k) can, in accordance with (17.41)
and (17.19'), be separated into a real and an imaginary part
1
d cM(k) = 7i S (k2+ M 2) —iP
k2+ M 2 '
On substituting this expression into (43.15) we obtain a simple relation
between the real part of Gv(k) and the function o
Re Gy(k) = 7iZ{d(k2) + Q ( - k 2) e { - k 2)}, (43.16)
where
1, x > 0,
6(x) =
0, x < 0.
From this it follows, in particular, that
= 0, k2 > 0,
ReGy(A:) (43.16')
>0, k2 < 0 .
Further, we obtain

1 C Q(M2) d M 2
Im Gy(k) = - Z P (43.16")
+ J k2+ M 2
Equations (43.16) and (43.16") imply the following relation between
the real and the imaginary parts of the photon propagator:

ReGy(k')dk'2
\m G y{k) = — (43.17)
71 I k ’2—k 2
—oo
INVESTIGATION OF SCATTERING MATRIX 579

From (43.15') we can also draw some conclusions with respect to


the asymptotic behavior of Gv(k) for large A:2. Since q(M 2) > 0, and
if the integral f q(M 2) cJM2— J converges, then Gy( k ) ~ ( J / k 2), where
k 2 -» oo. However, if the integral J does not exist, then Gv(k) falls off
as A:2 -» oo slower than 1/k2.
We now proceed to obtain the expansion of the electron Green’s
function Ge{p) in terms of the free particle Green’s functions.
In accordance with (43.12) we have for x 0 > y 0
Ge*P(x—y) = (01 (x)4>0 (y) | 0).

On utilizing (43.4) and (43.5) we obtain


GeaP( x - y ) = £ £(r)£'(r)ua(p)up(p)eipix- y), x0 > y0
T

or
Geap ( x - y ) = X s , T) r ( p 2, s ,I)2
p .S .I ft
(43.18)
where p, is the quantum number specifying the polarization, / is the
parity, and s is the set of the remaining quantum numbers.
Just as in the investigation of the photon Green’s function, we re­
place the four-fold summation over p by the three-fold summation
over p for a given p 0 = \ / p 2-{-M2, and by a summation over M,
i.e., we write (43.18) in the form
GeaP{x—y)
M.s.I ft.p

For x0 < y 0 we obtain, on utilizing (43.6), an expression which


differs from (43.18) by the replacement of x —y by y —x. Therefore,
in the general case of arbitrary x0 and v0 the function Ge has the form

p.v

M.s. 1

x (43.18')
p-t1
Here the first sum refers to the state of a real free electron, so that the
index s is absent from it, and / = + /.
580 QUANTUM ELECTRODYNAMICS

The sums over p and p appearing in (43.18') were evaluated in


§ 19. Indeed, for 7 = z we have

2«<r'(pysjrxp)
p./<
where

%,(*->)=(2^5J S f ^ p y ^ — '^p
and
1
SfM) (p)
p -iM - 0
The states with 7 = —z differ by the fact that for them the amplitudes
w satisfy the equation (ip—M ) u = 0, Therefore, the values of the
sum for 7 = —z differ from the values for 7 = z by the replacement of
SB
lU) by where
1
Sf-M)(P) =
p + i M + O'
Thus, on introducing the notation
M + dM
2 2 1 s, ± 0 £ '( —M 2, 5, ± z) = Z l(r(± M)dM,
M s

we can write the electron Green’s function in the form

Ge{ x - y ) = Z ,{ ^ mc)(.r-y )
CO OO

+ j o ( M ) S f M)( x - y ) d M + f a ( - M ) S ^ M) { x - y ) d M \ . (43.19)
"<c mc

The same expansion can obviously also be written for the Green’s
function in the momentum representation, i.e., for the Fourier coeffi­
cients in the expansion

Ge{x ~y) = J Ge(p)elplx- V)dip,

viz.:
z J ___1_____ C o(M)dM P o ( —M~)dM
Ge(p) (43.20)
l \ p — i m c—0 ^ p — i M — 0 J p + z'M +0
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 581

We see from (43.20) that Ge(p) has a pole at p 2-\-in2c = 0 corresponding


to the state of a real electron.
Expansions of the type (43.15) and (43.20) can also be obtained
for the vacuum expectation values of the commutators (0| [A^(jc) ,
A„(y)]j0) and the anticommutators (0|{i]>a(x:), «4>/7(jj/)}|0). By proceeding
in the same fashion as in the derivation of formulas (43.15) and (43.20)
we obtain
CO

(OKA/*), A1,(y)]|0)= Z {/>„„(*—;y) + \ q(M2)A Wllv( x - y ) d M 2) ,


0
co

(0|{^o(*)> ^,iO;)}l0) = z i {sa/,( * - jO + J [a(M)SiM)aft( x - y )


mc

+ a ( - M ) S l_„)ap(x-y)]dM }, (43.21)
where the functions A {M) and.S(M) are defined by formulas (17.37) and
(19.5). In particular, for x0 = y0 the following commutation relations
follow from the above expressions:
(0| [A„ (*, t), A r(y, /)]|0) = 0,
oo
dAv(y, t)
A /* , t), o )= - 1 5 p i v 5 ( x - y ) z ( \ + § e W d M 2},
( 1 dt
' 0
OO

(oK^C*. 0 . typiy, 0}|oJ=y4^ ^ ( JC- 3 ’) z i{ 1+ /[ff(A /)+<r(—M)] dM).


n,c

On comparing these formulas with the commutation relations (22.5)


we obtain expressions for the constants Z and Zx
CO

Z ~ 1 = 1+ \(){M2)d M 2, (43.22)
0
oo

Z f ! = 1+ f [a(M) + a ( - M ) ] d M . (43.22')
mc
Since q(M 2) is an essentially positive quantity, it follows from this
that 0 < Z < 1.
Equation (43.22) can be interpreted as the condition that the system
of wave functions for a given value of the momentum k be complete.
Indeed, in the case of a noninteracting electromagnetic field there
582 QUANTUM ELECTRODYNAMICS

exists only one value of the energy k 0 = \k\, and therefore Z = 1.


The interaction adds states of energies k 0— \ /k 2f - M 2, with their density
being given by Z q(M2). Therefore the condition for the completeness
of the system assumes the form (43.22).
Let us also consider the problem of the transformation of the electron
Green’s function under the gauge transformation A -> A^f-dyjdx^
(116), where % is an arbitrary scalar function.
The operator of the electron-positron field ip transforms in this
case in accordance with ip -> ipeix. Therefore, the gauge transformation
for the Green’s function has the form

G{e)(x, x') = G{f { x - x ' ) ( o i r O ^ ^ - e - ^ 'O l O )

where G{0e)(x —x') is the Green’s function corresponding to the transverse


gauge for the electromagnetic potentials, i.e., dl = 0 (cf. (17.21)).
On expanding %{x) in plane waves and on utilizing relations (17.12')
we can easily show that the last factor in the expression for G{e)(x, x') is
equal to exp(/e2[2?(0) —B ( x —x')]) where

B ( x —x') = dlk .

Therefore, the gauge transformation for the electron Green’s function


can finally be written in the form
Gief x - x ' ) = G f> (x-x ') exp(ie2 [ B (0 )-B (x -x')]) (43.23)

43.3. Connection between Propagators and the Scattering Matrix.


Integral Equations for Propagators
The foregoing matrix elements of the field operators A, 4*,4* and
their products were defined in the Heisenberg picture. They can be
determined most simply in a unique and relativistically invariant
manner in this particular picture, since only the operators depend on
the time, while the wave functions (the state vectors) are constant.
We now express these matrix elements in the interaction picture.
In order to do this we use the transformation formulas

0(t) = S (t, —oo)^°,


F (x )= S~1(x0,-oo) F(x)S(x0, —oo),
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 583

or
F(x) = S~17’(F(x)S), (43.24)
where S = 5(oo, —oo) is the scattering matrix and T is the chronological
operator. From this it follows that the matrix element (r]F(x)|r') has
the following form in the interaction picture:

(/■jF(x)|r') = (#°r, S-1r(F(x)S)0») (43.25)


where 0°r denotes the constant wave function for the system of fields
in the stationary state r. Since it is assumed that at time t = —oo there
is no interaction between the fields, the state vector &°r can be regarded
as coinciding with 0 T{—oo), where 0 T(t) is the wave function for the
system of fields in the interaction picture describing the state r.
In the case of stationary states 0 r(oo) may differ from 0 T(—oo) = 0°r
only by an inessential phase factor, i.e.,
0 r(oo) = S 0 T( - o o ) = Xj. 0 r ( oo) (43.26)
where the absolute value of the quantity ),r = (0°T, S0°r) = (r|S |r) is
equal to unity, \XT\-= 1.
Thus, we have

('•IFMIr') = (r!T (F W S) Ir') (43.27)

In the derivation of this formula we have made no distinction between


the instant at which the interaction between the fields is switched on,
and the instant when the wave function of the system in the Heisenberg
picture coincides with the wave function in the interaction picture.
However, a more consistent approach is a somewhat different one,
based on the adiabatic hypothesis, according to which at t = —oo
the interaction between the fields is switched on infinitely slowly, as
a result of which the wave functions in the two pictures — the Heisenberg
and the interaction picture—coincide not at t — —oo, but at some
finite value of the time, when the interaction has been completely
switched on. However, such an approach does not change the result
(43.27). We must only keep in mind that for the evaluation of the matrix
element (r| F(x)| r') in the Heisenberg picture we must utilize the eigen­
functions of the total Hamiltonian for the system of fields H, which
describe the states for r and r', while for the evaluation of the matrix
584 QUANTUM ELECTRODYNAMICS

element (rj r(F (x)S )|r'j in the interaction picture we must utilize the
eigenfunctions of the unperturbed Hamiltonian H0 which also refer to
the states r and r'.
Similarly, it can be easily shown that for a product of two operators
F^x) and F2(t) the matrix element connecting the two stationary
states r and r' has the form

( r |r ( F 1(x)F2(y))|r/) = (r | r(F (x) F(y) S )|rj (r| S| r)_1. (43.28)

In particular, the propagators defined earlier can be expressed in the


interaction picture in the following manner:

o i n U - y ) = ( o ir ( v .W V f O ) s ) |o ) , (43.29)

< % •(* -J>) = io \T {A /i( x ) A ll(y) S)|o). (43.30)

(Here, and in future, we omit the factor (01S | O)^1 where the index
0 denotes the vacuum state for the free fields.)
We can use these formulas to establish the relation between the
propagators and the scattering matrix.
In Chapter IV we have obtained a general expression for the scattering
matrix in the form of a chronological product of field operators. But
every specific electrodynamic problem is associated with expanding
the scattering matrix in terms of normal products. The only way of
doing this is to take the series expansion of the scattering matrix
and to expand its terms in normal products one at a time. However,
in principle we can consider an exact expansion of the scattering matrix
in terms of normal products. Such an expansion must be of the form

s = E ..., xn; y lt ^ ., ..., £„)


71. V

XN(Tp(Xl), . . . , :<f>(xn);yj(y1),...,y)(yn); A ( ^ ) , ..., A(£„)). (43.31)

For the sake of brevity we have omitted here the spinor indices
on -ip and y> and the vector indices on A; the quantity K (n-v) is a spin
tensor of the 2«th rank and a tensor of the vth rank, i.e., a quantity
having 2n spinor indices and v vector indices. (The number of times
that y) appears as a factor in each term is, in accordance with expression
(22.4) for the current, obviously equal to the number of times that yj
appears as a factor.)
INVESTIGATION OF SCATTERING MATRIX 585

The expansion coefficient K (n-v) obviously defines the exact expression


for the matrix element of a process in which the total number of electrons
and positrons in the initial and final states is equal to 2n, while the
number of photons is equal to v.
As an example we consider the Compton effect. By utilizing the
expressions for the matrix elements of the operators ip, yj and A we
obtain, in accordance with (43.31), the following expressions for the
matrix element:
s t^r = aa(ki)ap(k2)ti(p2){K$-2)( - p 2, Pl, k x, - k 2)
+Ka},2)( ~ P 2,Pi, ~ k 2 ,k 1)}u(p1), (43.32)
where K {1-2)(qx, q2, f 1, / 2) is the corresponding Fourier component
of K {1’V

K (1-2)(<h , ? 2 , / l , / 2)
= f K (1-2)(x,y, £2) *+»,»+/,{,+/,«,)d ^ x d ^ d ^ d * ^ ,
while u(p) and a{k) are the amplitudes of the electron and photon
states.
The matrix element for any process can be expressed in a similar
manner. Let there be n1 electrons and n2 positrons in the initial state,
and n—n2 electrons and n —nx positrons in the final state. Further,
we denote the total number of photons in the initial and the final states
by v. Then we have

S,-,r=
x W ( y x, ...,>-n) 0 ( f 1, ..., f,)*/4*! ...d 4f „
where ^ is the antisymmetrized wave function (in configuration space)
for the electrons in the initial state and for the positrons in the final
state, W is the antisymmetrized wave function for the electrons in the
final state and the positrons in the initial state, and 0 is the symmetrized
wave function for the photons in both the initial and final states

■■■,yn) = u (Pi) ••• ■■■ v ( - p n) £ ( ± ) e ’ ' 1,

W'{xx, ..., x n) = v ( - p [ ) ... v ( - p ' J u ( p nz+1) ... u(p'n) (± )e ^

0 = £ a(ki) ••• a(kv) e Ik}tii.


586 QUANTUM ELECTRODYNAMICS

The quantities K (m) and A7<0,2) which are called the electron and
the photon self-energy functions will be of particular importance for
us. We denote them in the following manner:
K$«Kx, y) = I ap(x, y) ; *£■«>(&, « = 7 7 ^ , f 2). (43.33)

The quantity 27 depends on two spinor indices a, /3 and on two four-


dimensional coordinates, and in the absence of external fields it obvi­
ously depends only on their difference: Z ( x , y ) = Z ( x —y). Its Fourier
components, therefore, depend on only one variable (momentum).
We denote them by 27(p).
The quantity 77 depends on two vector indices p, v, and also on the
difference of coordinates77 (x,_y) = I I ( x —y). Its Fourier components
are functions of one momentum. We denote these components by
n^{k).
The function 27 is simply related to the electron propagators Ge.
In order to establish this important relation we substitute into the
expression (43.29) for the function Ge the expansion for the scattering
matrix (43.31):
Gea&( x - y )

= (°| r {wa(*)v>p00■N [i - / Zys (*' —y')v>s (/) % (■«')<Px' d y h— ]} |o).


In accordance with the rules of § 24 the mixed T-products appearing
in the right hand side of this equation can be expanded in terms of
iV-products. Since the vacuum expectation value of any iV-product
is equal to zero, all the terms will vanish, except the first two for which
explicit expressions have been given earlier.
Further, by utilizing the equations

(oj T(ipa( x ) (y))|o) = S ‘p(x—y),


(0| T{y>a(x)y>p(y)N[ips(y')yjY(x')]} |0 ) = —S cay( x - x ' ) S cdp( y ' - y ) ,
we obtain
Ge( x - y ) = S c( x - y ) f f S c( x - x ' ) 2 ( x ' - y ,) S c( y ' - y ) d ix ' d y . (43.34)

In the momentum representation expression (43.34) assumes the


form
G°(p) = S ' ( p ) + S ‘(p)Z(p)S'(p). (43.35)
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 587

In a similar manner we can obtain the relation between Gy and 77.


On substituting (43.31) into (43.30), we obtain

G i M - y ) = D I M - y ) + f D U ( x - x ' ) n Ay( * ' - / ) D ' X y ' - y ^ x'd*/.


(43.36)
In the momentum representation relation (43.36) assumes the form

GjnXk) = D ^ ,k + D U ( k ) n Xa(k)D^(k). (43.36')


We have shown earlier (cf. (43.14)) that we can confine our discussion
only to the transverse part of the function Gy(k), which satisfies the
relation Gyvk v = 0, since the longitudinal part is an arbitrary function
which coincides with the corresponding function in the absence of
interaction. Therefore, the photon self-energy function must also satisfy
the condition of transversality, or, in other words, the quantities D°lp(k),
k k
Gyv(k) and 77 v(k) are proportional to the tensor J ^ =
K

G l,( k ) =

n r,(k) = n ( k ) ( d ^ kX ’)-

Since Ja?J?y = Jay equations (43.36) and (43.36') can be rewritten


in the form of relations between scalar functions
Gy{k) = Dc( k ) D c(k) IT(k)Dc(k), (43.37)
or in the coordinate representation

Gy(x—y) = Dc( x —y) + J Dc(x—x ’) n ( x ' —y')Dc(y'—y)d4x'd4y'.


(43.37')
We now define the quantities M (p) and P(k) related to E(p) and
ri(k) in the following manner:
1+ S c(p)Z(p) = (l + iS ° (p)M (p))-\
l + Dc(k)J7(k) = (1 + iDc(k)P(k))~1
588 QUANTUM ELECTRODYNAMICS

or
Z(p) = - i M { p )( \ + i& { p )M { p ))-\
(43.38)
IJ(k) = - i P ( k ) (\ + iDc{k)P(k)]-\
Then relations (43.35) and (43.37) assume the form
Ge(p) = S c( p ) - i S c(p)M(p)Ge(p),
(43.39)
Gy(k) = Dc(k)—iDc(k)P(k)Gv(k),
or

(43.40)
\-iP{k).
Gy(k) Dc(k)

Equations (43.39) for the Green’s functions were obtained by Dyson


(53). In the coordinate representation equations (43.39) assume the
form of integral equations
Ge{x—y) — S c{x—y) — if s c(x—x ') M ( x '—y')Ge(y'—y)dix 'd iy',
(43.41)
Gv( x —y) = Dc(x—y) —i j Dc( x —x')P(x' —y')Gy{y' —y)dix 'd iy'.

These integral equations can be rewritten in integro-differential


form, which enables us to establish their connection with the Dirac
and the d ’Alembert equation. In order to do this we multiply the first
of equations (43.39) by —1/ S c(p) — p —im. We then obtain
(p —im—iM(p))Ge(
—im— p ) = —1 (43.42')
In order to obtain the corresponding equation in the coordinate repre­
sentation we apply to the first of equations (43.41) the differential
operator p — i m = —iy d/dx —im. Since (cf. § 18) (p —im)Sc(x—y)
= —<5(x—y), then
( p - i m ) G e( x - y ) - i f M ( x - y ' ) G e( y ' - y ) d iy ' = - d ( x - y ) . (43.42")
If we denote by M the integral operator whose kernel is the function
M (x—y), i. e., MGe( x —y) = J M ( x - y')Ge( y'— y) d^y' then equation
(43.42") can be rewritten in the following form
(p—im—iM )Ge(x—y ) — —d(x—y). (43.42)
INVESTIGATION OF SCATTERING MATRIX 589

Equation (43.42) shows that the electron propagator is the Green’s


function for an equation similar to the Dirac equation (p—im —iM)y) = 0
but differing from the latter by the addition to the mass m of the oper­
ator M. For this reason the operator M is called the mass operator.
We transform the equation for the photon propagator in an
analogous manner. On multiplying the second of equations (43.39) by
1/Dc(k) = ik2 we obtain
i(k2+P(k))Gv (k)= 1. (43.43')
On applying to the second of equations (43.41) the operator i □
= io2/dxfl and on utilizing the equation (cf. §17) — i O Dc(x—y) = <5(x~y)
we obtain
i ( —n + P ) ^ * —y ) — <5(x—y ) , (43.43)

where P is an integral operator whose kernel is the function P ( x —y),i.e.,


PGy(x—y) = J P ( x —y')Gy(y'—y)d4y'.

We see that iGy(x—y) is the Green’s function for the equation


(— D+P)v4 = 0 which in macroscopic electrodynamics describes the
propagation of electromagnetic waves in a medium whose polar­
ization properties are described by the operator P. Therefore the operator
P is called the vacuum polarization operator or the polarization operator.
We can say that the mass operator M describes the interaction of
the electron with its own electromagnetic field. This interaction consists
of emission and absorption of virtual photons. In a similar way we can
say that the polarization operator P describes the interaction of a photon
with the electron-positron field. This interaction consists of creation
and annihilation of virtual pairs.
In conclusion, let us also derive the equation satisfied by the matrix
elements of the Heisenberg field operators corresponding to a transi­
tion from vacuum to a single electron or a single photon state. We begin
with matrix element (0 |iJ>(jc)| r) = (0\T(y)(x)S)\ r). On substituting for
S the expansion (43.31), and on carrying out the same operations which
lead to the equation (43.34) for the function Ge we obtain
(0|<]>(*)k) = (01y(*)|r)+ J S c( x —x')Z(x' —y')(0\tp(y')\ r) d4x 'd 4y '.
Going over to the momentum representation and utilizing the
notations
(0 |4 * (x )|r)= u (p) elpx, (0|y>(x)|r)= u (/?) eipx.
590 QUANTUM ELECTRODYNAMICS

we obtain
u(p) = u ( p ) + S c(p)Z(p)u(p). (43.44)

We obtain an analogous relation also for the matrix elements

(,r 1 ( x ) 10) = ue~ipz and (01A(x) | r) = a(k) eikI


u(p) = u(p)-ru(p)i:(p)Sc(p), (43.44')
a(k) = a(k) -f Dc(k) n (k)a(k).
On introducing the mass and the polarization operators M and
P we obviously get the equations for u(p) and a{k)
u {p) = u(p)—iSc(p)M(p)u(p),
(43.45')
a(k) = a(k)—iDc(k) P(k) a(k).

In the coordinate representation these equations are integral equa­


tions. Since (/p+m )^ = 0 and GA = 0 they are equivalent to the
integro-differential equations
(/p + m + M )(0|4»(x)|r)= 0,
(—n + P ) ( 0 | A (*)!#-) = 0.

We see that these equations differ from the equations for the correspond­
ing propagators by the fact that they are homogeneous, while the equa­
tions for the propagators contain inhomogeneities in the form of
^-functions. In other words, the propagators are the Green’s functions
for the equations for the matrix elements.

43.4. Electromagnetic Mass of the Electron


We have previously (cf. subsection 43.1) defined the electron mass
mc as the energy of the lowest state of the system of interacting fields
having a unit charge and zero momentum. Obviously, the quantity mc
need not coincide with the “ mechanical” mass m of an electron which
is not interacting with the electromagnetic field. Indeed, the mechanical
electron mass m plays the role only of a constant which appears in the
equations of quantum electrodynamics and which does not directly
determine the energy levels of the system of interacting fields. The
difference between mc and m may be called the electromagnetic electron
mass.
INVESTIGATION OF SCATTERING MATRIX 591

We now establish the relation between mc and m. In order to do this


we make use of the spectral representation of the electron propagator
(43.20). For values of p2 close to —m2 we can retain in (43.20) only the
first term, i.e.,

— ~G^(pj= P2+ "*2-> °- (43.46)

We compare this expression with (43.40)

— = p - i m - i M (p). (43.47)

The mass operator in the momentum representation M(p) is a matrix


which depends on the four-vector p. Such a matrix can be written as
M{p) = c1Jr c2p, where c1 and c2 are scalar functions depending only
on p 2.
The preceding expression can be written in the convenient form
—iM(p) = a(p2) ( p —imc) —ib(p2). (43.48)

When p 2+ m 2- ^ 0 we can replace the functions aip2) and b(p2) by


their values for p 2 = —m2c. On introducing the notation a(—m2) = a0;
b(—m2) = 8m, we obtain from (43.46), (43.48)
Z y y{p—im^) = p —im—i d m + a ^ p —im^
or
mc = m-f (5m,
(43.49)
Z r 2 = flo+1.

These formulas express the electromagnetic electron mass mc—m,


and also the constant Z x in terms of the mass operator M(p).
We now investigate the expansion of the photon propagator in the
neighborhood of the point k 2 = 0. On retaining in (43.15') only the
first term we obtain
= i Z -'k 2. (43.50)
GHk)
On the other hand, it follows from (43.40) that

= i(k2+P{k)). (43.51)
Gy(k)
592 QUANTUM ELECTRODYNAMICS

In the neighborhood of the point k2 = 0 the polarization operator


can be written in the following form :P(k)= P(Q>)-{-Bk2. On substituting
this expansion into (43.51) and on comparing with (43.50) we obtain
dP
Z "1 1 + 5 = 1+
~dk2 k2=0 (43.52)
P(0) = 0.
If the second of relations (43.52) were not satisfied, then the expansion
(43.50) would also not hold, and the photon Green’s functions would
have a pole not at k 2 = 0, but at k 2-\-P(0) = 0, i.e., the photon would
have a rest mass. The equation P(0) = 0 is associated with the gauge
invariance of the theory, whose significance lies in the fact that all
physically observable quantities can depend only on the electromagnetic
field tensor +,,„(&) = k Av(k)—k vA M(k), and not on the potentials
themselves. Therefore, physical quantities must vanish when
k h = 0. Since P(k) depends only on k2, we have P(0) = 0.
We can also associate this property of P(k) directly with the trans-
versality of the self-energy function IT/jv(k).
On assuming that the transverse tensor P^v = P{k){dflv— {k ^k Jk2))
has no singularities at k = 0, we can write it in the form
^ ( / c) = P 1(/cz)(/ cz «5m- ^ ^ ) ,
where the function Px{k2) is finite at k2 = 0. From this it follows that
P(k2) = k 2Px(k2) -* 0 for k 2 = 0.
In conclusion we return to'the investigation of the matrix elements
(0|4K*)|r) ar|d (0|A (x)|r). We have seen that they satisfy equations
(43.45). On utilizing the properties of operators M and P for k 2 = 0
and p 2+ m z = 0, we obtain
(/p+m c)(0|4*(x)|r) = °,
□ (0 1A(x)| t) = 0,
i.e., the equations for noninteracting fields. Since these equations are
homogeneous we cannot, of course, obtain from them relations between
the amplitudes u, u, a and the corresponding amplitudes for free fields
u, u, a. However, we have already obtained such relations earlier (cf.
(43.10)). These relations contain the universal constants Z x and Z
whose meaning and significance will be elucidated later.
INVESTIGATION OF SCATTERING MATRIX 593

§ 44. Structure of the Scattering Matrix

44.1. Self-Energy Parts of Diagrams


The higher order diagrams for the scattering matrix have a complex
structure. Therefore, it is very useful to break up these diagrams into
parts, and to investigate successively those more or less complex parts
which may appear in diagrams for different processes. A study of the

■* «*■

Fig. 55.

( 1)

(2 )

(3)

properties of such parts and the evaluation of quantities corresponding


to them enable us to obtain solutions for a large number of electrody­
namic problems in higher order approximations.
In doing this we must keep in mind that in higher order diagrams
we may completely leave out of consideration those parts which contain
no external lines, and which are not connected by electron or photon
594 QUANTUM ELECTRODYNAMICS

lines with the principal part of the diagram. Indeed the totality of all
such parts, which are called vacuum parts, will contribute to the matrix
element the factor (0 |S |0 )= ( , S&o), of absolute value equal to unity
by virtue of (43.26).
Of the greatest importance are those parts of the diagram which
are called the electron self-energy, the photon self-energy, and the
vertex parts. We now give their definitions.

Fig. 57.

~ ~ o - -

We call a part of the diagram which is connected to other parts by


two electron lines only an electron self-energy part.
An electron self-energy part is shown schematically in the form of
a square in Fig. 55. Fig. 56 shows electron self-energy parts of the
second and fourth orders.
We call a part of the diagram which is connected to other parts by
two photon lines only a photon self-energy part.
A photon self-energy part is schematically shown in the form of
a square in Fig. 57. Fig. 58 shows photon self-energy parts of the second
and fourth orders.
INVESTIGATION OF SCATTERING MATRIX 595

We denote by E w the quantity in the matrix element which cor­


responds to the electron self-energy part W. This quantity is a second
rank spinor (a four-rowed matrix in the spin variables) which depends
on the four-momentum p of the incoming or the outgoing electron
line (the momenta corresponding to the incoming and the outgoing
lines are equal by virtue of the law of conservation of momentum)
Z w~ Z w(p).
It can be easily shown that the totality of quantites corresponding
to all the self-energy parts W coincides with the self-energy function
which we have defined in the preceding section Z { p ) = ^ 2 7 w(p).
Indeed, the term containing 27 in the expansion of the scattering
matrix (43.31) corresponds exactly to the totality of all diagrams
terminated by two electron lines.
We consider a diagram containing an internal electron line joining
two points A and B. In higher order approximations diagrams will
appear in which there are inserted between A and B various self-energy
parts W. We can describe the whole set of sueh parts joining A and B by
means of a single effective electron line. We shall represent it by a thick
solid line. By definition such an effective line is determined by the

B A

Fig. 59.

graphical equation shown in Fig. 59 where the square represents the


self-energy function 27. From this equation we see that the effective electron
line represents the electron propagator Ge. Indeed, Fig. 59 implies the
relation Ge(p) = S c( p ) + S c(p)Z(p )S c(p), which coincides with (43.35).
The same result can also be obtained with the aid of the definition of
the propagator Ge( x - y ) = (o| T(ip(x) vO)s)|o), since if we expand
the scattering matrix S into a perturbation theory series we will obtain
from this definition a set of expressions corresponding to all possible
diagrams which begin with an electron line at the point A and terminate
with an electron line at the point B.
At this point, let us consider the external electron line of a diagram
which joins the rest of the diagram at the point A (Fig. 60). In higher
596 QUANTUM ELECTRODYNAMICS

order approximations diagrams will appear in which between the


point A and the external line there will be inserted various self-energy
parts. The set of all such diagram parts can be described by an
effective external electron line defined by the graphical equation shown
in Fig. 61. It can easily be shown that the external electron line represents

A A
■* •*

Fig. 61.

the matrix element of the Heisenberg operator ^ or rp. Indeed, Fig. 61


implies the relation u{p) = u{p)-\-S0{p)Z{p)u(p), which coincides
with (43.44). We can also obtain the same result by considering the
graphical representation of the matrix element (0|<|»(;c) | r) = (0|7y)S|r).
This result is very important since it eliminates the necessity of
considering those diagram parts which can be replaced by effective
external electron lines. The matrix elements of the operators <|> and rp
which correspond to effective external lines are defined by formulas
(43.10).
If we cannot break up an electron self-energy part into smaller parts
interconnected by only a single electron line we call it a compact electron
self-energy part. Examples of such parts are shown in Fig. 56 (1, 3, 4, 5).
The electron self-energy part shown in Fig. 56 (2) is not compact, since
it can be broken up into two compact parts.
We denote by Z ^ the quantity which corresponds to a compact
electron self-energy part W. We also define the entire compact self-
energy part as the sum of all the compact parts Z* = £ Z*.
w
The function Z*(p) is simply related to the mass operator M(p)
which we have defined earlier. In order to establish this relation we
construct the self-energy function Z from the functions Z* correspond
INVESTIGATION OF SCATTERING MATRIX 597

ing to the compact parts. This can be done in accordance with the
schematic diagram of Fig. 62, where the square corresponds to 27 and
a rectangle to 27*. This schematic diagram corresponds to the infinite
series 27 = 27*+27*Sc27*-|-27*S'c27*S'c27*-|- ..., on summing which we
obtain 27 = 27*(1—5'c27*)“1. On comparing this last equation with
(43.38) we see that 27* (p) = —iM(p).
The definitions given above and the resultant relations may be
transferred completely to the case of the photon self-energy parts and
photon lines.

F ig . 62.

We denote by 1JW the quantity which corresponds to a definite


diagram of a photon self-energy part W. The quantity I I W is a four­
dimensional second-rank tensor, and is a function of the momentum
of the incoming and outgoing photon lines. We call a photon self­
energy part compact if it cannot be broken up into parts joined by only
a single photon line. Figure 58 shows photon self-energy parts of the
second and fourth orders. The self-energy parts of Fig. 58 (1, 3, 4)
are compact, while the self-energy part of Fig. 58 (2) consists of two
compact parts. We denote by II* the quantity which corresponds to
a compact photon self-energy part W .
Just as in the case of electron self-energy parts, the totality of all
photon self-energy parts corresponds to the function Il(k) = ^ l l w(k),
W
which coincides with the photon self-energy function defined in the
preceding section.
The totality of all the compact photon self-energy parts corresponds
to the function II* (k) = £ll%,(k), which is related to Fl{k) by
W
11= n *+ n * D cI l * - sr ... = n * ( \ —Dcn*)~l, i.e., on the basis of
(43.38) we have TI*{k) = —iP(k), where P(k) is the polarization
operator.
We also define an effective photon line joining points A and B of
a diagram as the totality of all the photon self-energy parts together
with the two photon lines joining them to the points A and B. We
598 QUANTUM ELECTRODYNAMICS

represent an effective photon line in our diagrams by means of a thick


broken line.
The quantity in the matrix element which corresponds to an effective
internal photon line is the photon Green’s function Gy(k). Fig. 63 shows
the graphical relations between Gv, 17 and 77*, corresponding to equations
(43.37) and (43.39).

F ig . 63.

The external effective photon line is related to the self-energy parts


by the same relations which are shown graphically in Fig. 63 and which
are equivalent to equation (43.44'). Because of this we can leave out
of consideration those parts of the diagram which are equivalent to
external photon lines, since the latter correspond to the matrix elements
of the operator A which are determined by formula (43.10).

44.2. Vertex Parts o f Diagrams


We shall call a part of a diagram which is connected to the other
parts by two electron lines and one photon line a vertex part. In Fig. 64
the vertex part is shown schematically in the form of a triangle.

i
ii

F ig . 64.

Vertex parts can also be divided into compact and noncompact


parts. The former are those diagrams which cannot be divided into
parts interconnected by only one electron or photon line. Noncompact
INVESTIGATION OF SCATTERING MATRIX 599

vertex parts can be regarded as combinations of a compact vertex part


with effective electron or photon lines. Therefore, in future we shall
use the term vertex part to refer only to compact vertex parts. Figure
65 shows vertex parts of the third and fifth orders.
We denote by A y the quantity which corresponds in the matrix
element to the vertex part Y, and we denote the sum of all the compact
vertex parts by A = £ A y. We call the quantity corresponding to the
Y
totality of vertex parts and to an ordinary vertex diagram (which cor­
responds to the factor y) the vertex function, and we denote it by r ,
where r = y -\-A.

Fig. 65.

The matrix A , like the matrix y corresponding to a simple vertex


of the diagram, has one vector index (p) and two spinor indices (a, /5):
A
1*
= A = A aft„ and is a function of two electron momenta (py, p 2)
*

and one photon momentum (k) corresponding either to external or


to internal lines, by means of which the vertex part is joined to the
rest of the diagram: A = A ( p 1, p t ; k). In virtue of the law of conserva­
tion of momentum these quantities are related by P i ~ P 2 — k.
In the coordinate representation A is a function of two differences
of coordinates of the three points defining the vertex part of the diagram.
It is useful to divide the self-energy and the vertex parts defined
earlier into two further classes: reducible and irreducible. We call
600 QUANTUM ELECTRODYNAMICS

a part which does not contain within itself any self-energy or vertex
parts irreducible-, in the opposite case we call it reducible.
There exists only one irreducible electron self-energy part, viz.,
the second order sell-energy part shown in Fig. 56 (1). It can be easily
shown, for example, that the remaining diagrams of Fig. 56 are reducible.

i
ii

Fig. 66. Fig. 67.

Thus, diagram 2 is noncompact, diagrams 3, 4 contain self-energy parts,


while diagram 5 contains a vertex part. Diagram 5 can be represented
in one of the forms shown in Fig. 66 where the triangles denote vertex
parts.
Similarly there exists only one irreducible photon self-energy part,
viz., the second order self-energy part shown in Fig. 58 (7).

There exists an infinite number of irreducible vertex parts. Of the


diagrams shown in Fig. 65 diagrams (7) and (2) are irreducible. An example
of a more complicated irreducible vertex part is shown in Fig. 67.
By utilizing the concepts of irreducible parts and effective external
lines we can replace complex diagrams which we encounter in the
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 601

investigation of higher order approximations by skeleton diagrams


in which effective lines vertex and self-energy parts have been segregated.
Each one of these parts represents a set of diagrams of different orders.
As an example we consider the fourth order Compton effect dia­
grams shown in Fig. 68. The totality of these diagrams and of the second
order diagram shown in Fig. 21 reduces to the single effective skeleton
diagram shown in Fig. 69. This diagram corresponds to the matrix element

M = Z 1Ze2a2/ialvu2r /i(p2+ k 2, p 2; k 2)Ge(p2+ k ^ r v(p1, P i+ fq ; k ±) .


(44.1)
If we have second order approximation expressions for r and Ge
it is sufficient to substitute these expressions into M and to neglect
higher order terms in the product in order to obtain the sum of matrix
elements corresponding to the diagrams of Figs. 21 and 68.

Fig. 69. Fig. 70. Fig. 71.

With the aid of the vertex function we can construct a general expres­
sion for the mass and the polarization operators. In order to do this
we investigate the internal structure of the exact compact self-energy
part. It can be represented by means of the skeleton diagram shown in
Fig. 71. On the basis of this diagram we can write £*(p) in the form

Z-(p)=~,Jy„ 0 1( . p - k ) r , ( p , p ~ k ; k ) G i . ( k ) d ‘k . (44.2)

Similarly the compact photon self-energy part can be represented


by means of the skeleton diagram shown in Fig. 71. On the basis of this
diagram we can write the expression for 17* (k):

77*(k) = } s p J y l,G‘ ( p ) r iJ(p, p - k ; k)G’( p - k ) d‘p. (44.3)


602 QUANTUM ELECTRODYNAMICS

On substituting (44.2) and (44.3) into (43.39) we obtain two integral


equations relating the three quantites Ge, Gy and I 1

G*(p) = S % p ) + - ^ r Se(j>) J y ^ p -Q rA p ^ -k ^ Q G 'ip )


X Gyv(k) d*k,
(44.4)
Gy(k) = D°(k) + — ^ D ' ( k ) S p y llf Ge(p)
X ( p , p —k ; k)Ge( p - k ) <PpGv(k).
If there existed a third equation defining the vertex part we would
have a closed system of three exact equations of quantum electrodynam­
ics. However, there is no such exact equation, since the exact equations
of quantum electrodynamics are, in fact, not integral equations but
functional equations (cf. § 49).
j*=0
i

Fig. 72.

However, we note that for k = 0 there exists a simple relation


between the vertex part and the electron self-energy part known as
Ward’s identity (202)
dZ*(P) >
----------- (44.5)
A l i ( P , P ‘, 0 ) =
dp

from which on the basis of (43.40) it follows that


1
r u(p>p\ 0 ) = - (44.6)
dpu \ Qe(P)
In order to prove equation (44.5) we use the identity
dS c(p)
= - s c(p)
dPu M - m ) s % p ) = s c ( p ) y ^ ip )- (44-7)
The function S c(p) is represented graphically by an electron line. The
right hand side of this identity corresponds to a diagram consisting
of two lines with identical momenta and of one vertex (Fig. 72). This
establishes the relation between the diagrams corresponding to the
function Sc(p) and its derivative.
INVESTIGATION OF SCATTERING MATRIX 603

We now consider a compact self-energy part W, for example the


one shown in Fig. 73. We choose the momenta corresponding to the
variables of integration in such a way that the momentum p should
appear in each segment of the solid line joining the initial and the final
points of the diagram (Fig. 73). Differentiation of the factors correspond-

P -*i p-k^
■ <—r*---
y p
'k
Fig. 73.

ing to different segments of this line will give us the complete set of
vertex parts which can be constructed by adding photon lines to the
different electron lines forming the diagram W (Fig. 74). The electron
lines forming closed loops will not take part in the differentiation.
But such loops cannot lead to vertex parts. Indeed, on the basis of

iI I
II
I
◄---- -*---
\ " —— /
« .« «
\ ■— /
, < ■ \ " — "
s+ ;<■
/

Fig. 74.

Furry’s theorem (cf. subsection 25.4) we need take into account only
closed loops with an even number of vertices. But differentiation adds
an extra vertex, and, therefore, all such diagrams may be neglected
on the basis of Furry’s theorem.
We have thus shown that the set of all the compact self-energy
parts gives on differentiation the set of all the nonvanishing vertex
parts, i.e., relation (44.5).
Relation (44.6) also enables us to obtain an expression for r ^ p , p ; 0)
in the case p 2+ m 2c = 0, i.e., in the case when the vertex part terminates
by an external electron line. In this case Ge(p) is equal to expression
(43.46), on differentiating which we obtain

r fl( p , p ; 0 ) = Z x xy p 2+ m 2c = 0. (44.8)
604 QUANTUM ELECTRODYNAMICS

44.3. Renormalization o f Electron Mass


The difference between the energy spectra of the free electron-
positron field and of the electron-positron field interacting with the
electromagnetic field, which manifests itself in the difference between
the masses m and mc, introduces difficulties into the solution of prob­
lems of quantum electrodynamics, since the usual scattering theory
methods assume that the energy spectrum of the system is independent
of the perturbation.
However, in spite of the difference between the masses m and mc
it is possible to redefine the unperturbed system in such a way that
the noninteracting electron will also have the mass mc. This is carried
out by the following simple method, which is called mass renormali­
zation.
We write the Lagrangian density for the system of interacting fields,
as before, in the form L = Le+ L but introducing the following
change of notation. We interpret L e as the Lagrangian density for
a free electron of mass mc
L e = y)(ip-\-mc)y} = yj(ipJrm)ip-\-dmipy). (44.9)
We subtract from L, the term dmxpy added to L e
A

L i = ieyiAip—dmyTp. (44.10)
Then the scattering matrix assumes the form

S = T ( e f L *'x) = (44.11)
The matrix elements of S will contain in addition to the ordinary
three line vertices corresponding to the first term of (44.10), in which
two electron lines and one photon line meet, also two line vertices (in
Fig. 75 such a vertex is shown by a cross). In such vertices correspond-

Fig. 75.

ing to the second term of (44.10) only two electron lines meet. The addi­
tional term in Lt obviously leads to only one additional irreducible
diagram representing an electron self-energy part. Every diagram
containing an irreducible electron self-energy part now has a similar
diagram added to it in which the irreducible self-energy part is replaced
INVESTIGATION OF SCATTERING MATRIX 605

by the self-energy part shown in Fig. 75. Therefore, we can introduce an


irreducible self-energy part renormalized with respect to its mass which
consists of a sum of self-energy parts corresponding to Fig. 56 (7) and
Fig. 75. It corresponds to the function E {2)'(p) = (p)—idm.
We now consider an arbitrary reducible compact diagram W for
an electron self-energy part. We can evaluate E w according to the for­
mer by identifying m with mc, and by making each irreducible compact
self-energy part correspond to the function27(2)', and not to the function
Z {2). After this has been done, the self-energy part of Fig. 75 must
be added to the diagram as a whole.
The renormalization of the mass leads to the replacement of the
mass operator M(p) by M f p ) = M (p )—dm. For p = p Q, where
Pi = —m2c we have in accordance with (43.48) M x(p0) = a0(ip0+ m c),
where a0 = Z f 1—1.
The quantity M x(p) can be expressed in terms of the vertex part
A ( p , p ; 0 ) = A(p). On utilizing expressions (44.5), we obtain
p

(44.12)

where the integration is taken along an arbitrary curve joining the


points p 0 and p.

§ 45. Renormalization of Electron Charge

45.1. Physical Charge o f the Electron


We have already seen that because of the interaction between the
electron and the electromagnetic field the electron mass mc does not
coincide with the constant m appearing in the equations of quantum
electrodynamics. We shall now show that neither is the constant e
which appears in the equations of quantum electrodynamics and in
the scattering matrix the actual electron charge, and we shall obtain
the relation between the constant e and the charge of the electrbn.
The physical charge of the electron can be determined by investigat­
ing the scattering of electromagnetic waves of low frequency k by an
electron at rest. This process is a purely classical one and no quantum
corrections should be necessary in the limiting case k -> 0. It is, therefore,
natural to define the physical charge of the electron ec as the coefficient
606 QUANTUM ELECTRODYNAMICS

appearing in the Thomson formula for the effective scattering cross


section for a photon of k = 0
- I e* Y
° 3 \ Annie2 J
We shall show that the exact solution of this problem in quantum
electrodynamics leads to the Thomson formula with the following value
of the charge ec expressed in terms of the constant e:
ec = Z 1/2e. (45.1)
In order to prove (195) this fundamental relation we consider all
possible diagrams describing the scattering of a photon by an electron.
We divide such diagrams into three classes (Fig. 76): 1) diagrams which

1I
(3)
Fig. 76.

can be reduced to the skeleton diagram of Fig. 69. 2) diagrams which


represent compact electron self-energy parts, to the electron lines of
which two external photon lines are connected, 3) diagrams in which
the external photon lines are connected to closed electron loops.
Diagrams of the latter type give no contribution to the matrix element.
Indeed, diagrams of this type containing loops with an odd number
of vertices may be left out of consideration in virtue of Furry’s theorem.
Other diagrams of this type containing an even number of vertices
INVESTIGATION OF SCATTERING MATRIX 607

(cf., for example, Fig. 76 (3, 4)) are related for k = 0 by expression
(44.5) to diagrams containing loops with an odd number of vertices,
and, consequently, also give no contribution to the matrix element.
We now consider matrix elements corresponding to diagrams of
the first and second type. The matrix element corresponding to a dia­
gram of the first type can obviously be written in the form
M, = ZZ Xe2u2(F Ger v+ r , Ger j u, a, a2V
Since we are considering the case k = 0, then on the basis of Ward’s
identity the matrix element corresponding to a diagram of the second
kind is of the form
d2Z*(p)
M 2 = Z Z 1e2u2 Ui ai/ia2v'

By utilizing expressions (44.5), (44.6) it can be easily shown that


d2Z*{p) _ d2Ge~\p)
dpMdpv ~ dp^dpv
Further, on utilizing the expression

- ^ = a ‘( p ) r r ( p ,p ,o ) a ‘(p),
°PU ^
we obtain
d2Ge d2Ge
G T v Ger Ge+ Ger Ger v Ge - Ge - — — Ge
Sp„dpv dp^Pv
From this it follows that
d2z *
r ( ? r , + r f 6 i r <<= G ,' W 6 ,‘1
8pudpv 1 ’ ' ’ " " dPftdp
Therefore, the sum of the matrix elements M y and M 2 can be written
in the form
d^Ge i
M = M 1-)- M 2 = ZZ± e2u2Ge —— r G U\@ijjCi2t.
dPn°Pv
On utilizing the expression
Ge{p) = ZxS c(p) [1 + Cl(p2) + pc2(p2)], (45.2)

which follows from (43.20), and also the fact that


Ci(—m2) = c2(—m2) = 0,
608 QUANTUM ELECTRODYNAMICS

we obtain for M the expression

d2Sc
Ulahia2v
dPv

+2( — J Ze2u2(yfipv+ y yp/) u 1a1/Aa2y.


\ Op I v1——m*

If we choose the potentials a1 and a2 in such a way that the following


conditions are satisfied
aivPr= 0
then the matrix element will assume the form

d2S c(D1
M = Ze2u2S c 1(p)-~— s ° 1(P)uialfia2v.
opMdP,
On noting that
r)2 ri\
-P— 5 ~ = S c(p)y S c(p)yv+ S c(p)y„Sc(p)y
op,, op„ h
we finally rewrite M in the form
M = Ze2u2[yflS c(p)yy-\- Y , S c(p)y/i]u1a1^a2y.

On comparing this expression with expression (28.2), (28.3), we see


that for k = 0 the exact matrix element differs from its first order
approximation only by the factor Z. Since with the aid of the first order
approximation we have obtained in § 28 the Thomson formula for
k = 0 in which the quantity e played the role of the charge, we shall
now obtain the same formula in which, in accordance with (45.1),
the charge will be the quantity ec.

45.2. Renormalization of Propagators and Vertex Parts


The foregoing formula (45.1) elucidates the meaning of the constant
Z. Of great importance is the fact that in the expression for the matrix
elements of any arbitrary processes the constant Z can appear only
in the form of the combination Ze2 = e2. Our immediate problem is
to prove this assertion. We shall also prove that the constant Z 1
INVESTIGATION OF SCATTERING MATRIX 609

does not appear at all in the expressions for the matrix elements (53),
(203), With this aim in view we introduce the functions
G%= Z ? G e, (45.3)
Gl = Z~lG \ (45.4)
which we shall call therenormalized propagators. In addition, we intro­
duce the renormalized vertex function r c
rc= z,r, (45.5)
and also the function A 1
A x = Z XA. (45.6)
Between the renormalized propagators and vertex parts we can
establish a number of relations analogous to the relations which hold
between the original (unrenormalized) quantities.
First of all, we prove the relation
r c{px, p 2\ k) = y + A B(Pl, p 2; k), (45.7)
where
A R(p1, p 2; k) = A l (p1, p 2; k) —A x(pq, p 0\ 0)
and Pq is the momentum of the free electron which satisfies the condi­
tion Po+mc = 0- The quantity A R is known as the regularized vertex
part.
For our proof we use the definition (45.6) from which it follows
that
A l (Po,Po\V)= Z m p ^ p ^ - y ) .
On substituting into this equation expression (44.8) we obtain A x(p0,
p0l 0) = ( l - Z ^ y . Thus formula (45.7) assumes the form
r c{p1, p 2\ k) = Z t f + A ^ p ^ p i , k), (45.8)
i.e., it coincides with the definitions (45.5), (45.6).
Further, we show that the renormalized function Gec satisfies the
equation
Gec(p) = S c(p)—iSc(p)MR(p)Gec(p), (45.9)
where
P
M R(p) = iZ* = i J (A1/t(q, q; 0 ) - A lfj(Po, p 0; 0)) dq^. (45.10)
Po
610 QUANTUM ELECTRODYNAMICS

The quantity Z* is called the regularized electron self-energy part, while


M r is called the regularized mass operator.
In order to prove (45.9) we start with equation (43.39)
Ge( p ) = S c(p)—iSc(p)M(p)Ge(p)
and assume that the mass has been renormalized, i.e., we interpret
S c(p) as S cmf p ) and M(p) as M f p ) .
On substituting for M expression (44.12) we obtain from the preced­
ing expression
V

Z^Gec(p) = S c(p)-\-Sc(p) j A f q , q; 0)dqGec( p ) ~ i S c(p) M { p f) Z xGef p ) .

In order, to show that this equation coincides with (45.9) we subtract


it from (45.9). Making use of the definition (45.10) we obtain
V

0 - Z i)Gec(p) = s c(p) \ — $ A x(p i ,P i , 0) dq+iZ1M ( p 0)^G ec(p).


p "

Since A x(p0, p 0; 0) = (1 — Z x)y and M (p_) = (Z f1— l)((p0+ m c)(cf.


(44.12)), then it follows that the expression in the braces is equal
to —(1 —Z x) ( p —imc). Since S c(p) (p—imc) = —1, we obtain an identity
which proves the validity of equation (45.9).
The renormalized function Gvc satisfies the equation
Gl(k) = Dc(k) —iDc(k) PR{k)Gvc(k), (45.11)
where
k
P ,(k) = - i f V ^ (q )J q r ,
0

v , M ) = ZV M VM =
The quantity PR is called the regularized polarization operator.
In order to prove the validity of equation (45.11) we substitute into
it the definitions (45.12). On noting that on the basis of (43.53) we have

r L " ' z K L = 2,(i- z)4 - p(o)=o’


it can be easily shown that equation (45.11) is a consequence of equation
(43.39).
INVESTIGATION OF SCATTERING MATRIX 611

45.3. Three-Photon Vertex Parts


For subsequent development it is important to note that the quantity
Vpik) = i[dP(k)/dk2]2kfi= — [dn*(k)/dk^\ is analogous to the vertex
part A ^ p ) = A ti(p,p; 0) = —i [dM{p)jdp^J and that rules can be formu­
lated for evaluating it in accordance with corresponding diagrams.
We consider a part of a diagram which is joined to the other parts
by three photon lines. We call it the three-photon vertex part, and
denote the quantity corresponding to it by V , k 2, k 3), where k l ,
k 2, k 3 are the momenta of the photon lines leaving the diagram, with
k x-\-k2-\-kz = 0. The simplest example of a third order three-photon
vertex part k 2, k 2) is shown in Fig. 77. We recall that such,
parts have no real significance for finding the elements of the scattering
matrix, since together with a diagram of the type of Fig. 77 we must

Fig. 77.

always take into account a similar diagram but with the opposite sense
of circulation along the closed electron loop, and in accordance with
Furry’s theorem the sum of such diagrams gives no contribution to
the matrix element. Nevertheless, the quantity V ^ (kx, k 2, k 3) defined
in accordance with the diagram of Fig. 77 is itself different from zero,
and we can consider it formally.
Let k x = —k 2 = k, k 3 = 0. We can easily show by utilizing expression
analogous to (44.7) that in this case V/iafi(k,—k , Q ) = —dn<f(k)/dk/i
where T I ^ { k ) is the second order photon self-energy part shown in
Fig. 58 (/). Since I l afi(k) = TI(k) (dap—kakp/k2), the vector V/t (k) defined by
V (k) = -i- V/iaP can be expressed in terms of W 2)(k) in the manner
dfP2)(k)
*?>(*) = - dk fi
In order to generalize this relation to three-photon vertex parts
and to compact photon self-energy parts of arbitrary structure, we
612 QUANTUM ELECTRODYNAMICS

introduce one more definition, viz., in analogy with the first order
vertex part, i.e., with the ordinary vertex diagram which corresponds
to the factor y ~ —d(Sc(p))~1/dpfl, we formally define the first order
three-photon vertex part
dlD'ik ) ] - 1
2ik

and represent it by the fictitious diagram of Fig. 78 in which three photon


lines emerge from one point.
We now consider an arbitrary compact photon self-energy part.
We must distinguish between two types of corresponding diagrams.
In diagrams of the first type (for example, in the diagram shown in
Fig. 79) both external photon lines emerge from the same electron loop.

k+q'

Fig. 78. Fig. 79.

Therefore, the external momentum k refers only to the electron lines


of one half of the loop (for example, of the upper half-loop of Fig. 79).
Differentiation of the self-energy part of Fig. 79 gives the set of three-
photon vertices corresponding to the diagrams of Fig. 80.
//

In diagrams of the second type (for example, in the diagram shown


in Fig. 81) the external photon lines emerge from different electron
loops. In this case the external momentum must also refer to the photon
lines (for example, in Fig. 81 the upper dotted line joining the two electron
INVESTIGATION OF SCATTERING MATRIX 613

loops). In differentiating such self-energy parts terms will appear con­


taining dDc(k)/dkfi= Dc(k)2ik/jDc(k). In accordance with the defi­
nition cf the first order three-photon vertex part introduced earlier we
can likewise represent such terms by diagrams (Fig. 82).

Fig. 81.

Therefore we can make each diagram of a photon self-energy part


correspond to a set of diagrams of three-photon vertex parts VY . Further,
on introducing F/z = £ Fy/1, we obtain V^{k) = —[dIJ*(k)/dk/j]
— i[8P(k)/dk^[] in accordance with (45.12).
On defining, by analogy with the quantity r (k,k, 0), the three-
photon vertex function
A W = 2 ikr+ V f i ) , (45.13)
we obtain a relation analogous to (44.6)
d[Gy(k)]-'
4 .(* ) (45.14)
dk"
i*
We also introduce the renormalized three-photon vertex function
A= ZA M. (45.15)
By utilizing (45.15) and (45.13) we can easily show that

4„ = 2 * „+ V (45-l6)
Formulas (45.15) and (45.16) are analogous to the corresponding
formulas (45.5) and (45.7) for the vertex function r .

45.4. Renormalization of Matrix Elements


To evaluate the matrix element for a process it is sufficient to con­
sider the set of the corresponding irreducible diagrams, i.e., diagrams
which within themselves contain no self-energy or vertex parts. All
the reducible diagrams will be taken into account if we treat each irre­
ducible diagram as a skeleton diagram in which the internal electron
614 QUANTUM ELECTRODYNAMICS

and photon lines and the vertices correspond to the factors Ge, Gv and
r , while the external lines correspond to the amplitudes u, a.
Let a given irreducible diagram contain n vertices, Fe internal electron
lines, Fy internal photon lines, Ne external electron lines, and Ny exter­
nal photon lines. Since two electron lines and one photon line meet
in each vertex, and each internal line appears in two vertices, while each
external line appears in one vertex, we have

n = Fe+ - £ = 2Fv+ N v. (45.17)

We can write the expression for the matrix element in the following
schematic form:
M ~ e nf ( r ) n(Ge) Fe(Gv)Fy ( u f e ( a f r , (45.18)

which merely indicates the number of factors of each type appearing


in M. On substituting into (45.18) the renormalized quantities
r = Z XXF C; Ge = Z xG%; Gy = ZGy;
u = Z\l2u\ a = Z xna\ e — Z~ll2ec,
we obtain on taking into account (45.17)

M ~ e ? f ( r c)"(Gec)Fe(Gr)Fr (uf? (a)V (45.19)

We see that the constants Z and Z x do not appear in (45.19), i.e.,


the matrix element can be evaluated with the aid of the renormalized
quantities r c, Gec, Gvc in accordance with the same rules as in terms
of the original quantities r , Ge, Gy, if we at the same time replace the
constant e (which we can call the primary charge) by the true electron
charge ec.
In evaluating matrix elements corresponding to various processes
it is desirable to utilize the renormalized quantities from the outset.
In accordance with the results obtained in subsection 45.3, it is sufficient
for this to have rules for the evaluation of A x and Vx, since the quanti­
ties r c, G% and Gy can be expressed in terms of A x and Vx by means
of formulas (45.7)-(45.16).
In order to establish such rules we consider first of all an irreducible
skeleton diagram A. Let this diagram contain n vertices (n is an odd
INVESTIGATION OF SCATTERING MATRIX 615

number, for examples cf. Fig. 83). The expression for A can be sche­
matically written as
n —1
A ~ en 1J cn n ((j)"-1( g ^ . (45.20')

On replacing all the quantities by the corresponding renormalized


quantities we obtain
n —1
A 1 ~ e r 1J ( r y ( G $ * ~ ' (45.20)

Fig. 83.

i.e., A x can be evaluated by means of the same rules as A merely by


replacing the original functions and the quantity e by the renormalized
functions and the true electron charge ec.

Fig. 84.

In the same way we can easily show by considering the irreducible


skeleton diagrams of the three-photon vertex function (for example,
cf. Fig. 84) that the renormalized quantity Vx = Z V is related to the
renormalized functions Gec, Gvc , r c, A c by the same formulas which
relate V to the functions Ge, Gy, r and A with e replaced by ec.
616 QUANTUM ELECTRODYNAMICS

45.5. Formulation of Perturbation Theory as an Expansion in Powers


of ec.
The foregoing analysis enables us to evaluate the elements of the
scattering matrix in the form of an expansion in powers of the true
electron charge ec.
Each physical process can be made to correspond to a set of irreduc­
ible skeleton diagrams W with different numbers of vertices N.
A skeleton diagram W consists of effective external lines and of
an internal part. We denote the expression corresponding to the internal
part of the diagram W by K(W). Then the exact value of the matrix
element will be determined by the quantity K = ^ K ( W ) .
W
In accordance with (45.19) K{W) contains the factor e^ and can be
expressed in terms of the functions r c, Ge and Gvc , each of which can
in turn be represented in the form of a series in powers of e\. If we confine
ourselves in these expansions to terms proportional to any given power
of ec, then we shall finally go over from skeleton diagrams to elementary
irreducible diagrams, i.e., to diagrams in which the vertices and the
lines correspond to the factors y , S c, Dc which no longer contain ec.

F ig . 85.

As an example we consider the evaluation of the matrix element


for the scattering of a photon by an electron. In the first approximation
i.e., up to terms in e\, obviously only the elementary diagram of Fig. 69
is retained, and we obtain the formulas of § 28.
In the second approximation, i.e., up to terms in e*, firstly the ele­
mentary diagram of Fig. 85 is added; secondly, the diagram of Fig. 69
must be treated as a skeleton diagram in which Fc and Gc must be
evaluated up to terms of order e\, i.e., Fc = y +/1^3) = y -|-/1{1) —/1[3)
and Gc = S c-\-ScF {f )S c. The vertex part /l[3) is evaluated in accordance
with the elementary diagram of Fig. 83 (/). The regularized self-energy
part 2^2) can be directly expressed in terms of the self-energy part £ <2)
INVESTIGATION OF SCATTERING MATRIX 617

evaluated in accordance with the diagram of Fig. 56 (7), by means of


the formula

>(p) = ~ B
— ~ (P-Po,)- (45.21)

Indeed, it follows from the definition (45.10) that


P

Po

r dZ™(q) dZW(p)
J dq d% ~ I (P^-Po^)-
dPM P = Po

In the third approximation (up to terms of order e®) we must evaluate


the diagrams of Fig. 86 treated as elementary diagrams. The diagram

i
ii

i i
ii ii
Fig. 86.

of Fig. 85 should be treated as a skeleton diagram, in which the quanti­


ties r c, G% and Gy are evaluated up to terms of order e2. The first two
have been discussed earlier. For the photon Green’s function Gyc we have
similarly in this approximation Gy = Dc-\-DcI J ^ D c. The quantity
77jj2> can also be expressed directly in terms of the photon self-energy
part 77(2) evaluated directly on the basis of the diagram of Fig. 58 (7)
by means of
Id 77<2)\
77<2>(/c) = 77<2)( £ ) - 7 7 (2)( 0 ) - l - ^ - l a k \ (45.22)

which follows from the definitions (45.12).


618 QUANTUM ELECTRODYNAMICS

In the diagram of Fig. 69 when calculations up to terms of order


e® are being performed we must take into account in the expressions
for r c and Gc terms of order e*. The expression for r c has the form
r c — y A (R3>+ A (^ . The vertex part A[5) contains the factor el and,
therefore, must be evaluated directly in accordance with the elementary
irreducible diagram of Fig. 83 (2). The vertex part A[3) contains the
factor e\. Therefore, it must be evaluated on the basis of the skeleton
diagram of Fig. 83 (1) into which expressions for Fc, Gec and Gvc accurate
to terms of order e\ are substituted.
In order to evaluate M R up to terms of order e\ we can utilize formula
(45.10) after substituting into it the function A a on the basis of the dia­
grams of Fig. 83 up to terms of order e\. From this we can easily
establish the following rules for the evaluation of in terms of the
reducible elementary diagrams for the fourth order electron self-energy
parts Ej4\ Effl and Effi shown in Fig. 87.

a b

i n ///
F ig . 87.

In the case of diagram I which contains within itself the second


order self-energy part E (2) it is necessary, first of all, to regularize the
latter, i.e., to replace it by the quantity E {2) in accordance with formula
(45.21). Then, after calculating, with this replacement kept in mind,
the quantity corresponding to the whole diagram which we denote by
Ejiy, we must once again perform a regularization, i.e., replace E\iy by
the quantity E (fy in accordance with a formula analogous to (45.21)

ZtfHp) = <45-23>

For diagram II containing the photon self-energy part of the second


order 77<2) it is first of all necessary to regularize the latter, i.e., to re­
place 77(2) by 77<j2) in accordance with formula (45.22), and then, after
having evaluated on the basis of such modified diagrams the corre­
sponding quantity Z \ l y, to regularize it by means of formula (45.23).
INVESTIGATION OF SCATTERING MATRIX 619

In the case of diagram ///containing the vertex part /1<3) the following
more complicated expression is obtained:

^
TW
IIIR
— R ( p , ti, t ^ ) ~ R ( p Q, t1, t2) (p^—p0^)

dR(p, h, t2) \
x d* tx <i4t2, (45.24)
dP/t h.
where
R ( p , h , t2) = Ffl(p, h)Gvll(p, tu t2) Hv(p, t2)
~ Ffi(Po, h)Gv^(Po> h, 0)Hv(p, t2)
- F / S P ’ h)GVfl(po, 0, t2) H v(p0, t2),
f m(p , h) - Dc( h ) y ^ s c( p - t d ,
GVfl(p, tlt Q = iyvS c{ y - t x- t ^ ) y ii,
Hv(p, t2) = S c(p—t ^ y vDc{Q.

§ 46. Divergences in the Scattering Matrix and Their Removal

46.1. Divergences in Irreducible Diagrams


In the evaluation of higher order approximations to the elements of
the scattering matrix we encounter fundamental difficulties which
for many years gave rise to the opinion that the applicability of quantum
electrodynamics is, in general, limited to the first approximation. These
difficulties consist of the fact that the higher order approximations
to the matrix elements contain integrals which diverge for large values
of the momenta of the virtual particles.
In this section we shall examine in detail the problem of these diver­
gences and the methods of removing them. In the course of this discussion
we shall not investigate the divergences in the region of low momenta
(“ infrared catastrophe”) which are encountered in a number of cases.
The latter are associated with the inapplicability of perturbation theory
at low photon frequencies, and can be removed comparatively simply
by the method described in § 30.
We consider an irreducible part of a diagram. In the general case
it corresponds to a multiple integral over the momenta of the virtual
particles / = / F(qx ... qn)di ql . ... di qn. The integrand in this integral
620 QUANTUM ELECTRODYNAMICS

is a rational fraction. Let us determine the degree of this fraction.


In order to do this we note that each internal electron line of the dia­
gram corresponds to the factor S c(p) which contains the inverse first
power of the momentum, while each internal photon line corresponds
to the factor Dc(p) containing the inverse second power of the momentum.
Therefore, the integrand is a fraction of degree —(Fe-\-2F ), where Fe
and F are the numbers of internal electron and photon lines. We
now determine the number of variables of integration. The number
of different momenta is equal to the number of internal lines, i.e.,
Fe-\-F . But they are not independent, since the three momenta of the
lines meeting at each one of the n vertices of the diagram are related
by the law of conservation of momentum. One of the conservation
laws may be applied to the external lines, so that the total number
of independent momenta over which the integration is carried out is
equal to Fe+ Fy—n + 1. If we write the integral schematically in the
form

then on taking into account the foregoing calculation and utilizing


relations (45.17) we obtain
Kx = Fe+Fy- n + 1 = +
N
K2= 2 n - - £ - N y,

where N e and Ny are the numbers of external electron and photon lines.
Since we are considering an irreducible diagram, the integrand
cannot be decomposed into factors containing independent variables.
Therefore, the convergence of the integral / is determined by the difference
K = K . - 4 K X= I Ne+ N . - 4 . (46.1)
For K > 0 the integral is convergent, while for -K < 0 it is divergent.
It is noteworthy that the quantity K depends only on the number of
external lines.
It follows from (46.1) that there exists a limited number of types
of divergent integrals corresponding to the following values of the
numbers Ne and N :
INVESTIGATION OF SCATTERING MATRIX 621

1) Ne = 2 ,N r = 1, K = 0,
2) Ne = 2, N r = 0, K = -1 ,
3) ^ e = o,Nr = 4, K = 0,
4) Ne = 0 ,Nv = 3, K = -1 ,
5) Ne = 0, Nr = 2, K = -2 ,
6) N e = 0,Nr = 1, K = -3 ,
7) N e = 0,Nr = o, K = -4 .
Figure 88 shows the simplest irreducible diagrams corresponding
to these divergences. Not all of them can, in fact, appear in the expressions

( 6)

Ne =0
(7)

Fig. 88.

for the matrix elements. Thus, for example, the diagram of Fig. 88
(7) is a vacuum loop which, as has been explained in subsection 44.1,
need not be taken into consideration. Diagrams of the type of Fig. 88
622 QUANTUM ELECTRODYNAMICS

(4, 6) which represent closed electron loops with an odd number of


vertices, can also be left out of account in accordance with Furry’s
theorem (cf. § 25). However, we retain diagrams of the type of Fig. 88
(4) since they determine the auxiliary quantity—the three-photon
vertex V^—which we have introduced previously in subsection 45.6.
We see that the principal irreducible parts, from which the expression
for the scattering matrix in higher order approximations can be construct­
ed, are divergent expressions with values of K = 0, —1, —2. Among
them are included the irreducible self-energy parts Z i2) and /7 (2), the
vertex parts A i2n+1), and also the part corresponding to photon-photon
scattering (diagram of Fig. 88 (J)).

46.2. Introduction of a Cut-Off Momentum


The appearance of divergences in the scattering matrix shows that
the theory is unsatisfactory. It is clear that the perturbation theory
series is, strictly speaking, meaningless if the second term of the series
is infinite. Thus, the following problem arises which is of the greatest
fundamental significance.
On the one hand, the results of applying first order perturbation
theory, which were described in Chapters V and VI, are in excellent
agreement with experiment. On the other hand, in order that the first
approximation should have some theoretical significance, the subsequent
approximations should lead only to small corrections.
The idea of an escape from this contradiction is already suggested
by classical electrodynamics. As is well known, classical electrodynamics
is not a logically closed theory. Its consistent application leads to con­
tradictions, which, for example, manifest themselves in the infinite
electromagnetic mass of the electron. The content of these contradic­
tions can be expressed as the inapplicability of the equations of classical
electrodynamics at distances smaller than the classical electron radius
e2/mc2. (In fact, because of quantum effects classical electrodynamics
is already inapplicable at distances of the order of tijmc.)
If quantum electrodynamics, like classical electrodynamics, is also
inapplicable at sufficiently small distances, then without entering here into
a discussion of the problem whether this limitation is associated with
deficiencies of the fundamental ideas and equations of electrodynamics,
or merely with the use of perturbation theory methods, we may,
INVESTIGATION OF SCATTERING MATRIX 623

nevertheless, assume that our theory takes the interaction into account
correctly within the domain of space-time intervals |x2| > 1/L2, where L
is a sufficiently large quantity. Since in Fourier expansions large momenta
correspond to small |x2|, this means that the formulas of quantum
electrodynamics are valid only in the domain of momenta smaller than
some limiting momentum L. For \p\ ^ L the formulas of quantum
electrodynamics are incorrect, either because of the inapplicability of
perturbation theory, or as a result of the fact that the fundamental equa­
tions of quantum electrodynamics are inapplicable to processes for
which the domain of small space-time intervals plays an essential role.
We shall further assume that the domain of very large virtual momen­
ta \p\ > L actually plays no role at all in any process in which the
particles in colliding transfer to each other momenta Ap small compared
to L. With this assumption we can carry out integration in momentum
space over a finite region |/?| < L.
We shall show later that the cut-off momentum L can be chosen
in such a way that, on the one hand, the condition Ap L is satisfied,
and, on the other hand, the use of perturbation theory is justified. We
shall see that the expression for the matrix elements evaluated by methods
presented in § 45 does not depend on the cut-off momentum L which
appears only in the expressions for mass and charge renormalization.
We shall also see that the use of perturbation theory in quantum electro­
dynamics can be justified over a very wide range.
Thus, we shall assume that the integrals of interest to us, which
determine A , E and Z7 and other quantities, are evaluated over a finite
region in four-dimensional momentum space \p\ ^ L. This region
must obviously be invariant under Lorentz transformations.
Instead of introducing a finite region of integration we can also
use another equivalent method. It consists of introducing into the
integrand a “cut-off factor”, i.e., a certain factor f( q , L) which has
the properties f( q , L) = 1 for q < L and f(q , L) -* 0 for q > L, where
/ i s made to approach zero in such way that the integral converges. For
the cut-off factor we can choose, for example, a function of the form
(second of references (62))
L2
/(<?, L) = q2+ L 2
where n > 1.
624 QUANTUM ELECTRODYNAMICS

46.3. Convergence of Regularized Expressions for Irreducible Vertex


Parts and Self-Energy Parts
We shall now show that elements of the scattering matrix expressed
in accordance with formulas of sections 45 in terms of the true
electron charge ec and the true electron mass mc do not depend on the
cut-off momentum L, i.e., do not contain any divergences.
As we have seen, the matrix elements contain the renormalized
propagators and vertex functions Gec, Gvc, r c, which can be expressed
in terms of thfe regularized vertex parts A R and VR.
We begin by considering the simplest third order vertex part
A i3)(Pi,p2-,k) = j R ( p 1, p 2; q)diq, (46.2)
where

R ( P i , P i ’, q ) = ( f f f yas c ( P i - q ) y f, s c ( P 2 - q ) y [ ) Dcap(q')-

The regularized expression for the vertex part is, in accordance with
(45.7), given by
^ R ]{Pi,P2\k) = A f ( p 1, p 2 - k ) - A (f ( p 0, p 0; 0),
where p 0 is the momentum of the free electron: p\-rm2c = 0.
In accordance with the general results (46.1) the integral (46.2)
diverges logarithmically. Since the momenta p x and p2 appear in the
integrand R{pl , p 2\ q) in the form of the differences p x~q=^ tx and
P2 q ~ ^2
R(Pi,P 2 \ q ) ^ f { t i, t2,q),
the expression for R ( p i, p 2; q) can be written in the form
dR
R(Pi >P 2 j q) — R(Po> Po j 7 ) + (Plft Pot)
Pi = p ’

\4 ~ ) r(P 2 fl- P o fl) = R ( Po, Po; q)


\ r ' l n f Pi — p

df
(Plp—PoJ+i J, (P2fl P0f) s
I l=p'+<J '2d nt '/ l i ~ p u + q

where the derivatives are evaluated at some points p' and p" intermediate
between po and p± or p0 and p2. Since the function R is a rational fraction
of degree —4 with respect to q, then dfjdt1 and df/dt2 will be rational
INVESTIGATION OF SCATTERING MATRIX 625

fractions of degree -5, and the integral defining the regularized vertex
part
A r ] = § { R ( P i,P 2 \q )-R {p o ,P o \q )} d iq (46.2')
will converge. (This fact justifies the use of the term “regularized.”)
The foregoing proof of the convergence of can be easily general­
ized to the case of an arbitrary irreducible vertex part.
An irreducible vertex part of (2 « + l)th order yl(2n+1) (Pl, p 2; k)
(cf., for example, Fig. 83 for n = 2 and 4) can be written in the form
of a logarithmically divergent integral

Ai2n+1)(P1, p 2; k ) = j R ( p 1, p 2, q 1, ... , q n)d4q i ... d*qn.

The variables p1 and p 2 appear here only in the form of linear combina­
tions with the variables of integration = p x—q1; t2 = P i ~ q i ~ q 21 ■■■■
Therefore the difference is given by

R(Pi iP2->q\i ,q
n)
— R
(p
o,Po> q
i.
where the derivatives dR/dt1 are evaluated at some intermediate points
, and, consequently, the expression for
A Rn+1)(Pi,P2',k)

f
= {R(Pi,Ps, qt , .... qn) ~ R ( p 0,Po, qi, — , q n)} ••• d*qn
will be convergent.
We note that if we choose the transverse gauge for Dc, i.e., D^(q)
= Dc(q) (<5q/?—[qaqpjq2]), then the integral (46.2) will be convergent,
since the terms of highest degree in q will cancel in the integrand (second
of references (115)). However, such a simplification occurs only for the
third order vertex part yl<3), and does not occur for other irreducible
vertex parts. Therefore, in future we shall use the expression for Dc
in the following form Dcap{q)= 6a?Dc{q), since it leads to simpler algebraic
expressions.1
We now consider the simplest irreducible third order three-photon
vertex
Vl3'{k) = } Q J . k , q ) d ‘q, (46.3)

3 For each approximation the gauge of the propagator can be so chosen that
the expression for the vertex part will converge (24a).
626 QUANTUM ELECTRODYNAMICS

where

QtXk ’ ?) = J ^ c y SP yaS c( q ~ k ) y fi S c( q - k ) y aS c(q).


In accordance with (46.1) this integral contains a linearly divergent
part, i.e., it depends on L linearly. However, in view of the invariance
of the region of integration the terms proportional to L must vanish.
Indeed, we can assume for the investigation of the divergent part of the
integral (46.3) that k fi<s* qft. If we omit k in the integrand, then after
the integration and the evaluation of the trace have been carried out,
there will remain no vector parameter in terms of which the vector
quantity V might be expressed. A nonvanishing result can come only
from the terms of QMlinear in k , but they will lead only to a logarithmic
divergence.
From the foregoing it follows, in particular, that
V<3>(0) = f Q ( 0 , q ) d * q = 0 .
On substituting Q (k,q ) in

Q (.k ,q )= g (0 , ? H * x A k k
d k j o a ' \ dkadkp k' a p
we can easily show that the regularized expression for the three-photon
vertex part

k.3. = k -._ ) o*„ = / ( Qik, , ) - e ( p , , ) - #<,

is convergent, since the integrand which is proportional to the second


derivative d2Qjdqadq,i is a rational fraction of degree two units lower
than the degree of Q.
A similar proof can also be given in the case of an arbitraiy irreducible
three-photon vertex part.
From the fact that the expressions for the regularized irreducible
vertex parts A {3) and are convergent it follows that the expressions
for the regularized irreducible self-energy parts I {2) and I J {2) are also
convergent.
This can also be shown directly by using the explicit expressions
for 2^2)(p) and I I l2)(p). Indeed, in accordance with (45.21) Z {2) can be
obtained from U{2) by subtracting the two leading terms of the expansion
of £ {2) into a series in powers of p-p0, whence it follows that E (2) is
INVESTIGATION OF SCATTERING MATRIX 627

proportional to the second derivative of 27(2)(p) with respect to p evaluat­


ed at some point intermediate between p 0 and p. Since the momentum
p appears in the integrand of Z i2)(p) in the form of a sum involving
the variable of integration, then the integrand in ^ ^ ( p ) will be a rational
fraction of degree two units lower than the degree of the corresponding
integrand for 27<2)(p). Since in accordance with (46.1) the integral for
27<2) contains a linear divergence, the integral defining 27jj2) will be abso­
lutely convergent.
A similar situation exists in the case of the irreducible photon self-
energy part n {2). It follows from (45.22) that the expression for I I {2)(k)
can be obtained from the expression for 77(2)(k) by subtracting the two
leading terms in the expansion of 77(2)(A:) into a power series in A:. It
can be easily shown by repeating the arguments referring to 27jj2) that
this subtraction guarantees the absolute convergence of the integral
defining 77jj2).
However, here we encounter a peculiar contradiction. For the
derivation in § 45 of the fundamental formulas defining Gvc and PR we
used the property of the polarization operator P(0) = 0 which was
proved in § 44. For- the derivation of this property we naturally made
use of the fact that the quantity P(k) is finite. And yet, as we have seen
earlier, 77<2)(A:) contains a quadratically divergent integral, and after
the introduction of the cut-off momentum 77(0) ~ L2. We thus see that
the introduction of the cut-off momentum, which makes various quanti­
ties finite, violates the gauge invariance of the theory. Therefore, strictly
speaking, the simple replacement of the integrals appearing in the
theory by integrals taken over a finite region is not justified.
We can formally avoid this difficulty by means of the following
device. We consider the quantity Px(k) = P(k) —P(0). In an exact
gauge-invariant theory Px(k) = P(k). In our approximate treatment
we will introduce the cut-off momentum not in the integrals defining
P(k), but in integrals defining P^k). This will ensure both the finite­
ness and the gauge invariance of the resulting expressions.
We encounter a similar difficulty in the discussion of diagrams
corresponding to photon-photon scattering (cf. Fig. 88 (5)). In accordance
with (46.1) the expressions corresponding to such diagrams

M {k \ > , k 4) = J Q (k \ > k % , , k4 , (fa , ..., <7n) d (fa ... d (fn


628 QUANTUM ELECTRODYNAMICS

diverge logarithmically. The same divergence will also be contained


in the quantity

M (0, 0 ,0 ,0 ) = JQ (0 , 0 ,0 ,0 ,? ! . . . , q n) d iqx ... d*qn.

On the other hand, the gauge invariance of the theory requires that the
equation M(0, 0 , 0 , 0 ) = 0 should hold, since physical quantities can
depend only on fields, and not on potentials. We see that the divergence
of the quantity M ( k x, k 2, k 3, k x) is associated with the violation of
gauge invariance of the theory by the introduction of the cut-off mo­
mentum.
In an exact gauge invariant theory the following relation must
hold
M R(ki, k 2, k 3, kf) = M ( k x, k 2, k 3, k^),

where
M n(kx, k 2, k 3, fcj = M ( k x, k 2, k 3, k t) —M(0, 0, 0, 0). (46.4)

Therefore, we replace the matrix element M ( k x, k 2, k 3, k 4) by its regular­


ized value M R(kx, k 2, k 3, k x). The quantity M R(kx, k 2, k 3,k^), which
can be written in the form

M R(kx■
>k 2, k 3, k x)

= J {Q(k i , k 2, k 3, k i , q 1, qn) ~ Q ( 0 , 0 , 0 , 0 , ^ , . . . , q n)} d*qx . . . diqn,

will be finite, since ... qx ...)—Q(0 ... qx ...) is a fraction of degree


two units lower than the degree of the fraction Q(kx ... qx ...).
The problem of photon-photon scattering will be discussed in
detail in § 54.

46.4. Convergence o f Regularized Quantities in the Case of Reducible


Diagrams
We now proceed to investigate reducible vertex parts. In order
to obtain the regularized vertex part A R(Y) corresponding to a reducible
diagram Y, we must replace it by an equivalent irreducible skeleton
diagram.
INVESTIGATION OF SCATTERING MATRIX 629

We begin with the simplest example of a fifth order vertex part /1(5)
shown in Fig. 89. The irreducible skeleton diagram which is equivalent to
it is shown in Fig. 90, where the point A corresponds to a third order
vertex part A (3). In virtue of the results of § 45 in order to obtain A ^5)
we must first replace /1<3) by A ^ in accordance with formula (46.2'),
and we must then regularize the quantity evaluated in accordance with
the diagram of Fig. 90, which we denote by yF5)', by using the general
formula (45.7). Since we are interested here only in the problem of the

nature of the divergence of the quantity A (5y, then by omitting the


spinor matrices we .can schematically represent the expression for the
integral in A i5)’ in the following form:

A {5)'{Pi,P2\ k) ~ f R(p1,P 2 , q ) A (]?)( p - q , P 2 - r , k ) d iq, (46.5)

where R is the same function (a rational fraction of degree —4 in q)


which appears in the integral for zl(3) (46.2) and whose asymptotic
dependence on q for \q\2 >> \p\\, \p2\2 is of the following nature:

R(Pi,P*, q)diq ~ d { \ n q 2).

On the other hand, as we shall see later (cf. 47.4), for large
q /l<j3) behaves like In q2, and therefore the divergent part of the
integral defining A i5)' is of the form

J In q2d{In q2) ~ (In L)2,

i.e., A (5)' diverges logarithmically.


Since the integrand in the integral for /1<0)' differs from the cor­
responding expression for zl<3) only by a factor which depends on the
630 QUANTUM ELECTRODYNAMICS

momenta logarithmically, then by utilizing the same arguments which


were given earlier in subsection 46.3 we can easily show that the regular­
ized vertex part is convergent.
We now consider the fifth order vertex parts shown in Fig. 91. The
irreducible skeleton diagrams equivalent to them are shown in Fig.

92 where the thick lines correspond to t.he propagators Gec and GYCevaluat­
ed on the basis of the second order self-energy parts 2^2) and IJtfh
V

Since 2^2) = f A (£ \q)dq, then asymptotically for large p , F (2)~ p lnp2,


Po
whence we obtain for the asymptotic behavior of G%

Ge( ------------- -
cKP) p{\ + C ln/>2)

In order to determine the behavior of GY(k) for large values of


k we utilize the properties of the three-photon vertex part F (3). Since it
diverges logarithmically, it follows from dimensional considerations
that for large k it must be of the form, Vfi(k) ~ k fi In (L2jk2).
From this it follows that V (^J ~ k In k 2. Therefore

G<Y>
k2(l ■ ClnA-2)

(cf. the corresponding expressions in § 47).


If with the aid of these functions we evaluate the vertex parts A {5y
corresponding to the diagrams of Fig. 89-91, they will contain only
a logarithmic divergence. After regularization in accordance with
formula (45.7) we obtain a finite quantity. These results can be easily
generalized to the case of arbitrary reducible diagrams. On replacing
such diagrams by skeleton irreducible diagrams in which the lines and
the vertices denote respectively Gec, GY, Fc, and on reducing the latter
INVESTIGATION OF SCATTERING MATRIX 631

to the simplest irreducible diagrams, it can be easily shown that asymp­


totically for large momenta \p \2 > m2, |/c2| > m2

0i(p)~ sxp)fL -t\,

I,

where f s, f D and f r are dimensionless polynomials in the corresponding


logarithmic arguments. Thus, the diagram as a whole will correspond
to an integral the integrand of which will contain, in addition to rational
fractions of the type discussed in subsection 46.1, only logarithms of
momenta which do not alter the character of the divergence of the
integral.
In particular, these results also apply to expressions (45.23) and
(45.24) for the regularized fourth order self-energy parts. It can be seen
from formula (45.23) that after the removal of the divergence associated
with the internal self-energy part 27(2), the divergence in 27}4)/ can be
eliminated by subtracting the two leading terms in the expansion of
this quantity in powers of (p—p 0). In the case of the self-energy part
27)// the divergences are associated with the presence of vertex parts
at points a and b. Formula (45.24) eliminates the so-called overlapping
divergences associated with them.

§ 47. Evaluation of Self-Energy and Vertex Parts

47.1. Evaluation o f Integrals over Four-Dimensional Regions


The integrals which we encounter in the evaluation of quantities
corresponding to various diagrams are all of the form

F(p)dip
/ = (47.1)
J a1a2 ... an

where are polynomials of second order and F(p) is a polynomial


632 QUANTUM ELECTRODYNAMICS

of order n with respect to p . Such integrals can be conveniently evaluated


if we first utilize the identity

... H -M J*
1 X\

= (w—1)1 J dxxJ dx2...


o o

c __________________ dx „-i (47.2)


J [aiXn- 1+ a 2(xn- 2- x n_1) + ... + a B( l - * i ) ] n
or the equivalent identity
1 1
——-— —=(w—1)! f ux~2 dux f u2~3du2 ...
al a2 ••• an d J
i
f du 1_______________________
J Un~l [axui ... wn_ i+ a 2«i... wn_2( l —wn_ i) + ... + a n( l - w 1)]'1’
(47.3)
where x x = ux, x 2 = uxu2, . . . , x n_x = uxu2 ... un_x and x x, x 2, ..., x n_x
are scalar parameters. On substituting this expression into (47.1) we
obtain an integral in p-space

F(p) dlp
d(x i, ..., *n_i) — (47.4)
[(p-ay+ ir
where a and / depend on the parameters x x ... x n_x. The desired integral
I is obtained from /(jq ... xn_1) by integration over x x ... x n_x:
1 xi Tn-2
1 = ( n - 1)! / dxx f dx2 . . . f d x n_xJ(xx ... *„_,). (47.4')
0 0 0

We call this method of evaluating the integrals I, the method of


parametrization (second of references (62)),
First of all we demonstrate the validity of identity (47.3). For n — 2
it is self-evident, since
i
du
J _ = r _
oxo2 J \pxu +«*(! ~ u ) f
INVESTIGATION OF SCATTERING MATRIX 633

We now show that if it is valid for some n, then it is also valid for
n + 1. On noting that integrations over the different ut in (47.3) may be
interchanged we obtain
1 i
1
(n— 1)! dux
al °n + l
o
where A n is the denominator in the right hand side of formula (47.3). But

1 _ C______ Up 1du0_____
nAnar+1 ~ J [Au0+ a n+1( l - u 0)]n+1'
0
Therefore
j i i
-------------------- = n i l ug^duo ... ( du ^______ 1_-____ ,
J J " - 1 [ A u 0+ a n + 1 ( l - U o ) ] * + '

which is what we set out to prove.


We now proceed to evaluate the integral (47.4). We first assume
that the integral (47.4) is convergent. In this case we can shift the variable
p —a -+ p, as a result of which the integral assumes the form

K =
r f ( p ) d^p. (47.5)
J (p2+/)n
Obviously, we can confine ourselves to the evaluation of only the scalar
integral, since if f ( p ) = p j ^ p 2) , then K = 0; if f ( p ) = p fipvf 2(p2),
then
f P r P j i i P 1) p 2M p )
d*p
J (p2+l)n (P2+ i y
etc.
We now demonstrate the method of evaluating the integral (47.5).
In this integral the integration over p 0 is performed in accordance with
the rule for going around the poles (cf. § 17), i.e., along the contour
C shown in Fig. 93. But this contour can, evidently, be rotated by
n/2, as shown in Fig. 94. As a result p 0 will be replaced by ip'0 and
p2 = p 2—pl will go over into p >2,+p'02 (p'o is a rea* quantity).
Thus, rotation of the contour of integration C by n[2 corresponds
to the transition in p-space to the usual Euclidean metric, in which the
square of a length is equal to the sum of the squares of all four coordinates
with the fourth coordinate being real.
634 QUANTUM ELECTRODYNAMICS

Since
dAp — id^p = idp' dp’0 = 2n2iq3 dq = in2q2dq2, (47.6)
where dxp is an element of volume in p-space after the contour of
integration over p 0 has been rotated by n/2, p' = p, q = ] / p'2+pl,
we obtain the one-dimensional integral
OO

o
J
j f(j>2) d*p = i n 2 /(z) z d z . (47.6')

Fig. 93. Fig. 94.

In particular,
dip
w w 21 ’
dip n 2i
(47.7)
J 2(1- k * )

f Pad*p
J (p2jr l —2pky 2 ( l - k 2)
We now proceed to investigate divergent integrals of type (47.4).
Since in quantum electrodynamics no divergences higher than quadratic
occur, we shall be dealing with integrals of the type
d*p
o = J
(A--2pkAi?
p ^p
j p ( k , I) = f
(p^2pk + l f
_P,J>v dip__
• w * . o = J (p2—2pk+l)2

PMPV# P
• w * > o = / (jp-ipic+ fy
INVESTIGATION OF SCATTERING MATRIX 635

In all these integrals and in other similar divergent integrals we shall


carry out the integration over a certain finite and invariant region (N )
characterized by a number N , such that N oo corresponds to the
whole infinite four-dimensional p space (such a finite invariant region
may be defined, for example, by the inequalities |p2| ^ N 2, (ps)2/s2 ^ N 2,
where s is an arbitrary time-like four-vector).
Since integration over p 0 with p 2 kept constant does not lead to
a divergence, the contour of integration with respect to p 0 (cf. Figs. 93,
94) may be rotated, just as in the evaluation of the convergent four-dimen­
sional integrals (47.4), through n/2, i.e., it may be replaced by a vertical
straight line. After such a rotation we obtain a four-dimensional Euclid­
ean p-space with a real fourth coordinate, whose element of volume
d4p ' is related to the element of volume of the initial p-space by formula
(47.6). The region of integration in this case is a four-dimensional
sphere, whose radius we denote by L and which we identify with the
cut-off momentum introduced in § 46.
In proceeding to the evaluation of the divergent integrals of interest
to us we begin with the logarithmically divergent integral

On rotating the path of integration with respect to p0 by n/2 and on


utilizing formula (47.6) we obtain

The last integral can be evaluated in an elementary fashion and is


equal to
1
2
Therefore
(47.8)

Further, we consider the integral


636 QUANTUM ELECTRODYNAMICS

which diverges logarithmically, like the integral 7 (2)(0, /). On rewriting


this integral in the form
f dip _ f dip
J cp2- 2 p k + i y “ J 1 ( p - k y + i - W ’
we introduce new variables by means of p —k -> p. Then the integral
assumes the form (47.8) but with an altered range of integration, which
we denote by N'. In accordance with (47.8) the integral J {2)(k, I) is
equal to 7i2i(ln(L'2/l—k 2) —\). B u t!/ differs from L by a finite quantity
so that if L' is sufficiently great, In 1! differs from In I by a small
quantity of order 1/L, and we finally obtain
dip L2
- ji2i In (47.9)
I ( p * - 2p k + i y J^k2
We now evaluate the logarithmically divergent integral

JW )(k I) = f P a P r^P
°A , ) J (P2- 2 Pk + i r
Just as in evaluating the integral (47.9) we can shift the origin of coordi­
nates by introducing a new variable of integration p —k -> p :

f __ Pci
PcPrdiP__ _ J {Pc + K)(Pr+K)diP
J ((p2-
p 2- 2
: pk+ lf J {p2+ l - k 2f
(N) ( N')
On noting that

I
PcPrdiP
(.P2+ l*)
= ~ <5„ c p2d4p
1 „ r p2d*p
4 u°TJ ( p 2+ i y
_2-AlT
2J 1
d(i
\ q*+ i y -
snHL~L2 3\
d°r 4 [ln I 2}

PadiP = 0,
I (p2+iy
d*p n2i k <Jk T
X
k„k.
■J (p 2+ l f 2 I ’
we obtain finally

f PaPrd*P n 2i / 7,2 3 ^ j rfi K K


(47.10)
J (p2- 2 Pk + i y = T i ” ln l - k 2 “ 2-,
We now proceed to evaluate the integral

J (2)( k , l ) =
PadiP
J ( p 2~ 2p k + i y
INVESTIGATION OF SCATTERING MATRIX 637

On noting that
f PrPad*P dUa
(47.11)
J (p*-2pk+ lf dkz ’
where
Lt _ 1 r p*d*p
° 4 J ( p * ~ 2 k p + iy
we integrate (47.11) over k z from the point k z = 0 to the point kz:
TV
V dk_ =
4 J dK

where
A 0 = —lnL 2+ l .
The left hand side of this equation is equal to one quarter of the desired
integral, while the right hand side is equal to

Therefore finally we obtain

I = ” * ' Ml n7 (47' 12)


We note that if we had evaluated this integral by introducing a new
variable (retaining the same range of integration N) p —k -* p, we would
have obtained n 2ika(ln(L2/l — k2) —1) ; thus as a result of the transform­
ation p -> p —k, i.e., of the displacement of the origin of coordinates
by k, the divergent integral (47.12) acquires an increment nHka!2.
Finally, we investigate the quadratically divergent integral

PaPz d*P
I) =
(jP -lpk+ If ‘
On integrating formula (47.12) with respect to I we obtain

f / - i p k + i = n H k ° W-**> ln ~ ( l ~ k%)
+ (A,+ \ ) l + s ( N , **)}, (47.13)
638 QUANTUM ELECTRODYNAMICS

where e is a function of k 2 and N, containing a constant which diverges


quadratically as N -> oo, It can be easily shown that e is a linear func­
tion of k 2. Indeed, the left hand side of (47.13) acquires a factor n3
as a result of the transformation k -> nk. p -*■ np, / -> n2/, where n
is an arbitrary number, which is only possible provided e is of the form
£= I ■
/^2~\~k2(aA0-]-b), (47.13)
where A z is a quadratically divergent constant, while a and b are certain
numbers (A:2 is multiplied by the logarithmically divergent constant A 0,
and not by A z, since only under this condition will both terms in e
have the same dimensions).
On differentiating (47.13) with respect to kz, we obtain the desired
integral

/ 1 P^-TpJ+ W In( / - « - ( / - * » + , ( * + *)
+A:2(a^ 0+ 6 ) + l A z} ~ n 2ikakr [In (/—k 2) —aA0—b}. (47.14)
In order to obtain a and b and to express A %in terms of L we can
proceed as follows. We write J {2) (k , /) in the form

= K„(k, 0 + / - 7 # C T p (47.15)
where
K (k A _ . f P a P rd *P f P a P ,d iP
’ ’ J {p2- 2 p k + l f J (>2+ /)2 '
The second integral in (47.15) can be evaluated immediately:

The integral Kaz(k, I) can with the aid of

J _ _ _ J_ = _ f rtji—P)_ _ . (47.16)
an p* " J [{a- (i)z • p T 1 *

be brought into the form


1
/•£ A _ 4L. f f PaPrPfidiP _
INVESTIGATION OF SCATTERING MATRIX 639

Since

f PcPrP^ = 0
J (p2+ /)3
then by applying once again formula (47.16) we obtain

PyPyPaPr diP
d ,f [p2—2pkzt-\-l]i

The integral with respect to p appearing in this expression diverges


logarithmically, and can be evaluated in the same way as integrals
(47.12) and (47.14). By proceeding in this fashion we obtain a, b and
A 2, where a = —1, b = —a and A 2 = L2.
In conclusion we give a summary of the integrals which we have
evaluated:

d*p L2
— n2i In
/ \ p 2- 2 p k + l ) 2 l-k 2 '

PodiP L2
= n 2ik„ I In
/ (p2- 2 p k + l ) 2 l-k2 1

PaPr d*P 5k2-3 1


{p2- 2 p k + l f 6 T

+ l n ~ - ~ ' (47.17)

d4p n 2i 1
J (p2—2p k+l)3 = ~ T T ^ k * ’

PgdiP _ n2i K
I {p2- 2 p k + l f 2 l- k 2’

PaPr d*P L2 3 n 2i kakT


In
I {P2- 2 p k + i y l-k 2 2 ~ 2 l^W

47.2. Second Order Electron Self-Energy Part


We now undertake the explicit evaluation of self-energy and vertex
parts. We begin by investigating the second order electron self-energy
part E (2\p). This quantity corresponds to the diagram shown in Fig.
640 QUANTUM ELECTRODYNAMICS

56 (1). In accordance with Feynman’s rules (cf. subsection 25.5.) S (2){p)


has the form

Z m (.P) = yr S ‘( T - k ) Y f D‘(k) d‘k , (47.18)

where

= lp 2+ m 2 ’ = “ 'F f F '
The quantity 2 in the expression for Dc{k) which represents the photon
“mass” has been introduced in order to eliminate the infrared diver­
gence in the electron self-energy part, which appears as a result of
regularization (cf. § 30).
On substituting into (47.18) the expressions for S c(p) and Dc(k)
and on utilizing the formula

1 f dx
o
where
a2 = {p2+ m 2) x { \ —x ) Jr m2x2-\-X2{ \ —x'),

we rewrite (47.18) in the form

Z<2)(J)) = _ A e2_ ( C dx 2m + i ( p - k )
{P) ( 2 n y ) d k ) d [ ( k - p x ) 2+ a 2}2 •
o
Further, on introducing the change of variable k -*■ k-\-px and on
taking into account the resultant increment of the linearly divergent
integral (cf. formula (47.12)), we obtain

p i (47.19)
o
We write this expression in the form of an expansion in terms of
p —im
Z (2Kp) = (47.20)

where 27<z) and Z (02) are constants. On comparing this expression with
(43.48), (43.49) and (45.21), we see that S*2’ is the regularized value
INVESTIGATION OF SCATTERING MATRIX 641

of the electron self-energy part, Z{2> defines (in this approximation)


the electromagnetic mass and Z™ defines the constant Zl5 where
2 i2) = —idm, and Z (02) - Zp1—1.
It follows from (47.19) that
2e2 71 m (l+jc)
dx
2l (2tt)4j 4 lm + J d*k J [k*+m2x2+X2(l--x)]2
0

C p -im )T i« + Z » > (p ) = ( | ( p - im ) - ± f d ‘k
(2tc)4 14 7t ‘

i
X dx
+ a 2)2 [k*+m2x 2+X2( \ - x ) ] 2

+ ( p —/m ) - i- f d4k ( ----------- i— X---------- d x \ • (47.21)


n 2i J J [k2+ m 2x 2+X2( 1- x )]2 ,
0
For the evaluation of Z[2), Z ^ and Z)j2)(p) we utilize

I _ I = _ 2 I' _ ( ? - « _ .
a2 ^ J [/3 + ( a _ j5)z]3 «
o
and set in it a = &2+ u 2, and /? = &2+ m 2x2+ 2 2( l —;c). We then obtain
1 1
'j r J d ' k\ ( k - \-a-f [k2Jrm 2x 2-\-X2{\ —x)]2

x ( l —x){p2+ m 2)
dz
[k2+ m2x 2-f■A2(1 —x ) + zjc (1 —x)(p2+ m2)]3

dz
- i x ( \ - x ) ( p 2+ m 2) f' m2x 2+ X2( l - x ) + z x ( l - x ) ( p 2+ m 2) ’ (47-22)

On noting that
(p-f-im) [2m+ip{\—x)] l+ x
= 2m2/'
m2x 2-\-X2{ 1—x) + (1 —x)z{p2-\-m2) m2x2+A2(l —x)
2m2x(\-\-x)z j
m (l+ x )+ (/p —m )(l—x) £l
m2x2-\-X2{1—x)
+ ( p —/m)-
m2x 2+A2(1 —x ) + x (1 —x)z(p2 m2)
642 QUANTUM ELECTRODYNAMICS

and
p2+ m 2 = (p—im){p-\-im),
we obtain the expressions for Z {2)(p), 2^2), 27{2)

6
n i w * vi 2m2x ( l+ x ) z
i „ v + ) ‘ F x)
X <7z(p — im)2,
m ¥ i 2( l - j : ) + r ( l —x)z(p2+ m 2)
(47.23)

I
Z?)= S j - j - f r f(l- * ) ln frnV +4!(l-*)]*c
^ [ 0

- 2 f —
0 xr2+ - - ( l —x)
nr
. i .
Z[2) = {^ + J (l+^)Mo+ln(m2x2+A2(l—x))]Jxj>
0
where
/f0 = •—2 In L + 1 •
On assuming 2 < m we finally obtain the expression for the regu­
larized second order electron self-energy part
ia ( p — im ) 2
r<2>(p) = - - - -Xl-
} 1 A - 2, - — hip)
R VfV 2ji m 2(1 - e ) \ i-e /
ip— m T 1 / ? , 4 -- -8—
p + p 2 \ l l . A2
I n p ------ ——In — - (47.24)
+
m 2e(i —e) \ e 1— p / p p m 2

where
p 2+ m 2
P ^ <
nr
We see that the regularized value of 2^,2)(p) contains the photon
“mass” 2. This quantity did not appear in the original expression for
E {2)(p) and has arisen as a result of regularization.
INVESTIGATION OF SCATTERING MATRIX 643

In view of the “infrared divergence” this expression is inapplicable


for \p2\ < A2. For |g| > 1, i.e., \p\2 > m2 formula (47.24) yields

(47.25)

The divergent constants E {2) and E (2) have the form for A -> 0

(47.26)
From this we obtain the expression for the electromagnetic mass
of the electron,

(47.27)

Formula (47.24) has been derived on the assumption that the ar­
bitrary function d, which appears in Dcfiv(k) (cf. (17.21)) has been chosen
equal to zero. It can be easily shown that for an arbitrary gauge for
the potentials the quantity E {R2) is defined by the formula

(47.24')

where d°t = dt (0).

47.3. Second Order Photon Self-Energy Part


We now evaluate the second order photon self-energy part /7^2)(/c).
This quantity corresponds to the diagram shown in Fig. 58(7). In accord­
ance with Feynman’s rules the tensor 17$ (k) is defined by the formula

yJt ( [ p - m ) r v_{ i ( p - k ) - w) (47.28)


{p2+rnl)[{p—k f + m l]

This expression can be written as

+ i„A(k), (47.28')
644 QUANTUM ELECTRODYNAMICS

where

n«'(k) =

A(k) = - ^ n $ ( k ) .

In future we shall be interested in the transverse part of n j ^ \ which


is determined by the first term in (47.28') since only this term has a phys­
ical meaning.
First of all we evaluate L/7,(2):
O fJ-fx

i p —m j(p ~ k)-
3 n$(k) 3 (2^ 4 Sp
rp 2+ n ^ Vv [ ( p - k ) 2+ “m 2]
y> 2T d P

4m2-\-2p (p —k )
d*p.
3(27if (p2+ m 2) [(p—k)2+ m 2]

After having evaluated the trace of 4m2+2p{p—k) we rewrite \ T I {2^ in


the form
8e2 C {p-~k)2-\-m2+pk+rrfi —k2 „
m v(k)
W J ~~(p2~+m2j [ ( p - k ) 2T m 2] p

Ze* r d> p 8e» r _________ p _, _________

3(271)* J p2+ m 2 ^ 3(2tt)4 aJ (p2+ m 2) [ ( p - k ) 2+ m 2] p

8e2 1
+ 3{2j£) -(m2- k 2) f \ p 2+ m2) \(p—k)2jt-m2] dip.

The first of the integrals appearing in this expression diverges quadrat-


ically and can be written in the form

—in2m2 I n + n 2iL2.
ml

For the evaluation of the second and the third integrals we use the
formula (cf. (47.3))
1 r dx
(p2+ m 2) f ( p ~ k T - f m 2] =J Tp2~ 2 p K ; + Z F ’
INVESTIGATION OF SCATTERING MATRIX 645

where Kx = kx and lx = k 2x-\~m2. Moreover, on utilizing the values


of the integrals (47.9) and (47.12) we obtain

M v = & r ,r
1
= -rfijd x + + J,

Jf dx Jf [p*~2pKx+
[p2- :i ^ p l xY
1
= ~ n2ika j ln-^2-+ y + ln [ 1+ -^2-X(1~ ^ J-

Therefore

- * j - ( - in- ^ + 1 + ,n [ ' + £ x(i - 4 4 (47-29)

k k
We now evaluate p v TI(2){k)\
k2 ^

KK ip—m ^ i ( p ~ k ) —m
k 2 /7<,2)W = k 2 p2j\-m2 (p —fc)2+ m 2

On noting that
S p { fc (ip -m )£ [i(p -£ )-m ] = 4 (A:2m2+ k2p 2+ A:2(A:p)—2 (kp)2) >

we obtain
—k2Jr 3 k p —2(k p)2/k2
!Lh!S . n {2)(k) = -4— d*p\.
A:2 W (2tt)4 (p2+ m 2)[(p—k)2jrm2]
646 QUANTUM ELECTRODYNAMICS

The integrals appearing in this expression are equal to

/ “ 4 ~ 2- = in2((L2- ™ 2 In (L2/m2)),
1

Jf -(p2Jr
—----- - f
m ) [(p—k) ^ m2]
= —iJt2 f rfx f-In (L2/m2)
J
0
1 '
- f 1+ In (1 + (k2/m2) (1 —x) x)),
1
f ——---------------- - -— - — — i7i2k fx<7x I—In (L2/m2)
J (p2+ m 2) [ ( p - k ) 2+ m 2] aJ \ I J
0
+ | -f ln (l+ (£ 2//n2)(l —x)x)],
f ______P u P j ' P _____
J (p2+ m 2)[(p —fc)2+ m 2]
1
= — <5^ Jc /x [(ra2+/c2x(l —x)) In ( l 2/(m2+ fc2x (1 —x))j
0
-f l [5k2x2— 3 (m2-f- k2x)) —| L21
i
+ in2ktlkv f x2dx |ln[L2/(m2 A:2x ( l —x))]
o
Therefore
Ain2e2
-pnp(k) = l { L 2—m2)
lw ~

— J ( m 2—k2x Jr3k2x ( \ —x))\n(lJr (k2/m2) x ( l —x))dx . (47.29')


0 -*
On utilizing (47.29) and (47.29') we obtain the expression for I I (2){k)

n<» (*) = -- n s \ k ) - - /7«> (t)

= j ' A |] “ •') In (l - x))dx


INVESTIGATION OF SCATTERING MATRIX 647

This expression must be regularized. To achieve this we must subtract


from /7<2)(fc) the quantity l l {2)(0 ) + k 2( d l l (2)(k)/dk2\ 2=0 which is equal to
Lz 5\
(47.30)
m ~6 r
Therefore the regularized expression for n{2)(k) has the form
i
I I {2){k') = x) [l-\-(k2/m2) x ( l —x)]dx. (47.31)
o
On comparing expression (47.30) with (43.53) we obtain the second
perturbation theory approximation for Z~x (e2jef)

z - 1= 1
3n\ m2 6,
whence

Z = 1 - — (in L2 (47.30)'
hn \\ m2 6ii
On introducing'into (47.31) the new variable x = \ ( \ Jr rj) we
rewrite I J ^ i k ) in the form
i
la
n (i ){k) = — k2 J ( l - r f ) In dr\. (47.31')
-l
But

j In ^ + d v — —4(1 —0 cot 0),

J' ?]2 In £1 + 4-4 Ar (1 —Vs)j drj = — -- -(- \ (1 —0 cot 0) cot2 0,


m2 ^ " I 9 1 3

where 0 is related to k 2/m2 by the expression


k2
sin2 0 :
4m2
Therefore
Am2—2k2
(47.32)
3k2
648 QUANTUM ELECTRODYNAMICS

We give the expressions for I I (2)(k) in two limiting cases.


If | k 21<< m2, then

n {Rz)ik) = (47.33)
R v' 15n m2
If | A:21 > m2, then

n$H k)= i | k 2] ln^— 1- (47.34)


R w 2m m2
If we have the expression for I l [2)(k) we can determine the second
approximation to the photon Green’s function and the increment
SAjf) of the external potential of the electromagnetic field due to the
polarization of the electron-positron vacuum (cf. subsection 50.3).
The Fourier component of the increment of the external potential
due to vacuum polarization is determined by
d A f { q ) = Dc( q ) n ^ ( q ) A l e)(g). (47.35)
On substituting into this expression instead of IJ^2)(q) the regularized
value (47.32) we obtain

H ' 1( ? ) = f — ( ? „ ? - 8 „ « s) (1 - 8 c o t 8) - 1 ] A '« ( q ) .

(47.36)
On applying to 5A{txe){x) the operator — lH, we obtain the increment
in the external current j ^ i x ) due to vacuum polarization. The Fourier
component of this increment is equal to
d j ^ i q ) = qz <H e)(tf)
cl Um*—2a2 1T
= (47-37)

It can be easily shown that djfte)(q) satisfies the law of conservation of


charge:
q j j l , e)( q ) = 0. (47.37')
On noting that the Fourier component of the external current density
is given by j ifle)( q ) = qzA lff )(q) and that q/iA {^ )(q) = 0 we can rewrite
the expression for djjf^q) in the form

= - ^ [ 4m! 7 g2/ ( l - » c o t8 ) - | ] ? r (?)■ (47.38)


INVESTIGATION OF SCATTERING MATRIX 649

On expanding the quantity which appears in the square brackets


into a series in powers of q and on confining ourselves to fourth order
terms we obtain

dj» )(q) = T ^ q2j»e)(q)- (4739)


On going over from the Fourier components to the functions themselves
in terms of the coordinates and the time, we obtain

W '« = ^ - J j ^ □ /!" (* )• (47.40)

If in the expansion of (47.38) in powers of q we retain terms of


the sixth order, then in place of (47.40) we obtain (199)
a 1
<Uie)0 ) = 15S F c 4 “M j^(x). (47.41)
170tt m2

47.4. Third Order Vertex Part in the Case of External Electron Lines
We proceed to evaluate the third order vertex part A (f )(p1, p 2; q).
This quantity corresponds to the diagram shown in Fig. 65 (1). In
accordance with Feynman’s rules A {f )(p1, p 2‘, f ) can be written in the
form

~ -f r,S%p%- k ) y llS ‘(pl - k ) y , D ‘(k)cl>k. (47.42)

We shall see later that the vertex part A {f \ p x, p z\ q), like (p),
will contain after regularization an infrared divergence. (If p x and p 2
are momenta of a free electron, then/l{t3)(p1,p 2; q) contains an infrared
divergence before regularization.) Therefore, for the evaluation of
^ l 3)(Pi> Pzl q) we shall use, as we did in subsection 47.3, the expression
for Dc(k)
1 1
Dc(k)
i k2+A2
where X is the photon “mass”. On substituting the expressions for
Dc(k) and S c(p) into (47.42) we obtain

^ 3)(Pi, P*; q) = { y ^ i p i - n ^ y ^ i p i - m)yvj

— [Y&Pt—^ Y f y o y ' + Y v V o V f i p i —n b v W a —VrVoVpVTyJaT}’ (47-43)


650 QUANTUM ELECTRODYNAMICS

where
d*k
{k2—2pxk-\-p\Jr m2) { k 2—2p2k-\-pl+tn2) (/c2+A2)
ka d'lk
(47.43')
I (k2—2p1k + p l Jrtn2) {k2- 2 p 2k-\-p\-\-m2) (/c2+A2)
kgkr d*k
J =
f (k2- 2 p 1k-\-pl-\r m2) (k2—2p2k Jr p 2-lr m2) (k2-j-A2)
Of these three integrals only the third diverges (logarithmically) for
large | k \.
In the region of small k (for A = 0) only the first integral can diverge,
and only in the case when p 1 and p 2 are momenta of a free electron.
For this reason we can set A = 0 in the evaluation of the integrals
J a and Jdx.
We investigate first of all the case when p 1 and p 2 are momenta of
a free electron. In this case the integrals J, Ja, Jar can be evaluated
simply.
On taking into account the fact that p\ — p\ = —m2, and on util­
izing the formulas

_______ 1 _______ = f _ _
Jk2- 2 Plk) (k2 - 2 p 2k) J (k2--2pyk)2'

1 r 2x dx
w ~ 2 p y W ( .k 2+ t 2) = J
o
(k^-2Plk - f i ; r
where
Py = yPi-Vi} - y ) p,, px = xpy, lx = (1 -x )A 2,
we write J, Ja, Jdz in the form

0 0

(47.44)

t f j C i r k k Tdik
J" = j d y j ^
INVESTIGATION OF SCATTERING MATRIX 651

The integrals over k appearing in the foregoing expressions are, in accord­


ance with (47.17), equal to
r d^k n 2i 1
J 2pxk + i; y =
r k^k _ n2i
J = ^ M p ‘”
C k akTdik n 2i T L2 31 n 2i
J n ~ p f + ~ 2 ] W , p" p ” '
On carrying out the integration over x we obtain

In order to evaluate these integrals we note that


- p I = - {Pi + q ( l -y)f
= m 2-\-q2( 1 —y ) y
= m2[\ —4 y ( l —y) sin2 0]; q = p 2—Pi, (47-45)
where d is related to q by the expression
—q2 = 4m2 sin2 6. (47.45')

Further, in place of y we introduce the new variable f


tan f
2y-l
tan d '
Then
2 _ cos2 6 tan f
—P {Pla-P2a)'+^{Pla+P^)^ (47.46)
cos2 f ’ 2 tan 0

and the integral J assumes the form


652 QUANTUM ELECTRODYNAMICS

Since

( In COS ~ d £ = f £ tan £ </£,


J cos 0 J
0 0
then we finally have

■ /- - ™ l [ fta n fr f£ + 0 1 n A l (47.47)
m2sm 2 0 |J mj

The integrals Ja and J0x can be evaluated immediately after (47.46)


and (47.45') have been substituted into them:
i _nH 0 / , ^
J* - ^ ^ 2 d {Pl*+Pz°h
/, L _ l \ d(pla+ p 2a)(plt+ p 2r)
<rr\ m 4/ 2m2 sin 20

+ (5- + T ^ r ) ( 1- 0co t0 )}- (47-48>


We now return to formula (47.43) defining A (2) (pu p 2; q) and
simplify the coefficients appearing in front of the integrals J, Ja, Jgx.
We investigate first of all the expression which appears in front
of J; obviously it can be rewritten in the form
V&iPi—nbVpiiPi—m)?*= ^y^-lim [y J
~2im [yMP i+ P i?,} -2m 2y ^
Since in the evaluation of the matrix elements the quantity A (2) (px, p 2, q)
is always multiplied on the left by u2, and on the right by uu where
u2 and iq are spinor amplitudes for an electron of momenta p 2 and /q,
then in the last expression the matrix ipx appearing on the right, and the
matrix ip2 appearing on the left, can be replaced by -^m (if the matrix
iPi appears on the left instead of the matrix ip2, or the matrix ip2 appears
on the right instead of i/q, then before carrying out this replacement
we must make the substitution /q = f a —q, Pz = iq + $ ). After such
transformations the factor irt front of J assumes the form
y , 0 /»2 ~ m)yv = 4m2y^-2qy^q.
The last term on the right hand side of this expression can be replaced
by 2q2yM. Indeed, on noting that y j q = - q y ^ A - ^ q , and on multi­
plying this equation on the left by q, we obtain qy^ q = —q2yfiA-2qfiq•
INVESTIGATION OF SCATTERING MATRIX 653

But the last term gives no contribution in the evaluation of the matrix
elements since u2qux — u2(p2 — p 1)ul = 0 , and can therefore be
neglected. Thus, finally, the coefficient appearing in front of J in the
expression for A [2){px,p2\ q) can be written in the form
yr(ip 2 - m ) y fi(ip1- m ) y v = 4m2y u+2q2y ^
The factors appearing in front of the integrals Jg and Jaz can be
subjected to an analogous transformation:
y M h - m ) y lty ay t+ y ry ay lt(ip1- m ) y v =
Yr,YoYliY.Yv= ~ 2YtY^Ya-
On substituting these expressions and the values of the integrals (47.47),
(47.48) into (47.43) we obtain the expression for Ajf^px, p 2; q)

A ™ ( P x , P A q) = ^— 4-|(4m2+2^2)yi/ + |^4im(p1+^2)/ -2(A+A)y„?

J
+ 2 q y j , P i + P i ) + y ( A + P 2 ) Y M( P i + A) k i + 2 q y , t q K 2 — 4 y K ^ ,
where
n 2i 0 n 2i
k 2= (1 —0 cot 0),
m2 sin 26 "2a2

—0 cot 0 +

In this expression we can replace, as we have done previously, the


matrix ipx appearing on the right and the matrix ip2 appearing on the
left by —m. As a result of this we obtain
o
e2v I 20
A » )( P i ’ P A q ) = \~tan20 ( l n T - ! ) ^ 22 l / f t a n f r f f

0 1 1 , - ie“
£ 1 .. ,
n
(47.49)
^ t “ ^ r 4 V + 3 »

It remains now to regularize A {2) (pi,p 2; q). In order to do this we


must subtract from A ^ (p1, p 2; q) the quantity A f ' (p0, p 0; 0), where
p 0 is the momentum of a free electron. Since in the case under consid­
eration p x andp 2 are momenta of a free electron, then in order to regu-
654 QUANTUM ELECTRODYNAMICS

larize A f )(p1, p 2;q) it is sufficient to subtract from it its value at


q = 0. This value is obviously equal to

1 i L2 + In (47.50)
^ ( h . w O ) = - r ^ 4 n m2

The quantity L(3) is related to Zx by Z f1 = 1 + L <3). It can be easily


shown that
L<3) = — Zq. (47.51)

The regularized value of A<3J ( p 1, p 2; q) is of the form


u
a
A
$(p
1,
p 2
-
,q)=——
v tan 26 f
£ tan £ d£
K - 1
6 .I , -/a . a a . 26 .
- 2 Un9 + (47’52)
For |/)21 > m2, |g2| > m2

- ^ . ( ' n^ lnx + T lnI^ l - (47-53>

47.5. TVnrJ Order Vertex Part in the Case o f One External Electron Line
We now consider the third order vertex part in the case when only
one electron line represents a free state, while the second line represents
a state which may or may not be free. Insofar as the photon line is con­
cerned, we assume it to be an external line. Under these assumptions
the vertex part can be written in the form

ie2 C i ( p + k —q)—m
A ™ ( p , p + k \ k) =
(2ti)4 J 7v q2—2qp—2pk-\-xm2
i(p — q)—m diq
(47.54)
7fl q2- 2 q p 7v q2+22 ’
where
k2= 0 , p 2 = —m2, { p - f k f - f m 2 — 2pk = xm2.
In evaluating A f f p , p-\-k; k) we must keep in mind that on the side
of the free electron line, i.e., on the right hand side, the matrix A (2) (p,
p + k ; k) is multiplied by the bispinor u which satisfies (ip-fntyu = 0.
INVESTIGATION OF SCATTERING MATRIX 655

This fact enables us to utilize, for example, a relation such as pap


= 2impa+trPa, since
papu impau = im (2pa—ap) u = im (Ipa—ima) u.
Moreover, we must keep in mind that A ^ \ p , p-\-k; k) always
appears in the matrix element in the form of the product A ^ e ^ where
eMis the photon polarization vector which satisfies the condition = 0.
This enables us in the course of further calculations to utilize in the
expression for A ^ e ^ the relations
kek=0,
k e p - \ - p e k — 2(ep)k—xm2e,
imek-\-pke — xm2e—2(ep)k,
• AA AA^ ft A A ft ^ | A t A
imep—pek—m^e = - 2 m i e-\~pke.
On utilizing these relations we can write A ^ e ^ in the form

A<3)(P, P+k; k)e = {^yaeyrJ<fv+2(im h a- V ae k + p y ae)Ja


(2

—2[pke+2im(ep)]J0} ,
where
0 ,q
a,q
aqT
)
d*q
J 0 .a .a r /
(q2—2qp)(q2—2 qp~2qkJrxmz)q2‘
/ 2 _ :

and the symbol (1, qa, qaqT) denotes 1 in the case of J0, qg in the case of
Jg and qgqT in the case of Jgz.
All these integrals may be expressed in terms of the Spence function

Fix) = I l n ( l + w ) ~ , F(—1) =
6 ’
viz.:

1 j, “ lu .- '- ilf .+ k -
in x —1 X— l / X

1 1 x —2
\ A q~\~ n 3• ln|x| + 27,
in 2 at 2 1 x—1 °}
( x — 2)(«z+ 6 « - 6 ) 2x*+ 9x-\2\
lnUI +
+ k°k’ w \ n j»+— ix-\f x-1 J
656 QUANTUM ELECTRODYNAMICS

+ (* ./\+ L p ,)-^ -{< y „ + — — f l i , , |X|+ - - f - 5 j

1 I,, 3*—2 , , , , 1 1
+ P ^ 2 \ 2J<‘~ l ^ w ' t , M + l i = T i ’
where A0 is a logarithmically divergent constant.
In accordance with the general theory the expression (47.54) diverges
logarithmically for large q. In order to regularize the vertex part A {2) (p ,
p + k i k ) we must subtract from it the value of A i2) (p, p; 0) corre­
sponding to k = 0:

a ^ p .p -,0 ) = ^ . + x - i ” f )'

As a result of this we obtain the following expression for the regularized


vertex part A f y ( p , p + k ; k):
2
-im + fc -
m2(x—1)
In |x| pk-kp 3x—2
+ r « + P ^ ^ y ( k + i*m )
m2( x ~ 1)

+ - ^ - ( - 2 b r+ x m \ ) ( F ( y . - l ) - F ( W . (47.55)
On noting that
Aa a A A A
(ik p - p k ) y ^ - x m 2y ^ - 2 i m k y fi + 4kp^
we can write / ^ ( p , P + £; £) in the form

A^{p,p-\-k\k) = (m2Ayfi+ im B ky ix+imCpfi+ D k p i) , (47.56)

where

^ = - i » 4/rr. - 2 + - 2(- ^ r + - i n * - i ) - n - ■)],


In x
5 =
x -1 ’
1 ^ (3«—2)ln|x
x —1 (x —1)

Z) =
1 2 ( x - 2 ) ( 2 « - l) . . . 2 n n.
x —1 + - - - ^ 5 T - i )s— ,n l«l ^
INVESTIGATION OF SCATTERING MATRIX 657

§ 48. Functional Properties of Green’s Functions. Limits of Applicability


of Quantum Electrodynamics

48.1. Expansion Parameters of Perturbation Theory


We have seen that the perturbation theory series of quantum electro­
dynamics, which are formal expansions in terms of the parameter
e, have meaning only when we introduce the cut-off momentum L.
When L->oo the use of perturbation theory has no sense, since even
the second term in the expansion becomes infinite. It is therefore elear
that the true perturbation theory expansion parameter must depend
not only on e, but also on L. We now consider the problem of finding
this parameter.
With this aim in mind we investigate the structure of the perturbation
theory series which determines the vertex function r(p1p2;k). In § 46
we have shown that the terms of this series can be expressed in terms of
logarithmically divergent integrals, i.e., that they contain In L. Since
the vertex function is a dimensionless quantity it can depend only on
the ratio of L to some quantity having the dimensions of momentum.
Therefore, we can write the series for r(pt,p2;k) in the form
°o JJ
F(Pi,PP, k) = £ anln, where / = In—j-.
71=-0 171
and the coefficients an are functions of pf/m2, pf/ra2, fc2/m2 and e. Let
us determine the nature of the dependence of an on e. First of all, we note
that, as we have shown in § 46, irreducible diagrams of arbitrary order
correspond to quantities proportional to In L. On summing the contri­
butions of all the irreducible diagrams we obtain
OO

* 1 = 2 ? <hr(e2)r+1-
T —0

Reducible diagrams are equivalent to irreducible skeleton diagrams


whose vertices represent certain lower order vertex parts which also
depend on /. In the general case the vertex part of the (2o+l)th order
which is proportional to (e2)n, is a polynomial of the «th degree in /.
Thus,
OO

= £ anr(e2) n+r
r=0
658 QUANTUM ELECTRODYNAMICS

and
OO OO

n P, , p , - , k ) = 2 J P a ^yten y, (48.i)
t=0 n= 0
where the coefficients anr depend on p\lm2,p \jm 2 and k 2/m2.
The series (48.1) contains two parameters e2 and e2l, and the criterion
for the applicability of perturbation theory is the smallness of both
these parameters

e2< 1, e2l = e 2l n ~ < l . (48.2)


m2

The perturbation theory series for the propagators Ge(k) and Gy(k)
will obviously have a structure analogous to (48.1). If we write these
functions in the form
Ge(k) = s(k)S'(k),
Gv(k) = d(k')Dc(k),

where s(p) and d{k) are dimensionless factors, then


OO CO

m = I E s ^ y ^ iy ,
n=or=0
CO CO
(48.4)
d(k) = 2" I dnT(e2)T(e2l)n,
n= 0 t=0
where snT and drr depend only on k 2jm2.
We note that asymptotically the functions d(k) and s(k) cannot
depend on m. Therefore the cut-off momentum can appear in the expres­
sions for d and s only in the form lk = In L2/\k2\).
In this case the series for s(k) and d(k) will be of the form (48.4)
where / has been replaced by lk, and snT and dnT are constants
L2 71
s{k) = ^ s°nr(e2)T e2 In >
k2
n. r
(48.4')
L2 n
d(k) = £ d U e 2y e2 In I »
n, r

where \k2\ > m2, while d^T and s^T are constants.
The expression for the vertex function in the asymptotic case can
also be written in an analogous form.
INVESTIGATION OF SCATTERING MATRIX 659

48. 2. Zero Order Approximation in the Expansion in Powers of e2


The foregoing results ensure good applicability of perturbation
theory up to very high momenta satisfying only the condition ln(|p2|/m2)
< l / e 2.
Indeed, we can introduce the cut-off momentum L in such a way
that L2 !> |p 2| and e2 In (L2/m2) < 1 will hold. The criterion for the
applicability of perturbation theory (48.2) will then be satisfied, and
if the fundamental assumption made in subsection 46.2 is valid, we can
then utilize the method of renormalizations developed in §§ 45, 46 which
leads to results independent of L.
Therefore, in electrodynamics problems do not arise in practice
which require the applications of methods differing from perturbation
theory. Physically of considerably greater importance than the inac­
curacy of perturbation theory will be processes of interaction of electrons
and photons with electrically charged fields, for example meson fields,
which have additional strong nonelectromagnetic interactions. Never­
theless, from a fundamental point of view the problem of the existence
of exact solutions of the system of equations of quantum electrodynamics
is of considerable interest.
We consider the problem (115) of obtaining the zero order approxi­
mation for the propagators in the form of a series in terms of the one
parameter e, i.e., we shall attempt to eliminate the second condition
(48.2). The solution of this problem is equivalent to summing the series
(48.1)-(48.4) over/7 for r = 0. For large momenta |p2| > m2 this approx­
imation can be easily found by utilizing the property of the renormal-
izability of the propagators.
In order to do this we consider the renormalized photon propagator
G%(k) = Z ^ G ^ k ) . We write it in the form (48.3)
Gv(k) = dc(k)D'(k). (48.5)
Then
dc(k) = Z~'d{k). (48.5')
In accordance with the proof given in § 45 the function dc(k) does
not depend on L if it is regarded as a function of k and of the renormalized
charge e2 = Ze2. We call this property renormalizability.
The quantities Z and d(k) are functions of the cut-off momentum.
We emphasize this by providing these quantities with an index L and
660 QUANTUM ELECTRODYNAMICS

by denoting them as dL(k) and Z L. Since L may be chosen arbitrarily,


then for a fixed true electron charge e\ the following equation, which
expresses the renormalizability property, must hold:
Z - U l Lo( k ) = Z I ' d L{k), (48.6)
and weshall use this equation to solve the problem which we have
formulated (76), (29), (192).
We make the preliminary remark that as a result of (48.4') we have
dL( L ) = 1. (48.7)
We nowset k — L0. Then it follows from (48.6) and (48.7) that
Z ^ = Z ^ d L(L0). (48.8)
If L is close to L 0, so that e2(L2/L^)< 1, then we can utilize for dL
the first term of the perturbation theory series. In accordance with
(47.30') we have in this case

dL(L0) = l - - f L ( / - / 0), (48.9)


where

I= A> — ei Z ~ 'e 2C •

We emphasize that (48.9) is the first order approximation in terms


of the parameter e\ I and the zero order approximation in terms of the
parameter e\. On substituting (48.9) into (48.8) we obtain

Z z '-Z i,1 = Z ^ . i L (/-/„ ). (48.10)

From this relation follows the differential equation for Z L


d z i.\ = el
(48.10')
dl 12ril L
On integrating this equation we obtain an expression for Z L which
is valid for e\ < 1, but for arbitrary L. This expression is the desired
zero order approximation with respect to the parameter e\.
On integrating (48.10') we obtain

ZL= or Z, =
u W +c
1+
127 l2
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 661

The constant c may be obtained from the condition that when e2L-> 0
the zero order approximation with respect to e2h is given by Z L = 1.
Thus, c = 1 and
e2 [2
z‘ = ' - r ^ ln^ Zz'= 12ji2 m* (48.11)

On substituting (48.11) into (48.8) we obtain the expression1 for dL(k)


1
dLik) = (48.12)
e2 L2
1+ T 2 ^ l n l F

The expression for the renormalized function dc(k) can be easily obtained
directly from formula (48.5') on setting k — L. Since in accordance
with (48.7) dk( k ) — 1, then
dc( k ) = Z k \ (48.13)
On substituting into this equation the expression for Z (48.11), we
obtain
• 4(*)=— -*r- (48.14)
1— ‘12

We can also easily obtain the zero order approximation for the
functions Ge(p) and r ^ ( p , p , 0) for [p2|/m2 >> 1.
These functions have their simplest form in the case of the trans­
verse gauge for the function Dc^v:

Dc = Id — Dc
^ \V k2 I
As we have noted in § 46, in this case the expression for the vertex
part /1<3) contains no divergences.
This means that the series (48.1) for the vertex function contains no
terms proportional to e2l. By utilizing the method which we employed
earlier for making the transition from the approximation e2l << 1 to
the approximation <?2-< 1, it can be easily shown that Z x contains
no terms proportional to e2l, i.e.,
Zi = 1, (48.15)

1 T h i s r e s u lt is d u e t o L a n d a u , A b r i k o s o v a n d K h a la tn ik o v (115).
662 QUANTUM ELECTRODYNAMICS

Thus, the electron Green’s function and the vertex part are given
for e2 << 1 and for arbitrary L by the same expressions which correspond
to free fields. (This conclusion is valid if we take dt = 0.)
We can also obtain an expression for the electromagnetic mass
of the electron if we take e2 << 1 and L arbitrary. In accordance with
(47.27) for small e2l we have
9
dm ml.
4 12ti2
If the cut-off momentum L is changed the mass changes by the amount

9 e2
dm - -- —— - m dl
4 12n 2
or, after comparison with (48.10'),
dm
m

On integrating this expression, and on denoting by m(Z) the value


of the mass corresponding to a given value of the ratio e2/e2, we
obtain (115)
9

m(Z) = m (l)Z ~ 4, (48.16)

where m( 1) is a constant which evidently has the meaning of the mass


in the absence of interaction (Z = 1), i.e., of the “mechanical” mass.
The quantity e{k) = ec may be regarded as the effective charge
corresponding to the interaction in the domain of momenta smaller
than k, since e(k) can be chosen as the primary charge for the descrip­
tion of interactions in this domain.
Formula (48.11) which we can write in the form
C
e2(k) = (48.17)
(e2/12rt2) In (k2/m2)

shows the increase in the effective charge with increasing \k\, i.e., with
decreasing distances.
Since momenta of the order of k correspond to space-time distances
A~ 1/k, we obtain the following simple picture of the spatial distri­
bution of the electron charge. The primary charge e gives rise to po-
INVESTIGATION OF SCATTERING MATRIX 663

larization of the vacuum, i.e., to a certain spatial distribution around


the electron of charges of opposite sign which leads to a partial screening
of the electron charge. Therefore, for processes in which distances
smaller than A play no role the effective charge falls off in accordance
with formula (48.17) as A ~ 1/k increases. For A ~ 1 jm the effective

e(r)

_________I_____
r0 ft/mc z
Fig. 95.

charge assumes its minimum value ec, which we call the physical charge
of the electron determined, for example, from experiments on the
scattering by electrons of light of long wavelength (cf. § 45).
Schematically the dependence of the effective charge of the electron
on the distance z is shown in Fig. 95.

48.3. Integral Equations for the Zero Order Approximation


We have previously obtained asymptotic expressions for Gy, Ge
and r for e2 -< 1 and arbitrary L. In doing this we utilized the renor-
malizability of these expressions which was proved by methods of
ordinary perturbation theory employing an expansion in powers of
e2 \n(L2/m2) . It is therefore desirable to give a proof of these results
without assuming the renormalizability property a priori.
In § 44 we have obtained two integral equations connecting the
three functions Ge, Gy and F. As we have already noted, there exists
no third exact equation which would make this system complete. How­
ever, we can obtain an equation which is valid in the zero order ap­
proximation in terms of e2 (115). In order to obtain this equation it
is sufficient to make use of the skeleton diagram of the vertex part
shown in Fig. 83 (7).
Indeed, diagrams which are not taken into account by this equation,
for example, the diagram of Fig. 83 (2), contain intersecting photon
lines which in perturbation theory lead to terms proportional to (e2l)n
for n > 1, while the diagram of Fig. 83 (7) contains only terms ■ — e2l.
664 QUANTUM ELECTRODYNAMICS

In order to obtain the desired integral equation it is sufficient


to utilize the formula
r ^ P x i P i ; k) = y ^ + A ^ p ^ p z ; k)
and to obtain the expression for A on the basis of the skeleton diagram
of Fig. 83 (/) by making each of its vertices correspond to a vertex
function r and each of its lines to the Green’s functions Ge and Gy.
We thus obtain the equation
r^P uP tik)

x r ^ P i —q, Pi—q ; k)Ge(px—q)r„(Px,Px — q ’, q')Glp(q)diq. (48.18)


Together with the equations (44.4) we now have a complete system
of zero order equations for the propagators Ge and Gv.
It can be easily shown that the expressions for Ge and Gv obtained
in subsection 48.1 are asymptotic solutions of these equations. This
result is an independent proof of the renormalizability property of the
Green’s functions.

48.4. The Renormalization Group


In subsection 45.4 we have seen that the elements of the scattering
matrix (cf. formula (45.19)) are not changed if we go over from the
original values of the charge e and the propagators Ge(p), Gv(k),
r M(Pi,p2;k) to the renormalized values ec, Gpc(p), Gyc(k) , T ;C(p1, p2; k)
related to the original values by

We have utilized these transformations in order to remove in a unique


and physically meaningful way those divergences which are encoun­
tered in quantum electrodynamics. However, the role played by equa­
tions (48.19) is not limited to this, since they are related to more general
transformations admitted by the equations of quantum electrodynamics,
the investigation of which enables us to obtain a number of general
functional properties of the propagators1.

1 A detailed investigation of the functional properties of the propagators


is given in (30).
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 665

We start with the renormalized equations of quantum electrodynamics


containing no divergences and consider the two systems of finite quan­
tities e i , G i ( p ) , G l ( k ) , r i (p1, p 2\k) and e2, Ge2(p), Gy(k), A O i, Pz \ k).
If these quantities are related by the expressions

A = *iA ,
= 1 5
(48.19')
G\ = z3_1G\ ,

where z1, z 2, z 2 are arbitrary quantities, then on repeating the calcula­


tion carried out in subsection 45.4 we can easily show that the elements
of the scattering matrix calculated with the aid of both systems of
these quantities will be the same.
The transformations (48.19') obviously form a group which we
call the renormalization group, or the group of multiplicative renormali­
zation (191), (76), (29). We follow (29) in the following discussion.
On taking into account Ward’s identity we can without loss of
generality set zx = z2 and rewrite the group (48.19') in the form
G\ -> G \ = zxGi,
Gy -> G \ = zG\,
(48.20)
r1->r2= z?rx,
e\ -*■ 4 = z ^ e \ .

We first of all establish the functional equations of the renormali­


zation group. For the sake of simplicity we use the transverse gauge
for the electromagnetic potentials in which dt = 0, and we introduce
in place of Gy(k) and Ge(p) the quantities d(k2), a(p2), b(p2) which
depend only on the square of the momentum:

d(k2)
& (k) k2
(48.20')
a(P2)P + imbjp2)
p 2 + m 2

It is convenient to transfer the arbitrary factors which in virtue of


(48.20) appear in the functions d(k2), a(k2), b(k2) into the arguments
666 QUANTUM ELECTRODYNAMICS

of the functions by assuming that for a certain value of k2 = A2 these


functions assume the values
d(X2) = 1, a(X2) = a 0, b(X2) = b0.
We write d(k2), a(k2), 6(&2) in the form f ( k 2/X2, m2/A2, e2) . Therefore
the normalization condition for J(A:2) is of the form

d 1, = 1.
Az
It follows from (48.20) that
k?_ ___m1 I k 2 m2
zd = z'd
I 2"’ I 2 A'2 ’ A7^ ’ e

where z and z' are two values of the factor z corresponding to the values
of the charge e and e'. On assuming here that k 2 = X2 we obtain

z= z 'd \— 2 e'2 1-
u A/2 9 A'2 ’ c
Therefore

and

ev | £ , - - , <4 = e'vl ~ , £ , e-1 ■ (48.22)

Thus the quantity does not depend on the choice of X and is, as
is said, an invariant of the renormalization group. This quantity is
called the invariant charge.
For the actual charge ec the function d which we denote by dc(k2/m2,ef)
reduces to unity for k.2 = 0. Therefore, it follows from (48.22) that
, . k 2 m2
H2 ’e =

On setting k 2 = A2 in this expression we obtain

(48.23)

This relation defines the dependence of ex on the magnitude of A.


IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 667

On introducing the notation k 2/X'2 = x, m2/A'2 = y , 22/2'2 — t>


we rewrite (48.21) in the form

e2d ( x , y , e2) = e2d ( t , y , e2) d j y , —, e2d ( t , y , e2)j. (48.24)

This relation is the functional equation for the function d.


In a similar fashion we can obtain the functional equations for
the functions a and b, which determine Ge(p) (cf. (48.20')). On intro­
ducing for these functions the common notation s(p2/X2, m2/k2, e2)
we obtain the functional equation

s ( x , y , e2) s | l , y ’ e 2 d ( f > y> g2)j = e2> | y , y > e2 d ( t , y , e2)j-

(48.25)
If we regard relations (48.24) and (48.25) as equations determining
the unknown functions d, s, we can set ourselves the problem of finding
their most general solution. Such a solution contains arbitrary functions
of two arguments and therefore, generally speaking, is not of very great
interest. However, in the particularly interesting domain of high mo­
menta \k21, \p21 m2 considerable simplifications occur, since we can
omit the variable y from the set of arguments of the unknown functions.1
In this case, as we shall now show, the invariant charge is determined
by only one function of one argument.
On setting d(x, 0, e2) ~ d(x, e2) and on introducing for the in­
variant charge the notation g(x, f) = £d((x, f), we rewrite the function­
al equation (48.24) for the function d in the asymptotic region of
large momenta in the form

* ( ' .« ) • (48.26)

It can be easily shown that the most general solution of this equation is
S(x, f) = / ( « - < { ) ) . (48.26')
where f( x ) is an arbitrary function and x f ~1 is the corresponding
inverse function.
1 The a ssu m p tio n th at th e v a ria b le y can be o m itte d does not fo llo w au to­

m a tica lly fro m th e fact th at th e fu n c tio n a l e q u a tio n s (4 8 .2 4 ) and (4 8 .2 5 ) e x ist.

A b a sis for th is a ssu m p tio n can be fou n d in p e r tu r b a tio n th eory.


668 QUANTUM ELECTRODYNAMICS

Thus, for ]k2| > m2 the invariant charge is given by the expression (76)

= <48-27)

The general solution of the functional equation (48.25) for the


function s in the region of large momenta has the form

s ( ^ , e2) = \ p 2 \ > m 2, (48.28)

where A and F are arbitrary functions.


We can easily establish the physical meaning of the function d.
In order to do this we note that the potential <p(r) produced by the
charge e is related to the photon propagator by

<p0 ) = - ^ 3 - J ° y (k ’ ° ) eikr dk ■

On taking the Laplacian of <p(r) we obtain the charge density o ( r ) at


a distance r from the charge

Q(r) = —A<p = T j k*Gv(k,Q)eikrdk


(48.29)
-% 3
(2^)3 .
fd
(k,0 )
e'
krd
k.
The quantity g(r) obviously represents the charge density of the
cloud of pairs surrounding the charge e. We see that this density is
determined by the Fourier component of the function d(k).
For large values of \k2\ the function d(k) is defined by formula
(48.27). On utilizing this formula we obtain the following expression
for the charge density q(t) at distances small compared to ti/mc:

s i r ) = T2

On introducing in place of k the new variable (k/m) = k'


we rewrite Q(r) in the form
Q(r) = C j f ( k ' 2y k'r'dk', (48.30)

where C is a constant, r' = r{mj ] / / -1(e2)). This relation shows that


the form of the charge distribution at distances small compared to ti/mc
INVESTIGATION OF SCATTERING MATRIX 669

does not depend on the magnitude of the charge e which appears only in
the scale factor.
Relation (48.27) also leads to important consequences with respect
to the “bare” charge e 0.
On noting that in accordance with (48.23) e\ = e\dc(o o , ef), and
on taking into account that e2d(k2jm2, e2) is a positive increasing func­
tion of k 2jm2 we can draw the following conclusions. If as k2lm2 -> oo
the function f[{k2lm2\ f~ 1{e2')) -+ oo, then el is infinite, and the singularity
at the center of the charge distribution is stronger than a (5-function
corresponding to a finite point charge. However, if as k 2lm2 -* oo
the function f ( ( k 2/m2) f ~ 1(e2)) tends to a finite limit then e\ is equal
to this limit which does not depend on the value of e\.

48.5. Derivation o f Asymptotic Expressions for the Green’s Functions with


the Aid o f the Differential Equations of the Renormalization Group
The arbitrariness arising in the solution of the functional equations
(48.24) and (48.25) which manifests itself, in particular, in the fact
that arbitrary functions appear in (48.27) and (48.28), can be removed
if we seek solution's of equations (48.24) and (48.25) which in the limit­
ing case of sufficiently small e2 (for arbitrary values of the arguments
x,_y) reduce to the well known perturbation theory series. With this
aim in view we shall first of all obtain the differential equations for
the renormalization group which describe the infinitesimal transforma­
tions of the group (29).
On differentiating (48.24) and (48.25) with respect to x and then
on setting t = x we obtain the desired equations

e2d(x, y, e2)
e2 d ( x , y , e2) = < p l - , e 2 d ( x , y , e2) y
(48.31)

where

V(y, e2) = ( ~ l n s ( £ , y , e2)


670 QUANTUM ELECTRODYNAMICS

Thus, in order to obtain the functions d, s over the whole range


of variation of their arguments we must know the functions q> and xp
which are defined by the values of the functions d, s in the infinitesimal
neighborhood of the point x = 1.
The functions cp and xp can be obtained by using perturbation theory.
Since the quantity e2 appears in the definition of these functions in
the form of the invariant charge e2d, the method of obtaining d and s
with the aid of equations (48.31) requires that the condition e2 d 1
should hold.
We now show how the asymptotic behavior of the functions d(k2)
and s(k2) can be obtained for \k 2\ > m2 with the aid of equations (48.31).
On omitting in (48.31) the variable y we obtain

~ ^ e 2d(x, e2) = - ---- * ’~-<p(e2d(x, e2)),


X (48.32)
-^—\ns(x, e2) = —xp(e2 d(x, e2)),
Ma X

where

(48.32')

Integration of these equations yields


e‘ d {x ,e ‘ )
dz
= lnx,
I z<p(z)
(48.33)
e* d(x,e> )
s(x, e2) dz
s(x0, e2) I z<p(z)
c * d ( x 0,e * )

where e2d(x, e2) = z(x).


We obtain first of all d{x, e2).
In order to find <p(z) we use the perturbation theory expansion

d(x, e2) = 1 - - lnx 1 lnA'+r6 - 2 ln2x) + ... (48.34)


12n2 ^ 16tt2 4n 2
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 671

From this in accordance with (48.32') we obtain the expression for y{z).
3z
<p(z) + ... Z < 1. (48.34')
12 167l2

Substitution of this expansion into the first relation (48.33) yields


3e2
In* = 1 —d x— In d-\-e20 ( e 2, e2d).
\2 n 2 T in 2
whence

d^ x ’ ^ = 1- i ^ ln X + l ^ ln I 1— l | IT bl* l + ". (48-35>

On confining ourselves to the first two terms of this expansion we


obtain
k2
e2d(x, e2) = where x = —- (48.35')
m2
l ~ w ' nx

which coincides with the result (48.14) obtained previously (with ec-*e).
We note that from the method used for the derivation of formula
(48.35) it follows that \ — {e2l l l n 2) \ n x < 1 . Therefore, the logarithm
appearing in the third term of this formula is negative and, consequently,
the sign of the third term agrees with the sign of the second term. This
circumstance is not accidental, and is a manifestation of the general
situation which we shall discuss in greater detail in subsection 48.6.
As has been pointed out already, the method of integration of
equations (48.32) is based on the assumption that e2d << 1 since only
under this condition is the expansion (48.34') valid. Therefore, formula
(48.35') holds when the following inequality is satisfied:
e2 \k2\
1
Thr2 ° Tn2 > e2,
3n
whence it follows that \k2\ << m 2e a . This inequality is at the same
time the condition for the applicability of the series (48.35).
We see that in comparison with the usual perturbation theory
series , 2

A x ) - 2 f nme ^ n+m) lnBx, * = —5-.


t> m
672 QUANTUM ELECTRODYNAMICS

the series (48.35) has been rearranged and has the structure
fix ) = £ e2nf n{e2 In*),
Tl
where f n are functions of e2 \nx.
We can also obtain similar improved expansions of the functions
a ( x , y , e 2), b ( x , y , e 2) which determine Ge{x = p 2jX2, y = m2/X2).
In order to do this we must have the expansions of the functions a,
b by means of ordinary perturbation theory. They can be obtained
if we substitute into the expression Ge = S C+ S CE S C the expansion
(47.24) or (47.24') in place of 27. For dl = 0 we obtain

where c{y) is a function of y. On utilizing these expressions, and also


the relation (48.33) and the definition (48.32') of the function ip(z),
we can obtain the asymptotic formulas valid for large momenta (a* j),

In a(x, e2) = e2) - l } + ...,


(48.37)
In b(x, e2) = ^ lnt/^^A', e2) + ...,

where d{x, e2) is defined by the expansion (48.35).


With the aid of equations (48.31) we can also obtain the behavior
of the electron Green’s function in the region p2^ —m2. In this region
the function G2(p) has a singularity which is called the infrared singu­
larity (the existence of this singularity can be seen, in particular, from
formula (47.24') for 27(2) at *-►()).
In order to investigate the infrared singularity of the function
s we use the second of equations (48.31) and set y = —1 in it. Integra­
tion of this equation yields

six, — \ , e 2)
j(*o. — e2)
(48.38)
where a = —(p2/m2).
IN V E S T IG A T IO N O F S C A T T E R IN G M A T R IX 673

We determine the quantity which appears in figure brackets in the


integrand in the region „y ~ x 0 ~ 1. By utilizing the usual perturbation
theory expansion (48.36) we obtain in the case of the transverse gauge
(d, = 0)

3c2 x'
h Xns , x ' ~ l . (48.38')
8n x' 1
On substituting this expression into (48.38) we obtain the principal
part of the integral in the region x ~ 1
s(x, - 1 , e2) 3e2 , x — .
In In
5(a:0, —1, e2) 8n x 0— 1 ’
whence
3e2
s ( x , —1, e2) ^ (x—l) 8n (48.39)
i.e..
8 71
s I— ez I m (p2jrm 2) p 2 ^ — m 2.
m-
In the general case of an arbitrary gauge we would have obtained,

for E (2) as can be easily seen by evaluating I-f— In j ( f , —— e2d


[o£ \ x _
with the aid of formula (47.24'), an expression which differs from
(48.38') by the replacement of 3 by 3-t/?. Therefore, for d° ^ 0 the
infrared singularity of s has the form (1)

e- ph ( p 2-j-m 2) where -m2. (48.40)


mz
It can be shown (77) that the Green’s function must have a singularity
of this type not only for an electron, but for any charged particle of ar­
bitrary spin.

48.6. The Problem o f Closure o f Quantum Electrodynamics


As we have repeatedly noted, the theory which contains a cut-off
momentum L is not a closed one. The possibility of its practical ap­
plication was based on the assumption that we can choose L in such
a way that the region of momenta \k\ > L does not play an important
role. But only the limit L -» oo has a rigorous meaning.
674 QUANTUM ELECTRODYNAMICS

The perturbation theory method based on the smallness of the


parameter e2 In (L2/m2) is obviously unsuitable in principle for the
investigation of such a limiting transition. We have previously (cf.
subsection 48.2) obtained expressions valid for small e2 and arbitrary L.
For sufficiently large L2lk2 formula (48.17) assumes the form

^ « = T „ 7 Z ^ T (48'41>

In this case the renormalized charge ec = Z1/2e turns out not to depend
on the initial charge
e2 _ 2n
(48.42)
47i Tn (L 2~/m2)

Formula (48.42) is very instructive. The cut-off momentum L which


appears in the theory as an auxiliary parameter turns out for In (L2/m2)
> 1/e2 to be uniquely related to the physical charge ec. If we take for
e\ its actual value e2/4jr = 1/137 then we obtain L = we(3:i/2>137.
Large values of L lead to e2c/47i < 1/137, and in the limit L -> oo ec -> 0.
Thus, although the solutions obtained previously provide a formal
possibility for a limiting transition to the exact theory (L -* oo), this
transition cannot be actually carried out as it leads to a physically
incorrect result — the vanishing of the electron charge.
We shall apply the method described in subsections 48.2 and 48.3
only in the case of small initial charge (e2 << 1). Therefore, for the
time being we can only conclude from the foregoing that in a consis­
tent theory the primary charge e cannot be small, i.e., that at very
small distances a strong interaction must occur.
However, arguments can be given which show that the relation
(48.42) must be valid for arbitrary values of e2 (122).
In order to do this we consider the system of equations (44.4) which
define the Green’s functions. We note that in the first of equations
(44.4) defining the electron Green’s functions Ge the charge e2 and the
photon Green’s functions Gv appear only in the form of the product e2Gv.
The vertex function F cannot be defined by means of an exact integral
equation, but it can be characterized by the set of irreducible skeleton
diagrams which correspond to expressions containing e2 and Gy also
in the form of the product e2Dv (cf. (45.20')). Therefore, we can say
INVESTIGATION OF SCATTERING MATRIX 675

that the quantity e2d(k) — ie2k2Gy(k) is defined by an equation of the


type (43.43')

(^-2 + P 1(k)]j e2d(k) = 1, (48.43)

where Pt (k) is an operator which is independent of the parameter e2.


Formula (48.41) gives a solution of this equation for large values
of L2/k 2 and small e2. This solution corresponds to neglecting the term
1/e2 in (48.43) as a result of which the equation no longer contains e2.
Therefore, it is natural to suppose that for large e2 we can neglect this
term with even greater justification. In actual fact the situation is more
complicated, since the operator Px in (48.43), being an integral operator,
includes the region of integration over k close to L where the approxi­
mation (48.41) is not valid. However, this difficulty is associated only
with the sharpness of cut-off in momentum space. This could be avoided
by introducing other types of cut-off. We could investigate, for example
(154), a method of cut-off in which photon momenta are confined to
a considerably smaller region of integration than electron momenta.
In this case expression (48.41) remains valid over the whole range
of variation of k and is a solution of equation (48.43) valid for arbitrary
values of e2 (with the exception of very small ones).
A pictorial interpretation of the results (48.41) and (48.42) consists
of saying that at small distances a strong polarization of the vacuum
occurs. If the initial charge extends over distances of the order of 1/L,
then at large distances 1jk (for In (L2/k2) > 1) the effective charge
e(k) defined by formula (48.41),

= -71- e*d(k) = 3rc/ln (Lr/k2),


4jc 4n

will be small for arbitrarily large values of the primary charge e. However,
if the initial charge is a point charge (L-»oo), then it turns out to be
completely screened at any arbitrary finite ■distance.
We, therefore, conclude that quantum electrodynamics is not a closed
theory. It cannot in principle take into account interactions at distances
smaller than 1/L, where L is in any case greater than the value given
by formula (48.42). Therefore, we should expect that the value of the
cut-off momentum must have a deep physical meaning, expressing
676 QUANTUM ELECTRODYNAMICS

a significant change in the properties of space-time or in the prop­


erties of interactions at small distances.
As we have mentioned previously, in the case of electrodynamics
this question is not of great importance, since these distances are
so small that in the case of phenomena associated with even much larger
distances we must take into account phenomena which do not formally
enter into the content of quantum electrodynamics (for example,
meson processes).
The lack of closure of the formal structure of quantum electrodynam­
ics is of greater importance for theories modelled on quantum electro­
dynamics for the investigation of strong (meson-nucleon) interactions.
In such theories a relation of type (48.42) must also hold (155), and
for a large value of the renormalized charge e\ ~ 1 this yields L ~ m,
where m is a quantity of the order of magnitude of the masses of the
particles participating in the interactions (mesons, nucleons). Such
a value of L corresponds to distances of the order of 10~14-10 13 cm.
Of course, the method of investigation based on introducing a cut-off
in momentum space cannot have a rigorous foundation, since it is,
in fact, based on a nonlocal interaction of the type

\xp(x)k{y)xp(z)hfx—y) d f x —z) dx dy dz,

where <5X and <).z are functions of effective width of the order of 1jL
which go over into ^-functions for L->oo,
But it is not possible to formulate a noncontradictory system of
equations with such an interaction for finite L. Therefore, it cannot
be proved that solution of these equations followed by a subsequent
transition to the limit is equivalent to a solution of the equations involving
a local interaction.

§ 49. Generalized Green’s Functions


49.1. Green's Functions in the Presence of External Fields
So far we have been investigating Green’s functions in the absence
of external electromagnetic fields. But, as we know, in a number of
problems it is convenient to utilize this auxiliary concept. Therefore,
we shall now generalize the results obtained previously on Green’s
functions to the case when external fields are present. For this purpose
INVESTIGATION OF SCATTERING MATRIX 677

it is convenient to consider the field due to external currents J(x). The


external currents appear in the Lagrangian in the form of an additional
term L1( 7 ) = —J/i(x)A/i(x), where A ( a -) is the operator for the electro­
magnetic field. In the diagrams this term obviously corresponds to
vertices from which only one photon line emerges.

0
i
i
Fig. 96. Fig. 97.

The simplest diagram of this type is shown in Fig. 96. It represents


emission of a photon by external currents. The matrix element for
this process has the form

S^f = f J j x ) e - ikxd4x.
) / 2 oj J

Figure 97 shows diagrams for pair creation by external currents.


The matrix element for this process has in the first approximation the
form
S ^ f = —ie j uy/ive~i(Pl+p*)xJy(y) ^ v(x ~ T) dix d iy .

We note that this matrix element can be represented in the form

= —e j A lj f)(x)uy/ive~Hpikpi)x d4x,
where
Ale)(x) = i j J J y ) D ^ J x - y ) d*y.
The quantity Afte)(x) may be regarded as the potential due to the
external currents J (x). Indeed, since i Dc(x—y) is the Green’s function
for the wave equation the above expression for A (e) is a solution of the
equation l A {*] — —JM-
678 QUANTUM ELECTRODYNAMICS

In the general case the diagrams will contain parts joined to other
parts by only one photon line (cf., for example, the second diagram of
Fig. 97). Such a part is equivalent to a normal vertex which represents
the effect of the external field A (e),
A l' \ x ) = i f G\iv( x - y ) J v(y) f r y . (49.1)

Thus, the addition of the interaction is equivalent to the addition


of the interaction L[,
L'i = - j ^ A ^ i x ) ,
where ] (x) is the current operator, and A (Me)(x) is the external field.
All the quantities in which we are interested — the Green’s function
and the matrix elements—are obviously functionals of J(x) or of A ie)(x).
States which are stationary in the absence of external currents naturally
cease being stationary when J{x) ^ 0. However, we can consider the
set of states 0 n(J) which become stationary when J(x) = 0. In this
case the matrix element (r|S|r) = (0 r(J), S (/)^ r(/)) differs from unity
when 7 ^ 0 . In particular, this also applies to the “vacuum state”
(we retain this name in the presence of external currents only for
convenience).
In practice this is of no importance, since in actual problems we
encounter only weak fields constant in time.
We define the electron propagator in the presence of external fields
as
r t . v,OOs)jo)
(49.2)
(0jS]0) ~

■=—+—r
Fig. 98.

By introducing the denominator (0|S|0) which differs from unity in


the presence of an external field we exclude consideration of vacuum
loops. We can also use the former graphical definition of Ge(x, y) as
an effective electron line.
The equation for Ge augmented by a term which takes the external
field into account is shown graphically in Fig. 98, where the cross
INVESTIGATION OF SCATTERING MATRIX 679

denotes the vertex involving the external field. The corresponding


integral equation has the form

Ge(x, y) = S c(x—y) —i j S% x-x')M (x', y')Ge(y', y)dix' d*y'

- e J S c( x - £ ) A f £ ) G e(£ ,y ) d ^ . (49.3)
We define the photon propagator as

_(0|fiA ,W ^,(nS)|0)
GM
vfx, y) (0|S|0)
-A^(x)A^(y). (49.4)

It satisfies the same equation (43.43) as in the absence of external currents.


The difference consists only of the fact that Gy(x, y) and P(x, y), like
Ge(x, y) and M(x, y), are functions of the two variables x, y and not
of their difference x —y.
On applying to equation (49.3) the operator zp+m, and to the
equation (49.4) the operator Li, where pM= —id /d x, we can write
the equations for Ge and Gv in the following form:

(z'p+m + M —ieAie))Ge(x, y) — —i d ( x —y),


(49.5)
(\Z+P)Gy(x, y) = —i d ( x —y),
analogous to equations (43.42) and (43.43).

49.2. Green's Function for Two Electrons. Equation for Bound States
of the Electron-Positron System
We have earlier discussed electron and photon Green’s functions.
We can also define more complicated Green’s functions as vacuum
expectation values of different products of Heisenberg field operators.
Of the greatest importance is the Green’s function for two electrons,
since it enables us to obtain the approximate wave equation for two
particles, and, in particular, the equation for the bound states of the
electron-positron system.
We define the Green’s function for two electrons as

(o IT(yjfx,) ^ (x2) i p f x f (x4) S) Io)


GapytiXM X2XA) (49.6)
(01s 10)
Graphically the function G(~ 1is represented by a set of diagrams termi­
nating in four electron lines.
680 QUANTUM ELECTRODYNAMICS

We divide diagrams defining G'- "’ into compact ones (Fig. 99)
and noncompact ones (Fig. 100), calling those diagrams compact
which cannot be divided into parts interconnected by only two electron
lines.
It is also useful to introduce the self-energy part for two electrons
as the pait of the diagram connected to the other parts by four electron
lines. We denote the set of such compact self-energy parts by 7(12; 34)
where the indices denote both the coordinates and the spin variables.

\ x
\ /
X
✓ x
/ V

1 2

Fig. 99. Fig. 100

With the aid of this quantity we can construct the integral equation
for the Green’s function G{ K In order to do this we utilize the skeleton
diagram of Fig. 101 which shows the relation between G{ * and I.
In order to be able to describe further the bound states of the electron-
positron system we shall here use notation symmetric with respect
to the electron and the positron, viz., we shall denote by t p ( x ) and
y(x) the electron annihilation and creation operators, and by i p ' ( x )
and ip'(x) the positron annihilation and creation operators. These opera­
tors are interrelated by the equations

v>'(*) = c v(x), v'(x) — c M .y ),


where C is the charge conjugation matrix.
On denoting the Green’s function for the electron-positron system
by G~+ we have in accordance with (49.6)

Caffcyd(Xl X2> Yl) (0 1s | oi (ol vv?(*2) vv(-v3) s} | o)


1
CofsTo)- (° Ir {vh(*i)v^<>‘2)s} | o) (0 i S 10) (0\T{vv(x3)y'd(xt)S}\0).
(49.7)
INVESTIGATION OF SCATTERING MATRIX 681

The second term arises because of the absence of exchange diagrams


for noninteracting electrons and positrons. On the basis of Fig. 101 we
can write the integral equation (168)
G“ +(12; 34) = G-(13)G+(24)

+ G -(ir)G + (2 2 ')/(l'2 '; 1"2")G "+(1 "2 "; 34), (49.8)


where integration and summation over the variables corresponding to
repeated indices are implied, and G~ and G'{ denote Green’s functions
for the electron and the positron (in the general case in the presence of
an external field).

We can easily obtain the form of the kernel /( 12;34) for an electron
and a positron in the first approximation (proportional to e2), which
corresponds to the diagrams of Fig. 99 (1):

JaV.ydiXl X2>XZXi)
= -°C(*l-* 2 )< K * l-* 3)
+ e?(yllC)ap(C~1yl)syDc(x1—x J d ( x 3- x J d ( x l - x J . (49.9)

The matrix C appears in this expression for the reason that we are
describing the positron not by the operators y(x) and y>(x), but by the
operators y>'(x) and y/(x). The opposite signs of the terms in (49.9)
correspond to the opposite signs in the terms of (37.4); finally, the
(5-functions take into account the fact that in the diagram of Fig. 102(<?)
the point x 1 appears to coincide with the point a 3, while x> appears
to coincide with x4, while in the diagram of Fig. 102(6) the point x t
coincides with x 4, and a 2 with x 3.
682 QUANTUM ELECTRODYNAMICS

On utilizing the properties of the matrix C we can rewrite (49.8)


in the form
IaP;yd(Xl ’X2’ * 3 * 4)
= 5(x1- x3)<5(x2- x4)

+ e \ y C \ p{C-i y X D c{xl - x ,)d{x 1- x , ) b { x , - x i). (49.10)

On applying to equation (49.10) the operator

(/p(1) —zeAfo) + M (1) + m) (zp<2) + ieA(x2) + M +<2) + m),

(i) —
PV-

where M is the mass operator defined in § 43, and on taking into account
the fact that the single particle Green’s functions satisfy equation (49.5),
we obtain
KG + (xxx2; x3x4) = —6 (at—x3) b (x2—x4) ,
where
K = (zp(1) —zcA(at) + M (1) + m) (zp(2) + ieA(x2) + M+(2) + m) + /
(49.11)
and
IG~+(x1x2; x3x4) = J I (x4x2; y ^ G - ^ i y ^ - , x 2x^)diy 1uiy 2.
We see that the function G~+ is the Green’s function for the equation
K95(x1x2) = 0, which is the equation for the wave function (p(x1 x2)
which describes the state of an interacting electron and positron.
The kernel I in the absence of an external field can depend only
on the difference of coordinates. On introducing the new variables
x = x4—x2, x' = x3—x.,, X — -^(xj+Xj) and X' = -.l(x3+ x 4), we can
say that in this case I will be a function of x, x' and X —X'\
1 = 7(x, x'; X —X').
In the absence of an external field we seek the wave function <?(xt x2)
in the form 9?(x4x2) = <y(x)eiKX. On going over to the momentum
representation we obtain, in the approximation proportional to e2,

(49.12)
= Inhh
INVESTIGATION OF SCATTERING MATRIX 683

where

P i—PF-^K, p 2— ~ p-\- — K,

(p(p) = 2 J (p(x)e~ipxdix,

/ (2) (p, p ' \ K ) = J /<2>0, Y)e~i(^ ' I' ^ d ix d*x' d*Y

~~ (2jr)1 (p—p'ylZfQ ^ .■< ^A>v^2"j ‘

If we go over to the coordinate system in which the center of mass


of the particles is at rest, i.e., if we set K = (0,0,0, iK0), we obtain an
equation for the determination of the energy eigenvalues K0 of the
system of two interacting particles. In this case, if K0 is smaller than
2m, we shall obviously obtain a bound state of the electron-positron
system (in any coordinate system this condition has the form K 2 > (2m)2).
In the nonrelativistic approximation (up to terms of order v2/c2)
we can obtain from this expression the Schrodinger equation for the
electron-positron system which was used in Chapter VI.

49.3. Equations for Green's Functions in Terms of Variational Derivatives


The foregoing equations for the Green’s functions Ge and Gv define
these quantities if the polarization and the mass operators P = ill*,
M = i£* are known. Equations (44.2) and (44.3) express these operators
in terms of the Green’s functions Ge, Gy and the vertex function T.
As we have noted previously (cf. subsection 44.2), it is impos­
sible to construct an integral equation which would in turn express
.Tin terms of Ge and Gy, and which together with the first two equations
would give us a complete system of equations determining the Green’s
functions.
We shall now show that it is possible to find a relation connecting
the vertex function F with the electron Green’s function Ge. However,
this relation has the form not of an integral, but of a functional (varia­
tional) equation.
With this aim in view we investigate the change in the mass operator
M resulting from a variation of the external field A (e>. Let an infin-
684 QUANTUM ELECTRODYNAMICS

itesimal term bA(e) be added to the function A (e>. The corresponding


addition to the mass operator M is shown in Fig. 103. We see that
the diagram of Fig. 103 represents a certain vertex part A, and therefore

SZ*(x,y) = - e j A p( x , y ; ( ) 6A<-»(()d*S,
i.e.,
SZ*(x,y)
- e A J x , y ; £) = (49.13)
</>(£)

Fig. 103.

On utilizing (43.40) we obtain

r ^ x , y \ f) = H G ' i ^ y ) ) 1 (49.13')

where r ^ x , y; £) = A (x,y ; £)-\-yttb(x—y)b(y—f) and (Ge( x ,y ))-1 can


be obtained from the relation

j (Ge(x, JO) 1G(y, z)dxy = b(x—z).

On substituting this expression for r into (44.2), (44.3), and on noting


that in accordance with (49.1)

6 A ^ ( x ) = i I G*y( x , y ) d J r(y)d*y,

we obtain

f z * ( x , z ) G - ( z , y ) d > z ■=
(49.14)
i J F l* ( x , z)Gv(z, y)diz = eSpy/(- — — ^.
INVESTIGATION OF SCATTERING MATRIX 685

Thus, equations (49.5) assume the form


tP m —ieA{e)(x)—ey/j A l Ge(x,y) = —id(x—y).
a

[
(49.15)
8Ge(x, x)
\ J G v ( x , y ) = - i d ( x —y ) - e S p y ^
dJ»(y)
This system of variational equations for the Green’s function was
obtained by Schwinger (177).
Equations (49.15) can also be obtained in a formal manner without
resorting to graphical methods. In order to do this we evaluate the
variation with respect to J(x) of the expression for Ge(x, y). On utilizing
the definition (49.2) and the explicit expression for the scattering matrix
S in the presence of external currents
iJ j(i) + ■l(z))A(i)dix
S = T{e }
we obtain

= fo 7 A o ) ( oi r k ( * ) ^ ) A , M ) s | o )

_.(oihv.Mv,OOS)|o) (oir(^w s)io)


1 (o | s fo) ' (o ; s ; o) ' c ;
On defining the external field

(o| r(A (x)S)|o)


A(e)(x\ __
W (0 | S | 0)
we can write the second term in the right hand side of this equation
in the form iGe(x, y)AjJe)(x). The first term, which is equal to

i(o I T(4*a( x ) ^ ( y ) A ,/*)) | o ),


can be transformed by means of the relation

(49.17)
To prove this we utilize the fact that T{y)a(x)ip^{y)) represents a function
which is continuous for x0 > To and for x0 < y0, but which has a discon-
686 Q U A N T U M E L E C T R O D Y N A M IC S

tinuity where it jumps from 4»/jG0 4*a(x) at *o = To- 0 t0 ^p(>’)


at x 0 = To+0. The value of this discontinuity is equal to

+,»(y) (*)++»(*)+,» 00 = y A x - y ) 8 ap-


The derivative at the point of discontinuity is equal to the magnitude
of the discontinuity multiplied by <5(x0 — y0). Thus, we have

/p (r ( ^ ( * ) ^ 0 ') ) ) = i T { p ^a(x)i\tg(y))-id(x-y) dap.

Further, by utilizing the Dirac equation for the operator <]>(*)

(ip + m —iek (x)) (x) = 0,


we obtain equation (49.17).
On substituting (49.17) into (49.16) we obtain equation (49.15).
In a similar manner we can easily obtain the variational equation
for the photon Green’s function. In order to do this we must utilize

□ A „ (x )= - ie [*K*)»y t- O) •
from which it follows that n ^ e) = - J /J(x)-FieSpy/xGe(x, x). By varying
the last equation with respect to J(x) we obtain second equation (49.15).
If in equations (49.15) we replace the variation with respect to J(x)
by variation with respect to T (e)(;c) in accordance with (49.1), we obtain
the system of equations

(fp + m—ieA{e)(x))Ge(x, y)

- i e j Gy(x, Ge(z, y) d4z = - i d ( x - y ) , (49.17')

OGy(x, y ) = - i d ( x - y ) - i e S p J d4zGy(z, y ) y fi- ~ ^ ~ d4z.

"49.4. Expressions for Green’s Functions in Terms of Functional Integrals


Green’s functions can also be represented in the form of certain
closed expressions containing functional (continuous-dimensional)
integration (63), (74), (28), (54), (132). Although such a representation
has not in practice led to any new results, nevertheless, its study presents
a definite interest.
INVESTIGATION OF SCATTERING MATRIX 687

First of all we construct the expression for the vacuum expectation


value of the operator S = T{a), where

a = } IA (49.18)
and T is the symbol for the chronological product.
We also write T in the form T = Tv TA, where the operators Tv
and TA correspond to time ordering of the operators of the electromag­
netic field A and the electron-positron field ip.
The chronological product of operators of the electromagnetic
field can be expressed in terms of the normal product. In accordance
with (24.17) we have TA(a) ~ N^e^o), where NA is the symbol for the
normal product of the operators A and

A = '2 SAj^jSA^rf)
To simplify further calculations we formally replace integration over
four-dimensional space by summation over a discrete number of points.
On introducing the change of notation
j'(x)}/dix = j r ; A(x)}/dix = A x; Dc(^ —7j)\/di ^d‘ir] = D \^
we obtain

1 n^ c ______
32
A - (49.19)
2 dA^dA r\

T M = N \ e Ae zKl j
where it is understood that a summation is taken over repeated indices.
Now the operator a, which in accordance with (49.19) is a functional
of A(x) and j(x), is written in the form of a function of an infinite number
of variables A x and j x.
We evaluate the expectation value <7^(cr)> of the operator TA(o)
for the vacuum state of the electromagnetic field. Since the vacuum
expectation value of any normal product of operators is equal to zero,
we have
1 c d»
2 ^ dAt^A
<TA(a)} = e * n e xAx =0’
where Ax may be regarded as c-numbers.
688 QUANTUM ELECTRODYNAMICS

We evaluate the Fourier transform of the function a(jI5 Ax) with


respect to the variables Ax

o(A) = f or(j, A)e~lAt A; f j


J x
then we have

^ y J J C

The quadratic form appearing in the exponent can be transformed


to its principal axes. This changes the infinite-fold integral into a product
of one-dimensional integrals, and allows us to carry out the integration.
In this manner we obtain

<Ta (o)> = c f o(h A ) e ~ ^ D%' AiAn f [ ~ , (49.20)


J I
where
. 1 nc . 2
C = I J J e~* c cdA:

and D\ = D£c is the matrix D\n brought to diagonal form. This expression
contains the operators ip. Further, on averaging over the vacuum state
of the electron-positron field, we obtain

r — 1 A .A
(i3 e ) t 1 J—T dA
(01S10) = C J S \ A ) e 2 ^ c (49.21)
^ X

where

S* (A) = (Tv(<r))
and the brackets denote averaging over the vacuum of the electron-
positron field.
The expression for the Green’s function can also be brought into
a form analogous to (49.21). On carrying out the same transformations
as before we obtain
i Id% ' aaak dA,
(0\S\0)G ^(x,y) = J e • \ T,, {v,Xx)Wn(y)°(A, j ) f f l
2 jz

(49.22)
INVESTIGATION OF SCATTERING MATRIX 689

We see that the problem reduces to the determination oh the Green’s


function for an electron in the external field A
s °(A)Gap(x,y - A) — \ T j y j a(x)y}fi(y)a(j, A)}}. (49.23)
It follows from (49.23) that G(x,y; A) satisfies the equation
( i p - f m - ! 'd ( x ) ) G(x, y; A) — —i d(x—y).
If G ( x , y \ A) and 5°(/4) are known, then Ge( x , y ) can be expressed
in terms of these functions as

J G(x,y; A)e S°(A) f ] dAx


X
Ge(x,y ) (49.24)
f e x p ( - i( D % A eAjS°(A) f ] dA,
X

i.e., the Green’s function Ge(x, y) is obtained from the Green’s function
G(x, y; A) in the given external field A by averaging over all possible
fields A with a weighting function given by exp (—£ D^l]AtArJ)S°(A).
By utilizing a similar method the photon Green’s function can be
written in the form

/ 2 tV * nS ° ( A ) n d A z
GlXx,y) = — c----- ;-------------- ~ ------------- • (49-25)
/ e x p (-|C D c)f *,4f z lJ S 0(/i) [ ] dAx
X

We now obtain the expression for S°(A). We introduce the notation


A

X(x) = ey(x)A(x),
X(x) \/d*x = %x; y W V dix = % •

Then S0^ ) = (.Tv{e~/x'l'x)')• We expand this expression into a series in


powers of %xy x, and in each term of the series express the T-product
in terms of TV-products in accordance with the formulas of § 24. On
taking into account the fact that the expectation values of TV-products
vanish, we obtain

s ^ a ) = % — ^ d -(k}’ (49-26)
n
where Dn{K) is a determinant of the nth order constructed of the
quantities Kx.x. = y>xlXXj-
690 QUANTUM ELECTRODYNAMICS

The series (49.26) can be summed and yields


S°(A)=D(K), (49.27)
where
D { K ) = ||<30- K I4,.||
is an infinite determinant (the Fredholm denominator).
Expression (49.27) can also be written in a different form. We assume
that the matrix of the determinant D(K) is brought to diagonal form
in such a way that Kap = &upKa. Then .S0(/l) = f ] (1 —Ka) and
u
In S°(A) = ^ In (1 —Ka) = Sp In (1 —K ) . (49.28)
a
Since the last expression contains only the trace of the matrix, it is
invariant with respect to the choice of the representation.
The matrix l n ( l —K) can be written in the form
i
in(i - / o =- * /( ! --m
0
or, by utilizing the definition of K,

il nn( l - / f ) = —e C AdX . (49.29)


I 1 A

We compare the integrand in (49.29) with the Green’s function


Ge(x, y; A) for the Dirac equation for an electron in an external field A.
In matrix form such an equation has the form

( ( 5 T 1- ^ ) c ? ( ^ ) = 1,
whence
( ( ^ r ' - e i ) ' 1 = G(A). (49.30)
We see that
i
ln ( l—Ai) = - e A f G(XA)dX.
0
In the usual notation the last formula has the form
i
S p ln (l- K ) = - S p J dX J G(x ,x;X A )A (x)d ix, (49.31)
0
where the symbol Sp refers only to the spinor variables.
INVESTIGATION OF SCATTERING MATRIX 691

Finally we obtain
i
5t'(^) = e x p ( - e f dXG{x,x;XA)A(x)dix). (49.32)
o
Thus, the foregoing method reduces the general problem of quantum
electrodynamics to finding the Green’s function for the Dirac equation
in an arbitrary external field, and to subsequent averaging by means of
functional integration.
On solving equation (49.30) by expansion into a series in powers
of eA and on substituting these solutions into (49.32) and correspond­
ing integrals over A, we can develop a perturbation theory equivalent
to the usual perturbation theory of quantum electrodynamics.
It is obvious that in the zero order approximation Ge coincides with
S c and S° = 1. In this case we can easily obtain from (49.25) that Gy = Dc.
In the general case the factor S%4) in (49.25) takes into account the
polarization of the vacuum, while the factor exp (—-|(Z)c) j ^ fy4 ) corre­
sponds to the free electromagnetic field. These two factors correspond
to the two terms P(k) and k2 in the equation for Gv.
We note that in accordance with (49.30) and (49.32) S°(A) is a function
only of the product eA: S°(A) = F(eA). Therefore, on choosing for the
variables of integration in (49.25) in place of Ax the variables Bx = eAx,
we can write the expression for the function Gv in the form

/ Bti(x)Bv{y)e~TADC)^ B^ F(B) [] dBz


e'Glv{ x ~ y ) = -x - — • <49-33)
j e 2«* D ^ B*BV dB,
Z

We see that e2 appears in the right-hand side of (49.33) only in the


term corresponding to the free field, whose role becomes less important
as e2 increases. Thus, we obtain a result arrived at in a different manner
in subsection 48.5: the function e2Gv does not depend on e2 and therefore
must be of the form (48.41).
CHAPTER VIII

Radiation Corrections to Electromagnetic Processes

§ 50. Effective Potential Energy of the Electron. Radiation Corrections


to the Electron Magnetic Moment and to Coulomb’s Law

50.1. Energy o f Interaction of the Electron with the Electromagnetic


Field Taking into Account Corrections of Order a
We now proceed to study specific physical effects associated with
the higher order approximations of perturbation theory. Such effects
have the general name of radiation corrections.
We start by determining the radiation corrections to the energy
of interaction between an electron and an external electromagnetic
field.
The interaction between an electron and an external electromagnetic
field is determined in accordance with the Dirac equation by the quantity
U = e y ^ A f f x ) , where A f }(x) is the potential of the external field and e
is the physical charge of the electron (in this Chapter e and m will denote
the physical charge and the physical mass of the electron).
The matrix element of this quantity determines the scattering of
the electron by the field A f f x ) in the first approximation of perturba­
tion theory.
We wish to find the radiation corrections to the quantity U. For this
it is obviously necessary to replace in the expression for U the matrix y/t
by F/t and A \ f ( x ) by A \f( x ) + hA)f{x), where P , = y/t + A m is
the vertex part and S A f f x ) is the increment in the external potential
due to the interaction between the electromagnetic field and the
electron-positron field. The Fourier component of the first of these
quantities is given in the second approximation of perturbation theory
(in the case of external electron lines) by formula (47.52), while the
Fourier component of the second quantity is given by formula (47.36).
On neglecting the product A ^ S A f f x ) , and on utilizing these formulas,
we obtain the following expression for the Fourier component of the
[693]
694 QUANTUM ELECTRODYNAMICS

radiation correction to the quantity U in the second approximation


of perturbation theory:
5U(q) = e ( A ^ ){q)A{'lie){q)+yfl5A^ >(?)) = Fv{q) A {ve) (q),
where A ™ (q )= A™(plt p2 ! <l) and

Fv{q) = ™ 1 in - _ i
'(2ti)2 V, tan 20 /\ A
v
— -0 tan 6n -------2—:
2 tan 26 I £ tan £ d£ + - m- (? r - r , ? ) sT^ F

1 \ Arn^—2o2 1
-y.-i M - W > v/l[—- 3Tq2
T - (1 - 0 cot 0) - -n- (50.1)

If we write U(q)Jr 6U(q) in the form


U(q) + dU(q) = ey^A(/le)(q)-ir ieyAd(p(q) ,
where
eS<p(q) = ^ y4dU{q), (50. n

then the latter quantity can be regarded as the Fourier component


of the increment of the potential energy of the electron due to the in­
teraction between the electron and the vacuum. We call this increment.
the effective potential energy which determines the interaction of the
electron with the vacuum.
If we know the Fourier component ed<p(q) we can obtain the effective
potential energy which determines the interaction of the electron with
the vacuum as a function of the coordinates and the time:

e6(p(x) = ( 2 P F M ) A ^ ( q ) e ^ d * q , p = y 4. (50.2)

On substituting into this expression

Ale)( q ) = jA\,e)(y)e~imd!iy,
we obtain
= -p— j $ F V{q) A ^ ( y ) d^qd*y

= y j ^ F^ x ~ y ) Ale)(y) dy,
RADIATION CORRECTIONS 695

where

= J i l y ' f FM ) e'9Xdiq-

We see that the effective potential energy of interaction of the electron


with the vacuum is determined by an integral operator acting on the
external field.
On noting that qreiQX = (lli)(d/dxv)eiQX, we can rewrite expression
(50.2) in the form

ei<p<.x) = F, (i T^ r]> ’(«) W i = (4 ^ w ■


(50.3)
We assume that the external field Aj,e)(x) is a slowly varying function
of the coordinates and the time. In such a case the function Fv(q) can
be approximated by a power series in q\ the operator Fv(i ~i-d/dx) will
in this case be a differential operator acting on A (M e)(x).
On confining ourselves to only the linear and the quadratic terms
in the expansion for Fv(q) we obtain in accordance with (50.1)

(qyv- y vq)Y (50.4)


T y-+ i

Therefore, the effective potential energy of the electron will be given


in the third approximation of perturbation theory by

\ 4 1 m
In - T — 5— X

’ (50.5)
2 im
where
p p
p (vA— ________ a(«)
t *AX) ~ dxMAv dxv Afi ■
On introducing the electron spin matrices

r _ 1 v _ 1 r _ 1
696 QUANTUM ELECTRODYNAMICS

we can rewrite ed<p(x) in the form

edcp(x) = -

, e3 4 /, m 3 1 \ r_. , cr.
+ W w K - 8 - 5) Lj (50'6)
where H are the electric and magnetic fields.
The expression for the effective potential energy ed<p(x) contains
the photon “mass” X. This is associated with the fact that formula (50.2)
was obtained by means of perturbation theory, which, as we have
already pointed out in § 30, is, generally speaking, inapplicable to the
description of the interaction of an electron with long wavelength
photons. Therefore, formula (50.6) takes into account, strictly speaking,
only the interaction of the electron with short wavelength photons,
while if its interaction with long wavelength photons is required, it
must be evaluated separately.
The individual terms in expression (50.6) admit a simple physical
interpietation. Let us first of all consider the second term (207).
We assume for the sake of simplicity that the external field is an
electrostatic field. Due to the existence of zero point oscillations of
the field an electron experiences forces which result in its additional
displacement. This displacement, which we denote by dr, is associated
with fluctuations in the potential energy of the electron, given by
e<p(r+dr) —e<p(r) = e((5rV+^(c>rV) 2 9 ?(r).
On neglecting in this expression all powers of dr higher than the second,
and on averaging over all possible values of dr, we obtain
<■e<p(r+dr)-e<p(ry)Q= \e{dr)\A(p{r),
where (<5r)o is the average value of the square of the electron displace­
ment due to the zero point oscillations of the field.
In order to make an estimate of the order of magnitude of (<5i-)§
we start with the equation of motion for the electron, taking into account
the effect of the zero-point oscillations of the field

dr = — E0,
RADIATION CORRECTIONS 697

where E0 is the electric field due to the zero-point oscillations. On


expanding dr and E0 into monochromatic waves we obtain

\dr I = i2 m2o)i El,


urai I2

where drmand Emare the Fourier components of the quantities dr and E0.
The component Em is determined by the relation I El = a>/2 (the
normalization volume is assumed to be equal to unity), and therefore

On multiplying \drm\2 by the number of modes of oscillation in the


frequency interval dco, equal to (2/(2n)3)47ico2dco, and on integrating
over co, we obtain (^r)2:
e2 r da)
(<M02 =
27i2m2J co

This expression diverges logarithmically both at high and at low


frequencies. The divergence at high frequencies is associated with the
general divergence in quantum electrodynamics in the region of large
momenta of virtual particles; while the divergence at low frequencies
is a manifestation of the previously mentioned infrared divergence,
associated with the incorrect treatment of the interaction of the elec­
tron with long wavelength photons. We eliminate this divergence
by ascribing to the photon a “ mass” X.
We take the upper limit in the “regularized” value of the integral
to be equal in order of magnitude to m; thus, in order of magnitude
(dr)^ is equal to

MS = 2 n 2m2 \
In , - + C
X
(50.7)

where C is a numerical constant of order of magnitude unity, which,


naturally, cannot be determined in an elementary manner.
By utilizing this value of (^r)^, we obtain the expression for the
mean fluctuation of the electron potential energy

(e<p(r+dr)-ecp(r))0 = - |ln - - + c j z l cp(r). (50.7')


698 QUANTUM ELECTRODYNAMICS

On comparing this expression with formula (50.6) for ed<p(x) we see


that it coincides in order of magnitude with the second term in ebcp(x)
in the case of a purely electrostatic field.
We can easily obtain a relativistic generalization of formula (50.7')
for the case of an arbitrary electromagnetic field (which, however,
varies sufficiently slowly). In order to do this it is., obviously, necessary
to replace in the right hand side of (50.7'), firstly, the Laplacian by
the d’Alembertian operator and, secondly, the scalar potential <p by
(p—aA. We thus obtain the order of magnitude of the second term
in expression (50.6) for the effective potential energy of the electron.
Consequently, this term can be interpreted as the fluctuation of the
potentials associated with the displacement of the electron under the
action of the zero-point oscillations of the field.

50.2. Radiation Corrections to the Electron Magnetic Moment


We now investigate the first term in (50.6).
We assume that the electron is situated in a constant magnetic
field. Then the effective potential energy of the electron, taking into
account its interaction with the vacuum, will contain only the single
term
e3
ebw{x) = -----—T r -BUH.

On the other hand, the operator for the electron energy in a magnetic
field is of the form
Um = - r f Z H ,

where // is the magnetic moment of the electron (in the nonrelativistic


approximation the matrix reduces to unity). A comparison of these
formulas shows that as a result of its interaction with the vacuum the
electron acquires an additional magnetic moment which in the third
approximation of perturbation theory is equal to

t*' = (50.8)

where = etiflmc is the Bohr magnetron, and a = e2/47ific is the


fine structure constant. This additional term may be called the “ anom­
alous” magnetic moment in contrast to the “normal” moment ju0.
RADIATION CORRECTIONS 699

Thus, up to terms of order a2 the magnetic moment of the electron


is given by (this result is due to J. Schwinger (174))
a
A* — A*o ( 1 ^ (50.8')

In the presence of an electric field E the first term in (50.6) has the
form

n ' m H - i P a E ) = - — 7iy ^ F Mv, (50.9)

where / / = ^ 0— , = - jr (yfiYv—y„y^)and F ^ is the electromagnetic

field tensor. This expression agrees with the corresponding term in


the phenomenological Dirac equation (15.12) for a particle with an
anomalous magnetic moment p . The Hamiltonian for the electron
will in this case, up to terms of order v2/cz, contain terms proportional
to / / in (15.13).
The experimentally observed value of the “anomalous” magnetic
moment of the electron differs from the theoretical value (ajFjx)^
given by formula (50.8). It is therefore of interest to evaluate the mag­
netic moment of the electron taking into account corrections of order a2,
which correspond to fifth order perturbation theory effects. Such
effects correspond to the diagrams shown in Fig. 104.
For the evaluation of the “anomalous” magnetic moment of the
electron we can utilize the fact that the part of the effective potential
energy (50.6) containing the “anomalous” magnetic moment p', is of
the form (50.9) A part of the effective potential energy of the electron
will have the same structure also in higher order approximations of
perturbation theory.
The energy (50.9) corresponds to the part of the element of the
scattering matrix which has the form
M = —/ i ' u 2A ^ ( q ) y iivqvul , (50.9')
where
A lMe) (
q)= f Ale) (x) e~iQI d*x, q = p 2—p x

and and u2 are the spinor amplitudes of the initial and final states
of the electron of momenta and p2. Therefore, for the determination
700 QUANTUM ELECTRODYNAMICS

of the “anomalous” magnetic moment of the electron in fifth order


perturbation theory approximation it is sufficient to pick out from
the matrix elements corresponding to the diagrams of Fig. 104 those
parts which have the structure of (50.9').

\A<%) \A(?\q)
I I

(6) (7) (8) (9)


Fig. 104.

It can be easily seen that contributions to the increment in the


magnetic moment come only from diagrams 1 ,2 , 3 ,4 , 5, while dia­
grams 6, 7, 8, 9 determine corrections caused by polarization of the
RADIATION CORRECTIONS 701

vacuum to the first and third order vertex parts, and make no contri­
bution to the magnetic moment.
In accordance with the rules of subsection 25.5 the matrix elements
corresponding to diagrams 1 ,2 , 3 ,4 , 5 are given by

Mx e^ u f y i < e) (q)
(2ti)8 • J yM(Pi- k ) * + m t ^ J(Pl- k ) 2+ m 2
1 . d4k

2e3 i (p 2 —k ) —m
A/o = 8 "2 I y/I f „ _ zA2_L yyi2 ^ <2) (^2 k)
(In ) 11 (P2—k)2Jr m2
A

i(p2—k ) —m _ af4/c
x -/— —-T u n+—m2
(p 2—«)2 :)2+ m l2 yhu ~k T2 T ui ’
2 ^ <e> fa ) (p l —A»\o-r

M. - W U f /!<»(» n
~ {I :,)' -] " 'Pt,Pl ’ < p,-k)- h»l! 4»>(<l)
■"
i ( p L—k ) —m d4k
X 7------------- 2 - y „ -7 T « i»
(Pi—fc)2+ m 2 k2
(50.10)
;e2

i(P\ k ) —m d4k
Tpi ~ W + m * Yli k *+Xi Ul
i(P t-k)-m i ( p 2 —k —k ' ) —m -(e
^5 (2ti)8 " 2 (p2-A:)2+ m 2 ^ (p2-A:-A:')2+ m 2
i( P i—k —k ' ) —m i i ^ —k ' y - r n d4k d 4k'
* ( p ^ k - ' k y + m * Y l ( p ( - k if + m t y i 1 k* k ' 2 l ’
where the quantities 77£2)(A:), Z i2)(p) and A ^ ( p ,p ) A ( p ,p ,p p )
are defined by formulas (47.31), (47.24), (47.52).
We do not reproduce here any detailed calculations, but give only
the final expression for the correction to the magnetic moment of the
electron of order a2 (188), (102):

/1 (50.11)
P
702 QUANTUM ELECTRODYNAMICS

Thus, the magnetic moment of the electron taking into account ra­
diation corrections of order a2 is equal to

p = | l + - ^ - 0 .3 2 8 ^ J - j ^ 0= 1.0011596^0- (50. IT)

50.3. Radiation Corrections to Coulomb's Law


We now obtain the radiation corrections to the electric field pro­
duced by a point charge e at rest.
In accordance with (49.1) the general relation connecting the given
current Jft(x) with the field A^{x) produced by it, has the form

(x—x')Jv(x’)d*x\ (50.12)
where Gyv(x) is the photon Green’s function.
In the case of interest to us of a point charge at rest J (x) = ied^dfr) .
Therefore, the scalar potential cp(r) produced by the point charge e
situated at the point r — 0 is given by equation
OO

(p{r) = ie J Gv(x)dt. (50.13)


— OO

With the aid of the representation of the function Gy(x) in mo­


mentum space

Gy(x) = — j Gy(k) e>kIdik = J Gy(k, k 0)enkr- k°t) d*k

we can rewrite (50.13) in the form

<p(r) = — y j Gy( k , 0)elkrd k . (50.13')

On utilizing the general relation (43.37) defining the function Gv(k),


and formula (47.31) we obtain the following expression for the photon
Green’s function taking into account radiation corrections of order a:

G*(k) = jy(k) | l + ~ J * ( ! - * ) In [ l + - J - *(■-*)]<&}>

where
1 1
U'-(k)
/ k 2—i0
RADIATION CORRECTIONS 703

On substituting this expression into (50.13') we obtain the potential


of a point charge taking into account corrections of order a:
i
e C eikr I e2 r \ k2 1 1
?(r) = (2^)3J ( 1+ 2 ^ J +
(50.14)
On noting that
/’ 2n2i
J Dc(k,0)eikrd k = ---- —

and
1 1
1 f 2 1 2X 3 * k2
fx (l-x )ln \ + — x{\-x)\d x = - ^ V2— (1 —2x)dx
k2 , m2
0 0 1+ —T ^ (l—*)
m1
1
> 2p - y ® 2
= -V dv,
&m2 J k2
0 \ + -T—s ( \ —v2)
Am2
where v = 2 x—\, we rewrite <p(r) in the form

1 v 2\ \ oikr
^ = - h +cb I Am2 d k -
k 2+
\-v2
Further, on utilizing the relation
2mr
/yihr
A2+ (4m2/l —v2) r

and on introducing instead of v the new variable t = l / ] / l —v2 we


obtain finally the following general formula for <p(r) taking into account
corrections of order a (173):

1/ 2
<p(r) = —f _ { l - f _ f ! _ f g -2mrC / 1 _|_ 1l \\ (C2- l )
(50.15)
Anr | 6n 2 J \ 2£2 /
704 QUANTUM ELECTRODYNAMICS

The integral appearing in this expression can be evaluated in two


limiting cases, when mr < 1 and when mr > 1 :

— — —y — In mr, m r-< 1,
/ i \ ( « _ n i/2 6
e -2mrC I I _VL_ d t =
l1 + 2f £2 ^ 3 j/ tt e" ■2m r

2 , mr > 1
8 (mr)3'
(50.16)

where y is Euler’s constant. Therefore, the potential of the charge at


small and at large distances (compared to h/mc) is of the form

a
mr < 1 ,
4 nr 2>n - T ~ 2 r H n W
9>(r) = (50.17)
a o — 2 m r

mr >> 1 .
4?rr ' 4 ^ 1/2 ('mr)3/z f ’

We see that at large (compared to ti/mc) distances from the charge


the potential differs from the Coulomb potential by an exponentially
small term, while at small distances the deviation from Coulomb’s
law varies logarithmically with the distance.
Formula (50.17) may be interpreted in the following manner. If
we rewrite it in the form

= (50.17')

then we can say that the potential <p(r) is determined by Coulomb’s


law using an effective charge e(r). This charge coincides with the phys­
ical charge e at large distances r > (h/inc). But at small distances
r << (tijmc) it is equal to

e W ~ e [ 1+ 7 > w ] ' (50.18)

This expression agrees with the expression obtained in Chapter VII for
the effective charge for L ~ 1Jr.
RADIATION CORRECTIONS 705

§ 51. R a d ia tio n C o r r e c tio n s to E le c tro n S c a t t e r i n g

51.1. Electron Scattering by the Coulomb Field o f a Nucleus in the Second


Born Approximation
We now proceed to determine the radiation corrections to different
electron scattering processes. We begin with the radiation corrections
to the scattering of an electron by the Coulomb field of a nucleus,
which we regard as a perturbation.

Fig. 105.

The diagrams representing the process of scattering by an external


field in the first order and the radiation corrections to it in the third
order of perturbation theory are shown in Fig. 105. The scattering
matrix elements corresponding to these diagrams are given by
= ed2A <e>(q) u,, S™f = M <3>+ M <3>,
;<?3 i(p2—k ) —m * , . i(p, —k)—m d*k \
M<3>
{ I n f Ul\ } y»(p0- k f + m2 (<?) (p~-Jcf+ m * T 2' ^ 1’
(51.1)
ie*
A/<23>- -y
(27lY

J
rJ i(p+ J)-m ip —m
”,2 W.2 y* d*p ‘1>
\ ( p + q)2+ m2 1‘ p2+ m
where u1 and u2 are the spinor amplitudes of the initial and final states
of the electron of momenta p x and p 2. A (e)(q) is the Fourier component
of the external potential and q = p 2—p x-
706 QUANTUM ELECTRODYNAMICS

In the case of the Coulomb field of a nucleus of charge —Ze (in


Gaussian units)
A ^ ( q ) = 2jid(£l- e . 2 > ~ = 27iA<*'(q)d(£l- £.2) (51.1')

and
A (e) (q) = iyi — 2it d (£l —e2) ,

where £l and e2 are the values of the electron energy before and after
scattering. Therefore,
Ze2
SfVr = — ^r 2m(u2y l u1)d(£2—£l). (51.2)
9
Further, the sum of Mx(3) and M^3) is obviously the matrix element
(between the states wx and u^) of the effective potential energy of the
electron ed<p(x) which takes into account its interaction with the vacuum.
On utilizing formula (50.1') for the Fourier component, and on sub­
stituting (51.1') into it, we obtain the following expression for the sum
of M[3) and M^3):
e2 Ze
S?Jf = ~ i &7i q 2 u2y A 4 (1 —2 0 coth20) 11 + In
0
tanh.0 + 2 coth20 | utanhudu
0

+ 4(1 —0 coth0) 1 — coth20 j

4 i , 20_
£2), (51.3)
"9 m ^ sinh20
where 0 is related to the change in the momentum q and to the scattering
angle § by the expressions

q2 = \ q 2 = 4p2 sin2- - = 4m2 sinh20 , (51.3')

\Pi\ \Pz\ = | />| -

In order to obtain the scattering cross section taking radiation correc­


tions into account it is, however, insufficient to obtain the square of the
RADIATION CORRECTIONS 707

absolute value of + S ^ f and to sum it in accordance with the rules


of § 26 over the different initial and final states. Indeed, radiation correc­
tions are third order perturbation theory effects. Therefore, along
with them we must also take into account scattering in the second
approximation of perturbation theory (the second Born approximation).
We note that it is not necessary to take scattering into account
in the third Born approximation, since the matrix element corresponding
to this process is proportional to e6.

We investigate in greater detail scattering in the second Born approx­


imation (42). The diagram representing this process is shown in Fig. 106
The matrix element which corresponds to it is equal to
ip—m
A {e)( p * - p ) A {e)(p]. - p ) d ip \ u l . (51.4)
p2-\-m2
It can be easily shown that the integral appearing in this expression
diverges for a purely Coulomb field, so that we carry out the calculations
assuming the Coulomb field of the nucleus to be screened:

As we shall show later, we can make the limiting transition rj -*■0 in


the expression for the scattering cross section.
In the case of a screened Coulomb potential the quantity A {e)(q)
has the form
A {e){q) = ^ o ) - (51-4')

On substituting this expression into (51.4) we obtain


ip —m 7i
= 4/Zza2“2 -d*p ux^(«i-£2),
n + (P2~ P f
Jj V2 p2jr m 2 if-'rip—Pi)
708 Q U A N T U M E L E C T R O D Y N A M IC S

where p x = ie and e = ex = e2. On introducing the notation

f ____________ ds __ = j

f ____________ sds _ _ ^1 + ^ 2 j
J [ri2+ ( p 2~sY}[rj2+ ( s - p i ) 2}(pl—s2) 2 2
and on noting that
iyplUl = (yi e—m)u1,
u2iYP2 = u2(y4e—m),.
we write S (2) in the form
S™f = 4/Z2azM2{w(/l —/ 2) + y4e(/1+ / 2)} £2). (51.6)

This formula determines the correction to the scattering amplitude


in the second Born approximation. The effective cross section and the
polarization calculated taking this correction into account were given
in § 14.
In order to evaluate the integrals Ix and / 2 which appear in (51.6)
we first of all evaluate the integral
ds
L
f [(S^P)2+/12] (/72—52"+70)

s2 sin %
= 2ji (51.7)
IPs cos^ + P2+ r i2) (p2—j 2+ /0)

On introducing the variable t = cos%, and on noting that the integrand


is not altered by the substitution s -» —s, t -> —t we rewrite L in the
form
1 00

s2 ds
L = 71 dt
2Pst + P2+ A 2) (/72- 5 2+T0) ‘
-1

Further, on completing the contour of integration with respect to


i by a semicircle in the upper half-plane, and on applying the residue
theorem we obtain
7C2i p — P + iA
L= --ln (51.7')
p + P + iA
RADIATION CORRECTIONS 709

As we shall immediately see, the integrals and / 2 are related by


simple expressions to the derivatives of L with respect to A and Pr
ds
I [(s—P)2jr A 2]2 (p2—s2-\-i0)
dL
2A dA
71*
A ( p 2- P 2+ A 2+2ipA)
sTds
I [{s -P)2+ A 2\ (p2- s 2+ /0) =
1 dL
2 dPT
Pr dL
2A dA
1 / p —P-\-iA
= - 7 l 2P,
~ A ( p + P + i A j J - P + i A ) + ~2P* n p + P + i A

' ' Y (51.8)


2P2 \ p - P + i A + "p + P + i A /
Indeed, we use the formula

1__ 1 dz
ab 2 J jfl( l + z ) / 2 + 6 ( l - z ) / 2 j 2

and in it set a = rj2-\-{s—p 2)2, and b = rf-r{s—p ^ 2. Since

L \ ± i +bL A = [(j-p )> + ^ « r,

where
P = $ [(l+ z )/> i+ (l—z)/>d.
§
A 2 = r]2Jt p 2 sin2 —(1 —z2),

then we have
1 1 r dz
[ { s - p , ) 2+ri2} [ ( s - p tf + 7 ] 2] = "2_J [{s ^ P T + A 2}2 '
On comparing this expression with the definitions of Jx and / 2 and on
utilizing formula (51.18) we obtain
710 QUANTUM ELECTRODYNAMICS

On carrying out the integration with respect to z we finally obtain

71“
h = —~
\ p \ siny | / v4+ 4p21r f + p 2 sin2y j

r)\p\ s in y
x arctg
~ y j ?j4+ 4 p2lrf-\-p2 sin2—

j / v 4+ 4Pz (v2+P2 sin2y j + 202 sin2"2"


+ y ln
“J^/^4+ 4p2|^ 2+ p 2 sin2 —2p2 sin2y

. 0'
ps i n- -
arctg--------
_ T V2+ 2P2 OT2
12 — h - ‘ $ In -2———- —/■ 71 . (51-9)
$ 2pn~irj
2p2 cos2 — 2p3 cos2— sin-

When rj -» 0

0
/7T2 2| p 1sin—
A= — In
0 V
2 |p |3 sin2
(51.10)

A ITT /. 17} 171


/,=
,0 # l n 2| p| 0‘
cos2— 2 |p |3 cos2— 2 sin —
2

We note that for rj -> 0

.7 1
Rc/i -- 0, R e/2 1 (51.10')
0 . 0
4 1/713 cos2 sin—
RADIATION CORRECTIONS 711

51. 2. Differential Cross Section for the Scattering of an Electron


by the Coulomb Field of a Nucleus taking into Account Radiation
Corrections o f Order a
We now determine the differential cross section for purely elastic
scattering of an electron averaged over the orientations of its spin in
the initial and final states. In accordance with the general rules of
§ 26 this cross section is equal to

(51.11)

where J is the electron flux density, which is equal to their velocity


(the normalizing volume is assumed to be taken equal to unity); do
is the element of solid angle into which the electron is scattered; q{
is the density of the final states of the electron per unit energy and per
unit solid angle
Pz\z d \ p 2\ \ p 2\e2

Qf ~ (2n f de2 {Inf

and M6(e1—e2) is the sum of the matrix elements for the basic process,
the second Born approximation and the radiation corrections:
S % + Sf% + S^Jf = Md(e1—e2).
On utilizing (51.1), (51.3), (51.6), we can write this quantity in the form
M d f a —e2) = WoQwi<5(fi—£2) > (51.12)
where
Za
Q = (A o+ A fy^-B y^f-C , A 0 = 8tt2/

(1—0 coth 0) I 1—y coth2 0

<
t>

- 4 / Z 2a2£(/1+ / 2),
4yrZa2 0
mq2 sinh 20
C = —4/Z 2a2w (/1- / 2).
712 QUANTUM ELECTRODYNAMICS

Summation over spin orientations can be carried out with the aid
of the formula

£ lMl2== 4ele2
47TSpi Q V P i - n O Q O h - m ) }
M i. M i

1
Sp {[(A0+ A 1)yi + B y i q + C ] ( i p i — m )
4e1e2
X [(A* + A*)yt -\- B*yxq+ C*] (ipi-m)}.
On neglecting the terms |A ^ 2, \B\2, B*C, A*C, A*B and on utilizing
the formulas of § 26 we obtain

V \ M \ 2— - — {\A0+A,\2{m2- p [ p 2) + i m ( A * B - A 0B*)

+2me(A*C+A0C*)}, p [ = ( - Pl, iej.


On substituting into this expression
\A0+ A ^ = \ A J ‘ +2Rc(AZA,)

■z ' “° _ L » ^ j [ (1_ # c o t h # ) ( l _ * co.h>«) - | ]


= 64?r4-

+ £(1 —2 0 coth 20) 11+ In ^ tanh 0

+ 2 coth 2 0 J u tanh u ^wj | —64ti:2~ - e R e ( / j + / 2) ,

A * C + A qC* = 2Re (A* C) = ------- — m Re(1,-1,),

64rr3Z 2a3 0
A*B—A0B* = 2i Im (A* B) =
q^m sinh 2 0 ’
. d
m2—p'1p 2 = 2e2 1—v2 sin ,
we finally obtain

H \M \2
Hi, Hi

n w zw i „ . a
1—v sin2-- 2(1 —0 coth 0) I 1 — coth20
1 71
RADIATION CORRECTIONS 713

2 I ?
-- + 0 tanh 0 + 2 ( 1 - 2 0 coth 2 0 ) 1 + In
y \ m

0
v2 sin2 -
2 20
+ 4 coth 20 «tanh«<7«
. J> sinh 20
1 —it sin2—
2

. ft I . 0
7 iX a v sin — 1 —sin —
2\ 2
+ (51.12')
7}
1 —7; sin2 —
2
Therefore the differential cross section for purely elastic scattering
is given by

do „= I ----— — —\ (1 —t;2) 11 —v2sin2 —


2mv2 sin2— ' 2

X 11 — 2(1 — 0 coth 0) ( 1 — coth2 0 I — -2 + 0 tanh 0


71

+ 2(1 — 2 0 coth 2 0 ) ( 1 + In —

0
v2 sin2
2 20
+ 4 coth 20 j wtanhw du~
. 0 sinh 2 0
0 1 —v2 sin2 —
2

. o i o\ i o y 1
-\-7iaZv sin —I 1—sin y I I 1—v2 sin2 —I do. (51.13)

We see that the cross section doe contains the photon “mass” A; insofar
as the screening constant rj is concerned, it does not appear in dae,
as has been stated previously.
714 QUANTUM ELECTRODYNAMICS

We note that the integral appearing in dae can be written in the form
<K 1
1 \-v2 . , _
wtanh u du = -*- ————■sinh 2 0
C —
CIn (1- v 2? ) d t _
4 . J 2
0 sin - cos» ( l -«*{*) l / C
' 2 —cos2" 2
2 2

4> , 1
(51.14)
2 1 —
Indeed, we consider the integral

(51.15)
--1

where
Pz = \ ( } + z ) P i + W - z)Pz-
On introducing the new variable u
tanh u
z
tanh 0 ’

we rewrite R in the form

m cosh 0 \
R m2 cosh 2 0 1 ln ’ cosh
' u *^W'

After integration by parts we obtain

j u tanhudu = 0 ln cosh 0 — j lncoshu^w.

and therefore
<p
j u tanh udu sinh 20- R —In m .
0

Finally, on utilizing the formula


1
1 C dz _ 2 0

2 J —Pz m 2 cosh2 0
RADIATION CORRECTIONS 715

and on introducing into (52.15) in place of z the new variable £

c = \P.\ _ 1P . \
\Pi\
1
z-
0 " j / C2— cos2
sin

(ez is the fourth component of PJ, we obtain formula (51.14).

51. 3. Elimination of the Photon “Mass” from the Scattering Cross


Section
We now show how the photon “mass” can be eliminated from the
scattering cross section. In order to do this we must, as has been already
explained in § 30, together with the purely elastic scattering also include
in our investigation scattering accompanied by the emission of a soft
photon whose energy is considerably lower than the electron energy.
The cross section for the emission of such a photon is determined
by formula (30.25):
Za ---- I i i I 1 —n* "Jin*
da' =
& 2 / 71
2 mv2 sin2y

1—v2
X 2 (2 0 coth 2 0 —l)ln-^^- + — In cosh20-G(v,’&) do.
A v 1 —v "T T
v sin —
2
(51.16)
where

1 +at dC
v(G, '&)— In
1 —vC
(1 —v*C2) 1 / £2 —cos2
'l/{
The photon mass appears in da' in the form

— (2 0 coth 2 0 - l ) l n —f - ,
71 A

while in the elastic scattering cross section it appears in the form


o 7
— (2 0 coth 2 0 -l)ln — •
7i m
716 QUANTUM ELECTRODYNAMICS

Therefore, if we add the cross sections dae and dd' then the photon
“mass” A will not appear in the total scattering cross section, which
is the only one that has physical meaning.
We write this total cross section for scattering accompanied by an
energy loss not exceeding Ae in the form

/ ------------d ‘( l - ^ ) ( l - s ! sin *-|](l+ii,-^)rfo, (51.17)


I 2 mv2 sin2 — I

where the quantities SR and <5fl take into account radiation corrections
and scattering in the second Born approximation; in accordance with
(51.13) they are equal to

< 5 * =71- 2(1 - 2 0 coth20) 11 + In ) + 0 tanh 0

1 1
+ 2 (1 - 0 c o t h W i - y coth2 <Z>\ 1

• , 0—
v 22 sin-
20
+ 2 0 coth 2 0 In-—*—- +
1 —v i
1 - 22 ^ sinh 2 0
1 —v 2i sin -
2

(1 —w2)cosh 0 r T ln(l+wO _ ln (l—vQ 1 dC


v sin-
# J v&IL 1—vL,
COS —
1 + w ; J
C2 —cos2
&
2
1 2
(51.18)
d
(5R= navZ sin 2 “11— sin — 1 —v2 sin2

(in the expression for dR we have used formula (51.14)).


We emphasize that these formulas can be used only for sufficiently
small values of dR and dB. In particular, the nucleus must be sufficiently
light since the second Born approximation gives correct results only
for Z < 10-15 (radiation corrections to scattering were first obtained
by J. Schwinger (174); cf. also (55)).
Table 14 gives the values of dR and dB (in per cent) showing their
dependence on e , ■&, Ae.
RADIATION CORRECTIONS 717

T a b le 14

& 45° 90° 135° 45° 90° 135° 45° 90° 135°
e 2.5 MeV 4.0 MeV 9.5 MeV
<Wz 0.61 0.89 0.86 0.62 0.92 0.99 0.63 0.94 1.07
j A e — 10 keV 4.8 7.4 8.7 6.9 9.9 11.3 12.4 15.9 17.5
Ae = 25 „ 3.9 6.0 7.1 5.7 8.1 9.3 10.5 13.5 14.8
d*j.Jc = 50 „ 3.2 5.0 5.9 4.7 6.8 7.9 9.0 11.7 12.8
1As - 100 .. 2.5 3.9 4.7 3.8 5.5 6.4 6.7 9.9 10.8

Formula (51.18) for dR becomes inapplicable for As -> 0 since


as Ae -*• 0 the scattering cross section, strictly speaking, tends to zero.
We note that in the nonrelativistic domain the quantities dB and
dR tend to zero as v -> 0 .
51.4. Removal of the Infrared Divergence for an Arbitrary Scattering
Process
We have seen earlier that the radiation corrections to the cross
section for the scattering of an electron by an external field contain
the photon “mass.”
We shall now trace how the infrared divergence arises formally in
the matrix elements. Prior to regularization neither the electron self­
energy part, nor the vertex part (with internal electron lines) diverges in
the domain of low frequencies of the virtual photons. But the vertex
part and the electron self-energy part with external lines do contain an
infrared divergence (cf. § 47).
In the course of regularizing the internal vertex part A we subtract
from it the value of A for free electron lines, which contains an infrared
divergence. Thus, the regularized value of A R now begins to contain
the photon “mass” ).. In a similar manner X appears in the expression
for the electron self-energy part. For this reason the photon “mass”
begins to appear in the radiation corrections to a great variety of scat­
tering processes.
We shall now show, by generalizing the result obtained for the radia­
tion corrections to the scattering of an electron in the nuclear Coulomb
field, that if in the case of any arbitrarily complicated scattering process,
which is an nth order perturbation theory effect, we take into account,
along with the radiation corrections to this effect corresponding to
718 QUANTUM ELECTRODYNAMICS

0? + 2 )th order perturbation theory, also the emission of a.soft real


photon, then the total scattering cross section, which takes into account
both the radiation corrections and the additional emission of a soft
photon, will not contain the photon “mass”.
For the sake of simplicity we assume that only a single electron
participates in the scattering process. The matrix element for such
a process can be written in the form
M = u(p2)Q ( p 1, p 2) u ( p l), (51.19)

where px and p 2 are the electron momenta before and after scattering,
u(p{) and u{p2) are the corresponding spinor amplitudes, and Q is a
certain matrix.
We consider the lowest order radiation corrections to this scattering
process associated with the emission and subsequent absorption of
a virtual photon and leading to an infrared divergence. For example,
let us take bremmstrahlung from an electron as the basic process (diagram
Q in Fig. 107). Then the radiation corrections to this process which lead
to an infrared divergence will correspond to diagrams Wx, Vv V2, shown
in Fig. 107.
RADIATION CORRECTIONS 719

We show first of all that in the contributions made by the radiation


corrections the terms containing A, and arising from an equal number
of electron self-energy parts and vertex parts, mutually cancel.
We take, for example, the vertex A (Fig. 107). The matrix y in the
basic process Q corresponds to the contribution from the diagrams
for the -radiation corrections W1 and Vl
y llS c( p ) E ( p ) + A fi{ p , p ^ = y /lS c( p ) Z l + y ll( p — im) l Z 0( p - i m )
+ S ' ( p ) Z R{ p ) + y ^ L + A ^ R ( p , p 2) , (51.20)
where the notations of subsections 47.2 and 47.4 have been used. As
has been explained previously, the left hand side of this equation does
not contain an infrared divergence which appears only in the constants
2T0 and L. But these constants are related by the identity (47.51) L = —Z 0
and therefore cancel in (51.20). For this reason the regularized value
of the left hand side of (51.20) will not contain any infrared divergence,
as has been stated previously.
If the fundamental diagram Q contains n internal electron lines,
then in the diagrams for the radiation corrections W{ and Vt there
will be n internal electron self-energy parts and n — 1 internal vertex
parts. From all these parts there will remain only one uncompensated
photon “mass” A coming from one electron self-energy part. On taking
into account the existence of two free electron lines with which two
vertex parts are associated, we finally obtain from all the diagrams
W{ and Vt one “Uncompensated” A from the vertex part. In accordance
with (47.52) A appears in this part in the form of In (A/m). On taking
A -» 0 we can, therefore, write the expression for the principal part
with respect to A of the matrix element corresponding to diagrams
Wt and Vt
M^+v (51.21)
7i m

We now consider the last diagram for the radiation corrections Y and
determine the matrix element M Y corresponding to it. In accordance
with the general rules of subsection 25.5 we have

Q(Pi—k ,P i —k)
(Pz—k y + m‘
i ( P i ~ k) - m dxk
x (Px—k)2jrm 2 y M P i ) - k2+ A2 ’
720 QUANTUM ELECTRODYNAMICS

where because of the infrared divergence we have introduced into the


function Dc{k) the photon “mass” A. We are interested in the contri­
bution made to M y by small values of k, and, therefore, wherever
possible, we neglect quantities of order k.
On noting that

= u(p2) { - i ( p 2+ m ) - 2 i k n+2ip2o+ k y a} ^ 2ip2au(p2),


we obtain

M y a s -----=-(/>! Pi)u(p2)
71s

d*k 1
X Q(Pi,Pz)u(Pi)
I

or
My = May,
where
_ ia f d*k 1

k2-\-22 [(p2- k ) 2+ m2][pl - k ) 2+ ~^2] '


The integral

[ J K ________________ 1 ___________
J £ 2 +A2 [(p2- k ) 2+ m2][(Pl- k ) 2+ m 2]

appearing in the previous expression is, in accordance with (47.16'),


equal to

2 7i2i
m2 sin 26

Since we are interested only in the dependence of J on A, then,


without loss of generality, we can assume that the electron momentum
Pi before scattering is equal to zero. In this case
• 2A —e2+m
P1 P2 = — sin2 6 = ---- -— —,
2m
1
sin 26 = i — , 6= arccot v2.
m 2
R A D IA T IO N C O R R E C T IO N S 721

where v 2 is the electron velocity after scattering. Therefore I appears


in aY in the form of the term
, a arc tanh , X
a’y = - - In ,
7i v2 m
and in M v in the form of the term
a arc tanh
M \= M (51.21')
71 Vo

The total matrix element for the radiation corrections will contain
the photon “mass” in the form M ’R = M*,+v+M*, while the scattering
cross section, taking radiation corrections into account, will contain
it in the form
,\M + M
w ;L: 22 & M 2\ 1+, 2 a /arctanhi
----
{ n \
Therefore, the scattering cross section daR, taking radiation corrections
into account, will contain ). in the form of the added term
da\ = bRda0, (51.22)
where
, x __ la I arc tanh v
- 1 In (51.22')
71 \ V2 m

and da0 is the cross section for the basic process.


We now consider together with the radiation corrections also the
emission of a soft real photon of momentum k and polarization e.
The diagrams representing the emission of the photon k are shown
in Fig. 108. The matrix element corresponding to these diagrams is,
in accordance with subsection 25.5, equal to

ie i{p2+ k ) - m
Mk w’0 2) Q(p2+k,p{)
(2 7c)'i/2 11 2(l> ( P z + k f + m2
ijpx + fy —m e
+ QiPziPi+k) (Px + k y + m 2 u(Px).
}/2 a)
On assuming co << £x and on noting that
(ipl + m )u (p 1) = 0 ,
722 QUANTUM ELECTRODYNAMICS

we write M v in the form


1
M = _____ (51.23)
k ( 2t t )3/2 2 oj P\k p 2k I

The cross section for scattering accompanied by the emission of a soft


photon of energy not exceeding zle is evidently equal to
dok = bkdo0, (51.24)

where
e2 ST' /' dzk Pie _ f h e |2
h {Inf Zj J
^ /it
~2co Pi k p 2k\
Ae

On assuming, as we have done previously, that pi = 0 , we obtain in


accordance with subsection 30.3
, la / arctanh?;, As la /arctanhi^ , \ /, 2 /le 5\
’ ■ )( i 67
(51.24')
Finally, we obtain the total scattering cross section which takes into
account both the radiation corrections and the emission of a soft photon
do = doRA-dGk. The part do which can contain ?. has the form
do1 = bxdo0, where bx = bK RA-bk. But in accordance with (51.22') and
(51.24') this quantity does not contain the photon mass; therefore, as
has been asserted earlier, /. does not appear in the total scattering cross
section (96).

51.5. Scattering o f High Energy Electrons by an External Field


As the magnitude of the transferred momentum q increases, the term
(a/7r) In2 (9 2 /m2) becomes the principal term in the expression for the
quantity dR which determines the radiation corrections to the scattering
RADIATION CORRECTIONS 723

of an electron by an external field. This quantity attains values of the


order of magnitude unity when the transferred momenta attain values
q ~ m exp(£ \/jz/a). In such a case formula (51.17) no longer has meaning
since the quantity dR is no longer small. In order to obtain the correct
expression for the cross section for the scattering of an electron by an
external field taking radiation corrections into account in this energy
region we cannot confine ourselves to the approximation just considered,
but must sum the whole perturbation theory series. As will beeome
apparent from subsequent discussion, in each perturbation theory
approximation the largest term in the domain of large values of
transferred momenta will be an expression of the form [(a/n \n2{q2jm2)]n
where n is the order of the approximation.
The remaining terms of the type an[\n2(q2/m2)]s, where s < n , can
be neglected since

a" | In2-", | = an s| ln2 <^ < I n“ l"’A


mr m‘ ml

We shall refer to the largest terms as doubly logarithmic terms.


Our problem consists of picking out the doubly logarithmic term from
each successive perturbation theory approximation, and then to sum
all such terms. Such a summation can be carried out in the domain
of transferred momenta determined by (a/rc) \n2(q2/m2) < 1 , and the
expression so obtained can then be continued into the domain of trans­
ferred momenta defined by (a/7i) ln2(q2/m2) > 1.
In order to explain the manner in which the doubly logarithmic
terms are picked out we consider the matrix element Mj(3) defined by
formula (51.1). In the case of interest to us we are dealing with the
calculation of the correction to the vertex part for which q2 — (p1—p 2)2
> p\, p\. Such a relation between the vectors q, p u p 2 can evidently
hold only for non-Euclidean vectors p^, p.>, q (Euclidean vectors satisfy
the triangle rule).
The integral over k in the expression for Mj(3) does not differ in the
region ku > qa from the similar integral for the vertex part in the case
of Euclidean vectors /?,, p 2, q, and does not lead to the appearance of
doubly logarithmic terms (cf. subsection 47.4). Therefore, there remains
only the region of integration ka ^ qa, which, as we shall now see, leads
to doubly logarithmic terms.
724 QUANTUM ELECTRODYNAMICS

In place of k we introduce new variables of integration u,v, x related


to k by the expressions (first of references (192) and of (1))

k = p^u— v j + p 2{v — +*a (51.25)

k\ = x,
where the vector k L is perpendicular to the plane defined by the vectors
p u pz. Under the assumptions made with respect to p 1 and p 2 the vector
kL is space-like, i.e., x > 0. In terms of the new variables the element of
volume integrated over the angles in the plane of the vector kx has the
form for q2 > p \,p \

dik — q2 du dv dx

The quantities k2, (jp1 —A:)2 +a ??2 and (/?,—k)2+ m 2 are related to u, v,
x by the expressions
k2 = —q2uv-\-m2 {u2jr v 2)+ x,
{p1—k)2-\-m2 = k2+ q2v, (51.25')
(p2~ k ) 2j\-m2 = k2jr q 2u.
It can be easily shown that in integration over x a real contribution
comes only from going around the pole k2 = 0. Since x > 0, u and
v must vary in such a way that the point k 2 — 0 would correspond to
positive jc. From this it follows that the quantities u and v must have
the same sign.
On carrying out the integration over x we obtain in accordance
with (51.25') logarithmic integrals with respect to u and v.
We now establish the limits of integration over u and v. Since the
region of integration over .y is restricted by the conditions K < qa,
it follows from (51.25) that \u\ < 1, |z;l < 1. Further, on taking into
account the fact that the factor k 2 in the denominator of (51.1) must
in fact be replaced by k2jrk2, where X is the photon “mass”, which
is equal with logarithmic accuracy to the frequency of the accompanying
soft photon, we obtain the restrictions \u\ > X2/q2, |z>| ^>X2/q2. Finally,
we must take into account the requirement x > 0 at the point k 2 = 0 .
It leads to the inequality
q'7uv—m2(u2jr v 2) > 0 .
RADIATION CORRECTIONS 725

On carrying out in M[3) the integration over u and v taking the above
restrictions into account, we obtain the following expression for the
additional term to the vertex part in the third approximation of pertur­
bation theory:

This expression agrees, as it should, with formula (47.53).


We now consider the corrections to the vertex part in higher order
approximations. In diagrams of the nth order for the vertex part there
are n-1 virtual photon lines intersecting in an arbitrary manner. In this
case the numerators of the fractions in the matrix elements corresponding
to these diagrams coincide for q2 ;> m2, and only the denominators of
the fractions are different.
We first consider diagrams with two virtual photon lines. There are
two such diagrams —one with intersecting, and the other with parallel
lines.
On introducing in place of k 1 and k 2 the new variables uv vv ,xx
and w2, v2, x 2 in accordance with (51.25), and on carrying out the integra­
tion over x x and x ,, we obtain for these diagrams expressions propor­
tional to
{ (Pi h ) [(/?i kj) + ( M 2)] t(/>2 k i) + (Pz k 2)](p2k 0 } - 1

The integral with the highest power of the logarithm is obtained from
the first expression when the conditions (p-^) ;> (pxk ,); (p2ki) >> (p2k 2)
are satisfied, and from the second expression when the conditions
(Pik2) > (p^j); (Pzk2) > (p.>^i) are satisfied. In this case the expressions
corresponding to the two diagrams simply coincide. Addition of these
two expressions leads to the integrations over and u2 becoming
independent. On interchanging the variables k l ^ k 2, and on taking
half the sum of the resultant and the initial expressions we can easily
show that the integrations over v1 and v 2 also become independent.
As a result, we obtain the following simple expression for the addition­
al term to the vertex part proportional to a2
726 QUANTUM ELECTRODYNAMICS

Thus formula can be easily generalized to the case of the /zth approxi-

On summing all these expressions we obtain

(51.26)

The continuation of this formula into the region (ajn) In2 (q2/m2) > 1
obviously does not alter the form of F.
It follows from (51.26) that the cross section for the scattering
of an electron by an external field for large q2 taking into account
radiation corrections up to terms of order a differs from the cross
section without the corrections by the square of the exponential factor
appearing in (51.26) (first of references (1))

(51.27)

In order to elucidate the physical meaning of the quantity A it is


necessary to consider processes of scattering accompanied by emission
of an arbitrary number of soft photons of total energy not exceeding
a certain limiting value Ae. (It is assumed that Ae < m.) In this case
it turns out that with logarithmic accuracy ?. should simply be replaced
by Ae.

51.6. Radiation Corrections to Electron-Electron and Electron-Positron


Scattering
We now proceed to investigate radiation corrections to electron-
electron and electron-positron scattering (5), (163), (153).
Since electron-electron scattering cannot be regarded as the scattering
of an electron by an external field, then in order to determine the ra­
diation corrections to this process we cannot directly use the foregoing
results which refer to the scattering of an electron by the field due
to a nucleus.
The principal types of diagrams which determine electron-electron
scattering and radiation corrections to this process are shown in Fig. 109.
In order to obtain all such diagrams we must in addition to the diagrams
given in the figure also consider diagrams obtained from those given
RADIATION CORRECTIONS 111

there by means of the substitutions 1 )Pi+±p2, P^^-p'2\ 2 )


3) Pi pv
We denote by M t the matrix element corresponding to the zth
diagram of Fig. 109 (without the (5-function). We respectively denote
by M[, M u M[ the matrix elements obtained from M t by performing
the substitutions 1) p l <±p2, 2) p[<±p'2\ 3) Pi<±p2- The

. t*. u Pl-k V
P'l r Pi \ k+9\r r Pi P% "X / " Pi
\^p\-pi \k
__ L._____
*p\ Pi *P'i v fi* Pi P\ v pr k fi Pi
(D (2 ) (3)

P'l H-
*------ r*-------
P'i \ P%
P-9
vP\~k | Prk v
Pi h Pi
Pi
(5)

p\ ! Pt P* : p2
\* |9 !9 k
14 ■M----
Pi p\+k Pi Pi p.' pr k V Pi
(6) (7)
Fig. 109.
differential cross section for purely elastic electron-electron scattering
taking radiation corrections of order a into account can then be written
in the form

'V'7
5

I 1
where v is the electron velocity in the center of mass system, and the
summation is taken over the orientations of the electron spins in the
728 QUANTUM ELECTRODYNAMICS

initial and final states; do is the element of solid angle into which the
electron is scattered.
In order to obtain the cross section for electron-positron scattering
da'e we must in this expression setp 1= p_,p[ = p L ,p 2= —p'+, p'2 = ~P+>
where p_ and p + are the four-momenta of the electron and the positron.
The photon “ mass” A appears in the scattering cross sections dae
and da'e (it is contained in the matrix elements corresponding to dia­
grams 2 ,3 ,5 of Fig. 109).
In order to eliminate A it is necessary, as we have done in finding
the radiation corrections to the scattering of an electron by an external
field, to investigate the inelastic electron-electron and electron-positron
scattering accompanied by emission of a soft photon. The diagrams
corresponding to this process are shown in Fig. 109 (6 , 7) (together
with these diagrams we must also consider those diagrams which are
obtained from the ones shown in the figure by means of the substitu­
tions 1) 2) P ! ^ p 2, 3) p,<±p2, p [ ^ p 2).
We shall reproduce here only the expression for the total scattering
cross section do = dae+dau which does not contain any infrared di­
vergences in the limiting case of high energies.
If p > m, sin ^ -—' 1, where p is the electron momentum and ft is
the scattering angle in the center of mass system, then the differential
cross section for electron-electron scattering in the laboratory system
has the form
a2 ( 1 —v2) ( 1
^ = 4 F k l ~ ) | 2 <2 - 3 * + 3
m
+ 2 0 + 2 0 , - 2 & b) l n - - - ( 2 - 3 z + 3 f - f )
ZZl£

+ — (2&-4&x + 4 & f - \ l f ) + 0 a( 2 - 2 x + f-)


+ 0 b( 2 - 2 X+ 3 f - f ) - 0 l { 2 - X)
- 0 \ { 2 - z + 2 f - f ) - ^ ( 6 - 5 z+ 5 f-2 f)
+ 2 0 0 a(6 —7 z + 6 * 2 - 2 f ) - 2 0 0 b( 1 0 —1 1%+ 1 6X2- 5*3)
+ 2 0 a0 b( 2 - 3 x + 3 x * - X z)

— y^ ( 2 - 3 Z + 3z 2 - Z 3)-^ -% (2-% ) jdo + terms

obtained by replacing X ->■ \ —X, 0 ^ ± 0 b, (51.28)


RADIATION CORRECTIONS 729

where %= sin2 # / 2 and


(p'i~Pi)2 — q2 = 4m2 sinh2 0 ,
(Pi-P?)1 = ql = 4m2 sinh2 0 a,
(P'i—P2 )2 = q l = 4m2 sinh2 0 b.
In the same limiting case the electron-positron scattering cross
section is given by

a m
4 ( \ - 2 0 + 2 0 b- 2 0 a) In
71 Z Z ie

- 0 2( 2 - 9 Z+ 1 9 ^ 2- 1 5 ^ + 6 / ) - 0 2( 6 - 1 5 ^ + 1 9 Z2- 9 Z3+ 2 z 4)
- 2 0 2( l - 3 Z+ 4z2- 3 Z3+ ^ ) - 2 0 0 a(lO -17Z+ 24z2- 1 7 ^ + l O %4)
+ 2 0 0 i)(6 -1 2 ^ + 1 3 ^ 2- 6 z3+ 2 ^ ) + 2 0 Q0 fi( 2 - 6 z + 13z2

- 1 2 Z3+ 6 ^ ) + y ( 2 8 - 4 2 %+ 5 U 2- 2 3 %3+ 6 ^ ) + ^ - ( 6 - 2 3 %

^ 5 1 Z2- 4 2 z3+ 2 8 ^ ) + 0 6( 2 - 5 z + 6 Z2- 5 ^ + 2 / )


r2
” ( I - Z + Z !I! + - 4 - z ' ( 5 - 6 2 + V ) do. (51.29)

In the limiting case of high energies and small scattering angles


( p > m , pO ~ m) the electron-electron scattering cross section assumes
the form
a2(l —v2)
da = ( 1 —(3fl) do. (51.30)
4m2 sin4 —
2

where

d coth2 0 I ( 1 —0 coth2 0 )——

m
+ 4 (20 coth 2 0 - 1) I In ^ - - 1 I + 4 0 a(20 coth 20 - 1 )
20
40
1- x tanhx dx-j------r— r I xcoth.v dx
sinh40 tanh20. tanh2 0 J J
730 QUANTUM ELECTRODYNAMICS

In the case /? > m , pd ■ — m the electron-positron scattering cross


section is given by a similar formula.
In the limiting case of ultrarelativistic energies, when In (p/m) > 1,
s i n # ~ l , the electron-electron and electron-positron elastic scattering
cross section is determined by

dae = d<y0(l — dR), (51.31)

where da0 is the cross section for the basic scattering process and

8r —

- <5, = $9 = <5S= - - In ™ + 0 l \ .
71

_ 4a 0 a
o4 —
n 3

71 2 ZI£

(the quantity (5; corresponds to the /th diagram of Fig. 109; 8U takes
into account the emission of a soft photon).
Formulas (51.28) and (51.29) are no longer applicable if q2, ql or ql
become much larger than m2, since in such a case the radiation cor­
rections become of order of magnitude unity. In such a case we cannot
confine ourselves to the approximation just considered, but must
sum the whole perturbation theory series. For q2, ql, ql^> m2 such
a summation in the case of electron-electron scattering does not differ
essentially from the summation carried out in subsection 51.5 in
discussing the problem of the scattering of a high energy electron
by an external field. It leads to the replacement of formula (51.31)
by
dae — da0e~dR, (51.32)

where d'R is the principal term in dR containing squares of logarithms


(third of references ( 1 )).
A similar relation also holds for electron-positron scattering (third
of references ( 1 ).
RADIATION CORRECTIONS 731

§ 52- Radiation Corrections to Photon-Electron Scattering, to Pair


Creation and Annihilation, and to Bremsstrahlung

52.1. Radiation Corrections to the Compton Effect


We now proceed to determine the radiation corrections to scat­
tering processes in which photons as well as electrons take part. We
begin by considering the radiation corrections to the Compton effect (36).

Fig. 110.

Figure 110 gives diagrams showing the basic Compton effect (dia­
grams Y0 and Yf) and the radiation corrections to this effect of order a
(diagrams y is Y2, Y:1, Y4 and Y f Y', Y f Yf).
The matrix elements which determine the basic effect and the ra­
diation corrections are given in accordance with the rules of subsection
25.5 by
^Q i+fci—p 2~ k 2)
Si*f = ie2(27i)i (u2Q0ul)
2 |' <o1(o2
(52.1)
b(Pi+ki—P i ~ k 2)
s ^ f = — e^iu^Qu^)
2 |/a>1fo2
where

i f i ~ m e ^ - e , if, —in a
nVr x 2~e25
A A

Qo = e.,
m2'Kx

A = P i + k i = P-i+k2,
732 QUANTUM ELECTRODYNAMICS

A — Pi k-z — Pi >
m2x1 = f 2-\~m2,
m 2x 2 =11+™ °",

Q = 2 ( r , + r;),
i=l

i{p2—k)—m a i ( A ~ k ) —fn a i{px—k) —m dxk


W
y. (p2- k f + m 2 c? ( / , - k V*
) 2 : ni- e' (px- k ) 2+ m2 />rk 2+X2 '

A A

_ f d4A: i(px—k) —m a i ( f x—k) —m i/^—m a


2 J k2 (/i-^ j2+ ' « 2 m2 x 1gl>
A A A A A

v _ a i f i —m f i ( f x—k) —m a / ( —/c) —m a dik


Y z ~ e% m2x A J ^ ( f .- k Y - h ^ 2 ^ XpA-kY + in2 y >‘ k 2 ’

= a t'A - m f i ( f x—k ) —m d*k tf-m a

9
/ " ( / 1- i t ) 2+ ffl2 ' m 2x x

(we have adopted here the same notation as in § 28).


The matrices Y' differ from the matrices Yt by an interchange of
the photon momenta k x and k 2.
As we already know, the cross section for the basic process da0
is equal to
. _ 'Y02 co2_
II
o

m2x2
where

U0 = g Sp { 2 oOA—'h)0 o ( # 2 - ' ”)}

-4 :U -M - 4 TX , + -XM
'
- l ^X n+ - ?X l ) .

The cross section for Compton scattering taking radiation cor­


rections of order a into account can be written in the form
da = da0Jrdal
r$ <
d\ do
4 m‘ % Sp { [ a - / ( 4 y 2 ] { i k ~ 'm) [ e «+ ] (rp‘~ m)
RADIATION CORRECTIONS 733

(o: do
= ^ o + -4
;-- \lTCf Spi v 2 o ('> i-w )g ( i p - m )

On noting that

Sp {Qo(iPi-m)Q(ipa—m)} = (Sp {Q(ipt - m ) Q 0(ip2—m)})*,


we rewrite da in the form

(52-2)
where

u i = '|£-.■ Re Sp-t { [ 2 ,+ e ; ] (iPi-m)Q„(ip,-in)'j ,

Qi = 2 r , . Q i = £ y ;- (52.3)
i= l i=1

Further, we introduce the notation

A (*1**2) = -j^ 2 Re S p y { (^ (i^ -m ) GoO'.Pa—"*)}»


(52.4)
P2(yn ,K2)= Re S p y {eiO '^ —^QoO^a—wi)}.

It can be easily shown in analogy with subsection 28.4 that


P<i Oa, x2) = Pi(*2, «i). (52.4')
Therefore can be rewritten in the form
Ul = P P ^ 2 , y-\) • (52.5)
Thus, it suffices to evaluate Pi(xlt k^). This quantity can be written in
the form
4

P l ( X l , * J = 2 P {n), (52.6)
n= 1

where
734 QUANTUM ELECTRODYNAMICS

We first of all evaluate P{1). It is convenient to begin this calculation


by evaluating the traces of the matrices, after which the integration
variables ka will appear only in the form of scalar products k 2, kpx, kp2,
and this considerably simplifies the evaluation of the integrals ap­
pearing in P {1). The latter have the following structure:

a-, or. ate) J (1) (2 ) (3) (0)


where
(1) = (Pi~ k f + m 2 = k 2—2kp1,
(2 ) = (p2—k)2-\-m2 = k2~ 2 k p 2,
(3) — ( /1—k)2jrm2 = k2~2kf^-\-m2xx,
(0 ) = k 2+X2,
and the symbol ( 1 ; ka; kakT; k akTkQ) denotes 1 in the case of the in­
tegral J, ka in the case of the integral Ja, etc.
The simplification in the evaluation of the integrals due to the
evaluation of the traces can be illustrated in the example of the quantity
2p i g ' J m o- This quantity can be written in the form
= f l i k p ^ k ^ k = f [ k 2- ( k 2~ 2 k Pl)\kak T
= JI ((1) d*k = /£>. jJor
w5
1 ) ((2)
2 ) (3) (0)
(0 T J (1) (2) (3) (0)
where
k 'k ^ k
0 ) ”(2) (3)"'
(52.7)
yd) _ f KK<M
aT j (2)(3)(0)
(we also use analogous notation for the other integrals, by introducing
a superscript to denote the factor of the type (1), (2), (3), (0) lacking
in the denominator).
In this way we are able to reduce the quantities containing the
integrals Jarg to the evaluation of simpler integrals. Analogous simpli­
fications also arise in the evaluation of other integrals.
We list the values of the integrals which appear in the end result:
1 2/
j( ° ) = -------- )
n 2i xl Jr x 2

1 _/< 1) = JI2) = J [F(Xl l)-T '(-l)],


n 2i
RADIATION CORRECTIONS 735

W I J ‘3' = l ^ l y tM Z ^-A W -lnA ],

■~ ^ ( 2 y ) - 2 h ( y ) + \ a H
pU
n 2i ~ sinh2 y

J ( 0) _ J( 0 ) P l g ^ P z o
+ J<0)- Ah \ (ha+q2a
nH ° ^1 + ^2

7Z2i a
Jlv = U (1)- u " + — 4 |n "i - 2) - - ■«>,.
^ = t K + ( *1 1 /

1 ^ L . \ Pu + 12J<« + ^ - 1 In »r 2 U ■
n 2i a xx~ l s£x 1

7i' ( si nh2>' >P:^ P^>

(52.8)
7 l Ll 4 \" u 1 2

1
2y2f i af i z+ b [2 fla(qu + 7 2t) + 2 / 1t (7 1ct+ 9 2ct)
«x+ «2

U '0 ’ - -2 I (fcator+tflr • q2a)

n 2i J ™
^ = - M A ° + 12 - 3 + ^ ' nXl+ Z,m )

1 r
+ T P2aP2z
2
^ + *1-- t1 J1
(*l - l )

+ - ^ jr i 6 i» > + ^ + t r 6^ - ~ 2) m „
2 («1 - 1 ) 2

2 ^ + 9 « x- 1 2 1 1 , ,
— 2^ - ^ J +

2 ^ —9^ ! + 6
x | 6 y l l , + — S F T ) ^ lnx .+-■«i—i - 4 J
[<
736 QUANTUM ELECTRODYNAMICS

where the following notation has been introduced:


sinh2 y =
b- 1 —y cothy,
1 r
h{y) = — ucothudu,
y 0j
x
F ( x ) = ( In (1 + « ) — ,
J u
and
A ° = l - XnW '
In evaluating the traces of the matrices appearing in P a), we must
keep in mind that Yz is flanked by the matrices ip^—m and ip2—m,
and that both the matrix ip2 appearing on the left, and the matrix ipv
appearing on the right can be replaced by —m. This substitution,
together with the summation over p, enables us to rewrite the numerator
of the integrand in in the form

A A A a A A A A a A A ^

= - ' l i k y i k y vk - 2 i p 1k y vk y x+ 2 i k y J 1yvk + 4 m k 2bvX—2iyvkyxk p 2


+^Pxkyv{ifl —m)yl + 2 y v{ifl —m)yJcpi
A A

-4 (/h /> e )y J i(/i—fc)—m]y;i,


after which the evaluation of the traces presents no further difficulties.
(Here and in subsequent formulas for P in) we have replaced in accordance
with the rules of subsection 25.5 the vector ex by yx and e2 by y,,.)
In order to be able to express the traces of the matrices in terms
of the invariants and %2 it is useful to employ the following relations:
P\ = Pl = —
k \= k\= 0 ,
f\= xi-l,
f \ = x2—l,
2p ^ k i = l p 2k 2 = 2f \ k \ = 1f\.k 2 = Xj_,
2p i ^2 = 2p 2k x = 2f zk x = 2f zk z - x 2,
2 p i/i = 'l-Pzfx = 2 ,
RADIATION CORRECTIONS 737

^•Plfz — — ^2 2 ,
-P 1P 2 — *l + * 2 ---2 ,
/ = -22 1 /2

(in these relations we have set m = 1 ).


We finally obtain the expression for P {1)

\ X2 /
4 2*2 \
+ {J[a0)- J ^ ) P 2 a U + y<o,/i l a
»1 /

*1 *2 / \ *2

K “« a A . ( 1+ “ + ± )+ a / 2 + - ) ]
^1 ^2 x2
I( 1 \ Ia IT 12 4 2*2 , 2 *;
\ ^2 <^1 5<l ^2

X2

4 ^ +
^2 ^2 ^2

+ -/(0) ( 5 - 2 *,—*2- + — + —
\ Xi X2 X2 %X

+ 7( —2-j-x1-j-x2) ( 3+/c 2 -----------------1— ■- ] ■ (52.9)


\ Xi X2 *2 /

The integrals 2Japla and 2Jaf la appearing in this expression are


related to the integrals J (0), J n), J (2), 7 (3), J by the expressions
2Jap u = 2JaP2a = / ' “' - i ' " - i (01^ 12’
2Jaf i d — J {0)—J {z)+ x J .
We now proceed to evaluate P (2):

pt. 2) — 1 1| ifi-m r i(A -k)-™ i ( P i - k ) - ™ d*k_


16 tt2 P f \ 7v' m2xv J 7'lk2- 2 k f 1+ m 2x , yA k2- 2 k p \ ' 7fi k2

M<P l- r n ) [ y , r .+ r .~ m i- y > \ O p * - " ) j •


738 QUANTUM ELECTRODYNAMICS

The value of P <2) will obviously be unaltered if we carry out the


substitution px+±p2, («i and remaiii unaltered) and rewrite
all the matrices following the symbol Sp in the inverse order. If in addition
we interchange the indices v and A, we obtain P (2) = P (3).
On noting that the regularized value of T3 has the form
y _
i A —m ( 2 71^
9 * 93 ^ xr (Pi , P i + k x\ fri).
rtv-Kx
*
te 2
i V

where in accordance with (47.56)

^ { P i , P i+ k x\ = -~2 ( A + iBk,yk+ iCp?+ Dkxp?) ,

A = ~ ln 1 1m I) ~ 2 + -2 i(r^- —r1 )r ln xx (* - 0 - ^ ( - 1 )},


ln
B=
«i-
^ = ~------------
C 1 , 3 *i - rr^ln^,
2
1 — «1 l^ -l)
D = + I n * , - 27 { f ( Xl- l ) - F ( - l ) } ,
l —Kx Xx Xl&l —I) Jq
we obtain
/>(2)_j_p(3) _ 2 />(2>
1 I / f, —in **
— T -—i — (A y> + iBkirx + iCP>. + D k 1p ^ ( i p x- m
4
. i f i —m , i f i —m \,.~
X ' y' m ^ y'+ y '

—h / n 3 + ^ . — H—-■)—25(ji14'X2)
*1 I \ *1 *2 *2 /

+ 0 —8 + ^ + 8
*1 *2 Xo

D I /\ —xl-\-x2—x \ ~ x l x.1- 4*i U (52.10)


r
Finally, we obtain P(4). The regularized value of Y, has the form
RADIATION CORRECTIONS 739

where ^ \ p ) is defined by formula (47.24). We write S i2){fx) in the


form

Z r ](A) = ^ 2i{B1(ifx+m) + B2m] ,

where

x}
B‘2 _ H ,----t----- rr2-
*1 2)
-ln»x.
^ —l (p i-ir
On utihzing this notation we obtain in accordance with (52.6') the
expression for P {i)

P<4> = 16 1\m i S p \ /v ^ + - m) + £ 2m ) ( / / , - m )y ,

(
A A \
ifi—m ifz—m t..-

2 *1 2 / ' 8 %
8 %n
4^, yx?
Cn
— 8 + «2— ------— -j

+ - - 5 1(3 + x1- - - - - 4- + ^ (52.11)


2/<1 \

Finally, on adding P {1), P{2), P (3), and P w we obtain


4

Pi ( xj , x2) = ^ P {n) = (1—2y coth 2 y) • U0 ln2—2y coth 2 y [2h{y)


n= 1
2 — coth 2 j;
-h{2y)\-U0-\-h(y) I —4y sinh2y- ^- + 2y cothy

4coth2y «j_— 6 1 ^ 4
+ In I I\ 4y coth 2 y [ + „2
?^2 2^2 cosh 2 y xx

1 *2 jJ _|_J I 3x2 8
Xx 2 xx x2 X2 Xx 2 xx x\

7 3« 2 ( 2«x— —x\x 2 2x\-\-X2 4y2


X
xxx2 2^ 2 j^*2 (*i — 0 2x2{xx—I)2) xx~\~x2
740 QUANTUM ELECTRODYNAMICS

2
7 1
X
"4* 1
| * l ] - 4 ^ anh v ( i - ■ iW L H-----
/ \y-i *2

3 ___ 1 __/ J
+ — [/^ -i)
2 x2 x\ Xi — 1 \ x2 2
2 3
(52.12)
Xl «2 *2 xl X2

The photon “mass” I appears in the quantity Ux defining the radiation


corrections to the Compton effect. In order to eliminate it, it is necessary,
as was done in the investigation of the radiation corrections to the
scattering of an electron by an external field, to take into account in
the Compton effect the emission of an additional soft photon. We refer
to such a process as the double Compton effect.
In the laboratory system of coordinates {px = 0) the cross section
for the double Compton effect can be written in accordance with the
general result of subsection 51.4 in the form

daD= - ~ d a 0 y f , (52.13)
71 UQ
where da0 is the cross section for the basic Compton effect,

= 2(1—2y coth2y) | l n ^ - — + 4y coth2y [h(2y) — 1] (52.14)

and Ae is the maximum photon energy.


The total cross section for the scattering of a photon by an electron
taking into account both the radiation corrections and the emission
of an additional soft photon is given by

da = daQ1 1— --- ■ (52.15)


y 7c u0 )
As can be easily seen, this quantity does not contain the photon “mass.”

52.2. Limiting Cases o f Low and High Energies


The foregoing general formulas can be considerably simplified in
the limiting cases of low and high photon energies.
In the region of low energies, when iox£a a)2 = to << m the scattering
cross section has the form

da — — do( 1 - 2 co+ 2 eo2 ...) t/ 0 — { U ^ U D) (52.16)


71
RADIATION CORRECTIONS 741

where
U0 = 1+ cos2 ft,
4 I
Ux = —“'rco2( 1 —cos#)(7 0 In—
J m

-f- ( 1+cos # + c o s2# — —cos3 #)4co2 In — , (52.16')


3 m
4 Q \ F- r , 2 As
UD= —yco 2( l ——cos#)t/0ln
cos ft)U0 In —-—
A
and ft is the scattering angle.
In the region of high energies we consider three cases depending on
the value of the parameter (x1+ x 2)/x2 (in the laboratory system of
coordinates this parameter is equal to {xl -\-x^jx2= \ x x( \ —cos#):

In the first case xx ^ x2 > 1; here in the laboratory system we


haveo»2«^ c o 1 >> m , 1—cos# << m / a ^ . In the second case x2 :> 1,
;> 1 , and in the laboratory system co? cox m , 1—cos ft ^ m / c o v

In the third case 1> N i l «2, and in the laboratory system


c o 2 '— ' m , 1 — cos ft > > m / a j l .

In the first case

I
Ux = 4(1 —2y coth2y)ln — —8 y coi)\2y\2h{y) —hQy)]-\-4yh(y)co\hy

-fln |« 1 |(4y tanhy —1) —2_y2 —4y tanhy + 3 —In21xx I---- y (52.17)

In the second case

Xy
In 1 + - ln ' — i+jr 2 (52.18)
x2
742 QUANTUM ELECTRODYNAMICS

Finally, in the third case


Ki
*2

The quantity UD is determined in the second and in the third cases


by the formula

(52.20)

while in the first case the general formula (52.14) should be used.
The ratio —(a/n) (C/j/ C/0) attains its maximum value at d = 0;
this maximum value is equal to 0.04 for co = 50 MeV and to 0.09 for
a) = 1000 MeV.
In the high energy region due to the existence of terms containing
In2 (eo/m) the correction to the cross section for the scattering of a photon
by an electron becomes comparable with the basic cross section. There­
fore, formula (52.15) ceases to have meaning, and, as in the case of the
investigation of corrections to the scattering of an electron by an external-
field, it becomes necessary to take into account higher order approxi­
mations of perturbation theory. Such a procedure leads to the multi­
plication of the scattering cross section given by the Klein-Nishina
formula by an appropriate exponential factor. In this case doubly
logarithmic terms come from diagrams of type Fx and Y[ (cf. Fig. I ll)
with additional virtual photon lines (second of references ( 1)).
It should be noted that in the high energy region the multiple
Compton effect involving the emission of a large number of secondary
photons becomes important. In the center of the inertia system of the elec­
tron and the photon these processes have the character of the ordinary
Compton effect accompanied by emission of additional photons of
energy small compared to the energy of the principal particles. Therefore,
such additional photons have little effect on the momenta of the principal
particles. However, in the laboratory system these “additional” photons
RADIATION CORRECTIONS 743

can have energies exceeding the energy of the “principal” secondary


photon.
The following expression for the energy distribution of the secondary
electrons can be taken as characteristic for such processes:

dw(e) -
1 de/e
In"1- “ in* " I
I 2a
' 2a In’ “ 1
£
+1j expj — — In2 —
71

m 7i e
a , cd, , w, A / a , an , a>,\1
— In— In— +
+ k7 11 e x p ----- In— In—M •
Z £ m I \ 71 £ W/J
Up to terms of order a the total cross section for all such multiple
processes is given by the ordinary zero order Klein-Nishina formula.

k"2•i
!*,

\
1-P+k p-S
T

Y'
'2 >3

Fig. 111.

For the differential cross section this assertion does not hold for
very small scattering angles. (We note that the additional photons are
in this case very soft compared to the primary secondary photon).
The total scattering cross section for small angles of deflection of the
original photon is given by the following formulas (second of references
( 1 )):
/ m m
d c = d a 0e x p i r i n ’— for > 0>
V « : coi
m
da — da0 exp for 0<
m "i
744 QUANTUM ELECTRODYNAMICS

However, at large energies these angles make a negligible contribution


to the total scattering cross section.
A similar situation arises in the case of the scattering of a positron
by an electron for angles of scattering close to 180° in the center of mass
system (third of references (1)). The total cross section for all the processes
accompanied by the emission of any number of additional photons
has in this region the form
m „ 71
da — daQ exp---- In2 ------ -n for — < < n — 0 < --
7L JX (J 6 2

m
exp ——In2 - - for

V
K
da = da0 i

1
n m e

where s is the energy of the positron or the electron in the center of


mass system. For high energies this region makes a small contribution
to the total cross section. Therefore, the sum of the total cross sections
for processes accompanied by the emission of additional photons is
expressed by the zero order approximation formula (up to terms of
order a).
This also applies to the differential cross section for the scattering
of a positron by an electron for angles of scattering which are not too
close to 180°, and to the cross section for electron-electron scattering
for all values of the scattering angle.
52.3. Radiation Corrections to Two-Photon Pair Annihilation
If we know the radiation corrections to the Compton effect we can
easily obtain the radiation corrections to the process of two-photon
pair annihilation (84). Indeed, diagrams representing two-photon
annihilation including radiation corrections (Fig. Il l ) do not differ
topologically from diagrams representing the Compton effect with
corrections (Fig. 110). The difference consists merely of the fact that
instead of the two electron states in the Compton effect, one electron
and one positron state participate in the process of two-photon annihila­
tion and that, moreover, in the Compton effect one photon is absorbed
and another is emitted, while in the process of pair annihilation both
photons are emitted. Therefore, if in the formula which gives the cross
section for the scattering of a photon by an electron including radiation
corrections we change the sign of the four-momenta of the incident
R A D IA T J O N C O R R E C T IO N S 745

photon and of the scattered electron, we shall obtain the cross section
for two-photon pair annihilation taking radiation corrections into
account.
Thus, we must carry out in formula (52.15) the substitution
p x -> /?_, ky -> —ky, p 2 -> —jo.i_, and /q -*• k2, w h e r e a n d p + are the
momenta of the electron and the positron, and ky and k 2 are the
momenta of the two photons, and at the same time we must replace
the invariants xx, x 2 by xy = —2p _ k x = —2p+k 2, and x2 = —2p+k x
= —2p _ k 2. As a lesult of this we obtain the following formula for
the cross section for two-photon pair annihilation including radiation
corrections:

da = da0 1— Q |2(1 —2.vcoth2x) In - - —Ax coth 2x h(2x) —2h(x)


\ 71 ( m L

Xn ) G ( x 2j ' t"l) ^ , (52.21)


k ] ~ a<Kx-
where da0 is the cross section for the basic process of two-photon pair
annihilation and

U0G(xi, x2) = [h{2x)-h{x)\ x sinh2x(l +2cosh2x) + 2 * tanhx


I X\X2
X-2 Xi
In i xx; |4 x coth 1----— sinh2 x
2xy X<L

2x I xy—6 3x2 3 7 8 8
+ - + 1 - - ---- \~~
sinh2x \ 'U . x2 X^X^ X\ %i

x \ - 2 x x+ x \ x 2 _ 2x[ + x 2 y _ 4 x c o t h x i - :
2 x \ x 2( x x— \ ) 2 x 2{xy — l ) 2 I \ 2 Xy

71“
—x “ 12 3xy
: _ Ixy _ 2x\
4 '. + ' Xy
cosh2 .V \ xx 4 Ax 2 \ Xy X2 2x2

- 2- 1 + — 1-, ( ^ + * ) + - 2- [ n * l - l ) - ^ ( - 0 ]
x\ Xy — 1\ x2 2 I Xy

X
•JL
X2
X2

x{
. X\

Xo
x2
2
. 2—3-ll. %2
746 Q U A N T U M E L E C T R O D Y N A M IC S

2
* + +4 - L + i
x2 xx\ \ x x x2 \ *1 *2
4 cosh2 x = xx-\-x2-
This expression contains the photon “mass” . In order to eliminate it
we must add to the cross section da the cross section for three-photon
pair annihilation, in which in addition to the photons k x and k2 a soft
photon k 3 is emitted whose energy does not exceed Ae.
The cross section for this process is equal to

d&3 2 (lx coth2x—1)


2

+ 4 x coth2x[l—h.(2x)]\. (52.22)

We state the formulas for the total cross section in the limiting cases
of low and high energies.
In the nonrelativistic case we have
, , ( , , 7ia a I jr2 \ ) na .
</<7 = <fa0 ( l + — — — < 1. (5 2 .2 3 )

where v is the positron velocity, v <C 1; the electron velocity is equal


to zero. (The divergence of this expression for v -» 0 is associated with
the inapplicability of the Born approximation.)
In the relativistic case we have
2Ae
da dan{ 1—— I 2(1 —2x coth 2x) In------- \-X (52.24)
JJ

where X has the following values. If the angle between directions of


motion of the two photons is 180° and x x > ^ then

X = ( \ n x x) * - l n *, + 3 + -^ -.

If the angle between the directions of motion of the two photons is 90°
and xx = x2 > 1, then we have

x = - On x xy - A In 2xx+ 5.97+ [18.5 (In 2xx)2

67.25 In 2^+41.2].
R A D IA T IO N C O R R E C T IO N S 747

52.4. Radiation Corrections to Bremsstrahlung


We now investigate the radiation corrections to bremsstrahlung
from an electron in the Coulomb field of a nucleus (67).
The diagrams representing the basic process of emission and the
radiation corrections to it are shown in Fig. 112.
Since the matrix elements corresponding to the diagrams Yu Y'
contain an infrared divergence, we must in addition to the radiation
corrections also take into account double bremsstrahlung, in which
a soft photon (k‘, co') is emitted together with the photon (k , a>).
We do not reproduce here the calculation of the cross section taking
these effects into account, but confine ourselves merely to stating the
final results for several of the more interesting limiting cases.
We write the radiation corrections to the bremsstrahlung cross
section and the cross section for double bremsstrahlung in the form
CL Ct
doR = dRd(jQ, doD dpdoQ,(52.25)

where do0 is the cross section for the basic process.Then in the limiting
case coe! -< m2 (we Use the same notation as in § 51) we have

bB = 2(1 —x coth2x) | l n ~ + 1J + x tanhx

v2 sin2
2x
4x coth2x [h (2x) —h (x)] +
\ —vl sin2y

2(1 —x cothx) 11— coth2xj — y ; (52.26)

2Ae 1 , l —v , 1—v2
dn = 2(1 —2 coth2x)7 In —,-1 In -------- cosh2xC7(w, 0),
& V ?. 1 v)
> 7 \-\-v
M -7 1 1 . 0
v sm "2

where

r du 1-\-vu
G (v,0)= J - In
0 1—vu
eos-J- ( l ~ V 2U2 y U 2—cos2
2
748 Q U A N T U M E L E C T R O D Y N A M IC S

and x is related to Q = q2 = (—2Jr x l + x 2)m2jr 2(e1e2—p1p 2cosd)


by the expression q = 4m2 sinh2x (de is the maximum energy of the
soft photon).
As should have been expected, this expression agrees with the ra­
diation correction to the cross section for elastic scattering (cf. formula
(51.18)).

F ig . 112.

Further, we consider the relativistic case when > w, e2 > m,


but the energy loss is small, i.e., el5 £2. If, in addition, the scattering
angles are small, 6, 0l5 02 ~ (m/e) < 1, then

SR— 2(1 —2xcoth2x)[ln-----pi ) + x tanhx

+ 4 x coth2x[/?(2x)—/i(x)]-|-2(l —x cothx) ico th = * ) - § ;

(52.27)
8 9p
<5^ = 2(1 —2x coth2x) In ------\-2 In —(2xcoth2x—1)
A ttl
—4x coth2x [//(2x)—//(x)'j.

In the extreme relativistic case when not only the conditions


el; e2, to > m are satisfied, but also the condition l n j ^ j , In *2, In q ,
RADIATION CORRECTIONS 749

ln(e—x i ~ x 2) > 1 holds, but Inig/xj), \t\( qIx 2), In [fe—kx—x 2) / \ | ],


In [(£—xx—hz)/\ x 2\] ~ 1, the quantities dR and dD are equal to

dR= 2 ( \ - 2 y ) \ n - + 2y2- 3 y — —x,


m 3
(52.28)
dD= 2 ( 1 - 2 y) In — + 2y2- 2 y l n ^ * 1" - ^ - — - l n ^ 2 ,
A 4e1e2 mL
where

t = y ln (<?— x2)-

Finally, in the nonrelativistic case we have

<5* = y [ - ( / > i - / > 2)2| h i - ^ + y J + 2 ( / J 1- / > 2)Arln— j -


(52.29)

For /?!->■ 0 the corrections to the cross section vanish.


The variation of the cross section in the energy range for which
ay2/ji > 1 (1) is similar to that discussed earlier. Moreover, in this
energy range multiple bremsstrahlung begins to play a role. However,
the energy of one of the bremstrahlen is always considerably larger
than the total energy of all the other photons. Because of this the energy
lost by the electron in the form of radiation is expressed by the usual
zero order approximation formula.
The number distribution of emitted photons is given by the Poisson
formula. The total cross section for scattering accompanied by emission
of any number of photons of arbitrary energy agrees up to terms of
order a with the zero order approximation to the cross section for
scattering unaccompanied by bremsstrahlung.

52.5. Radiation Corrections to Pkotoproduction and Single Photon


Annihilation o f Pairs
Radiation corrections to the photoproduction of pairs in the Coulomb
field of a nucleus can be obtained from the radiation corrections to
bremsstrahlung by means of the substitution P i~ * p +, P2 ^>P-> and
k -> —k. We must keep in mind here that additional poles appear
in the integrals over the momenta of the virtual photon.
750 Q U A N T U M E L E C T R O D Y N A M IC S

As usual, we must add to the cross section for photoproduction


the cross section for photoproduction accompanied by the emission
of a soft phdton.
The total cross section for the photoproduction of a pair in the
field of a nucleus taking radiation corrections of order a into account
can be conveniently written in the form

da = da0 (52.30)

where da0 is the cross section for the basic effect and d is a quantity
which determines the corrections. We give here the expression for (5
in several limiting cases (81), (26).
Near the photoproduction threshold, when \p+\, \p_\ < m, <5 is
given by the formula
71* cos 8— cos 8T, cos 8
8 + 2 .4 - _
(52.31)
V -V
— sin2 0 , + - — sin2 8
P- + P+
where 8+ and are the angles between k and p +,p_; 8 is the angle be­
tween p + and /7_,and v ± are the velocities of the positron and the electron.
Here the first term takes into account the interaction between the
created electron and positron in the first Born approximation. There­
fore, the applicability of formula (52.31) is restricted by the condition
a/\v+~ r _ | << 1.
In the relativistic case of equal energies (e+ sw e_ >> m) and of small
angles (0±, 8 < ^ m / e ) d is determined by the formula
712 13 7l2
<5 (52.32)
p+P- ~2 4‘
y ~ 2- 2 m2
Here the first term is analogous to the first term in formula (52.31),
since the small quantity \/'(—2—2p+p_/m'2 represents the relative velocity
of the electron and the positron in the center of mass system for the pair.
Finally, in the ultrarelativistic case when In (kp+/m2) > 1, ln(/?_p+
/m2) > 1, In (q2/m2) > 1, but In (q2/kp±) ~ 1, In (p^p+/kp±) ~ 1, (5
has the form

13
<5 In 2p+P~ , £+£- In 2p+P- (52.33)
m2 11 ( / I f : ) 2 6 m2
RADIATION CORRECTIONS 751

The one-photon annihilation of a pair in the Coulomb field of


a nucleus is the process inverse with respect to the photoproduction
of a pair. Therefore, it can be easily shown that the radiation corrections
to the one-photon annihilation are determined by the same formulas
as the corrections to photoproduction, if we interpret the values of
p +, p_ and k appearing in them as the momenta of the particles being
annihilated and the momentum of the emitted photon.

§ 53. Radiation Corrections to Atomic Levels

53.1. Radiation Shift o f Atomic Levels


So far we have been considering radiation corrections to various
scattering processes. But the interaction of the particles with the
vacuum also affects their stationary states. This effect manifests itself,
firstly, in the radiation shift of levels and, secondly, in giving rise to
level widths.
We begin by discussing the radiation shift of an atomic level. In
order to do this we utilize the effective potential energy of the electron
introduced in § 50.
If we neglect the magnetic interaction between the electrons, then
in the absence of an external magnetic field the effective potential
energy of the electron h a s. the form

where <p is the potential of the self-consistent electric field in the atom
and E is the intensity of this field. If the electron is in the state ipn of
energy En, then the radiation shift of the level will be given by the ex­
pectation value of ed<p in the state y n:
dE’n = (e5<p)nn = (v»B, ebcfWn)

= _ L _l_jjL(in-^-_A — L \ (A <pf+ — < £ a £ > j ,


137 Any 3m2 \ 1 X 8 5/ V ^ ^ m / \
(53.2)
where <L> is the expectation value of L in the state y)n.
This expression contains the photon “ mass” X, and, as has been
explained in § 30, this indicates that the expression does not take into
752 Q U A N T U M E L E C T R O D Y N A M IC S

account the interaction between the electron and the long wavelength
photons. In other words, 6E„ determines only that part of the level
shift which is due to the interaction between the electron and the short
wavelength photons. We must, therefore, separately determine the
level shift due to the interaction between the electron and the long
wavelength photons, and add it to the shift determined by formula
(53.2) (19).
The process in which we are interested can be described in the
following manner: an electron in the state y n of energy En emits a vir­
tual photon of energy co and makes a transition to the state xpn., it then
absorbs the photon co and returns to the initial state ipn. The change
in the electron energy associated with this process is given, as is well
known, in the second perturbation theory approximation to which
we here confine ourselves, by
(53.3)

where Vnn, is the matrix element of the energy of interaction between


the electron and the photon, and the summation is carried out over
all the states of the atomic electron ri and of the emitted photon (sum­
mation over the photon states means integration over 4 7 i ( o 2d c o / ( 2 j r ) 3 , and
summation over the photon polarizations e).
Since we are interested in the interaction between an atomic electron
and long wavelength photons, we can utilize the nonrelativistic ap­
proximation in which the matrix element of the energy of interaction
between the electron and the photon is given by the formula

where v = (l//m)V is the operator for the electron velocity. On sub­


stituting this expression into (53.3), and on noting that
Ze I (ev)nn' I2 = S I v nn

we rewrite (53.3) in the form


K
<-* o 12
(53.4)
0
RADIATION CORRECTIONS 753

with the integration over co being taken here between the limits from
co = 0 to some large value K.
We must renormalize the mass in this expression. For a free electron
AEn is equal to its electromagnetic mass dm (or, more accurately, to
that part of dm which is due to the interaction of the electron with
photons of energy less than K). But we have already included dm in
the electron mass m, and we interpret the energy of the level n to be
the total energy with m subtracted from it. We must, therefore, subtract
from (53.4) the corresponding expression for the free electron.
Since for a free electron only the diagonal elements of the velocity
differ from zero, formula (53.4) in this case assumes the form

. 2 e2 r , v2
A t = —— - —- coaco— .
3 471- J co
o

On subtracting this expression from (53.4) and on noting that

2 \V n v '? =
nf
we obtain the following expression for the radiation shift of the level En,
due to the interaction of the electron with photons whose energy does
not exceed K:

2 e2
dE” = AEn—AE (En- E nr). (53.5)
3 4^ En- E n, - o j

We now transform this formula. We first carry out the integration


over co. On assuming that K is considerably greater than all the dif­
ferences between atomic energy levels En—E'n, we obtain
2 e2 K
8EL (Em En). (53.6)
3 4;t2 E
m-E
\
Further, on introducing the quantity £0 defined by the formula

2 K J ‘ (Em- E „ ) \ n \ E m- E .
m _ _ _ (53.6')
2 \ v nm\l (E,n- E n)
m
754 Q U A N T U M E L E C T R O D Y N A M IC S

we rewrite dE'n' in the form

m ( 5 3 ' 7 )

The sum appearing in this expression, as is well known (14), is


equal to

v nm\l (Em- E n) = f y%W(r)Vyjn dv


m "

= 2 ^ 2 J A v ( r ) Iy > n ( r ) \ 2 d v >

where V = eq>. Therefore, we have finally

(538)

So far we have made no assumptions about the magnitude of K.


We now choose K in such a way that the minimum energy of those
photons, the interaction with which is still taken into account by formula
(53.2), will be equal to K. But in accordance with (30.24) In A = In 2 AT—|
and the sum of dE'n and dE'n\ which determines the total radiation
shift of the level En, will contain neither A nor K.
In the usual units this shift is given by the formula (109)

e2 I 4 /1 m 3 5 1
6En = 8E'n+5E'n'
(4n)2 e \ 3m2\ n 2e0 8+ 6 5

+ ^ (ft,. £«£*>») I (53.9)

We now evaluate the level shift for a hydrogen-like atom. In this


case (in CGSE units) we have
Ze2
e<p = —
r
and
h2 ti2
(A&p) -^-^4 jte 2Z\y) (0)\2, (53.10)
m2c2 m2c2 ( V » , A e ( PWn)
RADIATION CORRECTIONS 755

but since

/ = o,
|^ n (0 )!2 = na / it
0, / # 0,
where n and / are the principal and the orbital quantum numbers, and
a = h2/me2, we have

fl2 Z-jT^Ry, 1=0,


2 < /% -> =
wrc
0, I # 0,
where Ry = i a 2mc2. Further, we can show that
Z4 Ry
4a2— y=/+-
efi n3 ( / + 1 ) (2 /4 -1 )

— = (53.10')
me
4a2—4 ^ - 7 = /-T.
«3 /(2/+1)
We see that the radiation shift of an .s-level is in order of magnitude
equal to a2E0, where E0 is the energy of the ground state.
We now obtain the shift of the hydrogen atom levels. In the case
of hydrogen the quantity In (mc2je0) is equal to 7.6876. On utilizing
expressions (53.10) and (53.10') we obtain the following values for the
radiation shift of the hydrogen levels (in frequency units):
Av(2si) = 1034 Mc/sec,
Zl^(2pi) = —17 Mc/sec, (53.11)
Av(2pi) — 8 Mc/sec.
As is well known, the undisplaced states 2^ and 2pi have the same
energy. But the radiation shifts of these levels are different and
the 2^ level turns out to lie higher than the 2pi level by approximately
1051Mc/sec.
Formulas (53.9), (53.10), (53.10') determine- the radiation shift
of a level of order a2E0. It can be shown that in the next approximation
of perturbation theory we would obtain a level shift for a hydrogen-like
atom equal to (101)
756 Q U A N T U M E L E C T R O D Y N A M IC S

For n = 2 and Z = 1 this gives 7 Mc/sec. Taking this correction into


account the difference in the 2s}. and the 2p h energy levels in hydrogen
turns out to be equal to 1057.19 Mc/sec, while the experimental value
for this quantity is equal to 1057.77±0.1 Mc/sec (167).
In concluding this subsection we note that by utilizing equation
(49.12) we can obtain the radiation shift of the levels of positronium.
The difference between the ground levels of ortho- and parapositronium
turns out in this case to be equal to (100)

AW = a2i? yj-~ I — + ln 2 j ~ - | = 2.0337■ 105Mc/sec.

The experimentally obtained value for this quantity is (91):


A W = (2.0333±0.0004).105Mc/sec.

53.2. Radiation Shift of the Levels of ju-Mesohydrogen


Polarization of the vacuum plays an insignificant role in the radiation
shift of the levels of an atomic electron. This is associated with the
fact that the Coulomb field of a nucleus is distorted by the polarization
of the electron-positron vacuum at distances of the order of a Compton
wavelength of the electron lilmc while the dimensions of the atomic
electron orbits are considerably greater than ti/mc. For this reason
polarization of the vacuum introduces into the radiation shift of atomic
levels a contribution of order a3, which amounts to approximately 3%
of the total radiation shift.
A different situation exists in the case of a //-mesic atom in which
//-mesons play the role of electrons. Since the //-meson mass is large
compared to the electron mass, the dimensions of the orbits in a //-mesic
atom are considerably smaller than the dimensions of the electron
orbits in an ordinary atom, and the polarization of the electron-pos­
itron vacuum plays a predominant role in the radiation shift of the
levels of a //-mesic atom.
We shall see that the radiation shift of the levels of a //-mesic atom
of hydrogen, or of //-mesohydrogen, is approximately given by aE0,
where E0 is the energy of the ground state of this atom, while for ordinary
atoms this shift amounts to only a2 of the energy of the ground state.
We now obtain the radiation shift of the levels of //-mesohydrogen
(70) (or, more accurately speaking, of a hydrogen-like //-mesic atom).
R A D IA T IO N C O R R E C T IO N S 757

In order to do this we use formula (50.15) for the radiation correction


to Coulomb’s law due to the polarization of the electron-positron
vacuum:
00

n /.c
-2 (t2- i ) 1/2
8(p(r) m r £
dt.
(m2r £2
1
The radiation shift of the energy of a //-meson which is in a state
ipn(r) of energy En is given by

8En = ed<p(r)\y>n(r)\2dr
ou
aZe2 i \(c 2- i y / 2 p~2mrC
1+ dC -~— \Vn(r)\2dr.
6tc2 I 2C2 C2
The ratio of this expression to the energy of the ground stare E0 =
(e2Z 2/4nafj) (a is the radius of the normal orbit of the //-meson in
mesohydrogen, equal to afi = 4nfm^e2) can be written in the follow­
ing form:
1 \(£2_ ni/2
2^ j - ^ - J r ^ s O d C , (53.13)

where

0
Rnl(r) is the normalized radial function and e = (m/m^aZ) (m is
the reduced mass of the //-meson, / is the orbital quantum number).
Approximate integration of (53.13) leads to the following values for
the shift in eV for different values of Z.

T able 15

" W . z
nI 1 6 20

1s 1.8 320 20000

2s 0.2 47 3250

2p 0.014 27 2550
758 Q U A N T U M E L E C T R O D Y N A M IC S

53.3. Natural Line Width


The interaction of an atomic electron with the electromagnetic
field leads not only to a change in the electron energy, but also to the
level acquiring a width, since due to this interaction there always exists
a finite probability for the transition of the atom to the ground state.
We now show how this width, which is called the natural width,
can be introduced.
We start with the Dirac equation containing the mass operator
M (3)(x, x') in the presence of an external field (cf. formulas (43.45),
(49.5)):

where A ^ \ x ) is the potential of the external field including the added


term due to the polarization of the vacuum.
As we already know (cf. § 43), the mass operator takes into account
the interaction of the electron with the vacuum of the electromagnetic
field, and in the first approximation with respect to a is given by the
formula
M (e){x,x') = /e2Dc(x ,x ,)y/i5'^)(x,x')y/i, (53.15)

where S£e)(x, x') is the Green’s function for the Dirac equation for
an electron situated in an external field.
If we substitute into (53.15) the expression (35.3) for S (ce)(x, x')
then the mass operator (in first approximation) assumes the form
M (e)(x,x')

— OO

(53.16)
where ipn(x) = y>n(r)exp(—iEnt) are the “stationary” electron wave
functions, satisfying the unperturbed Dirac equation

(here, as before, we must interpret A (*\x) as the external potential


including the added term due to the polarization of the vacuum).
RADIATION CORRECTIONS 759

We rewrite equation (53.14) in the form


/ — = H (0)i/j + J/9 M (e>(x, x')y(x')</4x' = 0, (53.17)

where H i0) is the Hamiltonian corresponding to the unperturbed Dirac


equation, and we seek solutions of (53.17) which depend exponentially
on the time
y(x) = y)(r)e~iEt.
On substituting this expression into (53.17) we obtain the equation
H (0)y (r)+ j ( 3 M (Ee)(r, r')ip{r')dr' = Ey>(r), (53.18)
or, in abbreviated form,
H Ey(r) = Ey>(r), (53.18')
where

M (Ee)y (r ) = f M ¥ ( r , r ' ) W(r')dr',


CO

M {Ee)( r , r ' ) = J eiExM {e){r, r ' ; r)dr. (53.18”)

The usual perturbation theory can be applied to equation (53.18),


and this yields
E = E l ° ' + 0 W s‘X „+ 2 l-"~ £(0)_£(0) l- + ..., (53.19)
n'^-n
where E ^ ] is the energy of the unperturbed state determined by the
Dirac equation without the mass operator, and
(pM(Ee))nn-= f vll°.)* ( r) p M {Ee)y)(n0)(r)dr= Jy $ )(r)M<g)y'£)(r)dr.
We confine ourselves to considering only the first order pertur­
bation theory approximation
(53.20)
On utilizing expression (53.16) for M (e>(x ,x '), and formulas (53.18”),
we obtain from the above expression
e2 r d*k y i (n |y/Yeiir|«') (n'l y^ e-^ln )
E ^ £«»-+
(2tt)47 J k2- / 0 2 j ' ^ X - l o y E ^ J - E ^ + kv

= ^ o,+ d k Q { E ^ \ E ^ , \ k \ ) { n \ y iielkr\ n ) { n \ y lle-ihr\n),


(53.21)
760 Q U A N T U M E L E C T R O D Y N A M IC S

where
(n\y^elkr\ n ) = J y)n(r )y neikripn.(r)dr,
u>->

dk0 (53.21')
e (£ r .£ ! ? M * i) = j 1(1-,- [(1 -/0 )£<?>-£<0)+*o] [ k * - k * - i 0]

We now evaluate Q i E ^ , E {°\ \ k |). Direct integration yields


711 1
Q(En0)> E n 'W k \) |£| £ (o )_ £(0) (J —/0) —| k | E°n,l\E°.\~'

On the other hand, the quantity l/( x —i0) appearing in the integral
can be replaced by

— L_ = nid (x) + {p
x —/0 x
1

and, therefore,

Q ( E E W , 1*1) = ^ E * . l \ E ° . \ d [ E ™ - E $ > - \ k \ E°,l\E°n,\\

rri 1
................... (53 22)
^ \ k\ £ (0 > - £ ( ? ) -|* |£ 0 /|£ 0 .| • ^ • >

Substitution of this expression into (53.21) shows that the quantity


A E = E —E ^ , which is the level shift due to the interaction of the
electron with the electromagnetic vacuum, is complex. On setting

AE = 6En- i ^ f , (53.23)

we find in accordance with (53.21) and (53.22) the expression for the
real and the imaginary parts of AE.

r dk y (n\y eikr\ri)
_ r w (vn \ y e~'kr\n)
?
6En T6tt3 JT k \4 f — —
£(0)_^(O). E&
|£i°>|

e2 r dk /r(o)
r =
71 88tt^2 J l^ r 2 j - n ^ ( n \YMJ kr\n')(n'\y Me~ikr\n)

/ F (0)
(53.24)
RADIATION CORRECTIONS 761

The real part of AE determines the radiation shift of the energy


of an atom found in subsection 53.1, while the imaginary part determines
the level width.
The imaginary part has a simple physical meaning: it obviously
represents the total probability of the electron making a transition
from the initial state by emitting a photon.
We note that for the determination of the radiation level shift we
could have from the outset utilized formula (53.24), but it is more
convenient to use this formula only for evaluating the contribution
made by long wavelength photons, as was done in subsection 53.1,
while the contribution made by short wavelength photons can be more
simply obtained with the aid of the effective potential energy of the
electron.

53.4. Photon Scattering near Resonance


In § 35 in considering the scattering of a photon by a bound electron
we saw that in the case of resonance, when the photon energy coincides
with the difference between the energy levels of the electron, the theory
yielding the first approximation based on the second order scattering
matrix becomes inapplicable. The usual method of taking radiation
corrections into account is also inapplicable here, since it depends
on the expansion of the scattering matrix into a series in powers of a,
while it can be easily seen that in the case under consideration the
expansion is made in powers of the quantity a/(E^0)—E[Q)—to) which
becomes infinite at resonance.
We must, therefore, take radiation corrections into account at
an earlier stage of the calculations which in the simplified derivation
of formula (35.14) corresponds to the introduction of complex fre­
quencies.
From the derivation of formula (35.14) for the cross section for
photon scattering we can easily conclude that the resonance denom­
inator in the matrix element is related to the .structure of the function
S {ce)(jq , x z) , which in accordance with (35.13) contains £ n+co in the
denominator. Therefore, we should expect that the replacement of the
function S^e)(x1, x i) by the exact electron Green’s function, which
takes into account the interaction between the electron and the electro­
magnetic field, must lead to finite results.
762 Q U A N T U M E L E C T R O D Y N A M IC S

We denote this function in the presence of an external electro­


magnetic field by Glee)(xlf x2). It satisfies equation (49.5):

Gie)(xi, x 2)
= S^>(xi, x2) —i j S<e)(*i, x 3) M (e) (x3, x^G [e)(x4, x 2) dix 3d 4x4.
(53.25)

By utilizing expression (35.3) for S^e) (jaq, x2) we can easily show
that the Green’s function G{ee)(x, x') also satisfies the equation

--- ieAiue)) + m) Gie)(X’ X')

+ J M M(x, x")Gle)(x", x') d*x" = - i d ( x - x ' ) . (53.26)

We now show that if we introduce into the matrix element for the
scattering of a photon by a bound electron in place of the function
(jjq, x 2) the function G(ee) (x 1, x2) , we shall obtain a finite expression
at resonance in accordance with the results of § 35 (128).
We expand the function G[e) (jq, ;t2) into a series in terms of the
complete system of functions ipn(r) which are the spatial parts of the
wave functions for an electron in the field Ajf^x).
We expand the time-dependent coefficients in the above expansion
into a Fourier integral. Such a combined expansion of G[e) (jq, xrj
has the form

G(ee)(* i,* 2) = 2 ^7 J ^ e ('"(1‘- t,)/„„(oj)xpn(r3)rpn(r2)

+ 2h j f (53.27)

We now obtain the equations satisfied by the expansion coefficients


f nm(a)). In order to do this we substitute the expansion (53.27) into
the integral equation (53.25). On noting that

oo
r i r -t,)
j vM P si'H x u x ,) * . = 2^7 J K ll ,
RADIATION CORRECTIONS 763

we obtain the system of equations for f nm(co)


1
/nnM = £„(! —/0)+a>

2 » ■ ( - < • ) /„ ( » ) } ,
m=£n )
(53.28)
1
fnm(p*) £ /0) + (U ^ ^nmi jmmC^)

- 2 0.)|,
P^m J
where

and

H n m{ 1) =fV n ( r 3) M ( e > ( r 3 , ri I 0 V m (r i) d r 3 d r i-

Introducing the notation


1 /mn(") = R„
= fi nn> n.
fn» 4 H
we rewrite the system (53.28) in the form
Hnm(-(o) , v Hnp(—a>)Rpm
— = (53.29)
E„ + co + p^m
2 En+w

Qnn— En+a) + Hnn{—o))+ Hnm(—(o)Rn (53.29')


m^n
In principle the system (53.29) enables us to determine R nm, as a result
of which Qnn can be obtained in accordance with equation (53.29').
By assuming that the quantity Hnm{—(o) is small, and by using the
method of successive approximations we obtain

H np(—(o)Hpm( oS)
Km E +0J jtfnm( w) + £ —o)—E„ +
71 I p^m
(53.30)

Qnm = Oi +I En
r? +i H
JJnn(~oS)+
( \ I 2\ j1---------—
^ii«( £ ------ "
m^n
764 Q U A N T U M E L E C T R O D Y N A M IC S

In the first approximation we have


Qnn^ M+En+ H nn(-(ti),
(53.30')
^ ______ ^mn( °>)____
ft
mn a)+Em+ H mrn(-co)
whence it follows that

nn o)-{-En-\-Hnn(—(jj)
(53.31)
r _ _ • • • _

mn [(ti+En+ H mm(—a))\ [o>Jr E nJr H nn( —o))\


We see | / mn|<< | / nn|; therefore in future we can neglect the nondiag­
onal elements f mn.
We now obtain G(ee)(x1, x 2). On substituting (53.31) into (53.27)
and on neglecting the quantities f mn we obtain
1 ^ r ^io(£i £j)
G[e) O i, x2) ph £ y n{ri) v>„(ra) J —
+ o>+HTn(—(o)
da). (53.32)

This expression differs from expression (35.3) for the function


S lce)(,xi>x 2) by the denominators of the fractions appearing in the in­
tegral. The quantity Hrn(—aj) which appears in G<,e) (xl , x2) obviously
coincides with the quantity ((3ME)nn, which determines the radiation
shift and the level width (cf. (53.23)):

= dEn- - r n.

Since this quantity is complex the denominators of the fractions in


(53.32) cannot vanish. On substituting expression (53.32) instead of
5^e)(x1, x 2) into formula (35.6) we obtain the photon scattering cross
section of the form (35.14).

§ 54. Photon-Photon Scattering and the Lagrangian for the Electro­


magnetic Field

54.1. Photon-Photon Scattering Tensor of the Fourth Rank


Interaction of electromagnetic fields with the vacuum of the electron-
positron field must lead to interaction between these fields, in partic­
ular, to interaction between photons. Indeed, let us consider two
R A D IA T IO N C O R R E C T IO N S 765

photons of momenta k± and k 2 whose energy is insufficient to create


a pair. Such photons can create two virtual pairs. As a result two photons
will appear of momenta k 3 and k 4 related to the momenta of the original
photons k 1 and k 2 by the single condition—the law of conservation
of energy-momentum k1-\-k2 = k 2-\-ki .
This process represents photon-photon scattering (56), (3), (103).
It obviously cannot be described by means of the classical field equations,
since in classical electrodynamics electromagnetic waves are propa­
gated independently without affecting each other.
In order that the field equations should be able to describe the
process of photon-photon scattering, and also other processes involving
interaction between electromagnetic fields, they must be non-linear,
in contrast to the linear Maxwell’s equations. From this it follows that
the Lagrangian density of the electromagnetic field cannot be a quad­
ratic function of the fields, i.e., it must differ from the function
l 0 = h e 2- h 2).
In the lowest order of perturbation theory to which we confine
ourselves here, the interaction between electromagnetic fields is described
by a part of the scattering matrix S w .
In accordance with (24.14) the matrix S U) has the form

SU)= J d*x2J d4x 3j J4x 4r(A ^ (x J A (x 0 v (-* i))


X N (y (x 2)A (x 2)y)(x2)) N( y(x3)A(x3) y ( x 3)] N f o ix J M x J y i x t ) ) } .
Since we are interested in processes in which only electromagnetic
fields but no real electrons take part, then in this approximation we
must replace the product of the electron operators y>(x)y)(x)by pairings
between them. The total number of pairings is, obviously, equal to
six (they are represented schematically by Fig. 113). All these pairings
produce equal contributions, so that the part of the scattering matrix S w
which is of interest to us has the form

S <4) = — d*Xt J d*x2J d4x 3J flf*jr4iV(A>l(jr1)A,(x2)AA(x3)A0(x4))

X Sp { y ^ c(x2—Xi)yvSc(x3— x3) ytfSc( x i - x 4)}. (54.1)

Here only the quantities AA(x)are operators.


766 Q U A N T U M E L E C T R O D Y N A M IC S

In momentum space this expression assumes the form

S {i) — j d*k2f d*k3f d ' k ^ Y d f a + k t + k t + k j

X,
{2nY (in y (2

*\ h * h
A * )

- :

"1 "2

£XI
F ig . 113.

where
A/fc) = f A /t(x)e~ikx d*x
and
TftvXaikl t k 2, k 3, k$)

= f d4p S p { y / i p + m ) - 1y v( i p - i k 2+ m ) - 1y t ( i p - i k 2- i k 3+ m ) - 1y a
X (ip—ik2—ik3—ik4-i-m)~1}. (54.3)
In place of the tensor TflvXa{kl , k 2, k 3, k4) we introduce the tensor
JwXofti, k2, k 3, k 4) which is fully symmetrized with respect to the
simultaneous interchange of the tensor indices and of the variables k t:
^fivXrr(^1 >^2 >k 3, ^ 4) = ' ^2> ^3> ^ 4)
(&i , ^2> ^4> ^ 3)“!“ Tit kra^k it k 3, k 2, k ^ . (54.3 )
It is clear that in expression (54.2) we can replace the tensor
T^xoikn k 2, k 3, k j by $JM,Xa( k u k 2, k 3, k j :

SM = - ^ ) fd 'k ijd 'k 'fd 'k tfM ^ y d ^ + k '+ k z + k ,)

XN\ A W W M W l ; (k k k k) (54 4)
(2 jr)4 (2 tt)4 (2 7 t)4 (2 tt)4 k * ' k *> k *>-
RADIATION CORRECTIONS 767

We call the symmetrized tensor J vXa(klf k 2, k 3, k j the photon-


photon scattering tensor of the fourth rank.
In accordance with the results of § 46 the tensor T Xa(klf k 2, k 3, k J
diverges for large momenta of the virtual electrons. In order to show
this we note that for large values of p the tensor T^vXa{kx, k 2, k 3, k J
behaves in the same way as ^ ( 0 , 0 , 0 , 0 ) :

T^ a(°> °> °> °) =


X Sp { y ^ ( i p - m ) y v( i p ~ m ) y x( i p - m ) y a( i p - m ) } .
On substituting into this expression
i S p { Y r ( i p —™)vMP—m) Y i O p - m ) y a(ip~m)}
= 0>2+ m 2)2(<V 6 ^ + 6 ^ <5„a+(5^ d j + 8 p ^ p vpxpa
- 2 ( p 2+ m 2) (pMp v dla+ p vpx d ^ + p ^ , dvx+pxpa d'J,
we obtain

W 0, 0, 0, 0) = - y (<v K + & * \ ' + \ a O / ^

in2 \ (j>2+ m2)3 ^kfjPv &Xa~^~PvPx ^fxa~^PfjPa ^rX^PxPo

, 32 r #p
+ CT2 J (/>2+m 2)4 P»P*PxP°-
But
/ P . p J ^ d ' p = i ^ f />2/(/>2)

$PtlPvPxPj(P'i) diP = + jV /A A V -
Therefore
1 f4 C d*p
W 0, o, 0, 0) = ^ j3 ^ + < 5 ^ <5„a- 2 ^ i j J (^ — ^ r

+ y (^ v O w2j (^ 2 ^ 2 )3

+ y ( 5/iV^<T+^<r +^ O ' ” 4j (^ 2 ^ 2 )4 j'

The first integral appearing in this expression diverges logarithmi­


cally.
768 Q U A N T U M E L E C T R O D Y N A M IC S

Insofar as the symmetrized tensor J ^ i k x , k 2, k 3, k ^ is concerned


the divergent integrals contained in its individual terms cancel, but,
nevertheless, the tensor J^xXki, k 2, k 3, k d must be regularized just
as the divergent tensor k 2, k 3, k A), since from considerations
of gauge invariance the tensors T flvxa(0, 0, 0, 0) and J ^ (0, 0, 0, 0)
must obviously vanish for k x = k z — k 3 = k i = 0, while the expression
J Xa(0, 0, 0, 0) differs from zero. As has been already explained in
§ 46, this loss of gauge invariance is associated with the existence of
divergences in quantum electrodynamics.
In order to obtain the regularized value of the tensor
J/jvXoikx, k 2, k 3, k x) we must subtract from it its value for k x = k 2
= k 3 = k x — 0. We denote the regularized value of J^xX ki, k 2, k 3, k x)
obtained in this manner by I vxa(kx, k 2, k 3, k J :
I/xviaiki> K kid = JfivxXki, k z, k3, k x) J^x X0, 0, 0, 0). (54.5)
The regularized matrix S(4) has the form

S<4>= - f d % J d * k 3J d ikX27tyd(k1+ k 2+ k 3+ k i)

„ N l A Xk i) K ( k ) 2 Aa(/c3) A0(/c4) (54.6)


^ ^ k i , k 2, k 3, k y .
\ (2tt)4 { I n f (2jt)4 (2jt)4

54.2. Photon-Photon Scattering


We now establish the connection between the tensor UXkx, k2, k3, K
and the photon-photon scattering cross section.
We consider the process in which the electron-positron vacuum
absorbs photons of momenta kx and k2 and polarizations ex and e3,
and emits photons of momenta k3 and kx and polarizations e3 and e4.
This process corresponds to the six diagrams shown in Fig. 114. Diagrams
1, 2, 3 and 4, 5, 6 differ only by the sense of going around the electron
loop, so that the matrix elements corresponding to them are the same.
The total matrix element for photon-photon scattering is determined
in accordance with subsection 25.5 by the following general formula:
s<4) = e4b(kx+ k 2- k 3- k d
l~*r 2 (w ^cugcoj1/2
X {elfie2ve3xeia j Sp {y/J(/p + m )-1yv(/p + / i 2+ m )-1

X y x { i p + i K - i h + m)~1y X iP - ik i + m)~1} d*P


RADIATION CORRECTIONS 769

+ ei„e?veiaen J Sp {yfi(/p+m) 1yv(i;p + ika+m) L


X r A ‘P + '* 2 - ikt +.m)~lyk( i p - i k x+m)~1} d*p

+ ei»ei?.eiv e-a\ —ikz+m)-1

X yvO'p — i h + ’k z+ m y iyg(ip — ik1+m)~1} d*p}. (54.7)

(1) ( 2) (3)

\ k
\*4 V \ \a Vk /
\* 4 /
\\ rt
P // s

X ---- \ P /
r — *
P~kA P-K1 P-*, P-Ar. P-k%

/ p - * 4-/r3 v / P ~ k A~k3 \ _/P ~ k A* k 2 \


/ v '
/k /r2\ . *2\
S A*
X/r2 V
(4) (5 ) re)
F ig . 114.

Utilizing expression (54.5) for the regularized photon-photon scattering


tensor we can rewrite this expression in the form

S<4) — J /_£2 \ 2 (2nyd(ksJr k i —k 1—k 2)


*--*/— 2 \ 4 ji) (a)1ai2cuna»4)1/2 ei,<e2i-e3Ae4<r
X / (^ ( - ^ i , - ^ A:8 , ^ . (54.8)

In order to determine the photon-photon scattering cross section


which in accordance with subsection 28.3 is an invariant, it is con­
venient to go over to the center of inertia system of the photons, in
which the total three-dimensional momentum of the photons before
770 Q U A N T U M E L E C T R O D Y N A M IC S

and after scattering is equal to zero: k 1Jr k 2 = &3+ £ 4 = 0. We denote


the four-momenta of the photons in this system by
k x (k, ico), k 2 (—k , ico), k 3 (k\ /co), k t (—k \ /co),
where |£| = \k'\. (It follows from the law of conservation of energy
that the frequencies of all four photons are the same in this system.)
On taking into account the fact that the flux density of photons k 2 with
respect to the photons k x is equal to 2, we obtain the following expression
for the differential photon-photon scattering cross section:
1 I e2 V 1
da = m ¥ iw I -* « • W *
' (54.9)
where do = 2n sin Odd is the solid angle which contains the momentum
of the photon k 3 (9 is the scattering angle, i.e., the angle between k
and *')•
Thus, for the evaluation of da we must know the value
of I^xai—kx, —k 2, k 3, kt), but we can obtain a simple analytic expression
for the photon-photon scattering tensor only in the two limiting cases
of low and high frequencies, when co < m and co > m. In these two cases,
which we shall now consider, we can also obtain simple closed expressions
for the photon-photon scattering cross section.
In the low frequency case co -< m the photon-photon scattering
tensor can be expanded into a series in powers of the frequency. The
first term in this expansion corresponding to co = 0 vanishes for the reg­
ularized tensor; therefore, the expansion begins with the term containing
all four frequencies linearly, i.e., which is proportional to co4 in the
center of inertia system. On the basis of this result, and of formula
(54.9) we can easily conclude that the photon-photon scattering cross
section is proportional to co13 for co m. On the other hand, in accord­
ance with (54.9) it must contain the eighth power of the electron charge
e. On taking into account the fact that the scattering cross section must
have the dimensions of an area, and that we have at our disposal only
the constants e, Ti, c, m, we obtain unambiguously the following expres­
sion for the integrated photon-photon scattering cross section for co •< m:

(54.10)

where a is a dimensionless constant.


R A D IA T IO N C O R R E C T IO N S 771

In the high frequency domain co >> m we can in practice neglect the


electron mass m in the expression for I Xa. Therefore, in this case the
mass m will not appear in the expression for the scattering cross section.
But from the quantities e, h, c, co we can construct only one expression,
viz., e8c2/co2, which contains the eighth power of the charge and which
has the dimensions of an area. Therefore, the photon-photon scattering
cross section for co >> m must have the form

(54.11)

where a is a dimensionless constant.

54.3. Connection between the Photon-Photon Scattering Cross Section


and the Radiation Corrections to the Lagrangian of the Electro­
magnetic Field
We now undertake a more detailed investigation of photon-photon
scattering in the low frequency range co < m. In this case the fields
F(E, H ) associated with the photons satisfy the conditions

We call such fields slowly varying.


Slowly varying fields cannot in practice create real electron-positron
pairs (if we exclude from consideration pair production by a large
number of soft photons); therefore when conditions (54.12) are satisfied
we can speak of a purely electromagnetic field.
The state of a slowly varying electromagnetic field can be described
by a Lagrangian which depends only on the components of the field,
and is independent of their derivatives with respect to the coordinates
and the time. Since the Lagrangian must be a relativistic invariant,
the fields can appear in the Lagrangian only in the form of the two
combinations E2—H 2 and (EH)2, which are the only two independent
field invariants.
If the fields are weak, then the Lagrangian density of the field L can
be expanded into a series in powers of E2—H 2 and (EH)2:
L — LqT L i , (54.13)
772 Q U A N T U M E L E C T R O D Y N A M IC S

where
L q= i ( E 2—H 2),
Lx = a(E2—/ / 2)2+ /9{EH)21 ...
= a{FlkFikf + b F ikFklFlmFm+ ... (54.14)
and a, 13, ...,a, b, ... are constants interrelated by
a = Aa-\-2b and = 4b, ... . (54.14')
The terms collected in L x are radiation corrections to the principal
Lagrangian L0. These corrections are associated with such effects as
photon-photon scattering. Indeed, the Lagrangian L = L0+ L x leads
to non-linear equations for the electrodynamics of the vacuum,
which, naturally, can describe photon-photon scattering.
We retain in the expansion of Lx only the first two terms, and show
how they may be related to the photon-photon scattering cross section.
We first recall that in the first approximation the scattering matrix
has the form
(54.15)
where L'(x) = j ^ x ) A^(x) is the part of the Lagrangian which determines
the interaction between the electromagnetic and the electron-positron
fields. We have seen previously that interactions between electromagnetic
fields are described by a part of the scattering matrix S(4). But the same
interactions can, obviously, also be described by the equivalent first
order matrix S<Jf>, without taking into account any virtual electron-
positron fields, if we interpret L' in (54.15) as an additional term Lx to
the Lagrangian of the electromagnetic field. Thus, we have

(54.16)
Such a treatment is obviously valid only for slowly varying electromagnet­
ic fields to the investigation of which we now confine ourselves.
On utilizing relation (54.14) we can obtain the form of the photon-
photon scattering tensor, or, more accurately, its dependence on the
photon frequencies. In order to do this we compare the matrix
RADIATION CORRECTIONS 773

with the first order effective matrix

Sffi = ' / A 'W F ,W F ,1W)! + i ( F i,W F „ W F „ W R ,( ,) ||f t .

On substituting into this expression

F '" ( * ) = ( 2 ^ /

where

and on noting that

J (FiMF*i(x))d*x A„(fc2)
(2tt)4
A;. (&,) ACT(/C4)
(2 7 t)4 " ( 2 7 c) 4
S pvio(ki,k2, k 3>^)(27r)4<5(^1+ ^2+ ^3+ fc4)»

where S k 2, k 3, k 4) is the symmetrized tensor:


S Mria(k i , k 2, k 3, k 4) = 4{k2/jk Xvku k 3aJr k 3flk 4vk u k2a-\- k iftk 3vk u k lg
&^k4Xk 3a(kxk ^ - dXak2/1k lv(k3k 4) - d MXk Ayk 2a(k1k a)
d„k3Mk u ( k M /jak^vkuikik^) 6v2k Xak4ii(k2k^)
+ b).a(ki k 2) (k3k t) + 6^ 6ra(kxk3) (k2k4) + 6^ 5vX(kxk4) (k2k 3)},
and (54.17)

J Fift O) F kl (x) Ftm(x) Fmi (*) dAx

= - J ' dAk x j"d4k2j d 4k 3J d 4k 4(2n;)2 d(k1Jt-k2+ k 3-\-k4)

w A J k J A v{k2) A x(k3) A a(k4)


R,tvto(kii k 2, k3, k 4),
(27r)4 (2ti)4 (2tt4) (2n)4

where R „Xa(kx, k 2, k 3, k 4) is the symmetrized tensor:


RprXa^ki, k 2, k 3, k 4) = k luk lllk 2?k 3cjJ\-k2/tk3vk ixk la-{-k3iik lllk 4xk 2a
~Fk 2fj k4vk xxk 3a~\- k 4/i k 3vk xxk 2a-\- k 3fik 4vk2xk Xrr-\-dflv\k2>kla (k 3k 4)
"Fk xxk 2a(k 3/c4) k2xk 3a{kxk 4) k 4Xk Xa(k 2k 3) k 4xk 2a{kxk 3)
k\A k 3a{k2/c4)] dXa\k4fi k 3v(kxk 2)-\-k3)i k4v(kxk 2) k 4/]k Xv{k2k3)
—k 2ji k 3V(ki k^) k 3^ k lv (k2k 4) k 2/t k 4v(kxA:3)] + [kXvk3a(k 2k4)
774 Q U A N T U M E L E C T R O D Y N A M IC S

+ ^ 3 v ^ l a ( ^ 2 ^ 4) ^ 4 v ^ 3 a ( ^ 1 ^ 2) ^ 4 * ^la ( ^ 2^ 3) ^ 1 v ^2<r ( ^ 3 ^ 4)

—k 3vk 2a{kx^4)H_ t^2^^3a(^l k^)-\-k3lik 2a(kx^ 4) k 2(k xa(^3 ^ 4)


~ k i/ik 3(i(kxk^)—k 3flk Xa(k2k^)~ k ifi k2a(k xk^)\+ 6va[k2fi k u (kxk 3)
+ k Xjlk2x(kxk 3) k 3fik iX(kxk 2) k 2(i k xx(k 3k^)—k i/jk u (k2k 3)
- k 3fl k2X(lcxfcj]+ 8 ^ [klvk iX(k2k 3) + k iv k xx{k2k3) - k Xvk 2x(k3k A)
—k 3vk ix(jcxk^) k 3l)k lx(k2k i) k ivk 2x(k1k 3)\-{-8^ 8Xa[(Jcxk^) (k 2k 3)
+ ( k xk 3) ( k t k ^ + 8 ^ 8va[(kxk 2) (k3k J + ( k M (k2k 3)\
+ <5„ 0 <U (*ik 2) ( k M + ( k xk 3) (k3k j \ , (54.18)
we obtain
4
^uvXaik1 , k 2, k 3, k 4) = — , 2 {^^/ivAaC^-i>k 2, k 3, fc4)
/ e2 \

+ 6^vAa (fci> k 2, k 3, f c « ) } . (54.19)

We now obtain the photon-photon scattering cross section.


On substituting (54.19) into (54.9) we obtain the expression for the
photon-photon scattering cross section in the center of inertia system
1 02\* J
da = (54.20)
(8 n:)2 \ TUT
An Tcot W * .
where
frf -- p(^i) p{Xt) p(h) p(Xi) T 4
cx c <j Ayvxa ( k x, k 2, k 3, k 4) (aS+bR),
(e2/4n ) 2
(54.20')
5 = - f c 2, k 3, fc4),
i? = e ^ e ^ e ^ e ^ R ^ - k x, - k , , k 3, fcj
(the index is used to denote the two different polarizations of the
photon k t) .
In accordance with (54.17) and (54.18) we have
5 = 4 { 2 ( e ^ k 3) ( e ^ k 3) ( e ^ k x) ( e ^ k x)
- ( eWi)ew.)) (e«*>fc3) ( e ^ k x) (kxk 2) + { e ^ eM ) ( e ^ k 3) ( e ^ k x) [(k2k 3)
- ( k 3ki) \ - ( e {X' )e ^ ) (eUl)fc3) (e{X'% ) (k 2k 3)
_)_(eu,)eua)) (eua)eu,)) (jcxk ^ (k3k i)
_)_(e(A1)e(A,)) (gunman-) (kxk^) (k 2k 4)
4-(eu,)e<A*)) ^e(Aa)e(Aa)-) (k2k^) (kxk 3)}.
RADIATION CORRECTIONS 775

R = 2(eiXl)k 3) (eiXt)k 3) (elXl)k }) ( e ^ k ^ —2(elll)eUl)) (e{Xt)kj) (<e(X>)k 4) (A:3A:4)


—2(e(Al)e(Al)) (eUl)&3) (eiXi)k 3) (k^k^
_j_(eu.)e(A,)) (e^ k 3) (e{Xt)k 3) [(kak ^ —(k2k 3)]
_l_(cw.)eM.)) ( e ^ k j (e{Xi)k 3) [(k2k 3) - 2 ( k 3k i)]
+ ( e (Aj)e,As)) (eul)k3) {eai)ki) [(k2k j - 2 ( k 3k t)\
+(e(Aj)e(A*)) (e{Xl)k 3) (e{Xl)k{) [(k2k3) + ( k 3k4)]
-|_(CU.>CM,>) (eU3)eU.)) [(^fc4) (fcgA^+fofca) (fc2fc4)]
+ (e(AJ)e(Al)) (eaj)ea4)) p ifca) {k3k 4)+{k4k4) (k2k 3)\
+ (« Wl)fiw*)) (c<Weu.)) p ^ ) (k3k^)-\-(k1k 3) (k2k 3)]
Therefore
4a)4
Af = {(8a+26) (e(A‘>/i') («<*■>«') (e<A«>/i) (g<A<)/i)

+ 46[(e(Al)e(Al)) (eWa)/i) (e(A<)/i) + (e(A:i)e(;i4)) (c(Al)/i') (e(Aa)/i')]


+ [4a(l —cos 0)—6(3 + cos 0)] [(e(Al) e Ua)) ( eUl)n') ( eiX,)n)
_|_(e(Ai)e(A,)) (e<Ai)/j') (e<A3>rt)]
+ [4a(l + cos0)—6(3 —cos0)] [(e(Xl) e{X,)) ( e{Xl)n' ) ( e(Aa)/i)
_)_(e(Aa)ea,)) (e^iifl') (e(A*)/,)]+2[8a+6(l + cos20)] (e(Ai)eaa>) (ea,)e(A4))
+ [4a(l + cos 0)2+6(5 —2 cos 0 + cos20)] ( e iXl)e {Xi)) ( e iXt)e {X3)
+ [4a(l —cos0)2+ 6 (5 + 2 cos0 + cos2 0)] (eUl)e{X3)) (e(X,) e{X,)) } ,
(54.21)
where n and ri are unit vectors in the directions k and k ' and 0 is the
scattering angle cos 6 = n - n'.
On substituting this expression into (54.20) and on averaging over
the photon polarizations we obtain
1
da = (48a2+ 4 0 a 6 + 1162)o)4(3 + cos20)2 do. (54.22)
(2 ^
On noting that a = 4a+26 and /5 = 46 we can rewrite this expression
in the form

da = (48a2—8a/0+ 3/02) a;4(3 + co sz 0) i/o. (54.22')


(Pn)
776 Q U A N T U M E L E C T R O D Y N A M IC S

We see that the frequency and the angular dependence of the photon-
photon scattering cross section for a> •< m does not depend on specific
values of the constants a and b, while the magnitude of the cross section
is determined by these constants.
Before evaluating the constants a and b we shall give the values
of M for different photon polarizations.
We denote by the polarization of the photon k t , lying in the
scattering plane (n, n'), and by e|4) the polarization perpendicular to
the plane (n, //'). On utilizing formula (54.21) for M we can easily show
that
M (e^, >, e<3>, e ^ ) = M(e\]\ e, , M3 ) M i (2)

16
(2tf+6)(3+cos26),
471
e<2>, e \ f \ eff>) = M ( e \ ^ , e \» , e[3\ e ^ )
8
2 (%a + 3b —b cos26),
- £ l\
4n I
(54.23)
ef2>, e<3>, e<4>) = M(eff>, e<2>, ef3’, e<4>)

(1—2 cos 0-i-cos20)-|—-——~-z(3-|-6 cos 0—cos26),


e2 \ 2'
l— Y
\ An! 4jt j
M(e^\ef2), ef3>, e<4>=
\ 2
(1 + 2 c o s0 -fcos20 )+ 2 (3—6 cos0—cos20).
471 471
We now proceed to evaluate the coefficients a and b. We utilize the
formula
1i l-y1
i-i/! 1-l/rl/j
i - y l- y 2

------ 1 I I /7 i;_ I /7 i;_ I /7 U I / i i > . —U -4 - f 1) _ i_ / >1 I 1 ). — -w v / ^


B C D = V ' \ d}>X\ ^
ABCD 2 / + +

= 3! J 4,
Vi+Vi+3/j+V^l, 2/j^O.
RADIATION CORRECTIONS 111

where diy = dyx dy2 dyz dy4, and we set in it


A = p2+ m 2, B = ( p —ko)2-\-m2,
C = (p - k 2- k 3)2+ m 2, D = { p + k t f + m 2,
Ayi + By2+ Cy3+D yi = [ p - k 2y 2- { k 2+ k 3)y3A-klyi}2+ m 2f l,
r _ 110 ^2^3 , 0 ^2^-4
f i — 1 + 2 -^ ;J- M .r i - 2 - ^ j ; 2 3 ,4-

With the aid of this formula we can write the tensor T ^ (kx, k 2, k3, kA)
in the form

T ^ J k ^ K , k 3, k 4) = - j ^ i - f d ' y f
X [iq + il— ik2—m\y} [iq+ H— ik2—ik3—m]ya[iq+d+ fkx—m]}, (54.24)
where
1= ^2T2 + (^2 + ^3)T3-^lT4-
In T/iv?a(kl , k 2, k 3, k i) we pick out the terms containing d ^ d ^ ,
d , <5 5 6
flA VO ’
We can show that the sum of these terms has the
/f < 7 V /

following form:

k'2' k^ = rfi f d4yf ~(q2+ mlyj4' {"3(<V 3'a


+ <5/1(J dv>-2d,u 3J [ ( q 2-rm2f 1)2+(q2+ m 2f 1)m2f l]

+ 3 (&,„ <5^ + ^ 3v; + Sm 3J m2f i + m 2(dM


V«5;iJ+ d /I(r i , r 3 ^ <5J

X [(^M-m2/ i H 5 - 5 / 1+ 6 ( a - /S)>;2- 2 a - /S)]+m2[ 4 ( l - / 0 2


+ 8( 1- / i ) ( a 8(1- / 1) a - 4 ( 1 - A ) /? + 8 a y i^3+8/3y2^ 4
+ 2 /1( a - ^ 2+ / 1(l —/x)—2 a/j—/3/J + 2 d/ir baXq2m2[$-2ay2- \ + / J

+ 2 <5/lo 3vr f m 2W y 2- 1+ / 0 ~ 2 ^ ^ 2/?}> (54.24')

where
_ 1 ^2^3 o _ _ 1 k 3k^
a 2 m2 ’ 2 m2 ’

1 k2k4
f 1 = 1—4a_y1_y3—4/3y2_y4.
778 QUANTUM ELECTRODYNAMICS

The regularized value of X ^ A k l ’ k 25k 3’ k i) iS eClUal t0


XlivAa(k l ’ k 2>k 3>k i) ^v;.CT(0> 0, 0, 0)

= ../^||-[5(l-/i)+6(o-ft^-2a-/3]
--'s[4(1 I10(1- .1-2a-fl-2(a-,<(j u-6(l -/,1a
J1
—3(1—/i)£+8a!^i>'j+8/5V2>’4]J+16'5(1,<5^j'—' t- 1+A) -(P ~ 2ay

+ 16V>rf / - ^ J?-

f d 'y i n f , -

Since we are interested in the low frequency case, we can take


a < 1, jjSi < 1 , jyj < 1,

ln/i= ■
- = l+ a -z j,
Ji

i
J1
= 1+2(1-A),
as a result of which XftvXg(k1, k 2, k 2, k^)—Xttvkn{0, 0, 0, 0) assumes the
form
X/ivXa(k l ’ k2’ k 3’ ^4) ’ 0) 0; 0)
= J r f V [ 1 4 ( l - / l) - 4 a - 8 / J
+ 1 3 ( l - / 1) * + 2 4 ( l- /1)7! ( a - » - 1 0 ( l - / 1)(t.+ /))+ 8 a V 1l'3

+8/^,]+16<5,„ /</>[/)- “-1 +/1+/5(1—«


K

—2a(l (1-/,)!J+16^4rfJ
—f i ) y z —

+2/5(1 - A ) * - ( l - A ) * J -1 6 4 ^ i „ f d ‘yfs( 1+ (1 - A))

+ 8(4,„ 4,„+4,„ <5,3-24,a S j f </>[(! - A ) + y ( l - A ) ! l •


RADIATION CORRECTIONS 779

From this it follows that


Xprtaiku k 2, ^3) k 4) —XftvXoi®} 0> 0> 0)+Ar/iVO/l(A:1, k2, kA, k3)
- ^ ( 0 , 0, 0, 0) + X tihJ k l , k 3, k 2, k J - X ^ i O , 0, 0, 0)
=246„, i „ f d‘A 2 ( l - / , ) + 2 ( l - / !) - 3 ( l - / , ) - / ) ]
+24a„„ / </‘y[2(l —/ i ) —3(1 - A ) + 2 ( l
+24a„, / < /M -3 (l —/ i) + 2 ( l —/ 2) + 2 ( l - / 0 - y ]

+ « , , « » , / ^ [ 1 0 ( 1 - / O H 1 0 (1 -/,)* -1 5 (1 - / a)!+ 6 ( l - / , ) 7
- 6 ( 1 - /0 / 1 - 6 ( 1 - / 0 1 3 - 1 0 ( 1 - / 0 0 - 1 0 ( 1 - A ) y + m - f , ) a
+ 2 4 (1 —/ 0 f c - S l t !- H ( l + J ( 7 - ( ! + + 4 ( 1 - /,) (a - y ) y s
—8a(l 8y (1 - f ^ y ^ ^ y ^ + S ^ y ^ )
< /M 1 0 (l-/0 ! - 1 5 ( l - / O H 1 0 ( l - / J)H 8 /? (l-/O y !
+ 8 y ( l - / 0 l '!+ 6 ( l - / 0 / l - 1 0 ( l - / 0 ( < i + / l ) - 1 0 ( l - / !) ( a + »
+ 24(1 - f 3) (a—y)y2—24(1- f 2) ( y - ^ ) y 2+16a2y 1y 3
+ 1 0 ( l - f 3)y + 2 4 ( l - f 1)(a -/S )y2]+4d/iAd„af ^ [ - 1 5 ( 1 - / ^
+ 10(1 —/ 2)2+ 10(1 —/ 3)2+ 4 ( 1 —/ 3) y —4/3(1 —/ i )
+10(1 - / , ) (a + f l —10(1 - / 2) (y+ /?)-10(1 - / 3) (a+ y)
-2 4 (1 - / 0 ( a - 0 ) y 8+ 2 4 (l - / 2) (y■
-/S)y2+24(1 - / 3) (a -y )y 2
—8a (1 —f 3) y 2+ 8/3 (1 - / 2)y 2] .
The integrals appearing in this expression can be evaluated with
the aid of

j+ = 4 -

f y ^ d' y = ^ a ’ f y ? dty = ^ o -
On utilizing these formulas we finally obtain
^fivXa(k1> k 2, k 3, k d Xflv^a(0,
^)+^wja(^:1> ^ 2, ^4> ^3)
- ^ ( 0 , 0 , 0 , 0 ) + ^ ^ , k3, k2, k j - x ^ i o , 0 , 0 , 0)

= ~ { 6 /lvd,a[l4(k2k 3)2+ l 4 ( k 2k 3) (fc3fc4)-3 (fc 3^ ]

+ V 5 VJ 14(fc3£4)2+ 1 4(fc2k 3) (k3k,) - 3 (*2fc3)2]


-<5,A<5,a[3(A:2fc3)2+3(A:3fc4)2+20(A:2^3)(A:3/:4)]}. (54.25)
780 QUANTUM ELECTRODYNAMICS

A comparison of (54.25) with (54.19) yields


e2 \ 2 1 e2 \ 2 7_
4a-{-b = 26 =
4n / 3 0 m * ’ 4tz I 4 5 m * ’

where it follows that

_ _ 5 / e2\ 21 h - - — l— Y '1 (54.26)


a 180 \ 4?r / m4 ’ 180 \ 4 tiJ m*'
On substituting these values of a and b into (54.14) we obtain the
expression for the radiation corrections to the Lagrangian density of
the electromagnetic field of order {e2j 4 n ) 2 (56)
1 e2 \ 2 1
Li = [14FikFkl FlmFmi 5 (FilcFki)2
180 \ 4:rc / m’
2 1
[(E2—Z/2)2+ 1(EH)2]. (54.27)
45 \4 n m 4
The differential photon scattering cross section will in this case
be given in accordance with (54.22) by
6
1 139 4 1 CO
(3 + cos2d fd o . (54.28)
(^ r)2 ~ W a "m2 m

The total cross section is equal to


_ 4 1 139 56 1 / co
where co << m. (54.28')
n 902 5 m2 \ m
If we know the Lagrangian density we can obtain the electromagnetic
energy density
r) T 1
vv = = 2 {E2+ H 2) + n ', (54.29)

where
w' = a(E2- H 2)(3E2+ H 2)+P(EH)2.

The second term gives the radiation corrections to the classical


energy density
w0 = -I (E2+ H 2).

The nonlinear electrodynamic effects can be described with the


aid of a field-dependent dielectric permittivity and magnetic permeability
RADIATION CORRECTIONS 781

of the vacuum. In order to obtain these quantities we determine the


electric displacement D and the magnetic induction B :
dL dL
D= B=
~dE’ dli
On substituting (54.27) into these expressions we obtain
1 1
D‘ = £ < + 4*
(54.30)
1 e4 1
B, = H,+ (2(£ 2- H 2)Ht- IE j(£77)}.
45 4n m4
In macroscopic electrodynamics D and B are related to E and H
by the following expressions
Di = ei k ^ k i &i — t^ ik ^ k -

A comparison of these formulas with (54.30) shows that the dielectric


permittivity and the magnetic permeability tensors of the vacuum
in the case of weak slowly varying fields are equal to

S {2(£2W “ h ^ + ih .h ,) ,
(54.30')
f t . = 6K + ^ - — — { 2 ( E > - H * ) S it- l E iEh}.

54.4. Exact Expression for the Lagrangian of the Electromagnetic


Field
In the preceding subsection we have obtained the radiation cor­
rections to the Lagrangian of the electromagnetic field proportional
to the fourth power of the fields. We now determine the Lagrangian
for any arbitrarily strong electromagnetic field which satisfies the
single condition (54.12).
For the solution of this problem it is convenient to employ the
concept of the electron-positron vacuum as a .system of electrons fill­
ing the negative energy levels (cf. § 18). We' shall in the future call
such electrons vacuum electrons.
Our starting point will be the assumption that the addition w' to
the classical energy density of the electromagnetic field
w0 = * ( £ * + //* ) ,
782 QUANTUM ELECTRODYNAMICS

due to the existence of the electron-positron vacuum, coincides with


the energy density of the vacuum electrons, i.e., of electrons filling
the negative energy levels, after subtracting their potential energy
in the external electric field. In other words, if (A, i<p) is the four-po­
tential of the electromagnetic field and is the set of wave functions
describing the states of a vacuum electron in such a field

i ~ W ~ = ( “ ’ \ v + e A ^v>\ r )+ ( P m + e(p W v~)’

then

w' = ~ ^ W - '*. (54.31)


V \ I p

where (ip, x ) = WxXx is the spinor index) and the summation is taken
over all the individual states of the vacuum electron p (the wave func­
tions i p are, naturally, assumed to be normalized per unit volume).
The first term in this expression represents the total energy density
of the vacuum electrons, which we denote by

(5 4 J2 )

We now show that the second term—the potential energy density


of the vacuum electrons—is related to the first expression by

^ l W - ’* . w l , - ,) = - ^ ' <54-33)

where E is the electric field intensity.


We first note that if the quantum mechanical energy operator H
depends on a parameter X, then as a result of an infinitesimal adiabatic
change of X by dX the energy eigenvalue ep = Hpp undergoes a change
equal to
dH
dep dX.
I jX p p

This relation is obtained immediately if we differentiate the equation

/ (y*(A), (H (A )-£„(A))yp(A)) dr = 0
with respect to X.
RADIATION CORRECTIONS 783

In particular, if H — H0-\-Xe<p, (<p does not depend on A), then


dH
K * P ) PP = * w
pp
In other words, if the scalar potential contains a constant factor A,
then the following relation holds:

J-f (.Vp }*. e q n p ^d r = A j wmdr.

where the integration is taken over the whole volume. In the case of
constant fields in which we are interested we can regard the electric
field E as the factor A and, moreover, we can go over from the total
energy to the energy density, i.e., we can omit the integral sign; we
thus arrive at relation (54.33).
It follows from (54.31) that

w ' = w m- E ^ - (54.34)

On comparing this equation with w = E (d L /d E )—L , we conclude


that the additional term L x in the Lagrangian density of the electro­
magnetic field compared to the classical Lagrangian L 0 = \ ( E 2—H 2)
agrees, except for its sign, with the total energy density of the electron-
positron vacuum in the presence of an external field:
L x = - w m, (54.35)
Since, on the other hand, L x is a function of only the two inde­
pendent invariants E 2—H 2 and (EH)2, the problem can be reduced
to finding the energy wm in special fields, viz., it is sufficient to obtain
the value of wm in a constant and homogeneous magnetic field H(0, 0, H )
and an electrostatic field parallel to it with the potential
<p = <p0eigx+(p^e-igI, (54.36)
where <p0 and g are constants (for small values of g we obtain a homo­
geneous field of intensity E = ig (cpQ—cpt) ■
This choice of electric field eliminates the difficulty associated
with the possibility of pair production in a constant arbitrarily weak
electric field extending over all space. It is clear that if the difference
of potentials of the form (54.36) between any two points does not exceed
the quantity 2me2, then the field (54.36) will not produce any pairs,
784 QUANTUM ELECTRODYNAMICS

We obtain first of all the value of wm in a constant and homogen­


eous magnetic field H(0, 0, H). In future we shall denote this quan­
tity by
In accordance with subsection 12.5, the possible values of the energy
of a vacuum electron are equal to

Enn(Pz) = ~ ] / ( m 2+ e H ( 2 n - p - \ - l ) + p l ) ,
where n — 0, 1, 2, ..., p. = ± 1 and pz is the component of the electron
momentum along the magnetic field; the corresponding wave functions
have the form (cf. 12.13))
l ( ~ £n S P z ) + m ) ®n(f)^
e‘V + ’V 10
p = 1,
V i z ( p z) [ e n ( p z) - m ) \ p * V n(*)’
\ - i ( 2 e H j v n_x(S)

e>pyy+ipzz
1° \
V*»lT>v(r) = — — p = — 1,
Y 2 £ nM( P z ) (£ n f i ( P z ) ~ m )
i(2eH(n+X)f v n+1(£)
\ — PzVn(£)
where
Py
( = }/ e H \ x + .e H

They satisfy the normalization conditions


J f wT n^~>*
p p y y z \( rY ) nw' ^pv ' \ p ypz (p)
\ dr = <
u5n n ’, 6u

f v U ? P z ( r ) W n( pTy pz i r ) d P y = eH

(it is assumed that the normalizing volume is equal to unity). There­


fore, the energy density of the vacuum can be written in the form

wm( 0 ) _ _ _
dPy dPz
K i p % ( 0 . v U vPa(r ))en „(/>*)
w -

(54.37)
^=±1 n= 0
where p = p z.
RADIATION CORRECTIONS 785

On noting that en_- y(p) = £n+1 +1(p), we rewrite (54.37) in the


form

LI-(O)
= ~ f ^ £oOO+ 2 J )e „ (p )} , (54.37')

where
sn(p) = I P '+ m '+ l e H n , n = 0 ,1 ,2 ,....
On using Euler’s summation formula

1 N
\ m + y F(kb) — l-F(bN)

i b N oo
1 I C B bZm \
= J j J F(x)dx+ 2 [F<2m~l)(bN) - / • ’(2m- 1)(0)]f , (54.38)

where B are the Bernoulli numbers: B, = - , B0 = ---- —,


1 6 2 30
B3 = — , . . . , and on setting b = 2eH in this formula, we obtain

r ^ oo

H’<°> =
-w J * / (2m ) ! o)

where
F(x) = |' p2-ir m2 x .

We must now regularize the quantity H’,(n0). This regularization can


obviously be carried out in the following manner. We must, first of
OO

all, neglect the integral J F(x)dx since it does not contain the magnetic
o
field intensity and therefore represents the energy of the free vacuum
electrons, and, moreover, we must neglect the first term with m = 1
in the infinite sum since it leads to an energy proportional to / / 2, which
has already been included in the unperturbed field energy w0. In neg­
lecting this term we are essentially carrying out a renormalization
of the field intensity associated with the renormalization of the charge.
786 QUANTUM ELECTRODYNAMICS

Thus, the regularized expression for the energy density wi°) has
the form

y "-n(o)dp
OO OO

(
w.mR
0 )
= —- - f f
0 ) “I (2")!
On noting that
F(2n-1) ( O ) ^ 2" 2 O
4 n -3
2 )/7r (/?2+ m 2) 2
and that
oc

/ _ dp___
(/>2+ m 2)5

we obtain
r ( 2n - 2 )
Hi°> = v z,2njgn(m2)2- 2n_ (^ r f (54.39)
y”R 2( n f ^
and since
r ( 2 n — 2 ) = J drie~ny ,2n- 3

we have
. 271-1
Z?m2 f drj \r\ 22nB n I br\
f(°l
Wv A= -I'll— I
mR 16tt2J
a0 «2
•1
/
71= 2 (2«)! \ 2 m2 /
On comparing this expansion with the expansion

c o t h x =x I++ 3- ++ £ i ,2 (2A:)! ’
(54.39')
fc—
we finally obtain
OO
t?2/ /* 2
H,(°> =
= — I (?;//* c o th (rjH*) — 1 (54.40)
87r2 J i f (
where
b _ eH
H* =
2m2 m2
Now in addition to the magnetic field we also turn our attention
to the electrostatic field parallel to it, the potential for which is given
by formula (54.36).
RADIATION CORRECTIONS 787

We expand vvm into a series in powers of E 2 or, what amounts to


the same thing, in powers of <p2:
CO

(54.41)
k= 1

It can be easily shown by repeating the arguments which lead to


(54.37) that
oo 00
eH
'm 2
/' = ±1 n= 0
H / dPen ^ { p ) , (54.41')

where e(2k)(p) is the correction to the electron energy eTl/1(p) in the


magnetic field proportional to E 2k.
We first of all consider w^2). By using perturbation theory we can
show that
OO

f ipeisiP ) = — w - r - x / ( p S . K ’- r * '
— oo

where
K 2 = m 2j\-eH{2n—(i,Jr \ ) ,

and where we have expanded q> into a series in powers of g and have
kept only the second order terms.
On setting in Euler’s formula (54.38)
2 g2 ___m2+ x __ j
F(x) = —e2\<p0
4 ( p 2+ m 2+ x ) 5/2

we obtain

1 e2E 2 C dx d2k~x 1 \ ]
Hm2
,( ) =

^r2" 12 Jo m2+x 2 1 (2/c)! dx2^ 1 m 2+ x / x„ 0J

where E is the intensity of the electric field equal to E 2 = 2g2|9?0|2.


Regularization of this expression reduces to neglecting the first
term containing the integral

dx
/ m 2Jr X
788 QUANTUM ELECTRODYNAMICS

and proportional to E2. By neglecting this term we, in fact, carry out
a renormalization of the charge, just as we have done earlier in evaluat­
ing the energy \v{^ ) in the magnetic field.
By utilizing the expansion (54.39') for coth x we finally obtain
the following expression for the regularized value of w{2):

= { f» * ■ (54-42>
0
where E* = 2E/m2.
We now consider further terms in (54.41).
The assertion can be made that for small g the quantity e $ has
the form
4 T
J ( P ) = g Te T \<Po\ tG t ,

where Gr is a function of K and P. Therefore


co °o qo oo

dx j GTdp— ^ b2n
< ) = ^ 1 TeT\ ^ r '■ r " n »1
and

2k dx Gzkdp —-y ^ g2ke2k\<p012k


Wm = 8n 2 <Po!
k - -2 k~2
oo cxj
.,(0) (54.43)
Tl—1 —ro
But the integral of Gr over p must have the dimensions of the (2r—2)th
power of the momentum and, therefore, can be only of the form

(54.44)
— CO

where f T are dimensionless constants.


In order to determine f r we consider in (54.43) the term proportional
to E2 and not containing H:
OO oo

fr
dx GTdp
( r — 2 ) m 2r i '
0 —
RADIATION CORRECTIONS 789

We compare it with the term proportional to H r in expression


(54.39). Since um is a function only of E 2— H 2 and ( E H ) 2, the coef­
ficient of E r must differ from the coefficient of H r by the factor (—l)r/2.
We already know this latter coefficient which is equal to (2e)2TBrmi ^ir
r ( 2 r —2)/2(2r)\, so that

hr ( ) (2r)i B,.

We now evaluate the first term in (54.43):

w » 2 “j j - i ) v r -r % )P -

= ^ |, £ * » .„ * • - ! + (54.45)
0 * ’
Further, we consider the second term in (54.43). On noting that
A 1 _ f w y -x-zn r(2n+k-2 )
dx 2"-1 (w 2+ x ) /£- 1 ix=0 ^ ' (A:- 2 ) ! ’

we obtain
1 OO OO
D rf271” 1
8tt2
k = 2 n- 1 (2«)! * <Fo
12*
/ c-* L
— oc

i oo oo n n

J T 6 2n~ , ( ^ ) 2fc (2^ ( ~ l ) ' f+ V 2)2^ A 2 ^ + 2 ^ - 2 )


fc = 2 n = 1

mi j £*(— 0 fc+1
m- ^ r i 2k(eE)2k
8ti2 (2/c)!
f c=2

. OO

X
n = 1

/ 2 /7*2 \
X e - v i r i E * cot t] E * — 1 -f- ——r— )(??//* coth ??//* — 1). (54.45')

Therefore

y ww = _ ^ F co t^ £ * -l + ^ — J?7//* c o t h r j H * .
k= 2 71 J V I
790 QUANTUM ELECTRODYNAMICS

On adding this expression to expressions (54.42) and (54.40) for w ^


and w<0), we obtain
rrf
w
•'n f — e v{rjE* cotr]E*r]H* coth.rjH* — l-\-^r)2(E*2—H*2)}.
JJ rf
n
0
(54.46)
On noting that
n cos]/B2—a2+ 2iaB+ co s 1/ B2—a2—2iaB
cot a coth B = 7----------------------- *' -----
cos |//32—a2+2/a/5—cos |//?2—a.2—2iafi
and on utilizing (54.35), we obtain the final expression1 for L x

L = J 2_ f e
1 S n 2 J rj3
0
. cos(rj }/E*2- H * 2+2i(E*H*))+cos(r]]/E*2- H * 2-2J(E*H*))
X
ir> cos {rj \ / E*2 H* 2+ 2i(E*H*)) - cos {?])) e *2- H * 2- 2 i(E*H*))
m n*
- e2
s- + — (54.47)
3 ( » 2~ £ 2) .
We now consider some limiting cases.
In the case of weak fields
,2
L1 = 2 / 1
[(E2- H 2)2+ 1(EH)2\ + .... (54.47')
45 \ 4n: J m4
This expression coincides with the one obtained previously (cf. (54.27)).
We now consider strong fields. We first take E* = 0 and H* >> 1.
In this case we write L x in the form

- /( * ) d x .

where
£= < 1
H*
and
m4 H*2 I 1
/( * ) = x coth .v—1—-- x2
8n 2
1 This expression was first obtained by Heisenberg and Euler (85). The deriva­
tion given here is due to Weisskopf (205); cf. also the work of Schwinger (171),
(172).
RADIATION CORRECTIONS 791

The function f{x) obviously tends to zero as * -*■ 0, and for x: > 1
behaves lik e /(x )= /(o o )+0(1/*), 1, where/(oo) = (e2/247r2)m4//* 2.
We now show that for functions having such behavior the following
relation holds asymptotically for e -»■ 0:
OO

J ( s ) = J " ~ —/(■*)dx «ss/(oo)ln—. (54.58)


o
In order to do this we utilize the identity
OO OO

§ ^ -f{ x )d x = ^ ^-f{x)d x
0 0
co cc
p—£2 p—£ /"*p—6^ p—x
/ ------ ----- [/(*)-/(« > )]dx + /( o o ) J ------ dx.
0 0
The first term in this expression represents a number which does not
depend on e; in the second term we can make the transition to the
limit e -> 0 as a result of which we again obtain a number independent
of e, while the third term is equal to
OO

/(oo) J - — ~ — dx = - / ( o o ) In £,
0
as can be verified by differentiation with respect to e. This term is the
principal term in the integral J(e) for e 0.‘
By utilizing these results we obtain the asymptotic expression for
L x in the case H* 5> 1, E* = 0
ynd 1
£ , = - 5 - ^ 8 ^ - / / * * Ini/*, H * > 1. (54.49)
&71* 3
Similarly it can be shown that in the case of strong electric fields
E* > 1, H* = 0 the function L x has the form

Lx= — e2-^-E*2 ln£*. (54.49')


0712 3
Thus the ratio of the added term in the Lagrangian Lx to its classical
value L 0 = i (E 2—H 2) increases at high field intensities only logarith­
mically with the field
Li a el FI
In——1- (54.50)
u 3ti m“
where F is either the electric or the magnetic field. Therefore, the
792 QUANTUM ELECTRODYNAMICS

non-linearity of the equations of electrodynamics represents a small


correction even in the case when the field is considerably greater
than the “critical” value mzc3/eh. If for E we take the field at the
“electron boundary” equal to
e
E= An-
Anri

then the ratio \L1/L0\ will be equal to


Lx 1 In 137
L0 3n 137
i.e., also in this case we have
L
~LL < 1.

§ 55. Photon Scattering by the Coulomb Field of a Nucleus

55.1. General Expression for the Cross Section for Photon Scattering
by a Constant Electromagnetic Field
Four electromagnetic field operators appear following the normal
product symbol in the scattering matrix Sl4) which describes non-linear
interactions between electromagnetic fields (cf. formula (54.1)). If
all these operators refer to photon states, then the matrix element
of S(4) will describe the photon-photon scattering process discussed
in the preceding section. However, we can also have two other cases,
viz., when only two of the factors A (x) refer to photon states, while
the other two refer to the external electromagnetic field, or when three
of the factors A^(x) refer to photon states, and one refers to the external
field. In the former case the element of the matrix S(4) describes the
scattering of a photon by an external electromagnetic field (82), (6), (105),
while in the latter case it describes the splitting of one photon into two,
or the coalescence of two photons into one in the external field. These
effects (together with photon-photon scattering) exhaust all the possibil­
ities for fourth order processes of interaction between electromagnetic
fields.
We confine ourselves to a discussion of the scattering of a photon
by an external electromagnetic field. If the field depends on the time,
RADIATION CORRECTIONS 793

then photon scattering will occur with a change in frequency; however,


if the field is constant, then scattering occurs without a change in photon
frequency (coherent scattering).
Of particular interest is the coherent scattering of photons by the
Coulomb field of a nucleus. The cross section for such scattering becomes
comparable with the cross section for the scattering of photons by
electrons at energies of the order of 1010 eV, but coherent scattering of
photons by nuclei can be detected at much lower energies due to its
characteristic angular distribution which has a sharp maximum in
the region of small scattering angles; we shall see later that in the case

Fig. 115.

of heavy nuclei and at a photon energy of co = 300 MeV the differential


cross section for such scattering at angles of-—'0.01° exceeds by three
orders of magnitude the corresponding cross section for Compton
scattering.
Figure 115 gives three diagrams representing the scattering of a photon
by an external field. The matrix element which determines this process
is given in accordance with the rules of § 25 by

^(c^coa)1^
J diq1diq2d(kl - k 2+ q 1^ q 2)elfie2v

A{e) 0/i) A[e) (q2)


X J d 4p ( S p ^ O p + m) 1y M p - ^ z ) + m) Va
(2tt)4 (2 j 4

x ( / ^ - / i 2+ h / i + w) hya0 p + iki+ m ) *] + Sp[yfXiP + m)


A A

X(i p + iq1+ m ) - 1yy(ip + iq1- i k 2+ m ) - 1ya(ip-hik1-hfn)-1]


+ S p [ y J i p + m ) - 1y J i p + iq1+ m ) - 1y x(i_p-l-iq+iq1+ m ) - 1y v
X (ip-\-ik1-\-ni)~1] },
794 QUANTUM ELECTRODYNAMICS

where A f ](q) is the Fourier component of the external field

A f \ q ) = f A ^ ( x ) e - i<1Idix,

while kx and k 2 are the photon momenta and ex and e2 are the photon
polarizations.
By utilizing the definition (54.5) of the photon-photon scattering
tensor I vXa{kXi k 2, k 2, k 4) we can rewrite this expression (after regular­
ization) in the form

S£f= 2 \ 47i I y co1a)2J


f diq ^ q 2( 2 n y d { k 1- k 2+ q xA-q2)

A ^iq ^A ^iq z)
2v~ (2n Y {kx, ~ k 2, q x, q 2). (55.1)
(2n)* luvlo

If the field is constant then Aj f f q) = Aj f )(q) (2ri)d(q0), the frequencies


of the incident and the scattered photons are equal, co1 = co2 = co, and

cm i ( e 2 Y 1 f j „ Aie)(q) A lae>(k2- k 1—q),n^ iX, ^ _ ^


5“ ' = " 2 te * f --------M{2j i y---------
(2 Tiy
* J dq- Q

X ^ e2vI ^ a(k i ’~ k 2, k 2- k x—q). (55.2)


On multiplying l ^ j ^ l 2 by the number of final states of the scattered
photon a>ldcL>2do2/(2ji)3, where do2 is the element of solid angle containing
k 2, and on dividing this by the flux density of incident photons 7 = 1 ,
we obtain the differential cross section for the coherent scattering of
a photon by a constant field
A\e)(q) A {ae)(k2 k x—q)
(2 tc)3 (2jt)3 eZv

Xl/tvioiki’—ki, q , k 2- k x- q ) *. (55.3)

55.2. Relation between the Forward Scattering Amplitude for a Photon


and Pair-Production by a Photon in the Field of a Nucleus
We shall now show that the matrix element for the coherent forward
scattering of a photon by a constant electromagnetic field is related
by a simple expression to the total cross section for pair production
by a photon in the same field.
RADIATION CORRECTIONS 795

We utilize the unitary nature of the scattering matrix


SS+ = l + e (S (1) + S (1)+) + e 2(S(2>+ S(2)+ + S(1)S(1)+)
+ e3(S(3, + S (3) + + S(1)S(2) + + S(2)S(1)+)
+ e4(S(4) + S<4, + + S(1)S(3) + + S(3)S<1, + + S (2)S<2,+) + ... = 1.
From this it follows that each of the expressions in brackets is equal
to zero, in particular
S<2)S<2) + + S (4) + + S (4) + S (1)S (3) + + S (3)S (1)+ = Q. (55.4)

We evaluate the matrix element of this expression corresponding


to coherent forward scattering of a photon k by an arbitrary constant
field
{k | S(4) + S(4)+1 k) = ~ { k | S(2)S(2)+|k ) - ( k \ S(1)S(3>+ j k)
—{k | S(3) S<1)+1k). (55.4')
The terms appearing on the right hand side of this equation can be written
in the form
(k | S(2) S<2)+|k) = V (k | S<2>| f ) ( / | S(2>+1 k),
r

{k| s (v s (3^ | k) = 2 (k| s (i) |/) ( / | s < 3 ) + 1k),


r
( k | S(3) S(1, + |A:) = ^ (A: JS<3) |/ ) ( / | S(1)+1 k),
r
where f denotes the final states of the matrix elements of S(1), S(2), S(3).
Obviously, only electron-positron pair states can be used as such states;
but the matrices S(1) and S(3) have no elements connecting single photon
states with states of the field containing one pair, so that the second
and the third terms in (55.4') vanish and, consequently,
(k | S<2) S<2)+ \ k ) = ^ (k \ S(2) | / ) ( / | S(Z)+| k) = —2Re(fc| S<4) | k). (55.5)
/
It can be easily shown that the left hand side of this equation gives,
up to a numerical factor, the total cross section a± (co) for pair production
by the photon k in the field under discussion. Indeed, on writing explicitly
the (5-functions in the matrix elements, i.e., on setting
(fl|S “ » | 6 ) = (fl!M «)|6)d(e0- e 6),
we can in accordance with § 26 write
796 QUANTUM ELECTRODYNAMICS

Therefore, it follows from (55.5) that

o±(co)= — —Re (A: | M (4) | k ) . (55.6)


71

Thus, the real part of the matrix element for the forward scattering
of a photon by a constant electromagnetic field is given, up to a numerical
factor, by the cross section for pair production by the photon in the
same field (98).
Equation (55.6) follows from the more general relation connecting
the total cross section for the scattering of particles by a force field
with the amplitude for their elastic forward scattering. This relation,
known as the optical theorem, has in accordance with (5.13) the form
at = 4n%lmf(co,Q>), (55.7)
where at is the total particle scattering cross section given by the sum of
the elastic scattering cross section (ae) and the absorption cross section
(cra), at = ae+ a a,/(fo , 0) is the amplitude for elastic scattering by
an angle 0, 1 is the wavelength of the particles, and co is their energy.
In the case of photons moving in a constant electromagnetic field
absorption is associated with the conversion of photons into pairs,
so that the cross section for photon absorption is equal to the cross
section for pair production by photons. The elastic scattering cross
section ae for the photons is considerably smaller than the pair production
cross section, and therefore
at ^ a ±(k). (55.7')
The amplitude /(co, 0) for the scattering of photons by a constant field
is related to the matrix element for scattering S f ^ r = M(co, 0) <5(co—cof)
by the expression

/(£ 0 ’ 0) = ~ " ( 2 i y £oM(" ’ 0 ) • (55-8)

Indeed, the differential scattering cross section has the form

do = 7 ^ 4 - 1M(co, 0) |2co2 d o . (55.9)

On the other hand, in accordance with the definition of the scattering


amplitude it is given by
da = |/(co, 6) |2 do. (55.9')
RADIATION CORRECTIONS 797

A comparison of these expressions yields the expression (55.8) up


to a phase factor. We define the scattering amplitude completely by
choosing this factor to be equal to —i; in this case both formula
(55.9) and expression (55.7) will be valid.

55.3. Momentum Distribution of Recoil Nuclei Accompanying Pair


Production by a Photon in the Field of a Nucleus
We now obtain the matrix element for coherent forward scattering
of photons and the total cross section for pair production related to it.
In the case of forward scattering the photon polarization is obviously
unchanged, e1 = e2, and, therefore, the value of the matrix element
averaged over the polarization states is in accordance with (55.2) equal
to

A/(4)(co, 0)

_J_ w _r A e)(q) A ff-q )


2\An 2oj J H (In)2 (2n f
,q ’ ^ k’
(55.10)
In accordance with (55.8) the forward scattering amplitude is equal
to

- k ’ q ’ ~ q y

(55.11)
On the basis of this expression and of expression (55.7) we can
write the general expression for the total cross section for the production
of pairs by a photon in the field A f f q )
o±(oj) = 4 n% Im/(a>, 0)
e
2 4 71 ) 3^ 7 Im \ ~ k>q>-q)
or
o±(oj) = J A f )(q)Ale)( - q ) T fiy(q, lc) dq, (55.12)
where
4 n I e2 \ 2
r , v(f . « = ( 2 ^ - - y R e [ ^ „ ( f , k ) + B ^ v{q, Q + B ^ X - q , k)].
798 QUANTUM ELECTRODYNAMICS

Sp[y„ {ip- m j y ^ K p —k ) - m ) y v(i(p- k - liJ -


{p2'+m2) [(p—k)2'+m2] [(p —k —q f + m 2] [(p—q)2+ m 2]

B „ ( i , k ) = j d ‘p

w Sp[y^{ip + i q - m ) y v(ip—m)yx( i p ^ i k - m ) y x(Jp—m)]


' [ { p + q f + m 2] [{p + k)2~+m2)(p2+ m 2Y ^‘ }

Since q is the momentum transferred to the nucleus as a result of


pair production, the integrand in formula (55.12) defines the momentum
distribution of the recoil nuclei.
We now obtain the tensor TMV(q,k). We first note that the cross
section for pair production must be invariant under gauge transforma­
tions, i.e., it must vanish if A ifie)(x) = dx(x)/dxM, where x(x) is an
arbitrary scalar function. From this it follows that

qMTMV(q,k) = 0. (55.13)

On the other hand, it is clear that the tensor Tflv(q, k) must have the
form
An / e2
TflM , k ) = (2n) 6----
co \ An ]
Re[Il dMV+ I 2kMk v+ I 3qMqv+ Ii (kMqv+ qr.k,,)\, (55.14)

where are invariant functions of q and k.


It follows from (55.13) that

k + h f + h q k = 0,
(55.14')
I 2 q k + 1\ q 2 = 0 ,

so that in order to obtain T^v it is sufficient to determine only two of


the four functions Ii.
We write /, in the form

7 i(? > k) = k) + Bt(q, k) + Bi{ - q , k), (55.15)


RADIATION CORRECTIONS 799

where A t and Bx are the contributions to /, made by and B In


accordance with (55.14) it is sufficient to obtain only the real parts of
these quantities.
We first of all determine 5,.
On evaluating in (55.12') the trace of the matrix product in accordance
with the formulas of § 26, and on utilizing the first of relations

i i
a bc * = 3 '^ dx^ d y 'y [ a + (b - a )x + ( c -b ) y \
0 0
4,
(55.16)
l r v

abed
= 3! J d x
0
j
0
d y j dz[a
0
+(6—c)x+ — +(c d)y (d— c)z]~\

we obtain
1 x

Bz(q, k) = -4 8 j d*p j d x J dy(x—y)y


0 0

p 1((1 + 4 ( x —y ))+2 [(x—y) (Q 2+ 3m2—2Qk ) + Q 2 + m 2]


X -
\p2+ m 2Jrq2{ \ —x) — Q2Y
(55.17)
1 X
B3(q, k ) = - 4 8 J d*p j dx j d y ( \ - x ) y
o o

p2( l - 4 x ) - 2 x ( Q 2+3m2- 2 Q k )
X
~ [p2+ m 2+ q 2(Y ^ x ) - <22]4
where
Q = q ( \ —x ) + k ( x —y).
Further, on noting that

f j4 1 rfi
J P (p 2~ + W ~ 6a* ’

f P2 _ nH
J P (p2+ a 2y 3a2 ’
800 QUANTUM ELECTRODYNAMICS

and on setting x — y = rj, 1—x = £ we rewrite (55.17) in the form


i i-f
1+4 rj
B 2 ( q , k ) = —16tt2i f d £ f d i j —
m2-\-qz£—Q2
0 0
rj (Q2+ 3m2—2qk £) + Q2+ m2
(m2+ q 2l;— Q2)2
(55.17')
i M
4+3
B 3 ( q , k ) = —16tez/ f <7£ f dr} £(! —£—??)
m2-\-q2£—Q2
0 0
( \ - 0 ( Q 2+ i ™ 2- 2 q k £ )
(m2 : q2tj- Q2)2
where Q = qt+krj. Integration over rj yields

B 2 ( q , k) = — I6n2i d£' _ ¥■H1


1 _ « [m*+q2H l — £)]z
pa£3

\m2+ Q 2t n _ t \ a a
+ 3 + 2 (l-£ )(4 m 2+ 4 r2£)
£ 4£4
m2+ q 2£(\ — £) [m ^ ^ W )]2
X - J - ^ - - - 1 3 ( « z+ ^ ) ^4"

( 1 - 0 /T 3 1 Z t4 T / T 2 1 +
1----^ r ( 2 m a+ ? ^ ) - 2 ( 2 m 2+ ? ^ ) ------- ^ -------

X ln(m 2+ 4 r z£(l — £))> + purely im aginary terms,

(55.18)
i
B 3(q, k) = - i — J ~ ^ { ( 3 £ - 2 ) [m2+ 4 r2£ ( l- £ )]
0
+ 4 (1 —£) (mz+ r 2£)}{In [4rz£ (1 —£)] —In [m2+ q 2H 1 -f)]} ,
where /? = 2g/c, 4r2 = (q - k )2.
We consider the quantity B3 (q, k). Since in the case of a constant
field q2 7^ 0, In (m2+ 4 rz£(l —£)) will be real, and, therefore, the whole
term proportional to this logarithm can be omitted.
RADIATION CORRECTIONS 801

The term containing In (m2+ 4 r2f(l —£)) gives a contribution to


BA( q , k ) different from zero only for r2 < —m2. In this case
m2+ 4 r2f (1 —|) has two real roots within the range of integration (0,1)

and since in2 contains an infinitesimal negative imaginary part, we have


ln |ra 2+ 4 r 2f ( l ~ f ) | , 0 < f < f x,
In [m2-r4r2f (1 —f)] ln |m 2+ 4 r zf ( l ~ f ) | —z' t t , f, < f < f 2,
ln[m 2+ 4 r 2f(l — f ) |, f 2 < f < 1•
Therefore
i
Im | -— /(£ ) In [m2+ 4r2f( l —f)] = —m J — / ( f ) ,
o f,

where / ( f ) has no singularities within the interval (0,1). By utilizing


this formula we can easily show that

32jr2m2
R eB(q, k) — In (55.19)

y r

In a similar manner we can obtain ReB2(q,k) and Re A^ q j k ) ;

/ m2
1
16rr3 ■ y
R e 5 2(^, A:) = q 2( \ 2 m 2q 2 — % m 2 —fi2) In
/. m2
- y > + ^

w
_ I / 1 -f- / |p ' ! 2 /M m ’ - q ‘)- 2 ^ '(S m I r ) ]
T2

(55.19')

ReA.,(q, k) = A'2(q, k)-\-A2(—q, k),


R e A 3(q, k) = A'3{q, k) + A'3( - q , k).
802 QUANTUM ELECTRODYNAMICS

1677 ^ / nfl
A'2(q,k) = - j 2- [q*- ? + 2m2] j / 1 + ^ -

m2
1-
16tt3 r-
A'z(q, k ) = 2(2m2-\-q*)q2(l3—qz-\-2m2) In —
1+

m
/
+ 1 / 1 + —r- [/53+/S2(2mz—^2)+ 6 ^ 2(2m2+ ^ 2) (?2—£)]
1
(55.20)

On substituting these expressions into (55.15), and on utilizing the


fact that is real, we finally obtain

Ke ' 2 [2(4m4+ 6 m 2q 2 — q *)+2 ( q 2 — 2 m 2) {5— /?2]

1 - 1\j/ H —r2
X In

]/>+^ K
f
/ m2
+ 1 / 1 + - T [4(92-i3 ) ( m ^ 2) - / ? 2]

(55.21)
m-
1- 1 / 1
32jt3 v7 . r- /rr
Re /, = - 2m2 In + (/? - » ' ) ] / 1
£2 mc
1+ 1/ 1

if r2 — Htf —k)2 < —m2 and I{ = 0, if r'2 > —m2.


We now consider in greater detail the particularly interesting case
when the field in which the pair is produced is a screened Coulomb
field
RADIATION CORRECTIONS 803

where F (q) is the atomic form-factor. In this case only one component
of the tensor 7^44 which is equal to

An I e2 \ 2
^ = ( 2 ? trV U ) R e ( - / , + / , CO2)

differs from zero and

<r±(co) = - f ^ R e C A - / ^ 2), (55.22)


2 co(2 jt)7 .

where
R e ^ - w 2^ )

wr
n, /
87T3 2~ mz


- y i+ -
Aq2 2w2 in ----- + 1/ l + ^ - ( i 9 - 9 2)
= ~ F
1+ 1 / 1
mz /■

/l2— Aq2co2

, m2

V
! + |/ 1
1 H— j-

mz (/52+ 2 (/5 - 92)(2m2- 9 2) - 8 m 2(92+ m

(55.22')

if r2 < —m2, and R e ^ —<u2/ 2) = 0, if r2 ^ - m 2.


On introducing the new variables
1 co
2 m Q = \q \, fi = (co2- \ q ~ k \ ) ^ z ^> y =
Am2 2m

we can write (55.22) in the form


y + / y a—l

cr±(co) = J d Q P ( Q ,c o ) , (55.23)
y— Vya- i
804 QUANTUM ELECTRODYNAMICS

where

m , a.) = [1 (55.23')

/ ( e ,i ' ) = A + ( i - 2 e 2) A + ( 2 e 4- e z- 4 e v - i ) ^ + 2 e 2/

X (1 +6<22—4(24)74+A'0+ ( l 4<22) ^ i + [(4Q2—1 ) 2 z 4Q2/y2\ Kz

+ 4Q2y%SQ2- l ) K :i+4Q*y2( l - & Q 2)K l

and

The function P( Q, w) determines the distribution of recoil nuclei


with respect to the absolute values of the momenta (98).

Fig. 116.

Figure 116 shows the momentum distribution of recoil nuclei for


different values of the photon energy. The quantity P/(Z2a3IZlnhrP)
is plotted along the vertical axes of these graphs (the letter A denotes
the asymptotic curve).
RADIATION CORRECTIONS 805

Formulas (55.23) and (55.23') are very complicated. Approximately


we can assume (17) that the function P(Q,co) is inversely proportional
to q between the limits

a
qm i n m
m2
co

and falls off very rapidly outside this interval (a is the atomic radius
in the Thomas-Fermi model). In other words
Q
m ax

ff,((0)=A f i
J q
^min
where A is a constant.
If screening can be neglected, then the lower limit of integration
is equal to mz/co and
m

&

A comparison of formula with formula (32.9) which determines the


cross section for pair production by a photon in the Coulomb field
of a nucleus shows that

A = ™ ( ± X —
9 \An m2 = —9 0 .

Thus, the momentum distribution of recoil nuclei in the absence


of screening is given approximately by

da (55.24)
<o

In the case of complete screening this formula is replaced by


806 QUANTUM ELECTRODYNAMICS

55.4. Angular Distribution o f Recoil Nuclei and Total Cross Section


for Pair Production by a Photon in the Coulomb Field o f a Nucleus
It is well known that when the following condition is satisfied

^0 m ax a>

(cf. subsection 29.6) screening of the nuclear field plays no essential


role, and the quantity F(q) can be neglected in (55.23'). We determine
the angular distribution of recoil nuclei in this case, i.e., their distri­
bution with respect to the angle 8 between the vectors q and k.
On introducing the variables

ri = y cos 8 = y ---- , where y = ---- -


qco 2m

we obtain for F(q) = 0

4rf 2r)Q

(55.25)

This expression can be written in the form


i
arcc o s —
y
= / d8P(8,co), (55.26)
o
RADIATION CORRECTIONS 807

where

P (e -m) = s in 6 1 ‘<“ ( , » ^ j w « { ( 1 - } ) [ ( ' - £ '


0

2 rf 2{rf— 1) r f—4 ,y
X
/«3 + + V / + r 2? - * + ?)i?

1/2

i-M
+ l n ----
m _r _ r2 2 f \ ( 1 , l-8)j2
r \ i/z l t^2/zz 2.rfn r f J \ 2r f A rfu

I6774—16??2—1 2?72H-1 r/2


(55.26')
rffji2

This is the function which determines the angular distribution of


recoil nuclei (98). ’

Figure 117 shows the angular distribution of recoil nuclei for different
values of the photon energy.
The quantity P(0)/(Z2a3/647r3m2) is plotted along the vertical axes
of these graphs.
808 QUANTUM ELECTRODYNAMICS

On integrating by parts the term containing In {1 —(l—(l//^))1/2}/{l


-}-(l—(1 //w))1/2}“\ and on introducing the elliptic integrals
j i /2

F (x ) = j (1—x sin29?)~1/2 dtp.

n/2
E(x) = J (1—x sin 2<p)ll2d<p,

we can write a±(co) in the form (160)


y
Z 2<?6 L(y)
(T± (CO) = dr)+2L(y)
64jz3m2y2 >/ 7]

+
(64+109y2) £

125 + — + 42y2 )F
[(-*n (55.27)

55.5. Small Angle Coherent Scattering of Photons by the Field of


a Nucleus (23)
It follows from formulas (55.6), (55.8) that the imaginary part
of the forward scattering amplitude for a photon in the Coulomb field
of a nucleus is given by

a2(a>, 0) = -^-<7± (a)), to > 2m,

a2(co, 0) = 0, co < 2m.


This expression can also be rewritten in the form
a2(co, 0) = 0, co sC 2m,

a ' (“ - 0) = 4 r ( ^ H ^ 2^ w - ^ W ) + - 2-7^ [ - ( i o 9 + 6 4 / )

x £ '((1 —T 2) 1/2) ) + ( 4 2 + 1 2 5 y 2 + 6 4 y 4) F ( ( 1 — y2)112)] I , co>2m, ( 5 5 .2 8 )


RADIATION CORRECTIONS 809

where y = ___ and


(X)

cosh-1x 1
sinh-1 dx.
x yx

1/3/ sinh -i 1
xy

On utilizing relations (5.19) and the preceding expressions we can


obtain the real part of the forward scattering amplitude for the photon:

0l(ro, 0) =

- ( 6 7 - 6 ^ ) (1 - / ) F1(>>)] - ^ 1, (55.29)

where

i/y .
„ , „ ^ f arc sinx , 1 , ,
Cx(y) = Re I ----- :----- cosh M---- ) dx,
o ^

i/y 1
cosh
yx
AOO — R-e J ^ _ x2y/2 dx,

1iv 1/2
1—y 2x 2
&i(y) Re dx,
r~x2
0

1/3/
dx
Fi(y) Re
o
810 QUANTUM ELECTRODYNAMICS

We note that the functions E ^ p ) , F ^ p ) , E(p) and F(p) are inter­


related by
Flip) = E(p), p < 1,

« f)= f£ (-) + ^ ( l ^ ] , p > i .


(55.30)
F1(p) = F(p), p ^ \ ,

r > u
We consider several limiting cases. If oo < 2m, then the cross section
for pair production is equal to zero and, consequently, a2(a>, 0) = 0.
For co < 2m we have
Z2
/(co,0) = «,(<», 0) = — --I- — ) • (55.31)
m 471 m
The differential photon scattering cross section for photon 0 = 0
and co < 2m has the form

^ = l ^ . 0 ) | * = ( ^ ) 2a ) 2Z . ( ^ ) 6^ - ; (55.31')

it is proportional to co4.
For co > 2m we have

Z2 / e2 \ 3 7 / co
a^co, 0) =
m \ 4rt 18 \ m
(55.32)
Z2 co 2co
a2(co,0) = - — — -s- — ln —
m \ 4n / 9n \ m m

In this case the imaginary part of the photon scattering amplitude


exceeds the real part in order of magnitude by a factor of In (co/ra).
Figure 118 shows the dependence of the forward scattering ampli­
tude on photon energy.
Since the scattering amplitude increases with co, the case of partic­
ular interest is the scattering of high frequency photons by a nuclear
field. We therefore consider this case in greater detail.
In order to do this we expand the photon scattering amplitude
f(co, 0) = (co, 0) + ia2(co, 0)
RADIATION CORRECTIONS 811

into a series in terms of amplitudes corresponding to different values


of the angular momentum / and parity I. In the high energy case the
scattering amplitude for a given value of / can be assumed to be inde­
pendent of /. Therefore, the expansion of the photon scattering ampli­
tude f(a>, 0) has in the high energy case the same form as the expansion
for the scattering amplitude for a scalar particle.

Fig. 118.

We write this expansion in the form

/(co, 0 ) = - y ^ ( 2 / + l ) ( C / i<5'- l ) P ; (cos0), (55.33)


I

where <3, is the phase at infinity, P,(cos 0) are the Legendre polynom­
ials, and C, is the absolute value of the amplitude of the outgoing
wave of angular momentum I. In the case of purely elastic scattering
C, — 1; however, if absorption of particles is taking place, then C; < 1,
and the quantity = 1—Cf determines the probability of absorption
of particles of angular momentum /.
The cross section for particle absorption is determined by the
formula oo
cr, = j t P Z ( 2 l + l ) y i, (55.34)
1= 0
812 QUANTUM ELECTRODYNAMICS

and the elastic scattering cross section is given by the formula


71 OO
ae = f |/(0 )|22ti sin ddd = 7i%2 ^ (2 /+ l)|C ,e 2i'5' —1|2. (55.35)
o i=0
Therefore, the total scattering cross section is equal to
OO
at = ae+ oa = 27rA22 , ( 2 / + l ) ( l - R e C ;e2ii0 . (55.36)
1= 0

We note that a comparison of this formula with the forward scat­


tering amplitude

/ ( ffll0 ) = - - l 2 ( 2 1 | l ) ( C , e“ ‘- l )
i
leads to relation (55.7).
For high energies small angle scattering predominates; in this
case large values of / are important, and the sum (55.33), which at
these energies contains many terms, can be replaced by an integral
over / or, what amounts to the same thing, by an integral over the
impact parameter b = l % .
For I > 1 and 8 < 1 the following asymptotic expression holds:
g
Pt (cos 6) ps J0(bs), and s = = cod,

where J0(x) is a Bessel function; on substituting this expression into


(55.33), and on replacing the sum by an integral we obtain

tfjXco, s) = co I bdba{b, <o) J0(bs) ,


(55.37)
a2(co,s) = coJ b db(i(b, co)J0(bs),

where a(b, <x>)= al and (i(b, a>) = are related to C, and d, by the
expressions
a; ps 2(5;, and /S,ps 1- C ,s s |y ,. (55.37')
The quantity y (b, co) = 2/3(b, co) represents the probability of
pair production by a wave packet of photons of frequency co, passing
at a distance b from the nucleus. This quantity can be easily obtained
by using expressions (55.23), (55.23') for the momentum distribution
of recoil nuclei. Indeed, the momentum transferred to the nucleus
RADIATION CORRECTIONS 813

is approximately equal to q = 1jb, so that the momentum distribution


of recoil nuclei at the same time determines the pair production prob­
ability as a function of the impact parameters b.
On utilizing the approximate formula (55.24) for the momentum
distribution of recoil nuclei, and on setting in it q = 1jb, we obtain
i

28 Z 2 / e2 \ 3 QT db
a±{oi)
9 m 2 \ 4 n) J b
_i
m

On comparing this formula with (55.6) and (55.37) we obtain the


desired quantity (3(b,a)) in the absence of screening:

28 Z 2 e2 \ 3 1 c a)
(Hb,a>) b.
9 m2 4n j 4 nb2 b2 1 m2
(55.38)
CO
P(b,a>) = 0, b>

where
7 Z2
K \\
9 nm2 \ 4n

In the case of complete screening the quantity co/m2 must be replaced


by the atomic radius a. If b ><x>/m2, then (55.38) yields ( 3 = 0 , but
it is more correct to take (3 = Cm2.
For a given value of b the quantity (3(b, co) can be regarded in accord­
ance with (55.38) as a definite function of eo. If 1/m <.b < a , then
we have
c
co > bm 2.
(3{b, co) = ~b2 ’ (55.38')
.0, co < bm 2.

Further,
10 for b > a, co arbitrary,
(55.38")
= for b < 1/m, co > m.
814 QUANTUM ELECTRODYNAMICS

If we know the quantity fl(b, co) as a function of the frequency,


we are able to obtain a(b, co). Indeed, a(b, co) and fi(b, a>) are interre­
lated for a given value of b by the dispersion relations (cf. subsection 5.4)

(o
a(b, co) = — 9
71 J (co —CO)CO f
(55.39)
a(b, co')
P(b,a>) 5> Jf 74 *— C
71 (co'-co)co'
(CO

Therefore, on substituting into these expressions (55.38'), (55.38”),


we obtain
C co-j-bm2
a(b, co) — <b <a,
Tib2 n \co—bm2\’ m (55.40)
a(b, co) = 0, b > a.

We now evaluate a2(co, 6) and a^co, 6). On introducing the no­


tation t = b m and x = s/m we obtain in accordance with (55.37),
(55.38)
1 eo
a2(s, co) = Cco f t dt J0(xt) + f — J0(xt) (55.41)
0 1

where we have replaced the upper limit in the second integral by infinity,
although in actual fact it is equal to
tmax = co/m in the absence of screening,
tmax = am in the case of complete screening.
(This substitution leads to an error only for very small values of x,
smaller than l//max, i.e., for angles considerably smaller than (m/co)2
in the absence of screening, and considerably smaller than \/coa in
the case of complete screening).
On noting that
i
Ji(x)
J" td tJ 0(xt)
x
o
RADIATION CORRECTIONS 815

and

r t/w 2 a :2 a4
* < 1,
2933
I
where In (2/y) = 0.116, we obtain

a2((o, 6) = CojI ]n — + l — ... j = CojF2(x ).


\ yx ' 2 1 16 28-3 1 2U33
(55.41')
In accordance with (55.37), (55.40) the real part of the scattering
amplitude is equal to

' 1 °° -i
T t ,/ co+mt , f dt T . co-fmt
al (co, 6) = —m t d t J 0 ( x t ) In--------- + — / o ( * 0 In -j-------- j- .
71 ./ ov 7 co—mt\ J t |co—mt\ I
"0 1
(55.42)
For co > m we can replace the logarithm in this expression approxi­
mately by 2tm/a), ,in consequence of which we obtain

°o I
[
y j J0(u)du-\-2 J' J0(u)du(u2x~3—a:-1)
Cm 1 2 4 x2 xi
... \ = C m F 1{x). (55.42')
T r y + ~ 3T “ 2‘d 7
The total photon scattering cross section is obviously equal to

cr(co) = 2ji J|cq(co, 0)-Ma2(cc>, 0)|2 sin0 c/0 = cq + cq, (55.43)


o
where

m
cq = 2ji {Cm)2 xdx,
co
(55.43')
oo

o2 = 2ji(Cm)2 J F\{x)x dx
816 QUANTUM ELECTRODYNAMICS

and x = ajd/m. Since to >> m, the part of the cross section which is
associated with the imaginary part of the scattering amplitude is consid­
erably greater than that which is associated with the real part, cr2 > crr
In order to evaluate a1 and a2 we note that

J J0(xt)J0(xt')xdx = t 1d{t—/').

Therefore

Moreover,

2 f x dx
I FI (x) x dx
71 J x2~ 4n 2 In-*- ,

where x ± is the smallest value of x for which (55.41) is still valid, i.e.,
x 1 = 11am in the case of complete screening, and at = m/co in the
absence of screening.
On utilizing these formulas and the definitions (55.43) we finally
obtain (23)

98 Z4 / e2 \ 6
ff> = 2- ( Cm)2 = w ^ - U ) > <55-44>

— | cr. In — in absence of screening,


to m
mV (55'44,)
— o2 \n.am in case of complete screening.

Within the range of applicability of these formulas, i.e., for to > m,


the total scattering cross section is almost completely determined by
the absorption of photons associated with pair production. This part
of the cross section may be called the cross section for potential scattering
in analogy with the scattering of fast neutrons by absorbing nuclei,
while that part of the cross section which is associated with the real
part of the scattering amplitude may be referred to as the cross section
RADIATION CORRECTIONS 817

for dispersion scattering. The ratio of the cross section for potential
scattering a2 to the cross section for pair production cr±(a>) is equal to

a2 Cm2 e- 1
Z2
18 \ 4tt / , o)
2 In °- In —
m m

Figure 119 shows the functions Ff(x) and F|(x) which determine
the angular dependence of the differential cross sections for potential
and for dispersion scattering. For comparison we have also given the
angular dependence of the diffraction scattering of particles by an
opaque sphere of radius r0 = Ijm.
The absolute value of the cross section cr2 in the case of urinium
is approximately 6 mbarn, while the ratio a2/a±(a>) is approximately
equal to 1/8000.

Up to energies of the order of 1010 eV the total cross section for


coherent scattering of photons by nuclei remains smaller than the
cross section for Compton scattering. However, the differential cross
section for coherent small angle scattering of photons by nuclei may
be considerably greater than the differential cross section for Compton
818 QUANTUM ELECTRODYNAMICS

scattering. Indeed, the differential cross section for potential scattering


per unit solid angle is given in accordance with (55.43') by the formula

In the case of uranium we have

For o j = 300 MeV and x = 0 .1 , which corresponds to 0 = 0.01°,


this formula gives 3000 barn/steradian, while the differential cross
section for Compton scattering under the same conditions gives only
7 barn/steradian. The Rayleigh scattering is also small in this case,
and, therefore, the scattering of photons for large energies and small
scattering angles is determined almost completely by potential scattering.
CHAPTER IX

Electrodynamics of Particles of Zero Spin

§ 56. Field Equations for Scalar Particles

56.1. First Order Equations


In this chapter we shall give a brief review of the electromagnetic
properties of charged particles of zero spin.
The electrodynamics of scalar particles is of interest because it
is the only example of electrodynamics of particles of spin other than
\ in which it is possible to eliminate all the occurring divergences with
the aid of renormalization.
For the investigation of electromagnetic properties of scalar particles
it is convenient to start not with the second order scalar wave equation,
but with the first order equations derived in § 20 and formally of the
same appearance as the Dirac equations (51), (104):

(56.1)

Here ip is the wave function for the particle, m is its mass and are
Hermitian matrices satisfying the relations
(56.2)
The linear equations (56.1) describe the properties of free particles
of either zero or unit spin, with the matrices ^ satisfying the same
relations (56.2) in both cases. The difference between the two values
of spin consists of the fact that in the case of zero spin the wave function ip
has five components, while in the case of unit spin it has ten compo­
nents. In accordance with this in the former case the matrices have
five rows, while in the latter case they have ten rows.
The five-component wave function which describes particles of
zero spin comprises the scalar wave function and its four derivatives
with respect to the coordinates and the time which transform like the
components of a vector.
[ 819]
820 QUANTUM ELECTRODYNAMICS

Since the matrices satisfy the same relations in the case of scalar
particles and in the case of particles of unit spin, the general formulas
for the matrix elements which determine the various processes of in­
teraction for particles of spins 0 and 1 are of the same form. The prob­
abilities of these processes are, of course, different for the two kinds
of particles, since traces of products of the five-rowed and the ten-
rowed matrices (3M are different.
We give here the formula required for calculating probabilities
of various processes which determines the trace of the product of n
matrices (3M:

spW M J - M M
_ (<*„. K - + Kk- i ,„ - if " is an even number,
[0, if n is an odd number.
We introduce the matrices = 2(3*—I where I is the five-rowed
indentity matrix. It can then be easily shown, by utilizing (56.1) and
the condition that (3^ are Hermitian, that the function Ip = y*^^ sat­
isfies the equation

~Q^Pli+niij>= 0. (56.4)

It can be easily shown that the quantity yip is a scalar, while yfi^y
is a four-vector.
With the aid of y we can construct the Lagrangian density L0 for the
free field corresponding to the scalar particles, and also the current
density and the energy-momentum tensor

1 dy dy \
L0 (56.5)
2 lx -- (.1 /

h = (56.6)

(56.7)

We now consider the interaction between the charged scalar particle


and the electromagnetic field. In order to obtain the equation describing
this interaction we must, as in the case of the Dirac equation, replace
in (56.1) d/dxM by djdx^ —ieA^x ), where A ^ x ) is the potential of the
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 821

electromagnetic field and e is the charge of the particle. As a result


of this we obtain
(56.8)

The analogous equation for y has the form

V> (56.9)

The current density for charged scalar particles in the presence


of an electromagnetic field is determined, as in the case of the free
particles, by expression (56.6). Therefore, the equation which deter­
mines the electromagnetic field A^ produced by charged scalar particles
has the form
d2A.
ox.. (56.10)

The equations for the scalar particles (56.8) and (56.9), like the
Dirac equations, are invariant under the transformation of charge
conjugation. This transformation has the form
ip' = ipC, y' = C-1y, (56.11)
where C is a matrix with the following properties:
c - % C = - f t,,
c = C+ = c ,
(56.12)
C2= 1.
If the representation of the matrices is such that the matrices j3k
are real for k = 1, 2, 3, while /?4 is a purely imaginary matrix, then
C = rj4. (56.13)
However, if all the matrices /? are real, then
C = r]1r)2r)3r]i . (56.13')
It can be easily shown that the quantities y ' and ip' satisfy the equa­
tions

(56.14)
822 QUANTUM ELECTRODYNAMICS

With the aid of the charge-conjugate functions ip' and ip' we can
write the current density in symmetric form

V>'PpV>')- (56.15)

Finally, we give the expression for the Lagrangian density in the


case of interacting scalar and electromagnetic fields

L = 1 ^ ° Ajl V> Z3" dx„ -Tm


2 dxv dxv P>‘T x ~ + m )y’' + J^
(56.16)
The last term describes, as in the case of electrodynamics for electrons,
the interaction between the fields.
If we treat ip, y> and A as operators, then equations (56.8),- (56.9)
and (56.10) describe the interacting quantized scalar and electromagnetic
fields in the Heisenberg picture.
We now demonstrate how the quantization of the scalar charged
particle field is carried out.

56.2. Quantization of the Free Scalar Field


We consider the state of a free scalar particle of momentum p (p, ip0)
ip(x) = cp(p)eipI.
From (56.1) it follows that cp (p) satisfies the equation

(ip+m)<p(p) = 0, (56.17)

where p = ft^p ■It can be easily shown that for any vector q the following
relation holds:
£(02- ? 2) = O - (56-18)

On utilizing this relation and (56.17) we obtain p2+ m 2 = 0, whence it


follows that p 0 = ± |/ p 2+ w 2.
Thus, a given value of the momentum of the particle corresponds
to two solutions cp (p) which differ in the sign of p 0. These solutions
may be referred to as solutions of positive and negative frequency.
We denote the amplitudes cp{p) corresponding to them by cp(+)(p) and
<p{- ](p).
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 823

In order to determine the physical meaning of these solutions we


define the energy and the charge of the field by utilizing the general
formulas (56.6), (56.7):

E = ~ j Titdr= ± p 0(<PU)(p)Pi<P{±)(p),
1 (56.19)
Q = j J j xdr =

(we assume the normalizing volume both here and later to be equal to
unity). Since the energy of the field E must be positive, it is natural to
normalize cp (p) in the following manner:

(y'‘ \ p W l<t'±'(p)) = ± 1. (56.20)

In this case E coincides with \pQ\ = j/ p2-{-m2, while the charge of the
field is equal to -\-e in the state <p(+)(p) and to —e in the state
Thus, solutions of positive and negative frequencies describe particles
of both signs of the charge.
We now proceed to quantize the scalar particle field. As in the case
of the electron-positron field, we must regard y) and ip as operators
operating in particle number space and obeying definite commutation
relations. In order to establish these relations we utilize expansions
of y> and y> in terms of plane waves:

y> = S { ^ P<p(+){ p ) eii,x+ ^ i ( p (~ ){ p ) e - ii,x},


p (56.21)
V = £ { a t v i + ) ( P ) e ~ i p I + b P ¥ ~ )( P ) e i p I } ,
p
where <pu ) (p) satisfy the normalization conditions (56.20) while ap,
a+, bp and b^ are certain operators in particle number space. Substitution
of (56.21) into (56.6),(56.7) leads to the following expressions for the
energy E, momentum P and charge Q of the field:

E = E /?o(a+aP+ bPb^)>
p
P = I> (a+ ap + b p b + ), (56.22)
p
2 = i> (a + a p -b pb+).
824 QUANTUM ELECTRODYNAMICS

Since these quantities must represent sums of the energies, momenta


and charges of the individual particles the operators ap and bp must
satisfy the following commutation relations:
[ap, aa ] dpq, [bp, b9 ] dp(],
[a,, a J = 0, [bp, b J = 0, [ap, bJ = 0, (56.23)
[a+ a+] = 0 , [b+, b+] = 0, [a+, b+] = 0.

Moreover, the eigenvalues of the operators a+ap and b+bp will be


integers, which we denote by N (p+) and while E, P, Q assume the
form
E = 2 p < t N '+ '+ N ' ~ '~ l ) ,
p
P = 2 p (NI+>+NI->- 1), (56.24)
P

= 2 < n ? ' - w - ' - i).


q
p
The quantities N^+) and are the numbers of positively and
negatively charged particles of momentum p.

56.3 Commutators o f the Field. Vacuum Expectation Values of Products


of Field Components
We now obtain the commutation relations for the operators y>(x)
and yi(x). On utilizing expansions (56.21) and the commutation relations
(56.23), we obtain
k W - V/»(*')] = Z F^ ( p ) e lp(x X'}- Z FaP(p)e —ip(x-x') (56.23')
p p
where
Faf ( p ) = <pi+)(p)¥P+)(p),
(56.25)
Ffp(p) = fPa~)( P) V(f \ P ) -

It can be easily shown that Ffp(p) and FaP(p) satisfy the equations
(ip + m )F +(p) = 0,
(—ip+m)F~(p) = 0.
Therefore, the left hand side of these equations must coincide up to
a numerical factor with the polynomial p(p2+ m 2) = 0:
F +(p) = a p ( i p - m ) ,
(56.25')
F~(p) = bp(ip+m).
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 825

The coefficients a and b can be evaluated from the normalization con­


dition (56.20):
[<pl+)(p)Pi<pl+Kp)) = SP (PiF+(p)) = 1 >
(?!' )(/’) ^ <' )W ) = Sp [PiF~(p))= —1.
On substituting into these equations expressions (56.25') we obtain
1
Sp(PiP(ip — ™)),
a

— = - S p ( £ 4p(;p+ m )).

Further, on utilizing formulas (56.3) which determine the traces of


the products of the matrices /S , we finally obtain
i
a= b
2m p0
and
ip(ip—m) /p O p + m )
F+(p) F~(P) (56.26)
2mp0 2mp0
On substituting these expressions into (56.23') we obtain the commu­
tators for the field:

[V>a(x),V>,l ( x ' ) \ = A (x ~ x ')- (56‘27)


To these relations we must add the obvious relations
bpa(x),tpp(x,) ] = 0 ,
[Va(x),^(x') ] = 0. [ }
We define the chronological product of the field operators

T (Wa(x)V>fl(x ')) = k[l+e(t, 0 ) v a(*)VV(*')


+ j ( l —e(t, t'))y>p(x')y)a(,x)t (56.28)
where
1, t > t\
e(t, /') =
-1 , t < t\
and calculate its expectation value in the vacuum state of the scalar
particle field <r(yia(x)^(x'))>o.
826 QUANTUM ELECTRODYNAMICS

The vacuum state is defined as the state characterized by the minimum


energy of the field, i.e.,
<a+a„>(, = <b+b,>o = 0 . (56.29)

On utilizing this definition and the definitions of the functions F +(p)


and F~{p), we can, first of all, obtain the quantities

= 2 F'-p(p)e<-'-r K

<Vf(*')v\,M>. = I Fi
p
Further, in accordance with (56.28) we obtain the quantity of interest
to us
1
<71Va(*)v^(*'))>o A w (x-x')
2m a/3

is (t , t')
A (x-x') (56.30)
2m aP

This expression can be rewritten in the form

< rfe (* )^ (x '))> o = T ^ ( x - x ' ) + ^ ( \ - p l ) d ( x - x r), (56.31)


m
where = —i/34 and

T%x) = P n - i n r i dx..
P r - x r - - 1 dx?
.2 w ]^ * ). (56.32)
2m
, y r " dx

The function T c( x —x') is the Green’s function for equation (56.1)

T c ( x ) = ~ i6{x)- (5 6 3 3 )

This function may also be called the scalar particle propagator.


We see that it differs from the vacuum expectation value of the chronolog­
ical product y>n(x)ip^x') by the quantity —(i/m)( 1—/?£) <5(x—x').
We recall that in the case of the electron-positron field the propagator
does not differ from the corresponding vacuum expectation value.
This is associated with the fact that not all the components of the field
y>(x) are dynamically independent (cf. subsection 57.1).
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 827

We also determine the difference between the T- and 7V-products


of two field operators y)(x) and y(x'). This quantity, called the pairing of
the field operators, is obviously equal to

WaO) V 0 0 =T(ip (x) y>(x')j-N(y> (x) y>(x')|

= r ( x - x ') + - - ( 1 0 ^ ( x - x ') . (56.34)

We also give here the expression for the Fourier component of the
function !P(x) which will be needed for subsequent development:

(56.35)
ip (/p —m) + p1+ m2
Tc(p) = (i p + m )-x =
w(/?2+ m 2)

§ 57. The Scattering Matrix in Scalar Electrodynamics

57.1. The Interaction Picture


Equations (56.8), (56.9), (56.10) in which rp(x), y(x) and A^(x) are
regarded as field operators are the fundamental equations for the electro­
dynamics of scalar particles in the Heisenberg picture. For the investi­
gation of the different processes of interaction between the scalar particle
field and the electromagnetic field it is convenient to go over to the
interaction picture. The state vector 0(t) in this picture is related to
the constant state vector &0 in the Heisenberg picture by the expression
& ( t ) = S ( t ) 0 o, (57.1)
where the Hermitian operator S(t) satisfies the equation

i* = V (l)S (i), (57.2)

where
V(/) = / H(x) dr
and H(x) is the Hamiltonian for the interaction between the fields in
the interaction picture. The relation between the operators in the Heisen­
berg picture and in the interaction picture / and f, is given by the expres­
sion
/ = S -W S (0 . (57.3)
828 QUANTUM ELECTRODYNAMICS

In electron electrodynamics the field operators rp(x) and A^(x) in


the interaction picture have the same form as the operators for the
corresponding free fields. But although this situation remains the same
in the electrodynamics of scalar particles in the case of A^(x), it no
longer holds for the operators yi(x). The point is that not all the com­
ponents of the field rp(x) are dynamically independent. Indeed, on
multiplying equation (56.8) from the right by 1—/5q we obtain

= (*=1,2,3). (57.4)

On utilizing the following specific representation of the matrix (3%

1 0 0 0 0
0 0 0 0 0
32
ro 0 0 0 0 0
0 0 0 0 0
0 0 0 0 1

we see that contains only two components, and correspondingly


(1—/?l ) y contains only three components. Thus, the components of
(1 —f i D f can be expressed in terms of the two components of fily
which are dynamically independent. On taking this fact into account
we can require that the dynamically independent components should
correspond to operators which in the interaction picture have the same
form as the free field operators, i.e.,

= (57.5)

where rp(x) are the operators for the scalar particle field in the interaction
picture and y.>{0)(x) are the free field operators for the same particles.
Expression (57.5) enables us to obtain a direct relation between
ip(x) and ip{0)(x). With this in mind we note that the field operators
y>(x) in the interaction picture, like the corresponding operators in the
Heisenberg picture, must satisfy the subsidiary condition (57.4). On
utilizing this condition, and also the analogous condition for free fields

a -/W " = <* = i, z, 3)


ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 829

and, finally, condition (57.5), we can easily obtain the expression relating
y)(x) to y>(0)(x) and A ^ ]{x)\

(57.6)

Similarly we can obtain

(57.7)

We now obtain the form of the Hamiltonian density H(x). In view


of the subsidiary condition (57.4) this quantity will differ from the
“natural” value —ie-ip{0)( x ) A w (x)ipw (x).
In order to obtain H(jc) we operate with operator /3^d/dx^+m on
the field operator ip(x) in the Heisenberg picture. On noting that in
accordance with (57.3) and (57.6)

+(*) = s -H 0 v (o,W S ( 0 + - - - - s - 1(0(i-i8g)A «o>(x)v «»)(jr)S(0,


m

we obtain, on utilizing the relation (1—/3o)/?o — 0,

( k = 1, 2, 3) .

Further, on utilizing equation (57.2), we obtain

(57.8)

On the other hand, it follows from equation (57.6):

S(0- (57.9)
830 QUANTUM ELECTRODYNAMICS

A comparison of (57.8) with (57.9) shows that the operator V(f) must
satisfy the relation

[V(0,/W °>(*)] = eA«°)(x)|l + - ^ ( l - ^ ) A < (»(x)JV<0'W

---- e| l r - + f f l | ( l - ® A ,0,W F <0)W ( k = 1 , 2, 3) . (57.10)


W \ VXk I
We can satisfy this equation and also the analogous equation for y(x)
by choosing H(x) in the form (141)

H (* )= —/eyi(0)(x) A<0)(x) J^l + — (1 —/3q) A<0)(x)J y){0)(x). (57.11)

This can be easily shown if we utilize the commutation relations for


y (0)(x) and yil0)(x) at the same instant of time:

[V>M(r> 0,V> (?', 01 = —i0 o)r ,& (r —r')+^(PoPk+PkPo)p*-J!T-&(r


f/l VJC —r ')'

57.2. Rules for Calculating Elements of the Scattering Matrix


We proceed to construct the scattering matrix S. This matrix, just
as the scattering matrix in electron electrodynamics, is given by the
value of S(f) for t = o o , and can evidently be written in the formI

I - i f H (i)d*x\
S= 7> /, (57.12)
where T is the symbol for time-ordering. In electron electrodynamics
the expansion of the last equation in terms of H(x) is at the same time
an expansion in powers of the charge e. In scalar electrodynamics the
situation is made more complicated by the fact that H(x) contains the
charge e not only as a common factor, but also inside the square brackets
of (57.11). Therefore an expansion in powers of H(x) will not be an
expansion in powers of e.
In order to avoid this difficulty we go over in the S-matrix from
T-ordering to A-ordering without utilizing the expansion of the S-matrix.
This can be done by means of the technique described in subsection 24.4.
By proceeding in this manner it can be shown that (7)

(57.12')
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 831

where
£' = z - v ,

On the other hand, it can be easily shown that

e —i jf H(x) d‘x — eve —ejf vi(z) A(x) y(i) d ‘x

On substituting this expression into (57.13) we obtain:

(57.14)
(Here and later we omit the index zero from the field operators.)
This expression is formally of the same form as the analogous
expression for the scattering matrix in electron electrodynamics. It
corresponds to the interaction Hamiltonian (106)

H (Aw = ~ iey(x)A(x)if(x) , (57.15)


and to effective pairings between the field operators given by
< ( * b ’/j(*') = T caft( x - x ' ) . (57.16)
We thus see that for the evaluation of the elements of the scattering
matrix we can start from the effective Hamiltonian (57.15), which
differs from (57.11) by the absence of the last term, and that we can
utilize the effective pairing between the field operators which differs
from (56.34) also by the absence of the last term. From this it can be
seen that the evaluation of the matrix elements of S can be carried out
with the aid of Feynman diagrams and rules of the same kind as in
electron electrodynamics, the only difference consisting of the fact
that now in the diagrams the internal lines corresponding to scalar
particles must be made to correspond to T c(p), and not S c(p); moreover,
the vertices of the diagrams must be made to correspond to the matrices
832 QUANTUM ELECTRODYNAMICS

(1^, and not y . Finally, we must keep in mind in evaluating the cross
sections for the different processes that scalar particles have no spin,
and, therefore, the cross section need not be averaged over spins.

57.3. Divergences o f the Scattering Matrix


We now proceed to investigate the problem of the divergences of
the scattering matrix. In order to do this we consider an irreducible
nth order diagram containing Nm external scalar particle lines, and
Np external photon lines. If the number of internal lines of the diagram
is equal to F, then the number of independent variables of integration
in the integral which determines the matrix element corresponding to
the diagram will be 4(F—n+1). In the integrand of the matrix element
we have a rational function the degree of whose denominator is equal
to 2F. Whether the integral is convergent or not is determined by the
difference between the degrees of the numerator and the denominator.
The integral will converge if K = I F —n —4(F—n + 1) ^ 1, where
n' is the degree of the numerator, and will diverge if K < 1.
Thus the problem reduces to the determination of n .
In order to determine n' we consider in our diagram an internal
scalar particle line which contains s vertices, s ^ n. This line corresponds
to the following expression in the integrand of the matrix element:

We determine the highest power of p in the numerator of this expres­


sion. It might seem that this power is equal to 2(s— 1), and coincides
with the degree of the denominator. In actual fact, however, because
of the properties of the matrices this is not so, and the degree of the
numerator turns out to be equal to s.
For example, take s = 3. Then A 3 has the form

A3=

1
( p { f m 2) O t + m 2)
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 833

The highest power of p in the numerator corresponds to the term

which vanishes for all pl and p2. Therefore, the degree of the numerator
is equal to 3, and not to 4.
The vanishing of X l is a consequence of the more general relation

xn = - M?2-?2)=°.
which holds if n is an odd number.
In order to show the validity of this last relation we introduce the
matrices
P = iW § $ . 'P = 2 X P/^> (57.17)
which differ from zero inthe case of thefive-rowed matrices (3/n and
which are equal to zero in the case of the ten-rowed matrices f}/t.
It can be shown that the matrices P and P satisfy the relations

M .p =

/?„P = P/V (57.17')


Pj8„ = A.P,
(P + P ) ^ , = ^ ( P + P).
The last relation shows that the matrix P + P is proportional to the
identity matrix. By taking its trace we can easily show that
P+ P = I . (57.18)

Finally the following relations hold:

if n is an even number,
(57.19)
if n is an odd number.
P^i^«s ••• ~ Pt‘n
We now consider the quantity Xn. By utilizing (57.17) we can rewrite
Xn in the form

= (P2- P 2) ^ A - P r t f ' - M P + n
834 QUANTUM ELECTRODYNAMICS

On utilizing the second of formulas (57.19) we obtain in the case of


odd n :
Xn = Cp'2- p 2)
but in accordance with (57.17) this expression vanishes.
By utilizing the vanishing of Xn for odd n we can easily show that
the highest degree of the numerator in the expression for As does not
exceed s. From this we can conclude that the degree of the numerator
of the rational fraction in the integrand of the matrix element cannot
exceed the order of the diagram, i.e., the number n. On setting n = n
we obtain for K the value
K = 2F—n —4(F—n + 1) = 3 « - 2 F - 4 .
On the other hand, 2F-\-NmJr Np = 3n. Therefore
K = N p+ N m- 4 . (57.20)
If 1, i.e., Np-\-Nm^ 5, the integral will be convergent. The
integral will diverge if the following inequality is satisfied (131):
Nm+ N p < 5. (57.21)
This inequality shows that the number of divergences which can be
encountered in the electrodynamics of scalar particles is finite.
We enumerate the different types of divergences corresponding to
the different values of Nm and Np:
1) Nm = 2, Np — 0, K= —2 quadratic divergence,
2) Nm — 0, Np = 2, K= —2 quadratic divergence,
3) Nm = 2, Np = 1, K= —1 linear divergence,
4) Nm = 0, Np = 2, K= 0 logarithmic divergence,
5) Nm= 2 , Np = 2, K= 0 logarithmic divergence,
6) Nm = 4, Np = 0, K = 0 logarithmic divergence.

The cases Nm = 0, Np = 1 and jVm= 0, Np = 3 need not be con­


sidered since the corresponding matrix elements vanish. This follows
from Furry’s theorem which is valid in virtue of the invariance of the
theory under the transformation of charge conjugation both in electron
electrodynamics, and in the electrodynamics of scalar particles.
The types of divergences enumerated above correspond to the
diagrams shown in Fig. 120. Diagrams 5 and 6 are new compared to
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 835

ordinary electrodynamics. The former corresponds to a radiation


correction to the Compton effect, while the latter corresponds to a cor­
rection to the scattering of a scalar particle by a scalar particle.
The elimination of divergences from the various matrix elements
can be carried out by means of the same regularization methods which

/ VV_,'
^ = 2 .^ 0
A
(1 ) (3 )
I I I I
II I1 II •I
it

<------' ' 1r

^=2-V2 ^ = 4 ,^ = 0
( 5) ( 6 )

F i g . 120.

were employed in electron electrodynamics. It is important to note here


that in actual matrix elements all the infinite constants, with the excep­
tion of two, cancel one another; while the elimination of the remaining
two infinities requires use of the idea of renormalization of charge and
mass of the particle.

§ 58. Scattering of Scalar Particles

58.1. Scattering o f Scalar Particles by the Coulomb Field of a Nucleus


We now proceed to investigate specific effects due to the interaction
between scalar particles and electromagnetic fields. We begin with the
scattering of scalar particles by the Coulomb field of a nucleus.
The probability of transition of a scalar particle from the state
of momentum p t and energy ex to the state <p(p2) of momentum
p 2 and energy e2 in the field A f \ x ) is equal to
w1 = 2 ne2\q>{p^)A{e){q)(p{p^\2b{El —e^, q = p 2— P i, (58.1)
836 QUANTUM ELECTRODYNAMICS

where
A(e) (q) = | P/iA(ff )(r)e~ieirdr.
In the Coulomb field of a nucleus

A(e)(q) =

and
Z 2e4
H’i = 2n — ^ \ ( p ( p 2) P tv ip J l 2d(et - e 2) .

The quantity | <p(p~>)Pi (p(Pi)\2 appearing here is a special case of the


more general expression

\ M \ * = \ v (±){Pr)Q'P{*)(py<t , (58.2)
where Q is a certain matrix. This last expression can be evaluated in
the following manner. First of all it is clear that

\M I2 = W KP f)Q uML)(Pi)^y~)(Pl) Qyd(ri^ )(Pf)^


where
Q = ViQ+v 4-
Further, on recalling that in accordance with (56.26) we obtain

(p{a 'K p )W K P ) = Fa(iKp)'

\ M \ ^ = S p { Q F ^ \ Pl)QF ^(P f)}

= 4fJ - e/ SpiQiPiOPrfrfQiPriiPrTm)}, (58.3)

where et and ef are the energies of particles of momenta p i and pf .


On setting Q = /?4 in this formula and on noting that /?4= ?/4/?4+?/4— Pi,
we obtain

\v(.Pt)Pi<riPi)\2 = £ S p { P J p d i h —m )P JP 2( iP i— fn)}.

On utilizing (58.3) we have

SP {PJPi(iPi—iri)PJPi(iPi—m)} = m2{ex+ e t f .
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 837

Therefore
ZV _(£l + £?)2
2nd{e1—e2). (58.1')
P i- P21 4^62
On multiplying \\\ by dp2/(2n)3 and on dividing by the flux of incident
particles J — we finally obtain the following formula for the
differential cross section for the scattering of a scalar particle by the
Coulomb field of a nucleus:
Ze2
da = (58.4)
2pv

where 9 is the scattering angle. This formula differs from the correspond­
ing formula for the scattering of particles of spin ^ by the absence
of the factor ( l —?;2 sin2 &).

58.2. Scattering of a Charged Scalar Particle by a Scalar Particle


We now determine the cross section for the scattering of a charged
scalar particle by a similar scalar particle.
The matrix element for this process has the form

S™r= ie2M(2jr)4rK/h+ A - P i~ P2 ) ,
(58.5)
M = {<p(P2)Pldp(Pi)){v(P2)P„<p(p'i)) , (v(p'i)Pr<p(.Pi))(v(pJP?<p(pD)
(Pi-Pi.)•“ ' +(K -a )2
where p x, p[ and p2, p'2 are the momenta of the particles before and
after scattering and 9o(p) are the corresponding amplitudes of the wave
functions of the particles.
We note that in contrast to the case of electron-electron scattering
when the two terms in M have opposite signs, in the present case both
terms have the same sign. This difference is due to the different statistics
in the case of electrons and of particles of .spin zero.
The differential scattering cross section is equal to

» ^ 1 *j 121Pz IA de2do2 - . , /cqc'x


da — ^ ~ j £2 ^2)’ (58.5)

where J is the flux density of the incident particles.


838 QUANTUM ELECTRODYNAMICS

On utilizing (58.3) we can write \M \2 in the form


1 ( 1
16e1e[ s2s2m4\ ( p 2- p j * Sp [fiu ifi (ip i —m) Pv ip* (iP - m)]
2

X Sp[P^ip[(ip[-m)Pvip2(ip2—m)]

1 4iSp[PMiPi(ipi—m)P,ip2(ipt-m)]
(P2 —P1 )
• */
X SplPpipl ( i P i - m ) p vip2(ip2- m ) \

2 Sp [PMipx(ilp1- m) ~pvip2(ip2- m) ^
(.P i- P i )2 (P2 - P 1 ) 2
X ip[(ip[—m)Pvip2(ip2- m ) ] \ .

Evaluation of the traces of the matrices with the aid of formulas (56.3)
yields
1 /( n. 4-n^ ( ( n. -1-n'^ ( 7r-l-n~V 2
\M \2 = \ f n_— n_ Y& 1 i
lee^ieafig \ (P2 - P 1 Y (P2 - P 1 ) 2
1

In the center of mass system this expression can be considerably


simplified:
\M\* = -J—i l ^ 2
4e2 s2—m2 sin2??
where ?? is the scattering angle and e is the particle energy. On substi­
tuting this expression into (58.5'), and on noting that in the center of
mass system J = 2p/e, we finally obtain the following expression for
the scattering of a scalar particle by a scalar particle (7):

-1
or mc e2—m2
da = 4m2zd sin2?? 2s2—m2
do. (58.6)

§ 59. Scattering of a Photon by a Scalar Particle. Bremsstrahlung


Photons from a Scalar Particle

59.1. Scattering of a Photon by a Scalar Particle


We now consider the interaction between scalar particles and pho­
tons. First of all we determine the cross section for the Compton scat­
tering of photons by scalar particles (32). By utilizing the diagrams
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 839

for this process shown in Fig. 21, and the general rules for writing
matrix elements we obtain the following expression for the matrix
element for the scattering of a photon by a free scalar particle:
lez i / l 0 / i ~ m) + / i + r a 2 ,
<p(P2)\e2
2 |/ ftj1w2

<p(Pi) (2tt)4d(px-\-kx—p 2—k 2) (59.1)


Cl m ( f \ + m 2)
Here <p(px) and (p(p2) are the amplitudes of the wave functions of the
particle in the initial and final states of momentum p x and p 2\ cox and co2
are the frequencies and ex and e2 are the photon polarizations; and
A = P i + k x = p 2+ k 2, f 2 = p x—k 2 = p2—kx (k x and k 2 are the photon
momenta).
The cross section for this process has the form
ei dp2 dk2
do 1<
p (p 2)Q(p ( a ) I2
4 ojx oj2 ~J(2n f
X d ( P i + k x — p 2 — k 2) e2— t o 2) , (59.2)
where

Q = ~ ^ - A (ifI (ifI - m) + f i + ™2) ex(if2(if2- m) + / 22+ m2) e2,

m2xx = f x+ m 2, m2x2 = f 2-\-m2,


and / = 1 is the flux density of the incident photons.
On utilizing formula (58.3) we write do in the form

da = ^ onr
~ r ~~b
xx SP {QiPiiiPi-™) QiPziiPz-™ )} > (59-3)

where r0 = e2j4nm and do2 is the element of solid angle containing


the momentum of the particle after scattering.
The traces of the matrix products can be evaluated with the aid
of the following formula:
i
SP ( I I [ifr(ifr~m)+f? + m2]er}
T — 1

= m4[(e!, / x + / 2) (e2, / 2+ / 3) (e3, f s + f i ) (e4, / 4+ fi)


~ ( e xe2) ( f $ + m 2) (e3, f 3+ f ^ (e4, / 4+ /i) -
840 QUANTUM ELECTRODYNAMICS

— (e2e3) ( / l + w 2) ( e i . / i r / a ) (c4»/ 4+ / 1)
- ( 63 ^ 4) ( / 42+ ™ 2) (<?i, / 1+ / 2) (e2, / 2+ / 3)
— (/1 +"**') (e2, / 2+ / 3) (e3, / 3+ / 4)
+ 0 i£ 2) (c3c j (/?+"**) (e2<?3) (/1 + m 2)( / 32+ m 2)].
(59.4)
(This formula can be easily generalized to the case of an arbitrary
number of factors).
On utilizing formula (59.4) we obtain
Sp { Q i P i ( i p i - m ) Q i h ( i h - ™ ) }

= Am21 (e1e2) - (e1, f x+ />J (e2, f x+ p 2)

~ 2nkr2(e2’f2+Pi) (ei ’ /2+/72)}


and
, „ co2 Jo2 P 2 . "I2 2
^ = r ~ri*x~2 ^ 2P2) ~ ~ ^ x 2 (eiP2)

(59.5)
If the particle was at rest before scattering, then p x = 0 and

da = r» { f r ) (e' e^ do*- (59-6)


In the case of unpolarized photons this formula yields

d a = \ r2\ ^ f ) (l + cos2^ do^ (59.7)


where ■&is the scattering angle.

59.2. Bremsstrahlung from Scalar Particles


We now proceed to determine the probability for the emission
of a photon by a scalar particle in an external field A[f{x) (39).
The matrix element corresponding to the bremsstrahlung process
can be written in the form

C(2) —
ie2
4W > .( ,)
y 2a» m 3x x

+ > ) (g) e (59.8)


m3x2
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 841

where p l and p2 are the four-momenta of the particle in the initial and
final states, cp( p j and <p(p2) are the corresponding amplitudes of the
wave functions, k and e are the momentum and the polarization of
the emitted photon, is the Fourier component of the external
field and

f i = Pi+k, f z = p1—k > q = p2-P 2 + k ,


fn-xi = f f + m 2 = 2p2k, m2x2 = f 2 = —2pxk .
If the external field is static, then
Aj,e>(q) = d ^ A [ e)(q)2jid(£2- E x+oj),

where ex and e2 are the values of the particle energy and co is the photon
frequency. The differential cross section for bremsstrahlung is deter­
mined by the following formula:
dk do2
da=-~ 2
8com2
M 'e,(?)i2 - j ^
| /?]|
SpF
"(2r t f '
(59.9)

where

F = Q ipi(ipi~m )Q ip2(ip2- m ) ,

i f i O f i —n f i + f i + m 2 if2(if2- m ) ^ f l - \ m l
Q = e /^4+ A e.
m3xx m3x2
The evaluation of traces of products of the matrices can be carried
out with the aid of formula (56.3). As a result of this we obtain

Sp F = 16m2 ( e , - J } - p i + - * i ~ - p i_ (59.9')
m2xx mix2

If the emission of radiation occurs in the Coulomb field of a nucleus,


we have A4(q) — — i(Ze/q2) and the bremsstrahlung cross section
assumes the form
2
4Z 2a3 \p2\ 1
e. — P 2 - \ ------- P i co c/co do do2, (59.10)
7i2 \px\ mlqi Xi x2
where do and do2 are the elements of solid angle containing the mo­
menta k and p 2.
842 QUANTUM ELECTRODYNAMICS

The cross section for the emission of an unpolarized photon can


be obtained by summing (59.10) over the two photon polarizations:
,2
, 4Z2a3 IP*! 1 2|/ £1
£i , £2
£2
da — ---- — ;coa
'21--- P2 H------Pi
71‘ \Pl I miqi *1 »2

<o j doj do do2. (59.11)


\ *1 *2 I

§ 60. Production and Annihilation of Pairs of Scalar Particles

60.1. Production o f Pairs o f Scalar Particles by a Photon in the Coulomb


Field of a Nucleus
We now determine the cross section for the production of a pair
of scalar particles by a photon in the Coulomb field of a nucleus (39).
The matrix element for such a process can be written in the form

ie2
S $ r = -----i = - y (+){p+) \ A {e){q) ifiO fi—m)+ fi + -
\2(x> m (ff-\-m 2j

>e 'J2^ / 2_ m ) + / 22f A {e)(q) <P{ H / O , (60.1)


m ( /|+ m 2)
where p + and p_ are the momenta of the two components of the pair,
<p(+)(p+) and ^ (-)( / 0 are their wave functions and q = p +-\-p_—k .
This expression can be obtained from the matrix element for brems-
strahlung if we carry out the following substitution in the latter:
p 1 -*■ —p~, P 2 P+, and k -> —k. Therefore, the cross section for
this process can be immediately obtained from formula (59.10) by
means of such a substitution. As a result of this we obtain
2
Z2a3 P+\ \P~ e
do = 8 P-
(2nf qioP *1
E+P- de+ do+ do (60.2)
x2
where

x2 o kP+
m2
and do+ and do_ are the elements of solid angle containing p + and p _ .
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 843

The total cross section for pair production is equal to (4)

y
5 + 6 j'2 —
:v^)\
where y = co/2m and E(x) and F(x) are the same functions which
(60.3)

appear in the total cross section for the production of electron-pos­


itron pairs (cf. § 32).
60.2. Production of a Pair o f Scalar Particles by Two Photons
The matrix element for the production of a pair of scalar particles
by two photons has the form

ifi(ifi-m )+ fi+ ™ 2
siS,= v '- ’o o j a m { fl+ m 2)

, -, t/2( / ^ - w ) + / 2g+/w2 .
(p+(p+) d(p++P-—k —k'). (60.4)
6 m ( f 2+ m 2)
where k , k ' and ’e, e are the momenta and the polarizations of the
two photons, co and co' are their frequencies and f x = p +—k ' and
A =P + ~k -
We give here only the final result of the calculation of the pair
production cross section (7). In the c.m. system this cross section is
equal to
2 (e'p+)(ep+)
do = do,

(60.5)
where r0 = a2/m and 0 is the angle between k and p +.
The cross section averaged over the photon polarizations has the form

do. (60.6)
844 QUANTUM ELECTRODYNAMICS

Finally, the total pair production cross section is given by

o = 7 ir * ~ { ( x 2+ l ) jc ) /j c ^ T - ( 2 x 2- l ) ln ( x + j/x 2^ I )} , (60.7)

where x - co/m.
60.3. Two-Photon Annihilation of a Pair of Scalar Particles
The process which is inverse to the one just considered is the con­
version of a pair of scalar particles into two photons. The calculation
of the cross section for this process is analogous to the calculation
of the cross section for pair production by two photons. We give here
only the final results.
In the c. m. the differential cross section for two-photon annihi­
lation of a pair of scalar particles has the form

(60.8)

where d is the angle between the momentum of one of the components


of the pair and the momentum of one of the photons, e is the particle
energy and do is the element of solid angle containing k.
The total cross section for two photon annihilation is equal to
(x2+ l ) x 2x2—1
a = Anrl ~Ai In (x— y x 2— ] (60.9)
a -2 - 1

where x = e/m.
60.4. Annihilation of Pairs o f Scalar Particles into Electron-Positron
Pairs and the Inverse Process
In concluding this section we investigate the process of the con­
version of an electron-positron pair into a pair of scalar particles, and
also the inverse process of the conversion of a pair of scalar particles into
an electron-positron pair (7). These processes correspond to the diagram
shown in Fig. 121; pe_ , pe+ denote the momenta of the components
of the electron-positron pair, while p™, p!£ are the momenta of the
components of the pair of scalar particles.
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 845

The matrix element for both processes can be written in the form
S & = ie2\M\2(271y (2Tty d i p ^ + p l - p t - p ”),
i (60.10)
M = (P%+p*y ^ ^ ^ ^ ^ ( ^ ( pD P ^K p D),

where u(pe_) and v (p +e ) are the spinor amplitudes of the electron and
the positron, and 91 'Kp'D and <f(+)(p™) are the amplitudes of the wave
functions for the scalar particles.

The square of the absolute value of M which determines the cross


sections for these processes summed over the electron and the positron
spins can be written in the form

fM ,2 = ( p ‘J p t r s v { r A i k - ™ y v , ( i p % + m r) }

1 A . A
x ^ s P{ft^P+0'P+—m)Pvip1(ip1+m)}, (60.11)

p e = r t,p% pm = PMp%,
where m e and m are the masses of the electron and of the scalar particle.
On evaluating the traces of the matrix products in accordance with
formulas (56.3), (59.4), we obtain
1 1
\M\2 = {(p%p1)(pe pi)
2£™£l£e,_£e~ (pt-i-pO*
-f (PIP® (Pe+P%) ~ (P lpl) (Pe+P'D ~ (PIP'D (P%Pl)
— [ ( P % P 1 ) — m l ] [ ( p l p m) - f m 2] ] . (6 0 .1 2 )

The cross section for the conversion of an electron-positron pair into


a pair of scalar particles is equal to

da = ei \M \2- J ^ j <5(/?+ + p t — P+ — Pl)> (60.13)

where J is the flux density of the incident particles.


846 Q U A N T U M E L E C T R O D Y N A M IC S

In the c. m. system we have


P% + P e- = P X + P - = °>
£% = el = £+ = e” = £,

(60.14)

where 6 is the angle between the momenta of the positron and one of
the scalar particles and do is the element of solid angle containing the
momentum of one of the scalar particles.
The total cross section for the production of a pair of scalar particles
is equal to
7ia2 / , w 2 \ 3/2/
(60.15)
1 6 e2 \

The cross section for the conversion of a pair of scalar particles


into an electron-positron pair differs from the cross section for the
production of a pair of scalar particles by the difference in the number
of final states, and also by a different expression for the flux density
which in this case is equal to

The final expression for the cross section for the conversion of a pair
of scalar particles into an electron-positron pair has the form
1/ 2
m2
do ^ - ) c o s 2eL/o. (60.16)
8£2 £-
The total cross section for this process is equal to
1/ 2 1/2
xca2
a (60.17)
4~fJ
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 847

§ 61. Polarization of the Vacuum in the Case of Charged Scalar Particles

61.1. Vacuum Polarization Tensor for Scalar Particles


In the preceding Chapter we have discussed the various effects due
to the interaction of the electromagnetic field with the vacuum of the
electron-positron field and of electrons with the vacuum of the electro­
magnetic field. In this section we shall discuss the analogous effects
in scalar electrodynamics. We begin by determining the radiation
corrections to the photon Green’s functions due to the scalar particle
vacuum. This problem reduces to the evaluation of the polarization
tensor
If we confine ourselves to the second order radiation corrections
the polarization tensor IJ^2) (k) associated with the scalar particle vacuum
will be given by the formula analogous to formula (47.28)

n f j ( k ) = —e2 f d*p Sp{fifi(ip-\-m) 1pv(ip + iic+m) '}. (61.1)


The difference between this expression and (47.28) consists of the
fact that this expression contains the matrices instead of y^ and
that, moreover, this expression contains an additional minus sign
associated with the different statistics for the scalar particles.
The tensor IJj^)(k) diverges quadratically; the regularized value
of the tensor can be found by means of the formula

On evaluating the trace of the products of the matrices ^ we obtain

where

f ___
J (p2~rm2)[(pJrk)2f - m 2Y
f PMd*P
J (p2+ m 2) [(p+k2) + m2Y
r = f ______ PjtPvfkP _____
^ J (p2+ m 2) [ ( p + k ) 2+ m 2Y
848 QUANTUM ELECTRODYNAMICS

With the aid of the identity

_L=f
ab
dx
J [ax-\-b{\ —x)]2

the integrals can be written in the form

J= f dx f Fdip ,
o
1

Ju = f d x f ptlFd*p,
o
i
Jllv = J dx j pflpvFd'p,

where
F = ----------1 -
(p2-\-2pkx+m2)z
We shall not carry out here any detailed calculations, but shall
directly give the final result for the regularized value of T I^ R{k):
i
in‘ v i dv
n™K( k ) = j (61.3)
k1
4m2
where v = \ —2x. On introducing the new variable 0 by means of
k1
sin & =
4m2’

we write 7 7 ^ (k) in the form1


( 4 m 2 I L-2 1)
n % ( k ) = 2 ^ i e H k „ k - d , „ k ‘) [ - ^ ( l- © c o t© ) + - J . (61.4)

When we have the expression fo rII^JR(k) we can obtain the radia­


tion correction to the external current J ^ i k ) due to the interaction be-

1 Feynman (second of references (62))-. (the factor 1/2 is missing in the formula
. / 2\ • ,
for n hvR given in this reference).
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 849

tween (,v)and the vacuum of the scalar field. The Fourier component
of this additional term is equal to

6A\:]{q) - - \ 2 n f n ^ R^ ’ (61.5)

where A f{le)(q) is the Fourier component of the external field.


In the case of slowly varying fields we can expand 77<2)(/c) into
a power series in k. On restricting ourselves to fourth order terms
we obtain
ffitM *) = 3 0 ( 6 1 . 6 )

In this case the additional term in the current is given by

(61-7)

61.2. Correction to Coulomb's Law


If we know the polarization tensor 77((^ we can also obtain the
radiaton corrections to Coulomb’s law. The radiation correction to
the external potential A (e)(k) is determined in momentum space by the
following general formula:
dA<;Hk) = n<?R(k)D',(k)A^(k).
In coordinate space this formula yields for a static field

dk
<5Aj;e)(i*)
(2n)7

In the case of the Coulomb field of a nucleus we obtain from this expres­
sion, on taking radiation correction into account and on utilizing
expression (61.3),

Ze e2 1 0ikr
ft4 dv
<p(r) + d<p(r) dk
4nr \ 4 n f 12n2 1 -0
(61.8)
or
<P(r) + 6<p(r) = <p(r) (1 + Xo00)> (61.8')
850 QUANTUM ELECTRODYNAMICS

where
a dv
XoOO k sin kr dk (61.9)
\2n? mi \k*( \ - v 2)

On carrying out the integration with respect to k and on introducing


in place of v the new variable f = (1 —v2)~yz we write %0(r) in the
form (3)
CO

The corresponding expression for the correction due to the electron-


positron vacuum has the form
CXJ

-- 2mjr£
Xll2
<'>=£/■ t2

where mlj2 is the electron mass.


In the limiting case mr << 1 and 1 the function %a(r) is deter­
mined by the expressions

« m J — ♦ where mr < 1,
07i \ mry 3
Xo(r) - - (61.10)
a
- (mr) 5'2e~'2mT where mr > 1,
16 | 71
where In y is Euler’s constant.

61.3. Photon-Photon Scattering. Radiation Corrections to the Lagran-


gian of the Electromagnetic Field
The scalar particle vacuum, just as the electron-positron vacuum,
leads to a number of nonlinear effects in electrodynamics, such as photon-
photon scattering, the scattering of a photon by the Coulomb field
of a nucleus, etc.
All these effects can be described in the electrodynamics of scalar
particles, as in electron electrodynamics, with the aid of the fourth
order scattering matrix:

S<4> = 4 I d 'X2 d'i x* d*x i N K) AOc) A f x f A f x f )


x Sp T f x . - x f f T f x 2- x a)Pi Tc(x, —v4)/?a T f x , - x f ,} .
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 851

In momentum space this matrix has the form

S (4) = f d*ki d*k2 d % dik 4 (2n'f

x n Ia ^ a ^ a ^ a m ) J ^ { k x, k 2, k 3, fcj, (61.11)
where

dnvXa^l ’ k 2, k 3, k 4)
'd’nVi.a(ki, k 2, k 3, kq)-\- Tfivaj (ki, k 2, k 4, k^-^T^xvaiki, k 3, k 2, k 4)

and

k 2, k 3, k 4) = ~ J dip Sp {PnQp+m) 1Pv(ip—ik2+m) 1

X p?(jp—i L — ik3+ m)-1Ba(ip- ik2—ik3—ik^+m) - 1}.


The quantity TflvXa(k1, k 2, k 3, k4), which is called the fourth rank
polarization tensor, diverges logarithmically for large momenta, just
as the corresponding quantity in ordinary electrodynamics.
Therefore, regularization can be carried out in the same way as in
electron electrodynamics.
On repeating the arguments of § 54 we can calculate the photon-
photon scattering cross section, and also obtain the radiation corrections
to the Lagrangian density for the electromagnetic field due to the vac­
uum state of the scalar particle field. We shall not carry out here any
detailed calculations, but shall confine ourselves to stating the final
results.
The photon-photon scattering cross section (in the center of mass
system) is given in the case of low frequencies by the formula (4)

where <w < m,


(61.12)
where m is the mass of the scalar particle and 0 is the scattering angle.
The total scattering cross section is given by
6
a4 7 136 1
a—— where i» < m . (61.13)
t T I o 'W f m *
852 QUANTUM ELECTRODYNAMICS

The correction to the Lagrangian density for the electromagnetic


field is determined by the following formula (176):

L'(x) = (61.14)

Concluding Remarks
The concept of an elementary particle in modern physics has un­
dergone a number of essential changes. At present it is already evident
that elementary particles are by no means the primary components
of matter, but are rather states of excitation of some dynamical system;
in the case of electrodynamics such a system is the combination of the
electromagnetic and the electron-positron fields which interact with
each other. Even in the absence of particles, i.e., of excitations, the
system which is in its ground state — the vacuum state — has definite
physical properties which manifest themselves in a number of phenomena.
The interaction between particles and vacuum manifests itself
in a remarkable manner in the radiation shift of atomic levels, in the
anomalous magnetic moment of the electron, in nonlinear electrody­
namic effects, and in other phenomena. These facts are of considerable
interest not only from a physical point of view, but also from a general
philosophical point of view. In them we find a new confirmation of
the well known thesis of V. I. Lenin regarding the inexhaustibility of
the properties of the electron and the infiniteness of nature.
In criticizing idealists, who in connection with changes in the concepts
of the structure of matter drew conclusions with respect to its disappear­
ance, Lenin wrote: “The ‘essence’ of things or ‘substance’ are also
relative; they express only an increase in the depth of human perception
of objects, and even though yesterday this increase in the depth of
perception did not extend beyond the atom, and today beyond the
electron and the ether, nevertheless, dialectical materialism insists
on the temporary, relative, approximate character of all these landmarks
in the perception of nature by man’s progressive science. The electron
is just as inexhaustible as the atom, nature is infinite, but it exists infin­
itely, and it is this uniquely categorical, uniquely unconditional rec­
ognition of its existence outside human consciousness and human
senses, that distinguishes dialectical materialism from relativistic agnos-
ELECTRODYNAMICS OF PARTICLES OF ZERO SPIN 853

ticism and idealism” (Lenin, “Materialism and Empiriocriticism,”


Gospolitizdat, 1948, p. 245).
The successes of quantum electrodynamics have demonstrated the
correctness of our basic physical concepts within a definite domain
of phenomena. However, these successes are relative. Electrodynamics
turns out not to be a logically closed theory, i.e., it cannot be developed
absolutely consistently without introducing auxiliary ideas which,
so far, are of a semiempirical nature. Attempts to carry over the methods
of quantum electrodynamics to domains of other phenomena have
so far not resulted in any serious successes. Apparently, the difficulties
of the present theory can be removed only by means of a new change,
and, moreover, perhaps a cardinal one, in the basic physical concepts.
It is quite probable that even the fundamental space-time concepts of
modern physics will undergo a change in this process.
Like any other science, quantum electrodynamics gives expression
to definite objective natural laws. “But this,—as V. I. Lenin points
out, — is not a simple, direct, complete expression, but a process involv­
ing a series of abstractions, the formulation and creation of concepts,
laws, etc., and it is these concepts, laws, etc., (thought, science = “logical
idea”) that encompass conditionally, and approximately the universal
regularity of eternally moving and developing nature” (V. I. Lenin,
“ Philosophical Notebooks”, 1947, p. 156).
These words of V. I. Lenin may be used to.give an excellent charac­
terization of the modern state of development of theoretical physics.
References

1. A brikosov, A . A ., JETP, 30, 96, 386 and 544 (1956), Sov. P hys. JETP, 3,

71, 474, 379 (1 956-57). 4 8 .5 , 5 1 .5 , 5 1 .6 , 5 2 .2

2. A cheson , L . K . , P h y s . R e v . , 8 2 , 4 8 8 ( 1 9 5 1 ) . 1 4 . 4
3 . A khiezer , A . I . , P h y s i k . Z . S o w j e t u n i o n , 1 1 , 2 6 3 ( 1 9 3 7 ) . 5 4 .1

4. A khiezer , A . I., A leksin, V. a n d V olkov, D . , D o k l a d y A k a d . N au k U SSR ,

104, 830 (1955). 6 0 .1 , 6 1 .2 , 6 1 .3

5. A khiezer , A . I. a n d P olovin , R ., D o k la d y A kad. N au k U SSR , 90, 55 (1 9 5 3 ).

5 1 .6

6. A khiezer , A . I. and P omeranchuk , I. I a ., P h y sik . Z. S o w je tu n io n , 11, 478

(1 9 3 7 ). 5 5 .1

7. A leksin, V ., P u b l. P h y s.-M a th . F a cu lty , K h a r ’k o v State U n iv e r sity , 7, 97

(1958). 5 7 .2 , 5 8 .2 , 6 0 .2 , 6 0 .4

8. Bartlett, J . H . a n d W atson, R . E . , P r o c . A m . A c a d . A r t s S c i . , 7 4 , 53 (1940). 14.4


9. Bf.restetskii, V. B., J E T P , 17, 12 (1947). 4.1
1 0 . Berestetskii, V . B . , J E T P , 1 9 , 1130 (1 9 4 9 ). 3 9 .2

1 1 . B erestetskii, V . B . , J E T P , 2 1 , 9 3 ( 1 9 5 1 ) . 1 0 .4

1 2 . B erestetskii, V. B. and L andau , L. D ., J E T P , 19, 6 7 3 (1 9 4 9 ). 3 8 .4

1 4 . B erestetskii, V . B . a n d P omeranchuk , I . I a . , J E T P , 2 9 , 864 (1955), S o v . P h ys.

JETP, 2, 580 (1956). 3 6 .4

15. B erestetskii, V. B. and S hmushkevich , I., JETP, 19, 591 (1949). 4 1 .1

16. Bethe, H . A . , Q u a n te n m e c h a n ik d er E in -u n d Z w e i-E le k tr o n e n p r o b le m e , H a n d b .

P h y s., 2 4 , (1), 1933, p. 273. 4 .2 , 3 1 .1 , 5 3 .1

17. Bethe , H . A ., Proc. C a m b r id g e P h il. S o c ., 30, 524 (1934). 5 5 .3

18. B ethe, H . A ., Proc. R oy. Soc. (L o n d o n ), A 150, 129 (1935). 3 3 .3

19. B e t h e , H . A ., P h ys. R e v ., 72, 339 (1947). 5 3 .1

20. B ethe, H . A . and F ermi, E . , Z . P h y s i k , 7 7 , 2 9 6 ( 1 9 3 2 ) . 3 8 .1

21. Bethe, H. A . and H eitler, W . , Proc. R oy. Soc. (L o n d o n ), A 146, 83 (1934).

2 9 .2 , 2 9 .6 , 2 9 .7

22. B ethe, H . A . a n d M aximon , L . C . , P h y s . R e v . , 9 3 , 7 6 8 ( 1 9 5 4 ) . 1 4 . 5 , 2 9 . 9


23. B ethe, H . A . , a n d R ohrlich , F . , P h y s . R e v . , 8 6 , 1 0 ( 1 9 5 2 ) . 5 5 . 5
23a. B habha , H . J . , P r o c . R o y . S o c . ( L o n d o n ) , A 1 5 2 , 5 5 9 ( 1 9 3 5 ) . 3 2 . 5
24. B habha , H . J . , P r o c . R o y . S o c . ( L o n d o n ) , A 1 5 4 , 1 9 5 ( 1 9 3 6 ) . 3 6 . 2
24a. B ialynicki-B irula , I . , N u o v o c i m e n t o , 1 7 , ' 1 2 2 ( 1 9 6 0 ) . 4 6 . 3
25. Biedenharn , L . C . , B latt, J . M . a n d R ose, M . E . , R e v . M o d . P h y s . , 2 4 , 2 4 9
(1952). 8 .2

26. Bjorken , J . , D rell , S . a n d F rautschi, S . , Ph ys. R e v ., 11 2 , 1409 (1958). 5 2 .5

27. B lokhintsev , D . I . , P r i n c i p l e s o f Q u a n t u m M ec h a n ic s, G o ste k h iz d a t, M oscow ,

1949. 7 .4 , 3 7 .3

28. B o g o l iu b o v , N . N ., D o k la d y A kad. N au k U SSR , 99, 225 (1954). 4 9 .4

[855]
856 QUANTUM ELECTRODYNAMICS

29. B , N. N. andSHiRKOV, D. V., JETP, 30, 77 (1956), Sov. Phys. JETP,


o g o l iu b o v

3, 57 (1956). 48.2, 48.4, 48.5


30. Bogoliubov, N. N. and S hirkov , D. V., Introduction to the Theory of Quantized
Fields, Interscience, New York, 1959. 48.4
31. B ohr , N. and R osenfeld, L., Kgl. Danske Videnskab. Selskab, Mat.-fys.
Medd., 12, No. 8 (1933). 17.1
32. Booth , F. and W ilson , A. H., Proc. Roy. Soc. (London), A 175, 483 (1940).
59.1
33. B reit, G., Phys. Rev., 34, 553 (1929). 38.1
34. Breit , G. and R osenthal, J. E., Phys. Rev., 41, 459 (1932). 13.3
35. Breit , G. and T eller, E., Astrophys. J., 91, 215 (1940). 35.4
36. Brow n , L. and F eynman, R. P., Phys. Rev., 85, 231 (1952). 52.1
37. C aianiello, E. R. and F ubini, S., Nuovo cimento, 9, 1218 (1952). 10.7
38. C asimir, H., Helv. Phys. Acta, 6, 287 (1933). 35.3
39. C hristy , R. F. and K usaka, S., Phys. Rev., 59, 414 (1941). 59.2, 60.1
40. C hurch , E. L. and W eneser, J., Phys. Rev., 103, 1035 (1956). 42.2
41. C ondon , E. U. and S hortley, G. H., Theory of Atomic Spectra, Cambridge,
1935. 4.1
42. D alitz , R., Proc. Roy. Soc. (London), A 206, 509 (1951). 51.1
42a. D ankoff , S. M. and M orrison P. P., Phys. Rev., 55, 423 (1939). 40.3
43. D arw in , C. G., Proc. Roy. Soc. (London), A 118, 654 (1928). 14.5
44. d e Shalit , A., Phys. Rev., 91, 1479 (1953). 8.2
45. D eutsch , M. and D ulit , E., Phys. Rev., 84, 601 (1951). 39.4
46. D irac , P. A. M., Proc. Roy. Soc. (London), A 114, 243, 710 (1927).16.1
47. D irac , P. A. M., Proc. Roy. Soc. (London), A 117, 610; A 118, 351(1928).
9.2
48. D irac , P. A. M., Proc. Cambridge Phil. Soc., 26, 361 (1930), 33.1
49. D irac , P. A. M., The Principles of Quantum Mechanics, 2nd ed., Oxford,
1935; 3rd ed., 1947. 23.1
50. D oggett , J. A. and S pencer , L. V., Phys. Rev., 103, 1597 (1956). 14.4
51. D uffin , R. J., Phys. Rev., 54, 1114 (1938). 20.6, 56.1
52. D yson, F. J., Phys. Rev., 75, 486 (1949). 23.4
53. D yson, F. J., Phys. Rev., 75, 1736 (1949). 43.3, 45.2
54. E dwards , S. F. and P eierls, R. E., Proc. Roy. Soc. (London), A 224, 24 (1954).
49.4
55. E lton , L. and R obertson, H., Proc. Phys. Soc. (London), A 65, 145 (1952).
51.3
56. E uler , H., Ann. d. Physik, 26, 398 (1936). 54.1, 54.3
57. F ano, U., Phys. Rev., 93, 121 (1954). 28.7
58. F ediushin , B., JETP, 22, 140 (1952). 29.10
59. F ermi, E., Rev. Mod. Phys., 4, 87 (1932). 16.4, 34.1
60. F ermi, E. and U hlenbeck, G. E., Phys. Rev., 44, 510 (1933). 33.3
61. F eshbach, H., Phys. Rev., 84, 1206 (1951). 14.4
62. F eynman, R. P„ Phys. Rev., 76, 749, 769 (1949). 25.5, 26.6, 46.2, 47.1, 61.1
63. F eynman, R. P., Phys. Rev., 84, 108 (1951). 49.4
REFERENCES 857

64. F ock , V. A., Z. Physik, 75, 622 (1932); Physik. Z. Sowjetunion, 6, 225 (1934).
7.4
65. F oldy , L. L., Phys. Rev., 83, 688 (1951). 15.3
66. F oldy , L. L. and W outhuysen , S. A., Phys. Rev., 78, 29 (1950). 15.2
67. F omin , P. I., JETP, 35, 707 (1958). 52.4
68. F owler , R. H., Proc. Roy. Soc. (London), A J29, 1 (1930). 42.2
69. F urry , W. H., Phys. Rev., 46, 391 (1934); 77, 550 (1950). 14.5
70. G alanin , A. and P omeranchuk , I. I a ., Doklady Akad. Nauk USSR, 86, 251
(1952). 53.2
71. G ardner , J. W., Proc. Phys. Soc. (London), A 64, 238 (1951). 27.5
72. G aribian , G., JETP, 24, 617 (1953). 29.10
73. G fl’fand , I. and I aglom, A., JETP, 18, 703 (1948). 20.5
74. G el’fand , I. and M inlos, R. A., Doklady Akad. Nauk USSR, 97, 209 (1954).
49.4
75. G el’fand , I. and S hapiro , Z., Uspekhi Mat. Nauk, 7, 3 (1952). G el’fand ,
I. M., M inlos , R. A. and Shapiro , Z., Representations of the Rotation
Group and the Lorentz Group, Fizmatgiz, Moscow, 1958. 8.3
76. G ell-M ann , M. and Low, F., Phys. Rev., 95, 1300 (1954). 43.2, 48.2, 48.4
77. G or ’kov, L. P., JETP, 30, 790 (1956), Sov. Phys. JETP, 3, 762 (1956-57). 48.5
78. G roshev, L. and S hapiro , I. S., Spectroscopy of Atomic Nuclei, Gostekhizdat,
1952. 40.3
79. G unst , S. B. and P age, L. A., Phys. Rev., 92, 970 (1953). 28.7
80. G upta , S., Proc. Phys. Soc. (London), A 63, 681 (1950). 16.5
81. G uzenko , S. and F omin , P., JETP, 38, 513 (1960). 52.4
81a. H all , D. and W ightman , A. S., Kgl. Danske Videnskab. Selskab, Mat.-fys.
Medd., 31, No. 5 (1957). 17.3
82. H alpern , O., Phys. Rev., 44, 855 (1933). 55.1
83. H amilton , D. R., Phys. Rev., 58, 122 (1940). 27.5
84. H arris, I. and B rown , M., Phys. Rev., 105, 1656 (1957). 52.3
85. H eisenberg, W. and E uler , H .,Z . Physik, 98, 714 (1936). 54.4
86. H eitler, W., Proc. Cambridge Phil. Soc., 32, 112 (1936). 6.3
87. H eitler, W., The Quantum Theory of Radiation, 1st ed., Oxford, 1936; 3rd
ed., 1954. 17.1
88. H eitler, W. and N ordheim , L., J. phys. radium, 5, 449 (1934). 32.5
89. H ori , S., Progr. Theoret. Phys. (Kyoto), 7, 578 (1952). 24.4
90. H ough , P„ Phys. Rev., 74, 80 (1948). 29.3
91. H ughes, V. W., M arder , S. and Wu, C. S., Phys. Rev., 106, 934 (1957). 53.1
92. H ulme, H. R. and B habha , H. J., Proc. Roy. Soc. (London), A 146, 723
(1934). 33.3
93. H ulme, H. R., M c D ougall , J., B uckingham , R. A. and F owler , R. H.,
Proc. Roy. Soc. (London), A 149, 131 (1935). 31.2
94. I offe, B. L., JETP, 31, 583 (1956), Sov. Phys. JETP, 4, 534 (1957). 5.4
95. J aeger, J. C. and H ulme, H. R., Proc. Cambridge Phil. Soc., 32, 158 (1936).
33.3
858 QUANTUM ELECTRODYNAMICS

96. J auch , J. and R ohrlich , F., Helv. Phys. Acta, 27, 613 (1954). 51.4
97. J o s e p h , J., Phys. Rev., 103, 481 (1956). 33.6
98. J ost, R., L uttinger , J. and Slotnick , M., Phys. Rev., 80, 189 (1950). 55.2,
55.3.
99. K allen , G., Helv. Phys. Acta, 25, 417 (1952). 43.2
100. K arplus, R. and K lein, A., Phys. Rev., 87, 848 (1952). 53.1
101. K arplus, R., K lein , A. and Schw inger , J., Phys. Rev., 86, 288 (1952). 53.1
102. K arplus, R. and K roll , N., Phys. Rev., 77, 536 (1950).50.2
103. K arplus,R. and N euman, M., Phys. Rev., 80, 380 (1950). 54.1
104. K emmer, N ., Proc. Roy. Soc. (London), A 173, 91 (1939).20.6,56.1
105. K emmer,N. and L udw ig , G., Helv. Phys. Acta, 10, 182 (1937). 55.1
106. K inoshita , T., Progr. Theoret. Phys. (Kyoto), 5, 473 (1950). 57.2
107. K lein, O. and N ishina , Y., Z. Physik, 52, 853 (1929). 28.3
108. K resnin, A. A. and R ozentsveig, L. N., JETP, 32, 353 (1957), Sov. Phys.
JETP, 5, 288 (1957). 36.3
109. K roll, N. and L amb, W., Phys. Rev., 75, 388 (1949). 53.1
110. K rutov , V. A., Izvest. Akad. Nauk SSSR, Ser. Fiz., No. 2,(1958). 40.5
111. L amb, W. E. and R etherford, R. C., Phys. Rev., 72, 241 (1947). 35.4
112. L andau , L. D., Physik. Z. Sowjetunion, 8, 487 (1932). 38.1
113. L andau , L. D„ Doklady Akad. Nauk USSR, 60, 207 (1948). 7.2
114. L andau , L. D., Nucl. Phys., 3, 127 (1957). 4, 22.4
115. L andau , L. D., A brikosov, A. A. and K halatnikov , I. M., Doklady Akad.
Nauk. USSR, 95, 497, 773, 1177 (1954). 46.3, 48.2, 48.3
116. L andau , L. D. and K halatnikow , L M., JETP, 29, 89 (1955), Sov. Phys.
JETP, 2, 69 (1956). 43.2.
117. L andau , L. D. and L ifshits, E. M., Physik Z. Sowjetunion, 6, 244(1934). 32.5
118. L andau , L. D. and L ifshits, E. M., The Classical Theory of Fields, Pergamon
Press-London-Paris, Addison-Wesley, Reading, 1951. 2.4, 16.1, 30.1, 34.1
119. L andau , L. D. and L ifshits, E. M., Quantum Mechanics, Pergamon Press-
London-Paris, Addison-Wesley, Reading, 1958. 2.4, 4.1, 7.4, 8.1, 9.1, 14.1,
29.8, 35.2
120. L andau , L. D. and L ifshits, E. M., Statistical Physics, Pergamon Press-London-
Paris, Addison-Wesley, Reading, 1958. 2.4
121. L andau , L. D. and P eierls, R. E., Z. Physik, 62, 188 (1930). 2.2
122. L andau , L. D. and P omeranchuk, L 1a ., Doklady Akad. Nauk. USSR, 102,
489 (1955). 48.6
123. L ee, T.D. and Y ang , C.N., Phys. Rev., 104,254(1956). 105, 1671 (1957). 9.4, 22.4
124. L ehmann, H., N uovo cimento, 11, 342 (1954). 43.2
125. L ifshits, E. M., JETP, 18, 562 (1948). 29.10
126. L ipps , F. W. and T olhoek , H. A., Physica, 20, 85 (1954). 28.7
127. L loyd, S., Phys. Rev., 83, 716 (1951). 27.5
128. Low, F., Phys. Rev. 88, 53 (1952). 53.4
129. LOders, G., Kgl. Danske Videnskab. Selskab, Mat.-fys. Medd., 28, No. 5,
17 (1954). 21.3
REFERENCES 859

130. M a j o r a n a E., N
, u o v ocimento, 14, 171 (1937). 9.3
131. M a t t h e w s , P., Phys. Rev., 80, 292 (1950). 57.3
132. M a t t h e w s , P. T. and S , A., Nuovo cimento, 12, 563 (1954).
a l a m 49.4
133. M K c , W. A. and F
in l e y , H., Phys. Rev., 74, 1759 (1948).
e s h b a c h 14.4
134. M V c , K. W., Phys. Rev., 110, 1484 (1958).
o y 29.4
135. M ic h e l, L. and W , A. S., Phys. Rev., 98, 1190 (1955).
ig h t m a n 10.6
136. M i g d a l, A., Izvest. Akad. Nauk SSSR IV, No. 2, 287 (1940). 32.5
137. M o l l e r, C., Ann. d. Phys., 14, 531 (1932). 36.1
138. M , N. F„ Proc. Roy. Soc. (London), A 124, 425 (1929).
o t t 14.4
139. M , N. F. and M
o t t , H. S. W., Theory of Atomic Collisions, 2nd ed.
a s s e y

Oxford, 1949. 14.1, 36.1


140. N e m ir o v s k ii , E., JETP, 18, 893 (1948). 32.4
141. N e u m a n , M. and F , W., Phys. Rev., 76, 1677 (1949). 57.1
u r r y

142. N is h in a , V., T , S. and T


o m o n a g a , H., Sci. Pap. Phys. Math. Res. Japan,
a m a k i

24, No. 18 (1934). 33.3


143. N is h in a, V. and T , S., Sci. Pap. J. Phys. Chem. Res. Japan, 27, 137
o m o n a g a

(1935). 32.5
144. N o r d s ie c k , A., Phys. Rev., 93, 785 (1954). 29.8
145. O , H. Phys. Rev., 99, 1335 (1955).
l s e n 29.9
146. O , A. and P
r e , J., Phys. Rev., 75, 1696 (1949).
o w e l l 33.5
147. P , L. A., Phys. Rev., 106, 394 (1957).
a g e 33.2
148. P , W., “Die Allgemeinen Prinzipien der Wellenmechanik” in Handb. Phys.,
a u l i

24, (1), 1933, p. 83. 2.2


149. P , W., Rev. Mod. Phys., 13, 203 (1941).
a u l i 20.4, 21.1
150. P , W., Rev. Mod. Phys., 15, 175 (1943).
a u l i 16.5
151. P , W., Niels Bohr and Development of Physics, Pergamon-London,
a u l i

McGraw-Hill, New York, 1955. 21.3


152. P ir e n n e , J., Arch. Sci. Phys. Nat., 29, 207 (1947). 38.4
153. P o l o v in , R., JETP, 31, 449 (1956), Sov. Phys. JETP, 4, 385 (1957). 10.7, 51.6
154. P o m e r a n c h u k , I. I a . , Doklady Akad. Nauk. USSR, 103, 1005 (1955). 48.6
155. P o m e r a n c h u k , I. I . , N
a cimento, 10, 1186 (1956). 48.6
u o v o

156. P o m e r a n c h u k , I. I . and S
a , I . A., J. Phys. (USSR) 9, 97 (1945).
m o r o d in s k ii a

13.4
157. P r e s e n t, R. a n d C h e n , S . , Phys. Rev., 8 5 , 4 4 7 ( 1 9 5 2 ) . 4 1 . 4

158. P , A., J. phys. radium, 7, 3 4 7 ( 1 9 3 6 ) .


r o c a 2 0 .6

159. R , G., Nature,


a c a h 129, 723 (1932). 13.3
160. R acah , G., N uovo cimento, 13, 66 (1936). 32.1, 55.4
161. R , G., Phys. Rev., 62, 438 (1942).
a c a h 8.2
163. R e d h e a d , M., Proc. Roy. Soc. (London), A 220, 219 (1953). 51.6
164. R , M. E., Phys. Rev., 73, 279 (1948).
o s e 14.4
165. R , M. E., Phys. Rev., 76, 678 (1949).
o s e 41.1
166. R , M. E., Phys. Rev., 82, 389 (1951), 13.4
o s e

166a. S a l a m , A., Nuovo cimento, 5, 299 (1957). 9.4

167. S a l p e t e r , E., Phys. Rev., 89, 92 (1953). 53.1


168. S a l p e t e r , E. and B , H. A., Phys. Rev., 84, 1232 (1951).
e t h e 49.2
860 QUANTUM ELECTRODYNAMICS

169. Sauter , F., Ann. d. Phys., 9. 217, 11, 454 (1931), 31.2
170. Schiff , L. I., S nyder , H. and W einberg , J., Phys. Rev., 57, 315 (1940). 12.4
171. Schwinger , J., Phys. Rev., 73, 416 (1948). 23.4, 54.4
172. Schwinger , J., Phys. Rev., 74, 1439 (1948). 23.1, 23.4, 54.4
173. Schwinger , J., Phys. Rev., 75, 651 (1949). 50.3
174. Schwinger , J., Phys. Rev., 75, 1912 (1949). 50.2
175. Schwinger , J., Phys. Rev., 82, 664 (1951). 61.3
176. Schwinger , J., Phys. Rev., 82, 914 (1951). 21.3
177. Schwinger , J., Proc. Natl. Acad. Sci., 37, 452, 455 (1951) 49.3
178. Shapiro , I. S., JETP, 19, 597 (1949). 41.1
179. Sherman , N., Phys. Rev., 103, 1601 (1956). 14.4
180. S chmushkevich , I. and P omeranchuk , I., Nuclear Phys., 23, 452 (1961). 34.1
181. S liv , L., JETP, 17, 1049 (1947). 13.4
182. S liv , L., Doklady Akad. Nauk USSR, 64, 321 (1949). 41.4
1 8 3 . S liv , L . , J E T P , 2 1 , 7 7 0 ( 1 9 5 1 ) . 4 0 . 4

184. Sliv , L. and B and , I., Tables of internal conversion coefficients for gamma
radiation, Publ. Acad. Sci., USSR, 1956. 40.3
185. Smirnov, V. I., Course of Higher Mathematics, vol. Ill, part 1, Gostekhizdat,
1956. 20.2
186. S morodinskii, Ia . A., JETP, 17, 1034 (1947). 13.3
187. Sommerfeld, A., Atomic Structure and Spectral Lines, Dutton, New York,
1923. 29.1, 29.8, 31.2
188. Sommerfield, C. M., Phys. Rev., 107, 328 (1957). 50.2
189. Sorokin , V., JETP, 18, 228 (1948). 3.2
190. Stearns, M., Phys. Rev., 76, 836 (1949). 29.3
191. Stueckelberg, E. C. G. and P eterman, A., Helv. Phys. Acta, 26, 499 (1953). 48.4
192. S udakov, V. V., JETP, 30, 87 (1954), Sov. Phys. JETP, 3, 65 (1956), and JETP
31, 729 (1956), Sov. Phys. JETP 4, 616 (1957). 48.2, 51.5
193. T amm, I. E., Z. Physik, 62, 545 (1930). 28.3, 33.1
194. T er-M artirosian, K., JETP 20, 925 (1950). 41.3, 42.2
195. T hirring , W., Phil. Mag., 41, 1193 (1950). 45.1
196. T issa, L., JETP, 7, 690 (1937). 32.5
197. T omonaga, S., Progr. Theoret. Phys. (Kyoto), 1, 27 (1946). 32.1
198. T ralli, N. and G oertzel, G., Phys. Rev., 83, 399 (1951). 40.5
199. U ehling , E. A., Phys. Rev., 48, 55 (1935). 47.3
200. V otruba, V., Phys. Rev., 73, 1468 (1948). 32.4
201. Vysotskii, G. L., K resnin, A. A. and R ozentsveig, L. N., JETP, 32, 1078
(1957), Sov. Phys. JETP, 5, 883 (1957). 29.4
202. W ard , J. C., Phys. Rev., 78, 182 (1950). 44.2
203. W ard , J. C., Proc. Phys. Soc., A 64, 54 (1951). 45.2
204. W atson, G. N., A Treatise on the Theory of Bessel Functions, 2nd ed., Cam­
bridge, 1945. 17.3
205. W eisskopf, V., Kgl. Danske Videnskab. Selskab., Mat.-fys. Medd.,XIV, 6(1936).
54.4
REFERENCES 861

206. W , C., Z. Physik, 88, 612 (1934).


e iz s a c k e r 34.1
207. W e l t o n T„ Phys. Rev., 74, 1157 (1948). 50.1
,

208. W , G., Quantum Theory of Wave Fields, Tnterscience, New York,


e n t z e l

1949. 16.1
209. W e y l , H., Z. Physik, 56, 330 (1929). 9.4
210. W i c k, G. C., Phys. Rev., 80, 268 (1950). 24.3
211. Wick, G. C., La Ricerca Scient., 11, 49 (1940). 41.3
211a. W i g h t m a n , A. S., Phys. Rev., 101, 860 (1956). 17.3
212. W i l l i a m s , E. J., Phys. Rev., 45, 729 (1934); Kgl. Danske Videnskab. Selskab.,
Mat.-fys. Medd., 13, 4 (1935). 34.1
214. W , L. and R
o l f e n s t e in a v e n h a l l , D., Phys. Rev., 88, 279 (1953). 23.3
215. Y a n g , L. M., Phil. Mag., 42, 1333 (1952). 10.7
216. Y u k a w a , H. and S a k a t a , S., Proc. Phys. Math. Soc. Japan, 17, 10 (1935).
42.2
217. Z e l ’d o v i c h , I a . B„ Doklady Akad. Nauk USSR, 83, 63 (1952). 28.7
Subject Index

A b so r p tio n c o effic ie n t, p h o to n s , 4 3 2 C o m m u ta tio n


A c tio n , 156, 223 r e la tio n s, 161, 164, 199, 824

A dvanced p o te n tia ls, 5 1 6 C o m m u ta to r, 162

A n g u la r c o r r e la tio n , o f y-ra y s, 3 6 2 o f fie ld c o m p o n e n t s , 1 7 6

A n o m a lo u s m a g n etic m om en t, o f p o ten tia ls, 1 7 4

e le ctro n , 6 9 8 o f s c a la r field c o m p o n e n t s , 8 2 4

n u c le o n . 150 C o m p to n effect, 325

A n tico m m u ta to r, 2 0 0 b o u n d e le c tr o n s, 4 9 2

field c o m p o n e n t s , 2 0 5 d o u b le , 7 4 0

A sy m p to tic b e h a v io r , w a v e C om p ton fo r m u la , 3 6 5

fu n c tio n s, 382 C o n v e r s io n c o effic ie n t, 541

A to m ic form fa cto r, 3 9 5 , 8 0 3 /C -sh ell, 5 4 3

A uger effect, 5 3 7 C u r r en t d e n sity v e cto r, 2 2 8 , 2 3 0

A z im u th a l a sy m m etry , 136 s c a la r fie ld , 2 3 7 , 8 2 1

e le c t r o n -p o s itr o n field , 2 1 1

C u t-o ff fa cto r, 623


B isp in o r , 75
C u t-o ff m o m e n tu m , 6 2 2 , 6 7 4
u n ita r y tr a n sfo r m a tio n s, 7 6

B oh r m agn eton , 698

B oson s, 241
D egree o f p h o to n p o la riza tio n , 17
B re it fo r m u la , 5 1 7
D iffe r e n tia l cro ss s e c tio n , 3 2 8
B re m sstr a h lu n g , 3 7 8
D ir a c e q u a tio n , 7 5 , 7 9
d o u b le , 7 4 7
in v a r ia n ce , 81
energy lo sses, 397
p o sitiv e and n e g a tiv e freq u en cy
from s c a la r p a r tic le s, 8 4 0
so lu tio n s, 87
scr e en in g , 3 9 2
D ira c m a tric e s, 75
D isp e rsio n fo r m u la , 4 8 4
C h a rge-con ju gate fu n c tio n , 89 D isp e rsio n r ela tio n s, 43
C h a r g e c o n ju g a tio n , 88, 2 6 3 D iv e r g e n c e s, ty p es o f, 6 2 0
C h arge-even state, 2 8 3 in e l e c t r o d y n a m i c s o f s c a la r
C h a rg e-o d d state, 2 8 3 p a rtic le s, 8 3 4
C harge p a r ity , 2 8 3

C h r o n o lo g ic a l p ro d u ct, e le c tr o m a g ­

n e tic field o p era to rs, 178 E ffe c tiv e charge, 662, 704

e le c t r o n - p o s it r o n field o p e r a to r s , 2 0 7 E ffe c tiv e e le c tr o n lin es, 595

C la ssic a l e le ctro n “ r a d iu s,” 367, 622 E ffe c tiv e p h o to n lin e s, 5 9 7

C le b sch -G o r d a n c o effic ie n ts, 6 4 E ffe c tiv e p o ten tia l e n e r g y , 693

C o h e r e n t s ca tte rin g , 7 9 3 , 8 0 8 E le c tr ic m u ltip o le m o m en t, 350

C o m b in ed in v e r sio n , 2 6 7 E le c tr ic m u ltip o le r a d ia tio n , 3 4 7

[8 6 3 ]
864 SUBJECT INDEX

Electromagnetic electron Green’s functions, 575


mass, 590, 662 d’Alembert equations, 185
Electron energy spectrum, Dirac equations, 211, 485
continuous, 117 equations for, 588
wave functions in Coulomb in variational derivatives, 683
field, 125 integral equations for, 602, 663
discrete, 117 for two electrons, 679
in Coulomb field, 124
Electron self-energy part, 594 Heisenberg picture, 268
irreducible, 600 “Hole,” 212
reducible, 600 Hypergeometric function, 403
regularized, 610 confluent, 123, 126
second order, 639
strongly connected or compact, 596 Indefinite metric, in Hilbert space, 170
Emission, of photons of long Infinitesimal operator, 217
wavelength, 413 “Infrared catastrophe,” 413
of two photons, 495 Interaction, between electrons and
EL-transition, 348 positrons, 523
Energy-momentum vector, of a exchange, 524
field, 225 Interaction picture, 269
of the electromagnetic field, 158 Internal conversion, of y-rays, 537
of the electron-positron field, 197 Intrinsic parity, 82
EO-transition, 565 Invariance, of Dirac equation, 81
selection rules, 568 of equations of quantum
Equivalent normal electrodynamics, 262
representations, 312 Invariants, of the electromagnetic
Equivalent representations of a field, 156
group, 216 Inversion, 245
Equation of continuity, 81 combined, 267
Exchange diagram, 512 total, 249, 263
Exchange interaction, 512, 524 weak, 266
Excitation, of nuclei by electrons, 562 Irreducible tensors, 62
Isomers, 356
Fermions, 241 Isotopic shift, 128
Feynman diagrams, 307
Feynman’s rules, 324 Lagrangian, 155, 223
Field equations, in Poisson bracket Lagrangian density, 156
form, 260 conditions for relativistic
for scalar particles, 819 invariance, 230
Fine structure, 124 for electromagnetic field, 158, 781
Fock’s functional, 61 radiation corrections, 780, 850
Furry’s theorem, 322 for electron-positron field, 195
for interacting fields, 256, 258, 822
Gauge transformation, 155, 262 for scalar fields, 236, 820
Gradient transformation. See gauge Longitudinal photons, 52
transformation Lorentz condition, 48, 155
SUBJECT INDEX 865
Lorentz group, 214 vacuum polarization, 589
full, 215 Optical theorem, 41, 796
irreducible representations, 217 Order of diagram, 309
proper, 215 Orthopositronium, 465
representation, 216 decay probability, 467
L-vector, 62
contravariant components, 63 Pair annihilation, 457
covariant components, 63 of longitudinally polarized pair, 461
Magnetic multipole moment, 352 one-photon, 463
Magnetic multipole radiation, 351 three-photon, 467
Majorana representation, 78 two-photon, 458, 844
Mass operator, 589 correlation of photon
regularized, 610 polarizations, 461
Matrix elements, of field radiation corrections, 744
with multiple photon
operators, 292, 572-575
production, 470
Maxwell’s equations, 2
Pair production, 438
Metastable nuclear states, 355
by two photons, 448
Metastable state, of hydrogen
in collisions of two charged
atom, 495
particles, 452
Method of equivalent photons, 473
in field of a nucleus, 438, 482
Mixed T-product, 296
in photon-electron collisions, 451
A/L-transition, 348 in radioactive decay, 457
Monoenergetic positrons, 564 of scalar particles, 843
/x-Meson, 508 Pairing, of electromagnetic field
Natural line width, 758 operators, 179
Nonlocal interaction, 676 of electron-positron field
Norm of a function, in Hilbert operators, 209
space, 169 of field operators, 297
Normal product, of operators of the of scalar field operators, 827
electromagnetic field, 178 Parametrization method, 632
of operators of the electron- Parapositronium, 465
positron field, 208 decay probability, 466
Nuclear magneton, 150 Parity, charge, 283
electron states, 109
Occupation numbers, 153 intrinsic, 82
Operator, charge conjugation, 280 photon states, 29
electron annihilation, 203 states of two-photon system, 56
electron creation, 203 Pauli equation, 145
infinitesimal, 217 Pauli exclusion principle, 198
mass, 589 Pauli matrices, 73
photon annihilation, 168 Pauli theorem, 241
photon creation, 168 Perturbation theory, 284
positron annihilation, 203 criterion for applicability, 658
positron creation, 203 for continuous spectrum, 378
projection, 99 PhotoefTect, 429
866 SUBJECT INDEX

Photon emission, 345 renormalized, 609, 631


Photon “mass,” 423 photon, 575, 584
Photon self-energy part, 594 in functional integral form, 689
irreducible, 600 in presence of external fields, 679
reducible, 600 renormalized, 609, 631
second order, 643 scalar particle, 826
strongly connected or compact, 597
Photon spin, 21
Quantization rules for fields, 154
Photon states, electric type, 30, 35
for the electromagnetic field, 161
magnetic, 30, 35
for the electron-positron field, 198
Physical electron charge, 605, 663
Picture, Heisenberg, 268 for the scalar field, 824
Picture, interaction, 269 Quantum electrodynamics,
Picture, Schrodinger, 268 fundamental equations, 255
Poisson brackets, 260
Polarization, of bremsstrahlen, 389 Racah coefficients, 67
of electrons, 95 Radiation, from a nucleus, 353
on scattering, 136 angular distribution, 358
of nucleus correlated with photon polarization, 359
emission, 361 Radiation corrections, 693
of photons, circular, 12 to bremsstrahlung, 747
elliptical, 14 to Compton effect, 731
linear, 12 to Coulomb’s law, 702, 849
longitudinal, 50, 159 to electron scattering, 705
scalar, 160 to Lagrangian of electromagnetic
of radiation, 356 field, 780, 850
of scattered photons, 374, 377 to two-photon pair annihilation, 744
Polarization density matrix, for Radiation shift, of atomic levels, 751
electrons, 97 of hydrogen-atom levels, 754
for nuclei, 356 of ^-mesic atom levels, 756
for photons, 13 of positronium levels, 756
Polarization moment, of nucleus, 356 Raman scattering, 486
Polarization operator, 589 Recoil electrons, 372
regularized, 610 Relativislically invariant field
Positronium, 527 equations, 229
decay, 465 Renormalizability, 659
fine structure, 530 Renormalization, of electron mass, 604
levels, 528, 530 of matrix elements, 613
Zeeman effect, 535 Representation of a group, 216
Potentials, of the electromagnetic dimension, 216
field, 46, 159 direct product, 220
Probability of a scattering process, 327 irreducible, 217
for polarized particles, 336 reducible, 217
Propagator, electron, 575, 584 spinor, 221
in functional integral form, 689 tensor, 221
in presence of external fields, 678 Majorana, 78
SUBJECT INDEX 867

Resonance scattering, of States, charge-even, 283


photons, 491, 761 charge-odd, 283
Retarded potentials, 514 single charge, 573
Rutherford formula, 139, 501 Stokes’ parameters, 15
Strong inversion, 249
Scalar photons, 51
Scattering, electron-electron, 499 Tensor, 221
of electrons, 131 angular momentum, 226, 230
in Coulomb field, 138 electromagnetic field, 154
polarized, 504 energy-momentum, 224, 230
photon-photon, 764, 850 canonical, 227
of photons, 36 for electromagnetic field, 158
by bound electrons, 484 for electron-positron field, 196
by Coulomb field, 792 for interacting fields, 258
by free electrons, 363 for scalar field, 237, 820
polarized, 373 symmetric, 227
by polarized electrons, 376 photon-photon scattering, 767
by scalar particles, 838 regularized, 768
positron-electron, 502 polarization, fourth rank, 851
of positrons, vacuum polarization, scalar
in Coulomb field, 138 field, 847
polarized, 506 regularized, 848
of scalar particles, 835 Thomson formula, 370
Scattering amplitude, 40 Time reversal, 245, 302
spinor, 131 Total inversion, 263
Scattering matrix, 290 Transition, electric 2L pole, 348
momentum representation, 316 magnetic 2L pole, 348
in scalar electrodynamics, 827 monopole, 565
symmetry properties, 302 Transition currents, 515
Scattering phase, 39, 134 Two-electron self-energy part, 680
Schrodinger picture, 268
Screening, 392
Second quantization, 153 Uncertainty relation, 177
Selection rules, for radiation, 355
Self-energy functions, 586 Vacuum, 572
Single charge state, 573 dielectric permittivity tensor, 781
Singular functions, 185 electromagnetic, 167, 174
S-matrix, 290 electron positron, 202
Spherical harmonics, 68 Vacuum electrons, 781
generalized, 72 magnetic permeability tensor, 781
spinor, 106 Vacuum parts of a diagram, 594
Spinor, 73, 221 scalar field, 826
contravariant components, 106 Variational derivative, 61
covariant components, 106 Variational principle, 155, 195, 223
inversion, 74 Vector spherical harmonic, 24
second rank, 85 contravariant components, 25
868 SUBJECT INDEX

covariant components, 26 third order, 649


longitudinal, 27 weakly connected or
transverse, 28 noncompact, 598
Vertex, of a diagram, 308 Virtual particles, 308, 318
two-line, 604
Vertex function, 599 Ward’s identity, 602
three-photon, 613 Wave function, for photons, 2, 11,31
Vertex part of a diagram, 598 arbitrary number of photons, 59
irreducible, 600 normalization, 7
reducible, 600 spin, 21
regularized, 609 two-photon system, 52
strongly connected or compact, 598 for positron, 92
three-photon, 611

You might also like