100% found this document useful (1 vote)
322 views439 pages

Nonlocal Quantum Field Theory and Stochastic Quantum Mechanics (PDFDrive)

Nonlocal Quantum Field Theory
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
322 views439 pages

Nonlocal Quantum Field Theory and Stochastic Quantum Mechanics (PDFDrive)

Nonlocal Quantum Field Theory
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 439

Nonlocal Quantum Field Theory and

Stochastic Quantum Mechanics


Fundamental Theories of Physics
A New International Book Series on the Fundamental Theories of
Physics: Their Clarification, Development and Application

Editor: ALWYN VAN DER MERWE


University of Denver, U.S.A.

Editorial Advisory Board:


ASIM BAR UT, University of Colorado, U.S.A.
HERMANN BONDI, Natural Environment Research Council, u.K.
BRIAN D. JOSEPHSON, University of Cambridge, u.K.
CLIVE KILMISTER, University of London, U.K.
GUNTER LUDWIG, Philipps-Universitat, Marburg, F.R.G.
NATHAN ROSEN, Israel Institute of Technology, Israel
MENDEL SACHS, State University of New York at Buffalo, U.S.A.
ABDUS SALAM, International Centre for Theoretical Physics, Trieste, Italy
HANS-JURG EN TREDER, Zentralinstitut fur Astrophysik der Akademie
der Wissenschaften, G.D.R.
N onlocal Quantum Field
Theory and Stochastic
QuantuIn Mechanics
by
Khavtgain Namsrai
Institute of Physics and Technology, Academy of Sciences,
Mongolian People's Republic, and
Joint Institute for Nuclear Research, Dubna, U.S.S.R.

D. Reidel Publishing Company


A MEMBER OF THE KLUWER ACADEMIC PUBLISHERS GROUP

Dordr.echt / Boston / Lancaster / Tokyo


*
Library of Congress Cataloging-in-Publication Data
Namsrai, Khavtgain, 1943~
aE
Nonlocal quantum field theory and stochastic quantum mechanics.

(Fundamental theories of physics)


Bibliography: p.
Includes index.
1. Quantum field theory. 2. Stochastic processes. 3. Space and time. 4.
Electromagnetic interactions. 5. Weak interactions (Nuclear physics) I. Title. II.
Series.
QC174.45.N33 1986 530.1'43 85-25617
ISBN-13: 978-94-010-8513-7 e-lSBN-13: 978-94-009-4518-0
DOl: 10.1007/978-94-009-4518-0

Published by D. Reidel Publishing Company,


P.O. Box 17,3300 AA Dordrecht, Holland.

Sold and distributed in the U.S.A. and Canada


by Kluwer Academic Publishers,
190 Old Derby Street, Hingham, MA 02043, U.S.A.

In all other countries, sold and distributed


by Kluwer Academic Publishers Group,
P.O. Box 322,3300 AH Dordrecht, Holland.

All Rights Reserved

© 1986 by D. Reidel Publishing Company, Dordrecht, Holland


Softcover reprint of the hardcover 1st edition 1986
No part of the material protected by this copyright notice may be reproduced or utilized in any
form or by any means, electronic or mechanical, including photocopying, recording or by any
information storage and retrieval system, without written permission from the copyright owner
Dedicated
to my father
Damdiny Khavtgai
Table of Contents

Preface xiii

Notations used in the book xvii

Part I: Nonlocal Quantum Field Theory 1

Chapter IIFoundation of the Nonlocal Model of Quantized Fields 3


1.1. Introduction 3
1.2. Stochastic Space-Time 7
1.3. The Method of Averaging m Stochastic Space-Time and
Nonlocality 11
1.4. The Class of Test Functions and Generalized Functions 14
1.4.1. Introduction 15
1.4.2. Space of Test Functions 16
1.4.3. Linear Functionals and Generalized Functions 19
1.4.3a. General Definition 19
1.4.3b. Transformation of the Arguments and
Differentiation of the Generalized Functions 21
1.4.3c. The Fourier Transform of Generalized
Functions 23
1.4.3d. Multiplication of the Generalized Functions
by a Smooth Function and Their Convolution 25
1.4.4. Generalized Functions of Quantum Field Theory 27
1.4.5. The Class of Test Functions in the Nonlocal Case 34
1.4.6. The Class of Generalized Functions in the N onlocal
Case 39

Chapter 2/The Basic Problems of Nonlocal Quantum Field Theory 46


2.1. Nonlocality and the Interaction Lagrangian 46
2.2. Quantization of Nonlocal Field Theory 47
2.2.1. Formulation of the Quantization Problem 48

Vll
viii Table of Contents

2.2.2. Regularization Procedure 51


2.2.3. Quantization of the Regularized Equation 52
2.2.4. Green Functions of the Field 4l(x) 57
2.2.5. The Interacting System Before Removal of the
Regularization 59
2.2.6. The Green Functions in the Limit 15--.0 61
2.3. The Physical Meaning of the Form Factors 64
2.4. The Causality Condition and Unitarity of the S-Matrix in
Nonlocal Quantum Field Theory 71
2.4.1. Introduction 71
2.4.2. The Causality Condition 73
2.4.3. The Scheme of Proof of Unitarity of the S-Matrix in
Perturbation Theory 78
2.4.4. An Intermediate Regularization Scheme 81
2.4.5. Proof of the U nitarity of the S-Matrix in a
Functional Form 86
2.5. The Schrodinger Equation in Quantum Field Theory with
Nonlocal Interactions 91
2.5.1. Introduction 91
2.5.2. The Field Operator at Imaginary Time 93
2.5.3. The State Space at Imaginary Time 95
2.5.4. The Interaction Hamiltonian and the Evolution
Equation 97
2.5.5. Appendix A 101

Chapter 3/Electromagnetic Interactions in Stochastic Space-Time 104


3.1. Introduction 104
3.2. Gauge Invariance of the Theory and Generalization of
Kroll's Procedure 105
3.3. The Interaction Lagrangian and the Construction of the S-
Matrix 111
3.4. Construction of a Perturbation Series for the S-Matrix in
Quantum Electrodynamics 113
3.4.1. The Diagrams of Vacuum Polarization 113
3.4.2. The Diagram of Self-Energy 115
3.4.3. The Vertex Diagram and the Corrections to the
Anomalous Magnetic Moment (AMM) of
Leptons and to the Lamb Shift 116
3.5 The Electrodynamics of Particles with Spins 0 and 1 120
3.5.1. Introduction 120
3.5.2. The Diagrams of the Vacuum Polarization of Boson
Fields 126
3.5.3. The Self-Energy of Bosons 130
Table of Contents IX

Chapter 4/Four-Fermion Weak Interactions in Stochastic Space-


Time 134

4.1. Introduction 134


4.2. Gauge Invariance for the S-Matrix in Stochastic-Nonlocal
Theory of Weak Interactions 135
4.3. Calculation of the 'Weak' Corrections to the Anomalous
Magnetic Moment (AMM) of Leptons 138
4.4. Some Consequences of Neutrino Oscillations in Stochastic-
Nonlocal Theory 140
4.4.1. Introduction 140
4.4.2. The p,-+3e Decay 142
4.4.3. The K2-+p,e Decay 144
4.5. Neutrino Electromagnetic Properties in the Stochastic-
N onlocal Theory of Weak Interactions 146
4.6. Studies of the Decay K2-+p,+ p,- and K2- and Kf-Meson
Mass Difference 152
4.6.1. Introduction 152
4.6.2. The K2-+p,+ p,- Decay 153
4.6.3. The Mass Difference of K2 - and Kf -Mesons 154
4.7. Appendix B. Calculation of the Contour Integral 155

Chapter 5/Functional Integral Techniques in Quantum Field Theory 157


5.1. Mathematical Preliminaries 157
5.2. Historical Background of Path Integrals 161
5.3. Analysis on a Finite-Dimensional Grassmann Algebra 165
5.3.1. Definition 165
5.3.2. Derivatives 166
5.3.3. Integration over a Grassmann Algebra (Finite-
Dimensional Case) 167
5.4. Grassmann Algebra with an Infinite Number of Generators 171
5.4.1. Definition 171
5.4.2. Grassmann Algebra with Involution 173
5.4.3. Functional (or Variational) Derivatives 174
5.4.4. Continual (or Functional) Integrals over the
Grassmann Algebra (Formal Definition) 175
5.4.5. Examples 177
5.5. Functional Integral and the S-Matrix Theory 179
5.5.1. Introduction 179
5.5.2. Functional Integral over a Bose Field in the Case of
Nonlocal-Stochastic Theory (Definition) 184
5.5.2a. Definition of Functional Integral 184
5.5.2b. Upper and Lower Bounds of Vacuum Energy
E(g) in Nonlocal Theory and in the
Anharmonic Oscillator Case 189
x Tableo! Contents

5.5.3. Functional Integrals for Fermions in Quantum Field


Theory 202

Part II: Stochastic Quantum Mechanics and Fields 21)

Chapter 6/The Basic Concepts of Random Processes and Stochastic


Calculus 213
6.1. Events 213
6.2. Probability 214
6.3. Random Variable 214
6.4. Expectation and Concept of Convergence over the
Probability 217
6.5. Independence 219
6.6. Conditional Probability and Conditional (Mathematical)
Expectation 220
6.7. Martingales 224
6.8. Definition of Random Processes and Gaussian Processes 225
6.9. Stochastic Processes with Independent Increments 229
6.10. Markov Processes 232
6.11. Wiener Processes 239
6.12. Functionals of Stochastic Processes and Stochastic Calculus 244

Chapter 7/Basic Ideas of Stochastic Quantization 253


7.1. Introduction 253
7.2. The Hypothesis of Space-Time Stochasticity as the Origin of
Stochasticity in Physics 255
7.3. Stochastic Space and Random Walk 260
7.4. The Main Prescriptions of Stochastic Quantization 263
7.5. Stochastic Field Theory and its Connection with Euclidean
Field Theory 268
7.6. Euclidean Quantum Field Theory 270

Chapter 8/Stochastic Mechanics 275


8.1. Introduction 275
8.2. Equations of Motion of a Nonrelativistic Particle 277
8.3. Relativistic Dynamics of Stochastic Particles 283
8.4. The Two-Body Problem in Stochastic Theory 287
8.4.1. The Nonrelativistic Case 287
8.4.2. The Relativistic Case 291
Table of Contents Xl

Chapter 9/Selected Topics in Stochastic Mechanics 296


9.1. A Stochastic Derivation of the Sivashinsky Equation for the
Self-Turbulent Motion of a Free Particle 296
9.2. Relativistic Feynman-Type Integrals 301
9.2.1. Diffusion Process in Real Time 302
9.2.2. 'Diffusion Process' in Complex Time 303
9.2.3. Introduction of Interactions into the Scheme 304
9.3. Discussion of the Equations of Motion in Stochastic
Mechanics 306
9.4. Cauchy Problem for the Diffusion Equation 309
9.5. Position-Momentum Uncertainty Relations in Stochastic
Mechanics 316
9.6. Appendix C. Concept of the 'Differential Form' and
Directional Derivative 321

Chapter lO/Further Del'elopments in Stochastic Quantization 324


10.1. Introduction 324
10.2 Davidson's Model for Free Scalar Field Theory 327
10.3. The Electromagnetic Field as a Stochastic Process 331
10.4. Stochastic Quantization of the Gauge Theories 336
10.4.1. Introduction' 336
10.4.2. Another Stochastic Quantization Scheme 337
10.5. Equivalence of Stochastic and Canonical Quantization in
Perturbation Theory in the Case of Gauge Theories 342
10.6. The Mechanism of the Vacuum Tunneling Phenomena in the
Framework of Stochastic Quantization 347
10.7. Stochastic Fluctuations of the Classical Yang-Mills Fields 352.
10.8. Appendix D. Solutions to the Free Fokker-Planck Equation 355

Chapter II/Some Physical Consequences of the Hypothesis of


Stochastic Space-Time and the Fundamental Length 357
11.1. Prologue 357
11.2. Nonlocal-Stochastic Model for Free Scalar Field Theory 359
11.3. Zero-Point Electromagnetic Field and the Connection
Between the Value of the Fundamental Length and the
Density of Matter 362
11.4. Hierarchical Scales and 'Family' of Black Holes 365
11.5. The Decay of the Proton and the Fundamental Length 369
11.6. A Hypothesis of Nonlocality of Space-Time Metric and its
Consequences 373
11.7. On the Origin of Cosmic Rays and the Value of the
Fundamental Length 384
xii Table of Contents

11.8. Space-Time Structure near Particles and its Influence on


Particle Behavior 396
11.8.1. Introduction 396
11.8.2. Stochastic Behavior of Particles and its Connection
with Stochastic Mechanical Dynamics 397
11.8.3. Soliton-Like Behavior of Particles 400

Bibliography 404

Index 421
Preface·

In this book we attempt to give a physical foundation for the concept of


nonlocality and to introduce the fundamental length in physics by the
hypothesis of stochasticity of space-time in the microworld. The structure of
space-time has always played an important role in physical theory. It is used
to simplify concepts connected with symmetry and to give understanding
about a great many complicated phenomena in a unified way, in terms of a few
simple principles. For example, Einstein's courageous idea of unification of
space and time was in fact a starting point for the general theory of gravitation
and became the cornerstone of modern relativistic quantum theory. At
present, further development of high energy physic~l experiments and theories
dictates a deeper level in the understanding of the structure of space-time and
its properties at small distances. However, our concepts of space-time are
confirmed experimentally to be valid to distances of the order 10- 15_
10- 16 cm.
Among the proposed properties of space-time at small distances [for
example, the concept of superspace and pregeometry, lattice (discrete) and
cellular structures of space-time, higher-dimensional geometry, etc.] an
important role is played by the idea of the stochastic character of space-time.
This idea is based on the fact that the quantum fluctuations in geometry are
inescapable if one believes in the quantum principle and Einstein's theory. A
possible deeper connection between the concept of nonlocality and the
hypothesis of space-time stochasticity, originally due to D. I. Blokhintsev,
stimulates the physical foundation for the nonlocal quantum field theories
constructed by G. Wataghin, A. Pais and G. E. Uhlenbeck, H. Yukawa, G. V.
Efimov, and others.
This book is devoted to the construction of a self-consistent model of
nonlocal quantum field theory and stochastic quantum mechanics based on
the hypothesis of space-time stochasticity and non locality in the microworld.
In that case the occurrence of form-factors in the theory (i.e., violation of the
concept of locality at small distances) and the random behavior of a physical
system are caused by the stochastic nature of space-time on a small scale. The
averaging of any fields independent of their nature (i.e., mass, spin, charge, etc.)

xiii
XIV Preface

over this stochastic space-time leads to the non local fields considered by G. V.
Efimov. In other words, stochasticity of space-time (after being averaged on a
large scale) as a self-memory makes the theory nonlocal. This allows one to
consider in a unified way the effect of stochasticity (or nonlocality) in all
physical processes. Moreover, the universal character of this hypothesis of
space-time at small distances enables us to re-interpret the dynamics of
stochastic particles and to study some important problems of the theory of
stochastic processes [such as the relativistic description of diffusion, Feynman-
type processes, and the problem of the origin of self-turbulence in the motion
of free particles within nonlinear (stochastic) mechanics]. In this direction our
approach (Part II) may be useful in recent developments of the stochastic
interpretation of quantum mechanics and fields due to E. Nelson, D. Kershaw,
I. Fenyes, F. Guerra, de la Pena-Auerbach, J.-P. Vigier, M. Davidson, and
others. In particular, as shown by N. Cufaro Petroni and J.-P. Vigier, within
the discussed approach, a causal action-at-distance interpretation of a series of
experiments by A. Aspect and his co-workers indicating a possible non locality
property of quantum mechanics, may also be obtained. Aspect's results have
recently inspired a great interest in different nonlocal theories and models
devoted to an understanding of the implications of this nonlocality.
This book consists of two parts. The first part is an attempt to give the basic
principles of non local theory of quantized fields and to construct the theory of
electromagnetic and four-fermion weak interactions of leptons from the
stochastic space-time point of view. In the second part, we present to the
reader the main ideas of stochastic quantum mechanics and stochastic
quantization methods for field theory. The core of the book is based on the
original results obtained during the years (1877-83) in collaboration with
Professor G. V. Efimov and his students; it covers nonlocal theory of
quantized fields, the relativistic and nonrelativistic dynamics of stochastic
particles, stochastic quantum mechanics and quantum fields. Since the book is
addressed primarily to theoretical and mathematical physicists, it emphasizes a
conceptual structure - the definition and formulation of the problem and
techniques of solution. Since the emphasis differs from that of conventional
physics texts, students may find this book useful as a supplement to their texts.
In both parts we include the mathematical fundamentals of quantum field
theory and stochastic processes: generalized function and functional methods,
and their applications in quantum field theory, the basic concepts of pro-
bability theory and random processes, and some new developments of field
theory and stochastic quantization methods. We have organized the book so
that it will be useful to new students as well as to experienced researchers. It is
neither self-contained nor complete, but it intends to develop the central ideas,
to explain the main results of a physical and mathematical nature, and to
provide an introduction to the related literature.
Physicists concerned with condensed matter may be interested in the
discussion of the stochastic solution to the classical Yang-Mills equation and
Preface xv

of vacuum structure; the behavior of vacuum energy density in the strong


coupling limit; the mechanism of vacuum tunneling phenomena in the non-
Abelian gauge theories; and some problems of cosmic rays (their acceleration
mechanism and the energy spectrum, and the ratio of the intensities of the
electron component to the proton component at the same energy level). The
central matters are functional integral techniques, stochastic quantization and
space-time metric fluctuational methods.
I would like to thank Professor G. V. Efimov (JINR) for helpful and
stimulating discussions, valuable comments, and for teaching me the principles
of scientific work. The author is truly indebted to Professors N. N. Bogolubov
and V. A. Meshcheryakov for their hospitality at the Joint Institute for
Nuclear Research, Dubna. I wish to thank Doctors M. Dineykhan (Mongolian
Academy of Sciences, Ulan-Bator, Mongolia) and P. Exner (JINR and Nuclear
Centre, Charles University, Prague, Czechoslovakia) for useful comments and
help in preparing the manuscript. I would also like to thank Professors Ch.
Tseren and Sh. Tsegmid (Academy of Sciences, Ulan-Bator, Mongolia); N.
Sodnom, E. Damdinsuren, and O. Lkhavga (Mongolian State Univ.
Ulan-Bator, Mongolia); B. M. Barbashov, V. G. Kadyshevsky and A. N. Sisak-
hyan (JINR); V. Ya. Fainberg and D. A. Kirzhnits (Lebedev Physical Insti-
tute, Moscow, USSR), Ya. A. Smorodinskii (Kurchatov Institute, Moscow,
USSR); N. E. Turin (High Energy Institute, Serpukhov, USSR); G. M. Zino-
vjev (Institute of Theoretical Physics, Kiev, USSR); P. N. Bogolubov (Insti-
tute for Nuclear Research, Academy of Sciences, Moscow, USSR); Hans-J.
Treder (German Academy of Sciences, Babelsberg, DDR); Alwyn van der
Merwe (Univ. of Denver, Colorado, USA); and J.-P. Vigier (Henri Poincare
Institute, Paris, France) for constant interest in my work and valuable
comments and encouragement in my investigations.
The author wishes to thank D. Reidel's referees for their valuable comments,
and Miss D. Bond and G. G. Sandukovskaya who have corrected the
grammar and style of the English version of the manuscript.
Finally, I wish to thank my mother S. Myadag, sister Kh. Tseveg, wife Dz.
Tserendulam, and children Nyamtseren and Tsedevsuren for the warmth of
their love and support.

KHA VTGAIN NAMSRAI


Dubna- V lan- Bator
September, 1984.
Notations used In the book

To introduce the general notation, all four-component vectors are chosen real.
The metric is given by the Minkowski one, gl'v (fl, v = 0, 1,2,3), for which
gl'v=O for fl#V,
goo = -gll = -g2Z = -g33 = 1.
The product of two four-vectors P and q with components
P = (Po,p) = (PO,PI,PZ,P3)'
is defined as
(Pq) = Pl'qvgl'v = Pl'ql' = Poqo - pq = Poqo - Plql - P2q2 - P3q3·
Summation is carried out by repeating indices, omitting the symbol of
summation.
Derivatives will be noted as follows:

i~
a = i 01' = PI' (fl = 0, 1,2,3),
uXI'
D = pZ = -o~ + OZ = _oz /ox~ + 02/oxi + 02/0X~ + oZ/ox~,
aS = 01'1·· . aI's = oS/OXI'I·· ·oxl',.
The equivalent notation
!(X) = !(x o, x) = !(x o ,Xl>X Z,x 3)

will be used for a function !(x) defined in four-dimensional space-time.


Sometimes Xo = t will be used.
The following equivalent notations

f d4X!(X) = fdX O fdX!(X o, x)= fdX Ofd 3 X!(Xo, x)


= fdXO fdX fdX
I Z f dX3 !(Xo, Xl' X2, X3)
XVll
xviii Notations used in the book

are used for integrals of some function f(x) over all four-dimensional space-
time.
Four- and three-dimensional <5-function will be read
<5(4)(X) = <5(x) = <5(x o) <5(3)(X) = <5(xo) <5(x) = <5(xo) <5(x 1 ) <5(X2) <5(X3),
omitting upper indices (3) and (4).
The Dirac YI'-matrices are chosen in the form:
YI' = (Yo, Y), (f.1 = 0,1,2,3),
Y; = 1,

fi = PI'YI' = PoYo - PY·


In this book the unit system (with rare exceptions) in which the light
velocity c and the Planck constant II divided by 2n (II = h/2n) are both equal
to unity, i.e., II = c = 1, is used. In this unit system, energy, momentum and mass
have the dimension of inverse length and time Xo = t has the dimension of
length.
Open sets in the real n-dimensional space /R" will be denoted by G, n, r, etc.,
and G, n, f, etc., are then their closure in /R". In this notation S = G \G means
boundary of the set G.
Sometimes the symbols V and j will be used, which mean: V - 'for any' and
j - "there exist". For example, V a E r, f(a) = 1 means 'for any a belonging to r,
f(a) = 1 always', and j a E r, f(a) = 1 means 'there exists an a belonging to r,
for whichf(a) = 1'.
The symbol [a J means the enteger part of the number a, for example, [5J = 5,
[5.3J = 5, [nJ = 3.
All other mathematical symbols and notations used in the book should be
self-explanatory.
Part I

Nonlocal Quantum Field Theory

Part I is devoted to the construction of a non local theory of quantized fields


and presents systematically the results of studies in this direction. In Chapter 1,
a method is formulated for introducing stochasticity into space-time with
indefinite metric (i.e., Minkowski space) and of averaging in it. This method
allows us to solve the problems of conserving relativistic invariance and of
eliminating ultraviolet divergences in constructing quantum field theory from
a stochastic space-time point of view. A class of test functions and form factors
(or generalized functions) arising from the averaging prescription in a mea-
sured stochastic space-time is also considered.
Chapter 2 deals with the formulation and proof of basic principles (the
quantization problem, the causality condition and unitarity) of the relativistic
quantum field theory in the case of non local interactions. The physical
meaning of the form factors is discussed too. It is shown that the finite and
unitary S-matrix describing the nonlocal interactions of quantized fields is a
solution of the Cauchy problem for the evolution equation (Schrodinger
equation in the interaction picture at imaginary time) with retardation.
The finite S-matrix obeying the basic theoretical principles is constructed for
the case of the electromagnetic and four-fermion weak interactions (Chapters 3
and 4). The different characteristics of electromagnetic and weak processes are
calculated and restrictions on the magnitude of the fundamental length are
obtained.
Functional methods and their applications in quantum field theory are
considered in Chapter 5. Here attention is focussed on functional analysis of
Grassmann algebra and on a definition of the functional integrals in quantum
field theory. Vacuum energy estimates have been obtained in the cp:l"-nonlocal
model and the anharmonic oscillator case.

1
Chapter 1

Foundation of the N onlocal Model


of Quantized Fields

1.1. Introduction

The fundamental problems of quantum field theory (QFT) are a unified


description of particle physics phenomena and a strict mathematical for-
mulation of the basic principles of the theory. One of the directions in which
these problems are investigated, is called the axiomatic approach in QFT.
Bogolubov, Medvedev and Polivanov's (BMP) formulation, for example, is
based on the variational derivative method of the S-matrix and on the
causality condition. Another field approach is connected with names of
Garding, Wightman, Haag, Lehmann, Symanzik and Zimmermann, etc. The
mathematical fundamentals of QFT were given in a series of monographs [see,
for example, Bogolubov et al. (1958), Streater and Wightman (1964), lost
(1965), and Bogolubov et al. (1975)].
Progress achieved in QFT in the last decade is connected with the develop-
ments of non-Abelian gauge jield theories (in particular, the one introduced by
Yang and Mills, 1954) and with the constructive quantumjield theory advanced
by Glimm and laffe and their followers. The former have opened a way to the
unification of all interactions existing in nature, in particular, the link between
electromagnetism and the weak interaction (made known by S. Weinberg,
1967, 1980; A. Salam, 1968, 1980; and Sh. Glashow, 1961, 1980) in what is now
being called 'electroweak' interaction. In the electroweak theory, the weak and
electromagnetic interactions are described in a unified way in terms of an exact
but spontaneously broken gauge symmetry. Evidence for the discovery of the
W±, ZO-bosons with the masses mw = 81 ±2 GeV and mzo = 93 ±2 GeV
(which coincide exactly with the theoretical predictions), which are indispensable
to the electro weak theory, was presented in 1983 by two large experimental
groups called UA1 and UA2 working at CERN.
The constructive quantum field theory has been enriched by new powerful
methods which rely on the ideas from Euclidean field theory and use pro-
babilistic techniques and concepts. A central role is played by the Euclidean-
Markov field qJ(f), introduced by Schwinger (1959), Nakano (1959) and
Symanzik (1966, 1969). Further, Nelson (1971, 1973) described the Euclidean

3
4 Chapter 1

boson quantum field theory as a Markov process: the Euclidean fields are
random variables and the Green functions are expectations of products of
them (see Part II). In Nelson's formulation, the Markov property is crucial
(Chapter 6) for the reconstruction of the physical theory from the Euclidean
theory and for the study of the Euclidean theory itself. Osterwalder and
Schrader (1973, 1975) and Glaser (1974) established the necessary and suf-
ficient conditions for Euclidean Green's functions to define a unique
Wightman field theory. Thus the Euclidean methods in QFT have generated a
constructive Euclidean field theory which is the synthesis of the Nakano-
Schwinger Euclidean field theory with the Feynman-Wiener functional in-
tegral theory at an imaginary time and, on a very deep level, represents the
joining of QFT (at least, the theory of scalar boson fields) with the theory of
the Markov process. This interrelation allows one to insert the theory of boson
fields into classical statistical mechanics.
Recently, a profound connection between quantum theory and the theory of
stochastic processes has been established, more explicitly with the develop-
ments towards stochastic mechanics and its relations with quantum mechanics,
originally proposed by Fenyes (1952) and Kershaw (1964), neatly formulated
by Nelson (1966, 1967), and successively extended by many authors: de la
Pella-Auerbach and Cetto (1975), Dankel (1970), Caubet (1976), Dohrn and
Guerra (1978), Guerra and Ruggiero (1973), Guerra and Loffredo (1980), Vigier
(1982), Davidson (1979-82), Lee (1980), Yasue (1979), and others. This
problem concerns Part II of the book.
Historically old approaches devoted to the construction of a finite theory of
quantized fields free from ultraviolet divergences exist. These approaches began
from the early stages of the development of field theory and were based on the
analysis of the fundamental principles (causality condition, locality and pro-
perties of geometry on a small scale, etc.) of modern local QFT at small
distances (or, equivalently, at very high energies). The locality condition (the
commutation rules), e.g., is one of the fundamental principles of the local QFT.
More clearly, this means that the commutator of operators of physical fields
disappears outside the light cone. On the other hand, this property of locality
ensures the independence of events separated by spacelike intervals, i.e. the
causality condition in space-time (usually called microcausality). A strict for-
mulation of the microcausality condition in QFT has been given by
Bogolubov and co-workers (see Bogolubov and Shirkov, 1980).
We present here (Part I) a mathematical method of constructing a nonlocal
theory of quantized fields by means of the hypothesis of space-time stochas-
ticity. Our model belongs to a class of approaches in which one assumes a
modification of the usual (local) theory of quantized fields, based on the
analysis of the principle of locality, and postulates that a possible violation of
locality at small distances may be conditioned by intrinsic problems of QFT,
like the ultraviolet divergences, the problem of electron self-energy, etc. This
problem appears especially inevitable as soon as we want to describe spread-
out (extended) objects (particles) within QFT.
Foundation of the Nonlocal Model of Quantized Fields 5

In attempts to reach a consistent construction of QFT with a reasonable


locality (macrocausality, i.e., maybe a slight violation of locality at small
distances) condition, a distinguished role is played by nonlocal QFTs
(Wataghin, 1934; Pais and Uhlenbeck, 1950; Yukawa, 1950; Kirzhnits, 1966;
Efimov, 1977; etc.). Moreover, constructing a nonlocal theory of quantized
fields is stimulated by the latest results from a series of experiments (Aspect et
al., 1982a, b) designed to test a fundamental aspect of quantum theory, which
show that:
(a) quantum theory holds good as far as we can tell,
(b) Bell's inequality is violated,
or, in a more concrete form, one may say that these experiments establish: (a)
the non-existence of a local hidden variable, (b) the existence of quantum
superluminal connections between the action of the independent parts of a
measuring device separated by space-like intervals. The last fact tells us that
quantum phenomena possess a nonlocality property. In other words, there
appears to be some form of 'nonlocal' effect, an effect that Einstein called a
'spooky action at a distance'. These results of Aspect have indicated that
quantum mechanics is adequate even for those situations where nonlocal
effects might seem to be necessary and have recently inspired great interest in
different non local theories and models devoted to an understanding of the
implications of this nonlocality.
There exist two different approaches to the construction of nonlocal theory.
Supporters of the first approach assume that in nature there exists some new
fundamental constant of the dimension of length, together with such constants
as the velocity of light c and the Planck constant h. Further, they assume that, in
a domain characterized by this new universal length I, one must expect a
principle change in our concepts of physical world and, in particular, the
concept of space-time and locality (causality).
Some possibilities of introducing the concept of fundamental length in
physics were considered by Heisenberg (1936); Born (1949); Pais and
Uhlenbeck (1950); Yukawa (1950); Takano (1961, 1967); Markov (1958, 19'59);
Tamm (1965); Kirzhnits (1966); Blokhintzev (1973b); Fubini (1974); Ginzburg
(1975); Efimov (1977a); Ehrlich (1976, 1978); Cheon (1978, 1979); Hsu and Mac
(1979); Lacroix (1979); Kadyshevsky (1980); and Brout et al. (1980), where
earlier references concerning this problem are cited.
The second approach is based on the assumption that the parameter I is not
a universal fundamental constant, but characterizes only the domain of
nonlocal interaction (or the size of extended particles, say, electrons) of the
considered quantum fields. In particular, such an opinion is the basis of
Efimov's (1977a) construction of a nonlocal theory, Thus, the parameter I,
with the dimension of length, arises inevitably in any attempt to introduce
non-locality into the theory. Recently, high-precision measurements (Bailey
et al., 1979; Van Dyck et ai" 1977, 1979; Robiscoe, 1968; Robiscoe and Shyn,
1970) in atomic physics, for example, measurements of the anomalous magnet-
6 Chapter 1

ic moment (AMM) of muons (electrons) and of the Lamb shift, have given
the following restrictions on the parameter 1:1~1O-15cm (l~1O-14cm)
and I ~ 10- 13 cm, respectively (see Chapter 3, and Efimov, 1977a; Hsu and
Mac, 1979). For discussion of various theoretical contributions, where earlier
references can be found, as well as for a review comparing the theory
and experiments, see Brodsky and Drell (1970); Lautrup et al. (1972); Calmet
et al. (1977); Kinoshita (1979), and Levine et al. (1982). From the high-
energy experimental data on e+ e- physics, it follows that I ~ 10- 16 cm (Fliigge,
1980; Wolf, 1980; Beron et al., 1978; Barber et aI., 1979a, b, 1980; Bartel et al.,
1980; Berger et al., 1980, and Ting, 1982). In other words, this means that the
leptons are point-like particles with radii smaller than 10 -16 cm. Recently, experi-
ments have been designed to probe distances up to 10- 17 cm, which allow one
to test some theoretical estimates of the value of 1 ~ 3 x 10- 5 fermi (Bracci
et al., 1983) and predictions about the structure of the leptons with radii I ~ 4 X
10- 17 cm (Li, 1982).
The problem of constructing the nonlocal theory satisfying the basic prin-
ciples of QFT, like the Lorentz covariance, finiteness, unitarity, causality and
gauge invariance, is now solved (Efimov, 1972, 1977a). One assumes in this
theory that neutral particles (for example, photons and neutrinos) are 'carriers'
of non locality, while the charged fields are considered to be local. However,
this way of introducing nonlocality into the theory does not remove all
ultraviolet divergences from the perturbation series for the S-matrix. There are
divergences in the so-called vacuum polarization diagrams constructed by using
propagators of the charged particles. Usually, in order to remove the diver-
gences in these diagrams, one uses the modified Pauli-Villars regularization
(see Bogolubov and Shirkov, 1980; Slavnov, 1974; Efimov, 1972). This method
of regularization of singular functions is to be understood as a formal
procedure only and has no definite physical meaning.
It is generally accepted at present that the essence of mathematical methods
for removing divergences is more or less explicitly connected with the proper-
ties of space-time at small distances, or with the very nature of high-energy
interactions which is inherent to all types of interactions. Thus, we believe that
the method of eliminating ultraviolet divergences must be the same for all types
of diagrams and must have a clear physical meaning.
In this book we give a proper foundation for the concept of nonlocality and
stochasticity which appears in physics and presents a basis for a non local
theory of quantized fields for scalar, electromagnetic and four-fermion weak
interactions (Part I), and for stochastic mechanics and stochastic field theory
(Part II) within the hypothesis of the space-time stochasticity in the micro-
world. We would like to introduce into our consideration the stochastic space-
time [R4 (x) with
x = (xo + ib 4, X + b), (1.1)
where xI' = (xo = ct, x) is the regular part of components x, and
Foundation of the Nonlocal Model of Quantized Fields 7

bE = (b 4 = cr, b) is some random vector with a measure w(biIF) obeying the


conditions

fdW(biI12) = 1, (1.2)

and where 1 is some universal length (a scale of errors). In our case, the
universal length 1 characterizes physically a certain domain within which the
existing space-time concepts and causality conditions may be violated and the
stochastic properties or nonlocality in the metrics can be manifested if they
exist.
It appears that the field obtained by averaging out the space-time ~4(X) is the
nonlocal field considered by Efimov (1966, 1977a) (see Section 1.3).
Equivalence of these approaches leads to the following hypothesis: the origin
of the form factors, which change the electromagnetic (Efimov, 1972) and weak
(Efimov et al., 1973) potentials at small distances, and the properties of vacuum
polarization may be connected with the stochasticity of space-time on the
microscale. Thus, in our model, apart from the universal length 1, there exists a
functional arbitrariness connected with the choice of a measure w(biI1 2 ) or of
a form of weak and electromagnetic potentials at small distances (see Section
2.3). This situation allows us to interpret our approach as a phenomenological
scheme having unknown parameters in the theory. Therefore, our model
belongs to the second-class approaches to constructing nonlocal theories
mentioned above.
Thus, two ideas, the introduction of stochasticity into spacetime in the
microworld and the method of averaging (which allows a transition to a large
scale of space-time) in it, lead to the construction of a QFT with nonlocal
interaction, satisfying the basic principles of the theory. The first part of this
book is devoted to the study of these problems.

1.2. Stochastic Space-Time

The structure of space-time and the physical phenomena within it, enter
inseparably into human cognition, and· their inter-relations are those of
dialectical unity. The simple mechanical motions and the majority of biologi-
cal and chemical processes, including human activity, are confined to the
concept of Euclidean flat three-dimensional space. To order events in this
space to each point-like event r!J, three numbers r = (x,y, z) are attributed, i.e.,
the co-ordinates of this event; the passage from one co-ordinate system to
another is fulfilled by the Galilean transformation:
r = r' + Vt, t = t',
where V is the relative velocity of the two co-ordinate inertial systems.
In the course of the advance of human cognition to deep levels of the
8 Chapter 1

structure of matter, the conception of space-time has been defined more and
more precisely and is still gaining further generalizations and developments.
For example, in the first years of our century, Einstein and Minkowski united
space and time as a single entity and introduced the new concept of a space-time
manifold - the pseudo-Euclidean four-dimensional space, which was a starting
point for the general theory of relativity and modern relativistic QFT. In the
first case, the Newtonian gravitation is due to the curvature of this space-time
manifold. In the Einstein theory of gravitation the ordering (or arithmetiza-
tion) of the events is fulfilled by means of certain agreements (Poincare, 1898;
Einstein, 1950; Fridman, 1965; also see Blokhintzev, 1973b) based on the
principle of the universal constant velocity of light and on an assumption about
the existence of a 'standard' clock (Marzke and Wheeler, 1964; Misner, Thorne
and Wheeler, 1973). This arithmetization leads to Minkowski space with the
indefinite metric
(1.3)

Transformation conserving the metric (1.3) and connecting co-ordinates of


events in different co-ordinate systems is the Poincare-Lorentz transformation:
X' + Vt' z=
t ' + (V /C 2 )X'
Y =y'. Zl,
t = (1- V2 /C2)1/2·

Further, the discovery of the quantum phenomena was initiated by in-


troducing an infinite-dimensional functional space, the Hilbert space, on which
the states of quantum mechanical systems are defined. For the description of
the subnucleon processes, it was necessary to introduce, simultaneously with
the usual space-time, supplementary so-called internal abstract spaces con-
nected with quantum numbers (spin, isospin, strangeness, charge, etc.) of the
elementary particles. Roughly speaking, the 'ordering' of events in these spaces
is reduced to group transformations and to their representations. This abstract
concept of spaces turned out to be very useful and tenacious of life in
classifications of the elementary particles and in the unified description of all
their types of interactions.
The theoretical and experimental successes of high-energy physics now
dictate a deeper level of understanding of the structure of space-time and its
properties at small distances. In the near future, distances up to 10 -1 7 cm will
be probed by high-energy experiments glided at CERN, Fermilab., Serpukhov,
etc. Moreover, as has been mentioned above, the intrinsic problems of QFT
and the hypothesis of fundamental length, and recent results in Einstein-
Podolsky-Rosen-Bohm-type experiments on single photon pairs, establishing
non locality in quantum mechanics, give a strong stimulus to the investigation
of the space-time structure of matter.
The concepts of supers pace and pregeometry (Wheeler, 1964; Akama, 1981;
Terazawa, 1981), the submicroscopic scale and lattice (discrete) structure of
space-time (Wilson, 1974; Kogut and Sussking, 1975; Vialtsev, 1965) and
Foundation of the Nonlocal Model of Quantized Fields 9

stochastic (or fluctuational) and higher-dimensional geometry (Snyder, 1947)


may become the arena of future physical theories. Friedberg and Lee (1983)
and Cole (1972), for example, discussed discrete quantum mechanics and a
cellular space-time structure, respectively. Finkelstein (1969, 1972, 1974) pub-
lished a series of papers developing a theory including space-time structure -
the so-called space-time code. As it has been supposed by Wheeler, the
electrical and nuclear charges give evidence for the presence of a submicroscale
structure of space-time resembling afoam-like structure which is on the whole
homogeneous.
It is assumed that quantum fluctuations in geometry are inescapable, if one
believes in the quantum principle and Einstein's theory. According to Misner,
Thorne and Wheeler (1973, pp. 1193, 1201) 'The fluctuations widen the narrow
swath cut through super-space by the classical history of the geometry. In
other words, the geometry is not deterministic, even though it looks so at the
everyday scale of observation.... The quantum fluctuations in the geometry of
space are so great at small distances that even the topology fluctuates, makes
'wormholes', and traps lines of force.' Thus, it seems, the fluctuations in the
geometry make a multiple connected topology which provides a natural
description for the electric charge as electric lines of force trapped in the
topology of a multiple connected space.
Thus, we believe that among the above-mentioned hypothesis of space-time
properties, the idea of the stochastic character of space-time at small distance
will play the most important role. In particular, as we shall see below, this idea
leads to the concept of nonlocality and thus it gives the very nature of nonlocal
QFT.
Some attempts were undertaken to construct quantum field theory in a
stochastic space-time (Markov, 1959; Komar and Markov, 1959; Takano,
1961, 1967; Ingraham, 1967; Blokhintsev, 1973a, 1975 and references therein).
A stochastic space-time, which can be used in constructing theories of elemen-
tary particles, was first considered by March (1934, 1937), Markov (1958) and
Yukawa (1966). Recently, this problem has been discussed by Blokhintsev
(1975), Roy Choudhury and Roy (1980), Cerofolini (1980), Prugovecki (1984).
Frederick (1976) and Roy (1979) considered mathematical spaces with stochas-
tic metric and quantized domain, respectively. A mathematical definition of
linear space with random metric, mainly in the Euclidean case, was given by
Menger (1942, 1949); Sherstnev (1963) and Schweizer (1967). Prugovecki's
(1978) papers are devoted to the construction of the relativistic kinematics of
massive and massless particles in a stochastic phase space.
In contradistinction to Markov, Takano, Frederick, Menger, Sherstnev,
Schweizer and others, we do not start here from metric space with stochastic
metric, but we consider the behavior of dynamic systems in space-time with a
measure (or distribution), say in the space-time 1R4 (x) with stochastic com-
ponents (1.1) and measure (1.2). This view is also the basis of Prugovecki's
statement: 'If we, however, accept the uncertainty principle and yet retain
10 Chapter 1

Einstein's operational definition for the structure of space-time, then we are


forced into describing space-time in stochastic terms ... ' (Brooke, Guz and
Prugovecki, 1981, p. 1731). This proposed way has also another justification:
stochasticity in physics appears more naturally on the stage of arithmetization
(or all reality is subject to an intrinsic stochasticity inherent in the measure-
ment process caused by stochasticity of space-time) rather than on the metrical
stage. The stochastic metric is introduced on a more deeper level, where one
should take into account the gravitational effects connected with introducing
stochasticity of space-time into physics. The study of latter problems is,
however, beyond the scope of this book, although some initial idea in this
direction will be discussed in Section 11.7 of the Part II.
Of course, stochasticity of space-time and random metric are connected with
each other as a single entity concept; for instance, any space, the points of
which have the measure, reduces to the space with random (or stochastic)
metric. Indeed, in our case of the space-time \R 4 (x), the metric averaged is given
by

s; = <S2>~4(x) = fd4bEW(bUF)[(XO + ib 4 )2 - (x + b)2] = x6 - x 2 - F. (1.4)

On the other hand, this result can be rewritten in the form


(1.5)
where a is a random number with a measure w(a/I) such that

fda w(a/I) = 1, fda aw(a/I) = 0,

We see that modification of the metric form (1.5) may be understood geometri-
cally as the introduction of the fifth dimension which is conditioned by the
probability measure. A metric form of the type of (1.4) (we call it nonlocal
metric form) will be considered also (Section 11.6).
An attempt to construct a quantum field theory, based on the metric form
(1.4) [or (1.5)] was done by Markov and Takano.
Now we concretize the form of the space-time \R 4 (x) and explain the
following question: why the points of space \R 4 (x) have to be of the form (1.1)
and the reason for it. The answer to this question will inevitably depend on the
dimension of space and the operation of the arithmetization of events. In the
nonrelativistic case (space dimension is three), it is sufficient to assume that
space is stochastic with respect to the spatial component x -4 X = X + b, where
b is a universal stochastic variable, subject to a probability measure with
w(b) = w( 1b 12), for example,
web) = (2nI2)-3/2 exp( -lbl 2 /21 2 ). (1.6)
This form of the measure follows from the homogeneity and isotropy of space.
Foundation of the Nonlocal Model of Quantized Fields 11

In the relativistic case, such an operation requires some explanation. In the


relativistic theory the space in which physical processes are investigated is the
Minkowski space, i.e., the space 1R4 (x) must be the Minkowski space. The
indefinite metric of this space leads to a number of specific problems, which are
not encountered in Euclidean space.
These difficulties peculiar to physical space are associated with the require-
ments of invariance and the condition of normalization of the probability of
some value (in particular, the measure w(b,,) itself) of the interval b 2 = b6 - b 2
in a space with indefinite metric (for more detail, see Blokhintsev, 1973b).
Thus, the invariance requirement means that the measure w(b,,) of the vector
b" must be a function of the interval b 2 = b"bl', and the normalization
condition gives the equation

The simultaneous fulfillment of these two requirements for w(bl') is actually


impossible in Minkowski space. The above difficulties can be avoided by
making the following assumptions:
1. The stochastic nature of the space 1R4(x) is manifested in the Euclidean
x
domain of the variables E = X E + bE'
2. Physical quantities are regarded as functions of the complex time t + ir in
the limit of microscale, where the stochasticity of space-time may be manifes-
ted [r is a stochastic variable which ensures the hypothesis of a stochastic
nature of the Euclidean space /E 4 (x, r) instead of the Minkowski space
[R4 (x, t)]. A justification of the importance of the method with t ---> t + i:r in
quantum field theory and quantum mechanics can be found in the papers by
Alebastrov and Efimov (1974) and Davidson (1978), respectively (see also
lannussis and Papatheou, 1982). Thus, in our model the real points of the
space-time [R4(X) consist of two parts (1.1), where b4 = cr, and any physical
quantity F(·) in [R4(X) depends on arguments of the type Xo + ib 4 , X + b with
the measure w(bUl 2 ), i.e., F(') = F(xo + ib 4 , X + b).

1.3. The Method of Averaging in the Stochastic Space-Time


and Nonlocality

Since, in accordance with our assumption, the real points of the space [R4(X) are
stochastic, these points cannot be used as the basis for a coordinate system,
and for the same reason derivatives with respect to them cannot be for-
mulated. However, the space of common experience (i.e., the laboratory frame)
is nonstochastic on a large scale. The stochastic nature of space is manifested
only in the microscopic world. Therefore, one can take a macroscale of
nonstochastic space mathematically continued from a stochastic microregion.
12 Chapter 1

Such a mathematical construction ensures a nonstochastic space to which the


stochastic physical space can be referred (see also Frederick, 1976). In our case,
this mathematical construction reduces to averaging with respect to the mea-
x
sure w(b~W) at each point = (xo + icc, x + b) of the space [R4(X). Thus, by
the physical quantity F(x, t), we understand <F(X)1R4(x)

F(x, t) = <F(x) )1R4(x) = f d 4 bE w(b~ 112 )F(xo + ib 4, x + b), (1.7)

averaged with respect to the measure w(b~/12) at a given time t.


It should be noted that with a suitable choice of the function w(b~/12) the
method of averaging (1.7), which ensures the transition from the small to the
large scale, leads to an entirely new object. In particular, a scalar field ({J (x)
after averaging in [R4(X) acquires the following form:

({JR(X) = <({J(X)R = fd4bEW(b~W)({J(Xo+ib4,X+b)


= (2n)-2 fd4peiPx K(p212)qJ(p), (1.8)

where

K(p 2 l2) = fd 4 bE exp(-ipb-p o b4 )w(bUF)

= (2n)2 f' dr r2 w(r2 IF)J l(r~=?)( _ p2) -1/2. (1.9)

Here p2 = P6 - p2, and J 1(x) is the Besselfunction. Further we are interested


only in the class of measures w(bUI 2) for which, K(z) in (1.9) are entire
analytical functions of the variable z with a finite order of growth 00 > p >t
and which decrease rapidly enough when z = p2 --+ - 00 (i.e., in the Euclidean
direction). Thus, expression (1.8) may be rewritten in the other representation
by means of a generalized function (or distribution) (see Section 1.4)

((JR(X) = fd4YK(x-y)({J(y) = fd4peiPx K(P212)qJ(P), (LlO)

where K(x) is some generalized function, the Fourier transform of which is


given by formula (1.9). The following representation is valid for K(x):
02 02
D = - -;:;--z + -;:;--z (Lll)
uXo uX

and the coefficients {c n } depend on the particular choice of the measure


w(b~/12). In quantum field theory, such an object, say ({JR(X) has been
investigated carefully by Efimov (1968, 1977a) from the point of view of
generalized functions K(x) whose space-time properties depend essentially on
Foundation of the Nonlocal Model of Quantized Fields 13

different sequences of the coefficients {c n }. Efimov showed that an object


constructed by means of such generalized functions K(x) is smeared (nonlocal)
in space-time. Thus, the relativistic invariant generalized functions K(x) are a
basis of a correct description of extended objects (see Chapter 2). In this case,
roughly speaking, the parameter I may be identified with the size of an
extended object (a particle).
Thus, within the framework of our scheme all fields: neutral and charged are
extended (nonlocalized) over the space 1R4(X). This allows one to take into
account in a unified way the effect of stochasticity (or nonlocality) in all
physical processes. At this point, even the space-time metric g/1V (x) is changed
and becomes nonloeal:

g/1V(x) ~ G/1Jx) = f d 4 yK(x- y)g/1v(y). (1.12)

Some consequences of this interesting possibility will be discussed below in


Chapter 1I.
Now let us calculate the causal Green functions of the extended field CPR (x )
(bearing in mind that the T-ordering symbol concerns the field cp(x), where
cP (x) is a scalar field with a mass m) by the following formula:

Dc(x l -Xz) = <01 T[CPR(xdcpR(X z )] 10)

= ffd4bIEd4bzEW(biE/IZ)W(b~E/ZZ) x
x <OIT{cp(x lo + ib l4 ,X I + bdcp(xzo + ib z4 ,x Z + bz)}IO)

= ffd 4 b lE d4bzEw(biE/[2)w(b~E/IZ)i-I(2n)-4 x
x f d4p (m Z - pZ - iB)-1 X

x exp[ipo(x IO - xzo + ib l4 - ib z4 ) - ip(XI - Xz + bl - bz )]

= i- I (2n)-4 fd4peiP(Xl-X2)[K(pZIZ)]2(m2 - p2 - iB)-I, (1.13)

where K(p 2 12) is given by formula (1.9).


Summarizing this section we see that according to the hypothesis of the
stochasticity of space-time we come to the Efimov nonlocal theory differing
only in that the causal Green function S(p) of any charged particle is replaced
by (in the momentum representation):
S(p) ~ SR(P) = V( - p212)S(p),
where
14 Chapter 1

for which the Mellin representation (see Section 2.4.4)

1
V(_p2[2) =----:-
21
f- P - iOO

-P+ioo
v(O
d(-.-F;(m
smn(
2 - p2 - ill);
(O<P<l)

is valid. The form of the functions V(_p212) and v(O depends on the form of
the measure w(bi/1 2). For example (Namsrai, 1981c), let

n-21-4e-mln)-1/2 m212(1_y2)-1/4 J (ml(1_y2)1/2) o..;;;;;y<


wm(y) = { 2 4(sinml/ml- cosml) -1/2 ,
o y>
Then we obtain

(1.14)

(1.15)
Here m is some parameter (m212~1) which can be identified with the particle
mass. Notice that the form factor (1.14) in the case m = 0 describes the
extended electron as a uniformly charged ball of radius I (see Section 2.3 and
Efimov, 1977a). Now let us write the Gaussian measure:
o";;;;;y<oo.
The function VG ( - p2 F) corresponding to this measure acquires the following
form
(1.16)
The main restrictions in the choice of form factors V ( - p212) as entire
analytic functions arise from the fundamental theoretical principle of the
theory, i.e., from unitarity and causality (see Chapter 2 and Alebastrov and
Efimov, 1973, 1974).
The physical meaning of form factors V( - p2[2) consist in changing a form
of the potential between interacting fields (for example, the Coulomb and
Yukawa laws) at small distances, and in making the theory finite In
each order of the perturbation series of the theory in coupling constant.

1.4. The Class of Test Functions and Generalized Functions

In this section we give some elementary definitions of test functions and


Foundation of the N on local Model of Quantized Fields 15

generalized functions (or distributions), and explain their main differences in


the cases of local and nonlocal quantum field theories. Our exposition carried
out below is based on the works due to Bogolubov, Logunov and Todorov
(1975), Vladimirov (1979) and Efimov (1977a).

1.4.1. Introduction
The concept of generalized functions is the generalized classical definition of
the function. This generalization arose from physical requirements and enabled
some idealized concepts to be defined as a density of the material point, and of
the point-like charge or dipole etc. On the other hand, the concept of
generalized functions is based on the fact that it is practically impossible to
measure the value of a physical quantity at a given point, and that one can
only measure its averaged value in the relatively small surroundings of this
point. As a result, it is usually assumed that the limit of sequences of these
averaged values may be understood as a value of the physical quantity under
consideration at a given point.
For example, in order to explain the above-mentioned situation we attempt
to define the density generated by the material point with the mass m = 1.
Assuming that this point is the origin of the co-ordinate system, we distribute
this mass uniformly inside the sphere with the radius of e and the centre at the
point 0. As a result we obtain the averaged density Pe(x)

( ) = {(4ne 3 /3)-1, if Ixl < e,


PeX 0, iflxl>e.
However we are interested in the density at e = +0. Further, we denote the
following function

i5(x) = {+
0,
00, if x = 0,
if x # 0,
(1.17)

as a pointwise limit of sequence of the averaged densities Pe(x) at e ~ +0. It


should be required that

fi5(x) dx = 1.

However, for the function i5(x) defined by the formula (1.17)

fi5(x) dx = 0.

This means that this function is not to generate mass and therefore it cannot
be understood as density. Thus, a pointwise limit of sequence of the averaged
densities does not fulfil our purpose. Now we define another limit, the so-
called weak limit, of the sequence of the averaged densities Pe (x). It is easily
16 Chapter 1

verified that for any continuous function f(x)

lim fPc (x)f(x) dx = f(O). (1.18)


C;---l> +0

This formula denotes that the weak limit of the sequence of the functions Pc (x)
for e -> +0 is the functionalf(O) (but not function!) giving to each continuous
function f(x) the number f(O) - its value at the point x = o. We define this
functional as the density b(x) - the famous Dirac b-function. Thus, we can write
Pc(x)=b(x), for e -> +0.
weak

This limit is given by formula (1.18) for any f(x). The value of the functional b
onto the function f(x) is the number f(O) which will be denoted as
(b,j) = f(O). (1.19)
We see that functional b generates total mass. Indeed,

f dx b(x) = (b, 1) = 1

at the point x = 0 [the second equality follows from (1.19)].


Thus, the density corresponding to the point-like distribution of masses
cannot be described within the classical concept of the function, and for its
description objects of a more general mathematical nature are needed - linear
(continuous) functionals. Before defining the functionals we consider the spaces
of test functions.

1.4.2. Space of Test Functions


From the example of b-function we see that this function is defined by means
of continuous functions as a linear (continuous) functional on these functions.
It is said that continuous functions are the test functions for b-function. This
point of view takes as its basis the definition of any generalized function as a
linear continuous functional onto sets of sufficiently well-defined so-called test
functions. It is clear that with the smaller ,set of test functions there exists
larger linear continuous functionals. On the other hand, the reserve of the test
functions must be sufficiently larger.
First, some definitions are needed. Let I5IJ be the linear space in which the
commutative and associative addition is defined as:
1. u + v = v + u.
2. u + (v + w) = (u + v) + w
and also
u+O=u, u+(-u)=O
Foundation of the Nonlocal Model of Quantized Fields 17

for each u, v, WE I5IJ and 0 E 151J. Moreover,

l·u = u,
and
(A + Il)u = AU + Ilu, A(u + v) = AU + AV
for any u, v E I5IJ and some (real and complex) numbers A and 11.
The real function p(u) defined on I5IJ is called the norm if the following
conditions are fulfilled:
(i) for any number }" p(Au) = 1},lp(u),
(ii) p(u + v) ~ p(u) + p(v) (triangle inequality),
(iii) if p(u) = 0, then u = O.
From the conditions (i) and (ii) follows non-negativity of the norm

0= p(u - u) ~ p(u) + p( -u) = 2p(u).

The functions satisfying only conditions (i) and (ii) are called the semi-norms.
If the norm p is given in 151J, then one can define the distance between any
two elements u and v E I5IJ as p(u - v). We say that the sequence (u l , ••• , Un" .. )
converges to the limit u if the distance between Un and u tends to zero when
n ----> 00, i.e., if

lim p(u n - u) ----> O.

Convergence defined by such manner is sometimes called the convergence over


the norm or the strong convergence.
Linear space I5IJ in which the convergence is given by the norm p(u) is called
the normalized space. As usual, the norm p(u) of the element U denotes as
p(u) = Iluli. For example, the Hilbert space L2 [a, bJ of all complex functions
f(x) with the norm

(f,f) = IIfl12 = r dxt(x)f(x)

is linear normalized space. Other linear normalized space is the space C ([a, bJ)
of continuous functions on the interval [a, bJ with the norm

p(u) = Sup lu(x)l.


xE[a.bj

As a further example, we consider the space C(O', p, n) of complex-valued


functions of n-real variables x = Xl" .. ,xn , having continuous partial de-
rivatives up to the order 0' inclusively, and decreasing no more slowly than
Ix 1- P together with all derivatives at infinity. In other words, for the functions
18 Chapter 1

u(x) from C(a, p, n) all the products of the type


x"DPu(x), 1(XI~p, IfJl~a (1.20)

are bounded, where

I(XI = (Xl + ... +(Xn' (1.21)


The norm in the space C(a, p, n) is given by

pp,,(u) = max Suplx"DPu(x)l. (1.22)


lal';;p XE~n

IPI';;"

The spaces of the type of (;(a, p, n) play an important role in the theory of the
generalized functions.
DEFINITION 1.1. Let
Pl (u) ~ pz(u) ~ ... ~ p,,(u) ~ ...

be increasing the sequence of norms co-ordinated in pairs in the linear space


OU. Space OU is called countably normalized (with the norms {p,,(u)}), if the
convergence in it is defined as follows. The sequence {u v } E OU converges to the
element UE OU, if lim,,~oo p,,(u v - u) = 0, for any a.
We give some examples of countably normalized spaces:
(1) Space D(G). Let G be the finite region, i.e., the bounded open set in the n-
dimensional real space W. Denote D(G), the set of infinitely smooth functions
(i.e., functions having continuous partial derivatives of all orders) in IRn,
tending to zero outside the region G. Define in D(G) the countable system
norm q" by the formula

qo(u) = Poo(u) = Suplu(x)I, ... , q,,(u) = Po,,(u) = max SupID"u(x)l,


x Icxl<'a x

where symbol D" is given by formula (1.21).


The particular case of space D(G) is space D(a) of infinitely smooth functions
of the single real variable x, tending to zero outside the interval - a < x < a.
For example, the function u(x) defined by the equality

u (X)={exp[aZ/(xZ-a Z)] for -a < x < a (a> 0)


a 0 for Ixl ~ a
belongs to space D(a).
(2) Space S[or S(lRn)] consists of all infinitely smooth functions in the n-
dimensional real space IRn, which decrease rapidly any polynomials of
(xi + x~ + '" + X;)-l/Z together with all partial derivatives at Ilxll---> 00, i.e.,
Foundation of the Nonlocal Model of Quantized Fields 19

In other words, for these functions all the norms (1.22) take finite values. We
will define the convergence in S by the countable system norms

p"(u)=p,,,,(u)=maxSuplxaDPu(x)l, (J = 0, 1,2, ... (1.23)


lal <;" XE IR"
IPI <;"

In particular, Hermit-Chebyshev functions, and in general, all functions of the


type
P(X1> ... ,xn)exp(-xi!ai - ... -x;/a;)
may be used as functions of S-space, where P(.) is an arbitrary polynomial.
(3) Space S(G) is defined as a set of functions from S, which tend to zero
outside (the generally considered unbounded) region G. The convergence in
S(G) is given by the same system norm (1.23). We assume a possibility G = jRn,
so that the space S is a particular case of space S(G). Space D(G) is also a
specific case of spaces of the type S(G) for the bounded regions.

1.4.3. Linear Functionals and Generalized Functions

1.4.3a. General Definition. Numerical function defined in linear space IJlt is


called functional. The functional F(u) is called linear, if for any u, VE IJlt and any
numbers IX and f3
F(lXu + f3v) = IXF(u) + PF(v). (1.24)
For example, the integral over the space of integrable functions on the some
interval is the linear functional. However, the norm is nonlinear functional.
The functional F(u) is continuous if from the convergence of sequence {un} E IJlt
to u E IJlt follows the convergence of sequence F(u n) to F(u). In particular, if the
convergence in the space IJlt is given by the norm p(u) (i.e., if IJlt is the
normalized space), then it is said that the linear functional F(u) is continuous
(or bounded) in 1Jlt. If there exists a positive number CF such as that for any
u E 1Jlt, the following inequality

(1.25)
holds. The set of all linear functionals in normalized space IJlt is in turn linear
space, and we call it the space conjugate to IJlt and denote it through 1Jlt'. In 1Jlt'
one can also define the norm by the formula

p'(F) = Sup IF(u)l. (1.26)


p(u) <; 1

In other words, p' (F) is the smallest of constants CF for which the inequality
(1.25) is valid.
The space 1Jlt' conjugate to the normalized space 1Jlt, with the convergence
20 Chapter 1

defined by the norm (1.26) is called full normalized space irrespective of


whether the initial space IfIJ, is full or not.
Now we turn to the study of linear functionals in the countably normalized
space topology which is given by system neighborhoods: the neighborhood of
any element u E IfIJ, being
U",(u) = {v, p,,(v - u) < e},
where e and (J are positive and natural numbers, respectively. This formula
means that the element v belongs to the neighborhood U",(u), if p,,(v - u) < e.
The linear functional F(u) is to be continuous in 1fIJ" if for any e>O one can
show the neighborhood U"b(O) of zero in which JF(u)J < e. In this neigh-
borhood p,,(u) < (j for any element u E IfIJ, •
We have already seen that the space conjugate to the normalized space is
itself full normalized space. However, the space 1fIJ,' conjugate to the count ably
normalized space IfIJ, has a structure different from the structure of the space IfIJ, •
In this case construction of the space 1fIJ,' can be described as follows:
If the functional F(u) is bounded in the neighborhood of zero p,,(u) < (j, then
this functional is continuous with respect to the norm p,,(u), i.e., the inequality
of the type of (1.25) holds
JF(u)J « Cp,,(u). (1.27)
The smallest (J for which this inequality takes place, is called the order of the
functional F(u). We see that any functional from 1fIJ,' has a finite order.
Now we take some examples for functionals:
(1) Space D'(a) conjugate to countably normalized space D(a) consists of the
functionals of the type

F(u) = fa u(,,) (x) dl1(x),

where (J = 0, 1, 2, ... and l1(x) is a function with bounded variation (for


example, see Gel'fand and Shilov, 1968).
(2) Space S' consists of all functionals of the type

F(u)= r dnxf(x)Dau(x),
J~n
where D a is given by formula (1.21), and f(x) is a continuous function of the
polynomial growth.

DEFINITION 1.2. A generalized function is called any linear continuous


functional over the countably normalized space S defined above, i.e., any
element of space S'. The functions of space S are called test functions.
It is clear that the concept of the generalized functions depends on the
choice of the initial (linear topological) space of the test functions. For
Foundation of the Nonlocal Model of Quantized Fields 21

example, instead of S we would take D(G) as a base of test space. L. Schwartz


(1957, 1959) defined the generalized functions as the continuous functionals on
space D of all finite and infinite differentiable functions (D is the union of
the spaces D(G), when the region G is changed). Functions disappearing out-
side some finite region of space are called finite functions. Closure of points set
on which a continuous function u(x) =F 0 is called the support of this function.
We retain the term 'generalized function' for the functionals from S', while
the functionals from D' we call the distributions due to L. Schwartz (1957,
1959), who calls the generalized functions from S' 'distributions temperees' and
Gel'fand and Shilov (1964, 1968) call these generalized functions of 'tempered
growth'. It is verifiable that any generalized function is distribution while the
opposite assertion is not, i.e., the inclution S' c D' takes place.
Another definition of the generalized functions as classes of fundamental
sequences is given by Bogolubov et al. (1975).
Now we give some basic properties of the generalized functions.

1.4.3b. Transformation of the Arguments and Differentiation of the Generalized


Functions. Mutually synonymous transformation Y = cp(x) [Y1 = CP1(X)"",
Yn = CPn(x), x = (Xl, . .. , Xn)] or X = cP -1 (y) of the space IRn onto itself leads to

(f(cp-1(y)), u(y)) = L" dnxf*(x)u(cp(X)) IJ(cp) I = (f(x),IJ(cp)lu(cp(x))) (1.28)

for any u(x) ED and f is a usual (local integrable) function. Here

OCP1 OCP1
aX 1 aXn
J(cp) = =F o.
acpn
-
acpn
aX 1 aXn

Whenf(x) is the distribution, the right-hand side of (1.28) is understood as the


definition of the distributionf(cp-1(y)). By means of the particular case Y1 =
Xl - ~X1' Y2 = X2"'" Yn = Xn, it is easy to verify that

(1.29)
in the limit ~x 1 -+ O. Here we use the fact that in accordance with the
continuity of the functionalfthe right-hand side of (1.28) in this particular case
tends to the limit -(f, au/ax 1) at ~X1 -+ O. By definition we call this limit the
partial derivative of the generalized function f(x) with respect to the variable
Xl' We notice that the operation of the differentiation is continuous in S', i.e.,
the following equality takes place

lim (aIn/ax, u) = (of/ax, u)


22 Chapter 1

for any u E S and limn~ 00 (f", u) = (f, u). These simpler properties of the
differentiation operation in Sand S' are not fulfilled in the case of normalized
spaces. Therefore, the countably-normalized spaces play an important role in
the theory of the generalized functions.
Further, we notice that
(1.30)

i.e., for the generalized functions the order of the differentiation has no
significance.
We consider some examples for the generalized functions of the single
variable, which may be defined as derivative of usual local integrable functions:
(1) The derivative of the well-known discontinuous function

8(x) = {1o for x >


for x < 0
0, (1.31)

equals b(x). Indeed, in accordance with definition (1.29) and formula (1.19) we
have

(~~, u) = - f:oo dx8(x) ~~ = - Loo dx ~~ = u(O) = (b, u). (1.32)

(2) The functional (d/dx) In Ixl coincides with the principal value of l/x in
Cauchy's sense; in other words,

(d 1~~XI, U(X)) = f?f f:oo dx u~x), (1.33)

where f?f is the symbol of the principal value of the integral. Indeed, by
definition (1.29), we obtain

(d )
-d Inlxl,u(x)
x
= - foo
-00
dxlnlxl·u'(x)= lim
e~+O
{f-
-00
e
u(x)
dx-+
X
1
E
00
U(X)}
dx-
X

= foo dx u(x) - u( - x) = f?f fOO dx u(x) , (u' (x) == du/dx).


Jo x - 00 x
Defined thus, the generalized function dIn Ixl/dx is denoted as 1/x [or
sometimes as f?f(1/x)]. We notice that the usual function 1/x is not locally
integrable (in the neighborhood of the point x = 0) and it, therefore, is not
identical with the generalized function 1/x. However, these two functions
coincide at x # o.
Analogously, we define the function 1/x 2 as a derivative of the generalized
function -1/x:

(:2' U(X)) = (~, U'(X)) = LOO dx u'(x) -xu' ( -x)


Foundation of the N on local Model of Quantized Fields 23

= 100

o
d u(x)
x
+ u( -
X
x) - 2u(0)
2 • (1.34)

In general, we assume by the definition

_" (_1)"-1 d"


x = (n _ I)! dx" In lxi, n = 1,2, .... (1.35)

where the derivative must be understood in the sense of the definition of the
differentiation in S'.
(3) The usual locally integrable function In (x + iO) defines as

In(x + iO) = lim In(x + iy) = lnlxl +i lim arg(x + iy)


y~ +0 y~+O

= In Ixl + in8( -x), (1.36)

where 8(x) is given by formula (1.31). The derivative of the generalized function
(1.36) is denoted as 1/(x + iO). In accordance with (1.32), (1.35) and (1.36) we
have
1 d 1
--.- == -In(x + iO) = - - in b(x), (1.37)
x + 10 dx x

or in the general case

( '0)_"_(-1)"-1 ~l ( '0)- -" . (_1)" ~"-1()


x +1 - (n _ I)! dx" n x +1 - X + l1t (n _ 1)! u x. (1.38)

and
(_1)"-1 d" (_1)"-1
(x - iO)-" = (n _ 1)! dx" In(x - iO) = x-" + in (n _ I)! b"-l (x). (1.39)

According to formula (1.29) derivatives of the delta-function on the right-hand


sides of equations (1.38) and (1.39) are given by the expression

d"
(b", u) = ( -1)"u(")(O), b" = -d [b(x)].
x"

1.4.3c. The Fourier Transform of Generalized Functions. We consider


space IR" with the indefinite metric, i.e., the interval between two 'events' x and
y as given by
S2 = gijxiYj = X 1 Y1 + ... + X1YI - Xl+ 1 Yl+1-'" -x"y" (1.40)
and construct the Fourier transform with respect to this bilinear form.
24 Chapter 1

LEMMA 1.1. The Fourier transform of the test function u(x) E S is

ii(P) = Fu(x) = (2n)-n/2 r


J~11
dnx eipx u(x) (1.41)

and is also the test function (ii(p)E S). Here the bilinear form xp is given as (1.40).
We see that formula (1.41) gives the mutually synonymous and continuous
mapping of space S onto itself.

The inverse transform of (1.41) is given by

u(x) = Fii(P) = (2n)-n/2 r dnp e- ipx ii(P);


J~n
(1.42)

which possesses the same properties as F. The Fourier transform therefore


realizes an isomorphism S on S.
We define the Fourier transform of the generalized function f(x) as the
linear functional J(p) on the space of the Fourier images of the test functions
u(x) by the formula
(](p), ii(P)) = (f(x), u(x)), (1.43)
where ii(P) is connected with u(x) by formula (1.41). Formula (1.43) may also be
considered as a definition of inverse Fourier transform, if the generalized
function J(P) is given then its inverse image f(x) determines its equality (1.43).
According to Lemma 1.1 the Fourier transform J(p) of the generalized
function f(x) (f(x) E S') is also the generalized function of the same space S'.
Conversely, any generalized function J(p) E S' has the inverse Fourier image
f(x) E S'. Thus, the following theorem corresponding to the Lemma 1.1 takes
place.
THEOREM 1.1. The Fourier transform of the generalized functions defined by
formula (1.43) is the isomorphism of space S' onto itself. We recall that the
isomorphism of the linear topological space dlJ 1 onto the same space dlJ 2 is called
the mutually synonymous and mutually continuous mapping dlJ 1 onto dlJ 2,
conserving linear operations.
We notice that Lemma 1.1 and Theorem 1.1 do not hold for the spaces D
and D' of finite test functions and distributions.

(
For example, we give Fourier transforms of some generalized functions.

Fe(t) = (2n) -1/2 f dte(t) eiwt =


2
n
W+ 10
)-1/2'
. I,

F(j(n)(x) = (2n)-n/2 f~n


dnxb(n)(x)e iPX = (2n)-n/2,
Foundation of the Nonlocal Model of Quantized Fields 25

1.4.3d. Multiplication of the Generalized Functions by a Smooth Function and


Their Convolution. The product of the generalized functions represents one of
the most important problems in the theory of generalized functions. Since this
is a nonlinear operation,. the product of the two pairs of the generalized
functions cannot be defined in the usual manner. It is easy to verify that it is
impossible to give multiplication which would be associative. Indeed,

1
x x
1 x
(1 )
-(x b(x)) = -0 = 0,;. - x b(x) = b(x). (1.44)

Nevertheless, there exists a wide class of functions for which one can define
their product with the generalized functions from S' by the natural manner.
This class is defined as follows.
It is said that the function cp(x) is multiplier in space S of test functions if
from u(x) E S it follows that cp(x)u(x) E S. The space of all multipliers we denote
as QM. It is clear that if cp(x) is infinitely differentiable and a polynomially
bounded function of x (together with all its derivatives), then cp(x) is the
multiplier in S.
If cp(x) is the multiplier the product of cp(x) by the generalized functionfE S'
is given by the formula
(cp(x)f, u(x)) = (f, cp*(x)u(x)) (1.45)
for any u(x) E S. By definition we assume that the product is commutative, i.e.,
cpf = h·
In some cases one can define the product of two generalized functions.
However, construction of the general theory in this direction is not complete
and encounters definite difficulties. The production problem of generalized
functions is crucial in quantum field theory. Since the latter always deals with
the generalized functions, definition of those products requires new mathemati-
cal methods and specific regularization procedures. This problem will be
partially discussed in Section 1.4.4 (see also Sections 2.2.2 and 2.4.4).
We now consider the operation of convolution. This operation is widely used
in the theory of the generalized functions. The convolution for two usual
functions f(p) and g(p) is given by the expression

f(P) * g(P) = I dnqf(p - q)g(q) = I dnqf(q)g(P - q). (1.46)


J~n J~n
As in the case of the multiplication operation the convolution is not defined for
any pair of the generalized functions. Using the second equality in (1.46) the
convolution of the generalized function f with the test function u(p) E S is
defined as:

f * u(P) = (f(q), u(p - q)) = I dnqf(q)u(P - q). (1.47)


J~n
26 Chapter 1

The first equality in (1.47) is the definition of the convolution as the operation
of the functionalf (q) onto the test function u(p - q) considering the latter as a
function of q at fixed p. Clearly the convolution (1.47) is infinitely differentiable
and a polynomially bounded function of p (together with all its derivatives). In
other words, iff E S' and u E S then f * u(P) E QM' i.e., the convolution (1.47) is
the multiplier. Generally speaking, the function f * u does not belong to the
space S of test functions.
Now we go over to the problem of division and to the concept of the
support of the generalized function. First, we consider the division problem
which is inverse in operation to multiplication and leads to the study of the
equation
cp(x)f = g, (1.48)
where 9 E S' and cp(x) E QM are the given functions, and f is an unknown
generalized function. In the case when cp(x) "# 0 for any x, and it goes to zero
not too rapidly (i.e., the function cp -1 (x) is also a multiplier), equation (1.48) is
solved elementarily. The problem of division becomes complicated if the
function cp(x) has zeros. Consider here only the case of a single independent
variable x and assume that cp(x) in (1.48) has a discrete set of zeros of a finite
order. By this assumption the division problem in the space S'(~1) is in
essence reduced to the solution of simpler equations of the type of (1.48):
xf=g· (1.49)
Solution of which has the form
(f, u) = Cu(O) + (g, U 1 (x»,
where u(x) and U1(x) belongs to S(~1). While the general solution of the
homogeneous equation
xfo =0
is
(fo, u) = Cu(O), i.e., fo = C(j(x) (1.50)
Further, by induction it is easily shown that the more general equation
(1.51)
always has a solution with respect to fE S', where 9 E S' and m is any natural
number. An arbitrary factor in this solution is produced by a general solution
of the homogeneous equation
xmfo = 0,
that is

(1.52)
Foundation of the Nonlocal Model of Quantized Fields 27

where C v are arbitrary constants (for details, see Bogolubov et al., 1975).
Further we discuss other questions of interest including the local properties
of the generalized functions. As opposed to the usual functions which are given
in each point of some set, the generalized functions are determined whole as
values of the functional on the space of the test functions. Generally speaking,
they do not have definite values at separate points. It is said that the
generalized functions f and g coincide in the region (or in the open set) G if for
any test function u(x) with the support on G the following equality (f, u) =
(g, u) takes place. Thus, by definition the set of points on which the generalized
functionfturns to zero, is open. Complement of this set to the whole space ~n
is called the support of the generalized function f.
THEOREM 1.2. Let f be the generalized function located in the origin of co-
ordinate system (i.e.J(x) = 0Jor x =1= 0). Then f(x) is expressed by afinite linear
combination of /j-function and its derivatives /jV(x). This assertion is valid for both
the cases of the generalized functions of many variables and of the distributions
from D'( ~n), i.e., if f E D'( ~n);

f(x) = L CaDa/j(x), (1.53)


lal"N

where N is the order of f and C~ are some constants.


Proof The proof of this theorem is based on formula (1.52) and on the fact
that any generalized function has a finite order (see Gel'fand and Shilov, 1968).

1.4.4. Generalized Functions of Quantum Field Theory

Initial objects of the local quantum field theory are singular functions, for
example:
the causal Green function Ac(x) (x = X o , x) or the propagator of the
particle with mass m,
the positive-frequency part A(_)(x) of the Pauli-Jordan function
A(x - y) = [cp(x), cp(y)], where cp(x) is the field operator.
Explicit forms of these functions in the case of the scalar field cp(x), are given
by the following formulas (for detail, see Bogolubov and Shirkov, 1980):
A(_)(x - y) = <01 cp(x) cp(y) 10), Ac(x - y) = <01 T[cp(x)cp(y)] 10),

(1.54)
28 Chapter 1

and
Llc(x) = F((2n)-2i-l(m2 - p2 - ia)-I)

f
== (2n)-4i- 1 d 4p e -i Px(m 2 - p2 - ia)-1

(1.55)

and J 1, N l ' Hil), and K 1 are the Bessel, Neiman, Hankel and Kel'vin
functions, respectively. Notice that in the case of QFT space-time is the
Minkowski, i.e., R = 1R4.
We see that these functions have a higher singularity of the type of <5(,1.),8(,1.),
A. -1/2 on the light cone A. = o. These singularities are connected with the
ultraviolet divergences in the theory. Removal of the latter leads, in fact, to
regularization of the singular functions Ll(_j(x) and Llc(x). On the other hand,
the regularization procedure of the generalized functions Ll( _ j (x) and Llc (x) is
crucial for their multiplication, convolution, etc. Due to the presence of the <5-
function in expressions (1.54) and (1.55) for Ll(_j(x) and Llc(x) the definition of
their multiplication: Ll,7(x) and [Ll(_)(X)]2 is not clear. This operation is con-
nected with the multiplication properties of the generalized functions (for the
<5-function: <5 2 (x) = C <5(x), where C is an arbitrary constant) and requires some
regularization procedure (the Pauli-Villars regularization is often used) giving
finite values for the regularized functions Ll~(x) and Ll(_j(x), i.e., Ll~(O) < 00
and Ll(_)(O) < 00. For example, in the above-mentioned case (Section 1.3), the
regularization procedure played by the method of averaging in stochastic
space-time.
However, it should be noted that the situation for Ll(_j(x) is radically
different from the case of Llc (x). It transpires that the product of two positive-
frequency parts Ll(_)(x) may be defined as the Fourier transform of the
convolution
Foundation of the Nonlocal Model of Quantized Fields 29

= (8rcp2)-18(Po)8(p2 - (M + m)2) x
x {[p2 - (M + m)2][p2 - (M - m)2]}1/2. (1.56)

Here, integration was carried out in an elementary manner: choosing the axis
qo along the time-like vector p and introducing new variables q2, wand n;
qo = -w, q = (w 2 - q2)1/2 n, where n2 = 1, and after integration over q2 and
angles giving the three-dimensional unity vector n, we obtain expression (1.56).
We notice that the function (1.56) is usually continuous and bounded at
infinity.
On the contrary, in the case of de(x) its product contains a finite number of
arbitrary constants which are caused by the products of the type of c5 2 (x) and
c5(x)8(x). The reader may locate the general theory of such productions in
Bogolubov and Parasiuk's (1957) work.
Thus, two main functions de(x) and d(_)(X) are the generalized functions
and therefore all the coefficient functions in the expansion of the S-matrix in a
series of normal products of the field operator <p(x)

S= I
n; 0
~fd4Xl···fd4XnSn(Xl'
n.
... 'Xn): <p(X1)···cp(Xn):

or Wightman's (1972) functions:


Wn(x 1, ... , Xn) = <0 I <p(xd· .. <p(X n) 10), (x = x o, x)
become also the generalized functions. Since the functions Sn(X 1, . .. ,xn) and
Wn(x1, ... ,xn) are expressed by the product of de (x) and d(_)(X), respectively.
For example, we consider the structure of the functions Wn(x 1, ... , xn) which
. are constructed by W 2(x, y) = <0 I <p(x)<p(y) I 0). According to the translation
invariance, W 2 (x, y) represents in fact only a function of difference y - x,
which we have denoted by d(_). This function must possess the following
conditions:
(1) d(_) (Ax) = d(_)(X) for all the Lorentz transformations AEL+.
(2) Support of d(_) belongs to the future light cone f\.
(3) IS d 4 x d 4 yf(x) d(_)(y - x)f(y) ;;;. 0 for all fE L(~2).
(4) d(_)(X) = d(_)( -x), if vector x is space-like.
(5) If, besides <Olcp(x)IO) = 0, then
(6) lim)~CI) d( _) (Ax) = 0, if x is the space-like vector.

Conditions (1)--(4) and (6) are the consequences of the Lorentz covariance,
spectrality, positive definiteness, locality and group property in Garding and
Wightman's axioms, respectively. We recall that the generalized function 1(p)
(p = Po, p),

1(p) = (2rc) - 2 f
d 4 p eipx f(x)

is invariant under the proper Lorentz transformation in ~4 (P) if and only if


30 Chapter 1

the Fourier imagef(x) is invariant under these transformations in ~4(X). If the


Lorentz invariant generalized functionJ(p) is located at the point p = 0 then a
similar formula to (1.53) takes place

J(P)= L CaDa c5(4)(P),


0<; JaJ<;"f

Any generalized function satisfying conditions (1}-(3) and (6) can be repre-
sented in the form (Kallen-Lehmann representation)

L\(_)(x) = r
Jm2>o
d,u(m 2 ) L\(_)(x; m2 ),

where

L\(_)(x; m2 )=(4n)-1 f d 3 pE- l (p, m 2 ) exp[iE(p, m2 )t - ipx] ,

E(p, m2 ) = (lpl2 + m2)1/2


and,u is a polynomial (positive) bounded measure. In the case when d,u(m 2 ) =
c5(m 2 - m5), a field associated with the functions Wn(x l , ••• , x n) = [Xl'" xn] is
called a free field with mass mo' Here
[xl"'x n ] = 0 if n is odd,
[ X 1 " ' X 2n ]="[x.x.]···[x.x.]
~ '1 J1 In In (1.57)

where [xy]=L\(_)(y - x). A summation is carried out over all (2n!)/2 nn!
permutations of the numbers i l , ... , in; j l ' ... ,jn of {1, ... , 2n} such that
i l < '" < in and i l <jl"'" in <jn'
Thus, we see above that in QFT basic objects of study are different classes of
the generalized functions. The type of classes depends essentially on the
physical principles of the theory. In the local QFT concept of locality and
microcausality condition plays an important role, according to which the
character of the generalized functions and type of test functions are defined.
For example, the principle of locality (or microcausality condition) is con-
nected with a definition of the local properties of the test and generalized
functions, and requires among them the existence of functions with bounded
support. In other words, from the mathematical point of view, the concept of
locality is connected with the possibility of giving a definition of the support of
a linear functional on some space of test functions. In turn the concept of
functional support is directly associated with the existence of some subspaces
of functions in a test function space. The functions of these subspaces can be
used as an instrument allowing the study of the space-time property of
functionals in QFT. Usually, as such functions one can consider the functions
of D-space of infinite-differentiable functions with bounded supports. By means
of these functions the space-time properties of the functional are investigated,
Foundation of the Nonlocal Model of Quantized Fields 31

for example, the commutator of Heisenberg's field q>(x):


[q>(X), q>(y)]_ =? (1.58)
or the causality condition for the S-matrix in the Bogolubov form

b (bS S +) -_.
bq>(x) bq>(y)
? (1.59)

The requirement of strict locality due to Jaffe (1966) is directly formulated in


such a way that the space of test functions always contains functions with
bounded supports. According to the concept of locality QFT, as usually
formulated, is based on the assumption that all its generalized functions belong
to the spaces q})' or S' c q})'. The choice of the space S of test functions has
been dictated on the basis that the spaces Sand S' go over themselves under
the Fourier transform (see Section 1.4.3c), which in turn reflects the symmetry
between x- and p-spaces in QFT. Moreover, we concluded from that that the
space S' contains a sufficiently larger reserve of the functions (in particular, the
b-function and all its derivatives, and all the usual polynomial bounded
functions etc.).
It is important to notice that the tempered distributions become traditional
in the study of the local QFT and that the concept of locality is introduced
here in the simplest and most natural manner.
Lately it has become clearer that this class of tempered distributions is not
always an adequate instrument. It transpires that the class of local distributions
can be extended essentially. The most significant results were obtained by
Meiman (1964) and Jaffe (1966). It is noteworthy that the definition of the
concept of microcausality by Meiman and strict localizability by Jaffe takes
first place in the papers of these authors. They obtained different classes of test
functions depending on the definitions introduced. Meinman's and Jaffe's idea
was to choose the 'minimum' class of test functions, meaning that, when
introducing a certain definition of locality, only those distributions which
satisfy the definition introduced, must be defined on the 'minimum' class of test
functions. The other distributions must not be defined with the whole class of
test functions. The requirement of this 'minimality' permits one to obtain the
important physical consequences from this purely mathematical hypothesis
following from the definition of locality alone. We may now consider this
situation in more detail. It should be noted that the assumption of the class of
the generalized functions, is an independent hypothesis which has nontrivial
physical consequences. An important corollary of this hypothesis is the
polynomial boundedness of the S-matrix elements in the momentum space. We
show that the growth of the generalized function in the p-space cannot be
arbitrary if we want to conserve the local properties in the x-space, in
particular, the microcausality principle. Without loss of generality, we consider
the function f(t) of the single variable t. The function f(t) is a generalized
function, i.e., which is a linear functional on some (for the present unde-
32 Chapter 1

termined) space au of the test functions u(t). In order to study the local
properties of the function (in particular, the assumption thatf(t) = for t < 0, °
such type of assertion enters into the formulation of the microcausality
condition) we assume that the space au must contain sufficiently larger reserves
of test functions. It transpires that the local properties of the function f(t) are
connected with an acceptable growth of its Fourier image J(w). It can be
verified according to the following example.
Let the space au ~ of the functions J(w) contains the generalized functions

°
from S' and at the same time all usual functions having the behavior as ebw at
infinity, where Ib I < a (a > is fixed). The space iJlt~ can be regarded as the
space of the linear functionals over the countably normalized space iJlta of
infinitely differentiable functions u(w) with the norms:

Pn(u) = max Sup lealwldku(~)I, n=0,1,2,... (1.60)


k<;;n W dw
It is easy verify that D c iJlt a c S. From the boundedness of the norms (1.60) it
follows that the Fourier transform u(t) of the functions from au a is analytic in
the band
-a<Imt<a (1.61)
and decreases within this band rapidly any negative power of It I at infinity. We
denote this space of Fourier images through OTt a. It is impossible to define the
local properties of the generalized functions in OTt a since they do not contain
finite functions. Moreover, in the case of the generalized functions f Ed/1 ~ it is
irrelevant to say that in the given interval they coincide with some usual
function cp(t), in particular, the assertion of the type off = 0, for t < loses its °
significance.
Indeed, for the linear functionals over the space OTt a there exists no natural
definition of the support in the absence of nonzero-finite test functions in OTt a .
As an example, we consider the series
00 bn dn
f(t) = L ,£5 n (t), £5n(t) = ~d £5(t), (1.62)
n= 0 n. tn

for °< b < a. This series converges in the space iJlt~ because the series
1 dn
L
00
(I, u) = ,u(n)(o)( _b)n = u( -b), dn)(t) = ~ u(t) (1.63)
n= 0 n. dt n

is the Taylor series for the analytic function u(t) inside its region of the
analyticity (b < a).
Furthermore, in spite of the fact that every term of the series (1.62) is located
at zero (and, in particular, that it disappears along the negative semi-axis t) the
support of its sum is all points of any closed contour C b surrounding the point
-b and belonging inside the band (1.61), since, according to the Cauchy
Foundation of the Nonlocal Model of Quantized Fields 33

formula:

u( -b) = -.
1 f dz--.
u(z)
2m Cb z+b

On the other hand, as mentioned above (Section 1.4.3) the series (1.62) [or
(1.63)] diverges with respect to convergence in Sf in accordance with the fact
that the generalized functions from Sf possess definite local properties.
Thus, the question arises: what is the widest class of the generalized
functions which one can use to speak about their local properties? By means of
the Fourier transform, the answer to this question may be roughly formulated
as follows: the Fourier-image J(w) of 'localizable' functions f(t) must increase
slowly any exponent. We have already noted that the existence of the dense set
of the finite functions among the test functions in the x-space ensures the
possibility of the usual definition of the local properties of the generalized
functions. From the previous example for the spaceOZt~, we have seen that if
J(w) can increase exponentially, then it is impossible to define the local
properties of f(t) synonymously. Jaffe (1967) found necessary and sufficient
conditions for a function of the type of
ro
g(z) = L cvz v, Cv ~ 0,
v=o

and that the space S(g) of Fourier transforms of the test functions from S(g)
contained finite functions of x-variables (x = Xo, x). It appears that for this
case it is both sufficient and necessary that the integral

1'0 dw w -2 In g(W2) < 00 (1.64)

converges, for example, the function g(W2) ~ exp[(w/ln w)(ln In w)-l-eJ, e> °
satisfies this condition. Meiman (1964) considered a more general problem, the
essence of which concerns the attempt to define the local properties of the
generalized functions (in particular, the concept of their support) without
having finite test functions. In this more general case one can go to the class of
the generalized functions satisfying the weaker restrictions than (1.64), namely,
which lead to the function increasing not rapidly any linear exponent at
infinity:
I f(p) I < Ce eejpj for any s > 0. (1.65)

In contradistinction to the local theory, nonlocal QFT allows stronger


growth than (1.65) for the generalized functions in p-space. We will now discuss
this problem.
The development of the nonlocal QFT and of the theory with the non-
polynomial interaction Lagrangian has required attracting spaces of analytical
functions among which functions with finite supports are absent. However, the
34 Chapter 1

study of the causality conditions and the locality of the theory is needed to
construct mathematical methods devoted to the investigation of space-time
characteristics of operator-valued generalized functions of the types of (1.58)
and (1.59) in the case when they represent analytical functionals, i.e., they are
defined only in the spaces of analytic functions. This problem has been
considered in detail by Efimov (1968) (see also Fainberg and Soloviev, 1977,
1978; and Soloviev, 1980).
An answer to all the supplied problems connected with the definition of
local and nonlocal generalized functions and the corresponding spaces of test
functions can be obtained by studying the space-time properties of generalized
functions of the type (1.11) which have arisen from the results of averaging in
the stochastic space-time 1R4(X), as was done previously in Section 1.3. We see
that the function K(x) (1.11) is the generalized form of the well-known local
Dirac b-function.
We say that the generalized function (1.11) is given in some test function
space rlIi, if for any f Eillt the functional

(1.66)

is well-defined, i.e., the obtained series converges absolutely. Here cn are some,
generally speaking, complex number coefficients, and we insert the parameter I
into Cn • Passing to the momentum space in (1.66), we obtain

(K, f) = f d 4 pK(p2) J(p) < ro, (1.67)

where

(1.68)

and J(p) is the Fourier transform of f(x). In other words, the generalized
function (1.11) is given on rlIi if series (1.68) defines the function K(p2) for all p2
and the integral (1.67) converges for any f(X)ErlIi. Both conditions (1.66) and
(1.67) are equivalent. The space-time properties of the generalized function
(1.11) are as usual investigated for different sequences of the coefficients {c n },
which correspond to different generalized functions.

1.4.5. The Class of Test Functions in the Nonlocal Case


We choose the space Iff of all entire analytic functions f(z) = f(zo, z) depend-
ing on four variables zl' = xI' + iyl' in the capacity of the class of our test
functions. We define in space Iff the countable sequence of the norms

Ilfllm = max If(zo,z)l·


Iz,1 <m
Foundation of the Nonlocal Model of Quantized Fields 35

We will say that the sequence offunctions {f,,(Z)} , wherefn(z)E g, converges


correctly if it converges uniformly in any bounded region of real values of
arguments (xo, x).
LEMMA 1.2. If each function of the sequence {fn(z)}, where fn(z)E g is bounded
according to norm II fn II m by the constant M and this sequence converges
correctly, then its limit fez) belongs to the space g and the norm of fez) does not
surpass M.
This lemma is the other formulation of Vitaly's theorem concerning the
convergence of analytic function sequences (Titchmarsh, 1939). Therefore, the
norms introduced are compatible and the space g is complete and perfect.
DEFINITION 1.3. The sequence {In(Z)} , where fn(z)E g, converges in the
region G c (:4 if it converges uniformly in the closed region G.
The space g is not closed, relative to this convergence definition, but we will
make use of it for another definition of the support of a functional on the space
g.
The space of all linear continuous functionals defined on the space g is
denoted by g'. The explicit form of such functionals can be found according to
the usual procedure (Gel'fand and Shilov, 1968; Kantorovich and Akilov,
1964):

(F,f) = r
JZ/II<m
dIlF(z)f(z), FEg', (1.69)

where IlF(Z) is a complex and completely additive measure in the region


Iz/' I < m. This formula with all possible m gives the general form of linear
continous functions on the space g.
Let us introduce the concept of functional support, and give the following:
DEFINITION 1.4. The region G c (:4 is said to be the support of the
functional FE g' if this functional can be extended continuously on the space
g( G) of functions, which are analytical in the closed region G. If the region G is
a point Z = (zo, z) then functional is said to be local.
It follows from given definitions that the sequence of numbers (F, fn)
converges to zero if the support of the functional F is a region G and {fn}
wheref"E g is any sequence which converges to zero in G.
DEFINITION 1.5. It is said that the compact BeG is the defining set of the
functional FE g'(G), if for any open m :::> B the functional F is continued on
gem), i.e., if '1m :::> B3Bw is compact in m and 3Cw such that 'If E g(w),

I(F, f)1 < Cw Sup Ifl.


Bw

If FE g'(G), then there exists at least one compact defining F and by the
36 Chapter 1

Hahn-Banach theorem it follows (since $(G) is the closed subspace of the


space I[(G) of continuous functions on G) that there exist compact BeG and
the measure /1 with the support in B, such that

(F, f) = L f d/1, V fE $(G).

Some further definitions are needed here. We denote by 1R 4 n and 1[4n the
spaces of real and complex n vectors x/1 and z/1 = x/1 + iY/1' /1 = 0,1,2,3. Let
G, r, 0., ... , and G, f, Q, ... be open sets in 1R4 and 1[4, respectively. For
denoting different spaces of functions we use the following symbols:
I5lt is some arbitrary space of functions;
d/i' is the space of linear functionals defined on d/i;
C(q) = C(q, 1R4n) is the space of functions f(x 1 , ••. , xn) defined in 1R4n, where
{f(XI , ... , xn)} are q times differentiable and continuous functions together
with all partial derivatives up to qth order inclusively;
C(q, r) = C(q, r, 1R4n) is the space of functions of C(q) for which all the following
products

I ...
Xi v
11
Xi v
apf(x
"ax.J 1/11 . .. aX.xn)
1 , ••. ,

Jp/1p
I< 00

are finite (s = 0, 1, ... , r; p = 0, 1, ... , q; v, /1 = 0, 1,2,3);


Lp = Lp(1R4n) is the space of the functions for which

f f
d4Xl··· d4xn! f(x 1 , · · · , xnW < 00;

Z is the space of all entire functions for which

fd 4 xl ... Jd 4 Xn !f(x 1 + iY1,···, Xn + iYn)! < 00

for any Y1, ... ,Yn;


$ = $(1[4n) is the space of all entire functions;
$(0) and $(0) are the spaces of analytic functions ill regions 0 and 0.,
res pecti vel y;
$a is the space of entire functions of the finite order of growth rx, i.e., for any
g(z) E $a there exist such positive numbers band B, as

!g(ZI,···, zn)! <:;; B exp[b ~ !Zi!aJ


The space Z~ (rx :;;;. 1) consists of all those entire functions f(ZI' ... , zn) of n
complex variables, which satisfy the following requirements:
(1) For any f(ZI, ... , Zn)E Za' there exist such positive numbers C > 0 and
Foundation of the Nonlocal Model of Quantized Fields 37

Aj > 9(j = 1, ... , n), that

If(z1,···,zn)1 .;;;cexp [jt1 AjlzjlaJ

(2) For any Y1, ... ,Yn

The number IX is chosen depending on the interaction Lagrangian [exactly, on


the measure w(bUZZ)] under consideration and the way of introducing non-
locality into the theory.
On the space Za (IX ~ 1) one defines all nonlocal generalized functions K(x),
for which the Fourier transform K(p2) is an entire analytic function in the p2_
plane, the order of which is

p< h = 1X/2(1X - 1).


Thus, Za = Z n Iffa is nontrivial space for rx ~ 1. Z~ is the space of entire
functions such that for any g(z) E Z~ there exist some positive numbers b, c,
and B, that

Ig(z)1 = Ig(X1 + iY1,"" Xn + iYn) 1.;;; B exp { ~ [ -clxY + blYi n }.


The space Z~ is nontrivial in the cases: IX = 1, f3 < 1 and IX > 1, rx ~ f3 > O. The
spaces Z~ coincide actually with the spaces S~ investigated by Gel'fand and
Shilov (1968).
Further, we denote by dfJ, S, fj, Z, and Za the spaces of the Fourier
transforms of the functions belonging to iJ!1, S, ... , and Za, respectively. For
example,

iff(x)EiJ!1, and we have S = S, fj = Z1 nS.


The space Za' which is the space of Fourier transforms of functions f E Za
consists of the differentiable functions 1(p1 , ... , Pn) which satisfy the condition:
(3) There always exist positive numbers C and Bj (j = 1, 2, ... , n) that

11(p1,···,Pn)1 .;;;cexp { - jt 1 BjIPjIY},

where y = IX/(rx - 1) > 1.


Let us now introduce some definitions concerning the entire analytic
functions of the single complex variable z = x + iy IC = IR + i IR E1 1 1.
38 Chapter 1

DEFINITION 1.6. Let g(z) be an entire function and

Mir) = max Ig(z) I


Izl ~ r

then the number


. In In Mir)
p = 11m -------"'---
r~oo Inr
is called the order of the entire function g(z). If the order of the entire function
g(z) equals p, it means that, roughly speaking, g(z) increases as exp[lzIP]' All
entire functions increasing slowly with respect to exp[lzle] (for any e > °
chosen arbitrary small) are called entire functions of zero order p = 0, but all
functions increasing rapidly with respect to exp[lzI N ] for any N > are the
functions of infinite order p = 00.
°
DEFINITION 1.7. Let h(r) be the monotonously increasing function for the
sufficiently large r. If

lim In Mg(r)/h(r) = 1,

then we say that the exact order of g(z) is h(r). If h(r) = ar P , then the number a
is called the type of the function g(z).
The function A(r) is called slowl~' changing or slow, if

lim rA' (r)/A(r) = 0, A' = dA(r)/dr


r~ 00

for example, A(r) = (In r)~, Va or A(r) = In In r. The concept of the exact order of
growth of the function g(z) yields a finer gradation order of growth. The entire
functions with the exact orders h(r) = ar P and h(r) = ar P ' A(r), where the
function A(r) is slowly growing have the same order p.
Let
w
g(z) = I gn zn
n=O

be the entire function represented in the form of series. Let us introduce

mg(r) = maxlgnlrn.
n#O

Then, the following relation is valid in the limit r ~ CI)

In Mk) '" In mk),


whence h(r) ~ Inmy(r).
This relation allows one to estimate with sufficient accuracy an order of
growth of an entire function.
Foundation of the N onlocal Model of Quantized Fields 39

DEFINITION 1.8. The following Fourier representations hold for the entire
functions. If h(r) = ar for g(r), then

g(z) = f dlli() eiz~, (1.70)

where Ilk) is the complex-valued absolutely additive measure in the complex


(-plane, located in the region 1(1 < a + 8 (8) 0). This representation was
proved by Gel'fand and Shilov (1968).
DEFINITION 1.9. If g(z) E $h [$h is the space of the entire functions such that
for any g(z) = g(z 1, ... , zn) there exist such numbers (j > and B > 0, that °
Ig(zu···, zn)1 ~ B ex p{ ~ h[(1 - (j)lzjIJ, Zh = Z n $h}'

where h(r) > 0, h'(r) > 0, hl/(r) > 0, then the representation (1.70) holds when
Il g (O is the complex-valued absolutely additive measure in the complex (-plane
for which there exists such 8 > that °
f1dllg(OI exp{H[(1 + 8)I(IJ} < 00. (1.71)

Here

H(s) = max Est - h(t)]. (1.72)


,?O

The representation (1.71) belongs to Efimov (1970). If h(r) = ar a, then


H(s) = sa/(a- L)(,x _ 1)(,xaa)-1/(a-l).

1.4.6. The Class of Generalized Functions in the Nonlocal Case


We study now the space-time properties of the relativistic invariant genera-
lized functions of type (1.11) or (1.66), for different sequences of coefficients
{c n }. We will distinguish only the following possibilities (for detail see Efimov,
1977a)
(1) cn = 0, for any n > no > ° (no is arbitrary),

(1. 73)

(1.74)
n~ ro

Thus let us consider the functional

(K, f)(x) = f ro
d 4 xK(x)f(x) = n~o (2~)!
c
on f(x). (1.75)
40 Chapter 1

If there exist functions with bounded support in the space IJlt of the test
functions to which the function f(x) belongs, then one can give the following
definition.
DEFINITION 1.10. The generalized function
c
I on b(4)(X)
(fj

K(x) = _n (1.76)
n = 0 (2n)!

is called local and is located at the point x = 0, if (K, f) = 0 for any f(x) E 1Jlt, a
support of which does not contain the point x = O.
We assume that a sufficiency of functions with a bounded support exist in
the space 1Jlt. It means that if some function g(x) is locally integrable and

f d 4 x f(x)g(x) = 0

for any f(X)E D(G) n 1Jlt, then g(x) == 0 almost everywhere. We now consider
the generalized function K(x) for three cases:
Case 1. Local distributions when Cn = 0, n > no > O. For the existence of the
functional (1.75) it is sufficient to require the functions of the test space 2no
times differentiable with respect to all arguments. Thus, we arrive at the spaces
of the type C(q), where q = 2n o. If one considers such K(x) for which no is
arbitrary but is finite for any K(x), then the functions of the test space must be
infinitely differentiable and among those there must be the functions with a
bounded support. So, we come to the spaces D and S. The spaces D, S, and
C(q, r) c C(q) are investigated in detail on the basis of which usual local
quantum field theory' is constructed. In this case the generalized functions K(x)
satisfy Definition 1.10. Therefore, they are local and are usually called distri-
butions of tempered order of growth.

Case 2. Local distributions when limn ~ (c n )l/n = O. The distributions of this


(fj

type in the case of single variable function were considered first by Meiman
(1964) and next by Soloviev (1971). We now define the conditions for which
the functional (1.75) exists. At first we carry out some formal transformations.
We make use of the following integral representation for the operator 0:

(1.77)

where

11 __ 00
(1.78)

The integration in (1.77) is performed over the Euclidean globe


b 2 = bi + b~ + b~ + b~ = b 2 + b~ -< 1. Let us introduce the function of a
Foundation of the Nonloeal Model of Quantized Fields 41

complex variable ~:

Wm = L
00

enan~ -2n-l. (1.79)


n~O

It follows from conditions (1.73) and (1.78) that W(~) has no singularities in
the complex ~-plane except for the essential singularity at ~ = 0 only.
Therefore,

where the integration is performed over any closed contour around the point
~ = O. Making use of (1.77) and (1.79), the functional (1.75) can be written in
the form

(K, f)(x) = f
n~o (2~)! 2~i d~ ~2nw(~) X

(1.80)

In the last term the summation is performed over all n because the integral
over b of the odd powers of (ib 4 a/ax o + b a/ax) is identically equal to zero.
Observing that there is a displacement operator with respect to the arguments
(xo, x) in (1.80), one can get

(K, f) (x) = ~ i d~ W(~)


2m J r
Jb' <; 1
d 4b f(x o + ib4~' X + bOo (1.81)

This expression is well defined because of f(x) E $0 or ZOo


The integration over ~ can be performed over a contour with an arbitrary
small radius because W(~) has the only singularity at the point ~ = O.
Therefore, the functional (1.81) can be extended continuously to the space of
functions which are analytic at the point x = (x o, x). Hence, the generalized
functions (1.76) with the condition (1.73) are local according to Definitions 1.4
and 1.5.
The Fourier transform of the generalized functions (1.76) can be written in
the form
K(p2) = (K, e ipz ) = ~ i d~ W(~)
2m J r
Jb 2 <; 1
d 4b exp( - POb4~ - ipb~)

(1.82)
42 Chapter 1

The function K(p2) is an entire analytic function of the order p < t in the
complex p2-plane, i.e. [see formula (1.65)],

for any 8 > O. According to the theory of entire analytic functions it is known
that for these functions there is not any tlirection in the complex p2-plane,
along which they could decrease. It means that they cannot play the role of the
cutoffJunctions in the construction of the perturbation theory for the S-matrix,
if we want to make use of the quasi-local arbitrariness in constructing the
coefficient functions Sn(X 1, ... ,xn) in the perturbation expansion of the S-
matrix.
The 'minimum' space of test functions in this case forms the class Iffo of
functions which are analytic with respect to each argument z/l = x/l + iY Il in
some band lylll < d ll where d ll depends on the function J(z) and can be an
arbitrary quantity. The fixed band lylll < do, where all functions J E Iff are
analytic, does not exist. In this way we got the relativistic generalization of
Meiman's class Co.
We notice that local distributions, when L{l/n) (c n )1/2n < 00, correspond to
the strict locality due to Jaffe (1966). In this case the theory will be strictly
local, if the space of test functions contains functions with a bounded support.

Case 3. Nonlocal generalized functions when limn~ co (c n )1/n = 12 , where 1 is


some parameter. We consider functional (K,f) (x) for which the formula (1.75)
is valid. Let us make use of the integral representation (1.77) and introduce
the function Wl(~) of the complex variable ~ according to (1.79). In this case it
follows from conditions (1.74) and (1.78) that the function Wl(~) is analytic in
the complex plane ~ outside the circle I~ I = I. Inside this circle it has some
singularities, the positions of which depend on an explicit form of the
coefficients Cn" Therefore,

Repeating the calculations of Case 2, one can obtain for the functional (1.75)
with condition (1.74) the following representation:

This expression is well defined for any J(x) E Iff.


Let us demonstrate that the representation (1.83) is a relativistic covariant on
Iff. We have to prove that the integral in (1.83) over the Euclidean space is
covariant under co-ordinate transformations from the homogeneous Lorentz
Foundation of the Nonlocal Model of Quantized Fields 43

group, I.e.,

= ( d 4 bf(Ao" x" + bo e, Ai" x" + bi e)' (1.84)


Jb 2< 1
where A is a representation of the homogeneous Lorentz group L. The
summation is performed over J1 = 0, 1,2,3. Here, we have introduced the
notation bo = ib 4 and i = 1,2,3.
Any A EL can be represented in the form A = Al A2 A3 where Al and A3 are
usual rotations and A2 is the purely Lorentz transformation along third axes.
The integral (1.84) is obviously covariant with respect to rotations in the three-
dimensional Euclidean space. Therefore, it will be sufficient for us to de-
monstrate the covariance of the representation (1.84) under the purely Lorentz
transformation A2 . Then, the relation (1.84) can be written

( d 4bf((x o + ib 4 e)ch 8 + (X3 + b3e)sh8,x 1 + ble, X2 + b2 e,


Jb 2~ 1
(Xo + ib 4e)sh8 + (X 3 + b3 0ch8)=

1b2 ~ 1
d4bf(xoch8+x3sh8+ib4e,xl+ble,

(1.85)

e
where th 8 = vic, I < I I < l + Il. Let us go over to the polar co-ordinates in
the plane (b 4, b3) and put b4 = r cos cp, b3 = r sin cp. Then, the left-hand side of
(1.85) can be written as

(2n dcp (I dr r f· ( db l dbd(x~ + ire cos (cp + i8),


Jo Jo Jb21 + b22 = I - r2

Xl+ b l e, x 2 + b2 e, x~ + re sin (cp + i8)), (1.86)


where x~ = Xo ch 8 + X3 sh 8, x~ = Xo sh 8 + X3 ch 8. In so far as the function
f(zo, z) is entirely analytic with respect to the arguments z" it will be the
entirely analytic function with respect to the complex argument cp. In doing so
the integrand in (1.86) is periodic along the real axis in the complex plane cp

°
with the period 2n. Therefore, in (1.86) one can displace the integration
contour over cp from to 2n by any purely imaginary number, i.e.,

f 21t dcp A(cp) = liB + 21t


dcp A(cp) =
f21t
) dcp A(cp - i8), (1.87)
o iO 0

where A(cp) is the integrand in (1.86). Thus, the right-hand side of (1.85) is
equal to the left and our assertion is proved.
44 Chapter 1

Now let us study the space-time properties of the generalized functions K(x)
in (1.83), i.e., consider the possible defining sets of the functional (K, f)
according to Definition 1.5. In the representation (1.83) one can give a
somewhat definite element of the Lorentz group A = A(8, cp, lj;) with the fixed
parameters 8, cp, and lj;, and then the functional (K, f) (x = 0) has the
following equivalent form:

(K, f) = ~i d~ W,(~) r2 d4bf(AoJlbJl~' AjJlbJl~)' (1.88)


2m JI<;I~I <;1+, Jb <;1
This functional can be extended to the space C(0/(8, cp, lj;)) of functions which
are analytic in the region
0/(8, cp, lj;) = {z: 1AoJlzJl 12 + 1A1JlzJll2 + 1A2JlzJll2 + 1A3JlzJll2 < [2 + c}.
In particular,
0/(0,0,0) = {z: Izo 12 + IZl12 + IZ212 + IZ312 < [2 + e},
-
0/(8.0,0) ={z:lz o ch8+z 3 sh81 2 +lz11 2 +lz21 2 +lz3ch8+
+ Zo sh 81 2 < [2 + e}.
Projections of these regions onto the plane (x o, x 3 ) are shown in Figure 1.

Fig. 1. The projections of some possible defining sets of the functional (1.88) on the (xo x 3 )-plane.
1 is the projection 0, (8, 0, 0) and 2 is the projection 0, (0, 0, 0).

Thus, the sets 0/(8, cp, lj;) are defining for the function (1.88); there is none
among them, that can easily be seen, which would be contained in all the
others. Therefore, a support of the generalized function K(x) cannot be given
synonymously. In this case the Fourier transform of K(x) acquires the form

i d~ W,W
Jr POb4~ -
K(p2) = 1.
(K, eiPz ) = -2 d 4b exp( - ipbO
nl J I <; I ¢ I < I + , b2 <; 1
Foundation of the Nonlocal Model of Quantized Fields 45

=~! d~ Tt;(02n2Jl(~(_p2)l/2)(_p2)-1/2/~. (1.89)


2m JI<I~I<I+f.
From this, it follows that the function K(p2) is an entire analytic function in the
complex p2-plane and has the order of growth p = t and type I, and for which the
following estimation:
(1.90)
is valid. Such functions may have only one direction in the complex plane along
which they may decrease. This is just the case which will be used below.
The 'minimum' space of test functions in this case is the class 01 of functions
which are analytic in the region r 1 C (;4. The point ZE r 1, if _12 « Y6 - y2 « 12.
Chapter 2

The Basic Problems of Nonlocal


Quantum Field Theory

2.1. Nonlocality and the Interaction Lagrangian

In this chapter we present the main ideas of constructing the nonlocal theory
of quantized fields. Our basic postulation states that in quantum field theory
nonlocality arising from the stochastic nature of space-time is manifested only
in interaction processes between fields. In other words, it means that we must
introduce nonlocality into the interaction Lagrangians of different fields. This
proposal has some basis. Indeed, if we maintain the Einstein relation between
momentum and energy E = (p2 + m 2 )1/2 to be valid for extended fields
constructed by means of the relativistic invariant distributions KWO) (1.11),
then they will satisfy free field equations, for example, in the case of the one-
component scalar field IPR(X) = </J(x):

(2.1)

According to our construction, the solution to equation (2.1) has the form

(2.2)

From this it is easy to see that due to the form factor K(p2[2) and the function
b(p2 - m2), the extended field (2.2) is, indeed, free and local (the normalization
condition K(m 2 F) = 1 is used here). Therefore, the free field case does not give
new results in our proposed scheme.
Thus non locality is introduced into interaction Lagrangians ;t'in(X) while free
Lagrangians ;t'o(x) of systems are considered to be local. Within this proposal,
the self-consistent nonlocal model of QFT was formulated by Efimov (1977a)
in the framework of the Bogolubov, Medvedev and Polivanov (1958) axio-
matics (see Bogolubov and Shirkov, 1980).
The outline of the construction of the S-matrix in the nonlocal theory
consists of the following. The general axioms are formulated, then a certain
nonlocal Lagrangian of the classical field is considered. For example, in the

46
The Basic Problems of Nonlocal Quantum Field Theory 47

case of the scalar field cp(x) the total Lagrangian of this kind can be

(2.3)

where K(x - y) is our nonlocal form factor (generalized function). In order to


construct the S-matrix in perturbation theory, the correspondence principle is
used, which states that for infinitesimal coupling constant 9 the S-matrix has
the form

(2.4)

where cp(x) is the quantized scalar field which satisfies equation (2.1). In
constructing the highest perturbation orders, we then introduce the nonlocality
method averaged in (1.8), according to which the propagator of the scalar
particle changes to the following [see formula (1.13)]:

1 [K(k2F)] 2 _ 2 2 1
2 . -+ m 2 - k2 - Ie. - V( -k I) m2 - k 2 - Ie
k2 - Ie . (2.5)
m -

in the momentum space jR4(k). Further, there is indicated the class of functions
to which the form factor V( - P 12) belongs (see Chapter 1), and an in-
termediate regularization (see below) is introduced in the framework of which
the finite S-matrix is constructed. At the end of the calculations one can
remove an intermediate regularization, and thus prove that the S-matrix
satisfies all general axioms: unitarity, causality, covariance, gauge invariance
and so on. In the nonlocal theory one should construct a special method
of quantization of nonlocal classical fields of type (2.3).
We present in the following two sections a canonical quantization method
of nonlocal classical fields, and the physical meaning of the form factors K (1 2 0)
of the theory. In Section 2.4 we formulate the causality condition and unitarity
for the nonlocal S-matrix in Bogolubov-Efimov axioms and prove the un-
itarity of the S-matrix in a functional form. Further, in the case of QFT with
non local interactions, we obtain the Schrodinger-type dynamical equation and
show that the finite and unitary S-matrix describing non local interactions of
quantized fields is a solution of the Cauchy problem for this evolution equation
with retardation.

2.2. Quantization of Nonlocal Field Theory

In this section we propose a procedure of canonical quantization of nonlocal


fields which are described by the Lagrangian of type (2.3). According to the
48 Chapter 2

idea due to Efimov (1974) we introduce a certain regularization into the


classical Lagrangian in such a way which permits us to carry out the usual
canonical quantization. This quantization leads to the appearance of additional
ghost states with indefinite metrics. The ghost states disappear when the
regularization is removed, but a trace remains, namely the propagator of the
scalar particle becomes nonlocal according to (2.5).
Distant analogy of the procedure proposed is the method of quantization of
the electromagnetic field (see, for example, Bogolubov and Shirkov, 1980). The
real physical photon can be in two states with transverse polarizations only.
However, these states cannot lead to the correct propagator of the photon. In
order to get the propagator of the virtual photon in the form

(2.6)
it is necessary to introduce non-physical scalar and longitudinal quanta with
indefinite metrics for consideration. The Lorentz condition
(\A~-)(x)lany physical states) = 0

and the current conservation o,J /l(x) = 0 guarantee that the scalar and
longitudinal quanta do not appear in any interactions of physical transverse
photons with charged particles, say, electrons. However, their role is that they
contribute to the propagator (2.6), i.e., the virtual photon consists of trans-
verse, longitudinal and scalar quanta.

2.2.1. Formulation of the Quantization Problem


Let us consider a one-component scalar field o/(x). The Lagrangian density
describing the classical field o/(x) can be written in the form with accordance to
our above proposals:
.ct(x) = !o/(x)(O - m2 )o/(x) + gU(KUZO)o/(x», (2.7)
here g is the coupling constant, U(z) is a function describing self-interaction of
the scalar field o/(x). We suppose that U(z) is analytic in the neighborhood of
the real axis in the complex z = u + iv-plane. In all other respects it is an
arbitrary function. The classe., of functions U(z) for which the finite nonlocal S-
matrix can be constructed, have been considered in detail by Alebastrov and
Efimov (1972, 1973).
The operator KWO) in (2.7) is non local and can be presented in the form of
(1.11), so that the generalized function K(x - y) = KUZO)<5(4)(X - y) belongs
to one of the spaces of generalized functions, considered in Chapter 1. In what
follows, for convenience, we consider the operator
V( -PO) = [K(ZZ0)]2 (2.8)
and suppose that the function V(z) satisfies the following
The Basic Problems of Nonlocal Quantum Field Theory 49

CONDITION 2.1
t ~ p ~ 00,
°
(1) V(z) is an entire function of the finite order i.e.,
:lC > 0, b >
lV(z) I ~ C exp{b I ziP},

(2) [V(z)] * = V(z*),

°
(3) V(-m 2 12 ) = 1,
(4) V(x) ?> for real x,

{
O(1/IZI2) Rez -+ + 00 (k 2 -+ - (0)
(5)
V(z) = O(exp(blzI P )) Rez -+ - 00 (k 2 -+ + (0).
In the expansion
00

(6) V(z)= L vn[-z-m 2 12 ]n thecoefficientsvnsatisfy


n=O

1, vn > 0,
°
(7) Vo = Vn,
(8) :lC > 0, A> that Vn < CAn/nn/p + 1).
Thus we have described all the quantities in the Lagrangian density (2.7).
Notice that the generalized function K(x) = KUZO) <5(4)(X), where KWO)
satisfies Condition 2.1, is defined on the space Za of test functions for
rx < 2p/(2p - 1).
The wave equation for the classical scalar field fP(x) can be obtained
according to the principle of stationary action and may be represented in the
form
(0 - m2)fP(x) = -gKUZO)U'(K(PO)fP(x)).
Let us introduce the extended field (1.8)
<fJ(x) = fPR(X) = K(PO)fP(x)
Then the total Lagrangian density (2.7) for the field <fJ(x) is
5i'(x) = t<fJ(x)E(O)<fJ(x) + gU(<fJ(x)), (2.9)
where
(2.10)
Formally the wave equation is written as
E(O)<fJ(x) = -gU'(<fJ(x)) (2.11)
for the interaction field, and
E(O)<fJ(x) = [(0 - m2)/V( -120)]<fJ(x) = ° (2.12)
for the free field.
The problem is how to understand these equations, how to study and solve
50 Chapter 2

them and how to perform the quantization of the field cp(x). We proceed in the
following way in accordance with Efimov's method. Instead of equations (2.11)
or (2.12) we consider a regularized equation
(2.13)
or, in the case g = 0,
(2.14)
Here 0 is a parameter of the regularization such that

lim EO(O) = E(O) = (0 - m2 )jV( _12 0). (2.15)


O~O

Accordingly, instead of the Lagrangian (2.9) we obtain


5i'(x) = tcpO(x)EO(O)cpO(x) + gU(cpO(x)). (2.16)
Our regularization is chosen in such a way that the function
EO(k 2 ) = (k 2 _ m2 )jV O( _ Pk 2 )

has zeros in some set of points:

n= [k
C/O

EO(k 2 ) ~ (P - m2 ) 2 - mJ(o)] (2.17)


j 1

such that mJ(o) > 0 (j = 1,2, ., .) and mJ(o) -> 00 when 0 -> O. Then the field
cpO(x) (0 > 0) can be quantized by the methods with indefinite metrics
(Blokhintsev, 1947; Pais and Uhlenbeck, 1950; and Nady, 1966).
The Hamiltonian H~ and the vector space of states £0 with indefinite
metrics can be constructed for the free system when 0 > O. Further, the SO-
matrix, and also different operators: the current
J6(x) = i[bSOjbCP"(x)]SO+,
the Green functions
GO(x - y) = <0 1 T [cpO(x)cpO(y )SO] 1 0),
and the Wightman functions
WO(x 1 , · · · , x n ) = <01 cpO(xd'" cpO(x n ) 10),
etc., can be found for the interacting system.
By definition, we consider that when b -> 0 the limits of all these physical
quantities are the quantum field solution of the initial system (2.9). The
problem indicates such a regularization procedure which provides the exis-
tence of the limits of all operators and matrix elements for any physical
quantities at 0 -> O. It means that we have to obtain a self-consistent theory in
the limit b -> O. This section is devoted to this problem.
The Basic Problems of Nonlocal Quantum Field Theory 51

2.2.2. Regularization Procedure


The regularization is introduced in the following way. Instead of the function

L vn(k Z -
00

V( _kZIZ) = mZ)"lzn, (2.18)


n=O

we introduce the regularized function


00 v (kZ _ mZ)"IZn
V b( - kZF) = n ~0 IIi! i (1 n_ b/j" (kZ _ mZ)jmZ) , (2.19)

where (J < lip « 2. It is convenient to denote

Then

Vb(-ZA)= I {n+z un(z-l)n }


n =0 TI [1 - bjj<1(z - 1)]
j = 1

-1)" un(z - 1)"[(n + 2)!Y}


L
00 {(

n+2 '
(2.20)
n= 0 bn + Z TI [z - J.lj(b)]
j =1

where J.lj(b) = 1 + fib. Let us consider the function

(2.21)

The function

(2.22)
is a meromorphic analytic function in the complex z-plane and has the simple
poles at the points
z = J.lP) = 1 + fib U= 0, 1,2, ...)
and decreases as [except the ray (z: arg z = 0)]
1
L
00

-b- ~lzl-3 (-1)"b- n - Z un [(n+2)!y,


E (z) n =0
when Izl ~ 00. Here the series converges according to Condition 2.1. For the
meromorphic function (2.22) the following representation is valid

1
Eb(Z)
I {( -l)j Aib)}, (2.23)
j =0 z - J.lib)
52 Chapter 2

where
oco

L
n ~ max{O,j- 2)
(2.24)

[(n + 2)!]" 1 (kG _ 'G) (2.25)


en, j = '" j - 1 '" n +2 } .
""'i~O ""'k~j+1

The numbers A/b) satisfy the following conditions


00

L (-I)i Aj(b)[,u/b)]N = 0, for N = 0, 1,2. (2.26)


j ~ 0

The inequality (Efimov, 1977a)


A/b) « const· exp{ B(j" I by} Ir( (1 I P - (J )j)
is valid for the numbers Aj(b) at the large j, where B is some number and
(JP < 1.
Let us consider now the function
Dii(k2) = _ [Eii(k2)]-1

which has the properties:


(1) It is a meromorphic analytic function in the complex k2 -plane and has
the simple poles at the points
k 2 = m;(b) = m2 . ,uj(b) = m2(1 + fib) U = 0, 1,2, ... )
(2) The residues of Dii(k2) at these points are

(3) When I k2 1 --> 00 in all P-planes, except the ray [m 2 , (0),


Dii(k2) = Vii(-k2[2)!(m2 - k 2) = O(l/lk 213).
(4) The function Dii(k2) may have zeros at the points k2 = ar(r = 1,2, ... )
(5) lim Dii(k2) = V( _k 2 [2)/(m 2 - k 2).
ii~ 0

2.2.3. Quantization of the Regularized Equation


Let us consider the classical system described by the Lagrangian density 5f'ii(X)
(2.16), where the regularized operator Eii(D) satisfies the properties enumer-
ated above. According to the principle of stationary action the wave equation
for the system described by (2.16) has the form (2.13). It is the differential
equation of the infinite order, i.e., it is an integral equation. In order to solve the
The Basic Problems of Nonlocal Quantum Field Theory 53

Cauchy problem we have to know the values of the function q/(x) and all its
derivatives in the initial moment of time.
We analyse the solution of this equation following the scheme proposed by
Pais and Uhlenbeck (1950). Let us introduce a system of fields
¢1(x) = [Ak5)r J2 {E b(O)/[O - mJ(b)]}¢b(x), (2.27)
where the coefficients A k5) are introduced by formula (2.24), and
mJ(b) = m2(1 + j"/b) U = 0, 1,2, ... ) (2.28)
According to definition (2.27), the fields ¢1(x) are not independent for different
j and they satisfy the relations

[A j(b)r J2 {[O E~(~J(b)]}¢f(x) = [A;(b)]lJ2 {CD E~(~;(b)]}¢J(X). (2.29)

The field ¢b(X) can be expressed as


cc
¢b(X) = I (_1)i[A/b)]lJ2 ¢1(x). (2.30)
j = 0

In fact, on the one hand, the following chain of equalities

¢b(X) = j~O (-1)j A/b){O E~(:J(b)}¢b(X)


= [1/Eb(O)]Eb(O)¢b(X) = rfJb(X)
is valid. On the other hand, using (2.29) it is possible to obtain

¢1(x) = [A/b)]l J2{O E~(:J(b)} i~O ( _1)i[AJb)]lJ2 rfJf(x)

= Jo (-1)i[A Jb)]l J2[Ai(b)]l J2{O E~(:I(b)}¢1(X)

= f
i= 0
(_1)i{
0
~i(b~
m (b)
i
}Eb(O)¢1(x) = rfJ1(x).

On the basis of the correlations (2.27), (2.29) and (2.30) the Lagrangian
density can be expressed in terms of the fields rfJ (x) 1
1 00

5£b(X) =:2 j~O (-1)j rfJJ(x)[O - mJ(b)]rfJ1(x) +

(2.31 )

Assuming the fields rfJ1(x) are independent variables and making use of the
54 Chapter 2

principle of stationary action, from equation (2.31) we obtain the following


infinite system of equations

[0 - m;(J)]4>1(x) = -g[AiJ)] 1/2 u{ to ( -1nA;(J)]1/2 4>t(X))


(j = 0,1,2, ... ). (2.32)
The system of equations obtained here is equivalent to (2.13). Indeed, first, the
relation (2.29) results indirectly from the system of equations (2.32), secondly,
equation (2.13) is fulfilled if 4>°(x) is connected with the fields 4>1(x) by relation
(2.30) and making use of (2.30), (2.29) and (2.32) it is easy to verify
00

EO(O)4>°(x) = EO(O) L (-1)j[A/J)r/ 24>1(x)


j =0

~ . 1/2{0 - mf(J)}
= j~O (-1)l[AP)] [A;(J)] 1/2 x

x [A.(J)]1/2 EO(O) ,J..O(x)


, 0 - mf(J) '+' J

where i = 0, 1,2, ....


Thus the Lagrangian (2.31) and the system of equations (2.32) are com-
pletely equivalent to the Lagrangian (2.16) and equation (2.13). Therefore, one
can assume that our initial system (2.16) is described by the Lagrangian (2.31)
where the fields 4>1(x) are independent and satisfy the equation of motion
(2.32).
The method used here is well known in the theory of differential equations.
It is usually used when a differential equation of the highest order is. replaced
by a system of differential equations of the first order.
All the above-stated arguments concerned the classical field theory. The
quantization of the system of the classical fields {4>1 (x)} can be performed
according to the canonical procedure of quantization. Let us introduce a momentum
field, conjugate to 4> jo(x, 0)

rr1(x,0) = [J/J¢1(x, 0)] f d 3 Y 5£O(y, 0) = ( -1)j ¢1(x, 0). (2.33)

The dot denotes the differential with respect to time


4>j(x,O) = at ° t)lt =0'
a 4>j(x,
The Basic Problems of Nonlocal Quantum Field Theory 55

We treat ¢1 and IT1 as operators with the commutation relations:

[¢1(x,0), ¢?(y,O)]_ = [IT1(x,0),ITt(y,0)]- = 0,

(2.34)

or
(2.35)
It is seen that the indefinite metrics is to be employed to quantize our system in
a regular manner (see Nady, 1966).
As we are unable to solve the system of equations (2.32) exactly, our
problem is to construct the perturbation series for the S-matrix and we
perform the quantization of the non-interacting system of fields {¢1(x)}, i.e.,
the case when g = 0. Instead of (2.32) we have
[0 - mJ(b)] ¢1(x) = ° (j = 0,1,2, ... ). (2.36)
The solution of these equations can be written in the form

¢1(x) = (2n) - 3/2 f 3k[2OJ~'kr


d 1/2 (d jk e -ikx + d; eikx ), (2.37)

where
OJ~'k = [k 2 + mJ(b)] 1/2 = [k 2 + m2(1 +j"/b)r/ 2.
The quantization conditions (2.34) and (2.35) lead to the operators dJ'k and
dJt to satisfy
[dJ'k' dw ]- = [djt A,t,]- = 0.
[djk , d/k ,] - = ( -1)j bji' b(3) (k - k'). (2.38)
The Hamiltonian of the non-interacting system can easily be obtained. Let us
write it in the normal form

f
Hg = j~O (-1)1" d3kOJ~'k dJt djk. (2.39)

The system under consideration consists of quanta with the following mass
spectra
m2 j = 0,
{ (2.40)
mJ(b) = m2(1 + j"jb) j = 1,2, ...
Let us denote
(2.41)
56 Chapter 2

When (j ---> 0, the masses of quanta with j = 1,2, ... go to infinity, according to
(2.40). These quanta are called ghost states or ghosts. The quanta with j = 0
have the finite mass m. We call them normal particles or scalar particles with
mass m.
The space of states Yfb is a vector space with indefinite metrics. It consists of:
(1) a vacuum state 10>, that is unique, defined by the conditions
djk 10> = 0 and normalized by <010> = 1;
(2) one-particle states Ij, k> = dJi; 10> which are normalized by
<j, kif, k' >= ( _l)i (jjjl (j(3) (k - k')

and are eigenstates of the Hamiltonian H~


H~ Ij, k> = W~K Ij, k>;
(3) many-particle states. If there are n particles with momenta k l , ... , k n and
among them there are V l , V 2 , •.• , V~ (n = V l + V 2 + ... + v~) identical particles,
i.e., with the same index j, then the following is given:
In> = Ijl, k l ;· .. ;jn, k n> = (v l ! ... v.!)-1/2
,....
d+
J1 k 1
... d+k 10>.
ill ,;

These states are also eigenstates of H~:


H~ln> = (wtkl + ... + wJ.kJln>.
All these states generate a complete system of eigenstates in the vector space
Yfb, i.e.,

0=10><01+
Dr
~.
n-1J" ... ,Jn-O
I._
(-1)il+ ... +jnfd3kl···fd3knln><nl=1.

What happens with the space Yfb when (j ---> O? At (j ---> 0 the masses of all
ghosts increase according to (2.40). Therefore, if any physical state is characte-
rized by a definite value of energy, then in the limit (j ---> 0 no physical states
with arbitrary but finite energy can consist of ghost quanta. In this sense we
have
lim Yfb = Yf, (2.42)
b~ 0

where Yf is the Hilbert space which contains:


(1) a vacuum state 10>, aklO> = 0,
(2) single- and many-particle states
In> = Ikl, ... ,kn> = (n!)-1/2~~ "·~:IO>.
All these states generate the complete system in Yf:

® = 10><01 +
Dr n
I= 1
fd3kl .. ·fd3knlkb ... ,kn><kl, ... ,knl = 1.
The Basic Problems of Nonlocal Quantum Field Theory 57

2.2.4. Green Functions of the Field ¢O(x)


First of all, let us consider the commutator

LlO(x _ y) = [¢O(x), ¢O(y)]_. (2.43)

Substituting the representation (2.30) into (2.43) and using (2.37) we obtain
00

LlO(x) = I (-1)j A/6) Llix). (2.44)


j = 0

Calculation of the explicit form of the different two-point Green functions


carried out in any textbooks of field theory (see, for example Section 1.4.4 and
Bogolubov and Shirkov, 1980). We use here their results. Thus we have

LlJ(x) = (2n)-3 fd 4 ks(ko) 6(k 2 - mJ(6)) e- ikx

= 2~i s(x o) 6(x 2) - [ ~~~) }X2) -1/2 8(x 2)J (mj (6)(X 2)1/2),
1 (2.45)

where
1
s(t) = { _ 1
t
t
>0
<0 ' 8(t) = {1o t>
t
0.
<0

Because the series (2.44) converges absolutely we have


LlO(x) = 0 when x 2 < O. (2.46)
Thus the operator ¢O(x) satisfies the local commutation relations.
Now let us introduce functions Llt+) (x) according to
Llt-)(x - y) = Llt+)(y - x) = <OI¢O(x)¢O(y)IO).
We have
00

Ll1±) (x) = I i
(-1) jA 6)LlJ(±)(x), (2.47)
j =0
where

For x 2 -> 0, according to Bogolubov and Shirkov (1980), one can obtain

° . 1
I _ 2
2
mj 2 2
Llj(t)(X) = - 4n s( +xo) 6(x ) - 4x 2n 2 + 16n 2ln (mj Ix 1/4) +
+ (imJ/16n)s(=t xo)e(x 2) + O(x 2 ln x 2).
58 Chapter 2

Substituting this expansion into (2.47) we obtain


2 00

~t±)(x) = ~
16n
L0 (-1)i A/b)Jlj In Jlj + O(X2 In x 2).
j=

Here we have used the correlations (2.26), hence it appears that the function
~t±)(x) is finite at x = 0 and
m2 00

~t±) (0) = 16n 2 j ~o ( -1)i A j{b)Jlj In Jlj < 00. (2.48)

It means that the operator cpb(X) is well-defined because of


<Olcpb(X)cpb(X)IO) = ~t-)(O) < 00.

Let us consider the causal Green function


00

~:(x - y) = <01 T[cpb(X) cpb(y)] 10) = L (-1)j Aj(b) ~1c(x - y), (2.49)
j= 0

where

~1c(x) = -i(2n)-4 f d 4 k e-ikx[mJ(b) - k 2 - ier 1 •

Otherwise

(2.50)

where
00

Li:(k2) = L (-1)j Aib)[mJ(b) - k 2 - ier 1


j =0

The function Li:(k2) is analytic in the complex k2 -plane for 1m k 2 ~ 0 and has
simple poles at the points k 2 = mJ(b) - ie U = 0, 1, ... ).
The retarded and advanced Green functions can be defined in the following
way
~~et(x) = O(xo) ~b(X) = ~:(x) + ~t+)(x),
~:dV(X) = -O( -x o) ~b(X) = ~:(x) - ~t-)(x). (2.52)
They satisfy the conditions
X2 < 0
~et(x) = 0 for { 2 0
X > , Xo < 0,
The Basic Problems of N onlocal Quantum Field Theory 59

X2 < 0
~dv (x) = 0 for { 2 0 0 (2.53)
X > , Xo > .
Thus we can see that all Green functions satisfy all requirements of the local
quantum field theory. It means that the field 4J~(x) is local. The following
correlations are valid for the regularized Green L\:(x) and L\1±)(x) functions:

L\:(x) = O(x o) L\1-)(x) + O( -xo) L\1+)(x),


(2.54)
L\:*(x) = O(xo) L\1+)(x) + O( -xo) L\1-)(x).
An additional correlation should be mentioned because it is important when
proving the unitarity of the S-matrix regularized

In other words, the T-product in the Wick sense coincides with the T -product
in the Dyson sense, i.e., symbolically Tw = Td • Indeed, it is easy to obtain for
the fields 4Jj(x)
(02/OX/L oy.) <01 T{4Jj(x)4Jt(y)} 10)
= <01 T {O~/L 4Jj(x) O~. 4Jt(y)}O) + i( -1)j c5 ji c5/LO c5. o c5(4)(X - y).

Therefore,

O 020 <OIT{4J~(x)4J~(Y)}IO) - <OIT{-oO 4J~(X)-oo 4J~(Y)}IO)


x/L Yv xp y.

L
0()

=i c5/LO c5.o c5(4)(X - y) (-1)j Aic5) =0


j=O

according to (2.26). Thus the following correlations take place


o/Lo. L\:(x) = O(xo) o/Lo. L\1-)(x) + O( -xo) o/Lo. L\1+)(x)
(2.56)
o/Lo. L\t-)(x) = O(x o) o/Lo. L\:(x) + O( -x o) o/Lo. L\:*(x)

2.2.5. The Interacting System Before Removal of the Regularization


The interacting system is described by the Lagrangian (2.31). The total
Hamiltonian of this system has the form
H~=Hg +H!, (2.57)
where Hg is given by the relation (2.39) and

H! = -g fd3X:U(4J~(X'0)):. (2.58)
60 Chapter 2

In the interaction picture, we have

H!(t) = e- itHg Hfn eitHg = -g f d3 x:U(¢"(x, t»:. (2.59)

Though the operator ¢o(x) is well-defined as we have shown in the previous


section, the Hamiltonian H! (t) is not defined, since the theory is trans-
lationally invariant and, thus, there are difficulties connected with the Haag
theorem (see Wightman, 1964). It is necessary to introduce an operation of
'switching on' and 'switching off' the interaction g:
g ---+ g(x/L, tiL) = g(xIL).
Such regularization takes into account at the same time that, first, our system
is situated in a box, hereby violating the Euclidean in variance and, secondly,
the interaction is adiabatically switched on and switched off at infinity, i.e.,
when t ---+ ± 00. The large parameter L defines the intensity of switching on the
interaction. The function g(x) satisfies the conditions:
(1) 0 <:;;; g(x) < g,
(2) g(O) = g
(3) dng(x)lx = 0 = 0, for n = 1,2,3,4; where d n = 0lll .. , 0lln

(4) f d 4 xg(x) < 00,

(5) g(x) E Z~, for IX < 2pl(2p - 1),


where p is the order of growth of the nonlocal form factors.
Removal of the regularization corresponds to the limit L ---+ 00, at which

g(xIL) ---+ g(O) = g. (2.60)

Thus the Hamiltonian (2.59) in the interaction picture takes the form

H~,L)(t) = - f d 3 x g(x/L, tiL): U(¢O(x, t»:. (2.61)

We recall again that when (j is positive, then the Hamiltonian Hi~,L)(t) is well-
defined, so that it is not necessary to introduce any ultraviolet cutoff.
The regularized S-matrix can be written in the standard form

so,L = Tex p{ - i f~oo dt H~'L)(t)} = Texp{i fd4Xg(XIL): U(¢O(X»} (2.62)

Here T has the meaning of strict ordering operator of the quantized fields
¢O(x) with respect to time. The chronological pairing or the Green function of
the operators ¢O(x) is given by formulas (2.49}-(2.51). Since the operators ¢O(x)
and Hi~,L) (t) are well-defined and local then the So,L-matrix on the vector space
The Basic Problems of Nonlocal Quantum Field Theory 61

,J't6 is unitary and microcausal, i.e.,

(2.63)

The typical problem of the nonlocal quantum field theory consists in proving
of the existence of the following sequence of limits in each perturbation order

SL = lim S6,L, (2.64)


6_0

Here the first limit fJ --+ 0 means removal of all 'ghost' states from the theory.
The second limit L --+ 00 means passage to infinite volume and switching on
the interaction over all four-dimensional space.

2.2.6. The Green Functions in the Limit fJ --+ 0


The Green functions in the limit fJ --+ 0 are generalized functions which are
defined on a space of test functions Za(oc < 2p/(2p - 1». Therefore we have to
consider improper transitions to the limit, i.e., investigate the limits

(2.65)

where G6(x) is a Green function and f(x) E Za' First of all, let us consider the
commutator ,!\6(X). We have

L
00

x (-l)j Aj(fJ)B(ko) fJ(k2 - mJ(fJ»


j= 0

where

Q6 = (2n)-3 d 4 f kJ(k)j~l (-l)j A}fJ)B(ko) fJ(P - mJ(fJ»

= j~l ( -l)j Ap)(2n)-3 ff d 3kdkoJ(k, k o)B(ko )fJ(k 2 - mJ(fJ». (2.66)


62 Chapter 2

The following estimation can be obtained

It = I(2n)-3 ffd3kdkoi(k,kO)8(kO)b(k2 - mJ(b)) I

~ const f
d 3 kwjk O exp{ -B[lkI Y + (w:'Y]},

where y = IX/(IX - 1) > 2p ~ 1. B is some positive number. Making use of the


inequality

Ikl Y + (W~lY ~hlm{ + h2lkl Y,


where hI = 2 -1 + y/2, h2 = 1 + 2 -1 + y/2, further we obtain

/1.;;;; const·exp( -Bhlm)) f d 3k[k 2 + mJ(b)]-1/2 exp (-B·h 2 IkI Y)

.;;;; const·bFGexp{ -B 1(f/bF/ 2 },

where Bl = Bhl mY. Now QO is evaluated to be


00 00

IQol';;;; L Aib)IJ.;;;; const. L F" 15 exp[B(r /b)P - Bl (jG /15)1/2] X


j=1 j=l

x r -1«1/p - a)j).
Because of p <!y, we have
00

IQol';;;; b·const. L F"r- «1/p - 1 a)j)';;;; b·const


j=1

and finally

Thus in the limit 15 -.0 the commutator aO(x) changes into the commutator of
the scalar field <p(x)

lim aO(x) = a(x) = (2n)-3 fd4k8(ko) b(P - m2) e- ikx • (2.67)


0_0

In the same manner it is easy to show that in the improper sense

lim at±)(x) = a(±)(x) = (2n)-3 fd4kO(=Fko) b(k2 - m2) e- ikx • (2.68)


O~O
The Basic Problems of Nonlocal Quantum Field Theory 63

The existence of these limits means that there exists a weak limit (Section 1.4)

lim qi(x) = <p(X), (2.69)


O~O

where <p(x) is the scalar field of mass m satisfying the free wave equation
(0 - m2)<p(x) = o.
Consequently, all ghost states disappear in the limit b ~ O. This result confirms
the statement that

which was accomplished in Section 2.2.3.


Now we consider the causal function ~~(x). We have

lim fd 4 X
o~ 0
~~(x)f(x)

= lim i-1(2n)-4fd4 kj(k)


o~ 0 j
i:= (-1YAib)[m;(b) -
0
k 2 - ier 1

= i- 1(2n)-4 fd4kV(-k212)j(k)[m2 - k 2 - ier 1

because of f(x) E Za.


Thus the causal function ~~(x) changes in the limit b ~ 0 into a nonlocal
propagator
lim Li:(k) = DAk2) = V( _k 2 12)[m2 - k 2 - ier 1 . (2.70)
o~ 0

The function DAk) has a single pole at k 2 = m 2. This pole corresponds to the
scalar particle with mass m. The poles corresponding to all ghost states have
disappeared. The form factor V ( - k 2 12 ) is an entire function and it does not
correspond to any real state. This form factor describes the non local character
of the interaction of our scalar particles.
Thus in the limit b ~ 0, our theory becomes nonlocal. In other words, the
'ghosts' (which disappear in the limit b ~ 0) as a self-memory make the theory
nonlocal. Consequently, the nonlocal character of the interaction of the
classical field (2.7) is revealed in quantum field theory as a residual effect of the
nonphysical ghost states (the origin of which is caused by the stochastic nature
of space-time) when these ghosts are removed by the transition to the limit
b~O.
64 Chapter 2

2.3. The Physical Meaning of the Form Factors

We now give the physical meaning of the form-factor introduced in (1.9). On


the basis of the simple model of quantum field theory, we show that the form-
factors of the class investigated above give the relativistic invariant description
of extended objects in QFT. Let there be an infinitely heavy source situated at
the point x = 0. A field is generated by the radiation and absorption of scalar
particles with mass m. We assume that the interaction Lagrangian density of
this system has the form
5£:n(x) = gcp(x)KWO)J(x), (2.71)
where cp (x) is the field of the scalar particles with mass m and the functiun J (x)
describes the source. Further, let us assume the source as point-like (situated at
the point x = 0) and the switching on and switching off as at the infinitely
remote past and future, respectively. For definiteness, we read
(2.72)
Here T is a large parameter characterizing the time interval during which the
source is switched on. At the end of calculations, it is necessary to pass to the
limit T ~ 00.
The function K(l2 D) characterizes the region of the nonlocal interaction
near the source. We assume that the function K(k 2 [2) satisfies all of Condition
2.1 of Section 2.2.1 for p = t except for the normalization condition which we
will discuss below. The following representation is valid for K(k 2 [2) when
p=t
K(k 2 [2) = ib 2 <12
d 4 bw(b 2)exp(ko b4 - ibk). (2.73)

When ko = 0, we set

K ( - k 2 l2) = f d 3 ba(b 2) eikb • (2.74)

The connection between the functions w(u) and a(u) can be found by the
relations

K(k2F) = fd4bW(b2)eikEb = fd3ra(r2)eikr,

where k 2 = -k~ = _k2,kE = (iko,k), or

K(k 2 [2) = 2n 2 Loo dt.}t J 1[(t( _k2))1/2]( _k 2)-1/2W(t)

roo sin(s( _k2))1/2


= 2n Jo dsa(s) (_k2)1/2 .
The Basic Problems of Nonlocal Quantum Field Theory 65

From these correlations, it is easy to obtain

a(s) = f dtw(t)(t - S)-1/2,

w(t) = -(2nt)-1 f
ds(s - t)-1/2[a(s) + 2sa'(s)]. (2.75)

The, potential generated by the source at the point x = r is equal to the


Heisenberg field operator cp(x) averaged over the vacuum:
W(x) = <01 T[cp(x)S]IO). (2.76)
In the given case, the vacuum state 10) containing no scalar particles is a state
describing the source. We will calculate the potential generated by the
field cp(x) near the source.
The method of calculation is as follows: We construct the total S-matrix
according to the methods developed above. It must satisfy the stability
condition <OISIO) = 1. However, in order to satisfy this condition it is
necessary to add a counter-term to the Lagrangian (2.71). This counter-term
AE is responsible for the renormalization of the energy of the basic state of the
source, or the vacuum energy, and describes the interaction energy of the
source with the field of the scalar particles. Since we assume that an operation
of 'switching on' and 'switching off' the interaction is adiabatic, this counter-
term must have the form
(2.77)
As shown above, we now introduce the regularization procedure associated
with introducing the nonlocality into the interaction Lagrangian. According to
the quantization procedure presented in the preceding section, we must
quantize two fields cp(x) and ¢(x) = KW O)cp(x) is such a way that the
quantized fields cpo(x) and ¢o(x) lead to the following causal functions:

Ac(x-y)=<OIT[cpO(x)cpO(y)]IO) = i- 1(2n)-4f d4k e2-ik:2-Y).'


m - -16

this can be done as shown above.


In accordance with the above deduction for the S-matrix we obtain

S = lim lim sO.T, (2.78)


T--H:IJ (j---+O
66 Chapter 2

where

The averaged values of the field operators, i.e., the Green junctions, are
calculated by the formula

= lim lim (01 T[(l(xd' .. <pb (Xn)Sb,T] 10). (2.80)


T....."oo 15--+0

Making use of the Wick theorem we obtain


Sb,T = St,T <OISb,TIO),

where

(2.81)

(2.82)

The problem is how to choose such I1E that

lim lim (OISb,T 10) = 1. (2.83)


T_ 00 b--+O

Let us consider expression (2.82). Using (2.72) and (2.77), and carrying out
some simple calculations, we obtain

lim (-tg2)ffd4Xl d4 x 2J(xl> T)D:(x 1 -X 2)J(X 2, T)


b~O

-i f oo

-00
1
dx o I1E(xo) = -iT(2n)1/2 ·-·I1E.
2
The Basic Problems of Nonloeal Quantum Field Theory 67

Thus the limits in (2.78) and (2.83) exist if

f..E = (2n)-3(tg2) f d 3k[K( _k 2/2)]2(m 2 +k2)-1 < 00. (2.84)

The regularized S-matrix is given by expression (2.81).


Now we consider the potential W(x). Substituting the S~,T -matrix (2.81) into
(2.76), we obtain

= lim lim ig fd 4 yq)~(x - y)J(y, T). (2.85)


T-+oo 0-+0

°
Passing to the limits c5 ~ and T ~ 00, and making some simple calculations,
we see that W(x) does not depend on Xo and takes the form

W(r) = W(r, 0) = (2n)-3 g fd3kK(-k2/2) eil<r (m2+k2)-1. (2.86)

Making use of representation (2.74), we have

W(r) = :n f d 3 ba(b 2 )exp ( -171~ ~Ibl)

= f:
{,~ {e-•• db~a(b')'hmb + 'h (mr) f dbba(b')e-"}' r <1
(2.87)
JL K(m2 [2) _e_, r> I
4n r

Here

K-( m 2/2) = fd 3 ra (2)sh


r
m(r2)112
( 2 112
mr )

Thus the potential generated by the point-like source J(x) = c5(3)(X) indeed
represents the potential of a sphere of the radius I. Therefore the function

describes the source which is the sphere of the radius 1. Distribution of matter
inside this sphere is given by the function a(r2). The interaction potential
generated by a single point-like element of matter is the Yukawa potential.
Notice that the potential W(r) outside the sphere of the radius I coincides
with a potential of a point-like particle, while the interaction force is given by
the constant geff = gK(m2 F). The normalization condition K(m 2 /2) = 1 means
68 Chapter 2

that the coupling constant g in the interaction Lagrangian is chosen in such a


way that it just defines the potential of the source at large distances.
Investigations carried out up to now have shown that in the statical limit the
relativistic invariant distributions (generalized functions) K(x - x') =
K(F D)b(4)(x - x') used here give the correct description of extended. objects. As
one might expect, these distributions yield a reasonable starting
point for the relativistic invariant description of the nonlocal interactions.
Finally, we write the Coulomb and Yukawa potentials generated by the
form factor (1.16) corresponding to the Gaussian measure wo(bUl 2 ) in the
stochastic space-time [R4 (x). In this case, the change of these potentials at
small distances is given by the nonlocal propagators of the photon and scalar
particle with mass m (i.e., n-meson) in the static limit:

epe(r) = e(2n)-3 f d 3 p e- ipr exp( _p 2 l2j4)jp2 = (ej4nr)¢(r/l), (2.88)

and

= g exp -4-
8nr (m2F){ ¢ (ml2 + Ir) - ¢ (12ml - Ir) - 2 sh rm } (2.89)

respectively. Here
- 2 rx
(fi(x) = erf(x) = In Jo dt exp( _t 2 ).

Now we consider some peculiarity of the form factor V ( - p2 l2) in the


quantum electrodynamical case. In this case, the function (2.84) has the form

(2.90)

Substituting the expression (2.74) for K( _k 2 (2) into it, we get

W(r) = ~
4n
ff d3r1 d3r2 a(d)a((r 2 - r)2).
Ir1 - (r2 - r)1
(2.91)

This formula is the electrostatical energy of two interacting spherically-


symmetric charges whose centres are situated at distance r from each other. In
this case a charge distribution is described by the function a(r2).
The function W (r) at r = 0 represents the so-called proper electrostatical
energy of the electron in the classical field theory. If we assume that the
electron is point-like [this means that it should be a(r2) = b(r) in formula
(2.91)] then we obtain the well-known classical electrodynamical result ~ the
The Basic Problems of Nonlocal Quantum Field Theory 69

electron's proper energy increases as W(O) '" eZ11 for 1---> 0, where I is the
electron size.
However, the interaction energy or potential W(r) of two electrons is finite
at r = 0 if the function a(rZ) is some smooth distribution. As shown by Efimov
(1977a) the case of form-factors with the order of growth p = 1 means that the
distribution a(rZ) is bounded, i.e., all charge is distributed inside the sphere
with the given radius l. If the order of growth is p > 1 then it means that a(rZ)
differs from zero in the whole space, but that it decreases rapidly as
a(rZ)=exp[ -const (Irl/l)Y],

where
'Y = 2pl(2p - 1) > 1.
An interesting extremal problem was done by Efimov (1977a) namely to find
such form-factors V( - pZIZ) for which the potential W(r) of two rest electrons
has the smallest possible value at zero, i.e., at r = O. Mathematically, the
following problem has the solution: among entire functions V(z) of the order of
growth p = 1 and of the type rI, satisfying conditions:
V(O) = 1, V(x) = IK(xW and on the real half-axis 0 ~ x < 00,

there exists a function minimizing the functional

,u[V] = IX) du u-1/ZV(u) = const W(O). (2.92)

Comparing (2.92) with (2.90) and (2.91) it is easy to see that the consideration
of the functional represents proper energy of electron.
This problem pertains to the class of extremal problems in the theory of
entire functions (for details see Ibragimov, 1962) and has a unique solution. To
solve the given problem rewrite condition (2.92) in the form

(2.93)

The function K (t Z ) is an entire function of the first order of growth and of the
type 1rI. Therefore, in accordance with Paley-Wiener's (1934) theorem, the
following representation holds

1 f(1/Z
K(t Z) = - - du eiut g(u),
Fn -(1/Z

where g(U)E L z( -1rI, 1rI). Since K depends on t Z, then the function g(u) is even
70 Chapter 2,

and therefore

K(t 2 ) = -Jfi-; Jor a/ 2


du g(u) cos (ut).

Making use of Parseval's equality (Titchmarsh, 1939), from (2.93) we get

Joroo dtlK(t W = Jor


a/2
}l[V] = 2 dulg(uW. (2.94)

Moreover, we have

-Jfi-; Jor
a/2
K(O) = du g(u) = 1. (2.95)

Let there exist a system of orthonormalized polynomials {P n(x)} with


weight 1 on the interval [O,!u], i.e.,

Then the function g(u) can be decomposed by this system


00

g(u) = L cnPn(u).
n=O

In this case formulas (2.94) and (2.95) acquire the form

-J-;f2 Jor
ra/2 f2 fa
Jo
00 a/2
}l[V] = dulg(uW= n~0IcuI2, du g(u) = -J-;co-JI = 1.

It is clear that

It means g(u) = ~ and therefore

2
K(t 2 ) = -;;
r
Jo
a/2
ducos (ut) =
sin (ut/2)
(ut/2) .

Finally, we have

V(z) = [sin(uJz/2)/(uJz/2)Y (2.96)


Consider what charge distribution is described by the form-factor (2.96)
obtained. For this, according to (2.74) we calculate the density of charge
distribution a(r2) corresponding to the function

K(k 2 [2) = sin[(u/J2)Jk2i2]


(u /2)Jk2i2
The Basic Problems of Nonlocal Quantum Field Theory 71

After obvious calculation we have

a(r2) = _1_ fd 3 kK(k2/2)e ikr = ~ b(lrl - (JI/2).


(2n)3 4n r2
This distribution describes a uniformly charged sphere with the radius (JI/2.
Now we define the energy of two interacting electrons by formula (2.91). As a
result, we get

~~(2-!1)
W(r) = { 4n (JI

e2 1
(JI
r < (JI,

r > (JI.
4n Irl
Turning to the following Mellin representation

V(kiI2)=~f-/3-ioo d(~(O (kiI 2);, (2.97)


21 -/3+ioo sm n(
(0</3< 1)

we write finally the form of the form-factor (2.96):

V (_k2[2) = [sin Fk2f2 J2 (2.98)


S Fk2f2'
For the function v(O in the representation (2.97) we obtain
vs(O = 21 + 2; /r(2( + 3). (2.99)
If we assume that the electron is a uniformly charged ball of radius I, then, it
is easy to show that

2 2 _ 9 [sin<_k 2 / 2)1 /2 2 21 /2J2


Vb(-k I )-(_k2/2)2 (_k 2/2)1/2 -cos(-k I)

and vb(O is given by formula (1.15).


Thus, we see that in the quantum electrodynamics the form-factors with the
order of growth p = t have greatest interest, when the rest electron may be
interpreted as a charged sphere (or ball) with radius I.

2.4. The Causality Condition and Unitarity of the S-Matrix in Nonlocal


Quantum Field Theory

2.4.1. Introduction
It is well known that the principles of causality and unitarity of the S-matrix in
quantum field theory are the basis of all approaches in the elementary particle
72 Chapter 2

theory which make claims to self-consistency and physical acceptability.


Therefore, the proof of the unitarity and causality condition is crucial in
constructing various models of QFT. These problems were considered in detail
by Bogolubov and Efimov (and their coworkers) in both the local and
nonlocal cases, respectively. In this section we formulate only briefly the
causality condition and prove the unitarity of the non-local S-matrix in the
functional form within the framework of the Efimov (1977a) approach.
First we turn to the mathematical formulation of the causality condition. In
the local QFT the causality is manifested as a requirement imposed on the
Heisenberg fields cp(x): they must be locally commutable
[cp(x), cp(y)]_ = 0 for x ~ y (2.100)
or as the microcausality condition for the S-matrix

_b
bcp(x)
(~+)_
bcp(y/ - 0 for x ~ y.
< (2.101)

Both the conditions are of a formal mathematical nature and they represent
the postulates describing the mathematical structure of the fields cp(x) in
(2.100) and of the S-matrix in (2.101) rather than the requirements of physical
causality. This is explained by the fact that the concept of point-like nature of
events suggested in (2.100) and (2.101) is incompatible with the ideas of
relativistic quantum mechanics
In the recent past, numerous attempts have been made to formulate the so-
called physical condition of causality (Blokhintsev, 1973b; Wanders, 1959;
Eden and Landshoff, 1965: Stapp, 1965, etc.), i.e., to find the minimal require-
ments on the amplitudes of physical processes which would guarantee the
absence of any obviously noncausal phenomena in the macroworld. However,
this problem remains as yet open and conditions (2.100) and (2.101) are
actually the only working formulations of causality ensuring well-known
results in different approaches to the local QFT.
It should be noted that the predictions of the local theory are in satisfactory
agreement with the experiment up to the presently attainable distances
(10- 15_10- 16 cm). This appears to indicate that, first, the mathematical
condition of causality guarantees without doubt the fulfilment of the physical
condition provided one is able to give a reasonable formulation of the latter.
Secondly, the concept of the point-like nature of events is a good approxima-
tion up to presently attainable energies.
However, the non locality which is caused by the stochastic nature of space-
time as well as by the fact that the concept of strict localizability of events is
only approximate, has stimulated many attempts to formulate a self-consistent
theory at the expense of locality, and obviously of microcausality too. In doing
so, we may consider the theory valid only in so far as the violation of
microcausality is localized in sufficiently small space-time regions. The in-
vestigation of the causal S-matrix properties in theories of this kind is of
predominant importance.
The Basic Problems of Nonlocal Quantum Field Theory 73

2.4.2. The Causality Condition


Our formulation of the causality condition is based on the Bogolubov and
Shirkov (1980) approach. We introduce an operation of 'switching on' and
'switching off' the interaction. Instead of the coupling constant g, we introduce
the function g(x) on the interval [0, g], which characterizes the intensity of
'switching on' the interaction (see Section 2.2.5). Let 9 1 (x) be different from
zero in some region G1 C ~4 and g2(X) - in G2 C ~4. Then the S-matrix of
theory satisfies the microcausality condition, if
(2.102)
for G2 <: G1 , i.e., if and only if all the points of the region G2 belong to the
future cone (or to the space-like region) with respect to all points of the region
G1 . The condition (2.102) can be written in the differential form

R(x, y) =
b (bS
bg(x) bg(y/
+) = 0 (2.103)

for x ;:5 y. This relation represents a formulation of causality in the differential


form and it has been given by Bogolubov and Shirkov (1980).
We now verify the condition (2.103) for the S-matrix which is constructed by
means of the interaction Lagrangian ~n (x) in the case of the local QFT. As
usual, the S-matrix is presented as a functional series over the powers of the
function g(x):

S[g] = 1+ n~l:~ fd4X1'" fd4XnSn(Xl, ... ,Xn)g(Xl) .. ·g(Xn), (2.104)

where the operator-valued expressions Sn(x 1 , ... , x n ) are given by


Sn(x 1 , · · · , x n) = T[~n(Xl)'" ~n(xn)].
First of all, we write the expression (2.104) in some convenient form. The nth
term of the series can be represented in the form

Thus, by introducing T-exponent we can formally summarize series (2.104)

S[g] = T{I+n~l :~ [fd4X~n(X)g(X)T}

= T exp {i f d4 x~n (X)g(X)}. (2.105)


74 Chapter 2

In order to check condition (2.103) we calculate the variational differential of


S [g] at the point y

-i~~f~~ = T{~n(y)exp[ifd4X~n(X)g(X)]}'
taking into account here that bg(x)jbg(y) = b 4(x - y) (for detail, see Chapter
5). Further, the four-dimensional space 1R4 is divided into two parts G + and
G _ by the space-like surface Xo = const. = Yo with respect to which G + lies to
'the future' and G _ to 'the past'. Thus, we have

-i~~~~ = T{~n(y)exp[i L+ d4X~n(X)g(X)] X

x exp[i L_ d4Z~n(Z)g(Z)]}

= T{~n(y)exp[i L+ d4X~n(X)g(X)J} X

(2.106)

On the other hand, we reach by analogy

S[g] = T{ exp[i L+ d4 x5i;n(x)g(x) + i L_ d 4Z5i;n(Z)g(Z)]}

and also

S+[g] = {T[exp(iL d4 X5i;n(X)g(X))]}+ x

x {T[exp(iL+ d 4 z5i;n (Z)g(Z))]} +.

From this, taking into account the unitarity property for

and making use of (2.106), we get

-i~~f~~s+[g] = T{5i;n(y)ex p [iL+ d 4Z5i;n(Z)g(Z)]} x


The Basic Problems of Nonlocal Quantum Field Theory 75

Therefore, the product


[bS[g]jbg(y)]S+ [g]
does not depend on the behavior of the function g(x) in the region G _, i.e., for
Xo <Yo. In accordance with the covariance principle of the relativistic QFT it
takes place also in the case when x~y. We recall that in this region (x~y) the
commutator of the field operator cp(x) (in particular, for the scalar theory):
[cp(x),cp(y)]_ = .1(x - y),
where

.1(x - y) = .1(+l(x - y) + .1(-l(x - y)

= (2n)-3 fd4ke-ik(X-YlS(ko)b(k2 - m 2 )

A = (xo - YO)2 - (x _ y)2


disappears, which ensures the independence of events separated by space-like
intervals, i.e., the causality condition in space-time (or the locality condition of
the theory). In the non local case the proof of causality condition will be carried
out in Section 2.4.5.

°
Now we consider how the causality condition is formulated depending on
the choice of spaces of test functions. If y = in (2.103), then it is verifiable
that for
M(x) = <exIR(x,O)IP),
(where lex) and IP) are arbitrary physical states) the integral

f d 4 xM(x)fdx) = °
for any functions fG(x) which differ from zero only in a space-time region
G c ~4 lying outside the future cone: G ¢ V +, here
V+ = { x: Xo > 0, x 2>} °, V-+ = { x: Xo :;;;. 0, x 2 °.
:;;;.}
Thus we arrive at the definition.

Microcausality condition. If the space I5fJ of test functions satisfies the condition
of strict locality, i.e., that in the space I5fJ there is found some subspace D of test
76 Chapter 2

functions such that its intersection is dense in D (for detail, see Jaffe, 1966) or
in short that D n i5IJ is dense in D, it is said that the S-matrix satisfies the
microcausality condition, if

f d 4 xf(x)<o:IR(x,0)1f3) = ° (2.107)

for any f(x) E D(G) such that G n V + = 0.


Thus, in order to verify the causality condition (2.107), it is enough to have
in the space of test functions on which the theory under consideration is
constructed, some subs paces of the functions with finite supports. Moreover, the
fact of the presence of the functions with finite supports in the space of test
functions was adopted as a principle of strict locality in QFT (see Jaffe, 1966,
1967; Jost, 1965).
The main trouble with formulating and verifying the causality condition in
the (nonlocal) stochastic QFT presented above, is the fact that the functions
with finite supports are absent in the space i5IJ == Z of test functions. If the
theory is nonlocal but causal, then its S-matrix should satisfy condition (2.102)
only for G2 > G 1 . For G2 ~ G 1 (i.e., they are space-like with respect to each
other), the matrix elements
(2.108)
should decrease rapidly enough at the increasing distance between the regions
G1 and G2 .
Since in the (nonlocal) stochastic theory the space Z of test functions
consists of analytical functions among which the functions with finite supports
are absent, then it is necessary to use the following condition.
DEFINITION 2.1 (macrocausality condition). Let <o:IR(x,O)If3) E Z~ (p> 1)
and let f(x) E Z~ C Zp. Then the nonlocal S-matrix satisfies the macrocausality
condition, if

¢(x) = fd 4 x'f(x - x')<o:IR(x', 0)1f3) (2.109)

decreases when x --+ 00 outside the region V + as


¢(x) = O[exp( -llxII P )], (2.110)
where Ilxll = (x6 + X 2 )1/2 is the Euclidean distance.
Now we discuss the macrocausality condition by means of the concept of
the generalized functions. From the physical point of view, the functions of a
test space describe a space-time distribution of some physical quantities. These
test functions we call conditionally wave packets which are given in real space-
time, and the functions describing them are assumed to fall off rapidly outside
some bounded region in the space ~4.
The Basic Problems of Nonlocal Quantum Field Theory 77

The generalized function K(x - x') used as the form-factor has the meaning
of spreading-out operator connected with the definite dynamics of the process
under consideration. In this case we are interested in its space-time properties.
Therefore it is necessary to choose some function cp(x) supported by a
bounded region G c 1R 4 , and to see the support of the function (see Section
1.4.3 )

¢(x) = CPR(X) = (K,cp)(x) = fd4XIK(X - XI)cp(X ' )

= f dp(O L~l d 4 bcp(xo + ib4 C x + bO (2.111)

in 1R4. According to the basic idea of the nonlocal theory (see, for example,
Block, 1952; Chretien and Peierls, 1953) the macrocausality condition is
equivalent to the requirement that the 'effective wave packet' ¢(x) will be
located in some bounded region Gcp C 1R4 as soon as the 'initial wave packet'
cp(x) is located in a bounded region Gqt C 1R4. Since, in our spaces of test
functions, finite functions are absent, it is necessary to choose such a 'wave
packet' which is described by a function falling off rapidly enough outside
some bounded region Gqt C 1R4. The fact that for x, t -> 00 the causal signals
attenuate as Itl- 3 / 2 (the property of smooth solutions to the Klein-Gordon
equation), it is a common assumption in physics that the function falls off
rapidly enough if it falls off exponentially, i.e., it satisfies the limiting relation:
lim 1cp(x) 1exp(allxln = o. (2.112)
IIxll~cn

for some a > 0 and N > O. Then the macrocausality condition is equivalent to
the requirement that 'effective wave packet' ¢(x) would fall off rapidly enough
if the function cp(x) describing the 'initial wave packet' falls off rapidly enough
outside some region Gcp.
Let us consider the functional (2.111) in the case when the functions of a test
space belong to Za(1 < IX < 2/a, 0 < a < 2). Wave packets fall off rapidly when
Ilxll-> 00, and therefore it is natural to describe them by the functions of the
space Z~ C Za. If cp(x) E Z~, then

Icp(x)1 ~ ex p { -const Iltolxllla},


when IXIlI-> 00. Notice that no functions exist in the space Za which would fall
off more rapidly. However, such order of decay is quite acceptable from the
point of view of the physIcal requirements [see formula (2.112)].
Let cp(x) in (2.111) be the wave packet which is described by a function of
Z~. Let us write the functional (2.111) in the momentum representation

¢(x) = (K, cp)(x) = f d 4 pK(P2)q5(P) e ipx •


78 Chapter 2

Let the function <J>(p)E Z~:, where (x' = (X/((x - 1) > 2/(2 - a) and
K(p2)EZ2/(2_a)' Since (x' > 2/(2 - a), the entire function K(P2) is the multip-
lier for the space Z~:, i.e.,
cp(P) = K(P2)~(P)E Z~:.

Therefore, ¢(X)E Z~ and it satisfies condition (2.112). Thus the 'effective wave
packet' ¢(x) falls off rapidly, and the same is true for the 'initial wave packet'
q>(x).
Thus, if the S-matrix satisfies one of the causality conditions formulated
above, we say that causality is fulfilled in the theory under consideration. The
causality condition in nonlocal QFT was studied by Efimov and Alebastrov
(see Efimov, 1977a).
It is necessary to note that if the Schrodinger equation in the interaction
picture

(2.113)

were to have mathematical meaning, the causality condition would be auto-


matically valid as a property of solutions to that equation. This problem will
be investigated in Section 2.5.

2.4.3. The Scheme of Proof of Unitarity of the S-Matrix in


Perturbation Theory
Now we turn to the formulation of the unitarity in the nonlocal QFT. If the
Lagrangian of interacting quantum fields is given, the S-matrix is then sought
in the form (2.104) of a formal expansion in powers of the function of
'switching on' the interaction g(x), while the operator expression Sn(x 1 , ••• , x n)
in (2.104) can be represented in a series of normal products of the quantized
fields q>(x):
(2.114)

where the coefficient functions {KJ are in turn expressed in terms of the
products
(2.115)

of the nonlocal propagators

f
Dc (X - Y ) = I'-1(21t)-4 d 4 k
V(_k2F)
2
m -
k2 . e
-Ie
-ik(x-y)
.

In accordance with the above deduction (see Section 1.4 and Condition 2.1 of
Section 2.2.1) the nonlocal propagator Dc(x) exists as a generalized function on
The Basic Problems of Nonlocal Quantum Field Theory 79

some space Z of test functions and their products - the coefficient functions
{Kd in (2.114) are therefore not defined mathematically. Thus, the formulation
of a working method for the definition of the products of type (2.115) is the main
problem to be solved in the construction of the finite S-matrix in the nonlocal
theory. As usual, the coefficient functions {Kd in (2.114) are constructed as a
limit of the locally integrable functions Kb . . m ... (x 1 , ..• , xn) by introducing a
regularization procedure given 'by the parameter b, so there exists the limit

in the improper sense or in the weak limit (see Section 1.4 and Bogolubov and
Shirkov, 1980), or otherwise

S[g] = lim Sb[g]. (2.116)


b~O

It is unclear whether the S-matrix obtained in such a way satisfies the initial
axioms and especially the unitarity relation:

(2.117)

In the local theory, the unitarity and causality conditions are directly used
to formulate the subtractive method of regularization. This circumstance
ensures the fulfilment of unitarity of the S-matrix in each order of perturbation
theory, at least in the case of renormalizable interactions.
In the nonlocal theory the regularization procedure is formulated in such a
way that the limit (2.116) exists and the S[g]-matrix is unitary in each order of
the perturbation theory (Alebastrov and Efimov, 1973). We present here the
outline of their proof of the unitarity of the S-matrix in the perturbation
theory. Suppose that the S-matrix is known in the form of a functional
expansion (2.104), where the operator Sn(x i , ... , xn) are given by expansion of
type (2.114). If the S-matrix is finite and satisfies the axioms of QFT (see
Bogolubov and Shirkov, 1980), then the coefficient functions K ... m .••
(Xl' ... , xn) satisfy the following requirements: }
(1) They are translationally invariant, i.e.,
K • ••
m.j ' " (Xl + a, ... , Xn + a) = K • ••
m.J ' " (Xl' ... , Xn)·

(2) They are integrable on some space of sufficiently smoothly varying test
functions lilt, i.e., for any f(x l , .. . , Xn) E lilt there exists an integral

f d4XI ··· fd4XnK ... mj",(XI, ... ,Xn)f(Xl, ... ,Xn) < 00.

If the coefficient functions K . .. m .... (Xl, ... , Xn) are known, then the expan-
sion for S + [g] is known as well. Following Bogolubov and Shirkov (1980),
one can show that the coefficient functions obtained by multiplying two
80 Chapter 2

operator functions with different independent arguments


Sn(XI,""Xn) ® S';;(YI,···,Ym) (2.118)
are of the form
K ... mj'" (Xl"'" xn)K(_)(x - y)K ... m," .(YI,'" ,Ym), (2.119)
where the functions

S, t

may be defined as generalized integrable functions on the space 1lIL. The sign ®
denotes the transition to a normal product of field operators cp(x) in the
product (2.118), according to the Wick theorem.
So, if the S-matrix in the form of expansion (2.104) is known, then the
product
S[g]S+[g] = S[g] ® S+[g]
Df

is given in each order of the perturbation theory as a generalized operator-


valued function on the space of the test functions 1lIL.
The problem of the nonlocal theory is to prove that
S[g] ® S+[g] = 1.
We shall start from the method defining a finite S-matrix through use of an
improper limiting transition, and we construct our proof in the following
manner. It is obvious that the S-matrix will be unitary, if there exists a
regularization procedure which possesses the following. features:
(i) The regularized functions K~ .. m' •••
]
(x I , ... , xn) are continuous and boun-
ded,and
lim Kb .. m.... (xI, ... ,xn) = K ... mJ.... (xI, ... ,xn),
b __ O j

i.e., the regularized S"[gJ-matrix is defined and there exists the improper limit
lim S"[g] = S[g].
"~O

(ii) The Green positive-frequency functions which determine the multiplication


in (2.118) and (2.119) can also be regularized and there exists the limit
lim KZ-)(x - y) = K(_)(x - y),
"~O

or symbolically
lim®=®.
"~O

(iii) In the relation


The Basic Problems of Nonlocal Quantum Field Theory 81

J[gJ = lim lim lim SO, [gJ ® 0, S03+ [gJ


0 1 ~O o,~o 03~O

the limit is independent of the order of limiting transitions to the points


15 1 = 15 2 = 15 3 = 0, i.e., the operator
J01, 0" °3 [gJ = SOl [gJ ® 0, SO) + [gJ
is continuous at the point 15 1 = 15 2 = 15 3 = 0.
(iv) Regularization is chosen so that
SO[gJ ® ° SH [gJ = 1.
Then there is the following chain of equalities
J [g J = S [g J ® S + [g J = lim lim lim SOl [g J ® 0, S03 + [g J
01 ~O O,~o 03~O
= lim SO[gJ ® °SH [gJ = lim 1 = 1.
O~O o~o

Hence
S[gJS+ [gJ = 1.
The main idea is to indicate a regularization procedure in the nonlocal QFT
which would satisfy all the requirements listed above. This will prove the
unitarity of the S-matrix.
We give here only a regularization procedure and prove the unitarity of the
S-matrix in the functional form. Proof of the unitarity in the perturbation
theory was given by Alebastrov and Efimov (1973).

2.4.4. An Intermediate Regularization Scheme


In Section 2.2.2 we introduced the regularization procedure which allowed the
performance of the canonical quantization of the non local interaction (2.7).
However, for the construction of the perturbation theory by means of the
propagators (2.5):
i5c(k 2) = V( _k 2 12)(m 2 - k 2 - is)-l,
and for the proof of the unitarity of the S-matrix, it is convenient to use a
somewhat different regularization.
Let us consider the regularization method for the propagators (2.5). Let
V(-k212) be the nonlocal form-factor which satisfies Condition 2.1 (or some
conditions of this type), then the Mellin representation

(2.121)

holds in the region m2 > k 2 • The function v(O possesses the following
82 Chapter 2

properties:
(1) It is regular in the half-plane Re, ~ - /3, and in this region,
Iv(e + i'1)1 « c(l + l'1I)-N exp(nl'1I)/r(lel/p + 1)
for VN and some number C > O.
(2) At the points, = -1, - 2, ... , - [/3], it has zero (at least of the first
order).
(3) v(O) = 1.
(4) v*(O = v(O.
In particular, for the following functions:
V~ = exp[ -F(mZ - k Z )],
Vl = [sin l(m Z - kZ)l/Z /l(mZ _ kZ)l/Z]4, (2.122)
V z = 21l ·r(1 + p,)Jil(mZ - kZ).1/Z)[l(mZ - kZ)l/Zrll,
we have
v~(O = 1/r(1 + 0,
Vl (0 = (24~ + 5 - 2z~ + 3)/r (2' + 5), (2.123)
vz(O = 2- zll r(1 + p,)/r(1 + or(1 + p, + O·
It is easy to see that these functions satisfy the above-mentioned conditions.
We emphasize once more that the representation (2.121) holds for k Z < mZ
only. For the passage to the region k Z > mZ , one has to go from the
integration in the ,-plane over the contour
Lo = {,:, = e+ i'1, e= - /3, - (fJ < '1 < oo}
to that over the contour (see Figure 2).
1m ~

Re 5

Fig. 2. Integration contours Lo and Le'


The Basic Problems of Nonlocal Quantum Field Theory 83

Lo = {(: (= -j3+rexp[±iHn-e)],O~r<oo} (O~e~tn)

Let us introduce the regularized function

D~(k2)=~
21 10r d( ~m 12~ exp(8(2)(m 2 _
SIll n(
k2 _ i8)~-1
,

where ° ~ e ~ tn. One can readily recognize that if 8 > 0, the function D~(k2)
is:
(a) regular and analytic in all the complex k 2-plane, except for the cut along
the half-line [m 2 , + 00),
(b) D~(k2)=O(1/lk211+fJ) for Ik 2 1-->00,
(c) lim D~(k2) = Dc(k 2) = V( - k 212)(m 2 - k 2 - i8)-1.
/i~O

For the function D~(k2), there exists the Fourier transform

D~(x) =i- 1(2n)-4 f d 4k e- ikx D~(k2)

(2.124)

where

(2.125)
. m 2 (1 +0
= i2~e-'1t~ 8nr(1 _ 0 Hn~ (mJx 2 -i 8) (mJx 2 - i8)-1-1;

H~2) (x) is the Hankel function, for which the expansion

(-x2) vH(2)(X)
v
= --
1
isinnv
{oo
e i1tv
n~o
(-It
(x/2)2n
n! r(1 + n + v)
-

- ( -2)2V x (x)2n 1 }
n~o (-It -
00

x 2 n! r(1 +n- v)

{ e i1tV
=. .
1
-
(2/X)2v
+ O(X2) } , (2.126)
1 SIll nv r(1 + v) r(1 - v)
holds at the neighborhood of the point x=O, when Rev < 0. Using this
representation, we obtain
(2.127)
84 Chapter 2

Therefore the function D~(x) is bounded at the point x = ° because of

D~(O)=_n_._
(4 )-2f- P - iOO (0
d(-~- m
2(1+0
[21; exp (Je) < 00. (2.128)
2z _P + ico sInn( W + ()
From the representation (2.124) and (2.126), it follows that the function D~(x) is
[1$ - 1J times continuously differentiable with respect to x at the point x 2 = 0,
where [fiJ denotes the entire part of 1$.
Thus, if J > 0, then the function D~(x) is locally an integrable function and
the products of the type I1 i,j D~(Xi - x) are also locally integrable. In the sense
of generalized functions, limb~o D~(x) = D/x), i.e.,

f
!i~ d4xD~(x)f(x) = f d4xDc(x)f(x) = f
d 4pDc(p2)}(p)

for any f(x) E za


[IX < 2p/(2p - 1)].
Now we regularize the function D(±)(k) corresponding to the propagator
Dc(k 2) (2.5). For the positive- and negative-frequence Green functions Li(+)(k)
the following equalities hold:

=~
2i 10r d(~[2"
sin n(
x

(2.129)

where °-< 8 -< tn. Let us introduce the regularized functions

-b _
1 -P - ioo
D1+)(k)----:- f
2z _ P + ioo
(r)
v." 21; 2 - 2
d(-.-l exp(J( )Dm(Cm ,k),
sllln(
(2.130)

where

D(±)((, m 2, k) = i- 1 {(m 2 - k 2 - ie)l;-l - (m 2 - k 2 + iek o),,-l}


= 2n8(+k o)e(k 2 - m 2)n- 1 sin (n()(k2 - m 2)"+- 1. (2.131)

The function (k 2 - m 2 )r;+- 1 may be understood in the sense of the generalized


functions (see Gel'fand and Shilov, 1964). Then it has poles of the first order at
the point ( = 0, -1, - 2, ... of the complex (-plane. Besides, it is important
that
The Basic Problems of Nonlocal Quantum Field Theory 85

In the x-space, we obtain

Dt±)(x) = i- 1(2n)-4 f
d 4ke- ikx 15t±)(k)

(2.132)

where

The first term of (2.133) coincides with the causal function DJCm 2 ,x) con-
sidered in (2.124). This function is bounded at the point x = o. The second
term is also bounded at the point x = 0 because of the expansion for the Bessel
function

x
( 2)V J -vex) = ( ~ )-2V k~O
W (x/2fk (x)-V
(_1)k k! r(k + 1 _ v) = 2" [1 + O(X2)]. (2.134)

Thus, the functions D(±j (C m 2 , x) as well as De((, m 2 , x) are bounded at the


point x = 0 and
D(±j (C m 2, 0) = DiC m 2, 0) = (4n)-2m 2(1 +°/((1 + O.
Comparing the representations (2.132) with (2.124), it is easy to see that there
exist relations between the functions D:(x) and Dt±) (x):
D:(x) = 8(x o)Dt-)(x) + 8( -xo)Dt+)(x),
Dt-)(x) = 8(x o)D:(x) + 8( -xo)D:*(x). (2.135)
Moreover, the following relations
oVDt(x) = 8(xo)ovDt_)(x)+ 8(-xo)ovDt+)(x) (2.136)
hold, when v <;;: 2[f3J + 2 for 13 > [f3J and v <;;: 213 + 1 for 13 = [13],
In order to show equation (2.136), we consider the difference
RvCx) = o~Dt(x) - {8(xo) o~Dt _lex) + 8( - x o) o~Dt +) (x)}
86 Chapter 2

Here we have used the equalities (2.135) and the relation


1
aol 8( ± x o ) = ±"2bvl - 1(x o ).

When x 2 -> 0, we have, according to (2.133) and (2.134)

D~(x) = Dt-l(x) - Dt+l(x) = e(Xo)8(X 2 )(x 2 )P- 1 0(1).

Therefore, the function

for any f(X)E S, behaves as

Fh(xo) = e(xo)1 Xo 1 2P + 10(1).

when Xo -> 0, i.e., it is 2[13] + 1 times continuously differentiable at the


point Xo = 0, and when 13 > [13], all its differentials are equal to zero at the
point Xo = O. If 13 = [13], then only the 2f3-differentials of the function Fh(xo)
are equal to zero at Xo = O. This means that the functional equals

(Rv,f)= f f 3
dxo d xRv(xo, x)f(x o, x) = Lv

vI=1
(-1)
v +1
I
v!
I( _
VI' V
)1 X
VI'

for any V < 2[13] + 2 for 13 > [13] and v < 213 + 1 for 13 = [13] and any
f(x o , X)E s.

2.4.5. Proof of the Unitarity of the S-Matrix in a Functional Form


After the preliminary calculations carried out above, we verify the unitarity
of the nonlocal S-matrix presented in the functional form [see formula (2.105)
and Chapter 5]:

S",JCP] = Texp{i f d 4x g(x)5tin(X)}, (2.137)

where the function 5tin (x) = U(cp(x» can be represented in the form:

Here {uo, u 1 , . •• , Un" •• } are some sequence of real numbers giving the form of
the interaction. We assume that for the sequence of real numbers {un}, there
The Basic Problems of Nonlocal Quantum Field Theory 87

exists in the complex (-plane a measure a(() such that

Un = in f(n da(().

The function U(ep) can then be written in the form

U(ep) = f da(() ei~cp. (2.138)

Notice that the interaction Lagrangian is usually chosen in the normal


ordering form. To take this circumstance into account when the formal
function transformation is performed, it is necessary to pass to the usual
product of operators from the normal product. This means that instead of
(2.138), we must consider

U(ep(X)) = f da(() exp [DAO)(2/2] exp(i(ep(x)). (2.139)

Denote

The T -ordering operation is given by the formula

T = exp { 2.I If d 4 Xl d 4 X2 Dc(x l -


2
x 2) bep(x l b) bep(x 2) } .

The multiplying operator of two (normally-ordered) operator-valued functions


is
NB[ep] x NC[ep] = NE[ep],
where
E[ep] = B[ep] ® C[ep]

The Fourier transform Li~;) (k) of the function Ll~j-)(x) is equal to


Li~j-)(k) = 2nO(k o) b(k 2 - m2 ).
Further, we assume that the functions Dc(x) and D(_)(x) are bounded and
continuous, and DAO) = D:(O) = D(-iO) = D(+)(O) < 00, according to our re-
gularization scheme presented above. All these results are valid for the
regularized or averaged functions DJ(x) (j = c, -, + ) at b > O. For simplicity
we omit the index b for the D-functions in this section.
88 Chapter 2

We show that the S-matrix (2.137) satisfies the following relations


St2tl [ep] ® StltO[ep] = St2tO[ep],
(2.140)

T{'I'[ep(x)]S[ep]} = Sw,xo[ep] ® 'I'[ep(x)] ® Sxo, -w[ep], (2.141)

St2 t l[ep] ® St:tl[ep] = 1, (2.142)

b;X) (b:(Y) S[ep] ® s+[ep]) = 0, for Xo < Yo· (2.143)

We notice that relation (2.143), the so-called causality condition (see Section
2.4.2), is a consequence of the first three equalities. Indeed, let formulas
(2.140) - (2.142) be valid then we have

b
bep(y) S[ep] = ig(y)T {U [ep(y)]S[ep]}
f

= ig(y)S", ,yo [ep] ® U'[ep(y)] ® SyO' - '" [ep].


Further, making use of expressions (2.140) and (2.142), we get
b
bep(y) S[ep] ® S+[ep] = ig(y)S""yo[ep] ® U'[ep(y)] ® SYO,-",[ep] x

x ® S Yo,
+ _ 00 [m]
'f'
® S + ,Yo [m]
't' IX)

= ig(y)Soo,yo[ep] ® U'[ep(y)] ® S!,yo[ep].


Condition (2.143) is obviously fulfilled, since the obtained expression does not
depend on the field ep(x) at Xo < Yo'
Now consider relations (2.140). Owing to our definitions, we have

St2 tl[ep] ® StltO[ep] = ex p { ~ II d 4xl d 4 xz[Dc(Xl - Xz) bepl(Xl~:epl(XZ) +

+ Dc(Xl - Xz ) b
Z
bepz (x 1 )bepz (x z )
l} x

x exp{i 1: 2
d 4 xg(x)U(epl(X)) +

+i L d 4 X9(X)U(epz(x))} 1'1'1='1'2='1'
The Basic Problems of Nonlocal Quantum Field Theory 89

Since the defining region of the function qJ1(X) is the interval [t 1, t 2J and for
the function qJ2(X) it is the interval [to, t 1J, then first, one can put
qJ1(X) = rp2(X) = rp(x) and secondly, D(_j(x 1 - X2) = Dc(x 1 - X2) in the second
term of the first exponential, because of XlO > X20 for X10 E [tl' t 2J,
X20E [to, t1l Therefore we have

Relation (2.141) is proven simply; to this end it is sufficient to use the definition
of T and ®-operators.
Now we turn to relation (2.142) for the unitarity. We have

(2.144)

where
I; = T[b!brp;(x)] (i = 1,2),

E t2 rJrp1' qJ2J =exp {i f d 4 xg(x)[U(qJ1(x)) - U(qJ2(X))]}. (2.145)

We make use of the equality:

for

f(x o) = fd 3 xg(x o,x){U(rp1(x o, x)) - U(qJ2(XO, x))}.

Substituting this equality into (2.145) and using representation (2.139) for the
interaction Lagrangian, we obtain

St 2 dqJ J ® St~tl [qJ J = 1 + if f


d 4 X d()(OT 1 ® 12 T i{ exp[i'rp1 (x)] -

- exp[i'rpix)]}exp[,2Dc(O)/2J x

(2.146)
90 Chapter 2

One can easily obtain the following relations:

T1 exp[i(<p1(x)]Tl1 = exp { _(2tDc(0) +

+ i(<P1(X) + i( f d 4x'Dc(x - x') b<P~(X'J, T2 exp[i(<pz(x)]T;l

= exp { _t( 2D c(0) + i(<P2(X) +

+ i( f
d4X 'D:(X - x') b<P:(X')}' ® 12 ex p[i(<P1(X)] ® li

= eXP{i(<P1(X) + i( f
d 4x'D(_)(x - x') b<P:(X,J,®12 exp[i(<pz(x)] ® II

= eXP{i(<P2(X) + i( f
d 4x'D(+)(x - x') b<P~(X')}' T1 exp[i(<p2(x)]Tl1
= exp{i(<P2(X)},
T i exp[i(<p1 (x)] (Tn -1 = exp[i(<p1 (x)].
By means of these relations, one can write formula (2.146) in the form

St2t1 [<pJ ® S,;,! [<p] = 1 + i J:~ d x f dIJ(O{ exp[i(<p1(X)] exp [i( fd


4 4 x' X

x (Dc(X - x') ()<P~(X') + D(_)(x - x') b<P:(X'))]-

(2:147)

x exp[i( f d 4X'(D(+)(X - x') ()<P~(X') +


+ D:(x - x') ()<P:(X'))]} Et2tO [<P1' <P2J 1'1'1 ~ ~
'1'2 '1"

Now we note that the argument Xo - X~ < 0 for all the functions Dc(x), D:(x),
D(_)(x) and D(+)(x) in the curley brackets, since x~ belongs to the interval
[x o, t 2 ]. Therefore, according to (2.135), we have
The Basic Problems of Nonlocal Quantum Field Theory 91

Formula (2.147) takes the form

St2tl[CP] ® St;'l [cp] = 1 + i f d4 x f da((){exp [i(cp 1 (x)] - exp[i(CP2(X)]} x

f
x exp{i( d 4 X'[D c(X - x') bCP:(X') +

+D:(x - x') bCP:(X')]} x

X St2 X O[CPl] ®S,;xo[cp2]1q>1~q>2~q>.


On the right-hand side of this expression one can put CPl = CP2 = cP, since all
the operators depending on the variational derivatives stand to the right with
respect to the square brackets. Hence, the whole expression is equal to zero,
and finally, we have
(2.148)
Notice again that the presence of the derivatives in the interaction Lagrangain
does not change the correctness of the procedures undertaken. The fulfilment
of relations (2.136) guarantees that in this case the unitarity is fulfilled.

2.5. The Schrodinger Equation in Quantum Field Theory with Nonlocal


Interactions

2.5.1. Introduction
We have already noted above that our scheme of construction of QFT in the
stochastic space-time (1R4(X) with the measure w(bUZZ) lead to the quantum
field theory with nonlocal interactions. In this case, both the fields averaged in
1R4(X) and the interaction Lagrangian constructed by means of these fields
became nonlocal quantities. Now our main proposal is that the Schr6dinger
equation, describing the dynamics of physical systems, must be modified
according to the idea of nonlocality in nature resulting from the stochasticity
of space-time at small distances. The present section is devoted to the
investigation of this problem.
Thus, as shown above, the construction of a self-consistent theory of
nonlocal interactions of quantized fields became possible owing to the follow-
ing two ideas: First, the form-factors must be entire analytic functions in the
momentum space, and must decrease rapidly enough in the Euclidean space.
Second, the form-factor must be quantized, i.e., it is necessary to introduce
supplementary degrees of freedom, which determine the regularization, in
order to facilitate the transition to the Euclidean metric, and the re-
92 Chapter 2

establishment of the form-factor in the limit of the cancelled regularization.


The development of these ideas allowed the construction of the finite and
unitary S-matrix for arbitrary interaction Lagrangians in each perturbation
order (see preceding sections and Efimov, 1977a).
What is incomplete in this construction? The following problem arose from
investigations into causality conditions. The coefficient functions of the S-
matrix in the configuration space turn out to be analytical functions. Analytic
methods used in the study of local properties of analytical functionals make it
impossible, in principle, to determine a space localization of studied func-
tionals within an accuracy of a certain distance given by the nonlocality
(Fainberg and Soloviev, 1978; Soloviev, 1980). Obviously, the use of non-
analytical methods is needed. However, these methods are not developed. The
results obtained by the projecting sequences of functions (Efimov, 1977a) give
rise to doubts because, as is shown in the paper of Fainberg and Soloviev
(1977), there are examples of explicitly nonlocal functionals which are, as local
ones, characterized by the projecting sequences of functions. Therefore, the
existence of the microcausality condition, understood as a strict equality of the
corresponding functional outside the light cone, remains an open problem in
the theory of non local interactions.
On the other hand, causality is nothing other than the correctness of the
Cauchy problem of Schrodinger quantum-field equation (or the equation of
Tomanaga and Schwinger). However, the utilization of the regularization
procedure in the construction of the S-matrix both in the local and nonlocal
theories is reduced in that the S-matrix is not a solution of the corresponding
equation, and is determined by a series of limits. It seems therefore that the
natural properties of the Schrodinger equation solutions defined as unitarity
and causality are to be proved separately.
Difficulties in nonlocal theory arise usually when a nonlocal form-factor is
introduced into the interaction Lagrangian, but the Schrodinger field equation
(or Tomonaga-Schwinger equation) remains local

fJ
i fJu(x) ,¥[u(x)] = Ytf'n(x),¥[u(x)], (2.149)

where

The integrability conditions are violated within this approach and numerous
other difficulties arise (see Markov, 1958, for example).
We assume that while introducing nonlocality in the interaction
Hamiltonian the equation of Tomonaga and Schwinger must be treated also
non locally but with retardation. In this way, in our opinion, one can get rid of
difficulties caused by the integrability conditions in the non local theory.
The Basic Problems of Nonlocal Quantum Field Theory 93

In this section we shall show that the S-matrix describing non local in-
teractions of quantized fields solves the Cauchy problem of the evolution
equation (or SchrOdinger equation in the interaction picture at imaginary time,
i.e., in the Euclidean metric) with retardation. In this way supplementary
degrees of freedom with respect to the Fock space of physical particles need
not be introduced.
Strictly speaking, the formulation of such an equation with the correctly
stated Cauchy problem at imaginary time does not answer directly the question
about the causality condition of the S-matrix in Minkowski space. However, a
simple analytical connection between the S-matrices both in Euclidean and
Minkowski spaces means without doubt that causality of the evolution equation
must ensure the absence of any physically observable noncausal phenomena.

2.5.2. The Field Operator at Imaginary Time


We shall consider the theory of a one component scalar field cp(x) describing
particles with mass m. The field operator cp(x) may be written in a standard
way (see Bogolubov and Shirkov, 1980, for example)

(2.150)

where w = (m 2 + k 2 )1/2. Creation at and annihilation ~ Boson operators


satisfy ordinary commutation rules:
(2.151)
We assume that there exists a single vacuum state \1'0 = 10) which obeys the
conditions:
(010) = 1, a.. 10) = 0 Vk. (2.152)
State vectors of scalar particles are represented by rays in the Fock space
which is, as usual, constructed over the basis

(2.153)

where n = 0, 1,2, ...


Now we pass to imaginary time t -> - iT, or to the Euclidean metric. In the
constructive quantum field theory the physical Hilbert space F of a free field in
Minkowski space is considered as a subspace of the Hilbert space N of a free
Euclidean field (see Simon, 1974). Especially, in order to obtain Euclidean
Green functions and to construct the scattering matrix (Petrina et al., 1979), in
creation and annihilation operators at and ~ the supplementary degree of
freedom ak+ -> ak~. (~ -> ~,.) connected with imaginary time is introduced so
94 Chapter 2

that commutation rules are of the form

(2.154)

where kE = (e, k).


Here we shall not enlarge the number of degrees of freedom of a scalar field
and we shall construct a space of our Euclidean states over the same basis
(2.153).
So, we introduce the free field <!J(xE ), where X E = (I:, x) in the Euclidean
space, i.e., at imaginary time by the replacement t ~ -ir in expression (2.150)
for the field operator <p (x, t):

(2.155)

Let us introduce the T-product operation, i.e., the imaginary time r ordering
operation. In representation (2.155) the parameter r is introduced explicitly, so
the T-product operation is defined straightforwardly:

(2.156)

Further, we shall define the two-point Euclidean Green function which will be
called causal

(2.157)

where kE = (e,k), ki = e 2 + k 2 and K 1 (z) is the Mac'Donald function. The


obtained function de (xE ) represents an analytical continuation to t ~ - ir of
the causal Green function in Minkowski space:

Now let us consider the commutator of the fields <!J(xE ) and the function
The Basic Problems of Nonlocal Quantum Field Theory 95

~(-)(X1E - X2E ) = <01¢(X1E )¢(X2E )IO) (2.158)

= (211:)-3 f~:: e ik (x l -x2)-w(rl-r2).

We see easily that integrals on the right-hand side of formulas (2.158) do not
exist under an arbitrary choice of the difference ('1 - '2). Therefore, only the
T-product of operators ¢(xE ) has a reasonable mathematical meaning.
Let us introduce operator R [¢] of the following type:

R[¢] = n~o :,fd4X1E··· fd4XnERn(X1E, ... ,XnE)T[¢(X1E)···¢(XnEJ. (2.159)

Operators R [¢] are defined by a set of functions {Rn(X 1E , ... , x nE )} the


properties of which will be elaborated below.
Let us define the operation of conjugation
(2.160)
Then for the operator R in (2.159) we obtain the following expression

R*[¢] = n~o :,fd4X1E··jd4XnER:(X1E, ... ,XnE)X


(2.161)
N ow we determine the operation of multiplication of two operators R 1 [¢] and
R2 [¢] of type (2.159) by definition:

R 1 [¢]R 2 [¢] = T{R 1 [¢]R 2 [¢]} = L _1_ fd4X1E··· fd 4 Xn E X


Dr n I· n2 n 1 'n 2 ' 1

(2.162)

2.5.3. The State Space at Imaginary Time


Let us define the state space N of the system as a set of vectors of the type
'I' = R[¢]IO) (2.163)
where R has the form (2.159). It will be assumed that the state (2.163) is given
96 Chapter 2

if all the functions Rn(X 1E , ... , x nE ) determining the operator R [<fJ] in (2.159)
are known.
Notice that any state in the Fock space F must be represented as shown in
(2.163) in the case of the corresponding choice of the set of functions
{Rn(X 1E , ... , x nE )} in the operator R[ <fJ] since there exists a simple linear con-
nection between sets of basic vectors:

10) 10)

<1 0 ) <fJ(x1E)10)
a+a+IO)
k, k2
~ T[ <fJ (X 1E )<fJ(X 2E )] 10) (2.164)
+ T[<fJ(X 1E ) <fJ(X 3E )] 10)
ak , ak~ 10)
uk \ <1 0 ) T[<fJ(X 1E ) <fJ (xnE)]I 0)

The appropriate formula of this correspondence can be easily obtained, if


necessary (Appendix A).
The most important here is the following: the Fock space F consists of
vectors of the type

'PF ="~ ~n! fd k fd knJ,,(k


3 1 •·· 3 1 ,··· ,kn ) ~~+ ... a + 10),
y'n! 1
k
n
(2.165)

but the space N of vectors

'P = L~
n.
n
fd 4 X1E ... fd 4 x nE Rn(X 1E ,···, xnE)T[<fJ(X 1E )··· <fJ(xnE )] 10). (2.166)

Obviously, from representation (2.165) and (2.166) it follows that the space F is
a subspace of N, since some set of mutually different vectors [different sets of
functions Rn( ... )] from N corresponds to each vector from F.
In the scattering problem the initial state of the type (2.165) is given at some
'0 = - T, where T -> 00. In this case such a state can be written in the form

(2.l67)

where rn(x 1 , ... ,xn) is connected withJ,,(k 1 , ... , k n) in (2.165). The state (2.167)
can also be rewritten in the form (2.166), where
n

Rn(X 1E ,···, x nE ) = rn(x 1 ,···, xn) TI1 b('j -


j =
'0)·
The Basic Problems of Nonlocal Quantum Field Theory 97

The norm of state vectors from N is given by definition:

II'P 112 = ('P, 'P) = <OIR * [¢ ]R[¢] 10)

(2.168)

It is known [see formula (7.38) of Chapter 7 and Simon, 1974, and Efimov,
1979] that for a free Euclidean field there exists such a Gaussian positive
measure d/l J that

(2.169)

Then the Euclidean norm (2.168) is

11'P112 = fd/lJIR[f]12 < 00, (2.170)

where the functional R [f] is given by

Therefore, functions Rn(X 1E , •.• , x nE ) must be such that the norm (2.170) is
finite.
So, we have constructed the state space N, which includes all vectors of type
(2.163) for which the norm (2.168) [or (2.1 70)] is finite. Furthermore, the
operators R[¢] (2.159) have been introduced for which the operations of
conjugation (2.161) and multiplication (2.162) are defined.

2.5.4. The Interaction Hamiltonian and the Evolution Equation


The dynamics of a quantum field system is described by the Schrodinger
equation. In the interaction picture it has the form

(2.171)

with some initial condition


(2.172)
98 Chapter 2

Passing to imaginary time t ~ - ir we obtain

(2.173)

It is customary to call equations (2.173) evolution equations.


In quantum field theory the local interaction Hamiltonian are usually of the
form of polynomials in the field operators. For example, for the self-interacting
scalar field we have '

(2.174)

or in the imaginary time formulation

(2.175)

If we look for a solution of equation (2.173) with the Hamiltonian (2.175) in


the form (2.163), then we have the standard problem of the local quantum field
theory with all its difficulties.
Now we consider the quantum field theory with the non local interaction.
We introduce the extended-nonlocal field (see Chapter 1)

(2.176)

where YE = (Y4'Y), Y~ = Y~ + yl and w(yD is a real function-measure, the


properties of which have been discussed in Chapter 1.
Let us calculate the causal Green function of the extended field <Pn (xE ) in the
Euclidean space. According to formula (1.13) we have

Dc(X1E - X 2E ) = (2n)-4 fd 4P E e-iPE(X1E-X2E)[K( -pD]l(ml + p~)-\ (2.177)

where

(2.178)

Here for the sake of simplicity we omit the index I for the functions K (.) and
w(·). As usually, we shall assume that the function w(y~) is chosen so that the
K( - pD is an entire analytic function in the complex p2-plane and that it
decreases so rapidly as p~ ~ 00 that

(2.179)
The Basie Problems of Nonloeal Quantum Field Theory 99

Instead of the interaction (2.175) we now write

Hin[r,¢] = g fd3XT[¢~(r,x)]

= g fd 3X ill fd4YjEw(Yji)TL61 ¢(r + r.,x + Yn)}. (2.180)

Therefore, the interaction Hamiltonian (2.180) belongs to the class of oper-


ators R[¢] (2.159). Notice that the interaction Hamiltonian may be chosen in
the normal form:

Hin[r,¢] = g Jd x:T[¢~(r,x)]:
3

= g f d3xT[¢~(r,x) - 6¢;(r,x)Dc(0) + 3D; (0)], (2.181)

where Dc(O) is given by (2.179). In this case


<01: Hin[r, ¢]: 10) = O.
Using the representation of the state vector 'I'(r) (2.163) we shall write the
evolution equation (2.173) for the non local interaction (2.180) or (2.181) in the
following form:
8
8rR[r,¢]10) = -T{Hin[r,¢]R[r,¢]}IO). (2.182)

The operation of multiplication on the right-hand side of (2.182) is defined


by (2.162). The initial condition for equation (2.182) is
(2.183)
So, the obtained evolution equation is retarded, since the 'time' r in the
nonlocal Hamiltonian (2.180) or (2.181) may precede the times in field oper-
ators R[r,¢J. However, the T-ordering operation in equation (2.182) arranges
the times appropriately.
Equation (2.182) with the initial condition (2.183) may be rewritten as
8
8r R [r,¢] = -T{Hin[r,¢]R[r,¢]}, (2.184)

(2.185)
Notice that if the solution (2.184) is written in the form R(r, ro), it does not
satisfy the condition
R(r,r o) = R(r,r 1 )R(r b r O ) (ro < r 1 < r),
100 Chapter 2

where multiplication is understood in the ordinary sense. This is an immediate


consequence of the non locality of the theory.
Nevertheless, the Cauchy problem for equation (2.184) can be formulated. If
we introduce the space cut-offs g -* g(x) in order to eliminate the difficulties
caused by the Haag theorem, the Hamiltonian H in [r, ¢] represents an operator
in the considered space N. We have

Hin[r,¢] = fd3Xg(X): T{¢~(r,x)}. (2.186)

Considering the norm of the state

we obtain

x 6D;(Xl - x 2 ) ~ 6D;(0{fd3xg(X)T < 00.


Therefore, the ultraviolet catastrophe is absent in the considered nonlocal
theory.
Let us turn now to the solution of equation (2.184) with the initial condition

t
(2.185). The solution is given by

R[r,¢] = T{ex p [ - dr'Hin[r',¢] ]Ro[¢]}. (2.187)

For the state vector we obtain


'I'(r) = R[r,¢]IO). (2.188)

t
The norm of state 'I' (r) (2.188) equals

11'I'(r)112 = f d l1 fex P{ -2 dr'Hin[r',f] }R o[f]1 2 • (2.189)

If the Hamiltonian H in [r,f] in (2.189) is chosen in the form (2.186) then


according to (2.181) we have

Hin[r,fJ ~ -6D;(0) fd3xg(X).


For the norm (2.189) we have

11'1' (r) II ~ 11'1'0 II exp { 6(r - f


ro)D;(O) d 3xg(x) }- (2.190)

Thus, we see that the solution exists in the state space N. The uniqueness of
The Basic Problems of Nonlocal Quantum Field Theory 101

the obtained solution follows from (2.187), which produces the S-matrix that
coincides completely with the S-matrix constructed and investigated in the
preceding sections.

2.5.5. Appendix A
Before obtaining formulas (2.164) [or (2.165) and (2.166)J we give some
definitions.
DEFINITION. Functions R n(x 1 , ••• , xn) (n = 1,2, ... ) belong to a class G(n) if
their Fourier transforms
Rn(k 1 , ... , k n) = F.(kl'· .. , k n)gn(k 10 , . .. , kno ) satisfy the following con-
ditions (E):
(1) F n(k 1 , ... , k n ) = j,,(k 1, ... , k n ) fI ~,
i= 1
Wi = (m 2 + k;)1/2,

(2) gn(W 1" " , W n ) = 1,


(3) g; (k 10 , ... , kno )g(k 10 , ... , k no ) = 1,

(4) Rn(x 10 , ... , XnO) = f dk 10 ... f dkno exp {i j tl kjOXjO }gn(k 10 , ... , knO )

is a regular function with respect to all variables X iO •

If 'I' is the state vector of the single particle system, then among the
components R n(x 1 , ••. , xn) in the Fock space only the amplitude R 1 (x) differs
from zero. Thus

and

'I'i = fd4XRt(X)<01 Tcp(x).


Df

Let R 1(x)E G(l), for example, be the function

(A.1)

f3 > 0, fl (k) E L 2 ( ~ 3), satisfies all conditions (E). Taking into account that
cp(x) = cp+(x) + cp-(x), where

cp±(x) = r
Jko > 0
d 4kexp(±ikx)b(k2 - m 2 )cP(±k),
102 Chapter 2

and according to (A.1) we have

'P 1 = I d 4xR 1(x)<p(x)10) = I d 3 kfl(k)a*(k)10)

here
a*(k) = a+(k, k o)(2wk ) -1/21 ko : 'nt,
The norm of the vector 'P 1 is given by

II 'P 1112 = 'P t ® 'P 1 = IId 4 x d 4yR T(x)R 1(y)<OI T[<p(x)<p(y)]lO).

In the momentum representation this formula takes the form

II 'P 1 112 = 2 Id 3 k lf (k)12


1
·-1
~ 1 0 m2
fro dk g*(ko)g(k o)
+ k2 _ k 2 _ is·
(A.2)
- 00 0

According to conditions (E) the last integral in (A.2) has the form

·-1 fOO dk g*(ko}g(k o} - ·-1 Idk ( 2 _ k 2 _ . )-1 - ~


I
- 00
0
m
2
+ k2 - k2
0 -
. -
IS
I 0 Wk 0 IS -
~

and therefore

11'P1112 = (2n) I d 3 klfl(kW.

Similarly, the two-particle state is given by

'P 2 = fi {II d 4xl d4X2R2(Xl,X2)T[<p(Xl)<P(X2)] - C 2 }IO) == R 210),


where the Fourier transform of the function R 2 (xl> x 2 ) C G(2) has the form
Rdkl' k 2) = f2(kl' k 2)(2wkl ) -1/2(2wk2 ) -1/2 g(kl0)g(k20}
and the constant C 2 is chosen as follows:

C 2 = i-I II d 4xl d 4x 2R 2(Xl> x 2) Ac(XI - x 2) = I d 3 pf 2(P, -p)WpQ,

here

Q = i-I I~oo dPog(Po)g( -Po)(w; - p~ - is)-I.

The choice of the function g(po) being such that the value IQI < 00, in
particular, for g(k o ) = exp[ -it(ko - w)] (t is some constant parameter), one
can easily verify that
Q = exp(2it wp)(n/wp).
The Basic Problems of Nonlocal Quantum Field Theory 103

Then

'1'2 = ft II d 3 k I d 3 k 2 f2(k l , k 2)a*(k l )a*(k 2)10)

and the norm of the vector 'I' 2 is defined as


11'1'2112 = <OIRi ® R210), (A.3)
where

Ri = {II d 4 xI dx 2Ri(x
4
l , x 2)T[q>(X I )q>(X 2)] - Ci }2- I/2
and

Ci = i-I II d 4 xI d 4 x 2R i(x l , x 2) Llc(XI - x 2)·

The symbol ® denotes the following operation


T[q>(XI)··· q>(xj )] ® T[q>(yd··· q>(Yi)] == T[q>(x I )··· q>(Xj)q>(YI)··· q>(yJ]
Then, the norm of the vector '1'2 defined by the equality (A.3) becomes the
following

11'1'2112 = i- 2 II d 4xI d 4x2 II d4X3 d 4 x 4 Ri(x l , x 2)R 2(X 3 , x 4 ) x

X Llc(XI - x 3 ) Llc(X2 - x 4 ) = (2n) -2 II d 3 k I d 3 k2 If(k l , k2W.

Similar calculations can be carried out for an arbitrary N-particle


(N = 3,4, ... ) state vector of the system. Thus, any normalized state of the
system in the Fock space F can be represented in the form (2.163) by means of
the operator R[ ¢] if its coefficient functions belong to the class G(n).
Chapter 3

Electromagnetic Interactions In
Stochastic Space-Time

3.1. Introduction

Quantum electrodynamics (QED) is a beautiful theory describing the dy-


namics of photon and fermion interactions (electrons, muons etc.) with en-
ormous accuracy in a very wide range of physical quantities (describing for
example phenomena occurring in a region extending from very large length
scales (5 x 10 1o em) to those of the microworld (10- 15_10- 16 em)). Up to
now no effect has been found experimentally which could not be confined
within the local QED.
The tremendous quantitative success of QED yields a fundamental criterion
for the construction of new physical models based on different assumptions
and hypothesis concerning the space-time structure at small distances (partic-
ularly, a possible violation of locality in QED and the existence of fundamen-
tal length).
Tests of locality are usually performed by the modification of the particle
propagators and vertex functions. In particular, tests of QED are carried out
in terms of the modified electron and photon propagators (see Kraus, 1975;
Magg et al., 1972; and Ringhofer and Salecker, 1980). On the other hand, it is
well-known that the introduction of nonlocality into the theory leads to a
change of particle propagators (Chapters 1 and 2). As shown above, however,
this change is not arbitrary and is determined by fundamental principles of
QFT: Lorentz covariance, finiteness, unitarity, causality and gauge invariance.
Roughly speaking, the aim of the nonlocal theory is to find restrictions on the
choice of a form for non local particle propagators (or the so-called form-
factors of the theory).
Thus, an analysis of experimental data for testing locality must be perform-
ed within a theory which satisfies the basic principles of QFT. The problem
of constructing the non local theory for the scalar field satisfying the above-
mentioned basic principles has been solved successfully (see Chapter 2 and
Efimov, 1977a).
According to the method presented in Chapter 1, both neutral and charged
fields are extended in our scheme over the large scale of space-time. The

104
Electromagnetic Interactions in Stochastic Space- Time 105

propagators of these extended-nonlocal fields are given by formulas of the type


(1.13). However the change of charged particle propagators (in particular, in
the momentum representation)

S(p} = SR(P) = V( - p2[2)S(p)

essentially complicates the proof of the gauge in variance of the theory.


This chapter is devoted to the construction of gauge-invariant QED of
particles with spin 0, t and 1 in the stochastic space-time ~4(X) considered
above. We introduce here a definite algebraical prescription which ensures the
gauge invariance of QED obtained in the perturbation theory. Further, the
electromagnetic processes are investigated and contributions to the anomalous
magnetic moments of leptons and to the Lamb shift due to nonlocality (or
stochasticity) are estimated. Here the considerations are mainly concerned
with the low-energy processes. Notice that at very high energies testing of
locality of the theory may be difficult because of interference effects between the
electromagnetic and weak interactions. For example, in the standard model of
electroweak interactions, the testing of QED will be disturbed by interference
from the weak effects due to ZO bosons.

3.2. Gauge Invariance of the Theory and Generalization of Kroll's Procedure

The concept of the gauge in variance in QED is crucial, within which the
interaction of fermion fields with photons is formed. Moreover, modern theory
(including unified theories of all type forces) of elementary particle interactions
is constructed in analogy with QED. A particularly striking example is the
electroweak theory of electromagnetic and weak interactions due to A. Salam
(1968, 1980); S. Weinberg (1967, 1980); and Sh. Glashow (1961, 1980).
We start this section with a short discussion concerning the method of
introducing the interaction Lagrangian between charged particles and pho-
tons. As an example, consider the field of electrons t{!(x). The free Lagrangian
of this field has the standard form

(3.1)

where

Yv = (Yo, y) are the Dirac y-matrices and m is the mass of the electron.
Lagrangian 5t'o(x) is invariant with respect to the global gauge transformation

t{!(x) -+ t{!'(x) = eiAt{!(x). (3.2)

Here A is an arbitrary real constant. It is obvious that the Lagrangian (3.1) is


106 Chapter 3

not invariant with respect to the local gauge transformation


l/J(x) -> l/J'(x) = U(x)l/J(x), (3.3)
where U(x) = exp[iA(x)] and A(X) is an arbitrary real function of x. this is due
to the fact that the derivative avl/J(x) is not transformed under (3.3) as the field
l/J(x) itself. Indeed, we have
avl/J'(x) = U(x)(a v + i avA(X))l/J(x).
It is well known that the local gauge invariance (3.3) can be maintained,
provided the interaction of the field l/J(x) with the electromagnetic field Av is
introduced. Consider the quantity (a v - ieAJl/J (eis the electron charge). We have
(av - ieAv(x))l/J(x) = U- 1 (x)(av - ieA~(x))l/J'(x), (3.4)
where

A~(x) = Av(x) - ~avA(X).


e
(3.5)

From (3.4) it is obvious that the Lagrangian which follows from (3.1) by the
substitution
avl/J-> (a v - ieAv)l/J (3.6)
is invariant with respect to the gauge transformations (3.3) and (3.5). To
construct the complete Lagrangian of the system under consideration, we have
to add the gauge invariant Lagrangian of the electromagnetic field. The tensor
of the electromagnetic field is given as
F/l v = a/lAv - avAil" (3.7)
Clearly
F~v = F/lv·
Consequently, the gauge invariant Lagrangian of the fields of electrons and
photons takes the form

(3.8)

The substitution of the derivative avl/J by the covariant derivative (a v - ieAv)l/J


in the free Lagrangian of the field l/J(x) leads to the following interaction
Lagrangian of electrons and photons:
5!;n(x) = ejvAv, (3.9)

where jv = lfI'Yvl/J is the electromagnetic current. Thus, the substitution (3.6)


fixes uniquely the form of the interaction Lagrangian. Such an interaction is
called a minimal electromagnetic interaction. In this connection, we notice that
the principle of gauge invariance alone does not fix the interaction Lagrangian
Electromagnetic Interactions in Stochastic Space-Time 107

uniquely. For example, the addition of the Pauli term f.1" Ilia I'vt/J F I'v to the
Lagrangian (3.8) does not mar the gauge in variance of the theory. Here
a I'v = !(1/i)(y I' Yv - Yv Y1') and 11 is the anomalous magnetic moment of the
electron (see Section 3.4).
All available experimental data confirm that the Lagrangian (3.9) is, indeed,
the true Lagrangian which governs the interactions of electrons and photons
in the local theory. It is also well known that electrodynamics with its minimal
interaction (3.9) is the renormalizable theory.
The application of the above-mentioned idea to the non-Abelian Yang-Mills
theory based on the gauge SU(2) x U(1) invariance has been made in the
construction of the unified theory of electromagnetic and weak interactions.
The exposition of this theory transcends the bounds of this book and its
fundamentals may be found in any monographs and textbooks devoted to
recent developments in quantum field theory (see, for example, Lai, ed., 1983;
Chaichian, and Nelipa, 1984).
In the language of the perturbation theory (or the Feynman diagrammatical
techniques) the gauge in variance of QED means that every matrix elements of
the S-matrix defining the concrete electromagnetic processes have a definite
structure, and algebraical relations exist between them. For example, in the
momentum representation, the so-called vacuum polarization diagram in the
second order of the perturbation theory, the following form (see below) exists:

(3.10)

as well as the following relation between the vertex function ['I'(p, q) and the
self-energy of the electron L(p):

(3.11 )

The latter relation is called the Ward-Takahashi identity. The explicit forms of
these functions may be written (for detail, see Bogolubov and Shirkov, 1980):

(3.12)

and

['I'(p, q) = (~:~4 f d 4
kl1((p - k)2)Y v S(q + k)YI'S(k)y" (3.13)

and the identity (3.11) proved where S(p) = (m - p)-l, l1(k 2 ) = (_k2 - ie)-l.
F or the proof of relation (3.11) consider the identity

as(p)
a = S( ')
p Yl' S( p,
') (3.14)
PI'
108 Chapter 3

where the vertex YJl is given by

YJl = - OPJl
o S-l( A)
p. (3.15)

Further, differentiating !(p) over PJl and making use of the identity (3.14), we
have

(3.16)

Choosing other momentum variables in (3.13) and assuming q = 0,


pi = P + q = p, we get

Comparing this expression with (3.16) we obtain the Ward-Takahashi identity


(3.11). The relations of the type (see Section 4.2):
qJl fJl(p, q)l pl 2 = p2 = m2 = 0

follows from the following identity


qJl{S(pdy Jl S(P2)} = S(Pl) - S(P2)
ifq=Pl-P2·
It is important to notice that the vertex YJl in any Feynman diagram is
connected with the propagator S(p) of the charged particle by relation (3.15). It
implies that if the particle propagator S(p) is changed, it should be to modify
the vertex YJl ~ uJl(·) at the same time in accordance with formula (3.15).
Now we discuss the gauge invariance of QED for the case of the nonlocal
theory. Thus, the hypothesis of the stochastic space-time leads to a change of
propagators of both neutral and charged particles. It is well known that any
modification of the local causal Green functions of charged particles results in
a violation of some algebraical relations, as mentioned above, and the
fulfilment of which grants the gauge in variance of the theory. There are
numerous papers (Kroll, 1966; Kraus, 1975; Magg et ai., 1972; and Ringhofer
and Salecker, 1980) devoted to this problem. Among these, Kroll's work plays
an important role. We generalize here his procedure, the essence of which
consists of the following (Dineykhan and Namsrai, 1977):
(1) To satisfy the conditions of gauge invariance for the modified theory
(with the changing propagators of the charged particles) one must change the
form of the one-photon vertex (for example, in the case of spinor QED):
(3.17)
Electromagnetic Interactions in Stochastic Space- Time 109

Fig. 3. The illustration of the change of the one-photon vertex in stochastic (nonlocal) QED.

(Figure 3) due to the Ward- Takahashi identity


kJ',.Jp, q) = SR(P) - SR{ti), (3.18)
where
(3.19)
Here d,/k) is some operator whose action on the entire functions is determined
below.
(2) Any theory with the modified propagators and the vertex functions
contains the minimal number of the many-photon vertices enu n satisfying the
condition
(3.20)
with U o = s; 1. If S; 1 and u1 are polynomial functions, the minimal number of
necessary multi-photon vertices equals (no - 1), where no is a number for
which the relation (3.20) turns to zero. Otherwise, S;; 1 and U 1 would be
non polynomial functions, the number no being infinite.
(3) The interactions of n-photons with the open charged propagators is
given by the following formula (see Figure 4)
(3.21)

Fig. 4. Interactions of n-photons with open charged propagators.


110 Chapter 3

where f n and Un are related by


_ • • /Jil n~l uiqj;kj+1, ... ,kn)
rn(q,kl,···,kn)=un(q,kl,···,kn)+i7n L. .'( .)' x
j~l J.n-J.

x SR(q)fn-j(q;k1, ... ,kj),


n
u1=f 1, qj=kj+qj-t> ql=k1+q, qn=
i
L:
~ 1
ki •

Here the symbol (J}n means summation over all permutations of the variables
k 1 , · · · , kn ·
(4) Now we shall determine the d-operation for entire functions: Acting on
the identity
£1V( _q2[2) = V( _q2[2)£1
by the d-operation and making use of its definition for polynomial function
(Kroll, 1966), we obtain

d/l(k)V( _q2[2) = [V( -(q + k)2[2) - V( _q2[2)] 7;. (3.22)

(5) The d-operation for the inverse of entire functions. Acting on the
identity
V- 1(_q2[2)·V(_q2[2) = 1.

by the d-operation we get

d/l(k)V -l( - q2[2) = - V -l( -(q + kf[2) [d/l (k)V( _q2[2)]V -l( _q2[2). (3.23)

(6) Calculation of d/l(k)SR(£1) where SR(q) is the modified propagator of


charged particles. From the identity
d/l(k)[SR(q)S;l(q)] = 0,
it follows that
d/l(k)SR(q) = -SR (£1 + k)[d/l(k)SR -l(q)]SR(q) (3.24)
or

where
f1/l(k, q) = u1/l(k, q) = -d/l(k)S;ol(q).
It is a particular case of (3.21) at n = 1.
(7) The proof of validity of the generalized Ward- Takahashi identity:
(P/l - q/l)f)p, q) = SR(P) - SR(q), (3.25)
Electromagnetic Interactions in Stochastic Space- Time 111

where
rip, q) = SR(p)u,Jk, q)SR(ti), k= p- q.
Taking into account the relation
u,/k, q) = YI'V- 1 ( _q 212) + (m - q - k)V-l( _p 212) x
x [dl'(k)V( _q 212)]V- 1 ( _q 212) (3.26)
and equation (3.22) we obtain after some calculations identity (3.25).
(8) Charged closed loop: first of all notice that due to Kroll (1966) the closed
loop in the local theory is given by

where

SR(qnyf'iq; kl' ... , kn)SR(ti) = (- Itd(k 1 ) d(k 2 )· •• d(kn)SR(ti)·


A generalization of this equation to the modified theory with entire form
factors represents no difficulty and the charged loop is determined by the
following expression

(3.27)

where

A - k (A ~ 1 A)
n j-:'t j!(n _ j)! SR(qn x
(Jlj
SR(qn)rn(q; k 1 ,···, n)SR q) = V

x uiqj; kj + 1, ... , kn)SR(qj) x fn - /q; kl' ... , k)SR(ti)


= (-It d(k 1 ) ... d(kn)SR(ti),
with r 0 = S; 1 (q). Tensor indexes are omitted here.
Thus we have generalized Kroll's prescription for our case and obtained the
necessary algebraic relations which provide gauge in variance for the S-matrix
in any order of the perturbation series. Investigations of gauge invariance of
the S-matrix for concrete interactions will be given below.

3.3. The Interaction Lagrangian and the Construction of the S-Matrix

According to the above deduction (Chapter 1), we must construct all physical
quantities (for example, interaction Lagrangian, causal Green function, etc.) by
112 Chapter 3

means of the nonlocal fields CPR (X) which are associated with the fields cp(x) by
formula (1.8). The causal Green function of the fields CPR(X) is determined by
expression (1.13):
D~(x - y) = <0IT{CPR(X)CPR(Y)}IO)R4(x)

in the physical configuration space, i.e., in the space of a large scale, where

D~(x) = i -1(2n) -4 fd 4 q e iqx V( _q2[2) K/q) (3.28)

is the Efimov nonlocal Green function if V( - z) = IK(zW is an entire function


and 3.c (q) is the Fourier transform of the local Green function (see Chapter 2).
The Lagrangian of a system of fields is constructed in terms of the averaged
<
fields CPR(X) = CP(X)R 4 (x) in the Minkowski space. Thus the initial Lagrangian
describing the electromagnetic and four-fermion weak interactions of leptons is
chosen in the form
5i'(X) = + 5i'em(x) + 5i'w(x),
5i'o(x)
5i'o(x) = -!: [ovAjx)] [ovAjx)]: + I: lfi/x)(i8 - m)l/!ix):,
j

~m(X) = e:IfiR(x)AR(x)l/!R(X):, (3.29)


G-
,f}
.;z,w(X) = )2: [l/!R(X)O/lVR(X)] [vR(X)O/ll/!R(X)]:,

where A~(x) and l/!7(x), vR(x) are the nonlocal fields of photon and leptons.
The summation in the expression of 5i'o(x) extends over all considered fermion
fields (j = e, /1, V e , V/l).
Formally, the S-matrix can be written in the form of the T-products:
00 1
S=l+i"
~
~S, n'
n = 1 n.

Sn = f f t01
t- 1 d 4 x 1 ··· d4 xn T d [5i'em(xj ) + 5i'w(Xj ) ] } . (3.30)

Here the symbol Td means the so-called Wick T-product or T*-operation (see
Chapter 2, and Bogolubov and Shirkov, 1980 and Efimov, 1977a) and the
lower case d corresponds to the algebraic prescription determined in Section
3.2.
In order to construct the perturbation series for the S-matrix (3.30) by
prescription of the usual local theory, it is necessary to change (in the
F eynman diagrams):
m+k m+k 22 k k 22
2 k2 . ~ 2 k2 . Vm( - k [ ), k2 . ~ k2 . Vo( - k I ),
m- -IEm- -lG - -IE --lG

2
g/lv( _k - iE)-1 ~ g/lYO( _k2(2)( _k 2 - iE)-l,
Electromagnetic Interactions in Stochastic Space- Time 113

and at the same time to insert the modified vertex (3.17) into the vertices at
the external photon lines. The calculation of the matrix elements for the
charged lepton loops will be undertaken using formulas (3.27) and (3.21).

3.4. Construction of a Perturbation Series for the S-Matrix in Quantum


Electrodynamics

The construction of a perturbation series for the S-matrix is possible only


within the framework of an intermediate regularization procedure. We shall
here use the regularization procedure presented in Section 2.4.4. The re-
gularization introduced there makes it possible to pass to the Euclidean metric
in any diagrams of the perturbation theory. We recall that the form factors
V( _q1Il) decrease only in the Euclidean direction, i.e., when ql ----> - 00.
Therefore we shall investigate the Feynman diagrams in the Euclidean mom-
entum space. At the end of all calculations it is nece~sary to remove this
intermediate regularization, i.e., to pass to the limit b ----> 0. Of course, final
results obtained in such a way do not depend on a concrete form of
regularization.
Let us calculate the matrix elements for the S-matrix corresponding to the
following primitive diagrams (see Figure 5) which are divergent in the usual
local quantum electrodynamics (Akhiezer and Berestetskii, 1965, 1981).

0) b) c);
Fig. 5. The primitive Feynman diagrams in stochastic (non local) QED.

3.4.1. The Diagrams of Vacuum Polarization


In the gauge invariant stochastic theory the vacuum polarization in the second
order of the perturbation theory is determined by diagrams sketched in Figure
5a. In the momentum representation the term of the S-matrix which cor-
responds to these diagrams is given by an expression of type (3.27):

n~v(kl' k2) = }~ 2(~~)4 f d4qV~( _ql F) Sp{f~v(q; kl' k2)S~(q)}, (3.31)

where kl + k2 = 0,
SR(Q2)fl"Jq; kl' k1)SR(fj) = (-1? dl'(kdd v (k 2)SR@ (q2 = q + kl + k2 = q).
114 Chapter 3

Equation (3.31) is simplified by the d-operation determined in Section 3.2. The


regularization procedure (j guarantees the possibility of passing to the
Euclidean metric. Thus making use of the definition of the d-operation for the
entire functions and taking the trace of y-matrices:
Spyl' Yv = 4gl'V'

Spyl' Yv YJY p = 4gl'vgJp + 4gl'pgvJ - 4gl'Jgvp,


we obtain

X [dv(k)VJ( _q2 [2)] + -1- dv(k) dl'( -k)VJ( _q2 F)}


= I1~~)(k) + I1~~(k) + I1~~)(k),
where q 1 = q + k,
fi~~) (k) = 4p(1 - 0 I dx(1 - x)-S f d 4 q5ts - 2 X

X [2q/lqv + 2kl'qv + gl'v(m2 - q2 - (kq))],

x f d q[(m2 -
4 q2 - it)S - (m 2 - qi - it)s],

Here we have used the representation for the form-factor (Chapter 2)

(3.32)

and the notation

P =~~f-P-iOO d(~12s.
(2n)4 2i _ P+ ioo sin n(
(0 < P< 1)
Electromagnetic Interactions in Stochastic Space- Time 115

Integrating over d 4 q and going to the Euclidean metric we obtain in the limit

f-
b~O

-R e2 2 1 P - iOO v(O
TIllv(k) =
n 2I·
-22 (kllkv - k gllvb.
-/l+ioo
d( -.- r x
sm n.,
(O<P< 1)

x (m 2[2)(
Jordxx(1 _ X)l - ( r(1r( -- 0() 5£( 0,

where

Assuming m 2 12 ~ 1 we obtain

fif,v(k) = 2e: 2 (kllkv - k 2gllv ) f dx(1 - x)x x

x {In X~o~x~) - ~ + v'(O) + In m212 + O(m 212)}.


We see that after the charge renormalization the value obtained for the
vacuum polarization coincides with the renormalized expression in the usual
local theory (see, for example, Bogolubov and Shirkov, 1980).

3.4.2. The Diagram of Self-Energy (Figure 5b)


The corresponding term in the S-matrix can be written in the form
-i: tlI(X)~R(X - y)t{!(y):,
where

Here
- . - .2f 4
Ie V 0b( - k 2 12) m + p - k~ A
b 2 2
~R(p)=hm
J_ 0
(2 n )4 d k - k2 - 18 . Yll m 2 - (p - kf - 12
. yIlVm(-(p-k) I).

We can use the representation (3.32) for the form factors V!(z) and vg(z) and
the general F eynman parametric formula
116 Chapter 3

f-
Then, after some calculations we arrive at:
- e2 1 f-[3-ioo V(() 2 2 1 Y- ioo ( )
LR(P) = - 2 ----: d(-,-(m I )~----: d1]~(m212)~ x
8n 21 -[3+ioo smn( 21 -y+ioo smn1]
(0<[3<1) (O<y<l)

r(-(-1])
x x
r(1 - or(1 - 1])
x
(1 (p2
Jo dx(l - x)~x-' 1 - m
)~+~
(3.34)
2x (2m - fix),

Assuming the value of m2 ZZ to be small, one can obtain for the self-energy
the following expression:

- 3v'(0) - 1 + -2'
n f-[3 - ioo
d(
v(()v(
'2 -(
() J
(3 - () +O(m212)
} + -16
e2
2 (m - p) x
1 -[3+ioo sm n n
(0 < [3 < 1)

The calculation of integrals of the type


n f-[3-iOO v(()v(-()
----: d(, 2 ( ... )
21 -[3+ioo sm n(
is simple for a concrete choice of the function v((). For example, in the case of
the form factor Vs (2.98) describing the extended electron as the uniformly
charged sphere with radius I (Section 2.3) the first integral in (3.35) is of the
form

~f-[3-ioo d(Vs(()vs(-() (3-()= _3~_ (n-3) ~ _229 I


2i -[3+ioo sin 2 n( 9 n=2 n(4n 2 - 1)(n 2 -1) ~ 90'
where vs (() is given by formula (2.99),
Thus the value obtained for the self-energy differs slightly from the value
calculated in Efimov's (1972) nonlocal theory.

3.4.3. The Vertex Diagram and the Corrections to the Anomalous Magnetic
Moment (AMM) of Leptons and to the Lamb Shift

Let us consider the vertex diagram shown in Figure 5c. In the momentum
representation it has the following form

f~(p, q) = j-'l e 2(2n)-4 f d4kDc( -(p - k)2 F)yv d/l(q)SR(k)yv


Electromagnetic Interactions in Stochastic Space- Time 117

where
dJl(q)SR(k) = (m - k - q)-l YJlVm( _k2[2)(m - k)-l +
+ (m - k - q)-l {Vm( -(k + q)2[2) - Vm ( _k 2 F)}ClYJlq-2.

The symbol (j for the intermediate regularization procedure is omitted here


and below. By using the identity

an - bn = n f: dx x n - 1 = n(a - b.) I dx[(a - b)x + b]n- \ (3.36)

the difference of the form factor values can be transformed to the form
Vm ( -(k + q)2[2) - Vm ( _k 2 12) =

- [q2 1
+ 2(kq)] -----;
21
f- P - iro
_p + iro
d( -.v(()
_.(. [2'
sm n(
11 dx
0
X

x [m 2 - k 2 - 2x(kq) _ q2 X ],-1

which is convenient for calculations.


Carrying out the necessary estimates we obtain the usual form for the
matrix element of the vertex functions between two free single-lepton states:

f~(p, q) = Ui P'){YJlF1(q2) + 2~j (JJlvqv F2(q2)}UiP), (3.37)

where

p' = q +.p,

Here
118 Chapter 3

e 2 n2f-~-iOO v(() 1 f- P - iOO v(1])


N(e 1]) = - - 4 ----;- d( -.- (mJ 12 )' ---:- d1] -.- x
(2n) 21 -0+ ioo sm n( 21 _P + ioo sm n1]

X f f fda df3 dy'P-'y-~ 15(1 - a - 13 - y),


o
5tdq2) = (1 - 13)2 - (q2jmJ)ay, mj = (me, m/,);
5t2(q2) = 5t1(q2) - (q2 jmJ) ty(l - ty) + 2(q2 jmJ)tay + (q2 jmJ)tyf3.
The first term of (3.37) in the limit q2 -+ 0 and with the assumption mJ [2 ~ 1 is

X [
1
-l"3v (0) I .v'(l)
+4 8
- - -In(m -2 1-2 ) - - 20]} -
v(l) 3 9

+ V(1)mJI2[tvI(0) + ~ ~g; + In(mJ12) - ~]}.


a = e2 j4n

(my is photon mass) and contains the terms corresponding to the charge
renormalization of the leptons. The second term of (3.37) at q2 = 0 contributes
to the AMM of the leptons by

aj
_
-
_ 4 a
F 2 (0) - (2')2 -2
1 n
f- iJ - iOO

-iJ + ioo
d(
f- Y - iOO
-Y + ioo
v(OV(1])
d1]. r'
sm n." sm n1]
22
(mj 1 'f
+, x
(O<y,iJ<l)

x (1 -1])(1 - () x

r(1 - 1] - or(1 + 2( + 1])


x .
r(3 - 1])r(3 + 1] + 0
Electromagnetic Interactions in Stochastic Space- Time 119

Assuming mJl z ~ l(mj = me' mJ we obtain

aj = ~{1 + mJ IZ[ -~
2n
n. f-
P- ioo d( v«().v(; - () (I - () -
3 21 _P + ioo sm n(
f V(l)]}

IX
~ 2n [1 - f mJZZv(l)]. (3.38)

The present experimental values of lepton AMM (Van Dyck et al., 1979, 1977
and Bailey et aI., 1979)
aexp(e-) = (1159652200 ± 40)10- 12 , aexp(ll) = (1165924 ± 8.5)10- 9
are reliably confirmed by quantum electrodynamics (Calmet et al., 1977 and
Kinoshita, 1979). It is natural to suppose that the contributions calculated here
should be of an order not greater than the experimental errors. This makes it
possible to establish the following restrictions on the parameter I:
<{1.1 x 1O-
1 '"
14
14
cm for V=V 1
(e-),
2.0 x 10 - cm for V = Vb

1< {0.76 X 10- 15 cm for V = VI


'" 1.40 x 1O- 15 cm for V=Vb (Il),
Here V = VI is given by

(3.39)
and V = Vb equals (1.14).
Similarly, for the level n = 2 of the hydrogen atom, the calculated shift
2S l/Z - 2P l/Z due to (3.37) is (see Efimov, 1977a and Faustov, 1972, for detail)

For the function vb(O determined by (1.15) this expression acquires the form

(3.40)
120 Chapter 3

where
ex 3 Ry = tme ex S = 1.25 x 10 3 MHz/sec.
The observed shift of 1057.912 ± 0.011 MHz is well explained by QED
(Brodsky and Drell, 1970, see also Scadron, 1980). Therefore
IAEl2S 1/2 - 2P 1/2) I ~ 0.011 MHz and substituting formula (3.40) into this
inequality we get I ~ 1.9 x 10 - 13 cm.
In the conclusion of this subsection, we note that the S-matrix obtained is
gauge invariant. Indeed, in the stochastic electrodynamics under consideration,
the Ward identity
o -
aLR(p) = -rip,
-R
0) (3.41 )
P/l
is valid, since this identity is a direct consequence of the identities (3.24) and
(3.17). Since we must not subtract any infinite counterterms, no dangerous
terms which can break the Ward identity (3.41) when formula (3.24) is valid
will appear in the perturbation theory. The diagram of the vacuum polari-
zation is gauge invariant due to our choice of the gauge invariant re-
gularization procedure of Kroll (1966).

3.S. The Electrodynamics of Particles with Spins 0 and 1

3.5.1. Introduction
Now let us construct within the framework of our approach the gauge
invariant quantum electrodynamics of particles with spins 0 and 1 on the basis
of the first-order Duffin-Kemmer equation (see Efimov and Namsrai, 1975;
Duffin, 1938; Kemmer, 1939; Kinoshita, 1950; Kinoshita and Namby, 1950).
We write the Duffin-Kemmer equation
(ia - m)t/I(x) = 0, (a = /3/lo/l)' (3.42)
with
(3.43)
The field t/I(x) consists of three irreducible fields t/I = {t/I 0' t/I(O), t/I(1)}. The first
of which is trivial t/lo = 0, while the other two are spin 0 and spin 1 fields,
respectively. From (3.43), one can obtain
[/3i' 2/36 - IJ _ = 0, (i = 1,2,3).
In accordance with this, the adjoint field !fJ(x) = t/I+(x)rJ with rJ = 2/36 - 1
obeys the equation !fJ(x)(ia + m) = O. The following operators

A(o) = ia - m and d(o) = ia - !(D - m2) + ~(ia)2


m
Electromagnetic Interactions in Stochastic Space- Time 121

satisfy the relation


A(8) d(8) = d(8)A(8) = 0 - m 2 , (3.44)
from which we deduce

The main problem of the theory is to find such matrices P/l which would
grant that equation (3.42) describes a field t/J(x) with definite spin s, namely
s = 0, 1 and does not contain any unnecessary components. By studying two

°
spinor basic representations we obtain here equations for a particle with spins
and 1. The wave functions in this representation can be written as 'PaP' where
rt and P are the Dirac spinor indices. The matrices ')! /l connected with these
indices denote ')!~l) and ')!~2), respectively. Under the Lorentz transformations

x~ = x/l + bell + bW/lvxv,


the spin matrix S/lV reads
SI'V = S~V + S~~) = [PI" Pv], (3.45)

where bell is a translation by an infinitesimal small real vector, bwl'v = - bWv/l


is the antisymmetric matrix for an infinitesimal Lorentz transformation, and
S(j)
f.lV
= .l[.y(j)
4
,,(j)] ,
J1' I v

It is easily proven that the matrices PI' satisfy the Duffin-Kemmer com-
mutation relations (3.43). As usual, the generators for infinitesimal translations
and for 'rotations' are formed by the Hermitian operators PI' = i8/l and the
matrices M I'V = S/lV + (xI'PV - XVPI')' respectively.
The scalar operators
(3.46)
commute with all the generators Ml'v and PI'. They are therefore invariants of
the group and they are multiples of the identity in any irreducible repre-
sentation of the inhomogeneous Lorentz group, and their eigenvalues can be
used for the classification of irreducible representations.
The eigenvalues of the operator p2 equal m 2 , where m is the mass of a
particle. The operator W is an invariant of the Lorentz group and therefore it
is convenient to consider it in the rest of the frame, where P = (m,O). Then
W = m2 S2 = m 2 (Si + S~ + S~).
where
Si = !eijkSjk (i,j, k = 1,2,3) and [Si' Sj] = ieijkSk.
The operators Sj obey the angular momentum commutation rules. The
eigenvalues of S2 are therefore s(s + 1), where s can take integer and half-
integer values and is called the spin of the field under consideration.
122 Chapter 3

Thus, the eigenvalues of the operator W acquire the form W = m2 s(s + 1).
We can therefore label each irreducible representation of the inhomogeneous
Lorentz group with a pair of indices (m, s). The wave functions for fixed m and
s are the eigenfunctions of the operators p2 and W:
(3.47)
In order to obtain the wave function with a definite value of spins s it is
necessary to construct a projecting operator choosing the spin representation
s. Such projecting operator is easily constructed:

- s'(s' + 1)p2
~
pes) = n s(s +W1)p2
Sf '" S - s'(s' + 1)p2
. (3.48)

In our basic representation there are only two spinor states: s = 0 and s = 1.
Both are irreducible. According to definition (3.48) we get
P(O) = 1 - W/2p2, (3.49)
P(l) = W/2p2. (3.50)
A wave function of the basic representation satisfies the so-called Bargmann-
Wigner (1948) equation:
(pii) - m)'P = 0 (i = 1, 2). (3.51)
Then the function
(3.52)
belongs to the irreducible representation (m, s).
In accordance with (3.52) the general solution of equation (3.51) can be
written in the form

'P(s) = n2 (~j)
p +m )P(s)'P, (3.53)
j= 1 2m
where

(
pij) + m)2 = pij) + m for p2 = m 2.
2m 2m
The problem now is to find an equation of type (3.42). For which the
obtained solution (3.53) is satisfactory. The operator A(o) (3.44) must be
constructed from the invariant operators with respect to the transformations
of the inhomogeneous Lorentz group, and must be linear with respect to the
mass parameter m outside the mass shell. For this purpose we transform the
solution (3.53), and introduce the projecting operator

(3.54)
Electromagnetic Interactions in Stochastic Space- Time 123

so that
[X(p)J2 = X(p).

This operator on the mass shell p2 = m2 satisfies the condition

X(p)(V 1 )
V2) +
+ m) = ( 2m
m)(V 1) + m).

Then, making use of these, relations (3.53) can be rewritten in the form

,¥(s) = X(p) P
A(1)
2; m P(s)'¥. (3.55)

Relations (3.53) and (3.55) give a general expression for the wave functions
from the irreducible representation (m, s). Thus, the Bargmann-Wigner equa-
tion (3.51) is equivalent to the following two system equations:

[1 - X(p)],¥~s) = 0, (i = 1,2). (3.56)


In turn, this system is equivalent to one equation
2(a,p)'¥~S) = 0. (3.57)
where
2(a, p) = (V 1 ) - m)x(p) + a[l - X(p)]
and a is an arbitrary parameter. In order to exclude the singular terms with
1/p2 we choose a = - m and obtain

2(m,p) = f3 Il PIL - m = p - m.
Wave functions with definite spin values are the solutions of the following
system equations:
2(a, p)P(s)'¥ = 0, P(S/)'¥ = 0, (S' =1= s).

This system is equivalent to the equation


NS)(p)'¥(S) = 0,

where
i\(O)(p) = (p - m)(l - W/2p2) + ao W/2p2, (3.58)

i\(l)(p) = (p - m)W/2p2 + a 1 (1 - W/2p2). (3.59)


Here the parameters ao and a 1 are arbitrary. The singular terms are removed if
it is assumed ao = a 1 = - m. Making use of the commutation relations (3.43)
we obtain the following expression for the invariant W:
W = B2p2 - 3Bp2 - 2Bp2 + 4p2, B = f3 v f3v.
124 Chapter 3

After simple transformations we have

(3.60)

where

P = --!(B - 4)(B - 1). (3.61)

Using relations (3.43) and (3.45) one can show that the matrix P satisfies the
equalities:

P(1 - P)fJv = o.
Therefore, the matrices fJ~O) and fJ~1) satisfy also the Duffin-Kemmer com-
mutation relations (3.43). Moreover, fJ~O) fJ~l) = O.
Traces of the matrices fJ~O) and fJ~1) are given by formulas following from the
representations (3.45) and (3.61) and from the properties of the y-matrix traces.
Thus, we have for s = 0

S {fJ fJ fJ fJ ... fJ fJ fJ } = {gllvg;.,,··· g,x + gv;.g,,··· gp,gxll' if n is even


p \ fl v ; . "• p, x
J ,
O1' f n 'IS 0 dd
n
and for s = 1

In our case, the divisor d(S)(P) is given by

Now to separate fields corresponding to spin 0 and 1 from the 16-component


wave function 'PaP' where (J( and fJ are spin or indices taking the values 0,1,2,3.
The field 'P in its basic representation is decomposed into the linear inde-
pendent 16-matrices:

(3.62)
J

where definition of QJ follows from the formula (3.62) itself. Here E = ysC and
C is a charge conjugate matrix, so that E obeys the condition EYfl E- 1 = yr.
Making use of formulas (3.45) and (3.61) we arrive at the following
Electromagnetic Interactions in Stochastic Space- Time 125

representation for the operators P and p:


P aal,Pf3' -- [1 + 4"1(y(1)y(2»
I' I' -
1(y(1)y(2»2]
S I' I' aal,{i{il
= + 4"(yp)aal(Cy
1
gaalg{i'{i
- --1
pC )f3'{i -
1 -
S(YpYu)aal(CyuYpC
--1
){il{i' (3.63)
1. [(1) + (2)]
Paal,{i{i' -
A _
2Pp Yp Yp aal,p{i'
= tPp{(Yp)aalg{ilp + gaal(CY pC- 1){iI{i}
The operation of P and p upon the matrix (QJ C)aP in the decomposition (3.62)
gIVes

P(QJ C) = L Paal'{i{i,(QJ C)al{il = {[QJ + hpQJ yp - h"YuQJYuy,,]C}afl (3.64)


a' (i'

and
p(QJ C) = L Paal'fJfJ'(QJ C)al{il = tp,,{[YpQJ + QJy,,]C}a{i' (3.65)
a' {i'

After simple calculations using only the properties of the Dirac y-matrices one
can reduce equation (3.42) with the operator A(0) (3.60) to the following system
equations:

(3.66)

from this it follows (p2 - m 2),¥ = O. Thus, we have obtained the usual Klein-
Gordon equation for spinless particles. The system equation (3.66) contains five
components so that for the matrices f3~0) one can find a representation in the
form of 5 x 5 matrices.
In the case of s = 1 we obtain

(3.67)

from this it follows

(3.68)

This is the Proc equation for spin s = 1. Equations (3.67) contain 10 com-
ponents and one can therefore find a representation in the form of 10 x 10
matrices for f3~1)-matrices.
For the function '¥ 5 in the decomposition (3.62) we get '¥ 5 = O. Thus, the
basic representation containing 16-components have decomposed into three
irreducible representations: the 10-component for spin 1, the 5-component for
spin 0, and the one-component for the trivial case '¥ 5 = O.
It should be noted that linear equations for the definite spin automatically
126 Chapter 3

permit the necessary supplementary conditions to be taken into account with


the exception of unnecessary spinor components.
The investigation of the perturbation theory for the S-matrix constructed on
the basis of Duffin-Kemmer equations will be formally identical to the QED
of leptons. Therefore, we should use the methods and procedures develop in
the construction of the gauge-invariant spinor electrodynamics in terms of the
stochastic space-time concept.
Hence, as was shown above, the averaged field t/lR(X) of a boson field in the
space 1R 4 (x) leads to the change of the free particle propagator (in momentum
representation):

T(A)
P
+ m)
= p(p (2 - p2 + m2 --+ Vm (_ P212)T(A)
P
== T (A)
R p,
m m - P2 - I/;
.)

where P = PI'/31' and /31' are the four l6-rank matrices which are split into 5-
and 10-rank matrices for particles with spins 0 and 1, respectively. In this case
it is also necessary to change the form of the one-photon vertex:
/31 --+ ul'(q, k)
1 = - dl,(k) T;; 1@.
Here the following identities hold:
dl'(k)TR@ = TR(q + k)uiq,k)TR@, (3.69)

(PI' - ql') r:(p, q) = TR(P) - TR@, (3.70)


where
r:(p, q) = TR(p)ul'(k, q)TR(cJ.), k = P - q;
uik, q) = /3Y -1( - q2[2)+(m - q - k)V -1(_ p2 F)[dl'(k)V(-q2 F)]v- 1(_ q2[2)
and
dl'(k)V( _q2[2) = {V( _p2[2) - V( _q2[2)}k- 2(kl' - 2/3l'k + 2k/3I'). (3.71)
Let us now examine the perturbation series for the S-matrix.

3.5.2. The Diagrams of the Vacuum Polarization of Boson Fields


Expression (3.31) for boson fields becomes

fr~v(kl> k2) = ~~ 2:;~)4 f d 4 qVm( _q~[2) Sp{r~v(q; kl' k2)TR(cJ.)},


where

TR(q2)rl'v(q, k1' k2)TR(cJ.) = (_1)2 dl'(k 1) d.(k2)TR(cJ.).


Making use of the definition of the d-operation (3.71) for an entire function we
Electromagnetic Interactions in Stochastic Space- Time 127

obtain

-R . ie 2 f 4 { Vm(-q2 12) 1 1
TIl'v(k) = hm (2 )4 d q Sp PI'
0-0 n
~ Pv --~
m-ql m-q
+ PI' --~-
m-ql
X

x [d v(k)Vm ( _q 212)] + tdl'(k) d,( -k)Vm( _q212)}

= I1~~)(k) + I1~~(l<:) + I1~~(k),


where

ql = q + k,

fi~~ (k) = ~ 1p~2~ ~i~ = gIl dy(l - y)~ (1 - y2)-~ f d 4 qE X

x {(kl'kv - k2gl'v)[0'1y2 + (to'1 - 1)(q~/m2)] +


+ g,'vO'l[q~ + 2m 2 - ik2(1 - y2)]} X

X [q~ + m2 _ tk2(1- y2)]~-2,

x (1 - y2)-~ fd4qE[q~ + m 2 - tk2(1- y2)]~-1,

I1~J(k) = :i[40'1(kl'k v - k2gl'J + 0'2kl'kv] x

x fd4qE{(m2 + q~)~ - (m 2 + qtE)~}'


for spin s = 0
for spin s = 1

Here

Further, making use of representation (3.32) and of the definition for the trace
of f3-matrices (Harish - Chandra, 1946, see also previous section), and
128 Chapter 3

ff(3)(k)
I'V
= 0,

for the scalar boson, and

x {3[ (gl'vk2 - kl v)(1 - 2X)2 x ;i-=-(~) 5t~ -

2 r( -1 - 0 ,£11 + SJ-
- 6m gl'v r(_O.:z,o +

+ 2(gl'v k
2 r( -1 - () 511
- kvkl') r(1 _ 0 0
+ s} ,

fi~~)(k) = 6m2gl'v~~f-Y-ioo d(-;'(O (m 2 / 2 )Sr(-1- 0 x


4n 2z _ y + ioo sm n( r( - ()

x I dx(1 ~ x)-S 516 +s,


ff(3)(k)
I'V
= 0,

for the vector boson, respectively. Here the variables x and yare connected by
the formula y = 1 - 2x.
From the above formulas it is easy to see that in both the cases of scalar and
vector particles the expression fi~2J (k) and the second term of fi~,lj (k) cancel
mutually.
Let us write the other terms:

x I dx(l - x)-S x (1 - 2X)2 5t~, (0 < ')' < 1)


Electromagnetic Interactions in Stochastic Space- Time 129

for the scalar boson, and

fi~~(k) = 41J(
n
1
(gJlVk2 - kJl kv)-2'
I
f- Y
-

- Y + ico
iCO

i
(1<y<2)

1 [ r(-()
x dx(1 - x)-' x 3(1 - 2X)2 () 5£'0 +
o r(1-

x
2 r( -1 - 0
r(1 _ ()
5£10 + 'J
for the vector boson.
Assuming the value mJ[2 (mj = ms , mv) to be small we obtain finally

fi~~(k) =
k2
2:n (gl"v k2 - kl"kJ { m2
r1
Jo dyy4[1 - !(k 2jm 2)(1 - y2)] -1 +

+ 2(v'(0) + In(m2[2» - t}.


ti:~~(k) = ~(gJlVk2 -
4n
kJlkJ{
Joedx[2 - 2x(1 - x)(k 2jm 2) - 3(1 - 2X)2] x

x In[1 - x(1 - x)(k 2jm 2)] + (1 - t(k 2jm 2 x »


(' I
x v (0) k -"34} '
b + 18m2
+ n m I ) + m2[2
2 2
2

b = v'( -1).
It is necessary to perform the charge renormalization of the scalar boson and
the charge and mass renormalization for the vector boson, and after these
procedures we obtain the usual quantum-electrodynamical quantities for the
vacuum polarization of these particles:

fi~~(k) = (gl"vk2 - k Jl kv){n(S)(k 2) - n(S)(O)} = 2:n (k2jm2)(gJlvk2 - kJlk v) x

x I dyy4[1 - !(k 2jm 2)(1 - y2)] -1,

ti:~~(k) = (gJlvk2 - kJlkv){ n(v)(k 2) - n(V)(O) - d~2 n(V)(k 2) b~J

= 4: (gJlVk2 - kllkJ x I dx{[2 - 2x(1 - x)(k 2jm 2) - 3(1 - 2X)2] x

x In[1 - x(1 - x)(k 2jm 2)] - 3(1 - 2X)2X(1 - x)(k 2jm 2) +


+ 2x(1 - x)(k 2jm 2)}.
130 Chapter 3

3.5.3. The Self-Energy of Bosons


Let us now consider the self-energy diagram, shown in Figure 5b (here the
solid line corresponds to the boson field). In our scheme the term of the S-
matrix corresponding to this diagram has the form
- i: !fr(X)L~ (x - y)!fr(Y):,

L~(X - y) = (2n) -2 f
d 4 p eip(x - y) l:~(p),
where
~b( ) = l'
'<"R p 1m
-ie 2
4
fd 4k Vg (-k 2[2)f3
2 11
m2 - (p - k)2 + (p - k)(p - k + m)
2 2 X
b~O (2n) -k - ie m(m - (p - k) - ie)

x f31l'V!(-(p - kn2).
Passing to the Euclidean metric in this expression, and substituting repre-
sentation (3.32) in a way similar to that used in the previous section, it is easy
to obtain the following formula:

A2 A}
-tB(1 - B)Ll
,
1 + (1 - B)L2 L2
m'
0 - .fLl
m '
0 ,

where
p= f3 Il PIl = f3oPo - f3p,

Here

tea + b) < Ya+b + f3a+b < tea + b + 1), a, b = 0, 1,2.


We calculate the electromagnetic correction to the boson mass in the two
limits m2[2 ~ 1 and m2[2 ~ 1. We have: (i) for m2[2 ~ 1
-b _ -~B -1 c5m _ {-(3a/4n)(m/m 2[2)c5 fors = 0,
LR(m) - 48n (B ) m2[2 - (a/4n)(m/m2[2)c5 for s = 1, (3.72)
E[ectromagnetic Interactions in Stochastic Space- Time 131

f
where

i5 =~ -P - ioo d(v«()v.( ~ 1 - () [1 + O(lnm2[2)], 0 < f3 < 1


21 -P+ioo sm n(
and (ii) m2 [2 ~ 1, i.e., in the classical limit

~~(m) = tr:xV(~/) m[1 + O(1jm2[2)] for s = 0, 1.

The latter means that contributions to the mass of the particles with spin 0
and 1 coincide with the classical limit. In general, as shown by Efimov (1977a)
in this limit contribution to the mass of charged particles with spin does not
depend on their spin value, although in the quantum field limit, i.e., m[ ~ 1 it
has a larger singularity with respect to the growth of spin.
It is interesting to calculate the mass difference Am(n+ - nO) ofn+- and nO-
mesons. Since nO-meson is electrical neutral, the value Am(n+ - nO) is caused
by the proper electromagnetic energy of the n+ -meson only. It is roughly
assumed that n+ -meson represents a uniformly charged ball, radius [ of which
determined by the relation

where
<r2>exp '" (0.6 x 1O-13 cm)2 =0.19jm;.
For the charged ball the function vb(O is given by formula (1.15) and therefore

_~f-P-ioo d1] vb(1])V~(;1 -1]) ~ 3.41 = 1.086.


2m -P+ioo sm n1] n

Substituting this value into (3.72) we get


3 1.086
Am(n+ - nO) = r:xm- - 0
3 = 2.6 MeV,
4 . 2

which gives a rough agreement with the experimental value of


Amexp(n+ - nO) = 4.6 MeV. It is clear that by an appropriate choice of a
charge distribution for n+ -meson it is easy to obtain a complete coincidence
with the experimental value of the mass difference Amexp(n+ - nO).
The vertex diagram and the static characteristics for the bosons were
investigated by Dineykhan et a[. (1977) within the framework of the stochastic
space-time, and by Efimov and Namsrai (1975) in the nonlocal theory. The
vertex function f I'(p, q) corresponding to the diagram shown in Figure 5c in
the boson field case takes the form
132 Chapter 3

After simple but long calculations, the function elf/(Pfl + fjp, q))ljJ is reduced
to the following form

elf/(pfl + rip, q))ljJ = elf/P fl ( 1 + 2~ Fo(q2))ljJ +


(3.73)

where MflV = (e/2m) tJisl'v(2B - 4)ljJ and QI'V = -(e/2m) If/sl'v(2B - 5)ljJ,
(Sfl V = PI'PV - Pv/3l') are antisymmetric tensors which define, as we see below,
the dipole and quadripole moments of the boson with spin S = 1. Functions
FJq2) (i = 0, 1,2) depend on the invariant variable q2 = (pi - p)2 and contri-
bute to the physical quantities of the boson field. Here we shall not present
them explicitly but recommend that the reader calculate them as an exercise.
N ow we briefly discuss the static electromagnetic characteristics of charged
bosons. Taking into account the identity

_ _{ I+I }
eljJ/3l'ljJ·A/l = eljJ 2m(P~ P/l) - 2ms/lvqv ljJA/l

= {e(P/l/2m)lf/ljJ - Mflvqv - Q/lvqv}A fl ,

expression (3.73) can be rewritten in the form

elf/(p fl + r /l)ljJ A fl
= {e r (P/,/2m)lf/ljJ - (1 - F 1 (0))M/ tv qv - (1 - F 2(0))Q/tvqv}A fl , (3.74)

where Afl is an external electromagnetic potential. In order to consider the


static characteristics of charged bosons, we pass to the static limit in ex-
pression (3.74). First, let us consider the case of scalar boson, i.e., S = 0. In this
case, through relations (3.62) and (3.66), we get

where ¢ is the field of scalar particles which denotes 'I' in (3.62) and (3.66).
In the static limit we have N/LvF/lv->v(oEjot)->O for v->O, where
F/l = oflAv - avAil' v is the velocity of a scalar particle and E is the electric
V

field. Thus, we see that the magnetic moment for scalar bosons is absent.
The vector ¢fl(X), denoted through 'I' /l5 in (3.62) and (3.67), has the
components

¢/l(x) = (2n)-3/2 f ~ (a/l e~ikx + b;


k e ikx ),

°
where aflk and b; correspond to particle and antiparticle, respectively. If
bilk = it means that antiparticle is absent in our consideration. In the static
Electromagnetic Interactions in Stochastic Space- Time 133

limit the vector a{lk has three components:


a = (I k l/m)a 3 0,
{ o
---+
a{lk = a = ela l + e 2 a 2 + (k/lkl)(wk/m)a 3 ,
where a = (a l , a2 , a3 ) is the polarization vector of the rest vector particle. The
vector spin S and the tensor Tij of the quadripole moment are given by
S=i[a+ xa] and Tij=ataj+atai-tbij(a+a),
respectively. In the vector boson case we have
lfIs{lv(B - 2)1/1 = -m(¢!¢v - ¢~¢{l)
and

where
G{lV = o{l¢V - ov¢w
In the static limit we obtain

tMI'J{lV = - 2~ SH and Q{lJ{lV = 2:2[~ Tij(oiEj + OjEJ],


where H is the magnetic field. Thus, the tensors M{lv and Q{lV define the dipole
and quadripole moments of the vector boson, respectively. Contributions to
these quantities are given by bill = - F 1 (0) and bll 2 = - F 2(0), calculations of
which have already been presented to the reader.
In conclusion it is easily verified that in the case of the electrodynamics of
particles with spins 0 and 1, the gauge invariance is held automatically by the
construction.
Chapter 4

Four-Fermion Weak Interactions in


Stochastic Space-Time

4.1. Introduction

The four-fermion theory (Fermi theory) plays, and it seems, will play a
fundamental role in the development of the theory of weak interactions. The
four-fermion V-A interaction describes in a unified way some of the weak
decays of leptons and fermions. Earlier success of this theory in the expla-
nation of muon and fi-decays has given the certain hope that within the
framework of this theory weak processes might be described at least in the
low-energy domain.
However, in four-fermion theory one meets the well-known difficulty caused
by the ultraviolet divergences and the renormalization problem. For this
reason, calculations of higher order corrections in the perturbation series in
coupling constant GF are difficult. Notice that a similar situation occurs in the
theory with intermediate vector bosons.
Ways of eliminating these difficulties were proposed in different models
which can be classified in two groups. One of them is connected with some
schemes and approaches aimed at constructing a new theory of weak in-
teractions within the framework of gauge theory [especially, the unified theory
of weak and electromagnetic interactions, i.e., the standard model of elec-
troweak interactions due to Weinberg, 1967; Salam, 1968, and Glashow, 1961,
1980]. The other approaches assume a modification of the usual theory of
weak interactions based on the analysis of fundamental principles (causality
condition, locality and properties of geometry on small scale, etc.) of modern
local quantum fic:;.ld theory at small distances (see for example, Alebastrov et
al., 1973; Kadyshevsky, 1980; Efimov and Seltzer, 1971 and Arbuzov, 1975).
The description of weak, electromagnetic, and also strong interactions
within the quark model framework based on gauge theories with a spon-
taneously broken symmetry, obviously represents a qualitatively new step in
understanding elementary particle phenomena and their internal structure.
May be that usefulness of the old (free of the above-mentioned difficulties)
and of the new weak interaction theories is revealed mainly in two limiting
cases. Indeed, when the energy is not high enough for the production of

134
F our-F ermion Weak Interactions in Stochastic Space- Time 135

intermediate particles (for example, Wi, ZO, Higgs bosons, etc.) which are
necessary in gauge theories, i.e., if the energy is small with respect to some
limiting value EI , the weak processes must be described by the four-fermion
theory. Contrarily, when E <: EI the gauge theory will play an important role
in weak processes. Here EI is the value of energy at which new particles W±,
ZO etc. will be produced (see Banner et aI., 1983). Of course, there has to exist a
reason!lble correspondence of both the theories at E ~ E I •
In the language of distance it means that starting from some small scale
Iw ~ 11EI the growth of weak-processes cross-section must be compensated by
corrections given by the intermediate bosons. We shall assume here that this
length of weak interactions is of an order of Iw ~ (hlmwc) ~ 10- 16 cm, and that
the unitary limit is reached at energies EI ~ mw ~ 100 Ge V depending on the
choice of the form-factor of the theory. Notice that it is quite possible that at
these energies the unification of electromagnetic and weak interactions, de-
scribed by the standard model of Weinberg-Salam-Glashow, is achieved.
It seems that in view of this aim, the investigation of four-fermion theory
within the framework of the second approach is undoubtedly interesting, and
can give new information about weak interactions. Moreover, recently great
attention has been paid to the low-energy weak interactions, connected with
the problem of neutrino oscillations and its consequences, and of the proton
instability in the grand unified theories (see, for example, Fiorini, 1979).
This chapter is devoted to the construction of a gauge-invariant theory of
weak and electromagnetic interactions of leptons in the stochastic space-time
1R4 (x}. In this case the investigation of the terms of the S-matrix is carried out
by the methods elaborated in the previous chapter.
Notice that in the stochastic (nonlocal) theory the concept of stochasticity
(locality breaking) is characterized not only by the length Iw ~ 1IEI , but also
by a form of the measure w(biI12} (1.2) (or the shape of a form-factor or of
potential) at small distances. It seems to us that in real physical processes apart
from the value of elementary length an important role may be played by the
form-factor of the theory. This occurred especially in the study of the decay
K2 -> fl + fl- and of the mass difference dm(K2 - Kf} in our scheme.

4.2. Gauge Invariance for the S-Matrix in the Stochastic-Nonlocal


Theory of Weak Interactions

The expansion of the S-matrix in powers of the normally ordered operators of


the electromagnetic field Aix} and the lepton fields tfJ(x} has the form

S= " -1-fd4k ... fd 4kn fd 4p ... fd 4pm


n,'S;, I n! m! I! 1 1
fd 4q ...Ij'dl
4q X

X FJl" ... ,JlJkl>"" kn; PI"'" Pm; q1,··· ,ql}:AJlJkd'" AJlJkn} X


X tfJipd'" tfJiPm}t!ij(Q1}'" t!ij(q/}; (j = e,fl, V., VJl)' (4.1)
136 Chapter 4

The requirement of gauge invariance means that the coefficient functions


IlJ··) in expansion (4.1) satisfy the following conditions:
Fill •... •

k I1j F f..ll.···, j1.j ••. • , J..ln (",)=0 '


kkF
lJ.i Ilj Ill,"" IIi.···. Jij •... ,J1n
("')=0 ' (4.2)

Let us remark that each of the conditions (4.2) are fulfilled when other
momenta in the function Fill •...• Il. ( ... ) are on the mass shell. The series of the
perturbation theory contains three types of diagrams: diagrams with purely
weak vertices, with weak and electromagnetic, and with purely electromagnetic
vertices. Investigation of the last type of diagrams will not be carried out here
because they represent gauge-invariant quantum electrodynamics constructed
by us in Chapter 3. Proof of the fulfilment of condition (4.2) for the diagrams
with mixed weak and electromagnetic vertices is simple. Indeed, in the
considered four-fermion theory of weak interactions the Ward identity

is valid since it is a consequence of identities (3.24) and (3.17) at k = O. By


definition:

Here r.R(p) and r~(p, q) correspond to the diagrams of the self-energy (Figure
6a) and the vertex (Figure 6b), respectively.

OJ b)
Fig. 6. The Feynman diagrams of the self-energy and of the vertex functions in four-fermion weak
interactions.

Proof of the gauge invariance shown in (4.2) in the series of perturbation


theory is quite simple and is based on the identity (3.18). The diagrams of
closed loops constructed by propagators of charged leptons are gauge invariant
due to the d-operation.
Now let us proceed to investigate these diagrams (see, for example, the
diagrams shown in Figure 7) in the series of perturbation theory. First
consider the diagram represented in Figure 7a. In the momentum representa-
Four-Fermion Weak Interactions in Stochastic Space- Time 137

a) b) c) d)

Fig. 7. The diagrams of some closed loops constructed by propagators of charged leptons in
stochastic (nonlocal) theory.

tion the term of the S-matrix corresponding to this diagram is given by


the following expression

F~~)(k) = ~eG 4 fd 4p Sp fJ;)(p, k), (4.3)


2 (2n:)
where
f~~) = d lt (k)SR(p)Oa' (G = GF )·
Taking into account formulas (3.18) and (3.24) we obtain
kit dlt(k)SR(p) = SR(P + k) - SR(P), (4.4)
From this it is easily seen that
kit F~2j (k) = O.
By using the main property (4.4) for the d-operation we can easily prove the
fulfilment of the gauge invariance condition for the other matrix elements of
closed loops, shown in Figure 7. The matrix element corresponding to Figure
7b is given by
(4.5)
where p, k are the external momenta, q is the internal momentum over which
one carries out integration. Making use of identity (4.4) we obtain
kltf~3jp = [SR(P + q + k) - SR(P + q)]OaSR(q)Op +
+ SR(P + k + q)Oa[SR(q + k) - SR(i})]Op
= -SR(P + q)OaSR(q)Op + SR(P + k + q)OaSR(q + k)Op. (4.6)
Elementary integration of the expression of type (4.3) with (4.6) over q gives
the following identity:
kItF~3jp(p, k) = O.
138 Chapter 4

Now we consider the fourth-order diagrams (Figure 7c). In this case the
terms of types f~2j and f~3j{J have the form

From this we easily obtain

kl'f~4j{Jy = [SR(q+ k) - SR(q)]OaSR(q + ql)0{JSR(q + ql - q2)Oy +


+ SR(q + k)Oa x [SR(q + ql + k) - SR(q + ql)]0{JSR(q + ql - q2)Oy +
+ SR(q + k)OaSR(q + ql + k)0{J x [SR(q + ql - q2 + k)-
- SR(q + ql - q2)]Oy
= -SR@OaSR(q + ql)O{JSR(q + ql - q2)Oy +
+ SR(q + k)OaSR(q + ql + k)0{JSR(q + ql - q2 + k))Oy.
In the second term of this expression, the substitution q + k ---> q and In-
tegration over d 4 q give
kI' F(41
l'a{Jy -- °
.
Finally, let us consider one more diagram, shown in Figure 7d. The term
f~~~~) (kl' k2' p) acquires the form

f~~~~) = d v(k 2){dl'(k l )[Oa SR(P + q)O{JSR@]}' (4.7)


From this we see that due to equality (4.6) we obtain the following identity
I'va{J --
k 11' k 2v F (2,2) °
.
Thus, the algebraic relations we have found show that within our model the
terms of the S-matrix satisfy the condition of gauge in variance in form (4.2) in
each perturbation order.

4.3. Calculation of the 'Weak' Corrections to the Anomalous Magnetic


Moment (AMM) of Leptons

In testing the locality of quantum theory, the calculation of corrections due


to the weak interactions to the quantum electrodynamical processes is always
very interesting. In the case of actual discrepancy between the local quantum
electrodynamics and an experiment, which is to be expected at very high
energies, the 'weak' corrections would contain, however, just the breakdown of
QED. From this the conclusion may be drawn that local QED can be
violated, and weak and electromagnetic forces can be equal in magnitude. It is
quite possible that in this domain of energies the process of unification of weak
and electromagnetic interactions starts.
F our-F ermion Weak Interactions in Stochastic Space- Time 139

The present section is devoted to the calculation of corrections to the AMM


of the leptons, based on the assumption that the parameter 1~ Iw ~ 10 -16 em,
which characterizes the domain of unification of weak and electromagnetic
interactions.
In the lowest order of the coupling constant G, the corrections to the
anomaly of leptons are given by two types of diagrams (Figure 8) correspond-
ing to both the diagonal and nondiagonal weak interactions, i.e., to two types
of terms of the interaction Lagrangian 5tw in (3.29). Within the framework of
the nonlocal theory the diagrams of such a type was discussed in detail by
Efimov et al. (1973). Therefore, we shall not calculate it in detail, but give only
the main result. Hence, in the stochastic theory the terms corresponding to
these diagrams, after a Fierz reordering, may be rewritten in the form

f~)(p, q) = lim (2n)-4


~~ 0
G2~1 fd4kN~p(k)OaS!(P + k)ut (P,k) x

x S!W + k)Op (j = e,/1), (4.8)


where the symbol /j means the intermediate regularization procedure used
above. Here p and p' are the external momenta of the leptons, p = p' + q and

SR(P + k)uip,k)SRW + k) = (mj - P - k)-1 yIl S RW + k) + (mj - P - k)-1 qy: x


q
x [Vm ( _/2(p'
J
+ k)2) - Vm.( - P(p
J
+ k)2)]. (4.9)
The function Nap(k) corresponds to the neutrino loop and is given by (see
below Section 4.4.2):

Nap(k)=~~f-~-iW d( .v(() (_Pk2)(~f-p-iW dl1 .V(I1) x


2n 21 _~ + ioo sm n( 21 _p + ioo sm nl1
(O<~<l) (O<p<l)

( 12k2) r(2 + ()r(2 + 11) 1


x - ~ r(1 _ Or(l -11) x r(4 + ( + 11) x

x [ - 2k a kfJr( -( -1]) - gap k 2 (2 + ( + 11)r( -1 -11- 0], (4.10)

Here Re( -1] - () = /j + P < 0 and Re (-1 - 11 - n


= /j + ( - 1 < 0 in first and
second terms of (4.10), respectively. Substituting (4.9) and (4.10) into (4.8),
integration over d 4 k and assuming mJ 12 ~ 1, we get the terms giving the
contribution of weak interactions to the AMM of the leptons:

f~)(q) = .f~)(p, q)lp2 =p12 =m2 =J


i fJ/lAVF(q2)
-2
mj
and
140 Chapter 4

Here the quantity B is given by

B =~-
1 n
122i
f-
-b
b - ioo

+ ioo
d (v(O
--- 1
sin n( 2i
f-
_P
P - ioo

+ ioo
v(1]) x
d1]--
sin n1]

r(2 + 0r(2 + 1])r( -1] - Ov( -1 - 1] - ()


x r(1 _ or(1 - 1])r(4 + 1] + () sin n(1] + () x

x 1
1
+1]+
([48 + 54(( + 1]) + 9(1] + 02 - (1] + (n· (4.11)

An integral of the type of (4.11) will be investigated in Appendix B. Shifting


in turn the contour of integration to the right we can reduce this integral to
the double series. The result of numerical calculations gives B = 2.4 for the
form-factor Vb (1.15).

b)
Fig. 8. The type of diagrams which give contributions to the anomalous magnetic moment of the
leptons (I = e,/i).

Thus, supposing I ~ Iw ~ iO- 16 cm, it is easy to show that the obtained


'weak' contribution aj is much less than the experimental errors for the value
AMM aexp(e-) and aexp(/l) of the leptons.

4.4. Some Consequences of Neutrino Oscillations in Stochastic-Nonlocal


Theory

4.4.1. Introduction
Many papers have appeared recently in which the problem of neutrino
oscillations are considered within different approaches (see, for example,
Bilenky and Pontecorvo, 1980a, b; Bilenky et al., 1980; Vuilleumier, 1979;
Barger et ai., 1980b; for earlier works see the reviews of Bilenky and
Pontecorvo, 1978; Wachsmuth, 1979; Morrison, 1980; and for a more popular
presentation see Thomsen, 1980; Sutton, 1980; De Rujula and Glashow, 1980).
The possibility of neutrino oscillations was first considered by Pontecorvo
(1967): he assumed that the oscillation may appear, if in addition to the usual
Four-Fermion Weak Interactions in Stochastic Space- Time 141

weak interaction, there is another interaction which does not conserve the
lepton number. Such a picture is similar to the oscillation in the system of
neutral kaons. Pontecorvo showed that massive neutrinos might change their
identities during the evolution of time. A particle which is born as an electron
neutrino in a beta decay may periodically behave as if it were a muon neutrino
or a tau neutrino.
Many recent papers (see, for example, Cheng and Ling-Fong Li, 1980; Kang
et al., 1980; Yanagida and Yoshimure, 1980; Neutrino 1982 conference and
references therein) devoted to the problem of neutrino oscillations, deal with
the unified theory of weak and electromagnetic interactions. Thus, in addition to
the standard hypothesis of lepton-quark analogy, in these works a new
conjecture is proposed that leptons, like quarks, are mixed. In such a theory
the oscillations ve~ vI' ~ Vt appear.
Possible indications for neutrino oscillations have been obtained in the
beam-dump experiments at CERN (De Rujuila et al., 1980) and in the
experiments with reactor antineutrinos (Barger et aI., 1980a; and Reines et at.,
1980). A number of experiments searching for neutrino oscillations stimulated
recent interest in the question, since they appear to give some indications for
nonzero neutrino masses. A direct experiment of Lubimov and co-workers (see
Kozik et al., 1980) on measuring the lie-mass from 3H-decay gives the mass of
the electron neutrino as between 14 and 46 eV, and most probably to be
36 ± 10 eV at 95% confidence level. The results of Reines, Pasierb and Sobel
(1980) gave no direct implication for the neutrino mass, although they implied
that the difference in mass between the two basic states lies in the region of
1 eV. It is worth noting that in the recent developments of grand unified
models, the possibility of the appearance of neutrino masses is also discussed
(Barbieri et al., 1980a, b; Gell-Mann et al., 1979; Georgi and Nanopoulos,
1979; and Witten, 1980).
Thus, we believe that a neutrino oscillation mechanism does exist in nature
and we will expect precise experiments on the properties of neutrinos.
Our aim here is modest, and consists only in considering this problem from
the viewpoint of the stochasticity of space-time. We believe that due to the
neutrino oscillation mechanism there exist mixed states of neutrinos which
give nonorthogonality between neutrinos, say, ve and vI" <velvl')=!=O (v t

oscillations are not considered here). In our case it is postulated that the
difference in behavior between Ve and vI' is caused by the internal properties of
these particles and depends on the non local effects of weak interactions. For
example,
VI' = Ve +fv = Ve - (1 - </J)v, (4.12)
where f(or </J) is some parameter which may be connected to a mixing angle
and a value of mass difference of neutrino states, and v is a massive neutrino
(basic state) which possesses the properties of both the Ve and vI' neutrinos. In
the representation (4.12) the transition propagator between Ve and vil has the
142 Chapter 4

form
(4.13)

where we put DVev.(x) = Dvev(x) since the v neutrino possesses the property of
Ve·
If the neutrino mixing is assumed, the exotic decays Jl ..... ey, Jl ..... 3e,
K± ..... Jlen±, K2 ..... Jle, etc., are principle, possible. Notice that these decays are
forbidden by the usual theory. These processes appear in the higher orders of
the perturbation theory. The present section is devoted to the investigation of
these decays within the framework of our approach. It is shown here that if the
parameter I for the weak interactions is of an order of I ..... lw '" 10 -16 cm and if
neutrino mixing takes place, then the probability for these decays is close to
the experimental upper bounds (see, for example, Bricman et ai., 1980; and
Fiorini, 1979). In this section, for example, we will consider the decays Jl ..... 3e
and K2 . . . Jle in detail.

4.4.2. The Jl ..... 3e Decay


Within the framework of our model of weak interactions the probability decay
is determined by the diagram shown in Figure 9a. Before we proceed to
estimating corrections from these diagrams to the probability of decays Jl ..... 3e
and K2 ..... Jle, let us consider diagrams of the type of neutrino-neutrino,
neutrino-lepton and lepton-lepton loops (Figure lOa, b, c). Here we study one
of these diagrams, say, the neutrino-lepton loop corresponding to the diagram

f
shown in Figure lOb. Its matrix element has the form
1 4 ~ ~ Vo(_k2[2) Vm(-(k + pfl2)
A

JIap(p) = (2 )4· d k Sp[kOim + k + p)Op]


nl .
- k2 -u;m-
2 (k +)2
p -Ie:
..

>'_--_'<
e
J'I
,--_ .....

a.)
e
e
b)
Fig. 9. The Feynman diagrams determining the probability of decays Jl .... 3e and Kf .... jle.

\~'
.... /
--~_.-#'
-'

a) b) C)
Fig. 10. The Feynman diagrams of neutrino-neutrino, neutrino-lepton and lepton-lepton loops.
F our-Fermion Weak Interactions in Stochastic Space- Time 143

After the standard calculations we have

ITaP(P)
1
= - 2 --:
1I-0-ioo v(O
d( -.- (m 2 z2)' --:
1I-Y-ioo d1} ~
() X
2n 21 _/j + ioo sm n( 21 _Y + ioo sm n1}
(0 < (} < 1) (0 < Y < 1)

x (m 2 [2'f 1
r(1 - or(1 - 1})
x i1 (l-X)'(
0
dx ~~
x
1 -p2)~+'
-
m2
x

x [( -2PaPp + gapp2)x(1 - x)r( -1} - () +


(4.14)
We notice that the neutrino-neutrino loop (4.10) in Section 4.3 is obtained by
substituting m = 0 in expression (4.14).
So, by definition (4.13) the matrix element corresponding to the /1-" 3e
decay has the form

M(/1-" 3e) = ¢(G/J2l(eOa/1)Nap(q)(eOpe),


where Nap(q) is the neutrino-neutrino loop determined by (4.10). Then the
square of the matrix element equals

L JMJ2 = 4(nl)-4G 4¢2c 2[(k1 k2)(kok3) + (kok2)(k1 k 3)], (4.15)


spin

where

;i I
-p - ico
d( v(Ov( -1 - 0 (4.16)
C = - -P+ico sin 2 n( ,
(0 < P < 1)

In expression (4.15) for the square of the matrix element we take into account
main terms of the order of (G 2 /[2). After integration over phase space of the
electrons and averaging over the muon spin we have
1 G4 ¢2C 2
W(II -.. 3e) - - 5
m /1"
r - 192 4[4 n 7

The probability branching ratio of /1 -.. 3e to /1 -.. evv constitutes


B(/1-" 3e) = W(/1-" 3e)/W(/1-" evV) = G2¢2c 2(n[)-4/4, (4.17)
where
2.3 for v = Vb,

c = { 1.4 for V = vs '


0.65 for v = VI.

Here Vb' V., and VI are determined by formulas (1.15), (2.99) and (3.39),
respectively, and J¢J~ 0.055 (see Lee et al., 1977). Assuming ¢ ~ 10- 2 and
144 Chapter 4

[ -> lw ~ 2 x 10 - 16 cm we get
1.3S x 10- 8 for Vb'
{
B(fl-> 3e):.5 4.9 x 10- 9 for vs, (4.18)
1.0 X 10- 9 for VI'

4.4.3. The Kf -> fle Decay


In the second order in Gthe decay Kf -> fle
is described by the diagram in
Figure 9b. The term of the S-matrix corresponding to this diagram can be
rewritten in the form

(4.19)
where

(mN +k- P-)Vm( -(k - p_)2[2)


x 2 2 0P' (mN ~ mA)'
mN - (k - p_) - is

Here rr~p(k) corresponds to the neutrino-nucleon loop determined by (4.14) at


mL -> mN • The hadron current is chosen in the Cabibbo form. After tedious but
elementary calculations we obtain

where
n
A=----;-
I-Y - ioo v(y) 1
dy-.-----;-
I-P - ioo vee) 1
d(-.-----;-
I-a -ioo v(l'/)
dl'/-.-x
21 -y+ioo smny 21 -P+ioo smn( 21 -a+ioo smnl'/

v( -1 - 1'/ - ( - y) r(2 + 0 r(2 + 1'/) r( -1 - 1'/ - 0


x , (0 < IX, {3, y < 1).
sin n(y + 1'/ + 0 r(l - 1'/)r(1 - Or(3 + 1'/ + 0
(4.20)
As above, after displacement of the contours of integration to the right we get
(see, Appendix B)

A = 2-
n I-P - ioo vee)
d(---.
1
x
2i _p + ioo sin 2 n( ((1 + 0(2 + ()

L
00

x [v(n)v( -1 - n - () - v(n - Ov( -1 - n)]. (4'.21)


n=O
Four-Fermion Weak Interactions in Stochastic Space-Time 145

Now the matrix elements (4.19) acquire the form

M = i}2 fKNII(G/}2)Z cos 8e sin 8e 1;;4~Z A ·CPK[,ii(p-)qe(p+)],


and its square is
IMl z = 2m;(1 - m;/m;)AZgZ,
where
GZ m m .4>
g = fKNII 2 cos 8e sin 8e ;6n~ F

Integration of the decay probability over the phase space of two leptons gives
m
W(Kf ~ fle) = -t[1 - (m;/m;)]2 x

xA 2 cosz8e sin z 8e .A-.


'P
2G4 f2 m2m2 n- 8 1- 4 /512n
KNII '" N (4.22)
where sin 8e = 0.23,f;NII/4n = 10.0 ± 2.8 (see, for example, Ebel, 1970; Nagels
et al., 1976), 4> ~ IO- z, and 1~ lw ~ 2 X 10- 16 cm;
- 1.8 for v = Vb'
A= { (4.23)
-2.7 X 10- 1 for V = vs.
Thus, the branching ratio of this exotic decay in our model is
_ W(Kf ~ fle) < {1.2 x 10- 8 for v = Vb'
B - ~ 10
W(Kf ~ all) 1.4 x 10- for V = vs.
We see that the exotic decays fl ~ 3e and Kf ~ fle depend on the form-
factors of the nonlocal theory. Our results, together with the calculations by

Table 4.1.

Processes Our Results of Experimental


calculations Cheng and Li upper limits

1.3 x 10- 8 V = Vb
{t-> 3e 4.9 x 10- 9 V = Vs 10- 12 1.9 X 10- 9
1.0 X 10- 9 V = VI

Kf -> {te 1.2 x 10- 8 V = Vb 10- 10 2 X 10- 9


1.4 X 10- 10 V = Vi}
Kf -> {t+ {t- 1.2 X 10- 6 V = Vs (9.1 ± 1.9) x 10- 9

~m(Kf - K?) 5 x 1011 h sec- 1 v= Vs 0.5 X 10 10 h sec- 1


2 x 1011 h sec- 1 V = VI
146 Chapter 4

Cheng and Ling-Fong Li (1977) and the experimental upper bounds [Bricman
et aI., 1980 (Particle Data Group)], are summarized in Table 4.1. These
numerical calculations are presented for a purely illustrative purpose. These
are important in a sense that they permit one to estimate a parameter of
mixing at a given value of elementary length l----* lw ~ 10- 16 cm. The in-
troduced parameter <p is connected to the mixing angle and the mass difference
of neutrinos v1 and v2 (or heavy leptons N 1 and N 2) of other models of weak
interactions by the formula
<p ~ sin qJ ·cos qJ. Ami, (i = Vi' NJ

4.5. Neutrino Electromagnetic Properties in the Stochastic-Nonlocal


Theory of Weak Interactions

In recent years great attention has been paid to the physical properties of
neutrinos. This is connected with the problem of neutrino oscillations and its
mass, and astrophysical consequences (Faber and Gallagher, 1979; Schramm,
1979; Schramm and Steigman, 1980; this problem has been discussed by many
authors at the Neutrino 1979 and 1982 Conferences), and also with the rapid
development of neutrino experiments carried out at CERN, Caltech-FINAL,
Serpukhov, etc. (see, for example, Busser, 1980; Winter, 1979; Baltay, 1979;
Arbuzov, 1975).
As usual, the neutrino is considered as a weak-interacting particle with zero
mass and without an electric charge. Therefore among the electromagnetic
characteristics of neutrinos the only nonvanishing quantity is its charged radius
r,. Possible experiments on measuring the charged radius of the neutrinos have
been pointed out by Andryushin et al. (1971).
However, it seems that the neutrino mass is not zero (see previous section).
Then the neutrino may possess the magnetic moment avo Recently, the magnetic
moment of a massive neutrino has been discussed by Fujikawa and Shrock
(1980).
Starting with the analysis of experimental data on inclusive reactions
v(v} + N ----* v(V) + anything and vI' - e elastic scattering, Bardin and
Mogilevski (1974); Kim et al., (1974) have investigated the electromagnetic
properties of neutrinos and calculated the correction to.the cross-sections of
these reactions due to Feynman diagrams involved in one-photon exchange,
and obtained the restriction on the charged radius and magnetic moment of
the neutrino.
The contribution due to the one-photon exchange calculated by these
authors is called electromagnetic, although, as is known, the charged radius
and magnetic moment of the neutrino must appear due to the effects of the
weak interactions. The calculation of these quantities in the usual theory of
weak interactions meets with difficulties because of divergences in the S-matrix
elements.
Four-Fermion Weak Interactions in Stochastic Space- Time 147

The present section is devoted to the calculation of contributions to r v and


av within the framework of our approach formulated in this book. In Feynman
diagrams of the order eG 2 giving the corrections to rv and a v there are closed
loops constructed by propagators of the charged leptons and neutrinos for
calculations of which it is necessary to apply the method of the stochastic
theory.
In the stochastic-nonlocal theory of weak interactions the electromagnetic
interaction of the neutrino, say, the muon one, in the lowest order of G is given
by the following Feynman diagrams (Figure 11). It should be noted that in the
stochastic-non local four-fermion weak interaction theory the Fierz reordering
is not yet studied so that in this section we consider all possible diagrams
which give contributions to the electromagnetic form-factors of neutrino.

VJ<
p --. tl,..
P':P+'1..

-~-~-:~
-- - ve
d)
ve
e)

Fig. 11. The vertex diagrams giving corrections to the electromagnetic form-factors of neutrino.

The matrix elements of the vertex functions corresponding to the diagrams


shown in Figure 11 between two-neutrino states which are used for the
calculations of the electromagnetic form-factors of neutrino have the following
general structure:
(4.24)
where
1
F 2 (O) = - av ' q = p' - p.
2me
Here r~ is the mean-square charged radius of the neutrino and a v is its
magnetic moment in the units of electron Bohr magnetons.
In order to calculate the contributions from the diagrams shown in Figure
148 Chapter 4

11 to rv and av we shall study them separately. First, let us consider the lepton
loop (Figure lla). The term for the S-matrix corresponding to this diagram
has the form
G
M 1 = ie J2 iiip')Oauv(p)Kap(q)AP(q), (4.25)

where the Kap(q) in the stochastic theory is given by

where
up(q,k) = -d p(q)S;1(k).'
On the other hand, taking into account the identity (3.24) for the d-operation
we have

Kap(q) = (2:)4i f d 4 k Sp[Oadp(q)SR(k)].

By definition
dp(q)SR(k) = (m - k - q)-1YpVm(-k2[2)(m - k)-1 + (m - k - q)-1 X

X [Vm( -(k + qf[2) - V m( _k2[2)] qyJ.


q
After some elementary calculations we obtain the gauge-invariant expression
for Kap(q):

where

K(q2) 1
= _----;-
2l
f- a
-
-a + ioo
ioo
d( ~ (m 2 [2)~ r( -() x
sin n( r(1 - 0
(0 <a< 1)

x Jor dx x(1 -
1
X)1 - ~ [1q
- 2
m2 x(1 - x) J~ .

Making use of the identity


r( - 0 = - nisin n(r(1 + 0,
we have
- 2 n
K(q ) = 2i
f- a - ioo
-a+ioo
v(() 1
d( sin 2 n( r(1 _ 0r(1
r dx x(1 - x)
+ () Jo
1 1-
~x
Four-Fermion Weak Interactions in Stochastic Space-Time 149

=
r 1
Jo dx x(1 - x) v'(O)
{
+ In m2 [2 + In
m2- q 2x(1 -
m2 (1 _ x)
X)} .

Here

r 1
Jo dx x(1 - x) In(1 - x) =
r(2)r(2)
r(4) [t/J(2) - t/J(4)] = -i6'
1
t/J(1 + n) = t/J(n) + -, t/J(I) = - C,
n
c ~ 0.577 is the Eulerian constant.
The first diagram considered gives a contribution to the mean-square radius
of the neutrino only:

r 21 v -6
-
8 F(I)( 2)1
8q2 1 q q2 = 0 -
_
J2
G 1 [2
2n 2 6 - V'(0) - In mZ[Z] . (4.26)

Assuming, as above, [~[w ~ 2 X 10- 16 cm we obtain


1 x 10- 33 cm 2 for v = Vb'

ri = { 0.9 x 10- 33 cm 2 for V = vs,


1.1 x 1O- 33 cm z for v = VI.

We see that the contribution from the diagram (Figure lla) to the mean-
square radius of the neutrino is of the order of 10 - 33 cm z.
Now we turn to discussion of the vertex diagrams. There are only eight
diagrams for each neutrino vi(i = e, f.1). Among them are diagrams of different
structure, e.g., those shown in Figure llb--e we shall begin with the calculation
of these diagrams. The terms for the S-matrix corresponding to these diagrams
are:
r 1p (p, q) = + p)]Op Sp{OaS~)(kl + kz)OpS~)(kz)},
Na[d/ -q)Sjf'>(k l
rzp(p,q) = NaSjf')(kl + p)Op Sp{Oa[dp(-q)S~)(l(l + kz)]OpS1')(I(z - q)},
r 3p(P, q) -_ N aSR(v) (k z - q)Op Sp{Oa[dp( - q)SR(e) (kl + kz)]Op SR(Il) (kl + p) ,
~ A ~ ~ ~ A }

r 4p(P, q) = N aS1 )(kz )Op Sp{OaS~)(kl + kz)Op[d p( - q)S~)(kl + p)]},


V

Na = i- 1(2n)-4 fd4kli-l(2n)-4 fd 4k 2 ·Oa' (4.27)

respectively.
Let us consider the first expression of (4.27). We are interested only in those
terms which give contributions to rv and a v in the limit qZ ~ O. The structure of
150 Chapter 4

type (4.24) is obtained in the usual way. After calculations we have

r 1p (p, q ) -- 16n4[2
1
3"1 [ ypq 2 F(ll . a F(1)] ,
1 + UJpaq mv 2 (4.28)

where

F\l) = P«(,11){2(3 + 211 + 2()r( -1 - 11 - 0 + r( -11 - m! - (1 + 11 + O]},


1+11+(
Fi1 ) = p«(, 11)r( -11 - O[ - 2 (1 + 211 + 20 - 13 - 611 - 6(].

Here

p«(, 11) = --:-


nf- fJ - ioo
d( -.-
v(Ox
21 _fJ + ioo sm n(

1
x-
2i
f- Y - ioo

_y + ioo
V(l1)
dl1--
v( -1 - 11 - ()r(2 + 11)r(2 + ()
sin nl1 sin n(l1 + ()r(1 - l1)r(1 - 0r(4 + 11 + ()
.

Similar calculations give the following structure for r 4p(P, q) and r 2p(P, q):
(4.29)

where
F\4) = p«(, 11){[ -2 + 2(1 + 11 + mr( -11 - 0 - lOr( -1 -11 - m,
F(4) = p«(, 11)[ -10 - 17(1 + 11 + 0 + 4(1 + 11 + 02]r( -11 - 0,
and

(4.30)

Here

F (2)
1 = p«(, 11)r(1 - 11 - 0{ 3+ x [x---1 - -2 -+ (x3 + x) - -5 2-+ -xJ+
(1-11)(1+11) 6 24 2 11

+ -1- x [2+((
- - - -7 - -x 53+X)
5 +- 1 - x)(4 + x) +
- - - -(1
2 - 11 1 + 11 6 6 2 11 12

13+x - 4x) + -
+ ---(1
31 +11
l1-X( 2+()
-- 1 - -
33 -11
- +
1 +11
J 1 -(1 - x) -
(1-11)(2-11) 3
[1
52 + x 1 [3 + xJ41 25 2+ (
-21+11(3+x) +(2-11)(3-11)x 1+11(-6-6x )+I+11'x
Four-Fermion Weak Interactions in Stochastic Space-Time 151

x ( -1 -!X + ~ 3 ~ X) + (2 + xWl + ix) + l2(2 - x)(l - x) J+


1] [1 3 + X 17 7 12+
+ (1 - 1])(2 - 1]) -2 1 + 1] (1 - x) + (2 + X)b4 + 24X) + 2 1 + 1] +
(J
2 1 2+((2) 2+x3+x
+- - - +x + - - - - - +
3 (1 - 1])(2 - 1])(3 - 1]) 1 + 1] .3 1 + 1]

F(2) = p(C 1]). r(1 - 1] - () [20 (5 +6 + 2) + 9 (1 _ ) +


2 6x(1 _ 1]2)(2 _ 1]) x x x 1] 1]

+ 21]X 2(1 - x) + 18X1](1 - 1]) +


+ 1]X2(( - 1]) - 8X1]2(1 + x) - 41]2 x 2(1 + x) -
- 7X1]2(X + 1]) - 1]2 x 2(3 + 2()J, (x = 1] + o.
It is easily seen that by substitution of the variables of integration, the third
term can be transformed to the second term of expression (4.27). Thus, in our
case of q2 -+ 0 we get
r 3p(P, q) = r 2p(P, q).
Now we proceed to the numerical calculations for the concrete form of the
function v((), say, for Vb which is determined by (1.15). The results of a
displacement of the contours of integration in (4.28}-(4.30) give
F\l) = -0.18, F~l) = 2.56,
Fi4 ) = -0.28, F~4) = 6.7, (4.31)
Fi2 ) = - 0.05, F~2) = 1.07.
Finally, we shall consider expressions (4.28}-(4.30) together with (4.31). Then
Mp = b(r 1p + 2r 2p + r 4p ) = b[YpF1(q2) + i(JPIXqIX mJ2(q2)], (4.32)
where
1
F 2(0) = 16n4 [2 (f1 + 212 + 14), b = ieG 2/2;

F1(q2) = (q2/16n 4 [2)(R 1 + 2R2 + R4).


Here
11 = tF~l)(O) = 2.56/3, 12 = F~2) = 1.07, 14 = i F~4)(0) = 1.1,
Rl = t( -0.18), R4 = i( -0.28), R2 = -0.05.
152 Chapter 4

Therefore
r~,= <rL> = 6q- z G z F 1 (qZ),

aVj = 4mvjme(Gz/2)Fz(0), Vj = Ve, vil


in the units of the electron Bohr magneton.
Recent experiments and data analysis [see, for example, Bardin and
Mogilevski, 1974; Kim et al., 1974; Daum et al., 1978; Bricman et al., 1980
(Particle Data Group)] established the following restrictions for v , < mv , a r; >,
and mVe "
m :5 0.57 MeV,
V"

m'e:5 6 x 10- 5 MeV.


Thus, our result gives the following restrictions:

and
<r~>:5 1 x 1O- 33 cm 2 ,

on the assumption 1--> Iw ~ 2 X 10- 16 cm. If (14 .,;;; mVe .,;;; 46) eV due to Kozik
et al., (1980), then
4 x 1O- z0 :5lavJ :5 1.34 x 10- 19 .
We notice in our model that the vertex diagrams, shown in Figure 11 b-e give
small contributions to <r~> of an order (l0-37_1O- 38 )cm Z with respect to
diagram 11a.

4.6. Studies ofthe Decay Kf --> p+ p- and Kf-and K~-Meson


Mass Difference

4.6.1. Introduction
Some time ago the rare decay Kf --> p+'p- was observed [see Bricman et aI.,
1980 (Particle Data Group)], whose branching ratio coincides, in order of
magnitude, with the unitary limit (Sehgal, 1969; Quigg and Jackson, 1968)
Bu(Kf --> p+ p-) = W(Kf --> p+ p-)!W(Kf --> all) :<; 6 x 10- 9 .

This process occurs essentially due to electromagnetic interactions (through


two-photon exchange). It is interesting to note that the well-known so-called
GIM mechanism (the hypothesis of the existence of the charm quark) orig-
inated from the investigation of this process within the standard model of
electroweak interactions (Glashow et al., 1970).
The calculation of 'weak' corrections to the B(Kf --> p + p-) and the mass
Four-Fermion Weak Interactions in Stochastic Space- Time 153

difference of Kf - and Kf -mesons within the usual nonrenormalizable theory


(the four-fermion theory and the theory with intermediate bosons) of weak
interactions gives a very small value for the cut-off momentum A of an order
of a few GeV. This contradicts the value of the natural cut-off
A ~ 10 2 -;- 10 3 GeV for the growth of weak interactions (see, for example,
Ioffe, 1973).
In this section we want to calculate matrix elements of the order O(G 2 ) in
decay Kf ~ 11 + 11- and mass difference ilm(Kf - Kf) within the framework of
the stochastic-nonlocal theory of weak interactions without introducing the
fourth quark. It is quite possible that suppression of these matrix elements may
be achieved by an appropriate choice of form-factors of the theory. However,
as shown below, change of these quantities is slow for the form-factors vs ' Vb
and Vi under consideration.

K~,~
7ZZZZZ£~-

;; F ?i
bj
OJ
Fig. 12. The Feynman diagrams which give contributions to the K f -> jl + jl- and the
L1m(Kf - K~).

In our model contributions of nonlocal interactions to the Kf ~ 11 + 11- and


the ilm(Kf - Kf) arise from the diagrams, shown in Figure 12.

4.6.2. The Kf ~ 11 + 11- Decay


Kf ~ 11 + 11- decay in the second order in Gis described by the diagram shown
in Figure 12a. The corresponding term has the form

M(Kf ~ 11 + 11-) = ififKNA (Gj fi)2 il(p- )r(p-, q)I1(P+ )q>K cos 8e sin 8e
where r(p _, q) was calculated in Section 4.4.3. Thus, the branching ratio of
this decay is
B(Kf ~ 11+ 11-) = W(Kf ~ 11+ 11-)jW(Kf ~ all) = l.7A2 x 10- 5 , (4.33)
where
m ( 1-2-" G4 m 2m 2 f2
m2)2 A 2 cos 2 8 sin 2 8 _~
W(Ko~I1+I1-)=~ KNA
L 8 mi [4 128n 8 4n
C C

and A is given by (4.20) or (4.21). Substitution of numerical value (4.23) for A


into (4.33) leads to the contribution
B(Kf ~ 11+ 11-) = 1.25 x 10- 6 for V = Vs· (4.34)
154 Chapter 4

4.6.3. The Mass Difference of K2- and Kf-Mesons


Let us now find the energy operator for the transition K2 into Kf. A typical
diagram of the order G 2 giving the contribution to the ~m(K2 - Kf) is shown
in Figure 12b. The expression corresponding to this diagram is
(4.35)

where

and

ITap(P) f
= j-1(2n)-4 d 4 k Sp[SR(k)OaSR(k + jJ)Op],

which is determined by the expression of (4.14). In the case mil 2 ~ 1 the


expression obtained for ITaP(P) on the mass shell of the K-meson acquires the
form
(4.36)

where c is given by (4.16).


The calculation of functions ITj(P) U= [3) is similar. As a result

f-
IX,

IT.( )=fKNAmNPj 1 b - iOO dr~12~~f-Y-iOO d ~12q x


J P 4n 2 2·1 _ b + ioo
...
sm n..r 2·1 _ y+ ioo rJ sm
. nrJ

The contour integration gives

IT /p) = - fKNA ::~j {V'(O) + 1 + In(mM2) +

+ 1o
1 [p 2 x(1 -
dx In 1 - -2
mN
x) ] - ----:-
n
21
f- b - iOO
- b + ioo
V(OV(-O}
d(. 2
sm n(
. (4.37)

Here it is necessary to carry out renormalization in the strong coupling


constant f KNA. After such renormalization expression (4.37) takes the form

r
IT/p) = -(mNP)4n 2 )fKNA· p,
(4.38)
p = dx In[1 - (p2/m~)(1 - x)x].

Substituting (4.38) and (4.36) into (4.35) we obtain the following expression for
F our-F ermion Weak Interactions in Stochastic Space- Time 155

Am(K2- Kf):

L.l.m
A (KOL -
KO)
S = (2 n )-6f2KNA G 2 SIn 2 C· p2
· 28ccos 28cmNymK' (4.39)

or

for v = Vb'
for V = vs, (4.40)
for v = v1 .

Table 4.1 presents the contributions calculated within our nonlocal theory
and experimental data on K2 --> f.1 + f.1- and the mass difference Am(K2 - Kf)
for different form-factors of the theory. From Table 4.1 we observe some
dependence of physical quantities under consideration from the form-factors of
the theory.

4.7. Appendix B. Calculation of the Contour Integral

In this appendix we give the method of calculation of the contour integral, for
example, the integral A given by formula (4.20):

n f-Y-;oo v(y) 1 f-P-;oo v(() 1 f-a-;oo v(1/)


A =-.- dy-.~-.- d(-.--.- d1/-.- x
21 _y +;00 sm ny 21 _p +;00 sm n( 21 -a +;00 sm n1/

v( -1 - 1/ - Y - ()r(2 + ()r(2 + 1/)r( -1 - 1/ - 0


X (0 < rx, [3, y < 1).
sin n(y + 1/ + ()r(1 -1/)r(1 - Or(3 + 1/ + 0
(B.l)

First we displace the contour y to the right. Then the poles will appear at
points y = 0, 1, 2, ... and y = n - '1- ( (n = 0,1,2, ... ). In the first case
(y = 0, 1,2, ... ) it is necessary to displace one of the other contours, e.g., [3-
contour. The calculation of the residues at points ( = 0, 1,2, ... and ( = N - 1/
(N = 0, 1, 2, ... ) give

n f- a -;00 v('1) 1
I
00
A 2- d'1-- v(k)v( -1 - k - '1) (B.2)
+ '1)(2 + '1) k = ° .
=
2i _a +;00 sin 2 n1/ '1(1

From similar calculations of residues at points ( = 0, 1, 2, ... and ( = N - 1/ in


the second case (y = n - '1 - 0 we get

nf-a-;oo v(1/) 1
I= ° v(k-'1)v(-1-k).
00
A= -2- d ' 12- - (B.3)
2i -d;oo sin n1/ 1/(1 + 1/)(2 + 1/) k
156 Chapter 4

After uniting expressions (B.2) and (B.3),


n
A=2-
I-a -;00
dl]--
V(I]) 1
L00 [v(k)v(-1-k-I])-
2i -a +;00 sin 2 nl] 1](1 + 1]) (2 + 1]) k = 0
- v(k - I])V( -1 - k)]. (BA)

The last integral (B.4) for the concrete form of form-factors of the theory is
calculated easily. For example, let
v= Vs = 21 + 2~/r(3 + 20.
Then
00
L Cv(k)v( -1 - k - 1]) - v(k - I])v( -1 - k)]
k=O

1 1
=2-2~ [ + +
r(3)r(1 - 21]) r(5)r( -1 - 21])

1
+ r(7)r( -3 - 21]) + .... - 1
r(3 - 21])
J
and

A = 4-
nI-a-;oo dl] 1 {
-
1+21]
+
2i -a +;00 sin 2 nl]· r(3 + 21]) r(3 - 21]) (1 + 1]) (2 + 1])

1 + 21] 1 }
+ 12(1 + 1])(2 + 1])r(1 _ 21]) [1 + TI(3 + 21])(1 + 1])] .
After calculation of the residues at points I] = 0, 1, 2, ... we obtain
13 {1 (() 1 1
A = - 45 + 4 12k~1 4k(k + W(2 + k) [1 + TI(3 + 2k)(1 + k)] -

-"8
1
J2
00
k(1 - k 2 )(1 - 2k)(1
1
+ k)(2 + k) ~
}
-0.27.
Chapter 5

.
Functional Integral Techniques In
Quantum Field Theory

In recent years, the investigation and application of stochastic processes and


quantized fields has concentrated on the different functionals of these processes
and fields. Such objects play, and it seems, will continue to play, a fundamental
role in the development of mathematical and physical theories. Substantial
success in the applications of functional methods has been obtained in quantum
field theory. In this chapter the main attention is focussed on this problem,
which was considered in more detail by Simon (1979); Glimm and Jaffe (1981);
especially in the cases of gauge fields by Nash (1978); Faddeev and Slavnov
(1980); and of stochastic fields by Langouche et al. (1979).
The study of the functionals of stochastic processes forms a new field uniting
the ptobability theory with mathematical analysis, that may conditionally be
called probability analysis. In Section 6.12 (Chapter 6) we discuss briefly some
problems related to the functionals of stochastic processes, and present an
elementary stochastic calculus. Recently there has been an increasing interest
in functional integral formulations of diffusion processes (see Leschke and
Schmutz, 1977 and references therein) and in the powerful synthesis of
functional path integration, Euclidean methods and stochastic methods (see
Guerra, 1981).
Before passing to consideration of the functional methods in quantum field
theory, we expound some mathematical problems of functional analysis. At
present, these problems are currently under intensive investigation, but a strict
mathematical foundation (as in the case of ordinary analysis) is not yet
available, so further careful investigation is required.

5.1. Mathematical Preliminaries

The mathematical methods which provide the foundation for functional


analysis are the idea of differentiation with respect to functions [or (quantized)
fields, or random processes] rather than with respect to real or complex
numbers; and integration with respect to a measure defined on a space of
functions (or paths and fields) rather than with respect to the Riemann or

157
158 Chapter 5

Lebesque measure used for real and complex numbers. These methods are
called functional differentiation and functional integration.
The rule of definition of functional differentiation and functional integration
depends on the properties of functions (or fields and processes) over which the
functional operations are carried out. An essential difference arises when
functionals depend on Grassmann variables. Therefore, in concrete applications
of functional methods one has often to choose a suitable method of defining
the functional operations for the object under consideration. We first give an
elementary definition of functional differentiation and the basic idea of func-
tional integration starting from physical intuition. Next we give a more
formal definition of the functional operations appropriate to quantum theory.
This analysis leads to finite- and infinite-Grassmann algebras. Application of
the method of functional analysis in the S-matrix theory will be considered in
Section 5.5.
We begin with functional (or variational) differentiation. We have to go
beyond the ordinary space of functions f(x) and to include generalized
functions [in particular the Dirac delta function b(x - x')], in order to define the
functional derivative. The Dirac b(x) function plays an important role in
functional calculus (see Section 1.4). The variational derivative of a function
f(x) with respect to itself is defined in terms of the delta function
bf(x)/bf(y) = b(x - y). (5.1)
Then we extend this definition in the case of a general functional. Let F(f(x))
be a functional of f(x), for example, we might have

F(f) = ff(X) dx or F(f(x» = f(x o), Xo fixed.

DEFINITION 5.1. The formal definition of the functional derivative (or


variational derivative) of F(f) with respect to f(y) is written bF /bf(y), where

b~~) = ~~G-l{F(f(x) + G·b(x - y)) - F(f(x))},

here G is a real positive number. If F(f(x» = f(x o), then

bff(x(y0) = limG- 1 [f(x o ) + G·b(x o - y) - f(x o)] = b(xo - y)


b ) .~O

which corresponds to definition (5.1). When F(f) = Jdxf(x),


~;~: = !~G-l{f dx[f(x) + G·b(x - y) - f(X)]} f
= dx b(x - y) = 1.
However, this result is easily obtained if we make use of (5.1) to write

bF(f) _ fd bf(x) - fd
bf(y) - x bf(y) - x
~(- y) --
u X
1
.
Functional Integral Techniques in Quantum Field Theory lS9

We may consider another example which is provided by the kernel term


occurring in a typical linear integral equation, i.e.,

F x(f) = Idx' K(x, x')f(x'),

where x is a parameter and not an independent variable. Then by (S.l) we have


C5F x(f)/C5f(y) = K(x, y). In the case of the functional derivative some rules of
ordinary differentiation are still valid. Let E = E(F) and F(f(x)), then the chain
rule which we are looking for reads

C5E _
C5f(y) -
Id ,[x
C5E ][C5F(X')]
C5F(x') C5f(y)·
(S.2)

IS
For example, let E 1 (F) = F and F(x) = F(f) = dX 1 dx 2 f(x 1 )f(x 2 ). Making

I
use of rule (S.2), we find that C5E(x)/C5F (x') = C5(x - x') and

C5F(f) (S.3)
C5f(y) = 2 dzf(z);

so the chain rule-tells us that

C5E(x)
C5f(y) = I
2 dx' C5(x - x')
,.
I dzf(z) = I
2 dzf(z).

...J
It is convenient to compute C5EdC5f(y) directly by using equality (S.l) and the
fact that

E 1 (f) = II dX 1 dxd(xdf(x 2 )·

Thus we have just shown that


C5F/C5f(y) = 1 and C5F 2 /C5f(y) = 2F
in the case of the functional F = Sdxf(x). It is easy to see that
C5F"/C5f(y) = nF" - 1.
In practical applications (particularly in the S-matrix representation) one
encounters functionals which can be represented in the following way:

Such a functional P(f) is called the functional power series in f According to


the above, we have
P(O) = Ko(x), C5P(f)!C5f(y) = K 1 (x, y) at f = 0,
and
160 Chapter 5

We see that if K 2(x,Yt>Y2) is symmetric in Y1 and Y2, then equation (5.4) has
the desired connection between the coefficient function and the functional
derivative. This means that the general statement is that

(5.5)

provided Kn(x, Yt> . .. , Yn) is symmetric under permutation of the variables


Y 1, ... , Yn' In the case of quantum field theory we shall consider such func-
tional power series below.
DEFINITION 5.2. Now we define an alternative method for defining the
functional derivative bFfbf(y) by which the functional Taylor series may be
easily obtained formally. The definition for the derivative is

fdYf'(y) bF~)
bf (y)
= lim A -1 [F(f
A~O
+ Af') - F(f)].

This definition gives the same result as the previous one. On the other hand,
this definition resembles the definition of the variational derivative of a
functional F(f), wherefbelongs to some real Hilbert space.Yt'. In this case, we
shall say that the functional F(f) is called differentiation at a pointfEJ't, if for
any hE J'f there exists an element bF fbf E J'f such that
IF(f + h) - F(f) - (h, bF fbf) I = o(llh II), (5.6)
bF f bf is called the variational derivative of the functional F at the point f Here

Ilhll = (f lh l2 dx)1/2,
is the norm of the function h, and the scalar product (h,f) is given by the
standard formula

(h,f) = fh(x)J(x) dx.


We now suppose that the ordinary function g(A) of a real variable A defined
by g(A) = F(f + Af'), has a Taylor series:

1
L
CD
g(A) = 1 g(n) (OW, (5.7)
n= 0 n.
then

g(1)(O) = l~~ A-1[F(f + Af') - F(f)] = f


dyf'(y) ~;~~;.
Functional Integral Techniques in Quantum Field Theory 161

By induction, it is easy to prove that

g(n)(o) =
f dx 1 ···
f dxnf'(x 1 )··
(5n F
-r(Xn) (5!'(X 1 ) ..• (5f'(xn)"

Therefore the power series (S.7) acquires the form

F(f + Af) = n~o


, 00 f dx 1 ···
' ,(X
fdxnf 1
,[ (5" F(f) JAn
)··"f (Xn) (5f'(xd··· (5f'(Xn) n!·

Now suppose that this power series converges when A = 1 and then set A = 1
andf = 0, and delete the prime on the functionf'(x). The result is the desired
functional Taylor series

(S.8)

Thus, F(f) has the functional Taylor series (S.8) when g(A) = F(f + At') has an
ordinary Taylor series convergent at the point A = 1.

S.2. Historical Background of Path Integrals

Functional integrals in the sense of integrals over a space of functions were


first considered by mathematician N. Wiener (1923). The first appearance of
the method of functional integration in quantum theory was due to Feynman
(see Feynman and Hibbs, 1965). In quantum theory, the term path integral is
often used interchangeably with the term functional integral. This is because the
integrals considered by Feynman can be described as integrals over a space of
functions each of which describes the path of a particle from point a to point b.
However, the functional integrals constructed by Wiener and those considered
in quantum theory are not quite the same.
In order to compare the two kinds of integrals consider first those due to
Wiener. They have their origin in the theory of Brownian motion. That is to
say, a particle diffuses at random so that its probability distribution is given by
the diffusion equation (see Part II)
(S.9)
with
P =(4nvt)-1/2 exp(-x 2 /4vt).

We would like to construct an averaging process of the posItIOns of the


particle. Let the position of the particle at time t be x(t) E IR 1. Divide the time
interval [0, t] into n equal parts by the points t 1, ... , tn _ 1, tn (tn = t). Let the
position of the particle x(tJ at time ti be in the interval [ai' bJ for
i = 1,2, ... ,n. Then the probability of this distribution of positions is given by
162 Chapter 5

the n-fold multiple integral (for algebraic convenience we have here set v = t)
I bdX Ib2 dx z '" Ibn dxnx
1
[2"n"tl(tZ-tl)···(tn-tn_l)]-l/Z 1
al U2 Un

(5.10)

here a l < bI> a z < bz, ... ,an < bn; 0 < tl < t z < .,. < tn' n ~ 1. This defi-

i
nition is correct because of

-
(2nt) l/Z exp [-(b2t- a)Z] =
n;l'
dc[2n(t - s)] -l/Z (2ns) -l/Z x

x exp{-(b - c)Z _ (c - a)Z}, (t > s).


2(t - s) 2c
The family of the variables {x(t): t > O} is Gaussian, and
lE{x(t 1 )x(tz)} = min(tl, t z ),
where the symbol IE denotes the (mathematical) expectation value (see Chapter
6).
Now we pass to the limit, where M = (tin) ~ 0, and we insert in the
integrand of (5.10) a function F(x 1 , ... , x n). Then we write

IEw{F(x(t))} = f dflw F(x(t))

= lim
n~ CJ)
(2n·~t)-n/z fdX 1 ... fdXnF(Xl,"" xn)x

Z l n"- 1(Xi + 1 - Xi )Z}


exp {- - - L.
- X (5.11)
X .
2M i=l 2~t

The symbol dflw stands for Wiener measure. We rewrite the right-hand side of
(5.11) as

ff
A1 ... F(x(t)) exp {-"2 Jor
1 1
dt' [dX(t')]Z}
~ IJ dx(t).
t (5.12)

This enables us to give a symbolic representation of the Wiener measure

dflw = A1 exp{-"2
I
Jo e [dX(t')]Z}
dt' ~ IJ dx(t).
t (5.13)

The factor l/A is included to normalize the measure so that dflw = 1. Notice J
that the integral (5.11) of F(x(t)) can be defined only in terms of this limiting
procedure defined on F(xl>"" x n) if F(x l , ... , x n) is well-defined itself.
Functional Integral Techniques in Quantum Field Theory 163

For example, we now consider the Wiener integral


lEw {exp[ - J6dt' U(x(t'))]}. Let U(x) be a continuous function defined on
the interval (- 00, (0) and U(x) > 0. Then the question arises: does the integ-
ral exist?
Since the functional under the sign of integral is bounded it is sufficient to
show that it is measurable. Since

Joedt' U(x(t')) = n
}~~ At k~ 1 U(x(k At)), At = tin (5.14)

[we recall that x(t) E ~1, and that we have assumed continuity of U(x)], and
the functional At I:Z = 1 U(x(k At)) is obviously measurable, then the measur-
ability of the functional J~ dt'U(x(t')) and hence of the expression
exp{ - J~ dt' U(x(t'))} is proved.
From equality (5.14) and according to the theorem about the convergence of
a bounded sequence of functions (see, for example, Titchmarsh, 1939) it follows
that

lE{ex p [ -1 dt'U(X(t'))]} = }~~ lE{exp [ -At kt1 U(X(kAt))]}.


and here the existence of the limit in the right-hand side is proved at the same
time.
On the other hand, from (5.11) it is easy to see that

IE{ex{ - At k t 1 U(x(k At))]} = f:'" -f:",


dx 1 •• dX n ex p [ - At k t 1 U(Xk)] x

where
P(Xi' x j ; i1t) = (2n At) -1/2 exp{ - (Xi - xY 12 i1t}, (At = tin).
From this we obtain the following result: the limit

}~~ f:", dX 1 ••• f:", dX n exp { -At kt1 U(Xk)} X

x p(O, Xl; At)P(X1' X2; At)·· . p(X n _ 1, Xn ; At) (5.15)


exists and is equal to

IE{ex{ - f dt' U(X(t'))]} (5.16)

Existence of the limit (5.15) follows here simply from the measurability of the
functional.
One can try to define the integral (5.16) as a limit of integrals (5.15); this
164 Chapter 5

procedure was indeed used by Feynman. The defects of such an approach are
obvious from the mathematical point of view. In the case of quantum field
theory, some definition of integrals of type (S.lS) will given in Section S.S.2.
Now we want to describe the integrals introduced by Feynman (1948).
Feynman wrote the integral (S.1S) in the form (xo = 0)

(2n.,1t)- n,2 f OO

-00
dx 1 ···f
-00
oo
dxnexp{-M[i
k=1
U(Xk)+~ I (Xk-Xk_1)2j}.
2 k=1,1t
From which it follows that in the exponential there stands nothing but the
integral sum for the integral - Sh dt'[t(dx/dt')2 + U(x(t'))]' Instead of (S.16)

r
Feynman writes

fex p { - fI~(:: + U(X(t'))jdt'}d (paths). (S.17)

This symbolic notation is physically more appealing since t(dx/dt)2 + U(x), is


the Hamilton function of a particle with mass m = 1, moving in the potential
field U(x).
Indeed, Feynman (1948) in his approach to nonrelativistic quantum mech-

r-
anics arrived at the integral (see Section S.S also)

f exp{ih -1 f I (:: ~ U(X(t'))}t'} d (paths), (S.18)

r-
where h is Planck's constant divided by 2n, and

fI~(:;, U(X(t'))}t'

is the classical action along the path x(t). In other words, the integral can be
written in the form

f exp[ - hi Jor dt'U(x(t')) j exp{i2h Jor dt' (dX(t'))2}


t
~
t
TI dx(t). (S.19)

Equation (S.19) would appear to be the integral over Wiener measure of the
quantity exp[ - ih -1 Sh dt'U(x(t'))]. However, actually this is not so, because of
the presence of the factor i/h in the exponential. The factor h is not important;
its presence is simply the matter of units. But the factor i is of fundamental
importance. This factor corresponds to a diffusion with diffusion coefficient v
which is purely imaginary. That is to say, the diffusion equation (S.9) is
replaced by

i otf; o2tf;
hat = v ox 2' (U(x) = 0).

This equation is recognized straight away as the Schrodinger equation. Because


of i = J=1 occurring in the exponential, one cannot strictly apply the
Functional Integral Techniques in Quantum Field Theory 165

Feynman path integral. Another approach to Feynman type integrals will be


presented by us in Section 9.2 of the second part of the book. Recently, very
delicate and useful new definitions for the Feynman path integrals have been
given by DeWitt-Morette (1972); Albeverio and H0egh-Krohn (1976), and
DeWitt-Morette et al. (1979) (see also Exner, 1984).
Before using the functional integral techniques in field theory, we consider
analysis on a finite- and infinite-dimensional Grassmann algebra.

5.3.. Analysis on a Finite-Dimensional Grassmann Algebra

5.3.1. Definition
A Grassmann algebra with n generators is an algebra with generators C i ,
n = 1, ... , n obeying the following relation:
(5.20)
In particular, ct = o. We denote the Grassmann algebra with n generators as
Gn • It follows from (5.20) that G n as a linear space has the dimension 2n. As a
basis in Gn , it is convenient to consider the following monomials:
(5.21)
The monomial C i 1 ..• CiP will be called monomial of the order p.
Each element f(C) of the algebra Gn can be represented in the form of a
linear combination of monomials

f(C) = fo + IJ1(k)C k + IJ2(k 1, k 2)Ckl Ck2 + ... +


k ki

+ I fn(kl> ... , kn)Ckl ••• Ckn • (5.22)


ki

The element of the type I.k.!,p(k


, 1 , • •• , k p) C k 1 ••• C k p will be called homogeneous
of the order p. It is assumed that the coefficients fn(k 1, . .. ,kn) are skew-
symmetric, i.e., changing sign with respect to rearrangement of any pairs of the
arguments k 1 , ..• , kn • This property of the coefficient functions makes the
expansion (5.22) synonymous. It is obvious that each element of Gn may be
synonymously represented in the form f = l' + 1", where l' E G~ and 1" E G~.
Elements1" and l' of sets G ~ and G ~ consist of linear combinations of monomials
of even

1" = fo + I f2(kl' k 2) Ck1 Ck2 + ...


ki

and of odd
f' = Ifl(k)C k + If3(k 1, k2' k 3)Ckl Ck2 Ck3 + ...
k ki

orders, respectively.
166 Chapter 5

5.3.2. Derivatives
N ow we proceed to define derivatives of elements of Grassmann algebras.
Unlike the case of ordinary functions, it turns out that, since the elements of a
Grassmann algebra do not commute, there are two sorts' of derivative. They
are called the left and right derivative and their difference becomes important
only when we are going to differentiate the products of generators. The
derivatives are linear and so we have to specify only the derivatives of
generators and their products. The left derivative of a product Ci1 ••• C;". of
generators is defined by

+ (_l)m-l (jjirn Ci1 '" C irn _ 1 (5.23a)


and the right derivative is given by

(5.23b)

We can express these definitions verbally by saying: in order to compute the


left (right) derivative of product {C;,}, s = 1, ... , m with respect to Cj we have
to commute each Cis to the furthest left (furthest right) place in the product
Ci1 ••• C irn and then to replace simply Cis by (jisj with the appropriate power of
-1. If the monomial Ci1 ••• C irn does not contain C j , both derivatives are zero.
The rules of differentiation for more general functions follow immediately
from the definition of derivatives of monomials. We illustrate this by simple
examples:
(1) Let

f(C) = f(C(y»·
p

Then

-f-
yp
f(C(y» = 2::f -
k t a~ k
f(C)]
C - C(y)
akp '

a ]
f(C(Y»-aa = I[f(C) ac akp '
Yp k k C- C(y)

(2) Let t be a real parameter and

Ck(t) = I akP(t)yp'
p
Functional Integral Techniques in Quantum Field Theory 167

Then

d
d/(C(t)) = -t~d~
Cit aC
aI = -t~(a
k I
)d~
Cit· aCk

element, and 12
Notice also the formula differentiating products. Let
be an arbitrary element of Gn , then
11 E G~ be an event

a~p (fd2) (a~/l )12 + 11 (a~/2}


=

(fd2) a~p = 11(12 a~J + (11 a~JI2.


If 11
E G~ is an odd element and
acquire the form
12 E Gn is an arbitrary element, the formulas

a~p (fd2) (a~pI1 )12 -11(a~/2}


=

(f21d a~p = 12(11 a~J -(12 a~Jft.


Let us notice finally the following properties of repeated derivatives:

a
( aC
1
I) a _ a
aC 2 - aC 1
(I a )
ac 2 ·

5.3.3. Integration over a Grassmann Algebra (Finite-Dimensional Case)


We shall present the definitions and show how they give rise to a satisfactory
integration calculus. A volume element is written as dC 1 ••• dC;, 1 -< i -< n,
where C; and dC; must satisfy
{dC;, dCd = {C k , dC;} = O. (5.24)
The (formal) integral is then defined by setting

(5.25)

Multiple integrals will be understood as repeated ones. Thus, formulas (5.24)


and (5.25) define the integral f 1(c) dC n ••• dC 1 on all monomials.
DEFINITION 5.3. The integrals defined by formulas (5.24) and (5.25) are
called the integrals over a Grassmann algebra with generators C 1, ... , Cn. It
168 Chapter 5

follows from the definition of the integral that for any element f( C) the formula

f f(C) dCm··· dC l = m! fm(1, ... , m), (1"';;;; m ., ;;;; n) (5.26)

holds, where fm(1, . .. ,m) is assumed to be anti symmetric under interchange of


the indices 1,2, ... ,m. As an illustration, take m = 3. Then the first formula of
(5.24) tells that

ff(C)dC 3 dCz dC l =f3(m l ,m Z ,m3) fCmlCm2Cm3dC3dCzdCl'

but
f3(m l , mz, m 3)Cm1 C m2 C m3
= f3(1, 2, 3)C l C z C 3 + f3(1, 3, 2)C l C 3 C z + f3(2, 3, 1)Cz C 3 C l +
+ f3(2, 1, 3)C z C l C 3 + f3(3, 1, 2)C 3 C l C z + f3(3, 2, 1)C 3 C z C l
= 6f3(1, 2, 3)C l C z C 3·
Hence

These examples should be sufficient to make the handling of integrals a quite


straightforward matter.
As in ordinary analysis, in the case of the Grassmann algebra there are
formulas of partial integration

ff(C)(a~p g(C))dCn···dC l = f[f(C) a~Jg(C)dCn.·.dCl'

f f(C{g(C) a~J dC n ·•• dC l = (-1)" + 1 f[ a~p f(C)] g(C) dC n •·• dCl·

For proof, it is sufficient to consider f and g as arbitrary monomials.


A useful formula is valid for the change of variables in integrals over the
Grassmann algebra. We restrict ourselves to the simple case when the change is
a linear one.

dC i = I iiik de k ,
k

where II iiik II is an inverse matrix to II aik II. The following formula is valid for
such a change of variables:

f f(C) dCn ... dC l = detll iiik II f f(C(e)) den· .. del" (5.27)


Functional Integral Techniques in Quantum Field Theory 169

Notice that in distinction to the usual formula for the change of variables, the
independent variables and differentials transform by means of mutually inverse
matrices. Now we shall make this more transparent. Let f(x I, ... ,xn ) be a
function of n real variables. Define the integral I of f by

I = f f(x) dXI ... dx n·

If a = aik is a n x n matrix, we can also write, x = ay, Xi = II aik IIYk'

I = detllaidlff(aY)dYI" 'dYn, deta =1= 0).

However, consider the integral over the Grassmann algebra

I = f g(C) dC n '" dCI,

and write the linear change of variables in the Grassmann algebra as C = Ae,
then we find that

I = (det A)-l f g(Ae) den'" del'

The 'Jacobian' is the inverse of what we are used too. The reason for this is
easy to find. We have CI"'Cn=(detA)el"'en but dCn"'dC I =
(detA)-1 den" 'de l so that SCidC i = 1 may be maintained. As an example,
we calculate the 'Gaussian integral'

I = fexP{taikCiCk}dCn· .. dCI, aik = -aki .

Let first II aik II be a real matrix. Applying a proper orthogonal transformation


II Sik II, the matrix II aik II can be brought into the form
o 0
o 0
o A2 0

Perform the change of variables Ci = Lk Sik ek in the integral I. Owing to


formula (5.27), we get

I = f exp{2[AI e l e2 + A2 e 3 e4 + ... + An/2 en - I enJ} den' .. del


170 Chapter 5

for even nand

1= f exp{2[A1 e 1e2 + ... + A(n - 1)/2 en - 2 en - 1]} den· .. del


for odd n, respectively (detlls;kll = 1 since Ils;kll is a proper orthogonal matrix).
Using formula (5.27), we find that for even n
1= 2n/2 A1 ... An = [det(112a;kll)]1/2,
and for odd n, I = o. Since for odd n, detlla;kll = 0, both these cases fall under
the following rule

f exp[t a;kC;Ck] dC n··· dC 1 = (detI12a;kIl)1/2. (5.28)

Formula (5.28) is valid not only for real matrices but also for complex ones.
DEFINITION 5.4. An algebra Kn with generators, k 1 , ••• , k n obeying the
following relations:
k;kj + kjk; == {k;, kj} = 0
for i =l= j is called Clifford or spinor algebra with n generators. A Clifford
algebra K 2n with a doubled number of generators is closely associated with the
Grassmann algebra Gn"
We recall that Bose systems with n degrees of freedom can be described by
means of the Hilbert space of analytic functions f(z 1, ... , Zn) = f(z) of variables
Z1> .. . , Zn with the scalar product

(fd2) = (2ni)-n f f dZ1 dz1 ··· f f dZ n dZn ex p ( - ~ Zk Zk)f1(Z)h(z).


Coordinate and momentum operators are realized as follows

It is easy to see that these operators satisfy the anti-commutation relations

Thus, the operators Pj' Qj turn out to be generators of the Clifford algebra K 2n .
At this stage, the operators Zk and O/OZk can be identified with the operators in
Gn , that is those of the left multiplication with the generator Zk and the left
derivative O/OZk with respect to Zk in Gn.
In conclusion, let us write the linear functionals and linear operators in Gn in
the following form

F(f)= ff(C)FrighJC)dCn···dC1 or F(f)= fFleft(c)f(C)dCn···dC1 (5.29)


Functional Integral Techniques in Quantum Field Theory 171

and

or

respectively. Here Fright and Fleft are fixed elements of Gn , and Kright and KIeft are
elements of G 2n ; C i , ek are generators in G2n [see Berezin, 1966 for the proof of
formulas (5.29) and (5.30)].

5.4. Grassmann Algebra with an Infinite Number of Generators

5.4.1. Definition
For use in field theories with fermions, we require an infinite-dimensional
algebra G. In quantum theory, a Grassmann algebra equipped with a scalar
product and with an involution are of particular importance. A strict mathemat-
ical definition of the algebra G and an analysis of it were given by Berezin
(1966) using the notion of a direct sum of a denumerable number of topological
linear spaces. Concrete applications of the analysis on an algebra of antic om-
muting variables will be considered in Section 5.5.
DEFINITION 5.5. As a Grassmann algebra G, we shall understand a direct
sum of spaces En possessing the following properties:
(1) EO is one-dimensional space with fixed basis element fo.
(2) For any f E EP, 9 E Eq, a product fg E EP + q is defined; and
(a) if f = exfo E EO, ex is a complex number, then fg = exg,
(b) fg = ( -1)Pq gf,
(c) if f1' ... ,In and 9 1, ... ,gm are linear independent elements of EP and
Eq, respectively, then /;gk(i = 1, ... , n; k = 1, ... , m) are linear inde-
pendent elements of EP + q,
(d) finite linear combinations of elements of the type fg, f E EP, 9 E Eq
generate a dense set in EP + q.
(3) The space G is closed with respect to a multiplication which is defined in
the following way if
f = fo + f1 + ... + In + .. " In E En,
9 = go + 9 1 + ... + gn + ... , g~ E En,
then
fg = h = ho + h 1 + ... + hn + ...
172 Chapter 5

where
n

hn = I !,.gn-k·
k=O

Every element In E En is represented in the form

In = f C(xd· .. C(xn)<p(X I ,· .. , Xn) dX I ... dx n, (5.31)

where <p(x I , ... , xn) is, generally speaking, a generalized function. Notice that
in the infinite-dimensional case we have labelled the anti-commuting variables
(or generators) by a continuous index, as C(x).
According to the definition of the Grassmann algebra G, each element
belonging to G may be represented by the form
f = fo + fl + ... + In + ... , fn E En.
Using the generic representation (5.31) for In, we get the following formula for
any element f E G:

f = f(C) = Co f
+ fdxC(x) <p(x) + fdX I dxzC(xl)C(xz)<p(x l , Xz) + ...

~o f
n f dX I ... dXnC(X I)· .. C(Xn)<P(XI' ... , x n ), (5.32)

where <p(x I, ... ,xn) are skew-symmetric (generalized) functions of n variables.


The elements C(x) we will call generators of the algebra G. They playa role
analogous to the generators in the finite-dimensional algebras.
Notice that formula (5.32) recalls a representation of an analytical func-
tional. In this connection, the elements f = f(C) of the Grassmann algebra G
with scalar product (f,f) can be considered as functionals of anti-commuting
variables (or generators).
We now consider two elements fl and f~ belonging to E1, i.e.,

fl = f C(X)<PI(X) dx and 1'1 = f C(x)<p'(x) dx,

and write their product

fil'l = f f
dx C(x)<p I (x) dyC(y)<p'(y).

Defining a product C(x)C(y) of generalized generators C(x) and C(x), we get

fJ~ = f
dxC(x)<pI(x) f dyC(y)<p'(y) = ffdx dyC(x)C(Y)<PI(x}<p'(y).
Functional Integral Techniques in Quantum Field Theory 173

According to the definition of the Grassmann algebra

It is evident that these equalities are equivalent to the relations


C(x )C(y) = - C(y )C(X), C(X)[C(y)C(z)] = [C(X)C(y)]C(z). (5.33)

5.4.2. Grassmann Algebra with Involution


Let G be a Grassmann algebra with scalar product in which mutually
homomorphous f ~ f* mapping onto itself is defined and which obeys the
following conditions:
(a) (f*)* = f,
(b) (fd2)* = fi ft,
(c) (f3f)* = fJf* (13 is a complex number).
(d) If for f, g, a scalar product is defined, then it is also defined for f*, g* and
they are related by (f*, g*) = (g, f).
(e) The space El is decomposable in direct sum El = F + F*. The mappingf*
satisfying these conditions is called involution in G.
Let us consider an orthonormalized basis C(x) in the space F with the scalar
product defined as (C(x), C(y)) = c5(x - y) where c5(x - y) is the delta function
Sdy<p(y) c5(x - y) = <p(x). In the space F*, we introduce a conjugate basis
C*(x), and extend the definition of the involution to linear functionals:

f* = f C*(x)cp(x) dx for f = f C(x)<p(x) dx.

The union of these bases is obviously the basis in El. As any basis in El,
{C, C*} is a system of generators of the algebra G. By means of the generators
C(x), C*(x), all elements of G can be written in the form:

f = I f<pmn(Xt. ... , XmIYl,'" ,Yn)C*(xd'" C*(x m) X


m,n

(5.34)
It is evident that the individual monomials in this sum are elements of the
spaces Emn. It is easy to see that if f has the form (5.34), then f* has the
following form

f* = I fcpmn(Xl, ... ,XmIYl, ... ,Yn)X


m,n
x C*(Yn)'" c(Ydc(x m)'" C(xd dmx dny.
We assume that the functions <Pmn are skew-symmetric separately with respect
to Xl"'" xm and Yl,'" ,Yn'
174 Chapter 5

5.4.3. Functional (or Variational) Derivatives


Let G be a Grassmann algebra with scalar product, and C(x) is the system of
generators of G. We define the left and right variational derivatives of the
element f E G with respect to C(x). We shall denote the left and right va-
riational derivatives of the element f by bf/bC(x) and f b/bC(x), respectively.
Both the derivatives are linear operators in G, hence it is sufficient to consider
their action on the monomials f" E En. First define the derivatives of a product
of generators:
b
bC(x) C(x 1 )· •• C(x n) = b(x - X1 )C(X2)·· . C(x n) -

- b(x - x 2 )C(xdC(x 3 )·· • C(x n) +

+ ... + (_1)n - 1 b(x - Xn)C(X 1 )· •• C(X n _ d, (5.35)


b
C( X1)· .. C(xn) bC(x) = b(x - Xn)C(X1)· .. C(x n - 1) - b(x - xn - d x

x C(x 1 ) · · · C(x n _ 2)C(X n) +

+ ... + (_1)n - 1 b(x - X1 )C(X 2)·· . C(X n).


The derivatives of products of the generators have an obvious meaning as
generalized functions; differentiating formally under the integral sign and using
formulas (5.35), we find that if

f = f f
dX 1 ... dXn cp(x 1 ,·· ., x n)C(x 1 )· •• C(x n),

then

(5.36)

f b~(X) = n f f
dX 1 ... dXn - 1 cp(x 1 ,· •• , Xn - 1, x)c(xd· .. C(Xn- 1)·

Here one has to take into account that the function cp(x 1, ... ,xn) is skew-
symmetric.
The derivatives bf /bC(x) and f b/bC(x) of general elements f E G do not
necessarily belong to G. One can consider them naturally as generalized
functions, i.e., instead of bf/bC(x) and f b/bC(x) one can consider the following
expressions:
Functional Integral Techniques in Quantum Field Theory 175

where cp(x) belongs to some set of test functions. In fact, it may happen that
there exists some function cp E E 1 (M), x EM such that

fdxcp(x)[(jfl(jC(x)] E G.

In this case, the element f will be called left-differentiable over the functions cpo
Similarly one can speak of right-differentiable elements over cpo If, however
Sdxcp(x) (jfl(jC(x) does not belong to G for some function cp(x) E E1(M) except
cp(x) == 0, the element f is called nondifferentiable from the left. Similarly one
understands an element which is nondifferentiable with respect to the right. If
the element f is differentiable over all functions cp belonging to some set
DE E1(M), it is called D-differentiable function.
Further we notice that from the definition of variational derivatives, the
following property can be deduced. Let us consider an algebra G with
generators C(x) and B(x). Let F(C)E G. Expanding F(C + B) in powers of B,
the coefficients of B in the first power are obtained as the left and right
derivatives of F;

F(C + B) = F(C) + f
B(x) (j~(X) F(C) dx + .. .

= F(C) + f
(F (j~X))B(X) dx + .. . (5.37)

This property may be used for the definition of variational derivatives as in the
case of the commuting variables [see formula (5.6)].
Let us consider a rule for change of variables under differentiation. Let now
f E G. The following formulas for change of variables under differentiation hold
(j
f (jC(x) =
f (j
dYK(x,Y)f(jB(y) ,

when the generators C(x) and B(x) are connected by the relation:

f dxC(x)K(x, y) = B(y).
Here the kernel K(x, y) is the continuous analog of a transition matrix from
system C(x) to B(y), and can be interpreted as a generalized function on some
space of test functions of x, y.

5.4.4. Continual (or Functional) Integrals over the Grassmann Algebra


(Formal Definition)
Let G be the Grassmann algebra with involution, and let E1 = F + F* be the
decomposition of the space E1 in definition of the algebra with involution. Let
176 Chapter 5

us embed a system of nested finite-dimensional subs paces F n with dimension


2n to the space F, such that their closure coincides with F: UFn = F. Denote
through GL a sub-algebra G obtained from the space Ln = Fn + F;. We
introduce a~ orthonormalized basis Cu ... , C 2n in each Fn in an arbitrary way,
and in F; the basis formed by the conjugate elements C!, ... , C!n. Take now
an arbitrary element f(C*, C) from G. Consider its value on GL . Denote
this value asf,.(C*,C) and consider the integral "

In = (f) = ff,.(C*, C) dqn dC 2n ··· dq dCl· (5.38)

It is easy to show that the integral (5.38) does not depend on the choice of the
orthonormalized basis in Fn. Indeed, let {B k } be another orthonormalized
basis in Fn, and {Bn the corresponding basis in F;. Then Bk is expressed by
Ck by means of a unitary matrix: Bk = Ls Uks Cs. It is clear that Bt is related to
Ct through the complex conjugated matrix: Bt = Ls uks C;. Using formula
(5.27) we find that

f f,.(B*, B) dB!n dB 2n ··· dB! dB l = det(u- l u- l )f fn(C*, C) x


xdC!ndC2n··· de! dCl.
Since u = II U ks II is unitary, det(u -1 U-1) = 1, the equality of the integrals is
proved.
DEFINITION 5.6. If there exists the limit: I(f) = limk~ 00 Ik(fd independent
of the chosen system of subspaces Fn, then this limit is called the continual (or
functional) integral of f on the algebra G. The continual integral over
Grassmann algebra in a physical context is called the Fermion continual
integral. We shall denote it as in the case of the usual functional integral

I(f) = ff(C*, C) Il dC* de. (5.39)

N ow we explain the rule of partial integration for this case. Let fl J2 E G be


two D-differentiable elements. Then the integral

(5.40)

has an obvious sense of a generalized function on D:

f dXqJ(X){f f{5~(X) f2) Il dC* dC} = f fl{fc5~(X) f2qJ(X) dX} Il dC* dC.
Then the following formulas for integrating by parts hold:

f fl(c5~(X) f2) Il dC* dC = f(fl c5~(x))f2 Il dC* dC,


Functional Integral Techniques in Quantum Field Theory 177

f f{f2 b~X)) n dC* dC = - f(b~(X) f1 )f2 n dC* dC,

f f1(bC:(X) f2) n dC* dC = f(f1 bC:(X))f2 n dC* dC,


f f1(f2 bC:(X)) n dC* dC = - f(bC:(X) f1 )f2 n dC* dC.
5.4.5. Examples
(1) Let G be the algebra with involution, E1 = F + F*, and C(x) be the
generalized orthonormalized basis in F. We consider the operator A in E1,
which given naturally by the matrix

A = (All(X, x') Adx, XI))


A21 (x, x') A 22 (X, x') ,
We construct the element f(C*, C) of space E2 by the operator A:

f(C*, C) = f{C(X)All(X, X')C(X') + 2C(x)Adx, xl)C*(X') +


+ C*(x)Adx, xl)C*(X')} dx dx'
or shortly
f = CAllC + 2CA 12 C* + C*A 22 C*
or in the form

f = (C, C*) (All


A21
A12)
A22 C*'
(C )
In particular, if A12 = E, All = A22 = 0, thenf = 2CC*. Let us calculate the
Gaussian integral

fexP(tJ(C*, C))TI dC* de.

Making use of formula (5.26) and of the definition of continual integral, we


find

(5.41)
178 Chapter 5

The determinant on the right-hand side of (5.41) is a generalized Fredholm


determinant. Power t means branching of square root which is positive for real
positive values of the argument.
(2) From (5.41), it follows, particularly, that

Iex P ( -C*C) Il dC* dC = 1.

Therefore one can define the Grassmann measure


dflG = exp( - C*C) Il dC* de.
To conclude this section, we notice that all the three types of functionals (in
which we are interested) namely, the functionals

(5.42)
and

A(C*, C) = A(C*, C) exp{ - I C*(x)C(x) dX}, (5.43)

corresponding to representations of operators .Ii in the matrix form and the


normal form, i.e., in each term of the type (5.43), creation operators stand left
from the annihilation operators, and functionals

<I>(C*) = I ~ I<I>n(X 1 , • •. , x n)C*(x 1 )· •. C*(x n) dnx,


n vi n!

~ II <l>n 12 d 4 x < 00, (5.44)

corresponding to vectors <1>, turn out to be elements of the same Grassmann


algebra G. The scalar product of operators <I> can be given by the formula

or

(<1>1' <1>2) = I<l>HC)<I>l(C*) exp( -C*C) Il dC* dC,

where <I>!(C) is the conjugate element (functional) to <1>2 (C*). Now we have
concluded the explanation of operations on functionals in the case of
Grassmann algebra, and turn to the application of methods of functional
analysis to quantum field theory.
Functional Integral Techniques in Quantum Field Theory 179

5.5. Functional Integral and the S-Matrix Theory

5.5.1. Introduction
There exist several approaches to the quantization of physical systems (in
particular, field theory). The operator method (or the usual canonical ap-
proach) of quantization is the most often used. In this method, the operators
obeying the canonical commutation rules correspond to physical field con-
figurations (see Chapter 2). However, as we will emphasize below, there is
another approach, the so-called stochastic quantization method, the basis of
which will be expounded in Part II. In Chapter 10 we shall discuss some
problems concerning the stochastic quantization of physical fields. There is
furthermore a third approach, in which quantum dynamics is described by the
continual (or path, or functional) integral, that is of the sum over all field
configurations. By means of this approach, Feynman was the first to formulate
a consistent and explicitly relativistic invariant perturbation theory for quan-
tum electrodynamics. This formalism turns out to be most suitable for the
quantization of gauge fields. In this case, integration is carried out not over all
field configurations, but over gauge-equivalent classes alone.
Recently, a very interesting link has been discovered between the stochastic
and canonical quantization in perturbation theory for any field theories includ-
ing gauge ones (Floratos and IIiopoulos, 1983). In the latter case, equivalence
is achieved provided one considers gauge-invariant quantities (see Chapter 10).
Guerra (1981) has pointed out that the path integral theory can be clearly
reformulated by means of stochastic mechanics considered in Part II of the
book. In this section we wish to make use of functional integral techniques in
the S-matrix theory. We know that the path integral of Feynman was
introduced to evaluate the sum of contribution from all possible paths that a
particle can trace moving from a to b. We know that in classical mechanics
just one path is important, for which the action is extremal. In quantum
mechanics this is not so. Every path is now important; it is only that that path
which makes the action attain its extremal value, and neighboring paths which
give an action within one unit h around the extremal value are the most
important. In fact, if A(x(t)) is the action for a path x(t) then its amplitude is
know to be given by exp{ih- 1 A(x(t))}. Then the integral over all paths is given
by

I = f exp{ ih -1 A(x(t))} TI dx(t),

where IT dx(t) as before. The action A is the time integral of the Lagrangian 5£
where

dx)2
5£ = tm ( dt - U(x).
180 Chapter 5

From the Hamiltonian dynamics point of view, quantum field theory is a


system with an infinite number of degrees of freedom. For example, in the case
of the neutral scalar field cp(x), described by the following Lagrangian (see
Chapter 2)
5t = !cp(x)(D - m2)cp(x) - U(cp), (5.45)
the points of the phase space are pairs of functions cp(x), n(x) forming an
infinite set of canonical variables. In this case state vectors of type (5.44) are
functionals <I>(cp(x)) of cp(x). In quantum field theory the main object is the S-
matrix and the generating functional Z for Green's functions. They are func-
tionals of field cp(x), an external source I1(X) and current J(x), or of a function
of 'switching on' the interaction g(x), respectively (see Chapter 2).
Representation of types (5.8) and (5.42) for these quantities is given by the
formulas:

(5.46)

and

(5.47)

In the operator formulation, the coefficient functions Sn(x 1 , ••• , x n) in (5.46)


arise in the expansion of the operator-valued S-matrix in terms of normal
products of free fields. The coefficient functions Gn(x 1 , ••• ,xn) in the expansion
of the functional (5.47) in powers of I1(X) (external source) define the Green
functions which are the corresponding vacuum expectations of chronological
products of the Heisenberg field operators.
Now we turn to derivation of the functional integral representation for the
S-matrix and Z. In quantum field theory the action A acquires the form

A = f 5t(x) d x,
4 (5.48)

and instead of the volume element TI dx(tJ of the space of functions (paths) the
volume element TIx dcp(x) defined on the space of classical fields appears. The
integral for S becomes then, with h = 1

(5.49)

where a normalization constant is omitted. The volume element TIx dcp(x) is


usually written 8cp or gJ[ cp].
The content of quantum field theory can be characterized by the set of all its
Green's functions <OIT{cp(Xl)···CP(xn)}IO). These Green's functions can be
written compactly as functional integrals. This is accomplished by using a very
Functional Integral Techniques in Quantum Field Theory 181

useful technique due to Schwinger. The technique consists of modifying the


action A (5.48) by adding a term Jd 4xl](x)cp(x) where I](x) is not an operator
but an ordinary scalar function. The resulting functional integral is defined
as Z(I]), a functional of the source I](x),

Z(I]) = f i5cp exp{ifd 4 x[5i'(x) + I](X)CP(X)]}. (5.50)

Usually, by substituting U(cp) -> U(cp + CPin), another representation of the S-


matrix as a functional integral may be obtained. For example, in the case of a
neutral scalar field described by Lagrangian (5.45) the S-matrix is given by

The obtained formulas (5.49) or (5.51), and (5.50) are attractive in view of their
compact form and clarity. So the representation (5.50) for functional Z(I]) in
the form of an integral allows simple formulas of analysis to be used:
integration by parts, change of the order of integration; change of variables
and calculation by the method of stationary phase.
Before we turn to the definition of functional integral, we shall compute Z(I])
for the trivial case U(cp(x)) = 0 and show that Z(I]) and its functional de-
rivatives (_l)n i5 nZji5I](Xl)' .. i51](xn) give the free particle Green's function. The
task of computing Z(I]) consists entirely of manipulation designed to reduce
the argument of the exponential to a quadratic form. The kinetic term is of the
form

Then, if we define ¢' = U I2 (O)cp, the resultant term is of the form ¢' ¢'. This
task is difficult due to the fact that L(O) is a differential operator and we do
not know how to take its square root. We therefore pass to momentum space
and define the function 3,1/2 (p) by the equation
(m 2 _ p2)1/2 Kl/2(p) = 1, or 3,1/2(p) = (m2 _ p2)-1/2.

Then we introduce the Fourier transform A1/2(X) of 3,1/2(p). Evidently,

Al/2(X) = (2n)-4 fd 4p eipx (m2 _ p2)-1/2.

The change of field variable that we require is now clear, it is

cp(X) = L -1/2(O)¢'(x) = f d 4 y A1/2(X - y)¢'(y).


182 Chapter 5

Then after some elementary calculations we get

fff d 4x d 4xl d4x24>'(xl) Al/2(X - xd(D - m2) Al/2(X - x 2)4>'(X 2)

We have now achieved the result that we required, and we can now change the
variable in the functional integral from <p to 4>. This means that

Z(1]) = K f f
J4>' exp{i d 4x[ -t4>'2(X) + f
d 4x l 1](x) Al/2(X - Xd4>'(Xd]},

where K is the Fredholm determinant for N/2. The last term to be dealt with
is the source term. It is straightforward to find a change of variable to
eliminate the source term. The choice needed is

4>'(X) = 4>"(x) + fd 4x l 1](x) Al/2(X - Xl)'

With this change of variable, the argument of the exponential becomes

where

H(y) = f d4xl1](Y) Al/2(y - Xl)'

We deal with this term in the same way as above:

with

Therefore we have

~ f
d4 xH(x)H(x) = ~ ff d 4x d4Y1](x) A(x - Y)1](Y),

where
Functional Integral Techniques in Quantum Field Theory 183

is the usual scalar particle Green's function satisfying


(0 - m 2 ) d(x - y) = - b(4)(X - y).
The functional Z('1) is now seen to be

Z('1) = expG ff d 4 x d4y'1(x) d(x - Y)'1(y) J. f


K· ()(Il' x

x exp { -~f d x¢'2(X)}


4 (5.52)

J J
The factor K . b¢' exp{ - i/2 d 4 x¢'2(X)} does not depend on '1 and is simply a
number. We eliminate it by redefining the integral that we are interested in as

f f
bcp expH d4x[cp(x)L(O)cp(x) + 2'1CP] }
Z('1) =
f bcp exp { ~
.
f d 4YCP(Y)L(O)CP(Y)}

= expH ff d 4 x d 4y'1(x) d (x - Y)'1(Y)}

Note that Z(O) = 1. The functional Z('1) has a functional Taylor series obtained
by expanding the exponential. The terms in this Taylor series turn out to be
the free particle Green's functions. For example, we have
b 2Z('1)/b'1(Xl) b'1(X2)1~ = 0 = i d(Xl - x 2 )·
To summarize the situation for a free field, we write

= f bcpcp(xd' .. x

x cp(xn)exp { ~ fd4XCP(X)L(0)CP(X)} x

x {fbcp exp [ ~ f d4YCP(Y)L(O)cp(y)Jr 1

The great importance of this property for free fields is the fact that it
remains true for interacting fields. The only change is that the free action is
replaced by

AF -> A = f
d 4x[tcp(x)L(0)cp(x) - gU(cp(x))]
184 Chapter 5

and the generating functional for Green's functions is given by.

f bcp exp{i f d 4 x[tcp(x)L(O)cp(x) - 9 U(cp(x)) + 1'}cp(x)] }


Z(1'}) = "--~--=---=--:,---.-------------=----"--

fbCP exp{i f d 4 x[tcp(x)L(O)cp(x) - 9U(CP(X))]}

Expansion of this integral in powers of 9 coincides with the usual perturbation


expansion in terms of Feynman graphs.

5.5.2. Functional Integral over a Bose Field in the Case of Nonlocal-Stochastic


Theory (Definition)

5.5.2a. Definition of Functional Integral. We shall investigate here the S-


matrix written in the form of functional integral in the Euclidean metric.
Therefore all the formulas and integration are written in Euclidean space. The
so-called Wick rotation between Minkowski space and Euclidean space for
Bose (or Fermion) fields was the subject of many studies (in particular, see
Schwinger, 1959; Osterwalter and Schrader, 1973, 1975). We will examine the
theory of a one-component scalar field cp(x) described by the Lagrangian
density (2.3) (Chapter 2):
f£(x) = tcp(x)L(O)cp(x) - g[KWO)cp(x)]\ (5.53)
where ¢(x) = KWO)cp(x) is a nonlocal field resulting from averaging over
stochastic space-time JR4(X) (see Chapter 1). The operator K(l20) is nonlocal
and satisfies the conditions (see Chapters 1 and 2):
(i) K(z) is an entire analytic function of the order p ;p t in the complex z-
plane;
(ii) K(z) decreases rapidly enough when z = 12k 2 = F(k6 - k 2 ) ~ - 00;
(iii) K( - m 2 12 ) = 1;
(iv) I is the fundamental length characterizing the region of nonlocal
interaction in the given case.
In the paper by Efimov (1977b), the representation of the S-matrix as a
functional integral was extended to the case of non-polynomial interactions. In
this section, we use the representation obtained above for investigation of the
theory described by the Lagrangian (5.53).
For this purpose, the S-matrix as a functional of the scalar field CPin(X) in
Euclidean space of dimension'd is written in the form of the functional integral

S[CPin] = N- 1 f bcp exp{ -~ f ddxcp(x)(m 2 - O)cp(x) - 9 x

x f ddx[¢(x) + CPin]4 }
(5.54)
Functional Integral Techniques in Quantum Field Theory 185

or

f
- g ddx[cp(x) + CPin]4}- (5.55)

Here the function A -1 (x) is defined by the relation

f ddy A(X1 - y) A -l(y - Xl) = c5(d)(X1 - Xl),

and normalization constants Nand C in (5.54) and (5.55) are chosen so that
S[CPin] Ig = 0 = 1, where CPin is an arbitrary scalar function (field). In expression
(5.55) one can carry out the change of variables

cp(x) = f ddx' A~/l(X - XI)cpl(X' )

where the function A~/l(X) is defined by the equality

A~/l(X) == A1/l(X) = (2n)-d f ddk e- ikx A1/2(kl),

L11/l(k2) = (L1(kl))1/2 = (m 2 + kl)-l/l


so that the Euclidean Green's function AE(X) is given by

AE(x - x') = A(x - x') = f ddy A1/2(X - y) Ai/l(y - x').

After this change the S-matrix (5.55) can be represented in the form

S[CPin] = f
C- 1 c5cp ex p { -~f ddxcp 2(x) - g f
ddx(cpix) + CPin)4},

(5.56)

The S-matrix in this representation contains a volume divergence connected


with the translation in variance of our theory. Therefore we have to introduce
the integration over a finite volume V E IRd in (5.56). It is supposed further that
in the volume V there is an orthonormal system of functions {g.(x)} ,
s = 1, 2, ... , such that

L
00

g.(x)g.(x') = c5(d) (x - x').


s = 1

The volume Vand the system of functions {gs} will be chosen in the following
186 Chapter 5

way
v= {x: -L -< Xj -< L,j = 1,2, ... ,d},

(5.57)

where

(2L) -1/2 cos -n(s-- - x 1) s is odd,


fs(x) = { L 2
(2L) -1/2 sin( I ~ x) s is even.

It is assumed that any function f(x) E L 2 (V) can be expanded over this basis
00

f(x) = L fngn(x),
n~O

After these preliminary remarks, we turn to the definition of functional integral


(5.56). In the functional integral (5.56) put
00

cp(x) = L gn(x)u n·
n~O

Then
00

¢d(X) = ¢(u, x) = L D!/2(X)Un ,


n~O

where

D1/2(X) = (2n) -d f ddkK(P k 2)(m2 + k 2) -1/2 e ikx .

Here hy{x) E D(V) is infinitely differentiable and a posItIve function and


limv~ 00 h~x)= 1. The function D1/2(X) is connected with the nonlocal causal
Green'sfunction DJx) in the following way

Dv(x, x') = I ddyD~/2(x,y)D~/2(y, x'),


Functional Integral Techniques in Quantum Field Theory 187

Dc(x - x') = }~~ Dv(x, x') = fdd y Dl/2(X - y)Dl/2(y - x')

= (2n)-d fdd p eip(x -x') [KWp2W(m2 + p2)-1.

f -
The system {gs} is chosen in such a way that
2 2 2
[K(l k )]
L
00 1/2
[Dn (x)]
2
= Dv(X, x) ~ DAO) = (2n)
-d d
d k 2 k2 < 00,
n=l v~oo m +

L
00

nM[D~/2(XW < 00, VM > 0, V < 00. (5.58)


n =1
It is easy to verify that

}~~ n~o D~/2(X)D~/2(X') = f ddy Dl/2(X - y)Dl/2(y - x') = Dc(x - x').

Let us introduce the following functions:


N N

CPN(X) = L gn(X)U n, 1>(N)(U, x) = L D~/2(X)Un


n=O n=O

and define the measure

dN /lu = nN dUn Fe exp( -


1 2
2:Un), (5.59)
n = 0 V 2n

so that

Then the functional integral (5.56) acquires the form

Sv[g, CPin] = fd/l u exp{ -g 1dd y [1>(U, x) + CPin(X)]4}, (5.60)

which is defined as (CPin = cp)

Sv[g, cp] = lim S~N)[g, cp],


N~ctJ

1
(5.61)
S~N)[g,cp] = fdN/luexp{ -g ddy[1>(N)(U, X) + cp(X)]4}
where d N/lu is given by expression (5.59).
It is shown (Efimov, 1979) that the functional integral (5.60) given by the
188 Chapter 5

limit (5.61) does exist for any V < (f) and defines an analytic function in the
complex g-plane singular at the point g = O. For example, we shall show here
that the functional integral (5.60) does exist in the case of Re g > 0, i.e., the
limit in (5.61) exists for Re g > O. In this region, we have
IstN)[g, cp]1 .:;( 1 (VN> 1).
Consider further the difference
Ll~+M = .s'[,,+M)[g, cp] - .s'[")[g, cp]

- exp[ -g I ddX(4)(N)(U, x) + cp(X))4]} (5.62)

We transform this difference in the following way

Ll~ + M = f f
d¢(1 - ¢) dN + M /1" :¢22 x

= f d¢(1 - ¢) fd N+ M/1u exp{ -gqu, ¢)}x

{ -12g I ddX[t/!(N,M) (u, x)] 2 X

x [4>(N)(U, x) + ¢t/!(N,M)(U, x) + cp(X)]2 + 16g2 [1 ddxIj}N,M)(u, x) x

x (4)(N)(U,X) + ¢t/!(N,M)(U, X) + cp(X))3J}'

where
Ij}N, M) (u, x) = 4>(N + M) (u, x) - 4>(N) (u, x),

qu,x) = I dd X[4>(N)(U,X) + ¢Ij}N,M)(U,X) + cp(x)] 4.

Making use of the Holder inequalities for integrals


Functional Integral Techniques in Quantum Field Theory 189

where PI + P2 = 1, one can obtain

I.-1Z+ MI <;;; f d~(1 -~) f d N+ Mflu exp{ -Re goC(u, mx


x {12IgICI/2(U,~) + 16IgI 2C 3/2(U, m[i ddx(t/I(N,M)(u, X))4J /2.

Using further the obvious inequalities


Cexp(-RegoC)<;;; [Regr i x rr x e- r ,

one can obtain finally

I.-1Z+MI<;;;const{fdN+Mflu i ddX[t/fN,M)(U,X)]4f
/2

= const {i [ I
V
dd x
N+M
s=N+ 1
D.(x)
]2}1/2
.

Because the series l: Ds(x) converges, there exists No for any e > 0, so that for
any M > 0 and N > No, the inequality I.-1Z + MI < e is valid. This means that
the sequences 51t') [g, CPJ for V < CfJ and Re 9 > 0 is fundamental.
Thus the limit in (5.61) existso This limit defines the functional integral (5.60)
for any V < CfJ and Re 9 > O. Proof in the case of Re 9 < 0 is given by
analytical continuation of the function 51t') [g, CPJ to the region Re 9 < O. It was
done by Efimov (1979).

5.5.2b. Upper and Lower Bounds of Vacuum Energy E(g) in Nonlocal Theory
and in the Anharmonic Oscillator Case. The representation of the type (5.60)
and (5.61) obtained above for the S-matrix as functional integral is very
convenient for investigation of many problems of quantum theory. In parti-
cular, for the problems of:
~ summation of the perturbation series, i.e., of going beyond the scope of
the perturbation theory;
~ strong coupling (g --+ CfJ), and of the phase transition in different quantum
field theory models;
~ the behavior of the vacuum energy density in the strong coupling limit,
etco
We shall demonstrate, for example, how to find the upper and lower bounds of
vacuum energy density.
. 1
E(g) = - hm ~ln <OISv[g, cpJIO),
v~w V
190 Chapter 5

when g -+ 00 in the nonlocal g<p~ -theory as a function of dimension d and of


an ultraviolet behavior of the causal Green's function i5Ak 2 ) in the Euclidean
region. We will follow G. V. Efimov (1979). first, the lower bound of E(g) will
be obtained. Let us consider the function

Sy(g) = f dJ.lu exp{ -g I ddX¢4(U, X)} (5.63)

The inequality

I dd X¢4(U, x) :;;. ~ [I dd X¢2(U, x) J


gIves

Sy(g) ~ f dJ.lu exp{ -v[L dd X¢2(U, X)J} (5.64)

This integral can be calculated in the following way. It can be represented in


the form

(5.65)

where

Fy(g, t) = f dJ.lu exp{ - 2itJ97V I ddX¢2(U, X)}

An integral of this type was calculated by Efimov (1979) and his result is given
(for V -+ (0) by

Qy = fc5CP ex p{ -~ I ddXcp 2(X) - ia I ddX x

= exp{ - ~ f(~~k In[l + 2iai5Ak 2)] - ~ f dd X f ddX'CX(X)G(x - X')CX(X')}

where
Functional Integral Techniques in Quantum Field Theory 191
Thus, when V is large enough, we get

Fv(g, t) = exp{ -~ f G~ yIn[ 1+ 4i{fy/2 Dc(k 2)]} (5.66)

Substituting this function Fv(g, t) into (5.65) and introducing the new variable
t = uV l /2, one obtains

When V ~ 00 this integral can be calculated by the methods of steepest


descent. The saddle point is on the negative imaginary axis in the (u + iv)-plane.
Putting U o = - ivo we have the equation

or

Vo = Jg f(~~Y Dc(k 2)[1 + 4voJg DAk2)]-1. (5.67)

Thus the following estimate for E(g) is valid

E(g) = -lim ~lnSv(g) >-E_(g),


v~oo V

(5.68)

where Vo is determined by (5.67). Formulas (5.68) can be written on the basis of


equation (5.67) in an alternative form

E_(g) = voJg f(~~)T _~k2D;(k2) - )Jc1 + 4voJg DAk2)]


DAk 2 -1 (5.69)

Let us now obtain the behavior of E_(g) for g ~ 0 and g ~ 00. When g ~ 0

Vo = Jg f(~~Y Dc(k 2) = Jg DAO)


and E_(g) = gD;(O). When g ~ 00, the behavior of Vo as a function of g is
determined by the ultraviolet behavior of the causal nonlocal function Dc(k 2).
Further let us consider some particular cases. First, we examine the case when
Dc(k 2) decreases as
(5.70)

for k 2 ~ 00, where 2(1 + a) > d as a consequence of condition D(O) < 00 in


192 Chapter 5

(S.S8). In this case, from (S.67) for 9 -> 00


Vo = const g(d/2)/l4(1 + a) - d]

follows, whence
E_(g) = C_l/l4 (1 +a)-d] (S.71)
where C _ is some constant.
In the case of anharmonic oscillator [d = 1 and a = 1 in (S.71)J the causal
Green's function is Dc(k 2) = (1 + k2)-1 (in units m = 1). In this case, the
function (S.68) can be calculated in the explicit form:

E _(g) = m;x{ - u; + ~[(1 + 4U)1/2 - 1J}-

This expression can be written after simple transformations


v(16 + 18v + 9v 2 )
E_(g) = 8(1 + v)(3v + 4) ,
where v(l + v)(2 + v) = 2g. Hence it appears that in the limit of large and
small 9

E (g) = {~g 9 -> 0, (S 72)


- i(2g)1/3 = 0.474 g1/3, 9 -> 00. .
Second, we consider the case exponential decreasing
DAk2) = O(exp( _k2Y)) (S.73)
for k 2 -> 00. The solution Vo of equation (S.67) is
Vo = const(ln g)d/4Y

in the limit 9 -> 00 so that


E_(g) = C_(lng)1+d/2Y[1 + O(lnlng/lng)]. (S.74)
Here C _ is a certain constant.
Now let us obtain the upper bound to the function E(g) for real positive g.
The following obvious inequality is valid for the function Sv(g) in (S.63):

Sv(g) ~ fbU ex p{ -t~1 (1 + qn)u~ - 9 I ddXC~1 D~/2(X)Un J}


n=1
TI 00
(1 f -- L
+ qn)-1/2 buexp{lOO
2s=1
u; - 9 X

(S.7S)
Functional Integral Techniques in Quantum Field Theory 193

Here qn are positive numbers. Then it is easy to get the following upper bound
for E(g):

· -In
E(g) = - 11m 1 Sv(g) -< -1 1im -1"
1...ln(1 + qs) - 1·1m -1 x
v~C() V 2v~C() V s v~C()V

-<"21 lim
V~WV
I
L +I
- + i [L
s
In(1 q.) 3g lim -
V~WV v
dd X
s
Ds(x)(1 + qs)-1 J2
The numbers qs can be chosen in such a way that

lim -1 L In(1 + q.) = ~(dk)d


~2 In[1 + q(k2)],
v~CX)V s n

lim L D,(x) = i(dk)d 15Ak 2) (5.76)


v~oo s 1 +qs J\2n 1 +q(k2)"
For example, in the case of the orthonormal system (5.57) these numbers
should be taken as
qS = q51 ... ,sd = q(L -2n 2(s21 + ... + S2))
d

and V = (2Lt
Because the numbers qs and consequently the function q(P) are arbitrary,
we have E(g) -< E +(g),

{ I f(dk)d
E+(g) = ~:~"2 2n In[1 + q(k2)] + 3g [i( dk)d 15 (k 2)
J\2n 1; q(k2)
J2} . (5.77)

When g --> 0, the minimum is reached for q(k2) = 0. In this limit, we obtain
(5.78)
i.e., the lowest perturbation order when the interaction Lagrangian is not
taken in the normal form.
The behavior of E+(g) for g --> 00 is determined by the ultraviolet asymptotic
of the Green's function 15c (k 2). In case (5.70), the function
(5.79)
can be chosen to determine the function E+(g) in (5.77), where A and [3 > d/2
are parameters.
Substituting (5.79) into (5.77) and putting g --> 00, one can see that the
minimum is realized for large A. We have

E+(g) <- min min{ C 1([3)A d + CzC[3)gA 2d - 4(1 + a)},


{3 > dj2 A
194 Chapter 5

where

Cz([3) = {f(~~J k2f3 ·k- 2(1 +a)(1 + k2f3 )-lJ.


Then it follows
E +(g) -< C +gd/[4(1 + a) - d1, (5.80)
where the constant C + can be calculated as follows
_. [ d'C 2([3) ][2(2(1+a)-d)Cz([3)]4(1+a)-d
C+-m;nC 1([3) 1+ 2(2(1+a)-d) d,C 1([3) .

Thus the two estimates (5.74) and (5.80) show that the nonlocal-stochastic
theory when Dc(k 2) = O«k2)-1-~) for k 2 -> 00 in the limit 9 -> 00, the vacuum
energy E(g) is
E(g) = Cl/[4(1 + a) - dl, (5.81)
where C is a constant satisfying C _ < C < C +. Formula (5.81) can be
rewritten in the form
E(g) = CgJ(d.a), (5.82)

f(d, a) = d[4(1 + a) - dr 1 = {1-[1- 2(1 ~ a)]}{1+[1- 2(1 ~ a)]r 1 < 1


because of 2(1 + a) > d according to the condition DAO) < 00.
Note once more that the behavior E(g) for 9 -> 00 depends on the dimension
of space-time d and the ultraviolet asymptotic of the causal nonlocal Green's
function i5 c(k 2 ) for k 2 -> 00.
Let us consider E+(g) for 9 -> 00 in case (5.73). This estimation can be
obtained by choosing q(k2) = A exp( - k 2Y ). After simple calculations one can
get in the limit 9 -> 00
E+(g) = C+(lng)l +d/2 Y [1 + O(lnlngjlng)].
Thus in the cpJ-theory with the propagator (5.73), we have
E(g) = C(lng)l +d/2 Y [1 + O(lnlngjlng)], (5.83)
in the limit of strong coupling. In the paper by Bervillier et al. (1978), the cp4_
theory with the propagator i5 c(k 2) = exp( _k2) [i.e., y = 1 in (5.73)] was
considered. These authors have calculated eight orders of perturbation theory
for different characteristics of this model (energy levels, Green's functions). The
knowledge of exact asymptotics (5.83) should be useful for different methods of
asymptotic summation.
Functional Integral Techniques in Quantum Field Theory 195

Finally, we turn to the function E+(g) in (5.77) for the anharmonic oscillator,
which reads:

In the limit 9 ~ 00, it is easy to obtain

E+(g) = mjn m!n[ 2 sin ~n/2(3) + 3gA -2(2{3 sin 2n(3) -2J
3[
= g1/3 { 2n m;x -u-
2
sin uJ -1}2/3
= 0.757 X g1/3. (5.84)

Collecting the estimations (5.72) and (5.84), we obtain


0.474 x g1/3 < E(g) < 0.757 X g1/3. (5.85)
As a second example, we shall present an elementary asymptotic method
which allows a more precise asymptotic value to be obtained instead of (5.85)
in the case of the one-dimension anharmonic oscillator. For this, let us consider
first a numerical function of the form

F(g) =f -
OO

00
du
Fe: exp{ -tu 2
v' 2n
- gu 4 }F(u)
[fOO
- 00
du
Fe: exp( -tu 2
v' 2n
- gu 4 )
J-
1

_f dJ.lu exp( - gu 4 )F(u)

- fdJ.lu exp( - gu 4)

where

dJ.lu =
du
Fe: exp( -IU ),
v' 2n
1 2
f dJ.l uu2m = (2m - 1)!!, m = 0, 1, ....

We now carry out some formal transformations


196 Chapter 5

and, passing to the variable u --+ u/ JQ, we get

ffo
du
1= Q-1/Z - - exp{ -tu z + t(1- 1/Q)u 2 - gQ- z u4 }F(Q-1/2U).

Hence

f d,uu exp (WQ(u)) F(Q -1/2U)


F~)= ,
fd,uu exp(WQ(u))

where
W Q= t(1 - 1/Q)u Z - (gQ-Z)u 4.
Now the question arises of finding the asymptotics of this function in the
limit g --+ 00, when F(u) is a non-simple function. The main idea of the solution
consists in expanding the exponential exp(WQ(u)) in a series, and studying the
extremal problem for this series. Then we have to find such a procedure of
minimalization which allows an expression that would converge to the true
asymptotic value of the integral in the limit g --+ 00 to be obtained. We shall
demonstrate this idea on the simple example when F = F1(U) = u z. Then the
exact asymptotic of the function F 1 (g) equals

g-1/2 {OJ dtt1/Z+1/4-1e-t

{OJ dt t1/4 - 1 e- t

= g-1/2ni) = 0338 x -l/Z.


ni) . g

Our goal is now to find such extremal procedure for the expression of F 1 (g),
defined by the series, which approaches the exact value 0.338. We put
WQ(u) = t(1 - 1/Q)u Z - (g/Q 2)U4 ~ W(u) = !U z - AU4,
Q}>1

Then

f d,uJ(u)(1 + W + tw 2 + iW 3 + ... )
F(g) = "------~--------
F + FW + tFW2 + .. .
1 + W + tw 2 + .. .
f d,uu(1 + W + tw 2 + ... )

= (F + FW + tFW 2 + .. ·)(1 - W - tw 2 - .•• )


Functional Integral Techniques in Quantum Field Theory 197

Collecting terms of the same order, we get

+ {i(Fw 3 - F W 3) - t(FW 2 - F W2)W + (FW - F W) x


x (_tW2 + W2)} + {l4(FW 4 - FW4) - i(FW 3 ) - FW3)W +

+ t(FW 2 - FW2)( - tw 2 + W2) + (FW - FW)( - t W3 +


+WW2_W3)}+ ...

For example, for F(u) = F 1 (u) = u 2 one can write some terms

W = f d.u u(tu 2 - ).u4 ) = t- 3)"

U2W 2 = fd.uuU2(!u4 - ).u 6 + ).2U 8 ) = t'3 x 5 - 3 x 5 x 7·). +3x 5 X

x7x9·).2,

Thus
F 1 ().) = )I{1 + [1 - 12A] + [1 - 36), + 12 x 32).2] + [1 - 72), +

+ 1920).2 - 19008).3] + [1 -12OA + 3 x 1920).2 - 9 x


(5.86)

In the first approximation we have minimalized the term 1 - 12A and find
).~! = /2' Substituting this value into the first term of expression (5.86), we get
Fil)(~!) = 0.289 x g-1/2.
198 Chapter 5

In the second approximation we find a minimum min;J1 - 36) + 12 x 32)Z],


J~! = i4.
Substituting this quantity in the first two terms of (5.86), we have
F1Z)(~!) = 0.343 x g-l /Z.

Further, as a third approximation, we take the minimum

min(1 - 72J + 1920J z - 19008J 3 ),


;.

and find
Fl3)(J~!) = 0.329 x g-l /Z.
In the fourth approximation,

min(l - 120J + 3 x 1920Jz - 9 x 14784J 3 + 36 x 34816J4 ),


;.

we obtain J~;!,...., lo which gives Fl4)(J~!) = 0.334 x g-l /Z. As a result of such
manipulations, we see that the formal series obtained after a sensible rninimali-
zation procedure may indeed converge to the exact (true) value 0.338 x g-l /Z.
Now we use this formal method for defining the asymptotic value of the
vacuum energy density for the anharmonic oscillator with the Lagrangian
density:

5t = t<pz + tw Z <pz _ g<p4.


In this case, we have

For simplicity we put WZ = mZ = 1 and find the asymptotics for


. 1
E(g) = - hm -In(<OISv(g)IO)), (5.87)
v~C() V

where

(5.88)

in the limit g --+ 00 (here dimension of space is one, i.e., d = 1). In order to
accomplish this, we multiply (5.88) by exp( -tqnu; + tqnu;) and make the
change Un --+ u~(1 + qn) -liZ. So we get

Sv(g) = fI (1 + qn)-
n=l
l/Z f d~
v2n
x

{
x exp -2Un
1 Z 1
+" qn Z
2 - - Un -
1 + qn
g i [L
V
dx
s
d s1/Z (X)
~Us
1 + qs
J4} .
Functional Integral Techniques in Quantum Field Theory 199

Let us put now

dflu = n°o dUn 1 2


M:: exp( - tun)
n= 1 V 2n

and take into account that

Here qn are positive numbers. The numbers qn can be chosen in such a way
that [see, also (5.76)]

lim ! L~ = fdk q(k2)


v~ooV n l+qn 2n l+q(P)'
Then the vacuum energy density (5.87) can be written in the form:

1 I In(l
E(g) = lim -2
V~OO V n
+ qn) - lim
V~OO
!V In S~(g).
We write an expression for S~(g) in a convenient form as in the case of the field
theory, namely

S~(g) = f dfl exp(Wq) = 1 + Wq + tw; + tw~ + ...


where

Here

CfJix) = L (qn)1/2gn(1 + qn)-1/2 un


n
200 Chapter 5

where gn = gn(x) is an orthonormalized basis. Then by definition

Aix - x') = <OIT[4>q(x)4>q(x')]IO) = f~~ e- ik (x-x')3.(k Z)[1 + q(kZ)]-l,

Cix - x') = <01 T{cpix )4>q(x')} 10) = f dk K1/Z(kZ)


2n e-ik(x-x') Jq(k Z) I + q(kz)"

Expanding In(1 + x), we get

V1 In Sv, = VI {-Wq 1 -Z
+ 2(W q -
-z
W q )
1-3
+ () W q -
1- -Z
2 Wq W q
1 - 3
+"3 W q + ... ,
}

Then one obtains the following expression for the vacuum energy density
E(g) = E'6 + E1 + E2 + E'§ + ... (5.89)
where

E'6 = lim -I L In(l


a)
+ qn) = fdk
-In[1 + q(kZ)],
v~ a)V n =1 2n

E2 = - lim
v~ooV
I - - If
-H(W; - W;)} = --
4
f
dxB; - I2gZ dxA:(x) +

+ 6g f dxC;(x)Ai O),

As in the previous example, assuming E'f = 0, we get

_~fdk[_q- - 6gJ ] - 0
2 2n 1 + q (1 + kZ)(l + q) - ,
Functional Integral Techniques in Quantum Field Theory 201

where

whence
6gI rdk
q(k 2)=1+k 2 and I=J2n(1+k 2 +6gI)-1=t(1+6gI)-1/2. (5.90)

Substituting this quantity into the first term in the expansion (5.89), we obtain

E(l) = ~foo
4n _ 00
dk In(1 + 6gI 2) = t[(1 + 6gI)1/2 - 1J
1+k
I ~ t 31/3gl/3
g~ 00

= 0.72 X gl/3.
The following terms E1, i = 2, 3, ... are calculated explicitly for a concrete form
for q(k 2 ), namely
q(k 2) = qc(k 2) = A2/k2, (A> 0).
Then, at this value of qc' we have

According to the above manipulation we must set Ei' = 0:


A 23 g2 3 g
E2Qc = -16 + 40 AS +"4 A2 = 0,

from which it follows


Al = {6[1 +(1 - ~6)1/2Jp/3g1/3.

Substituting this value of Al into the two previous terms E~c + E'fc, we have
E(2) = tAl - tAl + l(g/Ai) = 0.71 x gl/3,
in the limit g --> 00. The explicit form of the term Ejc for the quantity qc equals
to

As a third approximation we put Ei c + Ejc = 0 which gives


A2 = (7.5)1/3 X gl/3 = 1.96 X gl/3.
Corresponding to this value, the vacuum energy density in the limit g --> 00, is
E(3) = 0.685 X gl/3.
The exact calculations for the anharmonic oscillator using the Schrodinger
equation give (Parisi, 1977) E(g) = 0.668 X gl/3. One can see that our estimates
are accurate.
202 Chapter 5

5.5.3. Functional Integralsfor Fermions in Quantum Field Theory


The action for fermions is given by

A(t/J, If/) f
= d 4 xlf/(x)(i<3 - m)t/J(x), a= "Ill 8~1l
As might be expected the functional integral over the variables t/J, If/ is then the
integral of the exponential of in - 1 A(t/J, if/). In the functional integral, t/J and if/
are independent integration variables. This means that the analogue of Z(1'/)
(see Section 5.5.1) is the functional F((, ;:), where (n = 1)

F((,;:) = f f bt/J bif/ exp{i f d x[if/(x)(i a- m)t/J(x) + ;:t/J + if/n},


4 (5.91)

here ((x) and ;:(x) are the sources of if/(x) and t/J(x), respectively, and they are
elements of the Grassmann algebra. Using exactly the same techniques as
those used to evaluate Z(1'/), it may be shown that

F((,;:) = exp{i ff d 4 x d4 y;:(x)SF(x - y)((X)}K -1 x

(5.92)

where
SF (x) = (i a+ m)·A(x) (5.93)
is the free fermion propagator. Note that in all respects this is a very similar
formula to that for Z(1'/) in (5.52). An important difference is that the Fredholm
determinant K occurs as K -1 in (5.92), rather than K, as it does in (5.52). This
is because of the change of the variable formula given in (5.27). So that the
n
normalized F((, is given by

_ f f Jt/J Jif/ exp{i f d x[lf/(x)(i a- m)t/J(x) + ;:t/J + If/(]}


4

F((, () = (5.94)
f f Jt/J blf/ exp{i f d x[if/(X)(i a- m)t/J(x)]}
4

Now F(O, 0) = 1. We know that the functional integral technique gives rise to
the perturbation series of Feynman diagrams (for detail, see Nash, 1978). For
example, for quantum electrodynamics the Green function is given by the
formula
<01 T{A Il1 (xd'" A Il.(xn )if/(Y1)··· if/(Ym)t/J(Z1)'" t/J(zm)} 10 >
= (_ on + 2m In + 2m Z((, ;:, 1'/) (5.95)
J1'/lll(X 1)'" b1'/lln(X n )b((Yl)'" J((Yrn) b((zd'" J((Zrn)'
Functional Integral Techniques in Quantum Field Theory 203

with 1]/l = ( = r = 0, where


Z(C r,1]) =

= f f E>1jI E>lf/ f E>A exp{i f d x[ -iF/lvpv + If/(i a- m)1jI -


4

{f fE>1jI E>lf/ f E>A eX{i f d x( -iF 4 /lV Pv

+ If/(i a- m)1jI - t(O,lA,l)2 - e: i/f Y/lA/lIjI:)]}


(5.96)
In (5.96), 1]/l(x) is a c-number source for the photon field A/l(x). When e = 0, the
functional integrals over 1jI, If/, and A/l can all be reduced to the integrals over
quadratic forms with the result

Z((, r, 1]) = exp{i ff d 4 x d 4 yr(x)SF(x - YK(Y)} x

x expH ff d4 x d4Y1]ix)D~V(x - Y)1]v(Y)} (5.97)

The function D~V(x) is the photon Green function defined by


D~V(x) = _g/lV ~c(x), (5.98)
where ~c(x) is the Green function for the zero mass Klein-Gordon equation
(5.99)
In order to give a definition of the functional integral over fermion fields we
consider, for example, the Yukawa model which describes a pseudoscalar
interaction of the fermion field IjI (x) with the boson field q> (x) in the Euclidean
d-dimensional space !R~. It is assumed that d is even. The interaction
Lagrangian density is given by
5t(x) = iglf/(x)Ysljl(x)q>(x). (5.100)
The Euclidean y-matrices and spinors 1jI, If/ are defined so as to ensure, first, the
covariance of the theory with respect to the Euclidean rotation, and secondly,
complete the coincidence of the S-matrix both in the suggested and standard
approaches to the series of the perturbation theory.
In the Euclidean !R~ -space, there exist d anti-commuting hermitian matrices
204 Chapter 5

I'll (Ji = 1, ... , d) with the dimension d ® d(d is even)


YIl Yv+Yvy ll =2(j1lV' Ji,v=1, ... ,d, yll+=yJ1 ·
Introduce also the matrix Ys = id(d - 1)/2 1'1 ... I'd' which satisfies the following
properties:

For example, in the space ~2 as the yll - and Ys-matrices one can be chosen

where a j are Pauli's matrices. In the space ~4:

( 0ia -ia.)
Yj = j
0 J, U = 1,2,3)

Ys = -1'11'21'31'4 = G~)
t/J and VI-fields in the Euclidean space are independent and anticommutative.
The causal Green function for the Fermi fields can be written as

(5.101)

where k = YJ1k W The integration in (5.101) is undertaken over the Euclidean


space ~~. The propagator (5.101) satisfies the following conditions
S; (x) = SF(X), Ys SF (x)Ys = SF( - x). (5.102)
We introduce the following symbol
G(k 2) = ISF(ik)1 = [SF (ik)SF ( _ik)]1/2
and assume that
SF(ik) = [a(ikW·
Then the following relation holds

SF(X 1 - X2) = f ddya(x 1 - y)a(y - x 2),

where

(5.103)

The function a(x) also satisfies relations (5.102). The propagator of the Bose
field is

~(X1 - x 2) = fddy ~1/2(X1 - y) ~1/2(y - X2), (5.104)


Functional Integral Techniques in Quantum Field Theory 20S

where

N/2(X) = fG~y(m2 + k2)-1/2 e- ikx •

The S-matrix depending on the Bose field CPin(X) and the Fermi fields ljJin(x),
l/fin(X) can be written in the form of a functional integral by means of the
following formula:

S[ CPin' ljJin' l/fin] = C - 1 f bcp f f bljJ bl/f x

x (l/f(x) + l/fin(X))rS(ljJ(x) + ljJin(x))(cp(x) + CPin(X))} (S.10S)

Here the functions ~ -lex) and Silex) are defined by the relations

f dd y ~(Xl - y) ~ -ley - x 2) = b(d)(Xl - x 2),

f ddy SF(X I - y)Sil(y - x 2) = b(d)(Xl - x 2)·

The normalization constant in (S.10S) is chosen in such a way that


S[ CPin' ljJin' l/fin] I9 = 0 = 1.
As stated above, in the case of the Bose field we perform some formal
transformations. In (S.10S), we carry out the change of variables

cp(x) = f ddx' ~1/2(X - x')<l>(x'),

ljJ(x) = f ddx' u(x - x')'¥(x'),

l/f(x) = f ddx''JI(x')u(x - x'), (S.106)

where the functions ~1/2(X) and u(x) are defined by equations (S.104) and
(S.103). As a result of this change, the S-matrix (S.10S) can be represented in
206 Chapter 5

the form:

-
S[CfJ;n,l/J;n,l/J;n] = C -If<5<1> If - <5'P<5'P exp -"2 ddx<l> 2(x)-
{If
-f ddx'Ji(X)'P(X)} x exp{ig ddx['Ji(x) f + Vi;n(x)] x

x YS['P(X) + l/J;n(X)][<I>(X) + CfJ;n(X)]}, (5.107)

where the fields 'P(x), 'Ji(x) and <I>(x) are connected by relations (5.106) with
the fields I/J(x), Vi(x) and ¢(x), over which the integration is carried out.
The functional integral over the Bose-field has been defined above. Now we
turn to its definition over the fermion fields. We consider the functional
integration in (5.107) over the fermion fields in the language of integrals on the
Grassmann algebra (see above). Now we define the variables {C v , Cv } as the
generators of the Grassmann algebra. Here the index v = (n, ct), n = 1,2, ... ,
and ct = 1, ... , d is a spinor one. The generators {C v , Cv } satisfy the following
relations:
(5.108)
As above, we introduce the symbols (dC v , dC.) and assume that they are also
the elements of the Grassmann algebra, i.e., they satisfy relations (5.108).
Besides, they anti-commute with all the elements of the algebra {C v , C.}.
Integrals over the Grassmann algebra are

Omitting for simplicity the spinor indexes, we put into (5.107)


00 00

'P(x) = L gn(x)C v, 'Ji(x) = L Cvgn(x), (v = n, ct).


n=O n=O

Then, according to (5.106), we obtain


00 00

'P(x) = L O"n(x)C v, 'Ji(x) = L Cvan(x), (5.110)


n=O n =0
where

O"n(x) = I ddx'O"(x - x')gn(x'),

The following relations hold


00 00

lim L O"n(x)an(x') = SF(X - x'), lim L O"n(x)O"n(x') = G(x - x').


V--+OCJn=O v-OJ n=O
Functional Integral Techniques in Quantum Field Theory 207

We introduce the quantities


N N N
'¥N(X) = L O"n(x)C" 'PN(X) = L C,trn(x), <l>N(X) = L L\;/2(X)Un
n=O n=O n=O
and define the Grassmann measure
N d
dNJiG = TI TI dCnadCMexp(-Cn;Cna)
n=Oa=l
and

J
so that dNjiG = 1. Now we are able to define the functional integral in (5.107).
By the definition:

SV[qJin' I/Iin' Vlin] = :~~ f dNf1q, f dNjiG exp{ig 1 ddx('PN(x) + Vlin(X)) X

X YS('¥N(X) + l/Iin(x))(<I>N(X) + qJin(X))} (5.111)


Thus, the functional integral in the finite volume is understood as the limit of
the multiple integrals. From (5.111) it is evident that the obtained expression
coincides completely with (5.107) at the formal limit N -+ 00 and V -+ 00.
The series of the perturbation theory for the S-matrix elements obtained
from (5.111) coincide completely with the usual series of the perturbation
theory, since

lim fd JiG'¥(x)'P(x') = lim L O"n(x)trn(x') = SF(X - x').


V-+oo V-+oo n

As was mentioned above, the representation obtained for the S-matrix in the
form of the functional integral is convenient for investigation of the vacuum
energy density. In the given model, it is expressed by the formula

E(g) = - lim ~ In(<01 Sv(g)1 0»). (5.112)


v~oo V
<
According to the above definition, the value of 0ISv(g) 10) can be written in
the form of the functional integral

<OISv(g)IO) = fdJi<l> fdJiGexp{i g 1 ddX'P(X)YS'¥(X)<I>(X)} (5.113)


208 Chapter 5

Usually the problem is reduced to the study of the function E(g) in (5.112) for
the pseudoscalar interaction (5.100) beyond the scope of the perturbation
theory. Here we carry out an integration of expression (5.113) over the
Grassmann variables only. For this expression (5.113) can be rewritten in the
convenient form

n n dCnandCMn'
00 d

dCdC= (5.114)
n=Oan =1

n,a ",IX n',cx)

A matrix H equals to

H = 1 ddxO'(x)ysO'(x)<P(x) (5.115)

and its matrix elements have the following form:

Hna,nfa f = 1 dd X(O'n(X)Ys O'nf(x)<P(X))aa f'

Now we carry out the transformation identical with the one in the case of
the Bose-fields (see Section 5.5.2) in integral (5.114):

<OISv(g)IO >= f ff
dJl<J) dC dC exp{ - [C(l + q)C] + (CqC) + ig(CHC)} (5.116)

where q is a diagonal matrix with matrix elements

The numbers qn are positive and satisfy the conditions Ln ~ 0 qn < 00. By
changing the integration variables in (5.116) C na ~ C na(l + qn)-1/2, we get

where
Functional Integral Techniques in Quantum Field Theory 209

I.e.,

The functional integral over the Grassmann variables is easily calculated and
we obtain

<OISv(g)IO) = JIY + qn)d fdflq, det(I - Jq - igH q ), (5.117)

where J q = q/(l + q) is a diagonal matrix. The next problem is to estimate the


determinant in (5.117) and to obtain an estimation for the vacuum energy
density in different cases. This problem has been solved by Efimov and Ivanov
(1981).
Part II

Stochastic Quantum Mechanics


and Fields

Part II presents an elementary introduction to stochastic mechanics and the


stochastic quantization of physical systems. In Chapter 6 we give the basic
concepts of the probability theory, and random processes such as the Wiener,
Markov and Gaussian processes and martingales. Elementary stochastic cal-
culus which may play an important role in stochastic quantization procedure
is also given. In Chapter 7 we give the main ideas of stochastic quantization
proposed by I. Fenyes, E. Nelson, F. Guerra, etc. Here we have attempted to
give a physical origin of stochasticity in physics based on the hypothesis of the
stochastic or fluctuational nature of space-time and the physical vacuum in the
microworld. Chapters 8 and 9 deal with the construction of both the nonre-
lativistic and the relativistic dynamics of a stochastic particle including two
bodies, and with a study of the equations of motion and the Cauchy problem
for stochastic mechanics. They deal also with some specific problems, such as
self-turbulent phenomena, Feynman-type processes and position-momentum
uncertainty relations in stochastic mechanics.
New developments in stochastic quantization of different fields including
gauge fields are discussed in Chapter 10. Further, we briefly review other
methods of stochastic quantization of gauge fields, vacuum tunneling pheno-
mena, and stochastic solutions of the classical Yang-Mills equations obtained
recently by Matinyan et al. (1981).
In Chapter 11 we consider some consequences of the hypothesis of stochas-
tic space-time and of fundamental length. It is shown that the introduction of
the value of fundamental length leads to the restriction on the density of
matter and on the life-time of the proton. We close this book with some
comments on gravitational considerations connected with a hypothesis of
nonlocality of space-time metric and its possible consequences, and the origin
of cosmic rays and the value of fundamental length.

211
Chapter 6

The Basic Concepts of Random Processes


and Stochastic Calculus

In this chapter we briefly give some basic probabilistic concepts and de-
finitions of random processes such as the Markov, Gaussian and Wiener
processes (see Gihman and Skorokhod, 1975; Dynkin, 1965), and therefore
there is a large number of theoretically-probabilistic terms. There is a reason
for introducing the terms: a 'random variable' and 'random process' for the
equivalent 'measurable function' and a 'random field', namely an implied
change of view-point. We use the word 'random' although in theoretically-
probabilistic literature the words 'random' and 'stochastic' are used equally.
The basic concepts of probability theory are the notions of an experiment,
event and the probability of an event. At the formal descriptive level of these
notions it is accepted to start with the theoretically-set model of probability
theory, proposed by Kolmogorov (1933, 1956).

6.1. Events

In probability theory (stochastic) experiments are carried out by the obser-


vation of certain conditions. However, this set of conditions does not define
the outcome or the final result of the experiment. This means that in repeated
experiments with a precise observation of complex conditions, generally
speaking, the results will differ. The first fundamental assumption on the
formalization of theoretically-probabilistic concepts is that the results of the
given set of experiments can be described by means of some set Q. A definite
subset A E Q corresponds to each event.
At this, points WE Q play the role of atoms; any event is a sum of points but
each point W is not presented in the form of the sum of other events. Therefore,
points (or an arbitrary subset) of Q are called elementary events. However,
from a practical point of view there is no logic in regarding an arbitrary subset
of Q as events. Therefore, it should be taken in Q class of events of interest.
DEFINITION 6.1. A class of events is called an algebra of events, if it
contains the reliable Q and impossible (/) events, together with an arbitrary

213
214 Chapter 6

pair of the events A and B from this class (it contains also their sum and events
opposite to A). Two events nand (/) define the trivial algebra. Minimal
algebra containing the event A consists of four events: n, (/), A and A.
DEFINITION 6.2. The algebra of events is called a a-algebra, if together with
an arbitrary sequence of events it contains their sum. Usually, a a-algebra of
all considerable events in a given situation is denoted by the letter I:.

6.2. Probability

A probability (measure) space is a triplet {n, I:, f.1} of a set n, a a-algebra I: of


subsets of n, and a positive measure f.1 on n with f.1(n) = 1. Such a measure is
called a probability measure. A probability (a base or sample) space n is the
initial object in probability theory. Some simple well-known properties of the
probability measure are:
(1) f.1((/))=O.

(2) If Sk n Sr = (/), k =I- r then f.1( U ~ Sk) = I ~ f.1(Sk)·


(3) If S1 C S2, then f.1(S2\Sd = f.1(S2) - f.1(Sd.
(4) f.1(S) = 1 - f.1(S).
(5) If Sn C Sn+ 1, n = 1,2, .... , then f.1( U ~ Sn) = lim fl(Sn).

(6) If Sn ::J Sn + 1, n = 1, 2, ... , then f.1( n ~ Sn) = lim f.1(Sn).

6.3. Random Variable

Representation of a stochastic experiment conslstmg of numerical value ~


measuring, corresponds to the concept of a random variable. It is assumed
that for an arbitrary pair of numbers a, b(a < b) the event A(a, b) consisting of
~ c (a, b), is observed. Let A", - 00 < x < 00 denote the event ~ = x. It is
measurable. Indeed,

Ax = nA
n=1
00 (

n
1 + -1) .
x - -, x
n
At this, if x 1 =I- x 2 , the events Ax! and AX2 are incompatible (it follows from the
fact that the result of measuring is synonymous) and union of all the Ax,
- 00 < x < 00 gives n. We now define the synonymous real function f(w),
WEn by settingf(w) = x if x E Ax. From the definition it follows that in each
The Basic Concepts of Random Processes and Stochastic Calculus 215

experiment ~ = f(w) and at this a set


{w: a <f(w) < b} = A(a, b),
is measurable.
An important example of random variable is the indicator of events. The
indicator of event A is called a random variable (quantity) XA = XA (w):
I if WE A,
{
XA(W)= 0 ifw¢A.

If A E L, the X(w) is L-measurable.


We recall that the real-valued function f(w) determined on the measurable
space {n, L} is called measurable (or L-measurable), if for arbitrary real
numbers a and b we have {w: a < f(w) < 'b} E L. Thus, a real-valued measur-
able function on {n, L} is called a random variable. More precisely, we have
the following definition in terms of a measurable map language. Let {n, L}
and {S, I!,D} be two measurable spaces. A map g: w 1-+ x (x E S) is called a
measurable map {n, L} onto {S, I!,D} if
g-l(B) = {w:g(w)EB}EL,
for arbitrary set BE I!,D.

DEFINITION 6.3. According to the concept of the random variable ~ = f(w)


we define a measurable map from a measurable space {n, L} to another one
{S,I!,D}, i.e., ~:n 1-+ S such that C 1(I!,D)E L.
In other words, a random variable is a measurable function f mapping n to a
set of real numbers IR (where one regards Borel sets as measurable ones). For
the given random variable f, the measure /1 f on (- 00, (0) defined by
/1f(M) = /1U- 1 [M]), Me IR

is called the probability distribution of f. The (nonnormalized) Fourier trans-


form of this measure

cf(t) = I d/1f (x) eitx = 1 d/1 exp(iif(w))

is called the characteristic function of f. The characteristic function of the


random variable f is a wonderful object.
THEOREM 6.1. (The Bochner theorem). A function c (from IR or q is a
characteristic function of a random variable, if and only if it satisfies the following
conditions:
1. c(O) = 1.
2. A function t 1-+ c(t) is continuous.
216 Chapter 6

L ZiZ/ti - tj) ;;" 0.


i,j

(The proof of the Bochner theorem can be found in many text books, for
example, Reed and Simon, 1975.)
Moments of the randomfunctionfare given by their characteristic functions;
nth moments of f are called

In df.ll" = f df.l f (x)x n

if this integral converges absolutely. Iff has moments of all the orders if and
only if cf (t) is a function of the class IC 00, and in this case

In df.ll" = (-I t (:t)n cf(t) t = o' (6.1)

DEFINITION 6.4. A Gaussian random variable with mean value is called °


the random variable f, of which the characteristic function has the form
Cf(t) = exp( -at 2/2), a;;" ° (6.2)
Constant t is chosen so that

a = In df.lf2, (6.3)

[in accordance with (6.1)]. The quantity a is called the dispersion of f. Note
that the probability distribution density of the Gaussian random variable is
also the Gaussian function:

df.l (x) = {b(X) dx a = 0, (6.4)


f (2na) -1/2 exp( - x2/2a)dx, a > 0.
Expanding (6.2) into power series and using (6.1) it is easy to obtain the value
of all the moments of the Gaussian random variable with the dispersion a:

f df.lf2n = rn (2~)! an,


n.
n = 0, 1,2, ....

DEFINITION 6.5. Let fl J2' ... ,In be n random variables. Their joint pro-
bability distribution df.l f 1, ... ,fn and the joint characteristic function cf1, ... ,fn
(tl,"" t n ) on [Rn are given by the relations
df.lfl, ... ,f.(M) = df.l{[fl ® ... ®Inrl(M)},
and
The Basic Concepts of Random Processes and Stochastic Calculus 217

respectively, where the map f1 ® ... ® J,,: n H [Rn is defined by the formula
(f1 ® ... ® J,,)(w) = [f1 (w), ... ,J,,(w)].
DEFINITION 6.6. A set of the random variables f1' ... ,J" is called the joint
Gaussian random variables, if the joint characteristic function of this set has the
form

where {a ij }i. j ; 1 is (symmetric) real positive definite matrix.


From this, we obtain the analogs of the formulas (6.3) and (6.4):

au = fdl1f~,
dl1 ft, .. ·.!. = (2n)-nI2(det a)-1/2 exp{-~2L-.
'\' b .. x.x.} d"x,
[j [ j
[,j

where a = {a ij } is called a covariance matrix of the set {J;}, and the matrix a is
assumed to have the inverse one which is denoted through b == a -1.

6.4. Expectation and Concept of Convergence over the Probability

DEFINITION 6.7. Expectation or mean value ofthe random variable ~ = few) is


defined by

IE {n = 111(dW)f(W).

DEFINITION 6.8. If a sequence of random variables ~n = .f~(w), n = 1,2, ... ,


converges to ~ = few) for each w, with the exception of the set N such that I1(N) =
0, then it is said that ~n converges to ~ with the probability one; or that
~ = lim ~n (mod 11)'
It is said that the sequence {~"; n = 1,2, ... } converges over the probability to
the random variable ~: ~ = 11 -lim ~" if there exists the random variable ~
such that
11{ I~n - ~I> E} --> 0, at n --> 00 for any E > O.
In the theory of measure, convergence over measure corresponds to the
convergence over probability. A necessary and sufficient condition for the
convergence over probability of sequence of random variables is that: for any
E > 0 and b > 0 it may be found such no = no(E, b) that at nand n' > no;

11{ I~n' - ~n I > E} < b. This condition is called a fundamental condition over
the probability of sequence gn, n = 1,2, .. . }.
218 Chapter 6

DEFINITION 6.9. Let Lp = Lp(O, L, p), p ~ 1 be a linear normalized space


of the random variables ~ on {O, L, p}, for which IE{ I~ IP} < 00. In Lp a norm is
introduced by the relation
11~llp = [1E{1~IP}JI/P.

The convergence of the sequence ~n to the limit ~ in Lp (Lp-convergence) means


that
1E{1~-~nIP}~O, atn~oo.

The convergence over the probability results from the Lp-convergence. It


follows immediately from the Chebyshev inequality:
P{I~n - ~I > 8} ~ 8-PIE{I~ - ~nIP}.
The space Lp is dense. Of the spaces Lp, the spaces Ll and L2 are most
important. The space L2 = L2 (0, L, p) of complex random variables becomes a
Hilbert space if a scalar product of the random variables ~, 1'/ pair in L2 is
defined by the formula IE { ~fj}.
Two random variables ~ and 1'/ are orthodiagonal, if IE { ~fj} = O. In this case,
when ~ and 1'/ are real and IE {n = E{ I'/} = 0, the orthodiagonality means
noncorrelativity. The convergence of the sequence {~n' n = 1,2, ... } in Lp to
the random variable ~, means that
II~ - ~n112 = IE{I~ - ~nI2} ~O, at n~ 00.

Such a convergence form is called the mean square convergence and is written
~ = lim ~n' In some cases it is convenient to express the convergence condition
in Lp by means of the covariance of a set of random variables.

DEFINITION 6.10. The covariance B(tl' t 2 ), tiE T, of a set of random vari-


ables {~t: tE T}, ~tE L2 is called the function B(tl, t 2 ) = IE{ ~tl ~tJ.
THEOREM 6.2. Let Yt' be the (separable) real Hilbert space. Then there exists
the probability measure space {O, L, p} and for each ~ E Yt' a random variable,
¢( ~), so that ~ f-+ ¢(~) is linear and so that for any ~ 1, ... , ~n E Yt', [¢( ~ d, ... ,
<
¢(~n)] are jointly Gaussian with the covariance ~i> ~j = B(C ~j)[ (-, . is an > >
inner product on JIlJ. Note that {¢(m is called the Gaussian process with the
covariance <',' >.
Proof (for details, see Simon, 1979). Let ~ 1, ... , ~n be an orthonormal basis
for Yt'. Let c(t) be defined for any sequence t l' ... , t n , • •. eventually zero, by
c(t) = exp{ -t Li tf}· Owing to the above, when exp( -t Li,j aijt;tj) is positive
definite and ¢1"'" ¢n,"" jointly Gaussian with the covariance ~ij (this is
somewhat circumlocutory, one can just take 0 = ~ 00, ¢i = Wi, and p = 11:= 1 x
X (2n) -1/2 exp( -tw;) dW m directly). Now there is given a finite sum

N N
~ = L O:i~i' and set ¢(~) = L O:i¢i'
i == 1 i = 1
The Basic Concepts of Random Processes and Stochastic Calculus 219

Then

so, by continuity, cp extends from finite sum to a map from Jf to L 2 (n, dll). It
is easy to use continuity to see that

fdlleXP[iCP(m = exp{ -!II~ln,


so that cp's are jointly Gaussian with the proper covariance. The uniqueness of
the process follows from the uniqueness aspect of Kolmogorov's theorem and
the L 2 -continuity. •

6.S. Independence
DEFINITION 6.11. n random variables f1' ... ,fn are called independent if and
only if their joint distribution dllIi , ... ,in is the product measure
dill! ® ... @ dll ln · For two events A and B it is easy to see that an
independence means Il(A !l B) = Il(A)Il(B). From the definition it immediately
follows that:
1. n and A are independent, where A is an arbitrary event.
2. If Il(N) = 0 and A is any event, then N and A are independent.
3. If A and B i , i = 1,2, are independent, and B1 =:> B 2 , then events A and
B1 \ B2 are independent.
4. Let A and B i , i = 1,2, .... , n, be the independent events, while
B 1, ... ,Bn are incompatible in pairs, then A and u ~ Bi also are
independent.
5. .It is not dependent on A, if and only if Il(A) = 0 or Il(A) = 1. Let I be
some set and {Mi' i E I} be a set of event classes indexed by i E I.
DEFINITION 6.12. The classes of events {Mi' i E I} are called independent
(or independent in totality), if for any inequalities in pairs i 1, i2 , ... , in (i k E I)
and any A ik , Aik E M ik , k = 1,2, ... , n;
Il(A i! !l Ai2 !l ... ! l AiJ = Il(A i!) Il(A i2 ) ... Il(A iJ
Let ~i =/;(w), iEI be some set of the random variables in {Si,~J. The
quantities {~i' i E I} are called independent (or independent in totality), if for
any n, n = 1,2, ... and any Bk E ~ik' ik E I:

11(0 1 gik E BiJ) = ill Il(~ik E BiJ (6.5)

PROPOSITION 6.1 Two Gaussian random variables f and g are independent


if and only if their covariance (f, g) = Jdllf' g is zero.
220 Chapter 6

Proof The Fourier transform of their joint distribution

M(s, t) = IE {exp(iif + isg)} = exp( -tllif + sgll~)

= M(t, O)M(O, s) exp[ - ts(f, g)]


is a product if and only if (I, g) = o. •
6.6. Conditional Probability and Conditional (Mathematical) Expectation

First of all, we recall the definition of conditional probability and conditional


expectation in the elementary case. The conditional probability Cond(A I B) of
any event A under the hypothesis B, is given by
C ond(A IB) = .u(A n B)/.u(B)

if .u(B) 4= O. The conditional probability Cond(A I B) at fixed B is the norma-


lized measure defined on the same o--algebra of a set, as for the "uncon-
ditional" probability .u(A). In accordance with this, conditional expectation of
some random variable ~ = few) under the hypothesis B, is given by the
formula

IE {~ I B} = 1 f (W) Cond(dwIB). (6.6)

Using the conditional probability definition, this relation can be rewritten in


the following way:

.u(B)lEg I B} = 1dw~·
We note that if ~ = XA (w) is the (the characteristic function) indicator of the
event A then IE{ ~ I B} = Cond(A I B). Thus, conditional probabilities are the
particular case of the conditional (mathematical) expectations.
Let ;?,8 be a sub-o--algebra of ~ and ~ a random variable with a finite mean,
i.e., ~ E Ll (0, .u).
DEFINITION 6.13. A conditional expectation of ~ = few) with respect to ;?,8
is defined to be a ;?,8-measurable function IE { ~ I ;?,8} on 0 such that

1 ~
d.u IE{ I ;?,8} = 1 d.u(dw)f(w), (6.7)

holds for VB E;?,8. The conditional expectation IE g 1;?,8} is nothing but the
Radon-Nikodym derivative of the o--additive function on ;?,8

.u(B) = l.u(dW)f(W), VB E;?,8


The Basic Concepts of Random Processes and Stochastic Calculus 221

with respect to the probability measure J1. Therefore, the conditional expec-
tation, if it exists, is unique with the probability one. We prefer to write IE x
x gin in spite of IEgl~} if the sub-O"-algebra ~ is generated by a random
variable (. If (1, ... , (n are random variables, IE {~ 1(1, ... , (n} denotes the
conditional expectation of ~ with respect to the O"-algebra generated by
(1,· .. , (n·
Now we list the basic properties of the conditional expectation:
(1) If ~to, then lEg I~} ~ o.
(2) If ~ is a ~-measurable random variable, then lEg I ~} = ~.
(3) 1E{lEg I~}} = IE{ O.
(4) IE{ ~d to, i = 1,2, then

lE{a~1 + b~21~} = alEg11~1 + blEg21~}·


(5) If the random variable ~ is independent of the O"-algebra ~, then

lEg I~} = lEg}.


From the definition, the independence of the random variable ~ from the 0"-
algebra ~ means that the O"-algebra O"(~) and ~ are independent. Therefore, for
any BE~

1dW~ = IEgXB} = IEg·J1(B)}.

So that equality (6.7) holds, if to put IE{ ~ I~} = IE{ O.


(6) If '1 is ~-measurable, then
(6.8)
or

For proof of this property, it is sufficient to assume that '1 ~ O. If '1 = XBj'
B1 E~, then

rdWtJ·lEg I~} = JBnB


JB
r IdWlEg I~} = JBnB
r I~·dJ1 = JBrdW'1·~
so that equality (6.8) is satisfied.
Since the conditional expectation IE{ ~ I~} is a random variable, one may
consider the conditional expectation of this variable with respect to another 0"-
algebra ~ 1> which leads to the iterated conditional expectation IE {IE g I~}
I ~}.
(7) Let ~ c ~1' then
1E{lEg I ~1} I~} = lEg I ~}. (6.9)
222 Chapter 6

Indeed, if BE fJ, then BE fJ 1 , therefore

From this we have equality (6.9).


We now give some definition of the conditional expectation with respect to
random variables. Let ~ be a random variable taking the values Z1' Z2'···' Zn'
... , /1( ~ = zn) > 0, Bn denotes the event {~ = zn} and /1n (A) = /1(A n Bn)//1(Bn)
is the conditional probability of A under the hypothesis ~ = Z';. Then, the
conditional expectation of the variable 1] under the hypothesis ~ = Zn is given by
the formula

1E{1]I~=zn}=
Inrd/1n·1]=_1-i
/1(Bn) Bn
1]·d/1.

Let (= g(w) be a measurable map: {Q,~} 1--* {S, fJ}. Thus, ( is a random
variable belonging to S. Let:Fe, be a u-algebra given by the map (:

DEFINITION 6.14. The conditional expectation lEg I of the random n


variable ~ with respect to another random variable ( is called the random
variable lEg I~}. This definition is equivalent to the following: for any
BcfJ

fM -
=9
1
(B)
dJl IE gin = r _ ~ d/1.
JM = 9
1
(B)

THEOREM 6.3 The conditional expectation with respect to the random


variable ( is a fJ-measurable function of (; lEg I = s(O, where s(O is a fJ- n
measurable function.
Proof Let ~ ~ o. Then from the definition we have
(6.10)

It is evident that q(B) is a u-finite measure on fJ. Besides; q(B) = 0 if


/1(g -1 (B)) = 0, i.e., q is absolutely continuous with respect to measure /1g'
where /1 g(A) = /1(g -1(A)). On the basis of the Radon-Nikodym theorem there
exists such fJ-measurable nonnegative function s(w) that

q(B) = L /1g(dw)s(w).
The Basic Concepts of Random Processes and Stochastic Calculus 223

Using the rule for the change of a variable, we get

q(B) = r
Jg-IIBI
p(dw)s(g(w».

Comparing with (6.10) one obtains the equality IEglO = s(g(w)) = s(O. •
DEFINITION 6.15. (Conditional density). Let {gr, rJU, m} be some space with
measure m and ~ = g(w) be a measurable map: {Q, L}, to {Et, OU}. Then it is
said that a random variable has the probability distribution p(x) (over a
measure m), if for any A E OU

p(~ E A) = L m(dx)p(x).

According to the Radon-Nikodym theorem, the random variable ~ = g(w)


possesses a probability distribution if and only if the measure Pg is absolutely
continuous with respect to m.
PROPOSITION 6.2. ~ is independent of ( if and only if IE (ei'~ 10 is a constant
for each real rJ..
Proof If ~ is independent of (, then IE {F(~) lOis constant for any F.
Conversely if IE {eia/; lOis constant for all rJ., then the constant is necessarily (
reia/;} and
lE{exp(irJ.~ + if3O} = 1E{IE{eia/;IOlei/i,} = lE{eia/;}IE{e ia,}.
for any rJ. and f3 so, taking the Fourier transforms fl~.( = Pc ® p,'
We now define the conditional probability 1C0nd(A IB) with respect to the
O'-algebra [1J as a particular case of the conditional expectations, assuming
~ = XA(W).

DEFINITION 6.16. The conditional probability 1C0nd(AIB) at fixed A is a


[1J-measurable random variable satisfying the following equation:

1 dplCond(A I[1J) = fl(A n B)

at any BE [1J. Some properties of the conditional probabilities are:


(a) If event B is [1J-measurable, then lCond (B I [1J) = XB' which follows from
property (2) for the conditional expectations.
(b) Assuming ~i = XB' Bl n B z = (/), then from property (4) we get the
additivity of the conditional probability:
lCond(B l u B z I [1J) = lCond(B l I [1J) + lCond(B z I [1J).
(c) From property (5) it follows that if the event A does not depend on the 0'-
algebra [1J, then 1C0nd(A I [1J) = p(A).
(d) In the simplest case, [1J = {Q, (/), F, F} for the subset FE L such that
224 Chapter 6

o < p(F) < 1, one obtains

Cond(B I f!,B) = {P(B n ~)/p(~)


on F,
(6.11 )
p(B n F)/ p(F) on F.
THEOREM 6.4. (Bayes).
Cond(B I F) = p(B) Cond(F I B)/p(F).
Proof By equation (6.11), we have
p(F) Cond(B I F) = p(B n F) = Cond(F I B)p(B).

6.7. Martingales

DEFINITION 6.17. Let ~ 1, . . . . ~n be a sequence of random variables. We call


them a martingale (respectively, submartingale and supermartingale) if and
only if for each m, 1E{I~ml} < 00 and
IE gm I ~ 1, ... , ~m - 1} = ~m - 1

[respectively, IEgml ~1'" ., ~m - d ~ ~m - 1 and IEgmI~ 1, ... , ~m - d<


~m - 1]. It follows by induction that for j < m

(6.12)
(respectively, ~ ~j and <~) If
mean zero and are independent, then
'1, ... ,'n are random variables which have a
j

~j= Ie
i= 1

is a martingale.
THEOREM 6.5. (Doob's, 1953 inequality) Let {~n} be a submartingale. Then
for each A > 0 and n:

lE{max
O~j<n
~j ~ A} < r 1E{~n, 1

where ~: = max(~n' 0).


Proof Let Xj be the characteristic function of the sets Aj of points, where
~1"'" ~j-1 < A and ~j ~ A. Let X be the characteristic function of the set A
with maXo<j<n ~j ~ A. Since the A/s are disjoint sets with union A, we have
n n n
U{A} = I U{AJ < I 1E{~jxJ < I 1E{1E{~n I ~1' ... ,~j}lxJ
j =1 j = 1 j = 1

= I IE{ ~nXJ = IE{ ~nX} < IE{~:}.


j =1
The Basic Concepts of Random Processes and Stochastic Calculus 225

Above, we have used ~j :;;;> Ie in the inequality, the submartingale relation (6.12)
in the second inequality, the fact that Xj is ~1"'" ~j measurable and (6.8) in
the next equality and the calculation

in the final step.



DEFINITION 6.18. Suppose that {qt}, tE [a, b] is a family of random
variables. We say it is a martingale (respectively, submartingale and super-
martingale) if and only if
lE{qt I [qs I s « u]} = qu,

(respectively, :;;;>qu and «qJ for all a « u « t « b. Note that, a fortiori, one has
IE {qt I qSl"'" qsJ = qSn'
(respectively, :;;;>qsn and «qsJ if Sl « S2 « ... «sn « t so that for Sl < ... <sn
< .. " the sequence ~n = qSn is a (sub) and (super) martingale. As a result,

IE{ max qSn :;;;> Ie} «;'-1 IE {qs: }.


O<j~n

Taking the mesh of the points Si to zero we have the following:


COROLLARY 6.1. Let {qt}, tE [a, b] be a submartingale with continuous
sample paths (i.e., t 1--* qt is continuous almost everywhere; this is a 'version
dependent'). Then

IE { max qs :;;;»,} « A- 1 IE { q/} .


O<s";;:;t

The following basic properties of martingales are:


(i) If qt, - 00 < t < 00, is a submartingale (supermartingale), then m(t) =
IE{ qt} is an increasing (decreasing) function of t.
(ii) m(t) = const, if the function qt is a martingale.

6.8. Definition of Random Processes and Gaussian Processes

Let {Q, L, j,l} be a given probability space. If realization of some experiment is


described by a functionf(y) of a definite argument y, y E Y, then we say that a
random function is given on {Q, L, j,l}. Thus, a random function is a
map: w 1--* f(y) = f(y, w), WE Q. In addition it is required that the function
f(y, w) should be a random variable at fixed y.
DEFINITION 6.19. (General Definition). Let Y be some set and {S,:?lJ} be a
measurable space. The random map f(y):Y 1--* {S,:?lJ} is called a map of Y ® Q
226 Chapter 6

to S (this is a measurable map: {n, L} f-+ {S, ~} at any fixed y), such that
{w:f(y) E B} E L

for any B E~. Instead of a 'random map' the term a 'random function' is
usually used with a given value in S. In this case we call 'y' the argument of a
random function.
In the case when Y = T and y = t, t is time and a random function is called
a random process. If Y=T+={0,1,2, ... ,n, ... } or Y=T={ ... , -n,
- n + 1, ... , -1, 0, + 1, ... , n, ... }, then we deal with a random process with
a discrete time. If Yis a finite-dimensional Euclidean space IRd or a domain in
IR d , then f(y) is sometimes called a 'random field'.
Further, in this section we are interested in the case, when Y = T, i.e., we
consider only random processes. Notice that random processes are often called
stochastic processes.
In most cases it is convenient to identify the random process f(t, w) = w(t)
with the probability space {n, L, /l}. Indeed, we assume that n is a functional
space, w = w(t), t E T; and the IT-algebra L contains the whole set n of the type
{w: w(to) E B} for any to E T and B E~, and /l is any probability measure on n.
Thus, we see that one can naturally compare the random processf(t, w) = w(t)
with the probability space.
DEFINITION 6.20. Let q(t) = qt be a random process. Then by the notion of
random process qp - 00 < t < 00 in IRd we denote the triplet {n, L, /l}:

n= fl
-oo<t<oo

is the totality of paths y in IR d, i.e., y: IR f-+ IRd. For example,

for each SE T == IR, defines an IRd-valued random variable qs' where Ws denotes
a cross section of the sample path w at the time s. The random process qp
- 00 < t < 00, is equivalent to the family of random variables {qs}, sET == R
Any random process would be specified by giving the probability measure
on n; different probability measures define random processes of different
natures.
First we consider Gaussian processes. A Gaussian measure on Y == IR d ,
d = n = 1,2,. .. is a measure which is the transform of the measure with
density (2n1 2 )-dI 2 exp( -lxI2/212), under an affine transformation.
DEFINITION 6.21. The real random process qp t E T is called the Gaussian if
for any even m :;;;. 1 and any t k , k = 1,2, ... , m, tk E T; the sequence {qt" ... ,
qtm } has a joint normal distribution.
From the definition it follows that the characteristic function of this distribu-
The Basic Concepts of Random Processes and Stochastic Calculus 227

tion has the form:

and

Thus, all the particular distributions of the Gaussil:l,n process are defined by
two real functions: the mean value aCt) and the correlation function (covariance)
b(tl, t2):

The correlation function b(t 1, t 2) possesses the following properties:

(1) b(tl' t 2) = b(t2' td


(2) For any m, and real numbers Uk and points tk E T

L b(tk' tr)ukUr :;;?- O.


k,r = 1

Real functions possessing these properties are called the positive definite kernels
on T2. This definition is equivalent to the requirement that for any tr and tkE T
the matrix Ilb(tk' tr)ll, k, r = 1, ... ,m should be real symmetric and non-
negatively definite.
A similar definition of the vector Gaussian process with real components
can be easily formulated, the characteristic function of which has the form

where

aCt) = {a l (t), ... , an(t)}

are vectors belonging to ~n and b(tl, t 2) is a matrix with the elements


brs(tl' t 2), r, s = 1,2, ... , n.
For some problems it is interesting to consider the moments of the Gaussian
process. These moments may be obtained by the expansion of the characteris-
tic function to series. We restrict ourselves to the consideration of the central
moments of the scalar Gaussian process. Put
aCt) = 0; u = (u(l), ... , u(m)); B = Ilb(t k, tr)ll (k, r = 1, ... , m).
228 Chapter 6

Then

pZ p4 Z
= 1- 2 (Bu, u) + 2! 2Z (Bu, u) +
p2m
+ ... + (-1)m-
2m ,(Bu, u)m + ...
m.

from this it follows that:

m
IE { [ k ~ 1 d k ) qtk
Jzm-l} = O. (6.13)

Now we introduce the m-point momentum functions of t

. (t 1,···, t m )= 1E{[qt 1 Ji,


M·J 1 , . . . • 1m ... [q tm Jim} .
jl jz jm
The value of + + ... + is called the order of momentum function. The
momentum functions of the odd order equal zero
m
at L jk = 2s - 1, s = 1, 2, ...
k=l

Formula (6.13) can be rewritten in the form:


z . . 1
Mj, ..... iJt 1 ' t
•.• , m ) = (a m/au{' ... aU;;;')-2
m ,(Bu, u)m. (6.14)
m.
The momentum function of the second order coincide with the correlation
function
Mz(t) = M l l (t, t) = bet, t).
For the momentum functions of the fourth order, we have the following
expressions:

M 31 (t 1 , t z ) = 3b(tl' t1)b(t 1, t z );
+ 2b(tl' tZ)b(tl' t 3);
M Zll (tl, t z , t 3) = b(t l , tl)b(tz, /;3)
M ll11 (tt. t z , t 3, t 4) = b(tl, tZ)b(t3' t 4 ) + b(tl, t3)b(t2' t 4 ) +
+ b(tl, t4)b(tz, t 3)
etc.
The Basic Concepts of Random Processes and Stochastic Calculus 229

DEFINITION 6.22. The oscillator process is the family {q(t)}, - 00 <t< 00,
of the Gaussian random variables with the covariance
lE{q(t)q(s)} =texp(-It-sl).

LEMMA 6.1. Let SI' ... , Sn be real and ZI' ... , Zn E C. Then
n
L ZiZj[t exp( -lsi - sjl)] :;;.. 0.
i,j =1
Proof. By a direct calculation of the Fourier transform of exp( -I t I) and the
Fourier inversion formula,

texp(-It - sl) = (2n)-1 f dk eik(t-s) (1 + P)-1


so that

j,~ 1 ZiZj[t exp( -lSi - Sjl)] = (2n)-1 f Ijtl


dk Zj eiksj IZ (1 + kZ)-1

which is clearly positive.


Often, the process we have called the oscillator process is called the

'Ornstein-Uhlenbeck velocity process' since Uhlenbeck and Ornstein (1930)
introduced a process f(t) with differentiable paths so that df(t)/dt = q(t).

6.9. Stochastic Processes with Independent Increments

DEFINITION 6.23. A random process {q(t), t E T} E IRn is called the process


with independent increments, if for any m, tk E T; t 1 < t Z < ... < t m , the random
vectors q(a), q(t d - q(a), ... ,q(tm ) - q(tm _ d are mutually independent. The
vector q(a) is called the initial state (value) of the process q(t), and its
distribution is called the initial distribution of q(t). In order to define the
process with independent increments it is sufficient to give an initial distri-
bution P o(B) and a set of distributions P(t, h, B) (t :;;.. 0, h > 0, BE [IAn), where
[IAn is the O'-algebra of the Borel sets IRn and P(t, h, B) is the distribution of the
vector q(t + h) - q(t). Indeed, if these distributions are given, then any joint
distribution of vectors q(t 1 ), q(t z ),"" q(t m ) is synonymously defined by the
formula
230 Chapter 6

Here B - z denotes the set: {x: x = y - z, Y E B}. As concerns the initial


distribution P o(B), it can be arbitrary. On the other hand, it cannot be
guaranteed that to any given set of the distribution P(t, h, B) there corresponds
some process with independent increments. In order to have this, it is
necessary and sufficient that P(t, h, B) should possess the following property: at
any m and any a = to < t 1 < .... < tm = h + t; P(t, h, B) is a distribution of the
sum of the independent vectors q(t 1), q(t z ), ... ,q(tm), where q(tk) has the
distribution P(t k _ 1> tk - tk _ 1, B). Indeed, if this condition is fulfilled, then the
set of distribution (6.15) is satisfied by the conditions of co-ordination.
Therefore, the Kolmogorov theorem is used and there exists a random process
with finite-dimensional distribution (6.15). From the form of these distri-
butions it follows that this process has independent increments.
It is convenient to study the processes with independent increments by
means of the characteristic functions. Let us define

c(t, h, u) = r P(t, h, dx) exp[i(u, x)].


J~n
We have called the function c("') the characteristic function of the process with
independent increments. It defines fully a joint distribution of the differences:
(6.16)

Indeed, joint distribution of the sequence of vectors (6.16) has the characteris-
tic function C(t1,"" tm; u\ ... , um) determined by the expression:
m
c(t 1, ... , tm; U1, . .. , Um) = TI C(tk _ 1, iltk, Uk),
k =1

Thus, in order to give the process with independent increments, in the wide
sense, in addition to P o(B), it is sufficient to define c(t, h, B). The condition

c(t, h1 + hz, u) = c(t, h1' u) c(t + h1' hz, u)


is necessary and sufficient for c(t, h, u) to be the characteristic function of the
process with independent increments.
DEFINITION 6.24. The process with independent increments is called homo-
geneous, if the differences q(t + h) - q(t) have the distributions which do not
depend on time, i.e., P(t, h, B) = P(h, B). The homogeneous process is called
The Basic Concepts of Random Processes and Stochastic Calculus 231

stochastic continuous, if

lim P(h, 0.) = 0,


for any sphere U. = {x: Ixl < e}, e > O.
Now we give some of the properties of the characteristic function of the
homogeneous stochastic continuous process with independent increments:
(a) The characteristic function of the homogeneous process with independent
increments satisfies the equation:
c(hi + h2' u) = c(hi' u) c(h2' u). (6.17)
In particular, at any even n: c(n· h, u) = [c(h, u)y.
(b) The characteristic function of the homogeneous stochastic process is non-
zero everywhere and its value can be represented by
c(t, u) = exp[tg(u)], (6.18)
where g(u) is a synonymous and continuous function. From form (6.17) it
follows that

g(u) = lim [c(t, u) - 1]/t (6.19)


t~O

and convergence is uniform in each restricted sphere lui ~ N, 0 < N < 00.

THEOREM 6.6 Let c(t, u), t > 0, u E lR n, be a family of the characteristic


functions such that limit (6.19) will exist uniformly in an arbitrary sphere
lui ~ N, N > O. Then in {lRn, ~n} there exists ajinite measure M(B), BE ~'and

1 ['()
nonnegatively dejined in lR n operator b and vector a such that

g(u) = i(a, u) - z(bu,


1 u) +
~n
M(dz) e' u,z - 1- i(u, z) 2
1+lzl
J
X

x (1 + IZI2)lzl-2. (6.20)
Proof was given by Gihman and Skorokhod (1975).
THEOREM 6.7. If q(t)E lR n, t > 0 is a homogeneous and stochastic continuous
process, then the characteristic function c(t, u) of difference q(t + s) - q(s) has
the form
c(t, u) = exp[tg(u)], (6.21)
where g(u) is given by formula (6.20).
Now we consider some particular cases of formula (6.20):
(a) b = 0 and M(B) == O. In this case c(t, u) = exp[it(a, u)] which corresponds
to the characteristic function of the degenerated distribution located at the
232 Chapter 6

point ta E IR". Thus, with probability one, q(t) = qo + at and the point q(t)
would be found in the uniform motion with the velocity a.
(b) a = 0, b = 0, and M represents the measure located at the point Zo
mass of the value m. Characteristic function (6.21) in this case has the form

It is easy to see that the increment q(t) - q(O) can be represented in the form
q(t) - q(O) = zo[v(t) - mtlzo l- 2 ],
where vet) is the Poisson process with the mean value
lE{v(t)} = mt(1 + IZoI2)lzol-2.
(c) M(B) == 0. In this case the difference q(t + s) - q(s) has the normal
distribution with the mean value a and the (covariance) correlation matrix bt,
so that, if, for example, q(O) = 0, then the process q(t) is the Gaussian. The
considerable process is also called the process of the Brownian motion. Thus
the Brownian motion can be considered as a continuous process with inde-
pendent increments. It is also a Gaussian process. If q(t) is a one-dimensional
process, and b = 1, a = 0, then the process of the Brownian motion is called
the Wiener process.

6.10. Markov Processes

Before giving a formal definition of the Markov processes we recall a simple


model starting from the theory of the Brownian motion. In 1828, an English
botanist R. Brow observed under a microscope that small particles, pollen of
some flower, suspended in a liquid move chaotically, changing position and
direction incessantly. To describe such a phenomenon, we shall consider the
transition probability pet, x; to, r) that a particle starting from the position x at
time to belongs to the set r at a later time t.
The introduction of the transition probability (sometimes it is called the
transition function) pet, x; to, r) is based on the fundamental hypothesis that
the (stochastic) chaotic motion of the particle after the time moment t is
entirely independent of its past history before the time moment t. The
hypothesis that the particle has no memory of the past implies that the
transition probability P satisfies the equation

pet, x; s, r) = Is pet, x; u, dy) P(u, y; s, r)' (6.22)

for t < u < s, where the integration is performed over the entire space S of the
chaotic movement of the particle.
The Basic Concepts of Random Processes and Stochastic Calculus 233

The process of evolution in time governed by a transition probability


satisfying (6.22) is called the Markov process, and equation (6.22) is called in
mathematics (in physics) the Chapman-Kolmogorov equation (the
Smoluchowski equation). The Markov process is a natural generalization of the
deterministic process for which P(t, x; s, r) = 1 or = 0 according to y E r or
y ¢ r; that is, the process in which the particle at the position x at the time
moment t moves to a definite position y = y(x, t, s) with the probability of one
at every fixed later time moment s. The Markov process P is said to be
temporally homogeneous if P(t, x; s, r) is the function of difference (s - t)
independent of t. In such a case, we are dealing with the transition probability
P(t, x; r) that a particle at the position x is transferred into the set r after the
lapse of t units of time. Equation (6.22) then becomes

P(t + s, x; r) = Is P(t, x, dy) P(s, y; r) (6.23)

for t, s > o.
A more developed theory deals with the object of the motion trajectory
x(t) = x,. The random character of motion is expressed by the mathematical
assumption that X t = x/w), where w belongs to some set n on which there is
given a family of probability measures Ilx (as the set n above is called the space
of elementary events). A set A for which the value of Ilx(A) is determined is
called 'events' and the quantity of Ilx(A) is interpreted as the probability event
A by the assumption that the motion is started from the point x. Then a
transition probability is given by the formula: P(t, x; r) = Ilx (x t E r). In the
case of the Brownian motion the phase space, on which the value of the
function x(t) is given, is some domain of the three-dimensional Euclidean
space. Generally speaking, this space is the set S in which some system of
'measurable sub-sets' f!,D is picked.
The basic condition connecting the function x/w) and the measure Ilx is the
Markov principle: the future depends on the past only through the present. The
mathematical (equation) expression of this principle is the Chapman-
Kolmogorov equation (6.23). More precisely it means that at the given value
of X t prognosis for further motion of a particle does not depend on the
character of motion before the moment t. We recall that the Markov pro-
cesses, for which the condition of independence of the future on the past
known at the present is fulfilled not only for the constant moment but for the
definite class of random moments ((w), are called strict Markov processes.
A wider and more adaptable formulation of the Markov principle is obtained
if under 'the past' for the moment t one understands the family of all the events
L t which are observed before the moment t. Up to now we have assumed that
X t is determined for all t > O. However, many problems leads to the processes,
for which xt(w) is only given on some interval [0, nw)]. The random variable
((w) is called the truncation moment. Thus, we lead to a definition of the
Markov process as a family X t = {Xt, (; L t, Ilx}.
234 Chapter 6

DEFINITION 6.25. Let us give:


(a) The function '(w) E Q taking the value on [0, ex));
(b) The function x(t, w) = xt(w) defined for WE Q and t E [0, ~(w)] taking
the value from the phase space {S, ~};
(c) For each t » 0, the O"-algebra L t c Qt = {w: '(w) > t};
(d) For each XES, the function IlAA) defined on some O"-algebra LO c Q,
and L t C L ° at all t » 0.
Then we say that these elements define the Markov process X t = {Xt", L t ,
Ilx}, if the following conditions are fulfilled:
(1) If t .,;;;; s and A E Lo then {A:' > s} E Ls.
(2) {XtE r} E L t (t » 0, r E ~).
(3) Ilx is the probability measure on the O"-algebra LO.
(4) At any t » 0, r E ~ the function
P(t, x, r) = Ilx{X t E r} (6.24)
is the ~-measurable function of X.'
(5) P(O, x, S\x) = 0.
(6) For any t,h »0, rE~, Ilx{Xt+hErJLt} =P(h,xt,r) with the pro-
bability Qt, Ilx·
(7) For every WE Qt there is found such w' E Q that
Xs(W') = xt+s(w), (0 .,;;;; s < '(w') = '(w) - t).
According to (4) condition (6) is equivalent to the following requirement:
(4') For any t,h »0, rE~, AELt

Ilx{A,Xt+hEr} = Lllx(dW)XA·P(h,Xt(W),n.

The Markov process is called without truncation, if '(w) = + 00 for every


WE Q. At the fixed W the function xt(w), {t E [0, ,(w))} is defined on the
trajectory space S of the process corresponding to the elementary event w.
The O"-algebra L t can be clearly represented as the family of events which are
observed during the time interval [0, t]. From the above definition, we see that
the meaning of conditions (1)-(3), and (5) is clear. Condition (4) has a technical
character. Condition (6) means that the prognosis of the particle position
through the time moment h with the knowledge of its position at the moment t
depends neither on the value of t, nor on the phenomena which have been
observed during the time interval [0, t]. This is just the Markov principle: the
independence of 'the future' on 'the past' by the knowledge of 'the present'.
Condition (7) means some homogeneity of the family of trajectories of the
process. More precisely, let q>(u) E S be a function defined on some interval
[0, A), and let 0";;;; t < A. Then, we define the function Ttq>(u) = q>(t + u)
(0 .,;;;; u < A - t) by a displacement of the function q> on t. Condition (7) requires
that a set of trajectories of the process should be invariant with respect to all
The Basic Concepts of Random Processes and Stochastic Calculus 235

the displacements. In this connection we notice the following. In a suitable


function space F, P(t, x, r) gives rise to the linear transformation T t :

(Tt f) (x) = 1 P(t, x, dy)f(y), f E F,

so that, by (6.23), the semi-group property holds: T t + s = T t Ts (t, s > 0).


A fundamental mathematical problem in the statistical mechanics concerns
the existence of the time average limt ~ w t - 1 S~ ds Ts f. In fact, let S be phase
space of a mechanical system governed by the classical Hamiltonian equations
whose Hamiltonian does not contain the time variable explicitly. Then the
point x of S is moved to the point Yt(x) of S after the lapse of t units of time in
such a way that, by the classical theorem due to Liouville, the mapping
x --> Yt(x) of S onto S, for each fixed t, is an equi-measure transformation, that
is, the mapping x --> Yt(x) leaves the 'phase volume' of S invariant. In such a
deterministic case, we have (TJ)(x) = f(Yt(x)) and hence the ergodic hypothesis
of Boltzmann that the time average of any physical quantity coincides with the
space average of this physical quantity, is expressed, assuming Sdx < 00, by

lim t -1
t
dsf(y,(x)) =
1S1
dxf(x)

Jo
[
t~w dx
s
for all f E F, dx denoting the phase volume element of S. A natural generali-
zation of the equi-measure transformation x --> Ylx) to the case of the Markov
process P(t, x, r) is the condition of the existence of an invariant measure
,u(dx):

l,u(dX)p(t, x, r) = ,u(r)

for all t > 0 and all r.


We notice that the transition probability (or transition function) P(t, x, r) is
a considerably simpler object than the Markov process. Therefore, for the
Markov process it is very important to know its transition function P(t, x, r).
DEFINITION 6.26. Let us consider an arbitrary phase space {S, ~}. The
function P(t, x, r) (t ~ 0, XES, r E~) is called a transition function, if the
following conditions are fulfilled:

(i) For the fixed t and x, P(t, x, r) is a measure on the cr-algebra ~;


(ii) For the fixed t and r, P(t, x, r) is ~-measurable in x;
(iii) P(t, x, r) ~ 0, P(t, x, S) « 1;
(iv) P(O, x, S\x) = 0;
(v) P(t + s,x,r) = Lp(s,x,dy)P(t,y,r) (s,t ~O) (the Chapman-Kolmo-
gorov equation).
236 Chapter 6

From (v) and (iii) we have

P(u + t, x, S) ~ Is P(u, x, dy) = P(u, x, S) (u, t > 0).

Hence, P(t, x, S) is an unincreased function of t and there exists the limit


P( + 0, x, S). We can say that the transition function P(t, x, r) is normal, if
P( + 0, x, S) = 1 for any XES. We call it conservative, if P(t, x, S) = 1 for any
t, XES.
The function P(t, x, r) defined by formula (6.24) satisfies conditions (i)--(v).
It is called the transition function of the Markov process. We introduce the
following conditions:
Condition L(r): For any u > °
lim Sup P(t, y, r) = 0.
y --+ 00 t";;;; u

Condition N(r): For any e > °


lim t- 1 Sup P(t, x,
t-O XEr
0 e (x» = °
(through O.(x) we denote a complement of a-neighborhood of the point x, i.e.,
the set {y:p(x,y) >e}, where p(x,y) is called the metric or the distance
function).
THEOREM 6.8. Let P(t, x, r) be the normal transition function on the half-
compact {S, IC} and for every compact subset of r the conditions L(r) and N(r)
are fulfilled. T hen, there exists the continuous Markov process with the transition
function P(t, x, r).
Proof This can be found in any textbook on the theory of the Markov
process, in particular, see Dynkin (1965).
DEFINITION 6.27. Let J1 be some measure on {S, ~}. The function p(t, x,y)
(t > 0, x, YES) is called the transition density, if the following conditions are
fulfilled:

(1) p(t, x,y) > °


(t > 0; X,YE S).
(2) For fixed t, p(t, x, y) is ~ 0 ~-measurable function in x and y

(3) Is J1(dy) p(t, x, y) = 1, (t > 0, XES),

(4) p(u + t, x, z) = Is dy p(u, x, y) p(t, y, z), (u, t > 0; x, Z E S).


The Basic Concepts of Random Processes and Stochastic Calculus 237

It is easily verified that if p(t, x, y) is the transition density, then the formula

P(t,x,r) = {SrdYP(t,X,Y)
Xr(x)
att>O
at t = ° (XES,rE~)
defines the transition function. This function is normal if

lim
t~O Jsrdy p(t, x, y) = 1, (x E S),

and conservative if

I dy p(t, x, y) = 1, (t > 0, XES).

EXAMPLE 6.1. Let {S,~} be any measurable space and let TI(x, r)
(x E S, r E~) be the function satisfying the following conditions:
1. For any XES, TI(x, r) is a probability measure on ~.
2. For any r E ~, TI(x, r) is a ~-measurable function on S.
Define the function TIn(x, r) by formulas
TIo{x, r) = Xr(x),

TI.(X, r) = I TI{x, dy)TI n - 1(Y, r), (n > 1)

and put
antn
L -, e-
00

P(t, x, r) = at TIn (x, r), (6.25)


n= 0 n.
where a is some positive constant. It is verified that the series on the right-
hand side of (6.25) converges and determines the transition function. We call it
the Poisson transition function.
EXAMPLE 6.2. Let S be a numerical (axis) line, and let ~ be the (J-algebra,
obtained by a family of all the intervals, and v is some constant. The formula
P{t, x, r) = Xr(x + vt) gives the conservative transition function. It is said that
this function corresponds to a uniform motion with the velocity v.
EXAMPLE 6.3. Let S be the n-dimensional Euclidean space IRn and let ~ be
the (J-algebra of all the Borel sets of this space. Then put
p(t,X,y) = (2nt)-n/2 exp{ -(y - x)2/2t}. (6.26)
It is easy to verify that p(t, x, y) is a transition density with respect to the
Lebesgue measure. We call it the Wiener transition density and the cor-
responding transition function - the Wiener transition (function) probability.
238 Chapter 6

EXAMPLE 6.4. The formula


p(t, x,y) = (2nt)-1/2{exp[ -(y - x)2j2t] + exp[ -(y + x)2j2tJ) (6.27)
defines the transition density on half-axis [0, (0) (with respect to the Lebesque
measure). The corresponding transition function is conservative.
EXAMPLE 6.5. The formula
p(t, x, y) = (2nt) -1/2 {exp[ - (y - X)2 j2t] - exp[ - (y + X)2 j2t]} (6.28)
gives the transition density on the half-axis (0, (0). The corresponding tran-
sition function is not conservative but is normal.
Finally, we want to give some formal definition of the Markov processes and
the time-reversed one. The Markov process in IRn, X(t), - 00 < t < 00, is a
random process such that
Cond{XtE B I X(tm) = Xm,.··, X(t 1) = xd = Cond{XtE B I X(t m) = x m},
Xt == X(t)
holds with the probability one for any time series - 00 < t1 < ... < tm < t,
where B is a Borel set of IRn. Since one can regard the condition probability
with respect to a random variable as a function of this variable, then we
denote
P(t, x; U, B) = Cond{XtE B I Xu = X(u) == Xu = x}, t > u.
This is nothing but the transition probability or the transition function of the
Markov process defined above, which, therefore, satisfies the following
Chapman-Kolmogorov equation:

P(t, x; U, B) = fp(t, y; s, B) P(s, x; U, dy), t > s > u.

Further, we shall prove that the time-reversed process of the Markov


process is also the Markov one. The time-reversed process of the Markov
process is defined to be a random process Xi, - 00 < t < 00 such that
Xi = X( -t) = X -t holds with the probability one.
For technical simplicity we replace IRn by 7L n (7L denotes the totality of
integers): this corresponds to a lattice approximation.
THEOREM 6.9. Let X" - 00 <t< 00, be the Markov process in 7L", i.e.,
Cond{X(tm) = xmIX(tm- d = xm - 1> ••• , X(td = xd
= Cond{X(tm) = xmIX(tm- d = xm- d
for any - 00 < t1 < ... < tm < 00 and Xl' ... , Xm E 7L n • Then we have
Cond{X(t1) = xdX(t 2) = X2,··., X(tm) = xm}
= Cond{X(t 1) = x 1IX(t 2) = x 2}, (6.29)
The Basic Concepts of Random Processes and Stochastic Calculus 239

i.e., the time-reversed process xi, - 00 < t < 00, is also the Markov one.
Proof The left-hand side of equation (6.29) can be manipulated as
,u{X(t l ) = Xl , ... ,X(tm) = Xm}
,u{X(t Z ) = X Z "'" X(tm) = Xm}
= Cond{X(tm ) = Xm I X(t m - l ) = Xm-d'" x
x Cond{X(t z ) = X z I X(t l ) = xd x
,u{X(t l ) = xd
d ...
x~--~--~--~--~------~
Cond{X(tm) = Xm I X(tm - d = Xm - x
x Cond{X(t3) = X3 I X(t z ) = Xz} ',u{X(t z ) = XZ}
,u{X(t l ) = Xd
= Cond{X(tz) = X2 I X(t l ) = xd' ,u {Xt() }
z =X Z

= Cond{X(tl) = xII X(tZ) = Xz}


which is identical to the right-hand side of equation (6.29).
THEOREM 6.10. Let P(t, x; s, y), t > s, be a transition probability of the

MarkovprocessX t , - < t < oo;p(t, X) be aprobabilitydistribution ofXt . Then a
00
transition probability of the time-reversed Markov process Xi, - 00 < t < 00;
P*(t, y; s, x), t > s, is given by
P*(t,y; s, x) = p( -t,y)P( -', x; -t,y)p(-s, X)-l. (6.30)
Proof A straightforward manipulation yields

P*(t,y; s, x)= Cond{Xi = y I X: = x} = Cond{X -t = Y I X -s = x}

,u{X-t =y,X- s = x} ,u{X- s = x,X_ t =y} ,u{X-t =y}


,u{X -s = x} ,u{X-t=y} ,u{X-s=x}

_ { _} Cond{X -s = x I X - t = y}
- ,u X - t -Y ,u {X -s = X }

= p( -t,y)P( -s, x; -t,y)p( -s, X)-l .



6.11. Wiener Processes

We begin by giving some simple definition of this process, based on. its
transition probability (or function) P(·). We consider the Wiener transition
function P( .) (6.26) defined in Example 6.3 of Theorem 6.8. We show that for
every compact r this function is satisfied by the conditions L(r) and N(r) (see
240 Chapter 6

Section 6.10). Indeed, as is easily seen

P(t,x,r)<2(nt)-n/2 exp {_(!x!-a)2/2t} 1 dy,

for Ix! > a, where a = SUPYEf !y!. From this we see that the condition L(r) is
fulfilled.
Further

P(t, x, O,(x)) = (2nt) -n/2 f dy exp( - y2/2t).


y>e

After simple transformations we get


P(t, x, O,(x)) = cnF(c/J"i},
where is some constant and

1
Cn

dx xn - exp( - x 2/2).
00
F(r) = 1

By the Lopital rule


F(r) '" rn - 2 exp( -tr2), for r -+ 00

and therefore at t -+ °
Sup P(t, x, Oix)) '" cnc n - 2 t-(n - 2)/2 exp( -c 2/2t).
XES

From this we see that P(t, x, r) satisfies condition N(r). According to Theorem
6.8 some continuous Markov process without truncation corresponds to this
Wiener's function. We call it the Wiener process.
Transition function (6.27) from Example 6.4 of Theorem 6.8 is also con-
servative and satisfies the conditions of Theorem 6.8. The corresponding
continous Markov process without truncation is called the Wiener process
with reflection at point 0. We now consider transition function (6.28) de-
termined on the half-axis (0, (0) (Example 6.5). A direct calculation shows that
for every compact r the conditions L(r) and N(r) are fulfilled in this case.
From Theorem 6.8 we conclude that some continuous Markov process
X t = {x o " L t , ,ux} corresponds to function (6.28). We shall call this process
the Wiener process with a truncation at point 0. Thus, we arrive at a definition
of the Wiener process.
DEFINITION 6.28. We denote through S == ~n the n-dimensional Euclidean
space and through [YJ a (J-algebra of all the Borel subsets of this space. Then
under the Wiener process one can understand the standard continuous process
X t without-truncation in space {S, [YJ} with the transition function

P(t, x, r) = 1 dy Pt(Y - x), (t:;;;' 0, XES, r E [YJ),


The Basic Concepts of Random Processes and Stochastic Calculus 241

where
Pt(X) = (2m)-n/2 exp( -x2/2t).
According to Example (b) of Theorem 6.7 we can regard the Wiener process
°
as a Gaussian process with a mean value and the correlation t. By definition,
it is also the Brownian motion. To the Wiener process an object-definition
based on the notion of random walks can be given. It is known that the Wiener
process X t = q(t) is the limit of elementary random walks. Let Yn be the nth co-
ordinate function and Xn = L~ = 1 Ym. The family of random variables
{Xn},~)= 1 is called a random walk. Since IE{YnYm} = (jnm we have IE{X;} = n so
that the variable Xn = n -1/2 X n at least have a chance of having a limit. Due to
the well-known De Moirre-Laplace limit theorem X n approaches a Gaussian.
THEOREM 6.11 (Central limit thoerem). Let {Yn}:= 1 be a family of inde-
pendent, identically distributed random variables with IE{Yn} = 0, lE{y;} = 1. Let
Xn = n -1/2 L~ = 1 Ym. Then Xn approaches a Gaussian random variable Xw with
the covariance (correlation) one in the sense, that for any continuous, bounded
function on IR,

1E{j(Xn )} --> (2n)-1/2 f dy f(y) exp( _y2/2).


Proof By a simple limiting procedure (going through f E L as an In-
termediate step), it is sufficient to prove that
lE{exp(iIXX n )} --> exp( -1X 2 /2)
for all real IX. But

lE{exp(iIXX n )} = [F(IX/Jn)]n,
where
F([3) = lE{e iPY }.
Since IE{Yn} = 0, lE{y;} = 1, we have that
[F([3) - 1J[3-2 = -!
as [3 --> °
(by using the dominated convergence theorem) so that
lE{exp(iIXX n )} = [1 - !(1X 2/n) + 0(1/nW = exp( -1X 2 /2) + 0(1)
by the compound formula of interest.
Intuitively q(t) is lim,,~<XJ n- 1 / 2 X[ntj, where [aJ is an integral part of a. Thus,

each q(t) should be Gaussian with covariance t. To find the covariance, we
note that X n - X m and X m are independent if m < n. Thus, we expect that
q(t) - q(s) should be independent of q(s) for s < t, i.e.,
IE {q(s)[q(t) - q(s)]} = ° so IE {q(t)q(s)} = s for s < t.
Therefore, we define the following
242 Chapter 6

DEFINITION 6.29 The Wiener process (or Brownian motion) is the family
{q(t)}o <; t of Gaussian random variables with the covariance
IE {q(t)q(s)} = min(s, t).

LEMMA 6.2. Let 0';;;; S1 .;;;; Sz .;;;; ... .;;;; s" and Z1"'" Z"E C. Then

I" Z;Zj min(s;, s) > 0.


;,j ~ 1

Proof Note that (with So = 0)

,In Z;Zj min(s;, s) = ,2:"(S; - I ",I


Si _ 1) Zj IZ
'. J ~ 1 , ~ 1 J~ 1

is obviously positive.
This lemma can also be proved by noting that q(t) is really the limit of

n -l(Z X [Ilt] as far as the joint probability distributions of the finitely great
number of q's are concerned. Notice that
IE{ [q(t) - q(s)]Z} = t - s, (t > s).
Moreover, by writting q(t) = q(s) + [q(t) - q(s)] and using the independence of
q(t) - q(s) from {q(u)}o<;u<;s' we see that
lE{q(t) I {q(u)},O';;;; u';;;; s} = q(s), (t > s),
so that q(t) is a martingale.
°.; ;
°.; ;
THEOREM 6.12. If x(t) is a process for t with a continuous version so that
for s.;;;; t
lE{x(t) - x(s) I {x(u)}, 0';;;; u';;;; s} = 0,

1E{[x(t) - x(s)]ZI {x(u)},O';;;; u';;;; s} = t - s.


Then x(t) is the Brownian motion.
An interesting realization of the Brownian motion comes from the fact that
if XS is the characteristic function[O, s], then
<Xs. Xt> = min(s, t).
Thus, if ¢ is the Gaussian process associated to Lz(O, co), then q(s) = ¢(xJ is
the Brownian motion. ¢ is often called a white noise and the formal relation
dqjds = ¢(s) is often expressed by saying that 'the derivative of Brownian
motion is a white noise'.
Finally, we shall give some construction of the Wiener process in IR" due to
Nelson (1964). The base (sample) space of the Wiener process is taken
n = II, ~n, - 00 < t < 00, where ~" denotes a one-point compactification of
The Basic Concepts of Random Processes and Stochastic Calculus 243

~". With the usual product topology, 0 becomes compact Hausdorff. Let
Pt(x, d"y) = (4nvt)-"/2 exp{ -Ix - Yl2j4vt} d"y,
t> 0, be the n-dimensional Gaussian measure centered around x E ~", where v
stands for the diffusion constant.
By the concept of a cylinder function on 0 we define a functionf:O 1-+ R of
the form
f(w) = F(w(t-m),"" W(t-1), w(t 1), .. · "w(tm)),
for Lm < ... < L 1 < 0 < t1 < ... < tm . First, we define the Wiener integral of
a cylinder function f by

Iw(f) = fPt-m+l-t-m(X-m+1,dX-m)"'P-t-l(X,dX-d X

For continous F; 1R2m 1-+ ~, there exists integral (6.31) because we have
Iw(1) = 1).
Let C(O) be the totality of continuous functions on 0 and CCyj(O) that of
continuous cylinder functions on 0 (a cylinder function f is continuous if the
function F is continuous). C(O) with the usual supremum norm is a Banach
space. Secondly, we define the Wiener integral of a continuous function
9 E C(O) as follows.
By the Stone-Weierstrass theorem, we can approximate any function in
C(O) uniformly by those in CCyj(O). The mapping Iw:fl-+lw(f) defines a
bounded linear functional on CCyj(O) of a positive type, i.e.,

IIw(f) I .;;; Ilfll = Sup If(w) I, Iw(f) ~ 0


WEQ

if function f ~ O. Then, the mapping Iw has a unique extension to a bounded


positive linear functional on C(O). The extension should also be denoted over
Iw·
The Wiener integral of 9 E C(O) is defined as Iw(g). By the Riesz-Markov-
Kakutani theorem, we find that there exists a regular normalized measure Jlxw
on 0, indexed by x E ~", such that

Iw(g) = I Jl~(dw)g(w)
holds for \;/g E C(O). This is the Wiener measure.
Adopting the Wiener measure as a probability measure we can define the
Wiener process Wt, - 00 < t < 00 in ~" by the triplet {O, ~(Jl~), Jl~}.
Concerning the domain of the Wiener measure Jl~ we have the following
theorem.
244 Chapter 6

THEOREM 6.13 [Wiener (Nelson, 1964)]. Let ¢ be the totally of continuous


paths y such that y(t)E ~n for - 00 < t < 00; then J.1~(¢) = 1 for x in ~n.
Consequently, L(J.1~) is taken as a a-algebra of Borel sets of ¢, i.e., ~(¢).
We conclude this section with the following two theorems.
THEOREM 6.14. The function of the form

u(t, x) = fJ.1~(dw) uo(w t )

is a solution to the Cauchy problem of the following equation


au =
-
.
dlV grad u
at V

with the initial condition u(O, x) = uo(x).


THEOREM 6.15. (Feynman-Kac). A solution to the Cauchy problem of the
equation

~~ = v div grad u + U(t, u)u (6.32)

u(O, x) = uo(x), (6.33)


is given by the Wiener integral

u(t, x) = fJ.1~ (dw) uo(w t ) exp{L U(w" s) ~~ }- (6.34)

6.12. Functionals of Stochastic Processes and Stochastic Calculus

Interest in the study of the functionals of stochastic processes has grown in


recent years. One of the powerful methods in this field is stochastic integrals
and based on them stochastic differential equations. They allow wide classes of
random processes to be constructed starting from the simple process - the
Brownian motion - and give a wide possibility of applications of the pro-
bability methods to stochastic partial differential equations. The methods of
stochastic integrals are intensively used in the optimum control theory.
The theory of stochastic differential equations was constructed by K. lto and
I. I. Gihman. N. Wiener and S. N. Bernstein were their predecessors. The main
results of the stochastic integrals and stochastic differential equations were
expounded in a series of monographs (Doob, 1953; Dynkin, 1965; Gihman and
Skorokhod, 1972, 1975; McKean, 1969). The functional integration and its
application in quantum physics was considered by Simon (1979); Gel'fand and
Yaglom (1956); Glimm and Jaffe (1981). Therefore our consideration of these
problems will be very brief.
The Basic Concepts of Random Processes and Stochastic Calculus 245

Here we present an elementary treatment of some calculation aspects of


stochastic integrals which may play an important role in the stochastic
quantization procedure.
Notice that a stochastic integral may be understood as a functional of the
Wiener process. Thus, our interest in the Wiener process b(t) derives from the
fact that it is intimately connected with the definition of stochastic integrals via
the diffusion process. The diffusion process corresponding to the operator
Ho = tA (A is the Laplacian) was constructed by Wiener (1923) and it is often
called the Wiener process. According to Wiener this construction may extent
to any elliptical operator L which succeeds in constructing a fundamental
solution of the parabolic differential equation
aUt/at = Lut.
The simplest probability method for construction of the diffusion process
without truncation was proposed by Ito. Let X t be a one-dimensional Wiener
process and Xt be the diffusion process with a generating differential operator
1 2 d2 d
L = 2: a (x) dx 2 + a(x) dx'
Ito showed that both processes can be realized on the same space n of
elementary events so that ilx = {Lx and the trajectories x(t) and x(t) are-
connected by the relation

x(t) = Xo + I ds a(x(s)) + I dx(s)a(x(s)). (6.35)

On the right-hand side of this equation stands the so-called stochastic integral,
the general definition of which was formulated by Ito.
Equation (6.::: can be written in the differential form:
dx(t) = a(x(t)) dt + a(x(s)) dx(s). (6.36)
The last relation represents a stochastic analog of the usual differential
equation. The diffusion process theory based on Wiener's and Ito's ideas, was
given in many works, particularly in McKean's (1969) monograph. Roughly
speaking, the diffusion process theory comprises differential and integral
calculi based on the Brownian motion, with the only difference being that for
the calculation of a differential of a smooth function f depending on the
Brownian trajectory t ~ x(t) = bet) two terms of the power expansion should
be taken and db 2 replaced by dt:
df(b) = f'(b) db + tJ"(b) db 2 = f'(b) db + tJ"(b) dt
or

e
Jo
dbf'(b) = feb) /t _
0
~
2
e
Jo
dtf"(b).
246 Chapter 6

This calculus possesses some new outlines, for example, the role of exponential
is played by exp(b - tt) instead of the usual eb and the ordinary formal rules of
calculus do not hold for the stochastic integral, for example,

f db(s)·b(s) = Hb 2(1) - 1],

not b 2(I)j2. Note that b(O) = O. The main advantage of this method is.that any
smooth diffusion t ~ x(t) can be considered as a nonanticipatory functional of
the Brownian trajectory x(t), so that x(t) is a solution to stochastic differential
equation (6.36) with smooth coefficients a and (J.
The main goal of the stochastic calculus is to define dbf(s) or more J
J
generally dbg(b(s)). This integral cannot be understood as a Stieltjes integral,
since the curve b(t) is not rectifiable. Ito's idea plays an important role in the
definition of integrals of this type. The basic idea of Ito's definition is

This integral illustrates the difference between differentials pointing into the past
and the future. This fact is very important in Nelson's stochastic quantization
method. We define

I+(n) = m~l b(;)[b(;)-b(m; 1)}


L(n) = m~l b(m; 1)[b(;)_b(m; I)}
We have that for any n, I +(n) + I _(n) = [b(I)]2, since

and the sum telescopes. On the other hand, by

lim f(b, n, 2)
N~oo
= 2n b
lim I
n~oo k= 1
(k)
2
'(k - 1)
1 n - b ~ 12 ~ 1

we have I +(n) - I _(n) ~ 1 pointwise as n ~ 00. It follows that

This equality shows the difference between I + and 1_. From this we see that
the ordinary formal rules of calculus do not hold for the stochastic integral:

f dbb(s) = t[b 2(1) - 1], at b(O) = O.


The Basic Concepts of Random Processes and Stochastic Calculus 247

For suitablef's functions of s alone, Paley et aT. (1933) succeeded in defining


integral:

G(f) == f db(s) f(s),

in the following way. Iff E C 1 and f(1) = f(O) = 0, then we define

G(f) = - f ds f'(s) . b(s),

by a formal integration by parts. With this definition

IE{G 2 (f)} = ff ds dtf'(s)f'(t)min(t,s) = f dslf(s)1 2,

since o21os ot[min(s, t)] = <5(s - t).


Let f be the C 1 function M IR with I' = df Idt and f bounded. Then the
limit on the right-hand side of (6.37) exists in L 2 -sense and (11·11 = L 2 -norm)

(6.38)

This was proved by Simon (1979). Now we extend the definition of the integral
Sdsf(b(s)) from C I-function to an arbitrary function f with

which is finite; for example f is bounded near zero and L2 will certainly suffice.
This integral is in L2 for any such f. We can actually go one step further: Let
Xa be a characteristic function of ( - a, a) and suppose that

IE {f ds Xa(b(s)) j2(b(S))} < 00.

Then for each a, we can define S6 ds f ·Xa(b(s)). We claim that for virtually
every b(s) there exists an integral

al~~ f ds Xa(b(s))f (b(s)), (6.39)

pointwise, and which is finite. For b(s) integral (6.39) is continuous on [0, 1]
and so bounded. Thus for any fixed b(s) expression (6.39) is independent of a
for a ----* 00.
Now we can clearly use the d-dimensional Brownian motion and define
S6dbf(b(s)) and we can easily define g in place of S6. It is more important for
the applications we have in mind, that we can define S6doo f(oo(s)), for the
d)1w-Wiener (measure) paths. The construction is identical with the one stated
248 Chapter 6

above, but now

lE{f dS1f(W(S)W} = f ddxlf(xW,


where w(s) is the Wiener (process) path. This equality follows from the fact
that w(s) has the distribution of the Lebesque measure. This (random variable)
integral is called the I to integral.
ITO'S LEMMA. Let f(x, t) be afunction on IRd ® [0, tJ and use Vf, V2 f, .. .for
x derivative and f for t derivatives. Then

f(w(t), t) - f(w(O), 0) = I dw· (V f)(w(s), s) +

+ -1 ft ds(V 2 f)(w(s), s) + ft dsj(w(s), s) (6.40)


2 0 0
for almost every w.
Proof was shown by Simon (1979). This expression is often written in an
infinitesimal form as the formal one
df = Vf ·dw + (t.<'1f + j)dt, (V 2 == .<'1). (6.41)
It is surprising that the second derivative of the function f appear in the first-
order differential. Equation (6.41) can be expressed even more succinctly by
dW i dWj = bij dt. Expression (6.40) can be used for calculational purposes:
EXAMPLE 6.6 (d = 1). 1's with tI" +j = 0 will have particularly simple
stochastic integrals. For example, if
f(x, t) = exp(o:x - to: 2 t),
i
then df = Vf ·dw. Recognizing the Wick-ordered exponential
f(w(s), s) = : exp(o:w(s)):
and taking w to b, we see that

I db: exp(o:b(s)): = 0:- 1 : exp o:b(t): _0:- 1 • (6.42)

If (6.42) is expanded in one finds that

I
0:

db: bn(s): = (n + 1)-1: bn+ l(t):.

If we take n = 1 we find that

I db b(s) = t: b 2 (t): = Hb 2 (t) - tJ


The Basic Concepts oj Random Processes and Stochastic Calculus 249

a result already obtained. Ito's lemma directly provides the connection of the
Brownian motion and -t~ = Ho and the Feynman-Kac Jormula. A key
observation is that because the differentials point into the future, we have
IE {j(b(s)) db} = 1E{j(b(s))}IE{db} = 0
since IE {db} = 0; i.e.,

lE{f dbJ(b(S))} 0, =

more strongly

lE{g f dbJ(b(S))} = 0 (6.43)

if 9 is a measurable function with respect to {b(u) I u <:( a}. Choose


J(x,s) = {exp[-(t-s)Ho],g}(x),
so j = HoJ: Thus j + t ~J = 0, so that Ito's lemma implies that

J(b(t), t) = J(b(O), 0) + J: db· (Vf) (b(s), s).

Using (6.43), we see that


IE {j(b(t), t)} = 1E{j(b(O),O)}.
Since J(b(t), t) = g(b(t)) and J(b(O), 0) = (exp( -tHo), g)(O), we have proved
once again that
lE{g(b(t))} = (exp( -tHo), g)(O).
The proof of the Feynman-Kac Jormula is similar but more complicated (see
Simon, 1979). Here, we take H = Ho + U and
J(x, s) = (exp[ - (t - s)H], g)(x).
Then, at least for the sufficiently well-defined U's (say U E C g'), we have

J(b(s), s) = J(b(t), t) - f du U(b(u))J(b(u), u) - f


db· (V f)(b(u), u).

Notice that according to (6.43) the last term has a zero expectation.
Reiterating this equation, we obtain

J(b(O), 0) = J(b(t), t) - J: du U(b(u))J(b(t), t) +

+ J: du U(b(u)) r ds U(b(s))J(b(s), s) + G,
250 Chapter 6

where G also has a zero expectation due to (6.43). Continuing in this way, one
finds that

If one recognizes the object in [ ... J as exp{ - S~ ds U(b(s))}, this will prove that

(exp( - tH), g)(~) = IE {exp[ - L ds U(b(s)) }(b(S))}. (6.44)

Now we turn to stochastic integrals. A stochastic integral is a random


function c(b, t) obeying

c(b, t) = c(b, 0) + L db· f(b, s) + L ds g(b, s), (6.45)

where f and g are nonanticipatory functionals (i.e., functions f(b, s) with the
property that f(', s) is only a function of {b(t) I t < s}, i.e., is £$s-measurable)
with suitable Lp properties and c(b, 0) is independent of b. One writes (6.45) in
the short form
dc = f· db + g ds. (6.46)
The following result is crucial:
THEOREM 6.16 (Ito's lemma). If C1 , .. , Cm are stochastic integrals and u is a
C 2 -function on IRm ® [0, CX)) with some mild restrictions imposed on growth at
infinity, then x = u(c 1 , ... , cm , t) is a stochastic integral, and
::lu 1
L L
m::lu m ::l2U
dx = _v dc· + ~ds + - _v_dc. dc. (6.47)
i=lOYi ot 2 i,j=1 0Yi OYj' J
where
dC i = f i · db + gi ds, i.e., db k db i = bki ds.
For the proof of this theorem, see McKean (1969).
Now we consider some consequences of this theorem.
EXAMPLE 6.7 (Product rule). If dC i = fi ·db + gids, then
d(C I C2 ) = c l dC 2 + C2 dC I + flf2 ds. (6.48)
EXAMPLE 6.8 (Feynman-Kac formula). Let us compute the differential of

c = exp { - f~ ds U(b(S))} f(b(t)),


where U,f E CO'. Since the function z = S~ ds U(b(s)) is a stochastic integral
with dz = U ds, we have according to Ito's lemma and (ds)2 = 0 that
The Basic Concepts of Random Processes and Stochastic Calculus 251

d(e- Z) = _e- Z U ds. Again by Ito's lemma,


df = Vf . db + t(Af)ds.
Thus, using (6.48), we obtain

dc = exp{ - IdS U(b(S))} [Vf·db - (Hf)ds],

where H = -tA + U. Thus, if we define

(QJ)(x) = lE{ex p [ - IdS U(x + b(S))J(X + b(S))},

and we use (6.43) that IE {db} = 0, we see that

(QJ)(x) = f(x) - I ds[Q.(Hf)] (x),

which yields Qt = exp( - tH) and so we have a new proof of the F eynman-K ac
formula.

EXAMPLE 6.9 (Feynman-Kae-Ito formula). Let us compute the differential


of

e = exp{ - i I db . a (b(s)) - ti I ds(div a)(b(S))} f(b(t)) = e -iz f(b(t))

with a,fEC o. Then, according to Ito's lemma, we have


d(e- iz ) = -i e- iz dz - t e- iZ (dz)2

= -ie-iz{a·db+t{V·a)ds} -te- iz a 2 ds.


Further, we use (6.48)
de = e- iz {( .. ·)db + [tAf - ia·Vf - t{V·a)f - ta 2f]ds}.
So taking into account
Ho(a)f = t{ -iV - a)2f
and writing

(QJ)(x) = lE{ex p [ -i I db·a + ti I dsdiva ]f(X + b(S))},


we find that

(QJ)(x) = f(x) - I ds [Qs(Ho(a)f)] (x) (6.49)

which yields the Feynman-Kac-Ito formula.


252 Chapter 6

EXAMPLE 6.10 (Drift). Let a E C~ and

x(t) = b(t) + 1 ds a(b(s))

which for obvious reasons is often called the drift generated by a. Then
dx = db + a ds so that
d[f(x(t))] = V f·db + (a· V f + 1Llf) ds.
Following our usual procedure
IE {J(x(t))} = exp( - tX), f) (0),
where X is the differential operator Xf = -tLlf - a· Vf which is thus called
the generator of the drift.
EXAMPLE 6.11 (Cameron-Martin's (1949) formula). Let dz = a·db -
-ta 2 ds, and e = eZ f(b(t)). Then
de = eZ{[Vf + af] ·db + [a·V f + 1LlfJ ds},
so in terms of the generator X in Example 6.10, we obtain
1E{e" f(b(t))} = (exp( -tX),f)(O).
This is usually summarized in the Cameron-Martin formula

Dx (Rado~Ni~Odym) = exp{It db.a(b(s)) _.1 It dSa 2 (b(S))} (6.50)


Db denvatIVe 0 2 0

If one writes
Ho(a) = -tLl + ia· V + ta 2 + ti diva,
then one sees that H o(a) is the generator of a (complex) drift - ia and a
potential ta2 + ti· diva. Then the combined Cameron-Martin formula for

ta
a· V and the Feynman-Kac one for 2 + i ·div a yields the Feynman-Kac-
Ito formula. There is a connection between the Cameron-Martin formula and
the Feynman-Kac formula (see Simon, 1979).
REMARK 6.1. The F. Riesz-A. Markov-S. Kakutani, and M. Stone-C.
Weierstrass theorems listed in Section 6.11 are the fundamental theorems in
topological spaces and measures. For these theorems, the reader is referred to
standard text books on given fields, for example, Dunford and Schwartz
(1958), and Alexandrov and Hopf (1935).
Chapter 7

Basic Ideas of Stochastic Quantization

7.1. Introduction

In recent years interest has significantly increased in the investigation of


stochastic processes and fields. This is due in the first place to the fact that it
has been possible to establish an intimate connection between the theory of
stochastic processes, quantum mechanics (Fenyes, 1952; Kershaw, 1964;
Nelson, 1966, 1967; de la Pella-Auerbach, 1969; de la Pella-Auerbach and
Cetto, 1974, 1975, 1982; Davidson, 1979; Moore, 1979; Yasue, 1979; Lee, 1980;
Santamato and Lavenda, 1981; Claverie and Diner, 1973, 1975, and others),
and Euclidean quantum field theory (Nelson, 1973; Dohrn and Guerra, 1978;
Guerra and Ruggiero, 1973), known under the general name of the stochastic
quantization of systems (stochastic mechanics) stochastic quantum mechanics.
Studies have been made on the generalization of the ideas of the stochastic
quantization of Nelson and Fenyes for continuous systems and fields (i.e., for
systems with an infinite number of degrees of freedom) (Guerra and Ruggiero,
1973; Dohrn and Guerra, 1978; Dankel, 1970, 1977; Moore, 1980 and
Davidson, 1980-82), and also for particles with spin (Dankel, 1970, 1977; de la
Pella-Auerbach, 1971) and relativistic mechanics (Caubet, 1976; Aron, 1966;
Lehr and Park, 1977; Guerra and Ruggiero, 1978; Vigier, 1979 and Namsrai,
1980a).
From the mathematical point of view the most interesting of the results
obtained is that the dynamics equations of stochastic mechanics are nonlinear
partial differential equations which admit linearization and that the obtained
linear equations are formally identical with the Schrodinger equation if the
diffusion coefficient v is taken as equal to h/2m.
This common character of the mathematical formalism of the two theories
suggests the existence of a deep connection between the theory of stochastic
processes and quantum mechanics. This problem is not solved completely and
requires further detailed investigation (in this connection see Ghirardi et al.,
1978; Onofri, 1979; Grabert et aI., 1979; Lavenda, 1980; Aron, 1981, and
Mielnik and Tengstrand, 1980). With regard to the problem of the connection
between the stochastic (Markov) processes and the Euclidean quantum field

253
254 Chapter 7

theory, Nelson (1971, 1973) has finally formulated the Euclidean quantum field
theory in terms of random processes, and we now have the solution to the
problem of the unique correspondence between the Euclidean and pseudo-
Euclidean Green's functions (Osterwalder and Schrader, 1973, 1975 and Glaser,
1974), the study of which was already begun by Schwinger (1959); Nakano
(1959), Fradkin (1954), Symanzik (1966), Taylor (1966), and Efimov (1963).
Besides this approach, other directions are also being developed in the study
of stochastic processes and fields (see Chapter 10 and reviews of Skagerstam,
1975; Moore, 1979; Boyer, 1975; Surdin, 1971, 1978; Blokhintsev, 1975; de la
Pefia-Auerbach and Cetto, 1982). Some of them take as their starting point the
hypothesis of the stochastic properties of the electromagnetic vacuum (Braffort
and Tzara, 1954; Bourrett, 1964; Braffort et al., 1965; Marshall, 1963, 1965;
Santos, 1974; Rueda, 1981, and Cavalleri, 1981). This approach is called
stochastic electrodynamics. The reviews (Boyer, 1975; Surdin, 1971, 1978;
Sandos, 1974; Claverie and Diner, 1976, 1978; Rueda, 1978, and de la Pefia-
Auerbach and Cetto, 1978, 1982) with most of the relevant references in the
field are included and present the basic elements and ideas of stochastic
electrodynamics and its connection with quantum theory. Stochastic elec-
trodynamics is constructed as classical electrodynamics with radiation damp-
ing of charged particles and their interaction with a background electromag-
netic (vacuum) field with spectral density
p(w) = (h/2n 2 c3 )w 3 •

Other directions are based (to some extent or after a fashion) on the concept
of stochastic space-time (see Chapter 1 and references therein). In this case, it is
assumed that the origin of the random behavior of particles (like Brownian
motion) is the stochastic nature of physical space-time. In other words, the
stochastic nature of the processes is regarded as the result of the effect (or
influence) of space-time itself on the physical system. Our approach presented
in this part of the book is concerned with this direction in the theory of
stochastic processes and the main attention is devoted to a generalization of
Nelson's stochastic mechanics to the relativistic case.
This brief list shows that the theory provides the most attractive formulation
of quantum mechanics and relativistic quantum field theory by means of a
stochastic context, which still offers interacting possibilities for development
and application to the inquiring nonorthodox mind.
Roughly speaking, these directions arose from introducing the de Broglie-
Bohm-Vigier hidden variables in quantum mechanics, related to an attempt to
give the physical interpretation of quantum phenomena at the subquantum level,
or, at a deeper level, to the very nature of the quantum system, which sometimes
seems to force the renunciation of such basic concepts as objectivity or
causality. We shall not dwell on these questions which are widely discussed in
the literature (see Bunge, 1956; Jammer, 1974; Ballentine, 1970; de la Pefia-
Auerbach and Cetto, 1977; and Vigier, 1982).
Basic Ideas of Stochastic Quantization 255

7.2. The Hypothesis of Space-Time Stochasticity as the Origin of


Stochasticity in Physics

Before turning to a description of the methods of stochastic mechanics and its


relation to quantum mechanics, we discuss briefly the idea of stochasticity in
quantum mechanics and its possible origin (or physical cause).
The idea that the electron moves stochastically is almost as old as quantum
theory itself (for historical discussion of this problem see Jammer, 1974).
Schrodinger (1931) and Furth (1933) first considered the formal analogy
between Schrodinger's equation and the diffusion equation. The proposition
remained a purely formal one, until Fenyes (1952), Kershaw (1964), and
Nelson (1966, 1967) took it seriously and proposed to interpret quantum
mechanics as a Markov process in configuration space, i.e., that the quantum
mechanical system is assumed to execute a sort of Brownian movement which is
fundamentally a quantum theory of the same structure conceptually as the
SchrOdinger theory of the electron. This is the basic view of the first school
mentioned above. As a necessary ingredient of quantum mechanics, stochas-
ticity has also been considered by de Broglie (1967).
Accepting that quantum mechanics describes the behavior of matter only
statistically, the essence of the so-called statistical interpretation of quantum
mechanics (see, for example, Jammer, 1974; de la Pefia-Auerbach and Cetto,
1977; Ballentine, 1970; Rosen, 1974, and Ross-Bonney, 1975), strongly de-
fended by Einstein, Podolsky and Rosen (1935), consists in the assumption
that the description refers to an ensemble of similarly prepared systems, not to
an individual member of it, and that each one of these systems behaves
stochastically, thus contributing to unavoidable dispersions in all the dynami-
cal variables that are not uniquely fixed by the preparation process.
It deals with the meaning of statistics as well as the question of possible
limits on scientific knowledge. As is known, Einstein, de Broglie, and their
present followers never accepted Bohr's assumption that quantum probabi-
lities represented an ultimate limit to human knowledge. In this sense, they have
always held that probability distributions should result from the real action of
randomly correlated subquantum causal motions - a conception which implies
the introduction of subquantal hidden variables to interpret the probabilities
correctly described by field (and field equations) of quantum mechanics. This
view is the basis of Vigier's statement: 'New stochastic variables which imply a
superimposed random set of motions (reminiscent of Feynman's paths) result-
ing from the particle's interactions with a real 'vacuum'; a vacuum now
considered as a deeper (of course, covariant) subquantum distribution of
violent motions and thus alive with fluctuations and randomness .. .' (Vigier,
1982b, p. 924).
By leaving aside the philosophical aspect of the problem, we assume that if
the stochastic motion of particles is indeed responsible for the statistical nature
of quantum mechanics, then it should be possible to employ this fact as a
256 Chapter 7

cornerstone for the construction of a more explicit - and perhaps more


fundamental - theory of quantum syslems. The above-mentioned directions
along these lines have been developed with different degrees of success. Among
them, stochastic mechanics preserves the idea of random trajectories for the
configurational degrees of freedom of the mechanical system, but interprets
these random trajectories as effective trajectories of well-defined stochastic
processes, satisfying stochastic differential equations well-known in statistical
mechanics and control theory. The evolution of the random processes is ruled
by a stochastic version of the Newton second principle of dynamics proposed
by Nelson (1966, 1967) (see Section 7.4 and Chapter 8). The link with quantum
mechanics is provided by the proof that the distribution of the relevant
physical observables at each time is identical in stochastic mechanics and
quantum mechanics, as a consequence of the respective time evolution
equations.
Obviously, the next step in this direction is to search for the physical cause
(or the origin) of stochasticity. Thus the question: what are 'carriers' (or what is
the physical cause) of stochasticity, is of great importance in the stochastic
interpretation of quantum mechanics. Without answering this question the
theory would remain for ever a phenomenological and unsatisfactory one, at
least philosophically, as is the theory with the hidden variables.
We would like to emphasize from the beginning that the stochastic theory is
not a hidden variables theory and should not be considered as such. Moreover,
one would like to have to hand a more fundamental stochastic theory of
matter, from which quantum mechanics would emerge under certain circum-
stances and approximations. If such a theory could be constructed, it would
not only provide us with a deeper understanding of matter, but also broaden
the perspectives of the quantum theory itself.
As listed above, various independent efforts in this direction have contri-
buted to the groundwork of a theory of this kind: stochastic mechanics,
stochastic electrodynamics, the stochastic theory based on the hypothesis of
space-time stochasticity.
In stochastic electrodynamics the basic question concerning the origin of
stochasticity is answered by assigning a real character to the random zero-point
radiation field or vacuum field, which in usual quantum electrodynamics is
treated as a virtual- i.e., formal- field. This idea is already discussed in several
early papers, notably those by Welton (1948) and Weisskopf (1949) (for more
details, see de la Pefia-Auerbach and Cetto, 1982 and references therein).
With regard to the first school, i.e., stochastic mechanics, unfortunately it
cannot explain what causes the stochastic nature of the microscopic world. In
this approach either the problem was not discussed or no clear advantage of it
was obtained. There are differing points of view. For example, Bohm (1952)
interpreted the deviation from the Newtonian equations of motion as due to a
quantum-mechanical potential associated with the wave function. Fenyes
(1952) showed that instead of assuming a quantum-mechanical potential the
Basic Ideas of Stochastic Quantization 257

motion could be understood in terms of a Markov process. Weizel (1953) also


proposed a model for the random aspects of the motion in terms of the
interaction with hypothetical particles, which he calls zerons. Nelson's (1966,
1967) aim has been to show how close it is to the classical theories of
Brownian motion and Newtonian mechanics, and how the Schrodinger equa-
tion might have been discovered from this point of view. He examined the
hypothesis that particles in empty space, say aether, are subject to Brownian
motion. De la Pefia-Auerbach and Cetto (1975) dealt with particles owing
their stochastic behavior to the action of a stochastic surrounding medium
acting in addition to the external (applied) forces.
An interpretation of the Schrodinger equation in terms of particle trajec-
tories was first proposed by de Broglie. Bohm and Vigier (1954) introduced the
concept of random fluctuations arising from the interaction with a sub-
quantum medium. Recently, this idea has developed at a deeper level and
obtained a new formulation by Vigier and his team (see Vigier, 1982a, b;
Cufaro Petroni and Vigier, 1983b).
Another possibility concerning the third of the above-listed approaches
remains. The idea to use the stochastic space-time or random fluctuations of the
metric gl'v as the origin of stochasticity (i.e., the real quantum forces) is a most
attractive and interesting conjecture considered first by Einstein (1924) (see
Chapter 1 and references therein). We propose that the origin of stochasticity
in the microworld is caused by the effects due to space-time, i.e., gravitation
(exactly, 'gravitational vacuum') on material systems (particles). There exists a
standard technique for considering these effects when spinor fields may be
presented (Weinberg, 1972). In particular, these effects give rise to the cos-
mological term A9l'v in the Einstein equation:
(7.1)
as a representation of the energy momentum tensor distribution of the random
background gravitational aether or noise, i.e., the 'gravitational vacuum'. Here
K = 8nG/c 4 , G is the Newtonian constant.
We discuss the appearance of the cosmological term in (7.1). It is proved
that gravitation of homogenuous substances situated outside some sphere and
extending to infinity does not contribute to the force inside this sphere, and
thereby its action tended to zero. Thus attraction to the center of the sphere of
Galaxies situated on it is only caused by substances of the sphere and
acceleration of these Galaxies with respect to the sphere center is
ag = -GM/R 2 . Let us give the mass of the sphere through the mean density of
matter p: M = (4/3)nR 3 p. Then
ag = -(4/3)nGRp. (7.2)
Thus, even if at some moment Galaxies are stationary, at the next moment
they will begin to move with acceleration towards each other. As a result, the
Universe will compress and become unstable.
258 Chapter 7

In order to balance the gravitational forces and to make the Universe


stationary it is necessary to introduce a repulsive force independent of sub-
stance. Gravitational and repulsive accelerations must be equal and opposite
to each other with respect to modulus and sign: ag = - ar •
On the basis of such consideration Einstein introduced the cosmological
repulsive force which makes the Universe stationary. This force is universal: it
is not dependent on the masses of the bodies involved, but depends only on the
distance between them. Taking into account the fact that the mean value of
the matter density in the Universe is p ~ 2 x 10 - 29 g cm - 3 one can calculate
the numerical quantity of the repulsive acceleration which IS
ar ~ 5 x 10- 37 R cm sec-Z, or

where c = 3 X 10 10 cm sec- 1 .
Now we show that according to our idea of the space-time stochasticity
presented in the concrete form (1.1) in Chapter 1, a test particle acquires an
additional energy caused by the gravitational vacuum. As shown above
(Chapter 1), our proposal about the stochasticity of space-time has led to the
change of the space-time metric form (1.4) and (1.12) in the cases of flat and
curve space-time, respectively. On the basis of these formulas in Chapter 11 we
will obtain the following expression
(7.3)
for the rest energy of the test particle [see formula (11.54)J, in the units of
c = h = 1. Thus we see that the Einstein formula E = mc 2 is changed here.
Assuming 1---+ Igr = G1 / 2 [the case I P 19r formally corresponds to some strong
gravitation near particles which are regarded as 'small black holes' (see Salam
and Strathdee, 1976, and Li, 1982)], in the second term of (7.3) we get the
gravitational interaction energy of the particles:
Egr = tam 3 G = tam 2 G/A,
where A = h/mc is the Compton wavelength of the test particle with mass m and
a as some constant. In order to obtain the energy density we divide Egr by the
one-particle volume '" A3:

(7.4)

This formula coincides with the vacuum energy density proposed by the Soviet
physicist Zel'dovich. His argument for the occurence of non-zero vacuum
energy density may be presented in the following way: Let virtual particles
with mass m be created and annihilated in a vacuum. According to the
quantum theory, the mass of the particle defines the characteristic distance
A = h/mc. Virtual particles are generated from each other at this mean
characteristic distance. Then, by a usual formula the gravitational interaction
energy of surrounding particles is given by Egr = Gm 2 / A. This energy raises the
Basic Ideas of Stochastic Quantization 259

non-zero vacuum energy density or mass density: Pvac = evaclc 2 • Hence


evac = (Gm 2pc)jJ.3 = Gm 6c411- 4, (J. = II/mc).
We see that this formula differs from (7.4) by the constant a/2.
Moreover, it transpires that the theory requires that a 'vacuum fluid' should
possess an unusual negative pressure, according to our consideration above.
Therefore, it is better to speak about tension not pressure. Thus we assume
P = -e or, using Einstein's relation P = _pc 2 , we get P/c 2 = -po This last
equality is very important. Namely, it leads, as shown below, to the gravitation
repulsion of the vacuum.
We recall that the mass of a uniform ball with the radius R i1> equal to
M = (4/3)nR 3 p. But, in the Einstein theory it holds only in the case when the
pressure of matter is very small with respect to the value pc 2 • If it is amiss, then
M = (4/3)nR 3 (p + 3P/c 2 ). Usually p is much larger than P/c 2 and the term
3P/c 2 may be omitted. However, in the case of the gravitational interaction of
virtual particles, the pressure P and energy density are related: p = P/c 2 • The
'vacuum fluid' fills the whole space naturally. Therefore, for calculation of the
gravitation acceleration caused by the vacuum, one can use formula (7.2)
obtained for the usual matter, but p should be replaced by the sum
(p + 3P/c 2 ). Taking into account that Pvac = - P/c 2 , we get
a vac = -4nG(Pvac + 3P/c 2 )R = -4nG(Pvac - 3PvaJR = 4nG· 2p vac R. (7.5)
Thus, for the gravitational vacuum we do not obtain attraction as in the case
of usual matter [minus sign in (7.2)J, but repulsion [plus sign in (7.5)]. This
repulsion is caused by negative vacuum pressure which should be taken into
account in calculations of the gravitational force, according to the Einstein
theory.
With regard to the numerical quantity of the gravitational repulsion of the
vacuum, the theory is too crude to yield any precise value.
It should be noted that properties of space-time and physical vacuum
including the gravitational vacuum, and fluctuations and instabilities in them
have recently acquired the most important significance in contemporary
physics. Thus, for example, a vacuum is properly determined by evolutionary,
not merely energetic, considerations. This point is relevant for all proposed
models (especially in the Yang-Mills gauge theory) of elementary particle
interactions. The concept of the metastable and false vacuum, and gravitational
vacuum (Turner and Wilczek, 1982; Guth, 1981, 1982; Linde, 1982; Kirzhnits
and Linde, 1972), properties of vacuum as a physical medium (Lee, 1981), and
fluctuations in the Yang-Mills vacuum (Migdal, 1981) assist in understanding
very important physical problems:
phase transitions of grand unified theories;
very early evolution of our Universe described by the iriflationary model
which can solve the monopole, horizon, and flatness problems of the
standard hot Big-Bang cosmology;
quark confinement, and so on.
260 Chapter 7

It is important to remark that interrelation between particles and space-time


or exactly a profound connection between the particle-wave dualism and the
structure of space-time has developed further (see Fujiwara, 1980 and Efinger,
1981). According to Efinger, the idea is that the particles in question are
geometrical objects on a Riemannian space-time, i.e., one can take the view that
freely moving particles are represented by solitary waves which, by definition,
preserve their shape. Then one could hypothesize that associated with these
waves is a Riemannian metric gllV which is nearly singular (or has some
difference from the Minkowski metric) in a region corresponding to the width
of the solitary wave-amplitude (or a region of confinement of these particles -
an idea that recalls an old concept by Einstein, 1967; for detail see Chapter 11).
Outside this region one expects gJlV ~ I] JlV = diag(l, -1, -1, -1), where I] JlV is
the Minkowski metric. Then by our concept of space-time stochasticity, we
associate a nonvanishing random curvature R = gJlV RJlv = Ro + Rs (or
gJlv=I]Jlv+bgIlJ with every point in space-time, such that Rs--->O (for
<b9IlV)R ---> 0) outside the region of confinement of these particles. For ex-
ample, in Chapter 11 we will use the metric form gfV(x) = ¢2(X, bE)I]/Lv,
(bE = iT, b) in the cmiformally flat space-time. After averaging the space-time
[R4(X) (Chapter 1) this metric form will be

<g~V(x) )R4(x) = ¢2UZ /X2)l]llv'

in particular, a special form

gives an interesting result which will be presented in Chapter 11.


We close this section with some hope that the hypothesis of space-time
stochasticity will help us eventually to obtain a deeper and more intuitive
physical picture of the quantum system and stochastic processes.

7.3. Stochastic Space and Random Walk

Let us consider random walk from the point of view of stochastic space and
give some primitive reply to the question: Bow to understand that the particle
subjects Brownian-type motion is caused by the stochasticity of space? We
assume that the stochastic particle of interest is regarded as a classical
punctiform mass, occupying at every instant a single point in space and
traveling, therefore, along a trajectory. The probabilistic element, which is
essential to provide a link with quantum mechanics, is introduced by assuming
that this trajectory is continually influenced by a 'vacuum fluid', i.e., the
gravitational vacuum force induced around the particle. As a result, the
particle subjects random motion. While the exact properties of the gravi-
tational vacuum interaction are unknown, the fluctuations of the particle
Basic Ideas of Stochastic Quantization 261

position resulting from this interaction are presumed to be describable as a


Markov process.
Thus we consider the motion of the particle in the space [R3(X) (Chapter 1).
We assume that the particle executes N displacements; then its position Ro
after N displacements is given by the expression
N

Ro = L rJ.
j = 1

If the motion is uniform and rectilinear, then r7


= vOt j • We shall now assume
that in each displacement r7
the particle makes a random displacement
because of the stochastic nature of space: b j = IXjb, where IXj is an arbitrary
sequence of real numbers, and b is a universal stochastic variable, subject to
the probability distribution (1.6) (Chapter 1). Then after N displacements in
the stochastic space [R3(X), the position of the particle is determined by
N

<I> = Ro + B = L (r7 + IXjb). (7.6)


j=l

What is the probability WN(<I» d 3 <l> that the particle's coordinates after N
displacements lie in the interval between <I> and <I> + d<l> where <I> is determined
by equation (7.6)? In such a formulation, the problem can be solved by the
method proposed by A. A. Markov. The application of Markov's method to
the random-walk problem can be found in Chandrasekhar's (1943) paper.
The probability WN(B) dB that after N displacements the particle is in the
interval (B, B + dB)

(7.7)

is given by

WN(B) = ~
8n
fd 3p exp( -ipB) AN(p), (7.8)

where

AN(P) = ))1
N f d 3 bw(b) exp(iplXjb) = ))1 (2nI2)-3/2
N f (
d 3b exp iprLjb - Ib
12
2z2 ) .

To estimate AN(P), we calculate the value of the typical integral

I f
= (2nf2) - 3/2 d 3b exp(iplXj b - Ibl 2/2z2) = exp( - z21X] p2 /2)

and we then have

(7.9)
262 Chapter 7

where
N

P N = [2 x La;.
j =1
Substituting (7.9) in (7.8), we obtain
WN(B) = (2npN)-3 12 exp( -tIBI 2 jPN). (7.10)
We now determine v, which is called the diffusion coefficient. This vanishes,
v = 0, when I = 0, i.e., when the space is not stochastic. Setting

and denoting by v the quantity


lIN
lim N 2: L (~B~)2,
N---+oo j=l

where ~B~ = IJn aj , in which n is the number of displacements per unit time,
we obtain 2vt = Thus, B;.
liN .
V = - lim -
2 N _ ro N
L (~B;Y (7.11)
j =1

After the determination of v, expression (7.10) can be written in the form

p(B) = lim WN(B) = (4nvt)-3 12 exp( -IBI2j4vt). (7.12)


N-ro

Since B = eI> - R o,
p(eI» = (4nvt)-3 12 exp( -lei> - RoI2/4vt). (7.13)

As one would expect, our problem has been reduced to a random-walk


problem. In accordance with (7.12), a particle which makes n displacements
per unit time with each of these displacements bj = ajb subject to the
probability distribution w(lbI 2) is situated after time t in the element of volume
between Band B + dB with probability p(B) d 3B or p(eI» d 3<1> (when Ro =I- 0).
It is known from the theory of Brownian motion that the distribution
p(r, t; r o, Do), which determines the probability of finding the position r of the
°
Brownian particle at the time t if r = ro and = 00 at t = 0, is determined by
an expression just like (7.13)
p(r, t; r o, Do) = (4nv bt) - 312 exp( -Ir - ro 2 /.4v.bt),
1

for t ~ f3 -1, where Do is the initial velocity of the Brownian particle,


f3 = 6nayt/m, Vb = kT/mf3 = kT/6nayt, a is the radius of the Brownian particle, yt
Basic Ideas of Stochastic Quantization 263

is the coefficient of viscosity of the surrounding medium (the fluid), k is


Boltzmann's constant, and T is the absolute temperature. Thus, the in-
troduction of the stochastic nature of space leads to a Brownian type of
motion of a particle which in the absence of the stochastic properties of space
would move along a definite preassigned trajectory.
We now turn to the derivation of a differential equation for the probability
density p(r, t). The derivation of the equation is the same as in the random-
walk theory. Therefore, we shall here give only the results. As usual, we make
important assumptions:
(1) The time interval At is chosen sufficiently large for the particle to make a
large number of displacements in this interval, but small enough that during
<
this interval the mean-square increment of r, i.e., IM12) remains small; under
these conditions the probability that the particle is displaced by M in the time
At IS given by
P(~r, ~t) = (4nv ~t)-3/2 exp( -IM12j4v ~t), (7.14)
and does not depend on r. The function P(x, t; x', t'), where ~r = x' - x,
At = t' - t, t' ~ t, is called the transition probability (see Chapter 6). The
physical meaning of this quantity is that P d 3 x' represents the probability that
the particle performing Brownian motion will reach the volume d 3X ' around x'
at time t' if it started from the point x at the initial time t.
(2) It is assumed that the displacement of the particle in space at a given
time does not depend on the previous displacements; then

p(r, t + ~t) = fd3(~r)p(r - ~r, t)P(M, ~t). (7.15)

<
An equation of type (7.15) is called a Smoluchowski equation. Since l~rI2)
is small by hypothesis, we can expand p(r - ~r, t) in the integrand in a Taylor
series and integrate the obtained expression term by term. Going to the limit
At -> 0, we obtain
(7.16)
On the basis of these results obtained, we would like to construct stochastic
mechanics in the following chapter by means of Smoluchowski's equations of
type (7.15) for nonrelativistic and relativistic cases.

7.4. The Main Prescriptions of Stochastic Quantization

In this section we briefly present the main ideas of stochastic quantization of


dynamical systems and give a simple introduction of the probabilistic terms
used in our consideration (for more detail, see Chapter 6). We also refer the
reader to works of Nelson (1979) and Guerra (1981) for a deep and extensive
264 Chapter 7

review of the interplay between quantum mechanics and the theory of


stochastic processes. At the same time, an important step in this trend was
made by those who gave a precise mathematical definition of stochastic
mechanics and field quantization.
Thus, let us consider a system with n independent degrees of freedom
described by the classical Lagrangian
. 1 n. n •
L(Q, Q) =:2. L Qf +L AlQ, t)Qi - U(Q, t), (7.17)
1=1 1=1

where Qi are the canonical co-ordinates of the system. The Euler-Lagrange


equations of motion are given by
n
Qi = L (8Aj/8Qi - 8AJ8Q)Qj - 8AJ8t - 8U/8Qi, (i = 1,2, ... , n)
j = 1

and are invariant under the gauge transformation:


Ai(Q, t) --+ A;(Q, t) - 8f(Q, t)/8Qi' U(Q, t) --+ U(Q, t) - 8f(Q, t)/8t.
Here Ai and U are electromagnetic and potential fields, and f(·) is an arbitrary
scalar function. The canonical momenta of our system are
Pi = 8L/8Qi = Qi + Ai' i = 1,2, ... ,n.
The procedure of stochastic quantization is based on the (first) assumption
that each of the classical configuration variables Q(t) = {Q;}, i = 1,2, ... ,n is
interpreted as a Markov random process and satisfies a stochastic differential
equation of the form

dQ;(t) = bi(Q(t), t) dt + d W;(t), (7.18)

where bi(x, t), x E ~n, t E ~, i = 1,2, ... ,n is a velocity field to be determined


and the W's are Wiener process which satisfy
lE{dW;(t)} = 0, (7.19)
where v is the diffusion coefficient (as usual it is assumed that v = II/2m,
another possibility will be discussed below). The solutions to equation (7.18)
are called Ito's processes or the generalized Brownian motion. Here we assume
that d Wi(t) do not depend on Q(s) for s -< t.
If the b's satisfy a global Lipschitz condition (Nelson, 1967, p. 43) then a
unique solution to (7.18) is ensured in the finite-dimensional case. The
restriction to finite-dimensionality is not of major concern since any field
theory can be made finite-dimensional by imposing a momentum cutoff and
working in a finite volume (see Chapter 10). However, the global Lipschitz
condition presents more serious difficulties since it is not satisfied when the
quantum wave function has zero as it always does in excited stationary states.
This problem has been discussed by Albeverio and H0egh-Krohn (1974).
Basic Ideas of Stochastic Quantization 265

These complications can be avoided by working only with wave functions


which do not have exact zeros.
Further, it is convenient to introduce according to Nelson, the mean forward
derivative D + and the mean backward derivative D _ through the definitions

D+f(Q, t) = lim h- 1 1E{j(Q(t + h), t + h) - f(Q(t), t)IQ(t) = Q},


h----'1' 0+

D_f(Q,t)= lim h-11E{j(Q(t),t)-f(Q(t-h),t-h)IQ(t)=Q}, (7.20)


h~ 0+

where IE denotes the conditional expectation with respect to the random


variables Q;(t) (see Chapter 6).
It is easily seen from the equality
d
dt IE {j (Q(t), t) g(Q(t), t)}

= 1E{[D+f(Q(t), t)]g(Q(t), t) + f(Q(t), t)[D_g(Q(t), t)]},


and from definitions that the following equality holds for the expectations:

f:oo 1E{j(Q(t), t)[D +g(Q(t), t)]}dt = - f:oo IE{[D - f(Q(t), t)]g(Q(t), t)}dt
if f and g are sufficiently regular and at least one of them has compact support
with respect to the time variable t.
If we introduce the probability density p(x, t), x E IR n , t E IR = T, of the
random variables Q;(t), so that

1E{j(Q(t), t)} = f d"x f(x, t)p(x, t),

then we can immediately prove the following relations

with

b;- (x, t) - b+
; (x, t) = - 2v -aI n p(x, t)
ax;
and, more generally,

D + f(Q(t), t) =
a + .Ln
( ;;- a + v~ ) f(Q(t), t),
bt ~
ut ,= 1 uQ,

D - f(Q(t), t) = ( ata + ;~l


n a
b;- aQi -
)
v~ f(Q(t), t),
266 Chapter 7

where

~= ±(~).
;= 1 aQ;
The probability density p(x, t) satisfies the Fokker-Planck equations
ap a +
L -a (pb;) + v ~p,
n
-a = - (7.21)
t ; = 1 X;
or
ap
(7.22)
at
If we define

we have the equation of continuity


ap a
L
n
- = - -(pvJ = -div(pv) (7.23)
at ; = 1 ax;
obtained by averaging (7.21) and (7.22). We call v the current velocity. If we
make the definition of so-called stochastic velocity u; = ¥bt - b;-), we have

U; =
a
v-a In p = v(grad); ·In p.
X;

Therefore

au = vgrad ~ In p = v grad [aaP ~] = v grad[_-_d_i_V_P_V]


at at t p P

= - vgra{ div v + v gr:d p] = - v grad div v - grad(vu).

That is

-au = -v grad d'IV v - gra d(vu).


at
Now we define as mean acceleration
(7.24)
Then the second Nelson assumption is
ma;(Q(t), t) = F; (7.25)
by the Newton law; where F; is some external force, in particular,
Basic Ideas of Stochastic Quantization 267

Et)uations (7.23) and (7.25) lead to a system of equations which can be


transformed into Schrodinger's equation

a
in at ljJ(x, t) = (n2
- 2m V 2 +U )ljJ(x, t),
where
ljJ = exp(R + is), (7.26)
+
b; =2v;;-(R+S),
a (7.27)
uQ;
It was established by Davidson (1979) that Nelson's dynamical assumption
[(7.24) and (7.25)] is not the only possibility. The following equation [instead
of (7.25)]
t(D+ D_ + D_ D+)Q;(t) + tfJ(D+ - D_)2Q;(t) = -oU/oQ;(t)
also leads to Schrodinger's equation where
ljJ = exp(R + izS),
z = (1 - tfJ)-1 /2, (7.28)
Here fJ is a constant and bt and b;- are still given by equation (7.27) but with
the value of v in equation (7.28). If v in equation (7.28) is real, which is
necessary for a physical interpretation, then fJ must be real and satisfy
condition fJ < 2.
By choosing fJ appropriately, any value of the diffusion coefficient can be
used in a stochastic model of quantum mechanics. It is possible to choose v
arbitrarily small by letting fJ be large in magnitude but negative. In the limit of
zero v the model becomes deterministic and in fact turns to the familiar hidden
variable model of Bohm (1952), a point first made Shucker (1980a).
In the case of (7.17) with the formulas previously given, we can easily prove
that

a;(x,t)=-+
OV;
Ln a a
Vj_V;_2V2_(p-1/2~'p1/2).
at j = 1 OX j OX;
Nelson's second basic assumption is the validity of the Euler-Lagrange
equations of motion with the substitutions (L -> a;, Q; -> V;. Therefore we have
~ v·-v·
L..
a + L.. ~ (OA j OA;) oA; au
- - - v·----+
j = 1 J Ox j ' j = lOX; aX j J at ax;

+ 2 . v2 ~
a ( P -1/2 Ll
A.
P112), .-
I -
1, 2, .•• , n.
x;
The process is completely described by the probability density p(x, t) and the
268 Chapter 7

current velocity vJx, t), which satisfy a system of nonlinear differential equations
for the first order with respect to time [continuity equation (7.23) and the Euler-
Lagrange equations]. Therefore, if p and v are known at some initial time to, we
can solve the Cauchy Problem and know the Markov process Q(t) completely.
With the additional kinematical assumption that the generalized momentum
Vi + Ai is a gradient, therefore Vi + Ai = as/ax i for some function Sex, t), we
can show the complete equivalence of our scheme of quantization with the
usual one. In fact, if we introduce the wave function
t/!(x, t) = [p(x, t)] 1/2 exp[ih -I Sex, t)],
then we can check that when p and S envolve according to the stochastic
differential equations then t/! evolves according to the Schrodinger equation
. at/!
Ih -
1 In (h -a - )2 t/! + Ut/!,
~ Aj
at aX j
=; -
2j =1 I

associated with the same classical system. On the other hand, if we know the
wave function t/!(x, t), then we can recover the stochastic process.

7.5. Stochastic Field Theory and its Connection with Euclidean Field Theory

First of all consider the simple example of the harmonic oscillator with
Lagrangian
L(Q, Q) = !(F - !w 2 Q2.
The normalized ground state wave function is given by
t/!(x, t) = (w/hn)I/4 exp( - wx 2/2h) exp( - ihwt/2),
x E ~. Now we consider the stochastic process Q(t) associated with the ground
state. The density is
p(x, t) = It/!(x, tW = (w/hn)I/2 exp( -wx 2/h).
The current velocity vanishes because the phase function S = ihwt/2 does not
depend on x.
On the other hand, for the mean forward derivative D + Q(t) = b + and the
mean backward derivative D_Q(t) = b- we have
b+(x, t) = -b-(x, t) = -wx.
Therefore the basic stochastic differential equation of the process corresponding
to the harmonic oscillator is
dQ(t) = - wQ(t) dt + d w.
Taking into account the Gaussian form of the density we find that the process
Basic Ideas of Stochastic Quantization 269

Q(t) is a Gaussian process characterized by the expectations


IE{Q(t)} =0, IE{ Q(t)Q(t')} = (h/2w) exp( - wi t - t'I).
This Gaussian-Markov process is known as the Ornstein-Uhlenbeck process
(see Chapter 6).
We can consider now the Klein-Gordon equation for a real classical field
cp(x, t), x E 1R 3 , t E IR = T

(:t22 - L1 + m2) cp(x, t) = 0,

where m is a mass parameter with the dimensions L -1. The Hamiltonian of the
field is

H = ~ f[ (~~ y + (Vcp)2 + m 2cp2 ] d 3 x.


In order to use the stochastic quantization it is convenient to enclose, as
usual, the field in a cubic box of the length L, with periodic conditions at the
boundary. In this way we reduce the system to a family of the independent
harmonic oscillators, with canonical coordinates {cpdt)} and Hamiltonian

H = t L [4J;(t) + (k 2 + m2)cp;(t)],
k

where k = {ku kz, k 3 }, k i = 2nnjL, n i are relative integers. For each oscillator
of the frequency w 2 = k Z + m2 , we consider the ground state and associated
stochastic process explained before. Then going to the limit L -> 00, we find
immediately that the ground state stochastic process associated with the
Klein-Gordon equation is a Gaussian generalized process cp(x, t), with mean
zero and a covariance (for detail, see Chapters 6 and 10)

IE {cp(x, t)cp(x', t'n

(7.29)

where d 4 p is the Lebesque measure in 1R4 and p2 = p~ + pZ is the squared


Euclidean norm in 1R4.
Therefore we can recognize that the covariance is nothing but the two-point
Schwinger function of a free field with mass m (in a system of units in which
h = 1), i.e., the two-point Wightman function continued to a purely imaginary
time (see the following section).
The important fact, discovered by Guerra and Ruggiero (1973), is that for
the process we have constructed, the parameter t is the real physical time. In
conclusion, we would like to say that if the Euclidean field identified with the
270 Chapter 7

ground state stochastic process is associated with the classical field theory,
then the underlying four-dimensional manifolds (x, t) on which the field is
defined can be considered as the Minkowski physical space-time, and the
powerful method of the Euclidean field theory can also be applied to the study
of the stochastic field theory.

7.6. Euclidean Quantum Field Theory

Because of the fact that there is a profound connection between the Euclidean
quantum field theory and the theory of stochastic processes, we give here some
elementary definitions of the Euclidean QFT. The Euclidean QFT is based on
the idea of replacing the usual description of quantum field in the Minkowski
space-time with a description in some auxiliary Euclidean space. References in
this direction are presented in Sections 1.1 and 7.1.
First of all, we consider a free scalar quantum field fP(x) of mass m,
completely defined through its vacuum expectation values:
WZn = W(x1,,,·,X Zn ) = <0IfP(x1)"'fP(xzn)10)
Zn
= L
Wl(Xl,X)W(Xz,,,·,Xj-l,Xj+l,,,,,XZn), (x=,Xo,x) (7.30)
j; Z

with the expectation values of an odd number of fields vanishing. The two-
point function is given by (see Section 1.4.4)

W(x,y) = W(x - y) = (2n)-3 f d3k


2w(k) exp[ -iw(k)(xo - Yo) +

+ ik(x - y)], (7.31)


where w(k) = (kZ + mZ)l/Z >- m. If we perform an analytic continuation on x o,
Yo to the Schwinger points Xo = -ix 4, Yo = -iY4' X 4 >- Y4, then we obtain the
Schwinger function

+ ik(x - y)], (7.32)


(x E = X4 , x). The elementary formula

(7.33)
Basic Ideas of Stochastic Quantization 271

allows one to write (7.32) in the form

S(x,y) = f(~~Y exp[ik(x - y)](kZ + mZ)-l. (7.34)

where x, k are points with coordinates (X4' x), (k4' k), X, y E 1R 4 , kx = k 4x 4 + kx,
k Z = k~ + k Z, d 4k = dk 4 d 3 k, (dk)4 = d 4k. For convenience the index E is
omitted here.
According to the symmetry and the positive definiteness of (7.34) Euclidean
fields are introduced as Gaussian random fields with mean zero and covariance
lE{cp(x)cp(y)} = S(x,y), X,YE 1R4. (7.35)
Therefore, we see that the Euclidean theory can be described through stochas-
tic fields as opposed to the non-commutative Heisenberg fields of the
Minkowski theory.
As an example, we now consider the Schwinger function (7.34) in two-
dimensional space-time:

Sz{y,m6) = (2n)-Z fdZkeikY(kZ + m6)-1. (7.36)

An explicit form of this function is


Sz(x) = (2n)-1 Ko(molxl) = i(4)-1 HI}) (imo lxI),
where Ko(x) is the Mac'Donald function connected with the Hankel one by
the formula
Ko(x) = !niHi})(ix).
From formul::.>" (7.36) and (7.33) it follows, in particular, that

Thus, the Schwinger function Sz and the Wightman function W z coincide at


time t = s = O. Since the function Sz is real-analytic, then according to equality
(7.36) it co-ordinates with the analytical continuation of the function W z in the
extended future (cone) tube. For the function Sz all the Euclidean regions are
exhausted by this. Having an analytic continuation for the function W z , it is
easy to obtain it and the functions Wn given by formula (7.30).
THEOREM 7.1. Let [xy] = Sz(x - y, m6), where Sz is given by formula (7.36).
Then the functions (for notations see Section 1.4.4 in Chapter 1):
(7.37)
are the Schwinger functions for the free field with mass mo. The proof is given by
Simon (1974).
272 Chapter 7

Notice that the Euclidean field must commute at different points of space.
Indeed, from the locality it follows that [cp(X,O), cp(y, O)J _ = 0, because the
Euclidean field at an imaginary time must coincide with the Minkowski field
at the zero time. However, any two points of the Euclidean region can be
reduced to a zero (imaginary) time by means of the Euclidean (rotational)
motion.
Nelson (1971, 1973) pointed out the importance of the Markov property (see
Chapter 6) of the free Euclidean field and also of the interpretation of the
Euclidean field as the functional integrals (see Simon, 1974, 1979).
We notice that there exists a deeper connection between the field formulated
by Garding and Wightman, and the Euclidean QFT. Thus, one consequence of
the Wightman axioms is that Wn is the boundary value of the analytic function
W n(Zl" .. , zn) in a certain region G. G includes all the points of the form
Zj = (x j , is) with (x j , Sj)E jRd and with Zk =!= Z/ (all k, I). The functions

Sn(Y1, ... , Yn) = W n(Y1' is 1;· .. ; Yn' is n)


are called the Schwinger functions.
Rather than the construction of the Wightman field <1> (x), one constructs a
measure dfl on a space OU'(jRd) (OU is the Schwartz space of tempered functions)
whose moments are the candidates for the Schwinger functions; i.e., one
assumes {J ddy <1>(y)f(y)}(T) = T(f) as random variables and

Sn(Y1,· .. ,Yn) = f dfl <1>(Y1)"'<1>(Yn), (7.38)

that is

f ddy J1(Y1)'" f
ddyn fn(Yn)SiY1"" ,Yn) = f
TUJ'" T(fn) d{l(T).

Thus one needs some conditions on the measure d{l or on Sn which allow us to
reconstruct <1>. The earliest such conditions are due to Nelson (1973), but they
are difficult to verify in practice. Osterwalder-Schrader (1973, 1975) gave a set
of conditions on the Sn's whose modulo growth conditions with n are
equivalent to the Garding-Wightman axioms and, in particular, isolated a
condition now known as OS (Osterwalder-Schrader) positivity.
THEOREM 7.2 (Frohlich's (l 174) reconstruction theorem). Let d{l be a cylinder
measure on OU'(jRd) which satisfies the following conditions.
(i) Proper Euclidean motions [i.e., T(x) f-+ T(Ax + b), bE jRd, AESO(d)]
leave dfl invariant;
(ii) OS positivity, i.e., given a real-valued f E OU( jRd) with Supp f E{ x, s), S > O},
let (8f)(x, s) = f(x, - s). Then for real-valued f1" .. ,fn with the above
support and Z 1 .•• , zn E IC:
Basic Ideas of Stochastic Quantization 273

J
(iii) For any f E OZt(lR d), dJl exp[<I>(f)] < 00.
(iv) The action of the translations (x, s) ~ (x, s + t) is ergodic. Then, there is
a unique (scalar) field theory obeying the Garding- Wightman axioms
whose Schwinger functions are given by (7.38).

The most natural way to construct measures is to try Gaussian dJl's. It


transpires that these measures describe 'trivial' field theories in that they
describe particles without interactions - they are the analog of the harmonic
oscillator.
THEOREM 7.3. A Gaussian measure dp on 0Zt '(lRd) obeys conditions (i) (iv) of
J
Theorem 7.2 if and only if the covariance S2 (x, y) = dJl<l>(x)<I>(y) is of the form

S2(X, y) = (2n) -d f ddk eik(x - y) S2(k)

with

S2(k) = f dp(m2)(k2 + m 2)-1 + p(k2), (7.39)

where dp is a polynomially bounded positive measure on [0, (0), so that the


integral lies in OZt', and P is a polynomial which is positive on [0, (0).
Proof Sketch of proof (Simon, 1979) is the following: the condition (i) is
equivalent to S2(X, y) = f(x - y) withf rotation invariant, (iii) is automatic for
the Gaussian process, and (iv) can be seen to be equivalent to f(x) ~ as °
x ~ 00. That leaves the analysis of (ii). This is clearly equivalent to
n
I ZkZj exp{ -tSifk - ef j,!;, - efj)} ~ 0.
j, k =1
It can be shown that this, in turn, is equivalent to S2(ef, f) ~ 0, for allf with
the proper support property. This is an analog to the usual positive definiteness
condition that leads to Bochner's theorem. The unitary group of Bochner's
theorem (see Reed and Simon, 1975) is an unitary group in the x variables and

°
a self-adjoint semigroup in the s-variable. The temperedness of S2 leads to the
boundedness of the semigroup. Thus for s > and dp a tempered measure:

SAx, s; 0,0) = 2(2n)1 - d I dp(E, k) e ikx e- sE .


JE ;;'0
Since S2 is rotationally invariant, [Xi (8/8s) - s(8/xJ]S2 = 0. This translates
into Lorentz invariance of dJl. Since dp is supported in the region E ~ 0, it
follows (see, for example, Paley et al., 1933) that

dp(E, k) = Cob(E, k) + fdP(m2)[E-1 b(E - (k 2 + m 2 )1/2) dd-1k] dE.


274 Chapter 7

Using formula (7.33) we find that for s > 0

S2(X, s; 0, 0) = 2C o(2n)1 - d + (2n)-d f ddk e {fdp(m 2)(k 2 + m2)-1 }.


ikx

Since S2 -+ 0 as s -+ 00, Co = O. Rotation invariance yields the applicability of


the formula for all (x, s) #- O. All that remains is an ambiguity of a positive-
definite rotation invariant distribution supported at zero. This gives the
polynomial P(k 2 ) for the Fourier transform.
As usual, it is assumed that P = 0, dp(m2) = c5(m 2 - m~) d(m 2 ) in expression
(7.39). The resulting Gaussian process is called the free Euclidean field of mass
mo'
With this we finish a very brief consideration of Euclidean field. The detailed
study of this problem may be found in the monographs by Simon (1974),
Glimm and Jaffe (1981).
Chapter 8

Stochastic Mechanics

8.1. Introduction

As indicated by recent studies (Nelson, 1979; Guerra, 1981) stochastic me-


chanics (its name is probably due to Guerra) is a very simple and clear theory,
with enormous possibilities of physical and mathematical effectiveness in the
exploration of properties of quantum mechanical systems, especially when a
large number of degrees of freedom are involved, as the experience of the
Euclidean methods in constructive quantum field theory has shown. The main
advantage of stochastic mechanics, from the point of view of mathematical
control, relies on the possibility of exploiting the well-developed mathematical
methods of probability theory and stochastic processes (see Chapter 6).
However stochastic mechanics is not free from criticism. We would like to
stress a peculiar aspect of stochastic mechanics which is the basis of typical
misunderstanding and unjustified criticism. It is remarked that a very natural
objection arises in connection with any proposal of exploiting stochastic
processes for the formulation of a theoretical scheme related to quantum
mechanics. As usual, it is suggested that quantum mechanics may be in-
terpreted as a Markov process, but irreducible to a Brownian-type stochastic
motion. Or, more clearly it can be phrased as follows: 'Stochastic processes are
typical expressions of diffusion phenomena and therefore share a substantial
time irreversible character, while quantum mechanics is time reversible and
therefor~ cannot have anything to do with stochastic processes' (see Guerra,
1981). This distinction must be clearly borne in mind, since it is the failure to
take it seriously which leads to most of the criticisms (Kracklauer, 1974;
Ghirardi et aI., 1978; Grabert et aI., 1979; Mielnik and Tengstrand, 1980)
usually made against the stochastic interpretation. However, as has been
explained by Guerra (1981), the scheme of stochastic mechanics is time
reversible as it should be.
On the other hand, it should be noted that in stochastic mechanics there are
several open conceptual problems. Two of them are related to canonical
transformations and to the Fermi statistics (see Guerra, 1981). With regard to
the problem of the general covariance properties of stochastic mechanics, we

275
276 Chapter 8

believe that it may be successfully investigated in our approach based on the


hypothesis of space-time stochasticity.
Thus the attraction of the approach based on the hypothesis of stochastic
space-time to the description of stochastic processes is that we have succeeded
in generalizing stochastic mechanics to the relativistic case and defining
rigorously, in the mathematical sense, relativistic integrals of the Feynman-
type (Namsrai, 1980c). In addition, in this scheme, as one would expect, we
encounter the self-turbulent phenomenon (Namsrai, 1980b) which is character-
istic of a nonlinear system.
By assumption about the form of the stochastic space-time ~4(X) presented
in Chapter 1, one c.an construct an equation of the Smoluchowski type for the
probability density p(xl" s) in the relativistic case; for p(xl" s) there is a
representation of the form

p(XI" S + ~s) = f
d 4 bE a(bi, ~s)p(xo + ib4, X - b, s), (8.1)

where s is some invariant parameter which is, as usual, interpreted as a proper


time. The integration in (8.1) is over a four-dimensional Euclidean space
and a(bi, ~s) is some integrable function of the variable b~ =
b~ + bi + b~ + b~. Relativistic invariance is ensured by the fact
that the algebra of the Lorentz group and the algebra of the group of four-
dimensional Euclidean rotations are identical in the complex domain (see
Section 1.4.6 in Chapter 1).
There is a hope that if we begin the construction of a theory in the
stochastic Euclidean space e(:x, ' = b4 ), a relativistic invariant description of
the motion of particles in stochastic spacetime can be realized.
Before starting the construction of dynamics of stochastic particles, we give
here the physical meaning of the stochastic nature of space-time in the
nonrelativistic and relativistic cases. In the framework of our approach in the
nonrelativistic case, the stochastic nature of space leads to the impossibility of
determining the coordinates of a particle with an accuracy exceeding at least
the value of the particle's Compton wavelength. In the relativistic case, it is
harder to give this property a physical interpretation. Formally, it can be
interpreted as the presence of fluctuations of the four-dimensional co-ordinates
of the particle in the Euclidean space 1E4(X, ' = b4 ) (i.e., as the presence of a
random walk in the imaginary time). By itself, such an interpretation has no
physical meaning, but the method has value as one method of relativistic
description of random processes in terms of an equation of the Smoluchowski
type. It should be noted that equations (8.20) and (8.21) shown below do not in
general bear any relation to the ordinary Smoluchowski equations which
describe a stochastic consequence, and therefore the quantities P ± which occur
in these equations cannot be interpreted as transition probabilities (see
Chapter 6).
Stochastic Mechanics 277

By a universal (or fundamental) length, we shall understand the following.


Physically, the universal length l is some characteristic distance over which the
corresponding notions about spacetime and locality (causal connection) begin
to break down; in particular, stochastic properties and nonlocality can be
manifested if they exist. Recent experimental data show 1;51O- 15 -1O- 16 cm
(see Part 1). Other possibilities for introducing the concept of a fundamental
length in physics will be discussed in Chapter 11 (see also Chapter 1 and
references therein).
Our exposition is as follows. In Sections 8.2 and 8.3 we study the nonre-
lativistic motion of a particle in a force field and the motion of a relativistic
particle in a four-dimensional stochastic space-time, and obtain equations of
motion for a particle that are formally equivalent to the Klein-Gordon equation.
The two-body problem, Section 8.4, is studied in the nonrelativistic and
relati vistic cases.

8.2. Equations of Motion of a Nonrelativistic Particle

In constructing the dynamics of stochastic particles, it is customary to use the


mathematical concepts of left and right derivatives (or stochastic, or Ito's
derivatives) (see Chapter 6), by means of which the stochastic and current (or
systematic) velocities of the particle are constructed. Newton's law is used as
the dynamical equation (see the preceding chapter). We recall here that the
stochastic derivatives of a function f(x) are given by Ito's (1961) rule:
D+ f(x) = (a/at + v+'V + vV 2 )f(x)
and
D_ f(x) = (a/at + v_'V - VV2)f(X), (x = t, x)
where v + and v _ are called forward and backward velocities, respectively. In
this case the Newton law takes the form: ma(t) = F. Here a(t) = t(D + D _ +
D _ D + )Q(t) is called the mean second derivative or mean acceleration.
Q(t) is a stochastic process representing the stochastic motion of a particle
with mass m in an external force F.
However, Kershaw's (1964) approach (see also Lehr and Park, 1977) was
based on equations of the Smoluchowski type for the probability density
p(x, t) and the mean particle velocity v(x, t).
Despite the difference interpretation of the stochastic behavior of the system
(the particle), the mathematical method of description of the stochastic pro-
cesses in our scheme will be the same as in the studies of Kershaw, and Lehr
and Park which are based on the theory of Bohm (1952) and de Broglie (1964).
In Section 7.3 we considered the random-walk problem from the point of
view of a stochastic space. The greatest interest attaches to the study of the
278 Chapter 8

motion of a particle in a space whose properties are assumed to be stochastic.


We now turn to this question.
If it is assumed that the displacement of the particle in space at a given time
does not depend on the previous displacements, then the following relation
holds for the probability density p(x, t):

p(x, t + At) = fd 3(6X+)p(x - 6X+, t)P +(x - 6X+, t; 6X+, At), (8.2)

where P +(x - 6X +, t: 6x + , At) can be interpreted as the probability that a


particle at the position x - 6x + at the time t is displaced through 6x + during
the time At and, therefore, it reaches the point x at the time t + At. In the
stochastic theory (Nelson, 1966, 1967; Lehr and Park, 1977; Namsrai, 1980a)
with twice the number of transition probabilities, one uses P _(x + 6X_, t;
6X_, At), which is the probability that a particle has moved from the position
x + 6x_ by 6x_ in the time interval At prior to the time t and thus occupied
the position at the point x at the earlier time t - At. Thus, the equation
analogous to (8.2) takes in this case the form

p(x, t - At) = fd 3(6L)p(X + 6L, t)P _(x + 6L, t; 6L, At). (8.3)

For P ±' we can here choose an expression of the form

P+ =(4nv+.At)-3/2 eltP {-[6X+ -v+(X,t).At]2},


- - 4v±·At

where 6X± = v± (x, t)· At + Ax± is the total displacement of the particle in the
time At and v+ are certain constants of the type of diffusion coefficient. Setting
v + = v _ = v, expanding p and P ± in Taylor series, integrating over 6X±, and
retaining terms of order At, we obtain Fokker-Planck equations for p(x, t)
op/ot = -V·(pL) - v·V 2 .p, (8.4)
or
op/ot = -V(pv); u = vV In p;
(8.5)
v = t(v + +L ); U = t(v + - L ).

In stochastic theory, v and u are called the ordinary (current), and stochastic
velocity of the particle, respectively.
We now consider the motion of a particle in an external force field
F = -v U. Following Kershaw (1964), we can derive equations of the
Smoluchowski type for the mean velocities v + and v _ of the particles in
accordance with the formulas
Stochastic Mechanics 279

x p(x + t5x±, t)P ± (x + t5x±, t; t5x±, M) d 3(t5x±); (8.6)

v+(x,
-
t + M) = ~f[V+
N+ -
(x ± t5x_,
+ m f~(x
t) + At - ± t5x_,
+
t)] x

where f ± and f~ are certain external forces, and

N± = fp(X + t5x±,t)P±(x + t5x±,t;t5x±,M)d (t5x±) 3

and normalization factors. The upper and lower signs correspond to v + and
v _, respectively.
Expanding v±' p, P ±' f ±' and f~ in the Taylor series, integrating over t5x±,
and going to the limit M ---+ 0, we obtain from (8.6) and (8.7) four possible
equations:

(8.8)

Going over to the variables, v and u and adding and subtracting equations
(8.8), we obtain the following equations, which describe entirely different
processes:
dev - Itdsu = (1/m)F t; (8.9)
and
(8.10)
where
de = a/at + (v, V); ds = (u'V) + VV2; It = ± 1;
Fi = t(f+ + L); Ft-I) = t(f'+ + L); Fl = t(f'+ - L);
Fi~"I) = t(f+ - L); FI = t(f+ + L); F (_ I) = t(f'+ + L);
F'I = t(f'+ - f_); F'(-I) = t(f+ - L).
The left-hand sides of the obtained equations have a definite parity under
the time-reversal transformation. Indeed, under t ---+ - t

u ---+ u; ds ---+ ds
280 Chapter 8

and we readily conclude that the expression de" - Adsu does not change, while
deu + A d s" (A = ± 1) changes signs under the transformation t ~ - t.
Therefore, the right-hand side of the corresponding equations, which is the
force, must be chosen such that the individual equation as a whole remains
invariant under t ~ - t. This requirement is satisfied if we assume that
f'+ ~ f'_ as t ~ - t.
Then F; does not change, while F;:- reverses its sign, and F 1 ~ F( _ 1) and
F'l~ -F(-l) as t~ - t and, therefore, the four equations (8.10) actually
reduce to two equations. We also write the equation
au/at = -V(uv) - vV·(V·v), (8.11)
which follows from the continuity equation and expression (8.5) for the
velocity u.
We see that, on the basis of the hypothesis of a stochastic space and
Smoluchowski's equations, we have obtained in the framework of Kershaw's
approach the same fundamental equations that were obtained by Nelson
(1966-1967) and de la Pena-Auerbach and Cetto (1975) (see also the review of
Skagerstam, 1975 and the lecture of Santos, 1972) by different routes. Note
that Kershaw could not obtain these equations, since he considered only one
transition probability, namely P +(~x, At).
Detailed investigations in equations (8.9)-(8.11) are made in stochastic
mechanics. In Section 9.3 we consider a question associated with the require-
ments imposed on the equations of stochastic mechanics. In particular, on the
basis of some mathematical assumptions about stochastic processes, Nelson
(1966) obtained the first equation in (8.9) with A = 1 and (8.11), and showed
that if one considers a charged particle and takes the forces
F 7 = eE + !-v x Hand F 7 = - V U,
then these equations are equivalent to the Schrodinger equations

at/!
- = --(
i .
-lhV - e
-A)
2
t/! - .e
1- Ut/!
at 2mh e h

and

. -at/!
Ih = ( - -h2 V2 + U) t/!
at 2m

respectively. Here v = h/2m, t/! = exp(R + is),

R = tIn p, grad S = ~ (" + ;c A ),


Stochastic Mechanics 281

and A, the vector potential, is related to E and H by


1 aA
H = curl A, E = ---;;7ft -Vcp.
Further, on the basis of the description of the mean local motion of the
elements of continuous media (ensembles), which leads to two different kinds of
motion during a short time interval M, de la Pella-Auerbach and Cetto (1975)
obtained equations (8.9) with .Ie = 1. These equations can also be reduced to
Schrodinger equations for the wave functions t/I and t/I* if the Lorentz force is
chosen as follows:
+ e aA e e e
F1 = - eV cp - - ~ + - v x Hand F 1 = -u x H + - vV x H.
c at c c c
They also discussed the case .Ie = - 1.
Skagerstam (1976) investigated equations (8.9) and (8.11) from the point of
view of the theory of stochastic processes, defined them in a classical con-
figuration space. Recently, Davidson (1979) showed that in the Fenyes-Nelson
stochastic model there is an entire class of different dynamical schemes which
lead to a Schrodinger equation as the solution of a Markov diffusion problem.
As a result, the diffusion coefficient v in the Markov theory is an arbitrary
positive parameter (see Sections 7.4 and 10.2).
The remaining equations (8.10) have the same form as the Navier-Stokes
equations for 'fluids' with velocities v and u if v is formally identified with the
coefficient of viscosity, they are studied in Chapter 9.
Finally, we determine

lim ~ (L1x) 2 .
2 L1t
t.t~o

By definition,

where L1Xj = ZJn


(Xj' and n is the number of displacements per unit time (see
Section 7.3). We also have

(L1X+)2 = 2v+·M = (L1x_)2 = 2L·M


and hence
1 (L1X)2
lim - - -
M --> 0 2 L1t

We have arrived at Nelson's relation.


Thus, from the hypothesis of the stochastic nature of space we have
282 Chapter 8

obtained stochastic mechanics, identical with quantum mechanics, at least in


the mathematical formulation. The connection between the Schrodinger equa-
tion and the stochastic theory based on the hypothesis of a stochastic space
can be demonstrated by the following simple example. Suppose that at t = 0
the wave function has the form
(8.12)
where N is normalization constant, and a is a positive number; then for t> 0
the wave function is determined by means of the Green function of the
Schrodinger equation:

cp(x, t) = f d 3X' K(x - x', t)cp(x', 0),

where
K(x, t) = i- 1 (4nvt)-3 /2 exp(ix 2 /4vt), v = h/2m.

r
After simple calculations, we obtain

Icp(x, tW = N 2[ 1+ ~::: T3/2 ex p{ - :: (1 + ~::: 1}- (8.13)

We now consider the probability density corresponding to function (8.12), i.e.,


(8.14)
and we attempt to determine p(x, t) by means of the stochastic mechanics
considered above. Since u = v V In p the velocity corresponding to expression
(8.14) is UO(x, t = 0) = -(h/ma 2 )x and the solution of an equation of type (8.4)
for.uo = v +, (v _ = 0) is determined by the relation
[x - x' - uO(X')t]2 _ X'2}
p(x, t) = N 2 (4nvt)-3 /2 f, d x' exp {-----"-,------
3
4vt a 2

_ 2(
- N 1+
~)-3/2
2 4 exp
{_ x
2
~)-1} '
1+ 2 4
2( (8.15)
ma a ma

which is identical to the quantum-mechanical quantity (8.13). The kernel


G(x, t; x', t' = 0) = (4nvt) -3/2 exp{ - [x - x' - UO (x')t] 2 /4vt}
of equation (8.15) satisfies the Fokker-Planck equation
oG oxG 02
at = Pax + vox 2 G,

where
Stochastic Mechanics 283

8.3. Relativistic Dynamics of Stochastic Particles

As was point out in the Introduction (Section 8.1), in the relativistic case we
shall formally consider the motion of a particle which executes a random walk
due to the stochastic nature of the four-dimensional Euclidean space 1E4(X, c).
Suppose the particle executes N displacements; then its coordinates are
determined by the expression BIl = 'LJ = 1 f3 jb", where f3j is a sequence of N
numbers, and the vector bll =(b4 , b) is distributed with the probability
w(b ll ) d 4 b = w(bllbl') d 4 b.
Here b 2 = bllb ll = b~ + b 2 •
The probability WN(B Il ) d 4 B that BIl lies in the interval between BIl , BIl + dBfl
is given by

4
WN(Bfl) d B = d B(2nPN)
4 -2 (_B2)
exp 2PN ' P N=
N
L [2f3J.
j =1
Defining

2vs = B~ = B; = B; = B~
or
lIN
v=-lim _[2 L nf3J, (n=N/s)
2 N ~ cc N _j= 1

we obtain

where s is some positive invariant parameter which can be understood in the


present case as the proper time of the particle. Further, as in the three-
dimensional case, we define the 'transition probability'
P(LlYE' Lls) = (4nv· Lls) - 2 exp( - 1LlYE 12 /4v . Lls)
and then the 'Smoluchowski equation' takes the form

PE(X!, s + Lls) = f
d 4YE PE(X! - Y!, s)P(YE, Lls), (8.16)

where for convenience we have set YE = LlYE, and the diffusion equation
becomes
(8.17)
If P(YE' Lls) depends on the point x!, for example,
P(x!; YE' Lls) = (4nv· Lls) - 2 exp{ - eYE - Y~(XE)] 2/4v· Lls},
284 Chapter 8

where (y~)1l = u~' As and u~ = (u 4 , u) is the four-dimensional Euclidean ve-


locity of the particle, then instead of (8.17) we obtain an equation of the
Fokker-Planck type in Euclidean space:
all = (%x 4 , V).

The question of the transition to pseudo-Euclidean space (Minkowski space)


now arises. As shown above (Chapters 1 and 2), the idea of space-time
stochasticity and the construction of the non local theory of quantized (exten-
ded) fields can be realized by a displacement of the co-ordinates such that Xo
acquires a purely imaginary addition (xo ~ Xo + iT), while the co-ordinates x
remain real. It is found that the procedure for displacing the co-ordinate Xo is
deeply connected to the transition between Euclidean and pseudo-Euclidean
quantum field theories, and evidently it also has a direct bearing on the
relativistically invariant description of extended objects (see Chapters 1, 2 and
Efimov, 1977a).
Using this idea, we can write equation (8.16) in the form

where the variables x il = (xo, x) are pseudo-Euclidean and P ± can be chosen in


the form
P ± = (4nv± ·As)-2 exp{ -(YE - y~)2/4v± 'As},

Y~ = (±iu~ 'As, u± . As),

where ui are four-dimensional velocity vectors. From this we readily obtain


the two Fokker-Planck equations
(8.18)
or, in terms of v ll and u ll ,
u ll = t(u~ - u~) = - v all In p, (8.19)
where

We now attempt to obtain equations of the motion of a particle which in


accordance with the correspondence principle must take the form (8.8) depend-
ing on the choice of the force field in the nonrelativistic limit. In the framework
of our scheme, relativistic equations for ui are obtained from the integral
equations

ui(x v , s + BAs) = ;± f[ ull±(x - BY, Xo + iY4' s) +


Stochastic Mechanics 285

~s
+ 8-F~(x - 8y, Xo + iY4) ] x
m -

x p(x - 8y, Xo + iy 4, s)P ± (x - 8y, Xo

+ iY4,s;YE,~s)d4YE' (8.20)

~s
-8-F'f(X+8Y,X o +iY4) ] x
m -

x p(x + By, Xo + iY4' s)P ±(X + 8y, Xo


(8.21 )

where

8 = {I
-1
for u~
for uP:.

and F~ and F/~ are certain forces.


Expanding the expressions in (8.20) and (8.21) in the Taylor series, in-
tegration over d 4 YE, going to the limit ~s -+ 0, and making some calculations,
we obtain the equations

(8.22)

From this we obtain relativistic equations for vil and u il :

(8.23)

and

(8.24)
286 Chapter 8

where
Dc = a/as + v" a.. ; Ds = u" a.. + vD, A. = ± 1.
The functions <I»+)Il, ... ,<I>1' can be expressed in terms of F~, ... ,F'~ in the
same way as in the nonrelativistic case. <I»+)Il[<I>~-)"] does not (does) change
sign, and
<I>'t=><I>i-1) and <I>'r=> -<1>;':.1) under s-+ -so
Equations (8.23) and (8.24) in conjunction with equation (8.19) are the
covariant analog of (8.9), (8.10), and (8.5) in the relativistic case.
Note that the left-hand side of the first equation in (8.23) for A. = 1 is
identical to the expression for the acceleration obtained on the basis of some
assumptions in the framework of the mathematical approach of Nelson (see
Lehr and Park, 1977; Guerra and Ruggiero, 1978).
We can consider the special case when only the Lorentz force <I>~~r is
present; this is related to the electromagnetic field FIlVD by

<I>\+)1l = ~ FIlV vv = ~(a" A" - a" AIl)V...


c c
Here All is the electromagnetic potential, for which the Lorentz condition is
satisfied:
allAIl = O.
In this case, it is usually assumed that the generalized momentum is a four-
gradient of a world scalar S:

and then the first equation in (8.23) with A. = 1 and (8.19) are equivalent to the
Klein-Gordon equation

(8.25)

The proof is given in Nelson's (1966, 1967) and Leht and Park's (1977) papers,
and therefore we shall not give it here.
The equations of (8.23) with A. = 1 were also investigated by Vigier (1979) in
the framework of the approach of de la Pefia-Auerbach and Cetto (1975). The
external field in this case was chosen in the form

Such a choice of the field <I>\-)Il ensures consistency of the second equation in
Stochastic Mechanics 287

(8.23) with (8.19) for A = 1, as a result of which there is no overdetermination


of the physical quantities p and vI' (see Section 9.3).
The covariant analog of (8.10) is provided by equations (8.24), which will be
considered in Chapter 9.

8.4. The Two-Body Problem in Stochastic Theory

8.4.1. The N onrelativistic Case


We consider first two interacting nonrelativistic particles. We shall investigate
the problem in the framework of a stochastic theory based on Smoluchowski
equations for the probability density p (Xl' X2 , t) of finding the first particle at
the point Xl and the second at the point X2 at the time t with relative velocities
V l (Xl,X 2 ,t) and V 2 (X l ,X 2 ,t). It is assumed that the interaction potential

U(X l ,X 2 ) between two particles will depend only on the difference IXI - xzl
between their co-ordinates, i.e., U(xl,X Z ) = U(lx l - xzl).
Description of the two-body problem in stochastic mechanics is not only a
problem of real physical interest, but may playa very important role in the
causal action-at-a-distance interpretation of the Aspect-Rapisarda experi-
ments (Aspect et al., 1982a, b) on singlet photon pairs (see Cufaro Petroni and
Vigier, 1979, 1983a).
The problem we consider was first studied by Kershaw (1964) for the
transition probability P + (Ax, At), and he obtained equations that describe the
motion of the center of mass of the two particles and their relative motion. As
above, we shall investigate the problem of two bodies in the framework of two
transition probabilities P + and P _. Here, for the first particle
P~(Axr, At) = (2n'lAt/md- 3/ Z exp[ -m l (Axf)z/2'lAt],

and for the second


P;(AXI' At) = (2n,z At/m z )-3/2 exp[ -m Z (AxI)2/2'zAt],
where '1 = 2v l m l and '2 = 2v 2 m Z '
Without loss of generality, we take, 1 = 'Z = ,. In terms of the variables
r+ =xt -xi, R+ = (mlxt +m 2 xi)/(m l +m z ),
P~ and P~ take the forms P +(Ar+, At) and P +(AR +,At), where

,;, (2n, At//l)-3/2 exp[ -/l(Ar+)z/2, At].


Here o(xi)/o(r+) is the Jacobian of the transition from xi to r+, and
288 Chapter 8

)1 = m 1 m 2 /(m 1 + m 2 )· Similarly,

P+(i1R+,M) = ~(X~)) fd3(i1X)P~(i1X'M)P~(M


u(R m
i1R+ _ m 1 i1X,M)
m2
2

= (2n:r M/M)-3/2 exp { -M(i1R +f/2r M},


where M = m 1 + m2 is the total mass of the particles. Suppose
C+(r,R,t)=vt -vi,

here

R = [m 1 (xt + x~) + m 2 (xi + x 2 )]/M = (m 1 x 1 + m 2 x 2 )/M.


Then the total variations bR + and br + are determined by the equations
bR + = V+ i1t + i1R + and br+ = C+ i1t + i1r+
respectively. Analogously with the earlier investigations,
P;-(bR+,i1t;r,R,t) = (2n:r M/M)-3/2 exp{ -M(bR+ - V+ i1t)2/2r M},
and
Pr+(br+, i1t;r,R,t) = (2n:r i1t/)1)-3/2 exp{ -)1(br+ - C+ i1t)2/2r M}
and at the same time

satisfies the equation

P(r, R, t + i1t) = f f d 3 ((jr +) d 3 ((jR + )p(r - br +, R - (jR +, t) x

X Pr+(br+,M;r-(jr+,R-bR+,t) X

X P;- (bR +, M; r - br +, R - (jR +, t)

= p - M{Vr(PC+) + VR(pV+)} + M"2r{V2~P + V2}


~ ,

whence
(8.26)
where
VM = r/2M.
The corresponding equations for P obtained on the basis of the concept of
Stochastic Mechanics 289

the transition probabilities P~ and P~ take the form

p(r,R,t - At) = ffd3(br-) d 3 (()R -)p(r + br-, R + bR -, t) x


x Pr-(br-,At;r + br-,R + bR-,t) x
x P;-(bR-,At;r+br-,R+bR-,t)
or
t8.27)
where

From equations (8.26) and (8.27), we obtain


op/ot = -Vr(pVr ) - VR(pVJ;

(8.28)
Here

We recall that the potential U(r) acts only on the velocity C, then the
equations for C± and V± have the form

V±(r,R,t ± At) = ~± ffd3(br±)d3(bR±)V±(r =F br±,R =F bR±,t) x

x p(r + br±, R + bR±, tW; Pi-


where

N± = ffd3(br±)d3(bR±)p(r=Fbr±,R =FbR±,t)P;Pif.

As usual, expanding V±,C±,P;,Pjl', p and U in Taylor series, integrating and


290 Chapter 8

retaining terms only of order in L'lt, we obtain


M[dV± jdt + (V± VR)V± + (C± Vr)V±]

= ±M{VI'P-1[V;(PV±) - V±V;p] + VMP-l[V~(pV±)-

- V± V~P]};
Il[ac± jat + (V± VR)C± + (C± Vr)C±]
= -VrU(r) ± ll{vI'P- 1 x [V;(pC±) - C±V;p] +
+ VMP-l[V~(pC±) - C±V~p]}.

Going over to the variables Ve, Ue, Vr and U r we obtain the equations
(8.29)
where

Thus, the operators d~ and d~ decompose into a sum of two independent


parts. Accordingly, we can seek p(r,R, t) in the form of the product
p(r,R, t) = Pr(r, t)PR(R, t),
and the variables Ve,r and ue,r can be set equal to
Vr(r,R, t) = Vr(r, t); ur(r, R, t) = ur(r, t);

Ve(r,R,t) = Ve(R,t); ue(r,R,t) = ue(R,t),


where PR' Ve and ue describe the motion of the center of mass (as the free
motion of a particle with mass m 1 + mz ) and Pr' Vr, and U r describe the
relative motion of the particle [as the motion of a particle of mass 11 in the
centrally symmetric field U = U(r)]; then equations (8.28) and (8.29) take the
form
(8.30)
and
(8.31)
where
d~ = ajat + (Ve V R), d~ = (ue VR) + VM V'~,

d~ = ajat + (V r V r ), d~' = (u r V'r) + vI'V;,


Stochastic Mechanics 291

Following Nelson, equations (8.30) and (8.31) can be linearized by the


substitutions

We then obtain the following two equations for CPr and CPR:

(8.32)

and

(8.33)

respectively. The last two equations are formally identical with Schrodinger
equations for CPr and CPR if we set r = 11, i.e.,

VM = II/2M,

8.4.2. The Relativistic Case

We now turn to the problem of two relativistic particles. In this section, we


consider the stochastic behavior of two identical and correlated relativistic
scalar particles, since this problem has real physical interest. Our investigation
is based on equations of the Smoluchowski type for the probability
p(xr,xi,sbs2) for finding the first particle at the point x~ and the second
at the point xi at the 'time' Sl and S2 and for their relative velocities vi
(xi, x~, Sl, S2) and vi(xi, xz, Sl, S2), respectively.
For the direct generalization of the results obtained above for the single-
particle model to the relativistic two-particle case, it was found to be ma-
thematically more convenient to introduce the eight-dimensional configuration
space considered by Cufaro Petroni and Vigier (1979).
In this space, the positions of two particles and their relative velocities are
determined by the eight-component vectors Xi and vi(i = 1, ... ,8), respec-
tively, where

{ Xi}._
l-1, ... ,8
={XIlXV}
l'
_
2 ~.v-O.l,2.3'
.

Here, xi and Xz are the co-ordinate vectors of each particle. The metric tensor
292 Chapter 8

gij in this space is defined as

0 0 0 0 0 0 0
0 -1 0 0 0 0 0 0
0 0 -1 0 0 0 0 0
0 0 0 -1 0 0 0 0
gij =
0 0 0 0 1 0 0 0
0 0 0 0 0 -1 0 0
0 0 0 0 0 0 -1 0
0 0 0 0 0 0 0 -1

and
X z = XiX i = gijX iXl
' = (Xl) Z + (X )Z.
z
If X't(SI) and x~(sz)are the trajectors of each particle, then in the eight-
dimensional space their common trajectory will be Xj(SI'SZ)'
As in the three- and four-dimensional cases, we introduce the two-particle
analog P(X j, S), Sz, Lls I , Lls z ) of the singI(~-particle transition probabilities
P(x, t, M) and P(xll., S, Lls) used above. If the particles are not correlated,
P(xj, SI> Sz, Llsl> Lls z ) can be factorized:
P(xj,s),sz,Lls),Lls z ) = PI(x't,s),LlsdPz(x~,sz,Llsz)·
Without loss of generality, we choose the gauge Lls i = Lls z = Lls; then the
equation of Smoluchowski type for p(X\ Sl' sz) takes the form

p(xj, Sl ± Lls, Sz ± Lls) = fd8YEP(XI =F yl, Xl + iY~, X 5 + iYi, Sl' sz) X

x P±(XI =FyI, Xl + iY~, X 5 +


(8.34)
where Xl and yl (1 = 2,3,4,6,7,8) are the spatial parts of the vectors Xi and
yi, respectively.
Taking into account the explicit form of P±,
P± = (4nv± ·LlS)-4 exp{ - (Y~ - y~)z/4v± Lls},
y~ = (± iv~ Lls, ± iv~ Lls, v~ Lls), (8.35)
we obtain from (8.34) the differential equations
8p/8s 1 + 8p/8s z + 8Jpv i+) - V + Op = 0,
8p/8s) + 8p/8s z + 8i(pV i_) + LOP = 0,
8i = 8/8X i, -8 i 8i = 0 = 0) + Oz' (8.36)
Stochastic Mechanics 293

Here, we have set v _ = v + = v, where v is the diffusion coefficient, and vi+ and
vi_ are the forward and backward velocities. Going over to the variables

and adding and subtracting equations (8.36), we obtain


(8.37)
where Vi is the ordinary (regular or current) and ui the stochastic velocity of the
two-particle system.
In our scheme, conservation of mass (of the probability density multiplied
by the volume) means that there is no loss of mass through any hypersurfaces
characterized by the vectors vi and v~; in this connection, we adopt the
physical hypothesis that the total number of particles (i.e., the pair in the real
space-time) is conserved. Then we can write
a/as 1 + ap/as 2 = (dp'''l) + (dp'''2)
= (ap/ax~)v~ + (1 _/Ji)-1/2("1 ·ap/ax 1) +
+ (ap/ax~)v~ + (1 - /3~)-1/2("2·ap/aX2) = 0,

(/3; = v; /c 2 , i = 1,2) and our continuity equation in the configuration space


becomes
(8.38)
or, in terms of Vi and u i
-UiV i + v ·aiv i = o. (8.39)
In the case of a two-particle system, following Kershaw, we can derive
equations of type (8.34) for the mean velocities v~(xj, Sl' S2) in some external
field F~(Xj, S1> S2):
v~(xj, Sl + c; ,1s, S2 + c; ,1s)

= f
~± d8YE[V~(XI - c;yI, Xl + iYi, X 5 + iY~, Sl' S2)

(8.40)

where
294 Chapter 8

are normalization factors; M is some effective mass (it can be of a matrix form
with respect to m - mass of the scalar particle) of our two-particle system; and
1 for v~,
8- {
- -1 for vi_.

In our case, equations (8.40) lead to the differential equations

au + + --
--
i au~ . .
as as + u1± aj u'±
F~
-- ~
(2..
- v -v u aj u'±
M +
1 + Ou'±.) .
(8.41 )
1 Z

Adding equation (8.41), we obtain

. .1. . .
D v' - Ds u'
C
= -(F'
2M + + F'- ) = F'/M',

(8.42)
Equation (8.42) in conjunction with the continuity equation (8.38) is the
covariant analog of the single-particle case for a two-particle system.
Note that the left-hand side of equation (8.42) is identical to the expression
for the 'acceleration' obtained by Cufaro Petroni and Vigier (1979) on the
basis of some assumptions in the framework of the mathematical approach of
Nelson (1966, 1967), and Guerra and Ruggiero (1978).
The coupled pair of nonlinear differential equations (8.39) and (8.42) can be
linearized if we set

(8.43)

as before, where ¢(X i, S1' sz) is the phase function determined by the equation
(8.44)
Using expressions (8.37), (8.43), and (8.44), on the basis of equations (8.38) and
(8.42) we obtain the equation of Hamilton-Jacobi type
(8.45)
for the two-particle system in the case when there is no external force: Fi == O.
Here, we have set R = p1/Z, V = Ii/2m. From equation (8.45) there follows a
continuity equation of the form
2 aiR ais + R ai ais = O.
Finally, we have a formal equation for tjJ = R exp(ih -1 S):
(0 - 2m zcz/liZ)tjJ = O. (8.46)
In the nonrelativistic limit equation (8.46) leads to an ordinary two-particle
Stochastic Mechanics 295

Schrodinger equation for


t/J(x l , X2, t) = R(xI' x 2,t) exp(ih- I S)
which decomposes into two equations:
ap/at + VI(P V IS/m) + V 2(P V 2S/m} = 0,
where P = R2 = t/J* t/J, and
as/at + (l/m)(V I S}2 + (l/m)(V 2S}2 + Q = 0.
Here,
Q= - (h2/2m)(ViR/R + V~R/R)
is some potential, and it is usually called the nonlocal quantum potential of the
two bodies (see, for example, Cufaro Petroni and Vigier, 1979).
We note in conclusion that the physical consequences of the results obtained
above are discussed by Cufaro Petroni and Vigier (1979, 1983a).
Chapter 9

Selected Topics In Stochastic Mechanics

It should be noted that because of the fact that stochastic mechanics is a


nonlinear system, it will contain much physical information, and in particular
may admit some interesting solutions describing such phenomena as solitary
or soliton-like waves, turbulent and Feynman type processes, and so on.
Recently, studies in this direction have been initiated. A role of the Feynman
path integral theory in stochastic mechanics has been discussed by Guerra
(1981) and Moore (1980). We believe that the full clarification of the deep
connection between these two approaches will be a major step toward a better
understanding of the physical foundations of quantum mechanics.
There are some studies of solitary waves (Efinger, 1981) and soliton waves
(Gueret and Vigier, 1982) associated with the space-time structures near
particles and to the stochastic interpretation of quantum mechanics, respec-
tively. A possibility of the origin of processes of such types will be discussed
in Section 11.8.
Another interesting problem in stochastic mechanics is how to construct a
comprehensive stochastic scheme allowing stochastic processes in momentum
space to be considered. Some results are given by Shucker (1980b) and de
Falso et al. (1982, 1983), where it is found that quantum-mechanical momen-
tum can be read from the asymptotic behavior of Nelson's sample paths and
the position-momentum uncertainty product is obtained too.
In this chapter we shall investigate the equation of motion, the Cauchy
problem for stochastic mechanics, and some its specific problems such as the
self-turbulent phenomenon and Feynman-type processes. We also attempt to
describe quantum-mechanical momentum from the point of view of the
stochastic processes under consideration.

9.1. A Stochastic Derivation of the Sivashinsky Equation for the


Self-Turbulent Motion of a Free Particle

Sivashinsky (1978, 1980) noted a formal analog between the equation of


motion for a flame front and the Hamilton-Jacobi equation for the motion of

296
Selected Topics in Stochastic Mechanics 297

a free particle. He showed that if in the equation for the flame front one
introduces terms with higher derivatives, describing the structural perturbation
of the front, the front is unstable with respect to long-wavelength per-
turbations (Sivashinsky, 1977; Michelson and Sivashinsky, 1977). As a result,
the original deterministic equation can generate solutions of a stochastic type.
An attempt was made to interpret the equation with higher derivatives as a
Hamilton-Jacobi equation describing the motion of a 'quantum' particle.
However, in the framework of Sivashinsky's approach the choice of the
potential (the self-interaction potential of the particle) which 'generates the
turbulence' in the Hamilton-Jacobi equation is not unique and does not have
a clear physical justification.
This section is devoted to the derivation of a Sivashinsky equation for the
self-turbulent motion of a free particle in the framework of stochastic theory
on the basis of the hypothesis of a stochastic space-time (see Chapter 1). We
are therefore attempting to justify the occurrence of the potential which
generates the turbulence in the equation of motion of the particle.
We consider first the nonrelativistic motion of a single scalar particle in the
x
stochastic space ~3(X) with coordinates = x + b. In accordance with the
results in Chapter 1, by the physical quantity f(x, t) we shall understand
< f (x, t) u;l3(xl' averaged with respect to the measure w(b) at a given time t (in
particular, see formula (1.6) in Chapter 1). The assumption that the stochastic
component of the space ~3(X) is small means that

F(x, t) = <f(x + b, t)u;l3(X) = fd3bW(b)f(X + b, t)

= <f(x, t) + bJ%xJf(x, t)+±b;bp2/0X; oxj)f(x, t) + ... )


~ f(x, t) + [2 !J.f(x, t). (9.1)
It is assumed that w(b) = w( - b).
It should be noted that with a suitable choice of the function w(b) the
method of averaging (9.1), as shown above in Chapter 1, which ensures the
transition from the small to the large scale, leads to an entirely new object:

and in the relativistic case

where
c'
L Wo)n <5(4)(X),
OJ
K(x ll ) = _ n1
n= 0 (2n).
29S Chapter 9

and the coefficient Cn and (c~) depend on the particular choice of the measure
w(b) [w(b~)]. An object of this kind was constructed for the first time by
Blokhintsev (1973a) in quantum field theory for the special case when
bl' = ayl" where a is a stochastic variable and Yl' are the ordinary Dirac
matrices. In quantum field theory, such an object has been investigated in
Chapters 1 and 2 from the point of view of generalized functions (or
distributions) K(x) whose space-time properties depend essentially on different
sequences of the coefficient {c~}. It is shown (in Chapter 1) that an object
constructed by means of such generalized functions K(x) is smeared (non local)
in space (see also Efimov, 1977a).
Thus, the procedure for averaging (9.1) in the framework of our hypothesis
leads to a nonlocal object. However, such averaging cannot change the
physical nature of the object (for example, if it were stochastic before the
averaging, it remains so after it) but merely the spatial structure of the object is
changed, which is smeared in a certain region determined by I. In calculations
of local and nonlocal objects such as the velocity vex, t), the randomness effect
which arises from the fluctuation of the spatial coordinates accumulates with
the course of time, is manifested only in the dynamical aspect, and does not
depend on the intermediate averaging procedure (9.1). Then, in our case, the
localf(x, t), and the nonlocal, Fn(x, t), objects can be assumed to be stochastic
quantities. At the present time, it is not clear how one should write down the
probability equation (an equation of the Chapman-Kolmogorov type; see
Chapter 6) for a nonlocal stochastic quantity Fn(x, t). Therefore, in a rough
approximation, when the parameter I is infinitesimally small, we shall assume
that Fn(x, t) is determined by a finite sum of the form (9.1). The object
Fn(x, t) = f(x, t) + h(X, t), (9.2)
where fl(X, t) is determined by a finite sum of a series in the parameter I, as
shown above in Section 1.4.6, possesses like f(x, t) a local property and,
therefore, for Fn(x, t) in (9.2) we can write an equation of the Chapman-
Kolmogorov type.
As a first approximation in the parameter I, we shall ignore the second term
in (9.2), i.e.,
<f(x, t) >= Fn(x, t) ~ f(x, t).
It is to this approximation that our previous results apply (Chapter S), and in
this section we shall not ignore the second term in expression (9.2).
We now attempt to obtain the general form of the dynamical equations of a
scalar particle in the stochastic space when a term of order 12 is present in the
expressions for the velocity and the force in accordance with equation (9.2):

It is assumed that in equations (S.6) and (S.7) [(S.20) and (S.21)J of the
Smoluchowski type for v± (v~) the small corrections v~ and f~ (V~,l and F~,l)
Selected Topics in Stochastic Mechanics 299

occur only in a symmetric combination with respect to Llt and bx (Lls and
bx~ = yt), i.e., they are even functions under the transformations Llt ~ - M
and bx ~ -bx(Lls ~ -Lls and bx~ ~ -bxn In this case, v± and f±(v~ and
F~), which occur in equations (8.6) and (8.7) [(8.20) and (8.21)], are replaced
by the expressions

v+ + L v1+ and f± + L f~,


- {M,ox)- {ox)

where the symbol L{,,) denotes the operation of symmetrization with respect
to the variables { ... }; for example,

L v~ (x -
~} -
y) = L v~ (x + y) == t[v~ (x + y) + v~ (x -
~) - - -
y)].

As a result, we obtain equations analogous to (8.9), (8.10), (8.23) and (8.24):


dev -)~ d~u = (ljm)F;, deu + A d~v = (l/m)F)~, (9.3)
dev - ), d~v = (l/m)F;., deu + A d~u = (l/m)F~, (9.4)
and
DeVil - A D~UIl = (l/m)ll>~+lll, DeUIl + A D~VIl = (l/m)ll».-lll, (9.5)
Devil - )~ D~VIl = (l/m)ll>i, Deull + A D~UIl = (l/m)ll>'f (9.6)
in the nonrelativistic and relativistic cases, respectively. Here
(9.7)
Naturally, if we ignore the terms of order v[2V 4 (V[2 0 2 ) in expression (9.7) for
d~(D~), then, as one would expect, we obtain the old equations (8.9) and (8.10)
[(8.23) and (8.24)].
In the limit v + == L, i.e., u == 0 we obtain from (9.3) and (9.4) Newton's
equation de V = F /m and also the equation
de V- Jcv(V 2 + FV4)V = F)/m
for the particle.
For F}, = 0, these last equations are equations of the Sivashinsky type for a
free particle. For example, setting A = -1, we obtain
(9.8)
which is invariant under a Galileo transformation.
Note that for the system of equations (9.3) [or (8.9)] the condition v ~ 0 in
the 'classical' limit is consistent, since the first equation of (9.3) goes over into
300 Chapter 9

Newton's equation, and both sides of the second equation of (9.3) vanish in the
limit when v -> O. We recall that F+ is the external force, which does not
contain a stochastic component, and the second equation of (9.3) for A = 1
agrees with the continuity equation (see also Section 9.3) op/ot + V(pv) = O.
Therefore, if the Lorentz force is introduced in a natural manner in our
formalism, we can write
Fl = -m[u x (V x v) + vV x (V x v)].
In the case of the electromagnetic field, it takes the form
e e
F1 = - u x H + - vV x H.
c c
Conversely, in the system of equations (9.4) it is impossible to go over formally
to the condition v -> 0 in the 'classical' limit, since in the expressions for the
forces FA and F~ the ordinary and stochastic components of the force are
already present. It is entirely possible that the stochastic part of the force does
not necessarily vanish even as v -> 0, and on the other hand the stochastic
velocity u = vV In p vanishes in the limit v -> O. Then the stochastic terms on
both sides of equation (9.4) do not vanish in an equal way as v -> O.
A special method is needed to investigate the nonlinear equations (9.3), (9.4),
and (9.8) whose solution goes beyond the scope of the present book. In the
papers by Sivashinsky (1977); Michelson and Sivashinsky (1977), equation
(9.8) was studied numerically and it was shown that its solution behaves like a
turbulent processes.
In this section, we shall make some general comments about these equa-
tions. First, we note that equation (9.8) is equivalent to the Hamilton-Jacobi
equation only in the case when the 'flow' v = VS is irrotational, i.e., curl v = O.
Otherwise, the solution of equation (9.8) grows unboundedly with the time.
For example, let us consider the field v = (u, v, w), where
u(x, y, Z, t) = v(x, y, Z, t) = 0; w(x, y, Z, t) = <p(x, y, t).

Note that div v = O. For such a field, all the nonlinear terms in equation (9.8)
vanish, and we obtain a purely linear equation for <p(x, y, t):
o<p/ot + vVz<p + v[2v 4 <p = O.
This equation has spatially periodic solutions of the type
<p = A exp (at + ikl x + ikzy),
where
a = v(ki + kD - v[2(ki + kW.
If ki + k~ < [-Z, then the amplitudes of these solutions grow exponentially.
At present there are difficulties connected with the physical interpretation of
such a noncausal behavior of the amplitudes; perhaps only a hypothesis on the
existence of tachyons and a non ordinary version of the causal conditions will
Selected Topics in Stochastic Mechanics 301

permit this noncausal behavior of signals at least in a domain determined by


length I.
The second question relates to the choice of the sign of the parameter A. For
example, if the terms of order vP are ignored, the second equation of (9.3) for
A = - 1 does not agree with the continuity equation, and, therefore, the case
A = -1 for (9.3) is eliminated from consideration altogether.
When A = 1, the equation

(9.9)

of type (9.8) obtained from equation (9.4) is such that for it there is no
scattering (dissipation) in the short-wavelength regions. It is well known that
for such equations an important problem - the initial value problem - is not
sufficiently formulated. For this reason, the case A = 1, i.e., equation (9.9),
simply has no physical meaning if one does not assume the existence of terms
with derivatives of higher than the fourth order.
Since our approximation includes only terms with derivatives of the fourth
order, the case A = 1 [equation (9.9)J is eliminated. Of course, the questions
posed above remain valid in the relativistic case.
In the relativistic case, Sivashinsky's equation (9.8) now takes the form

(9.10)

which is obtained from equation (9.6) for ul1 == 0 and <1>11 = O.


Thus, adopting the hypothesis of a stochastic space-time, we have obtained
Sivashinsky's equation. We have shown that the resulting instability of the
uniform and rectilinear motion of a free particle leads to random fluctuations
of its trajectory. Despite the purely classical nature of the original equation,
typically quantum effects are imitated: the uncertainty relation, de Broglie
waves and their interference, discrete energy levels, and zero-point fluctuations
(for more details see Sivashinsky, 1978, 1980).
As we have seen above, the self-interaction potential of the particle [the
right-hand side of equations (9.8) and (9.10)J which generates the turbulence in
the motion of the free particle has a stochastic origin. In other words, the
stochasticity, which disappears in the limit u 0;> 0, renders the motion for v
unstable and preserving thus a memory of itself.

9.2. Relativistic Feynman-Type Integrals

In the scheme of quantum mechanics proposed by Feynman (1948) an


analog of the probability measure is used which allows one to describe the
behavior of quantum mechanical particles. Kac (1959) noticed that this
measure is a complex quantity. This complex measure in the Feynman path
integral corresponds to the presence of the factor i in the exponential for the
Wiener measure (see Chapter 5).
302 Chapter 9

Roughly speaking, the presence of i in the exponential for the Wiener


measure causes uncontrollable oscillations in the path integral. This makes it
sometimes difficult to understand the Feynman path integral as a well-defined
mathematical object. Despite this drawback it is a matter of history that the
path integral is an extremely important contribution to quantum theory.
There are many approaches devoted to the generalization of the Feynman
integrals to the relativistic case (see Feynman, 1950, 1951; Miura, 1979 and
references therein). Interrelation between the Feynman path integrals and
stochastic mechanics, and the path integral method as derived from a stochas-
tic variational principle have been presented by Guerra (1981) and Berrondo
(1973), respectively.
In this section we shall construct the Feynman-type integrals in the frame-
work of our approach using Smoluchowski equations. Equations obtained
allow the Schrodinger, Klein-Gordon, and Dirac equations to be derived
easily. The interaction in the Smoluchowski-type equations for fields cp(x, t) is
introduced here by means of Weyl's (1929, 1950) gauge theory.
In our model the Feynman process may formally be interpreted as a
stochastic process in complex time with a real probability measure of the
Gaussian type, which occurs in the Euclidean space.

9.2.1. Diffusion Process in Real Time


In the language of motion of a stochastic particle the property of the Markov
process means that the character of displacement of a particle at a given time
does not depend on the property of previous displacements. Accordingly, the
probability density p(x, t) must obey the Smoluchowski equation

p(x, t + M) = f d 3 (<5x)p(x - <5x, t)Po(x - <5x, t; <5x, ilt1 (9.11)

where Po is the conditional probability density that a particle at pOSItIon


x - <5x at time t will be displaced by <5x during the interval M, thus reaching
position x at time t + M. The simple form
Po = (4nv ilt)-3/2 exp{ -(<5x)2/4v ilt} (9.12)
reduces to the diffusion equation for p(x, t) (see Chapter 7)
op/ot = vV 2p, (9.13)
where v is the diffusion coefficient.
Following our model (see Chapter 8) in the relativistic case, we consider
formally the motion of a particle suffering the random walk due to stochas-
ticity of the four-dimensional Euclidean space 1E4(X, r). Then equation (9.11)
acquires the following form

p(xl" S + ils) = fd 4YEP(x - y, Xo + iY4' S)Pl(X - y, Xo +


Selected Topics in Stochastic Mechanics 303

(9.14)
where the variables xI' = (x o, x) are pseudo-Euclidean and P 1 can be chosen in
the form
P1 = (4nv ~s) - 2 exp( - y~/4v ~s). (9.15)
From (9.14) and (9.15) we have
op/os = vOp; (9.16)
Note that a more complicated form of the functions (measures) Po and P 1
make Fokker-Planck equations for p in both the nonrelativistic and re-
lativistic cases, respectively (see Chapter 8).

9.2.2. 'Diffusion Process' in Complex Time


Now a basic postulate is that a field <p(x, t) (probability amplitude) associated
with a particle suffers a transformation and is defined as a diffusion process in
complex time, whenever the particle displaces from point x - bx at time t to
position x at t + At. Then the corresponding Smoluchowski-type equation for
the field <p(x, t) becomes

<p(x, t - i ~t) = f d 3 (bx)<p(x - bx, t)p o(bx, ~t), (9.17)

where Po is given by (9.12). It is easily seen that from equation (9.17), we


obtain the Schrodinger equation assuming v = II/2m.
Substituting At --> - i At into equation (9.17) gives the Feynman integral,
and therefore in the nonrelativistic case our postulate does not make a new
resLlt in the method of the Feynman path integrals. An essential difference
appears in the construction of the relativistic Feynman-type integrals by using
the diffusion processes. We now pass to this question.
Roughly speaking, in the relativistic case the Feynman-type integrals for the
probability amplitude is formally replaced by Smoluchowski-type equations at
complex time. So, if the Feynman process <p(xl', s) is known at one value of s,
its value at a slightly larger value s - i ~s is given by

(9.18)

From expression (9.15) and equation (9.18), we obtain the Klein-Gordon


equation in the parametric form
i o<p/os = - vO<p. (9.19)
Sometimes, instead of equation (9.19), the following equation is considered
(9.20)
304 Chapter 9

which is obtained by using the measure:


d,u(YE' ~s) = d 4 JE(4nv ~S)-2 exp( -m 2 v ~s - y~/4v ~s).
Formal formulas (9.17) and (9.18) will be interpreted as well-defined ma-
thematical objects - some integral equations with a real probability measure of
the Gaussian type. Here quantity v is real always.

9.2.3. Introduction of Interactions into the Scheme


As is clear, in the nonrelativistic case our formalism is equivalent to the
Feynman integral if ~t ~ - i ~t. Then due to Feynman we can write equation
(9.17) in a potential field U(x) by the following formula

cp(x, t - i M) = f
d 3 (bx) exp[ - 2~V U(X)]CP(X - bX, t)Po(bX, t).

From this we obtain the Schrodinger equation

II ocp 1 (II-;-V )2 cp + U(x)cp


--;--=-
I at 2m I

if v = II/2m.
In the relativistic case we introduce the interactions into our scheme within
the framework of Weyle's gauge theory (see Miura, 1979). Following this
theory, the field takes the value cp(xl' + dxl" ~gil) after the transport connected
with the displacement of a particle from a world point gil(co-ordinates xI') to a
position gil' (co-ordinates xI' + dx,J, and therefore the variation bcp of cp made
by this transport is given by
bcp = cp(xl' + dx", ~gil) - cp(x,,) = dN cp(x,,). (9.21)
If dN is the total differential of a co-ordinate function ief(x,,)/lIc, i.e.,
. e of
dN=I-
II c ~dx",
uX"

then (9.21) affects only an arbitrary phase of cpo Generally, dN is given in the
form

(9.22)

where All is an electromagnetic potential. We assume dx" = x" - x~ and


rewrite (9.21) in the following form

cp(x ll , ~gil) = exp(dN)cp(x~) = exp( - ~: AI' dx" )cp(X~). (9.23)


Selected Topics in Stochastic Mechanics 305

By using this formula the Feynman path integral for a Klein-Gordon particle
may be defined in the form (Miura, 1979)

cp(x!', s + e) = fd 4 X expGso)
1
Q(±2nihe)-1/2cp(x!', Ll&),
where So is the action of a free particle.
Because the displacement of the variables x and Xo in our case is different,
i.e., it is of the Euclidean character, the corresponding formula (9.23) must be
changed in the following manner;

cp(x!', Ll&» = exp{2~VC A~y~ }cp(x - y, Xo + iY4), (9.24)

where A~ = (-iAo, A) and A~y~ = A 4 Y4 + Ay. Then we obtain the value of cp


at a space-time point after the transformation using averaging over cp shifted
by all possible Euclidean displacements with a real probability measure
P1(YE' Lls)d 4 YE of the Gaussian type and multiplied further by a
weight function exp {(ie/2mvc)A; y~} for an infinitesimal value of y~:

cp(x!', s - ills) = fdjl(YE, Lls) exp(~


2mvc
A~Y~)CP(X - y, Xo + iy4, s), (9.25)
where
djl(YE' Lls) = (4nv LlS)-2 exp( - yi/4v Lls) d 4YE.
From this we obtain the parametric Klein-Gordon equation in an external
field

. ocp ( a ie )2
I as = v ox!' - 2mvc A!, cp

if v = h/2m.
The generalization of our formalism to a Dirac particle does not present any
difficulty, for example, expressions (9.24) and (9.25) take the form

,I, {1. 1
'I'(x!"Ll &> )=exp 21e2mv}A~E Yv'Yv+Y!''Y!''A~'E ] } t/J(x-y,x o + lY4,S),
.
(9.26)

and

x t/J(x - y, Xo + iY4, s), (9.27)


where
A;,E = A~(x - y, Xo + iY4)
'Yare Dirac matrices.
306 Chapter 9

After some calculations we have the Dirac equation in the parametric form

here

Finally, notice that equations (9.18), (9.25) and (9.27) we have obtained,
generally speaking, have nothing to do with the Smoluchowski-type equations
which describe the probability consequence, and therefore the value P 1 (YE' Lis)
is not interpreted as the transition probability. On the contrary, these equa-
tions may be interpreted as a formal exposition of some mathematical objects
obtained by integrating with the real probability measure dll(YE' Lis) of the
Gaussian type. Feynman path integrals themselves are not obtained in our
scheme owing to the fact that variables x and Xo are shifted in a different way:
x -+ x - y and Xo -+ Xo + iY4.
However, our method is interesting as representing the possibility of re-
lativistic generalization of Feynman-type integrals.

9.3. Discussion of the Equations of Motion in Stochastic Mechanics

The description of a stochastic system is complete if its probability density


p(x, t) and velocity v(x, t) are known, since the stochastic velocity u is related
to p(x, t) by
u(x, t) = vV In p(x, t).
For them, we have obtained the equations (Chapter 8)
8p/8t + div(pv) = 0, m(d c v - dsu) = f, (9.28)
and in the relativistic case

where JI'(x) = pvl' is the four-dimensional current (see Appendix C III this
chapter). In addition, there are equations for u and ul':
(9.29)
On physical grounds, it can be asserted that these basic equations must
satisfy a number of requirements:
(1) To ensure that the physical quantities p and v are not overdetermined, it
is necessary to require that the last two equations (9.29) for u and ul' be
equivalent to the continuity equations for p and JIl, respectively.
(2) We shall assume that in stochastic mechanics the work done by a field
Selected Topics in Stochastic Mechanics 307

on a particle is associated only with the velocity v and does not depend on the
velocity u, i.e.,
<(d/dtHmv 2 » = (fv); d/dt = a/at + (vV).
In the absence of an external force, this condition amounts to conservation
of the kinetic energy in time.
(3) In the relativistic case, we must have v/lv/l = c l . Thus, on the basis of the
first requirement we can choose the form of the external forces fs and F~, and
the requirements (2) and (3) impose the further conditions
(v dsu) = 0; VII Dsu ll = 0. (9.30)

Note that in the nonrelativistic limit the last condition goes over into
v dsu = 0, so that these conditions are interrelated.
As an external field, we now consider the electromagnetic field and show
that equations (9.29) with external force
e e
fs = ~ u x H + ~ vV x H; H = curl A;
c c

e. e,.
F~ = ~pU'u;. + ~v aAFAIl (9.31)
c c
are equivalent to equations for p and JI1, respectively. In the non-relativistic
case, we set
e
mv + ~A = VS,
c
where S is the action. Then the first equation of (9.29) with allowance for (9.31)
gives
deu + dsv = -u x curl v - vV x curl v. (9.32)
Applying the operator V to the continuity equation for p and using the
formula u = vV In p, we obtain
au/at + V(u'v) + vV(V'v) = 0. (9.33)
Using the well-known formulas of vector analysis,
V(u·v) = (uV)v + (vV)u + u x curl v + v x curl u,

V(V 'v) = L'lv +V x (V x v), curl grad == 0,


we arrive at equation (9.32), which is what we wanted to prove.
As in the derivation of equation (9.32), the time component of the second
equation in (9.29) can be reduced to the form
(9.34)
308 Chapter 9

Here, we have used equations (9.31) and the equation (see Landau and
Lifschitz, 1971):
e 8S
mvl' + -AI' =
C 8x"·
Differentiating the continuity equation for JI', we find
Deuo + Dsvo = (uV)vo - u 80 v - vV 8 0 v + v dv o + (vV)u o - v 8 0 u.
Since Uo = v8 0 In p and u = vV In p, and therefore (vV)u o - v8 0 u = 0, from
which we obtain equation (9.34).
Similarly, we can easily show that the three spatial components of the
second equation of (9.29) are identical to the vector equation
Deu + Dsv = -voVu o + Vo 8 0 u - uoVvo + Uo 8 0 v-
- u x curl v + v 80 V Vo - v 8~v - v V x curl v,

which is obtained by differentiating the continuity equation for J" by the


operator V.
We now turn to the additional conditions (9.30). The equation
v·B = 0, (9.35)
where
B = dsu = (uV)u + v du
has a nondenumerable set of solutions (we assume that B is an unknown
vector), since it determines only the component of B in the direction of the
vector v, the value of which is Bv = 0; however, the component perpendicular
to v remains entirely arbitrary. Thus, if B is regarded as the radius vector of
some point M relative to the origin 0, the locus of the ends of all vectors B
satisfying equation (9.35) will be the plane perpendicular to v through 0. In the
relativistic case, this question was discussed by Vigier (1979) from the
geometrical point of view.
To conclude this section, we note that our equations (9.28), (9.29), and (9.33)
are nonlinear differential equations, so that their direct solution is an almost
insoluble problem in concrete applications. Finally, we give some special
solutions of e,quations (9.28) for the simplest case of the free and uniform
motion of a particle with the initial conditions
p(x, t = 0) = n- 1/2 ·Z-l exp[ - (x - xo?/F]; v(x, t = 0) = vo.
These solutions have the form
p(x, t) = n- 1/2 Z- 1 (1 + bt 2 )-1/2 x
x exp[ -(x - Xo - votf/F(l + bt 2 )]; b = 4v 2 Z- 2 ,
v(x, t) = [vo + bt(x - xo)]/(l + bt 2 ),
u(x, t) = - jb(x - Xo - vot)j(1 + bt 2 ),
Selected Topics in Stochastic Mechanics 309

and their mean value and dispersion are

<v(X, t) = f dx v(x, t)p(x, t) = Vo,

<u(x, t) = f dx u(x, t)p(x, t) = 0,

<x(t) = fdX xp(x,t) =_xo + vot,

. 2 _ 1[2
b2 t 2 .
DIS. V (x, t) - z b 2'
1+ t

2 2 1 2
b2 t 2
<V(X,t)=VO+ Z[ 1+bt 2'

In the case of a uniform translational and rotational motion of particles


(f = const) solutions similar to those above can be readily obtained.

9.4. Cauchy Problem for the Diffusion Equation

°
We consider the following equation in two-dimensional space-time
[first equation of (8.18) with u + = in Chapter 8]
(Xl and x o)

Since the variable sand Xo = ct are related by dXojds = c(1 - {32)1/2, and in
the rest frame of the particle they are simply equal; then this equation can be
written in the form (now it is taken Xl = X, Xo = t)
(0 2 - a 2 8j8t)p(x, t) = 0, a2 = 1jv.
We now investigate the Green function of the equation
(0 2 - a 2 8j8t)G(x - Xo , t - to) = -b(x - x o) b(t - to), (9.36)
where (x, t) is the point of observation and (x o , to) is the source point. To
construct the function G, we consider first the two functions

D 1(x, t) = 2~i fd2P8(P~) b(iC PvO - p2) exp(ipocr - ip1R), (9.37)


310 Chapter 9

where
Po = Po - ic/2v, Po
= Po + ic/2v, 't = t - to,
R = x - x o, P = (P1,PO).
Mter elementary calculations, we obtain (P1 = p)

D 1(x, t) = ~
4nz
exp[-c 2 t - to]
2v
f dPW - 1 exp(icw't + ipR),

D 2(x, t) = :n exp[ - c2 t ;vto ] f dpw -1 exp( - icw't + ipR),


where
w = (p2 _ c 2/4v 2)1/2.
We now consider the integral (see Morse and Feshbach, 1953)

D2 = ~ exp[-c 2 t -
4n 2v
to] r dp(p2 _
Jc
J1. 2)1/2 exp{i[Rp _(p2 _ J1.2)1/2C'tJ},

(9.39)
(1) If R > c't, the contour C can be closed along the semicircle in the upper
half-plane. Then the integral (9.39) is equal to zero, since there are no
singularities inside the contour (Figure 13).
(2) If R < C't, the contour C can be displaced so that it is stretched along the
negative imaginary half-axis (Figure 14). In the last case, the integral (9.39) can
be reduced to a well-known form. Let us introduce the new variable e so that
C't = L sh e,
L = I(R 2 - C2't 2)1/21,

Im p

c
Rep

Fig. 13. The contour C for integral (9.39).


Selected Topics in Stochastic Mechanics 311

Imp

Fig. 14. The contour C for integral (9.39) at R < cr.

and set p = J1 ch x. Then


i[ Rp - (p2 - J12)1 /2CTJ = iJ1(L ch 8 ch x - L sh x sh 8) = iJ1L ch (x - 8);
.
D2 = ~ e -I'""
l-in I 2+ 00
dx exp[iJ1L ch(x - 8)].
2n 3nil2 + 00

(3) Setting x - 8 = iy, then


1
D2 = _ _ e-/lCt
f-"/2 - ioo
dyexp(iJ1Lcosy).
4n 3"/2 - ioo

By the equation

--
1 f-"/2-ioo dy exp(iJ1L cos y) = J o(J1(R 2 - c2 T2)1/2),
2n 3"/2 - ioo

the function D2 takes the form


D2 = t exp( - J1CT)J o(J1(R 2 - c2 T2)1/2)8(CT - R).
Let us now consider the integral

D1 = ~ e-/lCt
4nz
r dp(p2 - J12)-1 /2 exp{ipR + iCT(p2 _ J12)1/2}.
Jc
This integral is equal to zero when R + CT > 0 (we recall that in the one-
dimensional case R may be negative), but it differs from zero when R + CT < O.
In this case we have
D 1 = -t exp( - J1CT)J o(J1(R 2 - c 2 T2)1/2)(1 - 8(R + CT)).
Green's function of equation (9.36) is given by the following expression

G(R,T) = -C(D2 - D 1) = tceXP(-J1CT)Jo(;v (R2 - C2T2)1/2 )8(CT -IRI).


312 Chapter 9

It is readily seen that as one would expect

lim -1 G(R, ,) = (4nn)-1/2 exp(_R2)


-- .
c~oov 4v,
Further, for comparison we write the Green function of the Schrodinger
equation in the proper-time representation:
i otjJ/os = -vDtjJ,
or

where

x = tca 2 (c 2,2 - R2)l/2,


and in the limit c --> 00 the function K(R, ,) is equal to the Green function
G = i- 1(4nn)-1/2 exp(iR2/4n)

of the equation
i otjJ/ot = -V(02/0X2)tjJ.
We now solve the one-dimensional initial-value problem

p(x, t) = a2 fdXo[PG]to = 0 + c- 2 fdxo[G oOP - P OOG] ,(a2 = 1/v),


to to to = 0
Whence

p(x, t) = te- {f(x + ct) + f(x -


lUt ct)
r+ct dxof(x o) x
+ L-ct

(9.40)
where
x = Jl[C 2 t 2 - (x - XO)2]1/2;

v(x o) = op(xo, t)/otlt=o'


Here J o(x) and Io(x) are the well-known cylindrical functions.
It should be noted that if functions f(x) and v(x) are strictly equal to zero
outside some interval [ - a, a], i.e., if they are functions with a finite support
situated in the interval [ -a, a], then the study of the integral (9.40) will not
represent any trouble and moreover, the causality condition (local) is strictly
Selected Topics in Stochastic Mechanics 313

((xl

Fig. 15. The illustration of a function with infinite support.

fulfilled (see Chapter 2). The situation changes if we consider the functionsf(x)
and v(x), support of which is infinite (see, for example, Figure 15). Our aim
now consists of studying the behavior of integrals of the type

T1
11
= - -
foo (_Z2)
dz exp -2- Io(X),
2c -00 a

and

T2 = const t f OO

-00 dz exp
(

7
- Z2) T'
I (X)

X = {l[C 2 t 2 - (z - X)2]1/2,

at the limits t -> 00 and Ixl-> 00.

In the case of Ix I -> 00, we assume the variable t belongs to the future
(forward) light cone, i.e., tEV+. For simplicity, let us take the point M of
observation with the co-ordinates (C 0), i.e., M = M«( = t, 0). Then the integral
T1 takes the form

T1 = const foo dz exp(-~2) Io(Q),


-00 a /
where Q = {l(c 2 t 2 _ Z2)1/2.
Making use of the integral representation for the Bessel function In(z):

I n (z)(z/2)-n = r- 1 (n + !)r- 1 f1
(t) dx(1 - x 2 t- 1/2 eZx ,

we obtain

f1
T1 = const -1 dp(l - p2) -1/2 _ 00 dz exp
foo (_Z2)
7 exp(pQ).
314 Chapter 9

Here

f~cc dz exp( ~~Z) exp(pQ) ~ exp( _ e:~z + ia zI1 zpZ ).


and, therefore

Similarly,

T z = const[ r(~;r(t)] f1 dp(1 - pZ) -liZ f~CXl dz ex p( ~~Z) exp(pQ)

Thus, in the case of functions f(x), v(x) with an infinite support the signal (as
possible, noncausal) attenuates as exp( _tz/a Z), when t ~ 00.
Let us now consider the behayior of the function p(x, t) at Ix I ~ 00:
(1) Let t be a finite value, then

Tl = ~ fet dz ex p [ - (z ~ X)Z]Io(Q).
2e -et a
Going over the substitution z = ety, we obtain

et
Tl =-exp
2e
(-
-z-
a -1
Zt ZyZ 2et ]
XZ) fl dyexp ---z---z
a a
[e
yx Io(l1et(1-yZ)1/Z).

Setting again x(y + 1) = p and y = pix - 1, yx = p - x, we obtain


Tl = (t/2x) exp( -xz/aZ)S(x),
where

S(x)=Jo dpexp
rZx {e
-7
Z t Z(p
~-1
)Z ---;;z(p-x)
2et }
x

and

lim S(x) ~ ( -aZ ) exp (2et eZ t Z)


- Z x - - Z 10(0).
x~oo 2et a a
Selected Topics in Stochastic Mechanics 315

Therefore,
T1 '" (a 2/4cx) exp[ - (x - ct)2/a 2],

i
for x -> 00. Similarly,

_.1 x + ct (-Z2)11(X) /l2c 2t 2 2 {-(X - ct)2}


T2 - 2/lct dz exp -2- - - '" - 8 - a exp 2 '
x-ct a X x a
where
X = /l[C 2t 2 - (z _ X)2]1/2.

(2) Let t ~ 1; then the region of integration over z can be extended over the
whole plane. In this case, the estimates obtained above are not changed. For
example,

T1 =
1 Jct
-2 dz exp
[:... (z +
2
X)2J10(Q)
c -ct a

=> -
1 Joo
dz exp {- (z + X)2} 10(Q) =
2 A(x) + A( -x),
2c _ 00 a
where

tt exp ( 7_X2) Joroo dy exp{C t2 2 2xctY }


A(x) = -~ y2 - ----;J2 10(/lctJl~)

-x 2 2xct)
=> (t/2x) exp ( ----;;z- +7 g(x);

2 t2(p)2 /
g(x) = Jxroo dpexp {C
-7 ~-1 2 10 ( /lct [ 1- (p~-1 )2J1 2) ,
- 2P act}

(x + xy = p).
It is easily verified that when x -> 00 the integral g(x) is less then any e chosen
arbitrarily small i.e.,

x~ 00

A similar estimation holds for T 2 • As a result, we have


Ip(x, t)1 « e exp{ -(x - ct)2/a 2},
when Ixl -> 00 and t ~ 1.
Thus, we see that instead of the condition p = 0 for the case of functions f(x)
and v(x) with a finite support, i.e., located in the interval [- a, a], we have
obtained the estimates of the type
p(x, t) '" exp{ -(x - ctf/a 2}
316 Chapter 9

when Ixl-+ 00 and t -+ 00 for the case of the functions f(x) and v(x) with an
infinite support. In this connection we notice that, in accordance with the
macrocausality condition (2.110) or (2.112) discussed in Chapter 2, non-causal
signals attenuating as exp{ -Ix + ctI2/a 2 } are acceptable for the practical
applications, in particular, for the description of physical processes taking
place in the macroscale.

9.5. Position-Momentum Uncertainly Relations in Stochastic Mechanics

In stochastic mechanics the concept of the configuration space and con-


figuration observables plays a specific role. In contrast with classical mech-
anics and quantum mechanics where all constituents of phase space have a
similar role, this feature is that the relevant random processes have always
been defined as taking values on the configuration space of the dynamical
system. In this section we make an attempt to construct momentum, and read
information about momentum observables corresponding to the stochastic
processes investigated in stochastic mechanics. Let us consider the simple case
of a point particle of mass m, moving on the real line in a potential U.
As we have seen above, stochastic quantization consists in fact in the
suitable reinterpretation of classical configuration observables as random vari-
ables whose time evolution is to be read from the Euler~Lagrange equations of
motion, rewritten as dynamic conditions for the corresponding stochastic
process Q(t) which is described by the stochastic differential equation (see
Chapter 7):
dQ(t) = v+(Q(t), t) + dW(t), (dt > 0). (9.41)
This equation describes a random distrubance d W(t) with expectation
IE {d W(t)} = 0 and covariance IE {[ d W(t)] 2 } = 2v dt. In stochastic quantization
procedure, the link between stochastic and quantum theories means that the
stochastic process Q(t) with diffusion coefficient v = h/2m is associated with a
quantum state of a particle of mass m descfibed by a normalized wave function
tjJ(x, t) = exp[R(x, t) + is(x, t)]. At this, probability density p(x, t), current
v(x, t) and stochastic u(x, t) velocities of the process Q(t) are given by
p(x, t) = ItjJ(x, tW, (9.42)
1 as(x, t)
1
v(x, t) = z[v+(x, t) + v_(x, t)] = -
m
ax ' (9.43)

h aR(x, t)
1
u(x, t) = z[v+(x, t) - v_(x, t)] = -
m
ax ' (9.44)

or

(9.45)
Selected Topics in Stochastic Mechanics 317

Here, we recall again definitions:

v+ (x, t) = l'1m !E{Q(t + dt) - Q(t) I },


Llt ~ 0+ At Q(t) ~x

v_(x, t) = lim !E{Q(t) - Q(t - dt) I }.


Llt~O+ At Q(t)~x

The operational meaning of the conditional expectation !E{ ... } appearing in


the definition is clear: v+(x, t)[v_(x, t)] is the mean slope with which those
sample paths at time t in x leave [enter] x.
On the other hand, this interrelation may be interpreted in the following
way. With each soluton ljJ(x, t) of the SchrOdinger equation for a particle, the
stochastic quantization procedure associates a diffusion process Q(t) whose
expectation, for every configuration observable f(x) satisfies
!E{f(Q(t))} = <ljJ(x, t),/(x)ljJ(x, t). (9.46)
We now discuss the following problem: how can one read information about
momentum observables from 'measurements' on the sample paths of the
process Q(t)? The study of the problem has been recently broached by Shucker
(1978, 1980b), and de Falso, Martino and Siena (1982, 1983) (see also de la
Pefia-Auerbach and Cetto, 1972). The given section is based on their studies.
They have presented an analysis of how information concerning momentum is
encoded in the stochastic processes Q(t), and they have given the quantum-
mechanical procedure for reading the momentum distribution from the
Fourier transform of the wave function. The basic ideas are:
(i) If r is the time at which the momentum distribution is requested,
construct the free process Qf(t) 'tanget' to the process Q(t) at time r, as the
process which evolves in the absence of potential starting from the initial data
at time r determined by the probability density p(x, t) and the mean forward
velocity field v +(x, t) associated with the process Q(t) at time r.
(ii) According to stochastical mechanical correspondence, the process Qf(t)
can be studied by means of the associated solution IjJ f of the free Schrodinger
equation which starts from the initial data IjJ f(x, r) = ljJ(x, r).
(iii) For almost every sample path w(t) of the process Qf(t) the following
limit exists:

1C (r,w ) -_ l'1m m [w(r + T) - w(r)j- . (9.47)


T~C() T
The existence of this limit is the observation that the mean forward velocity
field v~ (x, t) for Qf (t) behaves for large t as
v~(x,t) = x/t + O[t- 3 / 2 1IjJf(x,t)I- 1 ]. (9.48)
318 Chapter 9

(iv) Granted the existence of the limit (9.47), the fact that

( T)1/Z exp{i (n"4 -"21 2mh


;~ -;:;;
pZ )} _ (T )_
T t/J f -;:;; p; r + T = tfJ(p, r),

where IJ/(p, r) is the Fourier transform of tfJ(x, r), shows that the probability
density of the random variable
n(r) == n(r, co), (9.49)
is IIJ/(p, r)l z. The random variable n(r) is, therefore, to be identified with the
image of quantum mechanical momentum in stochastic mechanical
correspondence.
We now consider two simple examples which are interesting in canonical
transformation for stochastic mechanics.
EXAMPLE 9.1. First of all we recall that what would seem the most natural
definition of momentum in stochastic mechanics to be associated with the
random motion of the particle, namely m dQ(t)jdt, is in fact incorrect because
d W = O[ (dt )1/Z] in (9.41) prevents the existence of the relevant limit. Let us set
p_{t) = mv_(Q(t), t).
Then it is easy to see that multiplying both sides of (9.45) by mp and
integrating over x, we obtain
(9.50)
Thus, the two random variables P±(t) have the same expectation. Similarly,
multiplying both sides of (9.45) by mpx, integrating over x, and using (9.50), we
obtain

Schwarz's inequality then implies


.1Q . .1 (jp » mv, (9.52)
where .1Q and .1 (jp are the root mean square deviations of the random
variables Q(t) and (jp(t) = [p+(t) - p_(t)]j2, respectively. Taking into account
(9.44), we see that inequality (9.52) implies that the root mean square de-
viations of the random variables Q(t) and u(Q(t), t) satisfy
.1Q . .1(mu) ~ hj2. (9.53)

In order to connect this inequality with the position-momentum uncertainty


relations, we observe that, because of (9.42), we have in fact
.1Q = {1E[(Q(t) - IE(Q(t))f]}1/Z = [< tfJ(t), xZtfJ(t) - <tfJ(t), xtfJ(t) )Zr/2.
Further, we see that the relevant expectation of the momentum operator
Selected Topics in Stochastic Mechanics 319

p = (II/i) %x in
t1p = [< I/J(t), p21/J(t) - <I/J(t), pl/J(t) )2J 1/2,
can be easily read from the mean stochastic velocities of the process Q(t). We
have, indeed,

lE{p+(t)} f oS(x, t)
= lE{p-(t)} = dxp(x, t)~ = <1/J(t),pl/J(t),

f
and, similarly
2 2 oS(x, t) oR(x, t)
lE{p+(t)}=<I/J,pl/J)+211 dxp(x,t)~ ox .

Namely,
lE{p~(t) + p~(t)}/2 = <I/J(t), p21/J(t).
From these equalities, we have
t1p = [< I/J(t), p21/J(t)- <I/J(t), pl/J(t) )2r/2
= 2-1/2{IE[p~(t)J - [1E(p+(t))]2 + lE[p~(t)] - [1E(p_(t))yp/2. (9.54)
Equation (9.54) is compared with
/
m t1u -_ { IE [(p+(t) -2 p_(t))2]}1 2.

On the other hand, it is easy to verify that the following equality holds

(t1p)2 - (t1muf = 1E{[p+(t) ; P-(t)J} - lE{p+(t)}lE{p_{t)}

= fdxp(x, t)[oS(x, t)/OXJ2 - {fdxp(x, t) oS(x, t)/ox} 2.


From this, we can conclude that
(9.55)
In particular, t1p ~ t1mu. Thus, the Heisenberg position-momentum uncertainty
relations
[< I/J(t), x 21/J(t) - <I/J(t), xl/J(t) )2J 1/2 X
X[< I/J(t), p21/J(t) - <I/J(t), pl/J(t) )2J1 /2 ~ til (9.56)
can be traced back to the position-stochastic velocity uncertainty relation
320 Chapter 9

Finally, it should be noted that in the decomposition (9.55) of the root mean
square deviation of the quantum mechanical momentum there is a part due to
the current velocity and a part due to the stochastic velocity; it is just the
stochastic term that forces the position-momentum uncertainty. This problem
was first considered by de la Peiia-Auerbach and Cetto (1972).
EXAMPLE 9.2. To estimate the rate of convergence in (9.47) we consider an
explicit example due to Falso, Martino, and Siena (1983).
Let the wave function at time " = 0 be
ljJ(x, 0) = (2na) -1/2 exp( - XZ/4a), (9.58)
where a = (~xf. The mean forward velocity field for the process QAt) evolving
from the initial condition corresponding to equation (9.58) is

v~(x, t) = ~{R{ rjJ i 1 (x, t) a: ljJ f(X' ;)] + Im[:x In ljJ f(x, t)]}

= (hZt - 2mha)x/(h2 t 2 + 4a 2m2).

The process Qf(t) evolves therefore according to the stochastic differential


equation
(9.59)
where dt > 0, and d W represents increments of Brownian motion with dif-
fusion coefficient v = h/2m.
From equation (9.59) it follows that the two-point function of the process
Qf(t) satisfies for t > 0 the differential equation
d h2t - 2mha
dt IE{Qf(O)Qf(t)} = h2tZ + 4a2m2 IE{Qf(O)Qf(t)},

whose explicit solution, with the initial condition IE{ Q;(O)} = a is


IE{Qf(O)Qf(t)} = a[l + h2t Z/4a 2mZr/2 exp{ -arc tg(ht/2ma)}. (9.60)
Defining
(~nt)2 == 1E{[m(Qf(t) - Qf(0»/t]2},
from equation (9.60), we obtain
~x '~nt = th{l + 2(A/tf[1 - (1 + t 2/A2)1/2 exp( -arc tg(t/A))]} 1 / 2, (9.61)
where A = 2m(~x)2 /h. The main features of equation (9.61) are the following:
(a) As t -> + 00, ~x· ~nt goes to the quantum-mechanical uncertainty pro-
duct ~x· ~p (which in the particular case at hand is h/2);
(b) As t->O+, ~x'~nt diverges as t- 1;
(c) The limit (a) is reached from below;
(d) After a time of order A = 2m(~x)2 /h, ~x· ~n f is very close ~X' ~p (within
a few percent).
Selected Topics in Stochastic Mechanics 321

As to the generality of the intriguing point (c), we observe that a necessary


and sufficient condition for I1n t to reach its limit I1p from below is that for
every t large enough

(9.62)

Condition (9.62) lends itself t'o the suggestive interpretation that, at least in
some cases, such as the explicit example discussed above, part of the un-
certainty in momentum builds up as a cumulative effect of small fluctuations
over large times.

9.6. Appendix C. Concept of the 'Differential Form' and Directional


Derivative

An important question in relativistic dynamics is associated with the evolution


of the system on the transition from one hyperplane (J 1 to another (J 2' We
suppose the hyperplane (J is characterized by a single time-like vector
v = (va' v). The hyperplane (J is orthogonal to the world line f?f(s) of the
particle, and the four-velocity ull = df?f/ds = dxllids is the unit vector tangent
to the world line of the particle.
Hitherto, we have considered the evolution of the system from the point of
view of its proper time (see Chapter 8). For example, by means of this concept
we have derived the probability density p(x ll , s + I1s) for the 'time's + As from
the probability density p(xll's) for the earlier 'time's, and we have also
obtained a differential equation for P(XIl' s).
Now we attempt to derive an equation for p(x/ s) from a different point of
l ,

view, namely, on the basis of a geometrical object - a 'differential form' or '1-


form' (for more detail, see Misner et al., 1973); then the evolution of the
probability density P is determined by means of an oriented family of flat
surfaces k = dp. Using only surfaces, it is impossible to characterize fully the 1-
form k; it is also necessary to specify an orientation. We suppose k is
composed of the surfaces of the constant values, say, p = 7,8,9, ... ; then the
direction from surface to surface is assumed to be 'positive' if p increases in this
direction.
We now define the concept of the derivative along the direction of the vector
v = (vo,v). We take a certain vector v, construct the curve f?f(A), which is
determined by the equation f?f(A) - (l}o = AV, and differentiate the function p
along this curve:
o,p = (d/dA)I). = a p({l}(A)) = (dp/dA)1 9o '
The 'differential operator' 0, = (d/dA) for A = oalong the curve f?f(A) - {l}o = AV,
which carries out this differentiation, is called the 'operator of the derivative
along the direction of the vector v'. The derivative along the direction o,p and
322 Chapter 9

the gradient dp are intimately related. This is immediately seen by applying av


to the equation p(£!J» = p([!Po) + <dp, [!P - [!Po> + nonlinear terms, which de-
termines the 'gradient', and calculating the result at the [!Po, which gives
avp = <dp, d[!P/dA >= <dp'v >.
In terms of Misner, Thorne, and Wheeler (1973), this result reads: dp is a linear
machine for calculating the rate of change of p along an arbitrary vector v.
Introducing v into dp, we obtain at the output a.p, the 'number of intersected
planes', a number which for sufficiently small v is simply equal to the
increment of p between the base and tip of the vector v.
In the co-ordinate representation, the operator a, of the derivative along the
direction of the vector v is determined by the expression a, = vll a/axil. The
notation and concepts are taken from Misner et al. (1973).
Thus, the method based on the use of the concept of the derivative along a
direction warrants attention. For example, when this method is used equations
(8.20) and (8.21) simplify significantly by virtue of the elimination of the
dependence on u~ in the expression for P ±' which now takes the form
P ± = (4nv± AS±)-2 exp[ -(AYE)2/4v± ·As±]; (v+ = v_ = v).
Here, s + and s _ are the parameters which characterize the derivative along the
directions of the vectors u~ and u~, respectively. In this case, equations (8.22)
take the form

au~/as+ = 2u"a"u~ + vDu~ + F~ /m;


aU~/aL = -2u" a"u~ - vOu~ + F~/m;
au~/as+ = 2u"a"u~ + vOu~ + F'~/m;
aU~/aL = -2u" aAu~ - vDu~ + F'Vm.
By definition, the derivatives along the directions of u~ and u~ are
ll /a s+ -
au± - "
u+ all. Il /a s_ -
"u±' au± "a "u±'
- u_ Il

and we therefore obtain equations (8.23) and (8.24) with


Dc = VAa A; Ds = uAa" + vO·
Equations of form (8.23) with A = 1 and Dc = v" a" were obtained for the first
time by Lehr and Park (1977) (see also de la Pefia-Auerbach, 1971); they are,
of course, equivalent to equations (8.23) with Dc = a/as + vA a" and
Ds=u"a,,+vD.
Thus, the derivative with respect to s which occurs in this appendix can be
understood as the derivative along the direction of some arbitrary vector v. In
the special case when the particle's velocity ull is taken as the vector vll , we can
interpret s as the proper time of the particle. In our approach, conservation of
mass (of the probability density multiplied by the volume) signifies the absence
of mass loss through any hyperplane characterized by the vector u ll; it follows
Selected Topics in Stochastic Mechanics 323

from this in particular that the probability density for finding the particle
along the world trajectory is constant, i.e.,
op/os = (dp· u) = (op/oxo)U O + (1 - [32) -1/2 V . (op/ox) = O. (9.63)
This last equation expresses the conservation of the mass (the probability
density of the current) of the particle. Note that equation (9.63) recalls the
equation that reflects the absence of heat transfer and requires constancy of the
specific entropy (the entropy of unit mass) for a fluid particle (see Misner et al.,
1973):
d,u/dt = 0 or o,u/ot + (vV),u = 0,
and the last equation expresses the conservation of mass:
op/Ot + V(pv) = O.
We now obtain the continuity equation for the probability density of the
current in our case. We see from (8.18) that the condition (9.63) leads to the
following equations:
oJl(pu~) = vOp and oJl(pu~) = -vOp.

Further, we determine the vectors of the four-dimensional current:


JJl == ¥J~ + J~);
and then JIl satisfies the continuity equation
0I,JJl = O. (9.64)
Since vJlvJl = c2 , Ipl2 = JJlJJl/c 2 is a world scalar and, therefore, the equations
oJlJ~ = vOlpi and oJlJ~ = -VOlpi,
are already expressed in covariant form.
Chapter 10

Further Developments In Stochastic


Quantization

10.1. Introduction

The stochastic quantization programme of classical canonical systems, pro-


posed by Fenyes (1952), Nelson (1966, 1967) and their followers Cufaro
Petroni and Vigier (1979); Vigier (1979, 1982); Yasue (1979); de la Peiia-
Auerbach and Cetto (1975); Davidson (1979-82); Guerra and Ruggiero (1973),
etc. has been enriched by new methods in recent years. Th~se methods (see
Chapters 8 and 9; Baulieu and Zwanziger, 1981; Zwanziger, 1981; Parisi and
Wu, 1981; Niemi and Wijewardhana, 1982; Mc'Clain et aI., 1982, 1983; etc.)
rely on ideas of stochastic space-time (or vacuum fluctuation), and of the
stochasticity of gauge fields, and the use of new techniques and concepts. A
central role in these ideas is played by the fact that there exist stochastic
solutions of a simple variant of the Yang-Mills theory, obtained by Matinyan
et al. (1981a, b); Chirikov and Shepelyansky (1981).
We recall that the Nelson stochastic quantization procedure has already
been generalized and used for the description of continuous systems and fields,
in particular, in the case of the Euclidean-Markov scalar field (Guerra and
Ruggiero, 1973; Davidson, 1980 and Moore, 1980), and of the electromagnetic
(Guerra and Loffredo, 1980; Davidson, 1981), and gravitational (Davidson,
1982) fields. Moreover, the properties of physical vacuum in particular, the
mechanism of the vacuum tunneling effect can be illustrated within the realm
of the stochastic quantization (Yasue, 1979).
Recently, the relation between Euclidean quantum field theory and stochas-
tic differential equations has attracted considerable attention with the develop-
ment of new methods of stochastic quantization of any field theory, including
gauge theories. In particular, they offer a new method of quantization, which
sometimes presents conceptual advantages. For example, by considering a
stochastic equation in real space, one can show that a supersymmetry and
supers pace naturally arise (Parisi and Sourlas, 1979, 1982; Mc'Clain et aI.,
1983). In a more traditional approach (see Section 10.4 and references therein),
one may introduce a fifth dimension in the form of a new 'time' r and consider
a stochastic equation in this variable (Parisi and Wu, 1981).

324
Further Developments in Stochastic Quantization 325

The stochastic quantization method proposed by Parisi and Wu is to use


the solution of the Langevin equation to evaluate the correlation functions
according to the following prescription:
(i) Let 5i'[A(x)] (x = X 4 , x) be the Lagrangian density describing the dy-
namics of a physical system. 5i' is a function of a set of fields in d-dimen-
sional Euclidean space, which we denote collectively by A(x) (internal indices
are omitted here). In the Euclidean field theory (see Section 7.6) the Green
functions are the vacuum expectation values of products of fields and are
given by

(10.1)

where SeA] is the action of the system.


(ii) Introduce an additional dimension and consider fields which depend on
d + 1 variables A~(x, r). The evolution in this extra 'time' r is assumed to be
governed by a Langevin equation:

(10.2)

with the initial condition A~(x, 0) = O. It is shown (Floratos and Iliopoulos,


1983) that this condition can be replaced by a more general one
A~(x, 0) = C(x), where C(x) is an arbitrary function. Moreover, one can show
that the large time (r ---> 00) limits are independent of the choice of C(x). In
equation (10.2) IJ(x, r) is a random source with Gaussian distribution, namely
the random average over IJ has the following Wick decomposition property:

<IJ(x1,r1)IJ(x2,r2)"'IJ(X2n,r2n»~ = L TI <IJ(xi,r;)IJ(xj,rj )\,


possible pair
comb.

<IJ(x, r)IJ(x', r'»~ = 2 l5(d)(X - x')l5(r - r'),

i.e., IJ(x, r) is a white, Gaussian noise.


(iii) Correlation functions are now introduced as averages over IJ:

<A~(X1' r 1 )'" A~(xn' r n) \

= fl5IJex{ -~fddxdr'IJ2(X,r)J x A/x1,r1)"'A~(xn,rn)' (10.3)

A convenient way to rewrite (10.3) is to introduce the probability density


326 Chapter 10

peA, rJ which is defined by

peA, r J = f c51J exp{ -~ f ddx dp '1J2(X, P)} ry c5[A(y) - A~(y, r)]. (10.4)

In terms of P we can write the equal-time correlation function as

(10.5)

The probability density P satisfies the Fokker-Planck equation:

a peA, rJ
or =
f d c5
d x c5A(x) c5A(x)
{c5 c5S }
+ c5A(x) peA, rJ, (10.6)

with the initial condition peA, OJ = TIy c5[A(y)]. On the right-hand side of
(10.6) we assume that a suitable regularization has been employed. In parti-
cular, we can use the nonlocal method proposed in Part I. Roughly speaking,
in the given case our method of regularization leads to a smeared (nonloean
white noise:

1Js(x, r) = f ddyK(x - y)1J(y, r) or

1Js(x,r) = fdd y fdPK(X-y,r-p)1J(Y,P).

(iv) The correlation function (10.1) is then given by

<0IA(x 1 )' .. A(xn)IO) = lim <A~(xl,r)" 'A~(xn,r)\ (10.7)


,~oo

or, equivalently,

w. lim peA, rJ =peq[AJ = exp{ - S[AJ}, (10.8)


,~oo f c5A e-S[A]

i.e., at large times the probability distribution reaches the equilibrium one given
by (10.8). Here the limit is supposed to be taken 'weakly', as expressed in (10.7);
in other words, peA, rJ will always be applied to a string of fields (for details,
see Floratos and Iliopoulos, 1983).
The property (10.7) or (10.8) has been shown to hold for several systems
(Wick, 1981; Fox, 1978). It has been conjectured (Parisi and Wu, 1981) that it
is true even for gauge theories, provided one looks for gauge-invariant quan-
tities. In other words, it has been indicated that the limit (10.7) does not exist
for arbitrary gauge non-invariant Green functions, but it exists and gives the
correct result only for gauge-invariant quantities. This fact is very interesting
Further Developments in Stochastic Quantization 327

because it allows one to formulate a new quantization method which does not
require any gauge-fixing term and is therefore free from gauge ambiguities
(Gribov, 1978; Singer, 1978). In Section 10.5 we shall prove this statement
order by order in perturbation theory.
The purpose of this chapter is to report briefly on the progress in the
stochastic interpretation of quantum field theory. These results are based on
the studies (Guerra and Ruggiero, 1973; Davidson, 1979-81; Yasue, 1979;
Baulieu and Zwanziger, 1981; and Floratos and Illiopoulos, 1983) of a
generalization of the Nelson stochastic mechanics to systems with infinite
number of degrees of freedom and of the stochastic quantization method
introduced by Parisi and Wu (1981). In Sections 10.2 and 10.3 we present the
stochastic quantization method for scalar and electromagnetic fields in the
framework of Davidson's approach. Some ideas of stochastic quantization of
the gauge theories are given in Sections 10.4 and 10.5. The last two sections are
devoted to the study of the mechanism of vacuum tunneling phenomena and
stochastic fluctuations of the classical Yang-Mills fields, respectively.

10.2. Davidson's Model for Free Scalar Field Theory

In Section 7.5 it is shown that Euclidean field theory results from the
stochastic quantization of the free scalar field. We present here another
approach due to Davidson (1980a). This model is useful in our scheme where a
nonlocal Euclidean-Markov field is investigated (see Chapter 11). The basic
idea of Davidson's stochastic theory is the following: Let cp(x, t) be the free
scalar field with mass m. Then, consider the classical equation for the real free
scalar field in the Minkowski space:
PIlPllcp = m2 cp, PIl = -iha/axll'
where c has been set to unite and gOO = 1. Let us require periodic boundary
conditions on the field
cp(x, t) = cp(x + a, t), ai = niL,
ni is an integer, so that one may write

cp(x, t) = (hL 3 )-1/2 I exp(iqx) CPq(t), qi = 2nmJL, (10.9)


q

(mi is an integer).
Further, as usual, it is assumed that each component of the Fourier
decomposition (10.9) is a random variable satisfying the stochastic differential
equation
(10.10)
where bq is a smooth function of the type of a velocity field (see Chapter 7) and
328 Chapter 10

is determined by the ground state probability density Pq (<pq (t)) for <Pq(t)
(Davidson, 1980a). In equation (10.10) ~(t) is a Wiener process satisfying
lE{d~(t) d~l(t)} = 4v c)q,_ql dt,
here IE denotes the conditional expectation value with respect to the random
variable <Pq(t), and v is the diffusion coefficient. For illustration we write a
solution to simple stochastic equation of the following type (see Nelson, 1967;
Kac, 1959):

dQ(t) = - wQ(t) dt + d Wet), Q(t) = e- rot fro dW(s) eros. (10.11)

For the Wiener process d W(s) == yes) we have then,

= lim IE{Y(SdY(S2); -£ < y(P) < £}


Df HO P{ -£ < y(fJ) < £}
ro f.

J[2n(fJ _ s )]
- 00 -e

x ex [- (Yz - yd -112 exp[ - (Y3 - Yz)2/2(fJ - S2)]


p 2(sz-sd 2 (2nfJ)-1 /2 x

x exp { - [ -Y12
2s 1
+ ( Y2-Y1 )2 +
2(sz - S1 )
Yz2
2( fJ - S2 )
J}

Here s = 2vt, v = t. Assuming Sl = Sz = 2tv and fJ = 2v(t + dt) we obtain


lE{dW(t) dW(t)} = 2vt[1 - (1 - dt/t)] = 2v dt, lE{dW(t)} = O.
From the solution (10.11), we have
IE{Q(t)} = 0,

IE{Q(t)Q(t')} = lE{e- rot feo exp(wsddW(sdexp(-wt' ) fro eXP(WSz)dW(Sz)}


= (v/w) exp( -wit - til).
Further Developments in Stochastic Quantization 329

In the case of the Gaussian distribution, the ground state probability density
Pq(cpq(t)) and the function bq(cpq(t)) take the form
Oq(cpq(t)) ~ exp{ -(2wq/h 2)cp;(t)}, and bq = -4v(wq/h 2)cpq(t),
respectively, where
Wq = th(q2 + m2/h 2)1/2.
Thus, in the Davidson (1980a) stochastic model, the process CPq(t) is a
Gaussian stochastic process characterized by the expectations:
(10.12)
where rx = (4v/h 2 )wq .
Here, the diffusion coefficient v is an arbitrary quantity and undetermined with
respect to v = h/2 which is usually used in Nelson's stochastic mechanics. This
situation allows the diffusion parameter to be continued to imaginary values;
then the Markov expectations become the causal Green function of quantum
field theory which are Lorentz invariants. From this one can conclude that if
the diffusion parameter v is not measurable (undetermined), then the stochastic
model for the field is consistent with special relativity, although it is not
manifestly Lorentz covariant.
Now we obtain this result.
THEOREM 10.1. For non-coincident Xi (i = 1,2, ... , n) (x = Xi' tJ
hm . lE{cp(x1,td"'cp(xn,tn)} =
(2V 2v )
Sn x 1'-h t 1; ",;xn'-h tn , (10.13)
L-+ 00, A -+ 00

where A is a cutoffin q2, Sn(-) are the n-point Schwinger functions determined in
Chapter 7, in particular

Proof To show this result, it suffices to prove (10.13) for the two~point
function only because the expectations of (10.13) will satisfy the following
equation:

= L S2 (Xl' t 1; X2, t 2) ... S2 (X 2n - P t 2n - 1; X2n , t 2n ) (10.15)


"
(where the sum over n is a sum over distinct permutations of the arguments of
the S's), since cP is a Gaussian process. Using equation (10.9), we have
1 A
IE {CP(X1, t1)CP(X2' t 2)} = -h3 L exp(iqx1 + iq'x 2) IE {cpq(t 1)CPq,(t 2)}, (10.16)
L q,q'
330 Chapter 10

where a cutoff in q-space has been included in equation (10.16). Substituting


equation (10.12) into (10.16), we obtain
-3 A . fdq4 exp[iq4(t1 -t z )(2vlh)]
IE {CP(X1, tdcp(xz, tz )} =L Lexp[lq(X 1 -X Z )] -2
q n q4z + q + m zlhz .
Z

Taking the limits L --+ 00 and A --+ 00, this becomes

and the result is shown.


It is interesting to note that if v = h12, which is the value used by Guerra

and Ruggiero (1973), then the expectations of the Markov theory become
equal to the Schwinger functions (see Section 7.5). Theorem 10.1 shows that
the expectations of the Davidson model have the same analyticity properties in
v as the Euclidean field's Schwinger function does in t. It is believed that v is
not an observable and that measurable quantities are independent of v.
Therefore, one may consider continuing v into the complex plane, as has been
done by Davidson (1980a).
To illustrate the compatibility of the Davidson theory with special relativity,
we continue expectations to the point
v = ihl2. (10.17)
We denote the analytically continued expectations by lEv' One easily finds:

(10.18)

Now we compare (10.18) with the usual causal Green function of the ordinary
quantum field theory (see Part I and Bogolubov and Shirkov, 1980):
L1c(x) = <01 T {cp(x)cp(O)} 10)

= '-1 (2 )-4
I n
fd 4
q
exp(iqf1xf1)
Z Z ., (10.19)
m - q - Ie

From this we see that the two theories are related by


IEv=ih/2 {cp(x)cp(O)} = L1Ax) = <01 T {cp(x)cp(O)} 10), (10.20)
IEv= _ih/2{CP(X)CP(0)} = L1:(x) = <01 T* {cp(x)cp(O)} 10), (10.21)
where T and T* denote, as usual, time ordering (later times to the left) and
anti-time ordering, respectively.
Using the combinatorial rule of equation (10.15), which is satisfied by
quantum expectations as well as the Markov expectations and Schwinger
functions, we obtain immediately a generalization of (10.20) and (10.21) to
Further Developments in Stochastic Quantization 331

higher-order Green's functions:

IEV~ih/2{CP(xd···cp(xn)} = <OIT{cp(xd···cp(xn)}IO), (10.22)


lEv ~ -ih/2 {CP(Xl) ... CP(Xn)} = <01 T* {cp(X 1) ... cp(Xn)} 10), (10.23)

where we have stopped writing the t arguments explicitly on the left-hand side.
Since the right-hand side of equations (10.22) and (10.23) are Lorentz in-
variants, it follows that the Davidson model for the free scalar field theory is
manifestly Lorentz covariant when continued in the diffusion parameter v to
± ih/2.
It should be noted that the above results may be realized in the case of
electromagnetic fields. We now go over to this problem.

10.3. The Electromagnetic Field as a Stochastic Process

We consider stochastic quantization of the electromagnetic field and apply the


above obtained results for the scalar field to this case. Guerra and Ruggiero
(1973, 1978) and Guerra and Loffredo (1980) first analyzed the relativistic
scalar field and the free electromagnetic field using the quantization pres-
cription of Nelson. Moore (1979) has also considered the Maxwell field as
part of a general programme of the stochastic field theory. In these studies,
only the fixed value of the diffusion coefficient (v = h/2) of the original Fenyes-
Nelson theory was considered. As for the scalar field (Davidson, 1980a and in
the preceding section), the moments of the electromagnetic field (EMF) in the
ground state are equal to the Schwinger functions of the theory, but with the
times scaled by a common factor 2v/h which depends on the diffusion
coefficient v. In this section a stochastic quantization of EMF is carried out in
terms of vector potentials rather than directly in terms of electric and magnetic
fields as in Guerra and Loffredo's (1980) paper. Here we follow the approach
developed by Davidson (1981).
So, in an appropriate system of units the Maxwell equations take the form
V'E=p, v 'B=O; v x E = - oB/ot; vx B = J + oE/ot, (10.24)
where it is assumed c = 1. In the Coulomb gauge one has
V'A=O, E = -V1> - oA/ot; B = V x A; 111> = - p.

The Maxwell equations become


(0 2 / 0t 2 - L1)A = J - V o1>/ot. (10.25)
The energy density of the field in these units is given by u = -!eE2 + B2).
Quantization as in the scalar case shall be carried out in a cubic volume of
edge L with periodic boundary conditions in L. Consistency requires that J
and p have this same periodicity, but little generality is lost by this, since
332 Chapter 10

ultimately the volume will tend to infinity. Because of the periodicity require-
ment, the following decompositions are allowed

A = 2 -1/2 L - 3/2 I e(.1, k) eikx Q,l,k (t);


',k

¢ = Ie ikx p(k, t)/k 2 ; p = I eikx p(k, t); J = I eikx J (k, t),


k k k

where the components of k are restricted to be integral multiples of 2n/L and


e(.1, - k) = e(.1, k), e·k = 0, .1=1,2;
Q,l,-k = Qtk; p( -k, t) = p*(k, t); J( -k, t) = J*(k, t);
8p/8t = -ik 'J. (10.26)
It is customary to choose the transverse current by J T = J - V . 8¢ /8t, and it is
easy to show that its divergence vanishes V 'JT = 0 so that one may write

JT = 2- 1/2 L -3/2 '\' e(.1 k) eikx Z " ,(t)


L...' k'
,l,k
In this case we choose the real and imaginary parts of the Q's as the canonical
co-ordinates. Equation (10.25), then, leads to the following equations for the
Q's
(8 2 /8t 2 + k 2 )Q,l,k(t) = Zu(t)·
A suitable Hamiltonian from which these equations may be derived is

H = I a(IQ;.,kI 2 + k2IQ,l,kI2) -tRe(ZtkQ).,k)}' (10.27)


,l,k
which may also be written as

H = 1d3x[~A2 ttl + (V'AY - JT'A}


where V denotes the cubic volume of edge L.
Since the equations for the Q's (the canonical momenta are their time
derivatives) are those for forced harmonic oscillators, in principle the quanti-
zation of the field can be achieved by solving a Schrodinger equation for
uncoupled but forced harmonic oscillators. Thus, there is wave function in Q
space which satisfies Ht/J = ih (8t/J / 8t), and where the Hamiltonian operator has
the general form

H = ~ {-th2 8~f - ZiQi + tkfQf}. (10.28)

Here the Q's and Z's are real and represent the real and imaginary parts of the
Further Developments in Stochastic Quantization 333

Q's and Z's in equation (10.27). The sum in equation (10.28) is only over
independent Q's [in the sense of equation (10.26)J, which explains the factor-
of-two difference between equations (10.27) and (10.28).
So, we are ready to discuss the stochastic quantization procedure for the
electromagnetic field (EMF). According to the previous section, the stochastic
quantization of the EMF is straightforward. It is convenient to define

~(X,t)=r1/2L-3/2L:e(A,k)eikX( a + ia ).
i.,k aRe Q;,k a1m Q;,k
The forward and backward time derivatives can be expressed in terms of () as

D+ = :t - I d 3 x(E + + V¢ ) . () + v I d3 X () 2,

where E+ = -D+A - V¢, E_ = -D_A - V¢, The Maxwell equations


become
V'E+ = V'E_ = p, V·B =0, V x E+ = -D+B,
V x E_ = -D_B,
t[D+E_ + D_E+J + tf3(D+ - D_)(E+ - E_) = V x B - J, (10.29)
With the help of the following stochastic field,

W + (x, t) = 2 -1/2 L - 312 L: e(A, k) eikx W;~k (t), (10.30)

the stochastic differential equation can be written in the compact and suggestive
form
dA= -(E+ +V¢)dt+dW+, dA= -(E_ +V¢)dt+dW_, (10.31)
where the second equation of (10.31) is the 'backward' version of the stochastic
differential equations and W _ is another field of the form (10.30), but with
different Wiener processes.
In the special case v = th and when there are no charges or currents present,
equations (10.29) become precisely those derived by Guerra and Loffredo
(1980). It is interesting to note, as is pointed out by Guerra and Loffredo, that
the electric field does not exist as a well-defined stochastic process since the
time derivatives of A do not exist. Despite this fact, the two functions E+ and
E_ may be thought of as a stochastic replacement for the electric field.
For the free field, the scalar potential vanishes in the Coulomb gauge. When
the Schrodinger equation is solved in the ground state and the Markov
process for the field examined, it is found that the vector potential is a
334 Chapter 10

Gaussian Markov field with the following covariance:

dk)4(b .. - k.k.jk 2)
IE{Ai(x)Aj(y)}=h f( 2n 'Jk~~~2
[ ( ik02V)
exp ik(x-y)+ -h- IXo-Yol
J
(xo = t x , Yo = ty), (10.32)

where the infinite volume limit has been taken. The covariance for the
magnetic field is

dk)4 (b .. P - k.k.) [
IE{Bi(x)B/y)} = h f( 2n 'Jk~ + k~ J exp ik(x - y) +

ik 02V)
+ ( -h- Ixo - Yol J (10.33)

As has been stated before, the electric field does not, strictly speaking, exist.
However, because the following expressions are equal

(10.34)

it is natural to study the properties of the left-hand side of (10.34) and identify
these with an electric field. From this we have

a a
-;- -;-1E{Ai(X)Aj(Y)}
uXo uYo
=2v b(xo - yo)b;!(x - y) + (2v/WIE{Bi(x)B/y)}, (10.35)

where b (t) is the Dirac delta function and b:!(t) is the transverse delta function
defined by

15\ (x) = (2n) - 3 f d 3 k(bij - ki k j /k 2) eikx . (10.36)

Finally, terms with both electric and magnetic fields may be evaluated. The
result is

(10.37)

x fd 4 k k1exp[ik(x - y) + (ik o 2v/h)lx o - Yol],

where Sijl is the totally antisymmetric tensor. When both fields are at the same
Further Developments in Stochastic Quantization 335

point in space, one obtains the results


IE{B;(xo)Bj(yo}} = h(bjn2)(h/2v)4I xo - Yol-4,
IE {Ei± (xo)Ei± (Yo)} = h(bi)n2)(h/2v)2Ixo - Yol-4, (10.38)
IE{E±(xo)Bj(yo)} = O.
High-order correlations may be calculated from those above by means of
the well-known combinatoric rule for Gaussian processes. For example,
IE{A il (xd· .. A i2 JX 2m )} = L IE{A il (Xl )A i2 (x 2)} ... x
7[

(10.39)
where the sum-is over all distinct permutations and all odd correlations vanish.
As in the scalar field case (see Section 10.2) the Green functions of the usual
quantum field theory are obtained by analytically continuing the diffusion
coefficient v to imaginary values, as has been done by Davidson (1980a). The
analytically-continued expectations are denoted by a subscript. The following
results are found when the usual quantization is carried out in the Coulomb
gauge:
lEv = ih/2 {Ai(x)Aj(y)} = <01 T {AJx)Aj(y)} 10),
lEv = -ih/2{,ii~X)A/y)} = <01 T*{Ai(x)A/y)} 10).
The two-point function - Green function - for the quantum theory is given by

<01 T{Ai(X)A.(y)} 10) = ihf(dk)4 bij - k i k j (k 2 e-ik(x-y), (10.40)


J 2n k/Lk/L + IE
where k/Lk/L = k~ - k 2 , Xo > Yo, the metric convention gOO = 1 is used and E is
positive and is taken to zero after the ko integration is performed (for details,
see Chapters 2-4). The other time ordering is obtained by taking the complex
conjugate of equation (10.40). According to the Gaussian combinatories of the
quantum and stochastic theory, we may write in general
lEv = ih/2 {Ail (xd··· AiJX n)} = <01 T{A il (Xl)··· AiJXn)} 10),
lEv = - ih/2 {Ail (Xl) ... Ai. (X n )} = <01 T* {Ail (Xl)· .. AiJx n)} 10).
The conclusion to be drawn is that the stochastic theory is physically
equivalent to the Lorentz covariant quantum field theory, even though the
stochastic theory is not itself manifestly Lorentz covariant because the expec-
tations in equations (10.32}--(10.39) do not transform in a covariant way.
We notice that the generalized Fenyes-Nelson model has been applied by
Davidson (1982) to the weak field approximation of Einstein's general theory
of relativity, the so-called linearized gravitational field (Einstein, 1979; Misner
et al., 1973; Weinberg, 1972). We have not considered this problem here. The
stochastic quantization of supergravity has been discussed by Hori (1983).
336 Chapter 10

10.4. Stochastic Quantization of the Gauge Theories

10.4.1. Introduction
Interest in the stochastic regularization and stochastic quantization of gauge
theories has grown in recent years. The stochastic methods have already been
applied to gauge theories. Such approaches are based on the fact that gauge
theories possess stochastic properties. For example, it has been discovered that
simple variants of the Yang-Mills theory have stochastic solutions (Matinyan et
ai., 1981a, b). There have appeared a few stochastic models for gauge theories.
For example, Parisi and Wu (1981) proposed a method of stochastic quanti-
zation of gauge fields whereby Euclidean expectation values are obtained by
relaxation to equilibrium of a stochastic process depending on an artificial fifth
time parameter (for further developments see Nakano, 1983; Alfaro and
Sakita, 1983; Guha and Lee, 1982; Japan group, 1982). Within this approach
Zwanziger (1981) has obtained the equilibrium distribution directly without
reference to the artificial time, by a stationary condition which is an eigenfun-
ction equation in the Euclidean Hilbert space. Zwanziger's model allows one
to carry out a covariant quantization of gauge fields without Gribov' ambiguity
(Gribov, 1977, 1978). On the basis of this model Baulieu and Zwanziger (1981)
have proved the equivalence of stochastic quantization to the Faddeev-Popov
(1967) ansatz for covariant and axial gauges. Equivalence of stochastic and
canonical quantization in perturbation theory has been demonstrated by
Floratos and Iliopoulos (1983) (see Section 10.5).
Niemi and his coworkers (1982) have analyzed the equivalence between a
scalar quantum field theory (also gauge theories) in d-dimensions and its
classical counterpart in d + 2 dimensions which is coupled to an external
random source with Gaussian correlations. The main idea of this approach is
the following: A classical six-dimensional Euclidean scalar field <p(x) coupled
to an external source 1J(x) is considered. The Lagrangian of this system is given
by

The classical equation of motion


(5A[<pJ/(5(p(x) = -1J(x) (10.41)
is solved perturbatively. Here A is the classical action. This solution of
equation (10.41) is denoted by <p~ (x). It is assumed that the random sources
1J(x) have Gaussian correlations
IE {1J(x)} = 0,

1E{1J(x)1J(y)} = f(51J 1J(x)1J(Y) exp [ -~} f d 6 X1J2 (X)] = p.(5(6)(X - y),


Further Developments in Stochastic Quantization 337

(10.42)

In the momentum space (10.42) reads


lE{ry(k 1 )ry(k 2)} = (2n)6 p .(j(6)(k 1 + k2).
The parameter p defining the correlation strength has dimensions m 2 where m
fixes the mass scale.
This stochastic quantization scheme was started by Aharony, Imry and Ma
(1976). In this scheme the functional of the form

is investigated. Here x is four-dimensional and y is two-dimensional, and the


source J(x, y) of the form J(x, y) = (11m) (j(2)(y)j(x) is introduced (where j(x) is
arbitrary). It is proved that the four-dimensional n-point functions

IE{ cp(xd··· cp(x n)} = I(j'1m-ncp~(Xl'O) ... cp~(xn,O) X

x ex p { - 21p IId 4 xd 2 Yf,2(X,y)} (10.44)

generated by the expression on the right-hand side of equation (10.43), are


equal to the full n-point Green functions in the corresponding four-
dimensional quantum theory, order by order in perturbation theory (see
Mc'Clain et al., 1982; Niemi and Wijewardhana, 1982).
We briefly review the method of stochastic quantization due to Parisi and
Wu (1981) and Zwanziger (1981) in this section, discuss the equivalence of
stochastic and canonical quantization in perturbation theory, in the case of
gauge theories, and the Euclidean path integral description of the vacuum
tunneling phenomena from the point of view of the probability theory (Yasue,
1979) in Sections 10.5 and 10.6, respectively. We conclude with some remarks
concerning stochastic solution of Yang-Mills-type theory in Section 10.7.

10.4.2. Another Stochastic Quantization Scheme


Recently, Parisi and Wu (1981) have suggested that the Euclidean expectation
value lE{cp} of a functional 4>(A) of the Euclidean gauge field A~(x) be defined
as the limit to which the mean relaxes

lE{cp} I
= }~~ lE{cp}, = }~n;, (jA4>(A)P[A, ,],

where the probability distribution peA, ,] depending on an artificial fifth time


338 Chapter 10

parameter, is determined by a stochastic process, namely

. f
peA, r] =
4 b {b
d x bA~(x) bA~(x)
bS[A]}
+ bA~(x) peA, r],

~f
SeA] = d4 x I'~a (F~Y,
(10.45)

for some initial data

P[A, 0] = Po [A] :>0,

Here P = aplar, a is a colour index, J1 = 1, ... ,4 is a Lorentz index, and x is


an Euclidean four-vector. Our consideration is based on the approach de-
veloped by Baulieu and Zwanziger (1981). As mentioned above it will fre-
quently be convenient to replace A~(x) by discrete variables Qi' i = 1, ... , 00
according to

A~(x) = I QiGrl.(X),
i

where Grll (x) are a complete orthonormal set. Thus, an arbitrary functional
¢[A], for example

¢[A] = A~(x)Ae(y) or ¢[A] = tr P exp i f


AaAa(x) dx

will be written ¢[A] = ¢(Q). The stochastic method (Zwanziger, 1981) in-
troduces an artificial fifth time r and promotes ¢(Q) to a stochastic variable
¢(Q, r) governed by the diffusion equation with a drift force f(Q)

¢(Q, r) = - LQ ¢(Q, r), (10.46)


LQ = -ell + f·V] (10.47)
supplemented by the initial condition
¢(Q, 0) = ¢(Q). (10.48)
Here

and the drift force fJQ) is obtained from the classical action fJQ) = - as(Q)!
aQi' which for a gauge theory is the Yang-Mills action (10.45). Stochastic
quantization is completed by the ansatz for calculating Euclidean expectation
Further Developments in Stochastic Quantization 339

values

<</J) = lim </J(Q, r) (10.49)

which is equivalent to the proposal of Parisi and Wu (1981). Clearly, for this
formula to make sense, the limit must exist and be independent of Q. It is
shown (Baulieu and Zwanziger, 1981) that in a non-gauge theory it is reduced
to the standard functional integral formula

<</J) = SdQ</J(Q) exp[ - SeQ)]


(10.50)
J<5Qexp[ -SeQ)] ,

and that for a gauge theory it agrees with the Faddeev-Popov ansatz for
gauge-invariant </J.
As a trivial example, let Qi be the discrete Fourier components of A in a box
quantization, let SeA] be the free field action

SeA] = ~ L, (k? + m2 )Q?,


and take </J(Q) to be the two-point function </J(Q) = QiQj' One may easily
verify that the corresponding stochastic equation

. _ ,,{ jJ2 2
</J(Q,r) - ~ 8Q? - (k i +m
2 8}
)Qi 8Qi </J(Q,r),

with the initial condition </J(Q, 0) = Qi Qj has the solution


b ..
</J(Q, r) = k2 'J 2 [1 - exp( - 2(kJ + m2)r)] +
i+ m

One sees that limH 00 ¢(Q, r) = bijl(k? + m2 ), so the stochastic prescription


does give the familiar free field propagator <Qi Qj)'
Now we consider the case of a guage theory. The infinitesimal gauge
transformation bAofJ = DObwb
fJ'
where DOC = 8 bOC +jObC AbIt' (here the coupling
It!1
constant g has been absorbed in the structure constant jobc of the gauge
group), induces the change in ¢ given by

bA.[A]
'I"
= -fd4X wODOb_b_A.[A]
!l bA~(x) 'I" ,

so gauge-invariant functions are characterized by GO(x)</J[A] = 0, where


GO(x) = - D~b b/bA~(x) satisfies the commutation relations of the local gauge
group,
[GO (x), Gb(y)] = b(4)(x - y)rbcGC(x).
340 Chapter 10

With L, equation (10.47), written in the functional notation

L - -fd4x{_b- _ bScl[A]}_b_
- bA:(x) bA~(x) bA:(x) ,

it is easily verified that L is strongly gauge invariant, namely [Ga(x), L] = O.


Consequently, for gauge invariant ¢, the formal solution of equations (10.46)
and (10.48), namely,
¢[A, ,] = exp( - Lr)¢[A],
may also be written
¢[A, ,] = exp( - Lv ,)¢[A], (10.51)
where

Lv = L +f d 4 xva(x)G a(x) = L +f d 4 x(D ll v)a(x) bA~(X)" (10.52)

Here va(x) = va(x, A) is an arbitrary functional of A which, in the full enjoy-


ment of gauge freedom, we may specify at our pleasure. Reverting to the
discrete notation, we see that Lv is again of the diffusion form Lv =
- (1\ + fv' V) with a drift force fv(Q) that depends on v determined from

I !v,i(Q)Gfll(x) = -bScl [A]/bA~(x) - (Dllv)a(x, A). (10.53)


i

Let P(Q2' Q1' ,) be the kernel of the time-translation operator (10.51)

¢(Q1,')= fdQ2¢(Q2)P(Q2,Q1")' (10.54)

According to equation (10.46), it satisfies


?(Q2' Q1' ,) = -Lv,QI P(Q2, Qu c),
which is sometimes called the backward diffusion equation. From the com-
position law

f dQ2 P (Q3, Q2' '3 - '7)P(Q2' Q1' '2 - '1) = P(Q3' Q1, '3 - '1)'

the forward equation follows:


?(Q2' Q1' ,) = -L:,Q2 P (Q2' Q1' c),
where
L:,Q= -[1\-V·fv ] , L:,Qg= -[1\g-V(fg)]

and one has P(Q2' Q1' 0) = b(Q2 - Q1)' The form of the last three equations
allows one to interpret P(Q2' Q1' ,) as the probability that a Brownian motion
Further Developments in Stochastic Quantization 341

particle subject to a drift force fv will arrive at Q2 at time r if it was originally


at Ql'
The relevant theorem on stochastic processes has been proved only for a
finite number of degrees of freedom, and we suppose that our equations may
be regularized by keeping only a finite number N of modes Qi' i = 1, ... , N. In
this case the following theorem applies.
THEOREM 10.2. Given a drift force f (Q), if there exists a smooth positive
function, call it P eq (Q), which satisfies

(10.55)

where

L*Peq(Q) = -~Peq(Q) + V'(f(Q)Peq(Q)),


then Peq(Q) is unique and, moreover P(Q, Ql, r) relaxes to it:

lim P(Q, Ql> r) = Peq(Q)· (10.56)


T~ co

We call Peq (Q) the equilibrium distribution. More precisely, this theorem can be
reformulated as follows: let L * P eq (Q) = 0, where P eq (Q) ;;:. 0 is smooth and

f dQPeq (Q) = 1.

Let K be any compact set. Then

Sup fdQ1P(Q, Ql, r) - Peq(Q)I-+ 0, as r -+ 00.


QJEK

The existence of P(Q, Ql' r) is assumed by Varadhan (1980).


Thus, if there is a dist.~bution satisfying (10.55) then, according to equation
(10.54) the stochastic prescription (10.49) reduces to

<¢) = f dQ ¢(Q)Peq(Q)· (10.57)

In a non-gauge theory (v = 0) one has f(Q) = - VS(Q), and it is easy to verify


that
P _ exp[ - S(Q)]
eq (Q) - PQ exp[ -S(Q)J'

(suitable regularization is needed) satisfies conditions (10.55) and so, in virtue


of the theorem, the stochastic prescription (10.49) coincides with the standard
formula (10.50).
The important point for a gauge theory is that the above theorem holds
342 Chapter 10

even if f(Q) is not derivable from a potential, and consequently formula (10.57)
may be used, provided that for some va(Q) there exists Peq(Q) satisfying
conditions (10.55) [with L = Lv and f(Q) = fv(Q) given by equations (10.52) or
(10.53)]. Equivalence to the F addeev--Popov ansatz follows because the
equation
(10.58)
or

L* Peq[A] = - f d4xD~c c5A~(X) {va(x, A)Peq[A]},


does, in fact, hold for

v" (x, A) = - f oGab


d 4y -o~(x, y; A)Je(y, A),
Yv
with Peq [A] the F addeev--Popov distribution

Peq [AJ = PFP [A] = N det(o· D) exp{ - Sci - (21X)-1 fd 4 x(o· A)2 }.

Here Gab(x, y; A) is the Green function or ghost field propagator defined by


-oD~CGCb(X,y; A) = -(D. Oy)bcGac(x, y; A) = c5(4)(X - y) c5 ab ,
and the effective induced current Je(y, A) is given by
Je(y, A) = DtcF~v(Y) + IX -1 Ov o· Ab(y) + f bcd o(z)v GCd(y, z; A) Iz = y.
It is possible to verify equation (10.58) directly using properties of de-
terminants or the usual ghost and antighost fields (see Baulieu and Zwanziger,
1981).

10.S. Equivalence of Stochastic and Canonical Quantization in


Perturbation Theory in the Case of Gauge Theories
In comparison with field theories without gauge in variance, in theories
possessing a local invariance at the classical level, the proof of property (10.7)
presents some difficulties since, already for the free field, there is no limiting
distribution Po [A, CXl]. It is hoped that the limit exists only for gauge-invariant
quantities. Let F[A] be a functional of the gauge potential A~(x). The average
of F is determined by

<F)~ = f c5A· F[A]P[A, rJ, (10.59)

where P is again the probability distribution. It satisfies the Fokker-Planck


equation, (10.6), with SEA] the gauge-invariant classical action. The claim is
Further Developments in Stochastic Quantization 343

that, although peA, r] has no limit when r -> 00, the average (10.59) does
possess such a limit for gauge-invariant functional F[A] which, furthermore,
equals the result obtained by the usual method. We shall prove this statement,
order by order in perturbation theory, following the method of Zwanziger
(1981); Baulieu and Zwanziger (1981); Floratos and Iliopoulos (1983) (see
preceding section). The very elegant proof has been done by Floratos and
Iliopoulos (1983) and is based on a lemma which in the case of field theory
without gauge invariance can be formulated as follows.
Let us consider a self-interacting scalar field <p(x):
5£ = 5£0 + g5£in, 5£0 = -t(O,,<p)2 - tm 2<p2. (10.60)
In perturbation theory the Green functions are computed as power series in g.
Similarly, we expand the probability density P[<p, rJ:

P[<p, r] = I l Pk[<p, rJ.


k=O
The n-point equal time correlation function in the kth order of perturbation
theory is given by

<<P~(Xl' r)··· <p~(xn' r)\ = gk f


b<p·<P(Xl)··· <p(xn)Pk[<p, r], (10.61)

and is computed as the sum of the corresponding regularized Feynman


diagrams. Regularization may be carried out by means of the non local method
introduced in Part I. This is the meaning we attach to Pk[<p, r] and all our
arguments should be interpreted in this way.
The Fokker-Planck equation (10.6) now takes the form

(10.62)

with the initial conditions

PO[<p, O] = TI b[<p(y)], Pk[<p,O] =0, k= 1,2, ... (10.63)


y

We assume that, for 1 < I < k - 1, PI [<p, r] is uniformly bounded and


satisfies

(10.64)

In (10.64) w. lim again denotes the limit in the weak sense, as explained in
Introduction 10.1. We now go over to the following Lemma 10.1.
344 Chapter 10

LEMMA 10.1.

w.lim aa Pk[cp, rJ = O.
1"---+ co r

Proof We first transform equation (10.62) in the integral form:

Pk[cp, rJ = f L
<5cp' dr' Do[cp,cp'; r - r'J x

(10.65)

where Do[cp, cp', r - r'J is the Green functional of the free Fokker-Planck
equation

+ <5(r - r') TI <5 [cp(x) - cp'(X)J, (10.66)


x

with the boundary condition

Do [cp, cp'; r - r'J It t' =


= IJ <5 [cp(x) - cp'(X)]' (10.67)

By solving equation (10.66) we can construct Do [cp, cp'; r - r'J explicitly.


Similarly, we can construct Po[cp, rJ by solving equation (10.62) for 9 = O. The
results are (see Appendix D)

Do[cp, cp'; rJ = B(r) ~Ol exp { - ~ fd 4 kcp(k)D(r)cp( -k+ (10.69)

where No and ~o are normalization factors such that

In (10.69) cp(k) denotes [cp(k), cp'(k)] and D(r) is the following 2 x 2 matrix:
D11(r) = (k 2 + m 2)/{1 - exp[ -2r(k2 + m 2)]},
DJ2(r) = D21(r) = -D l l (r) exp[ -r(k 2 + m2)],
(10.70)
Further Developments in Stochastic Quantization 345

We can easily verify that, in the limit r -> 00, Po and Do satisfy the relations

P o[cp, r] ->
T~ co
PQ1 [cp] = (N'OQ)-l exp { - ~2 fd 4 kcp(k)W + m2 )cp( -k)}, (10.71)

We now look at Pk[cp, r] given by equation (10.65). By partial functional


integration we obtain

Pk[cp, r] = - f fd4x[b:~(:)J J:
bcp' dr' x

bDo[cp,cp';r-r'}p ["]
x{ bcp'(x) k-l cp ,r . (10.72)

We are also interested in the derivative aPk/ar:

aPk
ar = -fbCP' fd4x[~J ~ [t dr' x
bcp (x) ar Jo

c5D O [cp,cp';r - r'J}p ["J


x { c5cp'(x) k-l cp,1: . (10.73)

The exchange of the orders of differentiation and integration over cp' and r'
is harmless because, as we have already emphasized, (10.72) and (10.73)
represent the Green functions of the theory which are sufficiently regularized
even in the presence of local vertex insertions. We must first consider equation
(10.72) for k = 1, in order to verify the first step of the inductive hypothesis.
The reason is that Po [cp, rJ is not finite for r -> 0 because it satisfies the initial
condition [the first equality of (10.63)]. However, this b-function is multiplied
by bSin/bcp', which is a monomial in cp'. Therefore, using the explicit ex-
pressions of Po and Do given by (10.68) and (10.69), it is straightforward to
verify that PI [cp, r J exists, is bounded uniformly in cp for all values of r, has a
well-defined limit as r -> 00, and vanishes at r = 0, as required by the second
condition of (10.63). We must still prove that PI [cp, 00] = pen cp], but first we
prefer to complete the proof of Lemma 10.1 for arbitrary k. This is simple,
P k - 1 [cp', r'J is bounded and bDo/bcp' is integrable for all cp. The functional
integration over cp' in (10.72) is always possible because P k _ 1 contains the
Gaussian measure of the free field theory. Therefore Pk[cp, rJ, as given by
(10.72), is bounded for all cp and has a limit Pk[cp, 00]. A similar argument,
applied to equation (10.73), shows that aPk/ar also exists for all r and all cp
346 Chapter 10

and has a well-defined limit when r -+ 00. It follows that oPk/or -+ O. This
completes the proof of Lemma 10.1. •
Thus, we now return to gauge theories. Under an infinitesimal gauge
transformation with parameters ea(x) the fields A~(x) transform as
bA~(x) = D~beb(x) == bab 0lleb(X) - rbceb(x)A~(x) (10.74)
and the functional F[AJ as

bF[AJ = - fd4xea(x)D~b[b/bA!(x)JF[AJ. (10.75)

A gauge-invariant functional satisfies


D~b[b/bA!(x)JF[AJ = o. (10.76)
The crucial observation (Zwanziger, 1981; Baulieu and Zwanziger, 1981) is
that, if one wants to compute the average of such a gauge-invariant functional,
one may replace the stochastic equation (10.2) with the following:
oA~(x, r)/or = -bS[AJ/bA~(x, r) - D~bVb + I](x, r), (10.77)
where, in order to simplify the notation, we have omitted the subscript I] in the
stochastic fields. Dab is given by (10.74) and va = va[A, xJ is an arbitrary gauge-
noninvariant functional of A. Indeed, using (10.77), we can compute the time
evolution of a gauge-invariant functional F[AJ which satisfies (10.76):

of
or
= fd 4 x [OA~(X' r)
or
J[~-JbA~(x, r) ,
(10.78)

and we can see that, by virtue of (10.76), the term proportional to va does not
contribute. The average of F can be computed using a modified probability
distribution Pv[A, rJ which satisfies the following Fokker-Planck equation

oP
8r =
v fd 4 b {b
x bA~(x) bA~(x)
bS
+ bA~(x) + D~
b b}
v Pv[A, rJ. (10.79)

We see that, because of (10.76), the average value of a gauge-invariant quantity


F does not depend on the choice of va and is the same, irrespective of whether
one uses the distribution peA, rJ or the modified one Pv[A, rJ. Of course, this
is not true for gauge-noninvariant quantities.
We shall present here the special case va(x) = (1/0:) 0IlA~(x) with 0: an
arbitrary constant. Because S contains terms proportional to both g and g2,
the Fokker-Planck equation (10.62) will connect Pv,k with p v.k - 1 and Pv,k - 2
for k > 2. P v,l will only be connected to Pv,o' We again make the inductive
hypothesis (10.64) which we shall verify for Pv,o' The analogs of equations
(10.68) and (10.69) are

(10.80)
Further Developments in Stochastic Quantization 347

with
A:~ = (jab k- 2{((jJlV - kJlk v lk 2) [l-exp( -2rk2)] +
+1X(kJlk v lk 2)[1 - exp( -2rk 2IIX)]}. (10.81)
As expected, for IX = 1 we obtain the "Feynman propagator" for finite r. For
we recover the propagator given by Parisi and Wu (1981). We see that
IX -> 00

w. lim p v • o [A, r] = P';,;o [A], (10.82)


,~ 00

which verifies the inductive hypothesis. Similarly, we find for Do [A, A', r]:

Do = 8(r) AOl exp{ -~ fd4 k A~(k)D:~(r)A~( - k l (10.83)

(10.84)

where again A: denotes (A:, A:') and the 2 x 2 matrix D is the scalar matrix
given in equations (10.70) with m = O.
From this point the proof goes as before with the estimations which
correspond to a massless theory. From the explicit construction p v • o we first
prove the theorem for P v, 1; then using both, we proceed to P v, 2, etc. The
conclusion is that, in the limit r -> 00, we obtain a regularized perturbation
series which, however, is not the one we would derive from the usual theory
with a gauge fixing term proportional to (OJlAJl)2. This is obvious from the
absence of explicit Faddeev---Popov ghost terms whose contribution is replaced
by the more complicated structure of the vertices. There is a choice of va which
precisely reproduces the Faddeev-Popov perturbation theory, but in the
stochastic language it is a non local functional of A (Zwanziger, 1981; Baulieu
and Zwanziger, 1981).

10.6. The Mechanism of the Vacuum Tunneling Phenomena


in the Framework of Stochastic Quantization

This problem has been considered by Yasue (1979). We give here his method
of description of the vacuum tunneling phenomena. It is suggested (Polyakov,
1975a; 't Rooft, 1976, etc.) that a Euclidean path integral

f (jA exp { - ~ SE [A] + gauge-fixing term}.


where SE [A] is given by (10.45), provides a powerful tool to explore the
structure of gauge theory vacuum. Classical Euclidean solutions, which minim-
ize the Euclidean action SE [A], were shown to manifest tunneling phenomena
348 Chapter 10

between topologically inequivalent classical vacua within the WKB


approximation.
In the classical field theory, the SU(2) Yang-Mills field AI'(x, t) has an
isovector expression
3
AI' (x, t) = L A: (x, t)T a ,
a == 1

where {ra}, a = 1,2,3, is a basis of the Lie algebra of SU(2).


Dynamics of the Yang-Mills field is given by a Lagrangian

(10.85)

where the terms in the primed sum for /1, v = 1,2,3 are taken with reversed
sign. In terms of 'electromagnetic fields'
1 3
Bf = -2 L C;jkFjk'
k,j 1
=

f
Lagrangian (10.85) can be written as

51 = -21 ~3
L.... d3x(E~E~
1 1
- B~sa)
1 1·
(10.86)
i,a = 1

We shall work in Ao = 0 gauge. Then Ef becomes identical with GoAf = Af


and Bf does not contain Af. In this gauge Af's are dynamical variables of
the Yang-Mills field. Lagrangian (10.86) takes the form

51 = f
~2 i,a == fd3X(A~A~ -
I I I
BaBa)
1 l'

which gives us the following equation motion for Af:

;U =
I
- ~(~)
2 bAf
fd 3y BbBb
J J'

where the summation convention for all repeated Latin indices is assumed.
To quantize the Yang-Mills field, it is convenient to adopt the stochastic
quantization procedure (Yasue, 1979). This is because not only the structure of
the vacuum state but also the mechanism of the vacuum tunneling of the
quantized Yang-Mills field can be illustrated within the stochastic
quantization.
All the quantization procedures are the same as in Sections 7.4 and 7.5,
provided that the field equation is replaced by the following one

t(D+D_ + D_D+)Af(t) = -~(bA~(t)) fd3YB~B~. (10.87)

The quantized Yang-Mills field Af(t), - 00 < t< 00, is of course, a generalized
Further Developments in Stochastic Quantization 349

random process (for details, see Sections 10.3 and 10.4). The resulting
Schrodinger equation is

00.88)

where A is an abbreviation for {An, i, a = 1, 2, 3.


Following Yasue we shall investigate a structure of the vacuum state by
making use of stochastic quantization. Removing the infinite zero-point en-
ergy, we define a quantum theoretical vacuum state of the Yang-Mills theory
by a ground state of wave functional \f' (A) which satisfies the Schrodinger
equation

In the classical field theory, vacuum states of the Yang-Mills theory are
classical field configuration A;a(x) with zero potential energy

~f d 3 x Bf(A')Bf(A') = O.

They are pure gauge fields A; = (G') -I 0i G' E SU (2) Lie module on 1R 3 , where
G"s are unitary matrices such as limlxl~oo G'(x) = I. As limx~oo A'(x) = 0, we
can consider the pure gauge fields A' continuous mappings from IR 3 to SU (2).
Since ~ ~ S3 (three-dimensional sphere) and also SU (2) ~ S3, A"s can be
classified by the third Homotopy group of S3: n3(S3) ~ 71 (integers) with
respect to a fixed point belonging to 1R3. Namely, classical vacuum states of
the Yang-Mills theory consist of an infinite number of homotopy classes of S3:
(10.89)
where.[·J denotes a homotopy class to which the pure gauge field inside the
brackets belongs. Two pure gauge fields which can be joined by a continuous
gauge transformation in the manifold of SU (2) Lie module should be under-
stood as the same classical vacuum state (Jackiw, 1977).
In order to classify the homotopy classes, it is convenient to introduce the
Pontryagin index

For pure gauge fields, q is an integer which belongs to the homotopy group
n3(S3) ~ 71. Then, the classical vacuum states (10.89) can be rearranged as
{[AY = Gq- I OiGqJ}qEZ' (10.90)
where Ai is a pure gauge field with Pontryagin index q.
350 Chapter 10

In the quantum field theory the classical vacuum states (10.90) are rendered
unstable by the tunneling effect; they are metastable vacuum states. Let us
study the vacuum tunneling phenomena between metastable vacuum states
A{'s from the point of view of probability theory.
Within the conventional framework of quantum field theory, the wave
function '¥ (A) does not teach us the details of the vacuum tunneling. One can
describe the tunneling behavior of the quantized Yang-Mills field only within
semiclassical limits, i.e., within the WKB approximation.
It is assumed that the behavior of the quantized Yang-Mills field in the
vacuum state '¥ (A) is known to be a random process Af(t). Then, it can be
found that the random process Af(t), - 00 < t < 00, is a solution of the
stochastic differential equation
dAf(t) = Uf(A(t)) dt + dWf(t), (10.91)
where the transformation Uf(') is related to the vacuum state wave functional
by
a _ b
Ui(A) - h~( In '¥(A), (10.92)
uAf t)

and Wf(t) denotes a (generalized) Wiener process with the diffusion constant
tho Notice, that the stochastic differential equation (10.91) is an abbreviation
for the relation

Af(t) = Af(s) + 1 du Uf(A(u)) + Wf(t) - Wf(s), t > S.

Therefore, a transition probability of the random process Af(t) manifests the


tunneling process of.the quantized vacuum field configuration between metast-
able vacuum states A? The transition probability is given by an elementary
solution Q[A; s I A'; u] of the Fokker-Planck equation of the type (7.21) or
(7.22). Namely, we have

:s Q = -fd 3X b~?, [Uf(A)Q] + ~fd3 x bA~:A~Q,


, , (10.93)

and
lim Q[A; s I A'; u] = b(A - A'). (10.94)
s~u

To illustrate the mechanism of the vacuum tunneling, one needs to solve the
Cauchy problem of equations (10.93) and (10.94). This can be done by
introducing a relative transition probability F[A; s I A'; u] by
Q[A;sIA';u] = '¥(A)F[A;sIA';u]('¥(A'))-l. (10.95)
Equation (10.93) is transformed into a self-adjoint form

-h(:s)F = f d 3 x{ -th2CA;:Af) + tBi Bi }F. (10.96)


Further Developments in Stochastic Quantization 351

and the initial condition (10.94) into

lim F [A; s I A'; uJ = b (A - A'), (10.97)


s---+u

by the substitution (10.95). Equation (10.96) is nothing but an Euclidean


analog of the Schrodinger equation (10.88). The Feynman-Kac formula (6.32)
and (6.33) asserts that a solution of the Cauchy problem (10.96) and (10.97) is
given by the Wiener integral

F[A; s I A'; uJ

= fb[A(s) - AJ f.l{/ (bA(·)) exp { - 2111 f dt x

f
x d 3 x Bf[A(t)]Bf[A(t)] }. (10.98)

Here f.1':.v A ' de..notes a (nonstandard) Wiener measure with the diffusion coef-
ficient 11/2 defined on the totality of continuous paths A(') starting from A' at
t = u (for details, see Yasue, 1979). Correspondingly, the transition probability
of the random process Af(t) is found to be

n [A; s I A'; uJ = (If'(A)/If' (A')) f


b[A(s) - AJ f.1':.v A ' (bA(·)) x

x ex p { - 2111 ff dt d 3 x Bf[A(t)]Bj[A(t)] }. (10.99)

A tunneling probability of the quantized vacuun field configuration between


metastable vacuum states A[ and Af (q =l= p) from a remote past to a remote
future is neAP; 00 I Aq; - 00]. Noticing that neAP; 00 I Aq; - ooJ would coin-
cide with the conventional expression for the tunneling probability, i.e., the
ratio IIf' (AP)/If'(AqW, we find a vacuum tunneling amplitude of the quantized
Yang-Mills field in Ao = 0 gauge to be

x [A(t)]Bf[A(t)] }- (10.100)
352 Chapter 10

If we introduce a functional path integral expression of the Wiener integral

f.5[A(OO) - AP].u~OO,Aq(.5A(·)··· = NIP .5A(·) x


x ex p { - ;h f~oo dt fd XAf(t)Af(t)},
3

where bA(·) means to take a functional path integral and N is a normalization


constant, equation (10.100) becomes

= N IP bA(·) ex p { - 21h f~oo dt f d 3 x(Af(t)Af(t) + Bf[A(t)] Bf[A(t)])}.

In terms of the field strength tensor F~V' this can be written as

which provides an Euclidean path integral description of vacuum tunneling


phenomena in Ao = 0 gauge.
Thus, we have presented here Yasue's method of the Euclidean path integral
description of the vacuum tunneling phenomena in SU(2) Yang-Mills theory,
the validity of which has been proved from the point of view of probability
theory.

10.7. Stochastic Fluctuations of the Classical Yang-Mills Fields

Recently, the stochastic solutions to the Yang-Mills equations and the possi-
bility of derivation (based on these solutions) of effects connected with the
basic states of quantum chromodynamics have been discussed intensively
(Polyakov, 1975a; Gribov, 1977). In this connection, it is interesting to analyze
the classical Yang-Mills equations without external sources. The result could
be useful for the construction and study of a vacuum structure and a problem
of asymptotic states of the theory.
In this section we briefly discuss the study of a nonlinear oscillation in the
classical Yang-Mills theory. We consider the Yang-Mills equations without
external sources, corresponding to SU (2) group:
(10.101)
in the co-ordinate system, where the Poynting vector F~J'Ji turns to zero. The
Further Developments in Stochastic Quantization 353

field strength tensor F:v is given in Section lOA. For this case it is enough (in
the gauge viAf = 0, Af = 0) to satisfy one of the following relations:

(a) A f.1 = 0, (Fi,1 == VI F i ; i, 1 = 1,2,3).


(b) Af = O.
(c) Af,j - Aj,i = O.

(the point denotes, as usual, a differentiation with respect to time). In the case
of (a) when in the chosen co-ordinate system the potential depends only on
time, the Yang-Mills equations take the form:
(10.102)
The system described by equations (10.102) is naturally equivalent to the
discrete nonlinear mechanical system with the Hamiltonian
H = ¥Af)2 + ±g2 [(Af Af)2 - (Af Aj)2], (10.103)
which coincides with the energy of the field system (10.101).
It is natural to mention the (nonlinear) classical Yang-Mills mechanics
determined by equations (10.102). Notice that the corresponding classical
Maxwell mechanics is trivial and described by the equation Ai = 0, generating
the constant electric field Ai = - EJ, but due only to the nonlinearity of the
Yang-Mills equations (as shown by Matinyan et al., 1981), there appear
complicated nonlinear oscillations of colour.
In the simple case that allows an analytic solution, a nonlinear plane wave
arises. This plane wave is described by the elliptic cosine law with respect to
time. Indeed, if one searches for a solution to system (10.102) in the nine-
parameter form Af = Qf R(a) (t)/g, (over a there is no summation), where Qf is
an orthogonal matrix independent of time, QfQr = bab, then from (10.102) we
obtain the following system of equations
k(a) + R(a)(R 2 - R(a)') = 0, (10.104)
where
R2 = R(l)2 + R(2)2 + R(3)'.
The simple case (allowing analytic solution), when all three colours change in
the same form R(l) = R(2) = R(3) = R(t) with respect to time variable, was
investigated by Baseyan et al. (1979). In this case system (10.104) is reduced to
the nonlinear equation with one degree of freedom, the solution of which has
the following form:
(10.105)
where cn(x; k) is the Jacob elliptic cosine of the argument x and module k; to is
an arbitrary origin of the time moment, and /1 4 is an energy density in the
given co-ordinate system.
354 Chapter 10

Corresponding to expression (10.105) chromoelectric Ef = Qf R/g and


chromomagnetic
Bf = c;ijJabcQJQ~R2 /g2

intensities of fields in the 'rest' system of reference are parallel to each other,
but three vectors El, E2, E3 are orthogonal to each other. Such a picture
differs from the one, realized in the plane wave of classical electrodynamics,
where E ...L B. The intensities Ea and Ba change from time with the period

T = (3/8g 2 )1/4(4/.u)K(1/j2),
where K is a total elliptic integral of the first kind. Under the Lorentz
transformations the nonlinear wave [arising from (10.105)] with arguments
kx = kl'xll' where ko = .u{3, k i = .uVi{3 ({3-1 = (1 _V 2 )1/2, k 2 = .u 2) differ from
the non-Abelian linear plane wave due to Coleman (1977) (k 2 = 0) by the fact
that the value of the Poynting vector does not equal energy density (k 2 = .u 2)
(Baseyan et al., 1979).
So, we see that investigating the mechanical system (10.104) by the Lorentz
transformations, a solution to system (10.101) in an arbitrary co-ordinate
system can be obtained. Such a solution always contains a plane wave.
Therefore, our problem leads in fact, to the study of the classical Yang-Mills
mechanics (10.102). In connection with the above-mentioned, it is interesting
to study a further problem of the nonlinear oscillation of color in the classical
Yang-Mills theory. The solution of these problems with the number of degrees
of freedom N > 1 leads to the investigation of system (10.102) [or in the simple
variant of system (10.104)].
Matinyan et al. (1981a, b); Chirikhov and Shepelyansky (1981) carried out a
numerical integration of system equations (10.104) and obtained those sto-
chastic solutions in the two and three dimensional cases. They investigated the
nonlinear mechanical system with the Hamiltonian
n= 2 H2 = x2/2 + y2/2 + x 2y2/2,
n = 3 H3 = t{x 2 + y2 + 22) + t{X 2y2 + y2 z2 + X2Z2),
and with the corresponding equations of motion
n = 2 x + xy2 = 0; ji+ yx 2 = 0;
n=3 x + X(y2 + Z2) = 0; ji + Y(X2 + z2) = 0; z + z(x 2 + y2) = o.
The system when n = 2 is obtained from (10.104) for R(3) = O. Here, the
following notation is introduced

In conclusion we notice that the obtained connection of the Yang-Mills


mechanics with a stochastic process, is not surprising, since the Yang-Mills
mechanics is a nonlinear system described by some nonlinear partial differen-
Further Developments in Stochastic Quantization 355

tial equations. Thus, among the set of solutions to the Yang-Mills equations
generating a whole row of remarkable effects: monopole (Wu and Yang, 1969;
Polyakov, 1975b; 't Hooft, 1974; Nielsen and Olesen, 1973), instantons
(Polyakov, 1975a; Belavin et ai., 1975), and merons (de Alfaro et al., 1976;
Callan et al., 1978; Baseyan and Matinyan, 1980) there should exist stochastic
solutions responsible for the stochastic origin of all the physical phenomena
including the vacuum and space-time.

10.8. Appendix D. Solutions to the Free Fokker-Planck Equation

In this appendix we calculate the quantities Po[<p, r] and Do[<p, <p'; r] in


equations (10.62) and (10.66). By definition Po [<p, r] is

P o[<p, r] = f51'/ ex p { -~ f d 4 X dp '1'/2 (x, P)} ry b[<p(y) - <p~(y, r)], (D.1)

where <P~ satisfies the free Langevin equation


o<p~/or = (0 2 - m2)<p~ + I'/(x, r). (D.2)
The solution of this equation with initial condition <P~ (x, 0) = 0 is the
following

<p~(x, r) = f I d4 y dp G(x - y; r - p)l'/(y, p), (D.3)

where the Green function G(x - y; r - p) is given by (dimension of space


d = 4)

G(x - y; r - p) = (2n)-4 f d 4 k exp[ik(x - y) - (r - p)(k2 + m2)]. (DA)

We are also interested in the correlation function D,


D(x,y;r,r') == <<p(x,r)<p(y,r')\

= 2 fd 4 Z I~i~) dpG(x - z; r - p)G(y - z; r' - p). (D.5)

In order to find Po[<p, r], we write equation (D.1) in the form

Po[<p, r] = f bl'/ b~ exp { - ~ f d 4 x dp '1'/2 (x, P)} x

x exp{i f d 4 x ~(x)[<p(x) - <p~(x, r)] }. (D.6)

Then, we introduce expression (D.3) for <P~ and we do the Gaussian, 1'/ and ~
356 Chapter 10

integrations. The result is

Po[cp, r] = N o exp { - ~ff d


1 4 x d 4 x' cp(x)D- 1(x, x'; r, r')cp(x')}' (D.7)

where D -1 is defined as

f d 4 z D(x, z; r, r)D-1(z, x'; r, r) = b(4)(x - x'). (D.8)

From (D.5) and (DA) we find

x {l- exp[ -2rW + m2 )]}(k 2 + m 2 )-l, (D.9)


and so in momentum space (D.7) reduces to expression (10.68).
For the functional propagator Do[cp, cp';r] we use its definition
Do[cp, cp'; r - r'] == Po[cp, cp'; r, r']/Po[cp', r'], (D.10)

as the conditional probability density, where

Po [cp, cp'; r, r'] = f 15 11 exp [ - ~ f


d 4 xd p' 112 (x, P)] ry b[cp(y) - cp~(y, r)] x
x b[cp' (y) - cp~(y, r')]. (D.ll)
We follow the same method as for Po[cp, r] and forming the ratio (D.10), we
find expression (10.69).
Chapter 11

Some Physical Consequences of the


Hypothesis of Stochastic Space-Time and
the Fundamental Length

11.1. Prologue

As was noted above (Chapter 7), there are three approaches to the con-
struction of the theory of stochastic processes in physics. The traditional
approach describing the diffusion phenomena (also Brownian motion) is based
on the classical works due to Einstein (1956) and Smoluchowski (1923) (see the
reviews by Skagerstam, 1975; Moore, 1979; Yasue, 1979; Nelson, 1979; Guerra,
1981). Further developments in this theory are connected with the names of
Uhlenbeck and Ornstein (1930); Wang and Uhlenbeck (1945); Chandrasekhar
(1943); Kac (1959); Rice (1944, 1945). Mathematical foundations of the sto-
chastic theory for physical processes were laid down by A. Markov, N. Wiener,
A. N. Kolmogorov, K. Ito, N. N. Bogolubov, etc. New stochastic methods in
physics have been recently given in DeWitt-Morette and Elworthy (1981) (see
Chapter 10). The other two approaches are connected with the properties of
the electromagnetic vacuum (zero-point field) and with the postulate that the
random behavior of a physical system is caused by the stochastic nature of
physical space-time (see Chapters 7-9).
In this chapter, we would like to show that all the above-mentioned
approaches to the stochastic theory are connected to each other with the
hypothesis of stochastic space-time as the connection. On the basis of this
connection, it is shown that the spectrum density of the zero-point elec-
tromagnetic field in stochastic electrodynamics changes at small distances.
This fact gives a reinterpretation to the electromagnetic vacuum as a real
physical 'medium' and allows us to conclude that it is an energy density, i.e.,
the density of matter due to electromagnetic interactions may be restricted.
From our considerations presented below, one can therefore establish that the
introduction of the fundamental length [ into the stochastic theory leads
automatically to a restriction on the density of matter by u/ ~ hn - 2[- 4/C.
Further, an attempt to unify the scales of physical quantities (masses, lengths,
etc.) is presented in the framework of a hypothesis of the existence of a family
of small 'black holes'. On the semi-empirical level this proposal provides the
possibility of finding unique scale [characterized by the parameter ta (a being

357
358 Chapter 11

the fine structure constant)] which separates on different subscales including


the scale of the grand unification. It transpires that all the typical physical
scales (atomic, molecular, electromagnetic, weak, grand unification and gravi-
tational) of lengths are covered by this unique scale.
In the framework of the new approach to gravitation (de Alfaro et al., 1980)
the decay of the proton is discussed (Section 11.5) on the basis of the neutron
oscillations and of superfields. It is shown that a mechanism of the proton
decay is very sensitive to a value of the fundamental length 1. A possibility is
discussed that if the fundamental length is not larger than the Planck length
lpl ~ 19r ~ 10- 33 cm; then the proton decay is wholly due to the gravitational
interaction and its life-time is of an order of 10 43 years.
In Section 11.6 we presented some new results related to the hypothesis of
non locality and stochasticity of the space-time metric and its consequences. As
shown above (Chapter 1), the hypothesis of the stochastic nature of space-time
at small distances leads to the space-time metric nonlocality (1.12). The last
fact makes the theory nonlocal too, so the field equations and Green's
functions of test fields (particles) are changed according to results obtained in
Section 11.6. Moreover, in our proposal, Einstein's relation between the energy
and momentum of particles is also changed. Speculation concerning the creation
of elementary particles as some perturbation of the space-time metric resulting
from the nonlocality, is presented. A generalized mass operator (momentum
operator) of ground states of test bosons and fermions, and concrete forms of
the space-time metric nonlocality are constructed, which allow one to describe
the equidistant behavior of the mass spectra of observed particles in semi-
empirical level only.
Further, we deal with some comments on the physical mechanism responsible
for the acceleration of cosmic rays. It is suggested that this mechanism is caused
by the stochastic (or fluctuational) structure of space-time at small distances. A
method of introducing fluctuations in a cmiformally flat Riemannian space-time
metric due to ultra-high energy particles is presented. In this framework, a
nonlinear dynamics of particles and equations for the electromagnetic fields are
obtained. The former admits the acceleration mechanism for cosmic-ray par-
ticles and the fact that their energy increases extremely during the evolution of
the Universe. In our model the energy of cosmic-ray particles, say, protons and
their radius (the effective Schwarzschild), the age of the Universe and the value
of the fundamental length are interconnected, and are determined by a unified
formula, Einstein's relation for the relativistic particle energy. It allows one to
define experimentally the bound on the value of the fundamental length that is
lexp = 1.56 X 10- 33 cm for the maximum proton energy EOb = 10 20 eV observed
in primary cosmic rays.
Within the new acceleration mechanism of cosmic rays proposed here the
problem of the energy spectrum of the cosmic rays and the ratio of the
intensities of the electron component to the proton component at the same
energy level are discussed.
Physical Consequences of Hypothesis of Stochastic Space-Time 359

To conclude this chapter we speculate concerning an interrelationship


between the properties of the space-time structure near moving particles and
their dynamics. It is suggested that the space-time metric near particles
becomes curved g/lv(x, bE) depending on a random vector bE = (b 4, b) with a
distribution w(bi/F), the averaged space-time metric <g/lv(x,bE) over this
distribution gives a general effect on particle behavior. As a result the particle
motion in our scheme is described by a nonlinear equation. It transpires that
the nonrelativistic limit of this equation gives a simple connection between the
space-time structure at small distances and the dynamic behavior of particles.
Different types of the particle motion (nearly rectilinear, stochastic and soliton
like) caused by some concrete forms of the averaged conformally flat space-
time metric <g/lv(x, bE) are considered.

11.2. Nonlocal-Stochastic Model for Free Scalar Field Theory

In this section we consider the free scalar field in the space-time ~4(X) in the
framework of the Davidson (1980a) model (see Section 10.2). In the stochastic
space-time ~4(X), the field (10.9) may be represented in the fvrm:

cp(x) = (hL 3) -1/2 ~>iq(X+ b) cp/t + iT), (11.1)


q

where T is a real random variable. According to the above deduction (see


Chapter 1) we must average the field (11.1) with the measure w(bi/12). For this
purpose, we introduce intermediate Euclidean variables (h/2v)b 4 = iT, x4 = it
(generally speaking, v is a complex number), and average the expression (11.1)
with w(bi/1 2), where b4 = iT(2v/h); thus we obtain

= (hL 3 )-1 /2 I e iqx K(Q)·cpq(t), (11.2)


q

where

K(Q) = fd4bEWCneXp(ib~.Q) = 4n 2 Pa- 1 1'Xl dyy 2 J 1(aly)w(y2), (11.3)

b~ = (2in/h, b);

Here J 1 (z) is the Bessel function. As before, we are interested in such a class of
measures for which K(Q) are entire analytic functions of the variable Q2. For
example, in the cases of WO(y2) = limm~o Wm(y2), WO (y2) in Section 1.3, and
360 Chapter 11

Wl(y2) in Section 11.6, we obtain


Ko(Q212) = 3(Q2[2)-1[sin(Q2[2)1/2/(Q2[2)1/2 - cos (Q 212)1/2J,

KG(Q 212) = exp[ -Q 212/8J, KdQ 21) = sin 2 tjQ2/(tJ,{Pf


We now come to the main result which can be formulated as the following
theorem:

where S~ ( ... ) are modified n-point Schwinger functions determined by formula


(10.15) (see also Chapter 7), where for the scalar field the function S~ ( ... ) takes
the form

SM( t· t) - (2 )-4 fd4 exp[iq(Xl - X2) + iq4(t l - t 2 )J


2 Xl' U X2' 2 - n q q~+q2+m2/h2 x

(11.5)

The functions (11.5) with the form factor V (z) = [K(Z)]2, physical meaning of
which was discussed in Section 2.3, are called the non local Schwinger functions.
Proof The proof is in the literal repetition of the results obtained above in
Section 10.2. As a result, we obtain

[{ <ip(x ><ip(x 2) >} =


l ) L -3 ~ exp[iq(X l - X2)JK(I{ q2 - (::2) :ti J) 2
x

(11.6)

>
Here we assume that <ip(X) is a Gaussian process because, as discussed in
Section 9.1, the method of averaging (11.2) cannot change the physical nature
of the objects and may only change the spatial structure of the objects which
are extended (nonlocal) in some domain characterized by length l.
Operating on the exponential in (11.6) by the operator K(Q) and taking into
account that K(Q) is an entire analytical function of Q, we perform the limit
L ~ 00 and obtain

o
Notice that what is most important in the proof is the order of differentiation
Physical Consequences of Hypothesis of Stochastic Space- Time 361

with respect to the a/at in the operator K(Q) and of averaging with the
measure w(bi/12). It is necessary to perform first the averaging with w over the
field (11.1) at every point of the space-time 1R 4 (x) and then (at the last stage of
the calculation) to act by the operator K(Q). If this application of K(Q) took
place in some intermediate stage of the calculation, for example, if K(Q) would
act on the function exp{ -0:1t1 - t 2 1}, then the result obtained would cor-
respond to the usual theory (i.e., local theory). Indeed

K( Iz[ q2 - (::2) :t2i J) K(ZZ[ q2 - (4~2) :t2~ J) exp{ - 0: It 1- t21}

== K2( _~2ZZ) exp( -0:1t1 - t 21),

but the function K is normalized so that K( - m 2ZZ /h2) = 1.


One can easily see that for a particular value of the parameter v = h/2,
which is the value used by Guerra and Ruggiero (1973) we obtain the modified
(or nonlocal) Schwinger function (11.5). In the limit I -> 0 this function
becomes the local one therefore we reach the Guerra and Ruggiero result
exactly.
In our case, equation (10.18) takes the form

IEV=ih/2 <<p(x) <<p(y)} =


{
A A •

1(2n)
-4fd 4
qV(-q I )
22 exp[iq(x-y)]
2 2/h2 + Ie. , (11.7)
q - m
where V( _q212) = [K(q212)]2, q2 = q~ - q2.
Comparing (11.7) with the usual
Green's function of the nonlocal quantum field theory (Part I) we obtain
lEv = ih/2 {< <p(x) <<p(y)} = Dc(x - y) = <01 T{ 4>(x)4>(y)} 10).
Here

and 4> (x) are the nonlocal causal Green function and nonlocal field con-
structed by the nonlocal distribution (generalized function) K(x) given formula
(1.11). Thus we come to the nonlocal theory investigated in Part I. In the
nonlocal theory with the form factors V( - q212) belonging to the class of entire
analytical functions there exists an intermediate regularization procedure
(Chapter 2), which permits the changing of the contour integration in (11.7) to
a necessary domain in the complex plane qo (i.e., planes I and III).
In the nonlocal case the corrections similar to (10.22) and (10.23) are valid
lEv = ih/2 {< <p(x 1) ... <<p(xn ) } = <01 T{ 4>(xd· .. 4>(x n )} 10),
lEv = _ ih/2 {<<p(X 1) ... <<p(Xn ) } = <0IT* {4>(X 1) ... 4>(X n )} 10),
where T and T* are the Wick time and antitime ordering of operators 4> (x).
362 Chapter 11

Thus, the free scalar field in our model is equivalent to the nonlocal Markov
field whose covariances are obtained from the non-local Schwinger functions
in which all times are scaled by a common factor 2v/h. The hypothesis
concerning the indeterminate nature of v in quantum field theory makes it
possible to obtain the Efimov nonlocal theory. In the given scheme, the
Euclidean-Markov field and quantum field theory may be constructed in
which no cut-off procedure is involved.

11.3. Zero-Point Electromagnetic Field and the Connection Between the


Value of the Fundamental Length and the Density of Matter

In the previous section, in the framework of the stochastic quantization


method proposed by Fenyes (1952), Nelson (1966, 1967) and Davidson
(1980a), we have considered the free scalar field in the stochastic space-time
1R4(X) and have obtained the two-point non local Schwinger function for the
Euclidean-Markov field. We shall consider here the electromagnetic field in
1R4(X) (for the local case see Section 10.3 and Guerra and Loffredo, 1980). In
the case of the electromagnetic field expression (11.4) takes the form (in the
local case such generalization has been done by Moore, 1979):

Here it is assumed that v = h/2. It appears that in the framework of our


scheme the nonlocal Euclidean-Markov electromagnetic field (11.8) produces
the zero-point electromagnetic (ZPEM) field.
THEOREM 11.2. The covariance of the ith component of stochastic ZPEM
potential AI?) (i = 1,2,3) for the whole space is

(11.9)
where

By using the dipole approximation in the case of a local theory (I -+ 0) the


spectrum density of the zero-point electromagnetic field has been obtained by
Moore (1979) from equation (11.9). However, in this case frequency cutoffs are
employed, since 110(0) diverges. In our model 111(0) < 00, the spectrum density
may be calculated without the cutoff procedure, and the dipole approximation
is not involved. Let us calculate the spectrum density of the ZPEM field for a
specific form of the form factors V(q~ 12 ), say VG = exp( - q~ [2) see formula
Physical Consequences of Hypothesis of Stochastic Space- Time 363

(1.16). Then

IE{Al~!(x, t1)Al~l(x, t z )} = -tn- z foro dwQz(w)cosw(t 1 - t z ),

where
(11.10)
Here erfc(z) is the integral of probability (see Section 2.3). The second term of
(11.10) diverges in the limit 1-->0 which reflects the fact that a divergence
appears in the local theory. In order to get rid of the second term of equation
(11.10) in the nonlocal theory one also needs (finite) renormalization. Thus

IE{El~l(x,t)El~l(x,t)} = foro dw·pz(w)cosw(t 1 - t z ),

where
pz(w) = n- Z w 3 erfc(lw)!2,erfc(z) = 1 - erf(z). (11.11)
As is to be expected, in the limit I --> 0 the nonlocal spectrum density (11.11)
goes over to the spectrum density of the ZPEM field due to stochastic
electrodynamics (SED) (see Chapter 7). We see that the introduction of the
fundamental length into the spectrum density of the ZPEM field leads to a
restriction of the density of matter (or more exactly, electromagnetic matter)
by some value U z • We now calculate a value of U z • In our model the energy
density is finite in accordance with (11.11):

Uz = foro dw·pz{w) = ±<Ez + HZ) = (EZ).

For V = exp( -q~ IZ) we have Uz = (3/32)n- z l- 4 h/c in ordinary units. From
high-energy experimental data (see Introduction 1.1 and references therein) it
follows that I ;;:; 10 -16 cm and therefore Uz :<: t lO z5 g cm - 3. In view of this we
notice that the estimate of I due to the astronomical approach (see also Section
11.7) proposed by Ginzburg (1975) and the search for superdense matter may
be of great interest (see Shuryak, 1980).
Thus, in our scheme SED with the spectrum density (11.11) appears and we
call it a nonlocal SED. Therefore the concept of nonlocality may be introduced
by means of the properties of the vacuum fluctuations (or zero-point elec-
tromagnetic field).
It has been pointed out above that random or stochastic electrodynamics
(SED) (i.e., the theory of the motion of charged particles in the presence of the
electromagnetic vacuum) is constructed on the basis of the hypothesis about
the existence of the zero-point random electromagnetic radiation field E(O) with
the spectrum density
p(w) = hw 3 /2n z c 3 .
364 Chapter 11

If we neglect the force due to the magnetic radiation field; then the nonre-
lativistic basic equation of SED is
mr = eE(O) + f + mrr; r = 2r:t./3mc\ r:t. = e2 /4n, (11.12)
where m is the mass and e the charge of the particle, f = f(r, t, t) the external
force.
In the nonlocal SED obtained above, the equation of motion for the
charged particle has the same form as equation (11.12) with E(O) ~ ElO). Here
E\O) is the non local electric component of the zero-point nonlocal elec-
tromagnetic field (modified by the presence of the fundamental length 1) with
the spectrum density given by (11.11). Study of equation (11.12) with
E(O) ~ ElO) and its solution goes beyond the scope of the present book. We
notice only that in the case of the nonlocal SED arising from the nonlocal
Euclidean-Markov and Efimov theories, the fulfilment of the Lorentz co-
variance follows from the fact that the Lorentz and four-dimensional rotation
groups are isomorphic in a complex domain (see Chapter 1).
In conclusion we would like to note that our hypothesis of existing
superdense matter may provide some interesting consequences in the theory of
gravitation. For example, as suggested by Markov (1982), this hypothesis
allows the possibility of rewriting the gravitation equation in the following
form

R~ - 1R c5~ = 8;4G T~[ 1 - (~) 2]" _A (~yn c5~, (n ~ 1). (11.13)

Here it is assumed that a value of matter density U is always restricted by a


limiting quantity

Further, we distinguish two possibilities:


(1) In the case of (u/uplf~ 1, i.e., for h ~ 0 equation (11.13) turns to
Einstein's classical equation (7.1).
(2) In the case of (u/up1f = 1 equation (11.13) describes de Sitter's world
(u/uPlf = 1, R~ - 1R(j~ + A· (j~ = O.

On the other hand, this limiting principle (u';;;: upl ) may be expressed by
geometrical language, for example, by the relation:
Q = Rllvo;.RllvbA .;;;: 1/1~1'

Now equation (11.13) can be rewritten in the form

(11.14)
Physical Consequences of Hypothesis of Stochastic Space- Time 365

Denoting the left and right parts of (11.14) by G~ and B;,


respectively, we
obtain an additional equation in the form of the covariant derivative of the
right part of (11.14)
B~;v = O. (11.15)
This equation results from the well-known property of Einstein's equation
G;;v = O. Equation (11.15) leads to the fact that the covariant derivative of
energy and momentum tensor does not turn to zero T;; v =F O. The last
equation describes the process of the disappearance of matter (tensor T;) with
generating de Sitter's world. In other words, equation (11.15) gives the
transformation of matter into the vacuum state of the de Sitter world. As is
known, this world is unstable. The state equation in it is p + 8 < 0 that is a
macroscopic analogy of overheated liquid or overcooled steam. Equations of
the type of (11.13) and (11.14) are accommodated to the description of the
oscillating Universe (see Markov, 1984 and Walstad, 1980).

11.4. Hierarchical Scales and 'Family' of Black Holes

In contemporary physics some hierarchical scales for the physical quantities


(lengths, energies, times, etc.) are observed. (In particular, a problem of
hierarchy in physics has been discussed by Recami, 1981 and in references
therein). For example, atomic and molecular processes occur in the domain
characterized by length 0(10-8)cm, but electromagnetic processes: 0(10- 1°-
1O- 11 )cm. The length 0(10-13)cm is just a typical distance for nuclear
physics. The majority of physicists believe that weak and quantum-
gravitational effects will be manifested at distances of the order of
lw ~ 6.7 X 10- 17 cm and 19r ~ 5.72 X 10- 33 cm, respectively. We recall that
these lengths are connected with the Fermi GF and Newton G g constants by
the formulas:

Recently, some models (Pati and Salam, 1973; Georgi and Glashow, 1974;
Buras et al., 1978; Goldman and Ross, 1979; Marciano, 1979) have suggested
that grand unification, i.e., the link of electroweak and strong interactions
seems to take place at the energy scale of the order of mx ~ 6 x 10 14 GeV [at
length 0(1O- 29 )cmJ with the gauge coupling constant IXGUM = g2/4n ~ 1/40 at
this scale.
The present section is devoted to obtaining the parameters of mx and IXGUM
for the grand unification theories within the-framework of a hypothesis about
the existence of a family of black holes. It turns out that because of this
hypothesis all the above-mentioned typical lengths are confined to the unique
scale measured by the parameter tlX, IX being the fine structure constant. The
analogous possibility has been pointed out by Salam (1980).
366 Chapter 11

Following an idea of Tennakone (1974), we make the assumption that some


elementary particles are black holes (or monopoles) in the strong gravitational
field. Thus let us consider as a 'family' the chain of black holes;

i= 1·, 2·, 3·, (11.16)

(it is assumed that mli~ 1 ~ mli' i = 1,2, ... ) which are singularities in the
Nordstrom-Reissner solutions of the Einstein equation for a particle of mass
m Ii and charge e. This solution is given by the metric

(11.17)
where
(11.18)

If now we assume that the structure of space-time in the immediate


neighborhood of the charged heavy particles}; belonging to an ith pair of the
family (11.16) is determined by strong gravitation, then the constant Gg in
(11.18) should be replaced by the strong gravitational constant G~:

Then the singularities in the strong gravitational field of the charged particle};
are given by the metric (11.17) with
(11.19)
where mIP is the mass of an uncharged partner of the heavy particle}; which is
also assumed to exist.
Since G: mJ/e 2 ~ 1, the equation Ys = 0 has two real roots rIi ~ 1 and rIi" For
those one can easily obtain the following ratio (for details, see Tennakone,
1974):
(11.20)
or
(11.21)
where
rIi~l = e2 /2c 2 mIi·
0 (11.22)
Here rIi may be interpreted as the Schwarzschild radius of the singularity
corresponding to particle!;; m li and mIi~ 1 are masses of heavy and light
particles (black holes) belonging to an ith pair of (11.16), respectively.
It should be noted that the strong gravitational constant G;
emerges as a by-
Physical Consequences of Hypothesis of Stochastic Space-Time 367

product of the theory. Indeed, from (11.21) and (11.22) we have


G:. = e2/4mIi _ miP' 1
(11.23)
Now we determine a parameter which characterizes the hierarchical scales for
the physical quantities in 'black holes' physics. According to Tennakone (1974),
identifying fo with the electron and f1 with the proton (in given case
ff == neutron) we get from (11.20) (since mp '" mn )
e2 h h
r 1 = - - = tct- = 1.4 x 10- 13 cm,
2hc meC meC
or in terms of the energy scale
mWl = h/cr 1 = 140 MeV.
Further, assuming f1 = p in the next pair (i = 2) of the family (11.16) we have
mp = e2mh/2c2r2mI~'

By assumption mh :::::: mI~' Then


r2 = cth/2mpc = 7.7 x 10- 17 cm or mW2 = h/r 2c = 274mp = 257 GeV,
etc.
Thus, we see that number tct plays the role of scale in the transition from
one pair to another in the family of black holes, and it may be interpreted as a
parameter of hierarchy in physics. Further it is postulated that the mass ratio
of each hole belonging to the family (11.16) is given by parameter tct:
2
mIi =-m
ct Ii - 1 or mii = (2/ct)i - 1 mp' (i = 2,3, ... ). (11.24)

Typical scales for ri [mwJ corresponding to this parameter ct/2 are shown in
Table 11.1. Now with the same scale ct/2 and formally continuing the hie-
rarchical scales shown in Table 11.1, to the left we obtain the typical lengths
3.86 x 10 - 11 cm and 1.06 x 10 - 8 cm for the electromagnetic and atomic-
molecular processes. Further continuation of the hierarchical scales to the right
is impossible and the chain of black holes breaks at i = 9. Since in the case
i > 9 (we obtain an object with the mass larger than the Planck mass
mpi '" 1.2 x 10 19 Ge V) classical solutions of the Einstein equations are not
acceptable and quantum gravitational effects should be taken into account.
Let us calculate the gauge coupling constant at the scale of order
mx '" 4 x 10 14 Ge V. For this assuming mI ~ '" mh and taking into account
(11.24), from (11.23) we obtain

ct
GUM
g2 G7
= 4rchc = ;rch2 = 16
m
2 ct (m) 1
m~: '" 25'
From Table 11.1 we see that mx/mw '" 1.5 x 10 12 .
368 Chapter 11

Table 11.1

Black
holes

No. (i)
(;)
i=l
(~)
i= 2
G)
i= 3
U:)
4
(j:)
5

r; (em) 1.4 x 10- 13 7.7 x 10- 17 2.8 X 10- 19 1.0 x 10- 21 3.7 X 10- 24

mw,(GeV) 0.14 257 7.0 x 104 1.9 X 10 7 5.3 X 109

Comparison m, ~mw mmon

So, we notice that the scale corresponding to the seventh pair of black holes
may claim to be a scale of the grand unification (or, at least, a lower bound to
the grand unification scale). In distinction to the paper by Tennakone (1974)
where one pair (e, p) of black holes was considered, in this section we have
assumed that there exists a family of black holes in the strong gravitation field.
Within the framework of this hypothesis the number of pairs of black holes is
restricted and equals 9 = 1 + 8. Thus, in our case the 'natural' emergence of
mass hierarchies is [mpb (a/2)mpl' (a/2)2 mp1 ,'" ,mwJ or [mp, (2/a)mp,
(2/a)2 mp' ... ,mp1J.
Speculative extension of the unique scale discussed in the case of the
microworld to the large scale of the Universe gives the following hierarchy in
lengths Li = (l/a)iLo (i = 0, 1,2, ... ), where it is naturally assumed that
Lo = rear is the radius of the Earth. Quantities of this macroscale are presented
in Table 11.2.

Table 11.2

Scales in lengths (cm)


Number
of scales Theoretical Expected lengths from
calculating lengths Astrophysical Data

0 6.35 x 10 8 The radius of Earth rear'


1 8.7 X 10 10 r0= 7 x 10'0 em
2 1.2 X 10 13 A verage distance from Earth to Sun, i.e.,
rAU = 1.5 x 10 13 em
3 1.6 X 10 15 Radius of the solar system rss = 0.6 x 10 '5 em
4 2.2 X 10 17 Light year 9.5 x 10 17 em
5 3.1 X 10 19 Nucleus of Galaxy rNG = 10 parsec.
6 4.2 X 10 21 Distance from Sun to Galaxy centre rGC = 8
kilo parsec.
7 5.7 X 1023 Distance between centres of Galaxies
R ~ 18 x 10 23 cm
8 7.9 X 10 25 Average diameter of Meta-Galaxy R ~ 1025 cm
9 1.1 X 10 28 Radius of the Universe, i.e., Hubble radius
RH ~ 1028 em
Physical Consequences of Hypothesis of Stochastic Space- Time 369

Table 11.1-(Continued).

Black
holes (~) U:) (j:) (z)
No. (i) 6 7 8 9

r,(em) 1.3 x 10- 26 4.9 X 10- 29 1.8 x 10- 31 6.6 X 10- 33

mw,(GeV) 1.4 x 10 12 3.9 X 10 14 1.0x10 1 ? 2.9 X 10 19

Comparison mx mpl

11.5. The Decay of the Proton and the Fundamental Length

Recently the problem of proton decay was discussed intensively within the
framework of grand unified theories (see, for example, Ellis et al., 1980;
Weinberg, 1981; Langacker, 1981). The possibility that the proton decays is
not only important for grand unified theories but it is of great interest from the
point of view of cosmology and of the general philosophical problem of matter
structure and its stability. The hypothesis of neutron oscillation and the
effective many-fermion operators for the proton decay have been considered
by Kazarnovsky (1980); Mohapatra and Marshak (1980); Pati et al. (1981); L.
Chang and N. Chang (1980); Sawada and Fukugita (1980); Weinberg (1979);
Wilczek and Zee (1979); Smirnov (1980), etc. High sensitivity experiments in
search for proton decay are undertaken and projected in many laboratories.
In this section we study proton decay within the framework of the hy-
pothesis concerning the existence of the fundamental length in Nature, i.e., we
try to connect a mechanism of proton decay to the properties of space-time. It
is shown (from the grand unified theories) that if the proton decays, then it
occurs at least in an energetic domain determined by a scale M ~ 10 14_
1015 GeV (or I = h /M c ~ 10 - 2 8 - 10 - 2 9 cm). Therefore we make the assump-
tion that the structure of space-time (i.e., an effect of gravitation) can play an
essential role in the mechanism of proton decay.
Moreover, starting from the nonlocal stochastic electrodynamics construc-
ted by the hypothesis of stochastic space-time, it has been shown above that
the introduction of the fundamental length into the quantity of spectrum
density for zero-point electromagnetic field reduces the restriction on the
density of matter to be u, ~ n- 2 /- 4 h/c. From this, the conclusion can be made
that the existence of a superdense object may expose the space-time structure
in the general theory of relativity (see also Ginzburg, 1975). In this connection,
it seems to me, that study of the properties of the proton deserves great
attention. This is related too to the fact that there exists some indication
(Miettinen and Thomas, 1980) of the possibility of the existence of a high
density matter core (nucleus) in the proton with respect to its charge density.
370 Chapter 11

Furthermore, according to our present structure concept of matter, the proton


represents a superstable compound object from which our world is built.
Our basic assumptions are the following:
(i) Fundamental length exists in Nature.
(ii) The mechanism of proton decay is highly sensitive to the value of
fundamental length.
(iii) The structure of space-time appears in the phenomena of the micro-
world through an oscillation and acceleration mechanism of particles in
space-time (the acceleration mechanism will be discussed in Section
11.7).

Now we consider the oscillation mechanism of particles, say a neutron. Our


approach is based on the new method of gravitation proposed by de Alfaro,
Fubini, and Furlan (1980). Let us consider one component the scalar field
q>(x). The invariant action of q>(x) interacting with gravitation is

A = f xvf-g
d4 .ct(x), (11.25)

where .ctE = -R/4 and .ctM = tgl'V ovq> 0l'q> are the Lagrangians of gravitation
and matter field q>(x), respectively. Here gl'V is the metric tensor. The total
Green function for the scalar field q>(x) coupled with gravitation is
(11.26)

f
SM = Texp{i d4 xvf-g .ctM(X)}.

Following the idea of de Alfaro et al. (1980), we study the gravitation effects
in the flat space limit
(11.27)
where [yI'V is the Galilean metric and I is some universal constant of the
dimension of a length. We call it the fundamental length. For the present we
do not identify I with Igr determined by de Alfaro et al. (1980):
(11.28)
We notice again that there exists an enormous range between experimental
restrictions lexp ;:S 10- 16 cm (see Chapter 1) and Igr which must be investigated
in physics. It is quite possible that starting from some small distances I but
I ~ Igr the space-time structure may play an important role in particle physics
processes in the microworld. Such a possibility can be interpreted formally as
the appearance of some strong gravity with coupling constant Gs = Fc 3 /4nh.
Notice that it is necessary to make a change q> ---> mq> in expression (11.26) at
the same time with the transition to limit (11.27). Then the S-matrix for q>(x)
Physical Consequences of Hypothesis of Stochastic Space-Time 371

coupled with gravity takes the form

SM = T exp {it m2 F f d4 x oJlV ov<P OJl<p}- (11.29)

Making use of the modified Wick theorem (see Chapter 2 and Bogolubov and
Shirkov, 1980) we obtain from (11.26) and (11.29)

where

,10(X) = i- 1 (2n)-4fd4p 2 e- i: X •
m - p - 18

is the usual Green function for a free scalar field <p(x) and

,11 (x) = i- 1 (2n)-4 f d 4p p2 e- ipx (m 2 - p2 - i8)-2, (11.31)

etc. Here (m 212)n ,1n(x) (n = 1,2, ... ) are terms of higher order with respect to
,1o(x) which are due to the gravitational interaction of the scalar field <p(x). It
is interesting to calculate a contribution due to gravitation to the Yukawa
potential of two scalar particles coupled with gravitation. In the static limit
this potential is given by (11.31):

~Ff'
<p (r) = - ~F ~(2
- d 3pe-·pr p2(m 2 + p2)-2 = - ~- 1 - mr)e- mr , (11.32)
9 (2n)3 r 8n

Setting 1= Igr' we obtain a modified Newton potential

<pg(r) II = 19r =
Ggrm2(
-- mr) exp( -mr).
- 1- 2 (11.33)

We recall that the Yukawa potential of two scalar particles is given by the first
term of (11.30) (in the static limit):

<py(r) = (2n)-3 d 3 p f e-ipr


= _e- mr•
1
m2 + p2 4nr
We see that sum of the potentials <pg(r) and <py(r) is finite at the point r = 0 in
the case of the Compton wavelength Ie = h/mc.
Strictly speaking, the expressions (11.32) and (11.33) for potentials (and the
concept of potential at all) are valid at distances r ~ I and r ~ Igo respectively.
We see that the real corrections to the physical processes of the microworld
due to space-time structure (i.e., gravitation effects) are very small, of the order
m212 or even m414.
Now we consider a mechanism of proton decay within our hypothesis. For
this purpose we define first a generalized concept of particle oscillation. A
motion of a particle from one point to another in the space may be formally
372 Chapter 11

considered as repeated (multi) oscillations of a particle. The matrix elements


for transitions-into-itself of the particle coupled with gravitation during its
time evolution, are given by
M(p -> pi) = <Ola-(p)SMa+(p')IO) ~ (5(4)(p - p')(1 + im 2 [2).
Then, by definition, the oscillation time is
r ~ 1/[m(1 + m4[4)] = (1 - m414)/m.

The oscillation period of a free particle is understood as the time during which
this particle travels (or succeeds to go through) a distance equal to its
Compton wave-length, i.e., r = A/e, A = h/me.
Such a formal definition of the oscillation period is generalized easily to a
more general case, when the particle during its time evolution in space
accomplishes transitions to completely different particles. For example, n -> n,
Ve -> vI' -> Vt and even n -> nO, n -> KO, etc. For description of such strange
transitions it is necessary to introduce the concept of a superfield (concerning
superfields, supersymmetry and supergravity see the review by Mezincescy and
Ogievetsky, 1975; the book edited by Van Nieuwenhuizen and Freedman,
1979) responsible for the above-mentioned transitions. The particles (and their
antiparticles) with different spins and masses, etc. enter a superfield multiplet.
For example,

'I's = L «(J.i!.pi + f3i M - 1/ 2 t/1J,


i

where !.pi and t/li are boson and fermion field operators, respectively, and (J.i and
f3i are some real numbers (generally speaking, they are Grassmann variables).
The appearance of the factor M i- 1 / 2 in the second term of the superfield 'I's
follows from a dimensional argument. A simple Lagrangian for the superfield
coupled with gravity is

(11.34)

Then the total Green function of the type of (11.26) for the transition n -> n
takes the form
Dnn(x - y) = <01 T[t/I,,(x)Vln(y)SsJIO) <OISsIO)-l

(11.35)
Physical Consequences of Hypothesis of Stochastic Space- Time 373

where M p = mn and

~sn(x - z) = <01 T['¥s(x)tl/n(z) 10)


Dr
= i- 1 M;1/2 f
~)4
(2n
d4 e-ip(x-z)

mn - p -
A ••

IE

Here we have assumed f3n = f3ii = 1. Note that since, in a free state, direct
transitions between nand ii are impossible the term ~",i(X - y) of the type
~o(x - y) in (11.30) does not appear in expression (11.35) for the nn-
transitions.
In the case of neutron oscillations it is easy to see that the oscillation period
for the nn-transitions is
(11.36)
We see that if the n -+ n oscillation exists, then it is wholly responsible for the
space-time structure (i.e., gravity). The neutron oscillation period 'nii must be
~ 10 30 years, since it would otherwise contradict the stability of matter, as was
determined experimentally (Reines et al., 1979; Reines and Crouch, 1974; see
also the review by Goldhaber and Sulak, 1981). This is because, if the
oscillation mechanism in the presence of gravity changes the neutron into an
antineutron, the n can be destroyed with other nucleons in matter into
multi pions resulting in an effective-bar yon-number nonconserving decay. For
example, (A, Z) -+ (A - 2, Z) + n+n-.
Let us obtain restrictions on the va:lue of the fundamental length I. From
inequality 'nii ~ 10 30 years and (11.36) it follows that I;;S 10- 29 cm. On the
other hand, according to our semi-empirical approach (see previous section)
the next possible candidates for quantities of I are 18 ~ 1.8 x 10 - 31 cm and
19 ~ Igr = 5.72 X 10- 33 cm. Corresponding to these quantities of I the proton
life-time is ,; ~ 10 38 years and ,~ ~ 'gr ~ 1043 years, respectively. The last
possibility means that the proton decay is wholly due to the gravitation
interaction with the Newton constant Gg • In particular, the true value of the
fundamental length I ~ 10 - 33 cm results from our consideration carried out in
Section 11.7. So that the value, p ~ 10 43 years is more probable.

11.6. A Hypothesis of Nonlocality of Space-Time Metric and its


Consequences

In this section we consider some consequences of introducing nonlocality into


the metric of space-time given by formulas (1.4) and (1.12) in Chapter 1. We
again recall that introduction of the proposal about the stochasticity of space-
time inevitably leads to the change of the space-time metric, say (1.4):

(11.37)
374 Chapter 11

where T = b4 . A change of this form of metric was discussed by Markov (1959)


and Takano (1961) too. But there is another possibility
So = x~ - XZ = Sl = <[(xo + T)Z - (x + ib)Z] >~4(x)

= f d 4bEWcn[(X o + T)Z - (x + ib)Z] = So + [z. (11.38)

Now we consider another case when the space-time metric is determined by


the formula ds 2 = g/lv(x) dx/l dx v • In such a curved space-time we have the
following expression, according to (1.12):
ds z = g/lv(x)dx/ldx V = (dsDZ = G~v(x)dx/ldxV, i = A,C;

~v(x) = fd4bEWCn9/l.(xo + iT, X + b) = fd4YKA(X - y)g/lv(Y), (11.39)

G~.(x) = fd4bEwC~)g/lv(xo + T,X +ib) = fd4YKC(X - y)g/lv(Y),

where KA(X) and KC(x) are some distributions of type (1.11), Fourier trans-
forms of which are given by

_
KA(_pZ[Z) = (2n)2 Jorro drrzw (r2)
r ll() _r2pZ)(_p2)-1/2, (11.40)

j(C(pZ[Z) = (2n)2 Lro drr2wC:)ll()r2p2)(p2)-1/2, (pZ = p~ _ p2) (11.41)

Furthermore, as usual, we shall investigate only the class of distributions -


measures w(biIZZ) for which the form factors (11.40) and (11.41) will be entire
analytic functions of the variables z = p2[2 with a finite order of growth
00 > p ~ t. Then the metric tensor G~v in (11.39) can be rewritten in the form

G~v = Ki(l20)g/lv. (11.42)


For KiUZO) the representation of type (1.11) holds.
From expressions (11.40) and (11.41) it follows that the functions j(A and KC
have a completely different behavior in the limit pZ ...... ± 00, i.e., in the pseudo-
Euclidean and Euclidean directions, respectively. In order to present clearly a
general characteristic of this behavior, we now shall calculate these functions
for a specific form of the measures w(bil[2) under consideration (see Chapter
1). Let
1. wG(YZ) = 4n- 2[-4 exp( _2y2), 0 «y < 00,

"1 -2[-4y -Z(l _ y Z)-l/Z , 0<y < 1


Z) { 2n
2. w 1 (Y = 0 y ~ 1
Physical Consequences of Hypothesis of Stochastic Space-Time 375

3.
Z
wz(y ) -
_{n-0 I-
Z 4 (1+/l)(2+/l)(1- YZ )1' /l>-1, O<y<1
' y :;,. 1

Then the form factors j(A and j(c corresponding to these measures are equal
to
(11.43)

(11.44)

(11.45)
Here the upper (lower) sign corresponds to j(A(j(C). Whence, it is obvious that
the functions j(A,C have a different behavior, say
at pZ ~ + 00,
at pZ ~ -00,

j(C( z[2) = {O(eXP(IJi?T)) at pZ ~ - 00,


1 p O(1/lpZI) at pZ ~ + 00.
Such a different behavior of the form factors j(A,C plays, as seen below, an
important role in constructing a finite theory of interacting quantized (test)
fields from the hypothesis of nonlocality of the space-time metric point of view.
Now the following questions arise: How to realize the abovementioned idea
of nonlocality of the space-time metric in physical applications and what is a
physical quantity which represents a 'carrier' of this nonlocality? We shall
discuss these questions in detail. Indeed, it is quite possible that in Nature
there exist exclusive exotic objects (we call them test particles) with respect to
the space-time structure at small distances. These test particles, if they exist,
would give rise to a transition mechanism from the nonlocality of the space-time
metric to fluctuations in it, and, therefore, would act as a resonant factor in the
process of the creation of elementary particles by some perturbation (or
fluctuation) of the space-time metric (see below). However, the question of the
possible existence of such objects requires, it seems, a more fundamental
physical principle connected to the quantum theory of gravitation, which is
not clear in the present formulation. We assume that as such an object (test
particle), black holes (small) and other hypothetical particles like quarks, bags
(for example, MIT bag), monopoles (Dirac's, 't Hooft-Polyakov's and GUT's
super-heavy), maximons (particles with the Planck mass mpl ~ 1.2 x 10 19 GeV),
etc. may be taken. The role of these test particles in the process of the creation
of elementary particles and their specific differences from usual local particles
will be discussed below.
Further, we postulate that the square-energy-momentum operator of par-
ticles is constructed by means of the metric tensor G~v of the type of (11.42)
376 Chapter 11

and acquires the form:

A2 #2 Gi [P (11.46)
p =:> = - ltV axil ox v '
and in the flat space case one can assume that
()b2 _ _ i 2 02 (11.47)
';7 - K (l O)gl'v axil ax'"
where gl".~is the Mink ow ski metric. We assume that such a form of the
operator 1}2 results from the hypothesis of nonlocality of the space-time metric
at small distances or in a singular region of confinement of particles which are
represented by solitary waves (Efinger, 1981). In the last case one could then
hypothesize that particles under consideration are geometrical objects on a
Riemannian space-time metric gllv(x) which is nearly singular (or nonlocal) in a
region corresponding to the width of the solitary wave-amplitude (or in a
region of confinement of particles).
Moreover, the form we have proposed of operator (11.46) or (11.47)
naturally generalizes the idea of quantization of three-dimensional space
according to Fujiwara (1980). His scheme is derived as a particular case for
our consideration. Indeed, the new momentum operator #[ according to
(11.46)] constructed by the generalized function KCf, the Fourier transform of
which is determined by the form factor (11.44) in the static limit Po --+ 0,
coincides explicitly with the momentum operator proposed by Fujiwara
(1980). The difference in both the schemes is that in our model the relativistic
invariance is strictly conserved.
Thus, our postulation concerning the nonlocality of the space-time metric
leads to a profound connection between the particle-wave dualism and the
structure of space-time. In other words, such a structure of the space-time-
stochasticity or nonlocality (after averaging over a large scale) as a self-
memory would affect the Q)-k relation for the waves and E-p relations for
particles and makes the theory nonlocal. Notice that in the three-dimensional
case this problem has been discussed by Fujiwara in detail.
Now we consider some consequences of the non local theory of quantized
fields based on the assumption of the nonlocality of the space-time metric.
Since the important physical quantities such as the Hamiltonian (or
Lagrangian) and field equations of test particles are constructed by means of
the operator (11.46) or (11.47), these quantities are changed in accordance with
the change of the metric forms (11.42). For example, the field equation and the
Green function for a scalar field acquire the form:
(fj>2 - m2)cf>(x) = [KiUZO)O - m2]cf>(x) = 0, (11.48)

G(x) = i- 1(2n)-4 fd 4p e ipX [m2 - p2 /(A,C(p2[2)r 1, (11.49)


Physical Consequences of Hypothesis of Stochastic Space-Time 377

Here the Green function G(x) satisfies the following equation


Uj2 - m2)G(x) = -c5(4)(X).

The contour integration in (11.49) is chosen as in the case of local theory, so


that in the limit I ~ 0, the value of G(x) gives the usual Green function for the
scalar field with mass m. Similarly, in our model the propagators of photons
and spinor fields read:

D~Jx) =
f eipx

-i(2n)-4 g /lv d 4 p p2j{A,C(p2ZZ) , (11.50)

and

S( ) = '-1(2 )-4fd4 ,px m + p[j{A,C(p2/2)]1/2 (11.51)


eX I n pe m 2 - p 2K- AC
'P
( 2/2)

Now we discuss the choice of the form factors j{A or j{c in (11.40) and
(11.41), It is well-known that (see Section 2.3) the introduction of nonlocality
into theories in accordance with (11.46), (11.49)--(11.51) changes the form of
potentials between interacting fields (for example, the Coulomb and Yukawa
laws) at small distances. For example, the Coulomb potential of two interacting
electrons is given by expression (11.50) in the static limit:

rp(r) = f
e(2n)-3 d 3p p2j{A~~i:p2/2r (11.52)

From the condition rp(O) < 00 it follows that only form-factors of the type
j{C(p2{2) given by expression (11.41) are acceptable. In particular, for form-
factor Kg (11.43) the equality (11.52) reads

rp(r) = e(2n)-3 fd 3p p-z exp( -ipr - p;ZZ) = (4:r) erf(V;r)-


One verifies that for the form-factors j{C(pZ/2) the propagators (11.49)--(11.51)
are finite in the Euclidean metric. Thus, in our model all matrix elements of the
S-matrix are studied, as in the case of the nonlocal theory of quantized fields
(see Chapters 1 and 2), in the Euclidean domains p; < 0 of all the momentum
variables Pi' The passage to the case p; > 0 is done by an analytic
continuation.
A concrete scheme for the quantization of fields and the construction of a
nonlocal theory of electromagnetic and weak, etc. interactions with the use of
the propagators (11.49)--(11.51) will be presented in a separate work. Actually,
in our scheme the ultraviolet divergences are absent.
We now consider the more interesting consequences of the idea of the space-
time metric nonlocality, First, it transpires that in our scheme the Einstein
relation between the energy and momentum of a particle is changed. For this,
consider a simple case when the metric is defined by formulas (11.37) and
378 Chapter 11

(11.38). According to formula (11.42), these equalities can be written in the


unified way
(11.53)
where a is some number depending on the form of measures w(b~/12) and
taking two values, positive and negative, in accordance with the choice of
metric formulas (11.37) and (11.38) [or the form of form factors (11.40) and
(11.41)]. Then the equality '!J2 = m 2 from equation (11.48) in the given (free)
case yields the following equation
p2 _ m2 _ aF p4 = 0,
that has two solutions:
(11.54)
Here it is assumed that ml ~ 1. If a > 0, then both the solutions are positive,
which corresponds to the appearance (or creation) of a supplementary particle
with mass of an order of 1/a1 2 - m2, at which the initial mass m acquires a
small additional term of an order of m2 [2 resulting from the nonlocality of the
space-time metric. If a < 0, then a particle of the tachyon type corresponds to
the second solution, i.e., a particle with complex mass.
From the first equation (11.54) it follows that
E, = [p2 + m2(1 + am 212)]112 ;:::; Eo + (1/2Eo)am412,
Eo = (p2 + m2)112.
In particular, for the rest energy we have
EJ = me 2 (l + tam 2(2)
and in the classical limit
2
Eel, = ~
2m (1 - lam 2F)
2 •

Further, for the concrete form of form factors, say for KS given by (11.44) we
have
P6 = p2 + 41- 2[arcsin (tmlW or E, = [p2 + 41- 2(arcsin (tml))2rI2.
In view of this we notice that the test of Einstein's formula E = me 2 may be of
a great interest.
Another exceptionally important fact is that the space-time metric non-
locality, as can be seen from expression (11.54), admits the existence of a
corresponding partner with mass 1/a12 - m2 for any particle with mass m, even
in the case of the flat-space-time-metric nonlocality (11.53).
It turns out that the equidistant behavior of observable mass spectra of the
elementary particles may be understood within our hypothesis of the space-
time metric nonlocality, if the parameter 1 used here is assumed to be an order
Physical Consequences of Hypothesis of Stochastic Space- Time 379

of the size of the domain in which the quark confinement is reached, i.e.,
l-+ leff ~ 10- 13 cm. On the other hand, this quantity may be identified with the
effective Schwarzschild radius of the proton (see Li, 1982; Salam and Strathdee,
1976). As mentioned above (Chapter 1), in our scheme such a possibility is
based on the assumption that the parameter l is not a universal fundamental
constant but characterizes only the domain of the non local interaction of test
particles. In connection with this we notice that there is an another interesting
approach (Brooke et aT., 1982) in which elementary particles are regarded as
quantum space-time excitons within Born's reciprocity principle (Born, 1938).
Now we give some speculation concerning the possibility of obtaining the
equidistant behavior of the observable mass spectra of elementary particles. It
should be noted that in Nature there exist two basic different objects with
respect to the space-time description. Those of the first sort behave as point-
like particles and are described by the local quantum field theory, i.e., their
equations of motion satisfy finite order differential equations (for example,
Klein-Gordon and Dirac equations). In this case, according to the above-
deduction, the connection between the properties of these particles and of the
space-time metric nonlocality is expressed by the change of the energy-
momentum relation at small distances, in particular, for the rest energy of the
particles we have obtained
E? = mc 2 (1 + tam 2 (2).
Here we assume m2 l 2 ~ 1 for any free point-like observable particles and
lal ~ 1 for the measures (distributions) W;(y2) listed above.
The situation is radically changed for the extended (nonlocal) objects. A role
of such objects, as mentioned above, may be played by bags, strings, mono-
poles, different compound states of quarks and gluons etc. For definiteness,
we have called them above test particles. The features of these test particles are
the following: first they are described by integral (strictly speaking, integro-
differential) equations, i.e., differential equations of infinite order; secondly,
they have some internal structure and those physical quantities (charge, mass,
etc.) are characterized by some distributions. Namely, the presence of the
internal structure and distributions for the test particles plays an essential role
in their space-time description. In this nonlocal case, we should distinguish
between the internal and external mediums (in particular, space-time) around
the nonlocal objects and take into account their structures. The interrelation
of these mediums or their reaction on nonlocality with respect to the change of
space-time properties may be expressed by means of the concept of the
perturbative effect of test particles on the space-time metric non locality. Our
basic assumption is that this perturbative effect is described by some differen-
tial equation of infinite order and gives rise to equidistant behavior of
observable particle mass spectra, if we would regard the observable particles as
some definite states (excitons) of excited test particles. Below we consider a few
excited test particles which generate spectra of masses of observable particles.
380 Chapter 11

In the case under consideration, the parameter I and the measure w(b~/[2)
characterize the physical properties of the perturbative effect of test particles
on the space-time metric nonlocality. Roughly speaking, now the parameter I
loses its universal character and may be understood as an effective size leff of the
test particles (henceforth by I we denote just this value leff in this section). The
measure w(bUI 2 ) corresponds to an energetic (or mass) distribution of excited
test particles. Further, we assume that the effective size of the considered test
particles are comparable with the size of the domain in which the quark
confinement is reached, i.e., I -4 leff "-' 10- 13 cm. Therefore, in the test particle
case, M21;ff "-' 1 and E? = Mc 2, where M is the characteristic mass of test
particles.
Thus, we assume that basis (ground) states <PB and I/lF of test particles for
bosons and fermions satisfy the following differential equation of infinite order

and their eigenvalues are determined by the formulas


p2KB(p21~) - m~ = 0, p2KF(p21;) - M~ = 0, (11.56)
where I -4 IB and IF denote the proper radii of test particles for bosons and
fermions, respectively. The higher order equations (11.55) are reduced to the
Klein-Gordon and Dirac equations as IB,F -40. Quantities mo and Mo cor-
respond to some lowest (or unperturbed) values of the masses in elementary
particle mass spectra for bosons and fermions. On the other hand, equations
(11.56) can be understood as conditions for the quantization of masses.
As is seen above, only form-factors of type (11.41), say, like KC's in (11.43)-
(11.45), can describe the equidistant behavior of the elementary particle mass
spectra. It turns out that the parameters m o , M o , IB' IF and the form of form-
factors Kf(p 2 If) (i = B, F) are sufficient to obtain some theoretical mass
formulas which roughly correspond to the equidistant behavior of the spectra
masses of observed particles. For this, consider the simple cases (11.44) and
(11.45). The first equation of (11.56) in the case of (11.44) takes the form
m~ - (211~)[1 - cos (p2~)] = 0.
Assuming mo = filIB' we get the quantized condition cos(p21~) = 0. From
this we have a family of particles:
mn = (p;)1/2 = (nIIB)(t + n), n = 0, 1,2, ...
In the second case (11.45) the second equation of (11.56) yields

M~ - 4.21"·r(3 + f.l)p2(jP2i[)-2-I"J 2 +lJiW) = 0.


°
Setting M 0 = here we obtain the following condition for the fermion mass
quantization J 2 + l"(p21;) = 0, that corresponds to some family of particles with
Physical Consequences of Hypothesis of Stochastic Space Time 381

masses:
Mn = (l/IF )jz + I',n' n = 1,2""
where jz + I',n are zeros of the Bessel function J 2 + I'(x),
We assume that the basic (initial) metric perturbation gives rise to some
family of particles, but newly created particles of this family may in turn
generate the space-time metric, and in consequence, that there are possible
second (daughter) metric perturbations which are able to generate other
families of particles, etc, Such a process of metric perturbation is analogous to
the chain reaction of atomic nuclei (or analogous to the cosmic tree diagrams
according to 't Hooft, 1979), Of course, the mechanism of perturbations (or
fluctuations) of space-time metric proposed here differs from the quantum
gravitational concept of fluctuations in metric, this should be on length scales
which are far smaller than any at present probed, As a result of such
consecutive perturbations, the observed equidistant behavior of the particle
mass spectra may be explained, It appears that for a satisfactory description of
the equidistant behavior of the particle mass spectra, it is sufficient to know
two basic metric perturbations connected with the generation of the 1!-meson
and proton and their corresponding partners, Next, a few second (daughter)
metric perturbations are needed which arise from the K-meson and A-, L-, 3-
hyperons generated already by the initial metric perturbations,
So, assume that two metric perturbations are defined by the following
formulas:
1! 1 c
m(1) = _(1.
I 2 + n), - =- m(ol) = 90 MeV n = 0, 1,2, ' , ,
n
B IB h '

1
1= 36.138 MeV, n = 1,2, ...
F

where j201 n denotes the nth positive zero of the Bessel function J of the order
20t (Olv~~ (ed.), 1960). The corresponding theoretical spectra of particle
masses are presented in Table 11.3 (the experimental data were taken from
Particle Data Group, 1982).
Daughter metric perturbations caused by the K-meson and A-, L-, 3-
hyperons are given by
M~i) = (1/1JJ 20 1 n' i = 1,2,3,4,
"2'
where 1/11 = 19.07 MeV, 1/12 = 43 MeV; 1/13 = 46 MeV; 1/14 = 50.66 MeV; for
K-, A-, L- 3-family of particles, respectively, and the corresponding theoretical
mass spectra are shown in Table 11.4.
Since masses of the K-meson, proton and A-, L-, 3-hyperons are equal in
the order of magnitude, we assume that their form factors must be the same
and the corresponding values of the parameter 1/1 must grow with increasing
particle mass. It is natural because the generation of a particle with a larger
382 Chapter 11

Table 11.3

n m~l) Experimental m~2) Experimental


values values

0 141 n
424 K 938 p
2 707 ~p 1103 A
3 990 8*(975) 1249 1:
4
N~1440)
1272 { 1(1270),'1(1275) 1386
D(1285) {
5 1555 p'(1600) 1518 N(1520)
6 1838 4>(1850) 1647 n-, N(1650), N(1675), N(1680)
7 2120 n(21 (0) 1773 N(1700), N(1710), N(I720)
8 2403 .5(2450) 1897 N(1990)
9 2686 2020 N(2080)
10 2969 '1,(2980) 2141 N(2190)
11 3251 1/1(3100) 2662 N, N(2200), N(2220), N(2250)
12 3534 { 1/1(3415) 2381}
1/1(3510),1/1(3555) ~(2380~~(2450)
13 3817 1/1(3685),1/1(3770) 2501
14 4100 1/1(4030),1/1(4160) 2619 N(2600)
15 4382 1/1(4415) 2737 ~(2750), N(2700)?
16 4665 ? 2855 ~(2850), N(2800)?
17 4948 ? 2972 ~(2950)?
18 5230 B(5200) 3090 N(3030)?
19 5513 ? 3206 ~(3230), N(3245)?
20 5796 ? 3323 ~(3350)?

33 9472 Y(9460)
34 9755 ?
35 10037 Y(10020)
36 10320 Y(10350)
37 10603 Y(l0570)

285 80723 mw
290 82137
295 83551
320 90619
325 92033 mzo

mass requires a larger space-time 'curvature' R = 1/1 or smaller l. Further we


notice that the choice of the form-factors (11.44) and (11.45) describing the
equidistant behavior of the particle-mass-spectra in the given case need not be
unique, and other forms of the metric perturbation (correspondingly, of the
measure w(b~/12) or the chosen form-factor) are possible. Thus, in our ap-
proach the fundamental open problem is the question of a possible unique
choice of form-factors, which requires, it seems, another fundamental physical
Physical Consequences of Hypothesis of Stochastic Space-Time 383

Table 11.4

n m~K) Experimental n m~) Experimental


values values

495 K
2 582 '1 m~i\)
3 659} 1 1115 A
4 732 P 2 1312 A(1405)
5 801 w 3 1486 A(1520)
6 869 K*(892) 4 1650 A(1600, 1670, 1690)
7 938 1'1'(958) 5 1807 A(l800, 1820, 1830)
8 1001 0(980) 6 1960 A(1890)
9 1066 <1>(1020) 7 2110 A(2100, 2110)
10 1130 H(1190) 8 2257
11 1193 B(1235) 9 2403 A(2350)
12 1257 p' (1250), A 1 (1270), 10 2548 A(2583)
Ql(1280)
13 1320 { n(1300), A 2(1320)
£(1300) m~I;)
14 1382 k(1350) 1194 ~
15 1444 { Q2(1400), K'(1400), 2 1404 ~(1385)
K*(1430),E(1420) 3 1590 ~(1560,1580, 1620)
16 1507 1'(1515) ~(1660, 1670)
17 1569 L(1580) 4 1765 ~(1750, 1775)
18 1630 8(1640), K*(1650) 5 1933 ~(1915, 1940)
{ w(1670), A 3 (1680) 6 2096 L(2030)
19 1692
<1>'(1680), g(1690) 7 2257 L(2250)
20 1753 L(1770), K*(1780) 8 2415 L(2455)
21 1815 X(1850) 9 2571 L(2620)
22 1876 D(1870) 10 2725
23 1937 8(1915)
{ D*(2007,2010), m~3)
24 1999
0(2030), F(2020), F(2021) 1 1315 .=.
{ h(2040), p(2050) 2 1546 2(1530)
25 2060
K*(2060) 3 1751 2(1820)
26 2121 F*(2140) 4 1944 2(2030)
27 2181 p(2150), £(2150)
28 2242 K*(2200), p(2250)
29 2303 £(2300)
30 2364 p(2350)

principle. As pointed out above, formulation of this principle is not known and
needs deeper studies in field theories. Moreover, the mechanism of metric
perturbation (allowing one to create the particles) resulting from the space-time
metric nonlocality is not clear.
In conclusion, notice that the mass splitting connected with the internal
quantum numbers such as charge, isospin, charm, etc. would not be surprising.
In our model, we do not consider these quantum numbers, and consequently,
384 Chapter 11

the metric perturbation is not sensItIve to the changing values of these


numbers. Therefore, we associate some mass values of theoretical calculation
with experimental values that are closer to them.
From Table 11.3 it follows that if the equidistant behavior of the mass
spectra caused by the nonlocality of space-time metric is valid, then our
scheme predicts the existence of a family of bOSOliS with masses: 2700, 4700,
4900, 5500, 5800, ... , 9750 MeV, etc.
Finally, it should be noted that an attempt to explain the equidistant
behavior of observed particle mass spectra in the framework of our proposal
listed above, is indeed not of principle interest, but it plays an important role
when it deals with hypothetical particles (for example, maximons by term of
M. A. Markov) of mass of the order of the Planck mass M p /. In this case, it
seems to us that mass spectra of these maximons, if they exist, should be
caused by fluctuations in metric and quantized according to the above
deduction, and obeyed also the equidistance behavior as in the case of usual
particles observed recently.

11. 7. On the Origin of Cosmic Rays and the Value of the Fundamental
Length

The problem of the origin of cosmic rays plays an important part in high-
energy astrophysics. To solve this problem one should answer a set of
questions (see Ginzburg and Ptuskin, 1976) among which the acceleration
mechanism is of great interest. If the main sources (their spatial distribution) of
cosmic rays are pointed out and the intensity and spectra of cosmic rays
emitted by the sources are given, the problem of the cosmic-ray origin can be
divided into an external (acceleration mechanism) and an internal (source
theory) problem. The former, i.e., the acceleration mechanism that carries
cosmic rays (especially, protons) to the energies over 10 2 °_10 22 eV extent in
the primary cosmic radiation still remains unsolved (Ginzburg and Ptuskin,
1976; Hillas, 1975).
Here we discuss this important problem in the framework of the hypothesis
of space-time stochasticity and fluctuations in metric (see Chapter 1). It is well
known that a high isotropy and a rather larger content of secondary nuclei in
cosmic rays indicate an effective 'mixing' and a long wandering of high-energy
particles in the Galaxy. We assume that such a mixing and isotropization are
caused by the stochastic (or fluctuational) structure of space-time in which the
cosmic-ray propagation takes place. More exactly, we suggest that the physical
mechanism responsible for the acceleration of cosmic rays during their pro-
pagation in the Galaxy has the stochastic origin connected with the space-time
properties at small distances.
As shown above, the idea of using the stochasticity of space-time and
random fluctuations in metric as an origin of the stochastic and non local
Physical Consequences of Hypothesis of Stochastic Space- Time 385

interpretation of quantum mechanics plays an important role in constructing


the nonlocal theory of quantized fields. Now we obtain the dynamics of
cosmic-ray particles within the hypothesis of fluctuations in metric. The basic
idea is the following: there exists a profound connection between the structure
of space-time and the propagation mechanism of ultra-high energy particles.
To realize this connection we assume that freely moving particles perturb space-
time around themselves, and at the same time the space-time metric is generated
and becomes a Riemannian one ifl'v(bE, x) depending on a random vector
bE = (b 4 = r, b) with a measure w(bi/12) obeying conditions (1.2) in Chapter 1.
As a result of such fluctuations in metric, as shown below, a nonlinear
dynamics of particles arises and the maximum energy of particles, say protons
in the primary cosmic-rays depends on the value of the fundamental length at
which fluctuations in metric take place. This allows one to define the value of
the fundamental length from high-energy astrophysical experiments.
A simple form of ifI'V is chosen, that is ifI'V = ¢2(b~, x 2)1] 1'"' and the averaged
space-time metric acting on the behavior of particles is obtained by the formula

gl'vUZ/ X2 ) = <if"v(bE, X)~4(x) = f d4bEw(b~/12)¢2(b~, X2)1]I'V

(11.57)
where 1]I'V is the Minkowski metric.
Now we pass to obtaining cosmic-ray particle dynamics in this conformally

r
flat space-time. The action for a free particle takes the form

S = -mc ds, (11.58)

ds = [gl'vUZ/X2 ) dxl' dxv]1/2 = ¢ds o, ds o = (dx vdxV)1/2.


According to the action principle

6S = -mc ib 6ds = -mc


ib 6ds 2
- - = -mc
ib 6(gl'vdxl'dxV)
a a 2ds a 2ds

= -mc
r {I
Ja 2
b dx v dxl' og dx v d6XI'}
cis ds oxv: 6dx " + gvl'cis ~ ds = O.

i
Integration by parts in the second term gives

6S= -mcg -
VI' ds
V
dx6x l' Ib -mc
a a
b
ds {I v
dx dxl' og d (
----------.!J'..6x"-~
2 ds ds ox" ds
g -dXV) 6x l' } .
VI' ds

From this we have the following nonlinear equation

Du v
Ds
_
= g"vTs
du"
+ r v.I'''u I'u" -- 0
,
386 Chapter 11

or taking into account that g/lV = ¢ 2rt /lV' ds = ¢ ds o , we get


du~ /ds o - a In ¢/ax v + u~(u~a In ¢/ax A) = o. (11.59)
Here ub = dxv/ds o ,
rV,/lA = t(ag/lviaxA + agAviax/l - ag/l;./ax V)
is the Christoffel symbol.
In the nonrelativistic case equation (11.59) will be discussed in the next
section. We now solve it in the two-dimensional Minkowski space-time (x, t)
and in the simple case when ¢(x) depends only on the spatial variable x (now
the variable x is regarded as a parameter). Since u~ = -vAl - {32)-1/2,
ds o = cdt(l - {32)1/2, {3z = v;/c 2 , from (11.59) we get (v x = v):

dq
dt
2) ( = ~ = -c aIn
= e (1 _ q, q c' e ax
¢) . (11.60)

This is the Riccati-type equation (for example, see Bender and Orszag, 1978).
The substitution q = W/eW, W= dw/dt converts it into a second-order linear
equation for w(t): W - e 2 W= O. The solution to the last equation is
w = C 1 ch(et + c 2 ), and therefore q = th(et + cz ). Here the integration constant
C2 is given by an initial condition, say q(O) = vo/c.
So, the solution of the Cauchy problem for the Riccati-type equation (11.60)
IS

q(t) = th(et + Arth(vo/c», (11.61)


or
q(t) = e sech 2 (et + Arth(vo/c».
The next question concerns the choice of the metric form g/IV(F/X 2), i.e., the
function ¢W/X2) in conformally flat space-time. A simple form may be
obtained if one assumes that:
(i) Due to the small parameter I it is sufficient to determine the metric
form g/lVW/X2) in (11.57) by the few first moments of the random
variable bi
of the function ¢Z(bi/IZ).
(ii) According to an old concept by Einstein (1967) and recently by Efinger
(1981) the averaged Riemannian metric (11.57) may be remarkably
peaked (even nearly singular) in the region of the confinement of a
particle and outside this region one expects
g/lV ~ rt/lV = diag(l, -1, -1, -1)
(or the boundary condition g/lV --> rt/IV at infinity).

(iii) 9/lV --> rt /lV' for I --> O.


(iv) The size of the domain of particle confinement corresponds to the
Physical Consequences of Hypothesis of Stochastic Space-Time 387

particle radius r which may be interpreted as the effective Schwarzschild


radius (Li, 1982; Salam and Strathdee, 1976), for example, the experim-
ental value of the proton radius is r = rp = 0.8 x 10 -13 cm. Thus, the
simple form
(11.62)
satisfies the above-mentioned conditions automatically.

The perturbative effect of the particles on the metric leads to fluctuations


such that the deviation from '1I'V has a nonzero mean. In view of (11.61) and
the assumption (11.62) the particles are accelerated to large velocities v ~ c
during the evolution of the Universe. At this, the energy of the particle
acquires the form
E => El = mc 2 (1 -lm- 1/2 , (11.63)
where IN = q2(t) is given by equation (11.61).
Thus, measuring the maximum energy of the particle (the largest proton
energy value observed in cosmic-rays is about E'l~p = 10 20 eV, see Hillas, 1974)
in primary cosmic-rays, we can determine experimentally the value of the
fundamental length from formulas (11.61) and (11.63). We assume that the age
of primary cosmic-rays Ter in the Galaxy is of the order of the age of the
Universe Tu = 1/2Ho = 8.85 x 109 yr. Recently model-dependent calculations
(see Ginzburg and Ptuskin, 1976) have shown that a 'large' Ter :<: 108 yr is
more possible. Here the value of the Hubble constant is Ho = 55 km/sec x
x (megaparsec).-1
The simple extremum problem for equation (11.61) in the case of (11.62)
gives the maximum value of q(t) at x = r/J3 (I ~ r). At this the argument at of
q(t) takes the form:

for r = rp ' where L1 = 1/10' 10 = 10- 33 cm. So, assuming t = Tu , Vo = 0,


r = rp = 0.8 x 10 -13 em, x = rp/ J3 and El = E'l.hp, we obtain from equations
(11.61) and (11.63)
E'l,hp = 10 20 eV = mp c 2 texp(10.68 L1 2 ).
From this it is easy to find that L1 = 1.56, therefore
1= 1.56 X 10- 33 em, (11.64)
which is very close to the Planck length Ipi = (G g h/C 3 )1 / 2 = 1.62 X 10- 33 em
discussed in the quantum gravitational theory, where Gg is the Newtonian
constant.
It transpires that the very slighest change in the value of the fundamental
388 Chapter 11

length from the quantity (11.64) (or Ipl ) gives rise to the enormous range of
energy of the cosmic-ray proton (see Table 11.5).

Table 11.5

1/10 1.56 1.62 1.69 1.82 1.93 2.04 2.14 2.24 2.34 2.43 2.51 2.60

Thus, we see that if the true fundamental length that exists in Nature is of
Ipi= 1.62 x 10 - 33 cm, then the maximum energy of the proton is restricted by
the value E~:':, = 7.1 X 10 20 eV. We note that another value T~ = 1/Ro for Tu is
discussed in literature, that gives 1= 1.11 X 10- 33 cm for EY~p = 1020 eV. In
this case E~:':, '" 10 3 3 eV.
Now we consider the motion of a charged particle in an electromagnetic
field characterized by a potential AI' within our scheme. After some calcu-
lations analogous to the usual (gravitational) case (for example, see Landau
and Lifshitz, 1971) we have the following equation

Du v e
mc-=-F
Ds c VP uP , (11.65)

where

Fvp = Ap;v - Av;p == oAp/ox v - oAvlox p.

It turns out that in our model the equations for the electromagnetic field
acquire the forms:

and

(11.66)

where r is the four-vector current satisfying the equation of continuity

We recall again that according to conformally flat space-time considered


above, the determinant of the metric tensor gI'V is calculated from equation
(11.57).
Further, we write the equation of motion for the charge in an external force
fl' in accordance with equation (11.65):

(11.67)
Physical Consequences of Hypothesis of Stochastic Space- Time 389

where fv has components:

in particular, for bremsstrahlung f = -rv, -r = 1eZc-3. Then, the equation of


motion in the case of v ~ c, cf> = cf>(x) has the fmm (in the absence of the
electromagnetic field):

(11.68)

or in the two-dimensional case


4= (-r/m)ij + 8(1 - qZ), (q = vic). (11.69)
This is an autonomous equation if it is assumed that t is an independent
variable and x is a parameter. Then a standard trick is to express u = 4(t) as a
function of q and to find an equation for u and its derivatives. We thus let (in
the two-dimensional case)
4= u(q), ij = u'(q)u(q), u'(q) = du/dq,
and for u(q) we have the following equation:
u(q) - (-r/m)u'(q)u(q) = 8(1 _ qZ).
This equation is still too difficult to solve in a closed form. This is easy to show
if one assumes that 8 ~ zz (x ~ 1) is a small parameter, then a solution is
sought in the form
u(q) = uo(q) + 8U 1(q) + 8ZUz(q) + ...
First two terms of u(q) satisfy the following equations:
uo(q) - (-r/m)u~(q)uo(q) = 0,
u1(q) - (-r/m)u~(q)uo(q) - (-r/m)u~(q)u1(q) = 1 _ qZ. (11.70)
Solutions of the former are

Corresponding to these u~}l and UbZl solutions to the last equation (11.70) are
ui1l(q) = 1 _ qZ

and
U\Zl(q) = tqZ ~ c 1 -r ·q/m + (ci-r2/m Z - 1)ln(c 1 + mq/-r) + C z .
So, for equation (11.69) we have the following approximate solutions:
390 Chapter 11

where c;(i = 1, ... ,4) are integration constants which may be defined by the
initial and other physical conditions for the given problem.
We notice that the conditions of propagation of the electron-positron
component are different from those for protons and nuclei since relativistic
electrons (and positrons) undergo considerably larger synchrotron and
Compton energy losses. Therefore, the fundamental equation (11.60) describ-
ing the cosmic-ray proton-nuclear component propagation is changed for the
electron-positron component but for its propagation, it seems, equations of the
type of (11.67) and (11.69) are more suitable (see below). Thus, we assume that
the electron acceleration mechanism differs from the acceleration mechanism
for protons and nuclei. As is mentioned by Ginzburg and Ptuskin (1976), it is
most probable that the electron-positron component is completely secondary,
i.e., it is generated by cosmic-rays (protons and nuclei) in the interstellar
medium and perhaps in the sources.
There are a few approaches devoted to explanation of the particle accele-
ration mechanism in stochastic electrodynamics (Rueda, 1977; 1978; Rueda
and Lecompte, 1979) and to the motion of a particle with the shadow of the
vacuum fluctuation (Cheon, 1982). Unlike these approaches, in our model the
particle-acceleration mechanism is caused by fluctuations in metric and de-
pends on the value of the fundamental length 1that may be obtained from the
experimental data on measuring ultra-high energy protons. In other words,
ultra-high energy particles are 'carriers' of space-time properties at small
distances, i.e., they perturb space-time around themselves and make the metric
stochastic (fluctuational). This perturbative effect of the particles on the metric
gives rise to fluctuations such that the deviation from the Minkowski metric
1]1'" near the particles has a nonzero mean. As a result of the last fact, the
particles move with acceleration due to self-interacting gravitational field and
acquire very high energies during the evolution of the Universe.
The most important distinctive outline of our acceleration mechanism
presented here is its smooth character of acceleration by which a high final
energy of particles is achieved little by little ~ continuously or in small
portions. This circumstance is especially essential for the explanation of
existing ultrarelativistic heavy nuclei components in cosmic rays. It is clear
that compound nucleus may conserve its structure as a whole in the accele-
ration process if and only if it gains energy in sufficiently small portions. From
this point of view it is naturally assumed that acceleration has a macroscopic
character and is not the consequence of some elementary act or scattering
process as it might be, for example, if it occurred in ultrarelativistic percussion
waves.
Now we discuss other unsolved questions of interest including the problem
of the energy spectrum of the cosmic rays and the ratio of the intensities of the
electron component to the proton component at the same energy level, within
the framework of our scheme proposed above. It should be noted that these
two problems play an important role in the construction and choice of the
Physical Consequences of Hypothesis of Stochastic Space-Time 391

concrete model for the propagation mechanism of the cosmic rays, which must
reproduce all the experimental data. Thus, in order to characterize our
mechanism with respect to the others and the experimental data, we will
consider, first, the problem of the energy spectrum of the cosmic rays. It
appears that our acceleration mechanism leads to the generation of particles
with the power-law spectrum
N(E) dE = KE-Y dE, (11.71)

where N is a particle concentration and K is some constant. Indeed, by the


same proposition discussed by Ginzburg and Syrovatskii (1964) it is assumed
that the particles start to move with acceleration with equal probability at any
time moment, and the probability density of finding the particles with 'age' in
the interval t, t + dt, is
dW = (l/T) exp( - t/T) dt, (11.72)
where T is the average lifetime of the particles during which the particles
undergo acceleration mechanism (for discussion of T see below there T
denotes Ter). Then, according to the formula (11.63), we have

dE =
dt E 1 _1 42 qq,. (q = v/)
c. (11.73)

Here q and q are given by equations (11.60) and (11.61). Thus,


dE
dt = Be th(et), (11.74)

where we assume q(O) = vo/c = o. Solution of this equation is simple, that is


E = Eo ch(et), (11. 75)
here Eo is the initial energy. But in the ultrarelativistic case we can take v ~ c
or et ~ 00. i.e., q ~ 1 or th(et) ~ 1 in equations (11.73) and (11.74). In this
approximate case, instead of (11.74) and (11.75), we have
dE
dt = eE, E = Eo exp(et), (11.76)

where tmax is the maximum acceleration time.


Taking into account equation (11.76) it is easy to verify that
dt = dE/eE, t = (l/e) In(E/E o)

and

d W ~ N(E) dE ~ (l/eT)Eb/ eT E -(1 + l/eT) dE, (y = 1 + 1/eT). (11.77)


A more exact equation (11.74) gives
dW ~ (l/eT)r 1 jeT Eo 1 + 1/eT E- 1/eT(E/E o - 1) dE. (11.78)
392 Chapter 11

According to the above deduction and reference of Ginzburg and Syrovatskii


(1964), we obtain for the proton component

Gp = (9/8J3)c/2 1';3 and Tp ~ (1 --;- 3) x 108 year.


Thus, Gp Tp ~ 0.645 for Tp = 2.2 X 10 8 year and therefore
Yp = 2.55 (11.79)
which coincides with the experimental value of Yexp'
Within our scheme, it is naturally assumed that the acceleration mechanism
and the value of GN TN for the nuclei component are the same as the proton
component, so that in this case the energy spectrum and the value of yare
determined by the analogous formulas of (11.77), (11.78), and (11.79).
With regard to the acceleration mechanism of the electron-positron com-
ponent, energy losses should be taken into account, and it should be assumed
that, as mentioned above, the electron-positron component is secondary, i.e., it
is generated by means of the processes n ~ J1 ~ e [for detail see Ginzburg and
Syrovatskii, 1964)]. Here we study the equation of the type of (11.67) and
(11.69) in the relativistic case, when the electron undergoes energy losses
connected with bremsstrahlung in the external magnetic field
HG ~ (3 --;- 10) x 10 - 6 Oe (Oersted)
or (11.80)
H MG ~ (3 --;- 10) x 10- 8 Oe

of the Galaxy or of the MetaGalaxy (Ginzburg and Syrovatskii, 1964). In this


case the force (Landau and Lifshitz, 1971)
2 e4
f = - - - Z-5 (F/1vUv)2V (11.81)
3m e
IS directed to the opposite direction of the particle velocity v; choosing the
latter as x-axis and exposing the expression (11.81), we have
2 e4 qH~
Ix = -"3 m Z e4 (1 _ qZ)' (q = vic).
Here it is assumed that the electric component of F/1v is zero. Then, from the
equation (11.67) we obtain the following analogous equation of (11.69):

4= Ge (1
3 me 2
Z
- q Z ) - -2 ( - e )2 H 2 -
-L me
(q )
(1 - q) 1 IZ•

Thus, equation (11.73) in this case acquires the form


dE
-
dt = G Eq -
e
bE ZqZ , (11.82)
Physical Consequences of Hypothesis of Stochastic Space-Time 393

where

Here the energy E and H are expressed in electron volts and oersteds, res-
pectively. In the ultrarelativistic case q ~ 1 we have the approximate equa-
tion
dE 2
-=oE-~E
dt "e U.
(11.83)

This is the Bernoulli equation. The substitution


y(t) = [E(t)r 1 (11.84)
leads to the new differential equation for y(t),
y= -eeY + b, (11.85)
which is soluble because it is linear in y(t). The solution of this inhomogeneous
linear differential equation is

y(t) = r1(t{c1+ b f dp exP(eeP)] = r1(t{c1+ fexP(eet)l


where

l(t) = ex p { ee f dP} = exp(ee t)

is the integrating factor for equation (11.85). Further, according to formula


(11.84), we have
E(t) = exp(eet)[c1 + (b/ee) exp(eetr\ (11.86)
here the integration constant C1 is given by an initial condition, that is
E(O) = Eo. Thus,

E(t) = exp(eet) [(1jE o) + (1jEm.x)(exp(eet) - 1)r 1 . (11.87)


From this it is easy to see that
E(t)lt~ 00 ~ Em•x = ee/b.
We observe here the interesting fact that within our hypothesis concerning the
acceleration mechanism of cosmic-ray particles, the electron energy is restricted
by the maximum value of Em•x' This means that for some time Te during which
our acceleration mechanism gives rise to increasing the electron energy up to
Em • x , after which Te the electron energy is not increased and remains constant
at the order of E m • x ' In other words, it may be said that our acceleration
mechanism is switched off, when the electron energy reaches E m • x , i.e., for
394 Chapter 11

t ~ Te. This effect is caused by the ratio of energy gains and losses due to the
space-time structure near particles, and bremsstrahlung in the external mag-
netic field. When, the electron energy E(t) reaches Emax, this ratio becomes unit,
and energy gained compensates for energy losses. We now calculate the value
of Emax which depends on ee and H -L. As the proton case, the former is given by
ee = (918 J3)el 2 r ;3, where re may be understood as an effective Schwarzschild
radius connected with the internal structure of the leptons. Some theoretical
prediction (Li, 1982) gives re ~ 4 x 10- 17 cm. Therefore, ee ~ 8 x 10- 7 sec- 1
and by expression (11.80) for the magnetic field H-L' we have
E ~ {(24 -;- 2.1) x 10 18 eV for H~ ~ (3 -;- 10) x 10- 6 Oe,
max (24 -;- 2.1) x 10 22 eV for H trG ~ (3 -;- 10) x 10 - 8 Oe.
Now we turn to the calculation of the energy spectrum for the electron
component. Taking into account equation (11.83) and formula (11.87), we have

dt = dE
ee· E(l - EIEmax)'
Substituting these expressions into (11.72), we get

dWe~Ne(E)dE~T8E~Te"
dE (
1 __.°_
E)-liT",
x
ee Emax

(11.88)
Here we again observe the power-law spectrum for the electron-position
component. Assuming as above Teee ~ 1, it is easy to satisfy the experimental
value of y ~ 2 -;- 3.
Now it is relevant to discuss the ratio of the intensities li(i = e,p) of the
electron component to the proton component at the same energy level within
our scheme. In the case of the isotropic radiation the particle concentration Ni
and the intensity Ii are related by the formula (see Ginzburg and Syrovatskii,
1964).

V· = C
, ( 1- (-m.c2
E' -
)2)1/2
Here the intensity of the particle with energy larger than the given value E is
determined by

IJ > E) = i"l li(x)dx = 4C 1


rc 00 dx N(X{ 1 - (m~2 rJ/ 2
Thus, the intensities of the electron and proton components are given by the
following formulas
Physical Consequences of Hypothesis of Stochastic Space-Time 395

and

Ie "'~(E~)1/T'C,(_1_)(1_ E~ )-l/T,C, rEm., dxx- 1 - 1 / Th x


4n Tese Emax JE
X (1 - xlEmaxt 1+ l/T,c, [1 - (mc 2/x)2J 112 ,
respectively
It is easy to verify that the following asymptotic estimations may be obtained
for both cases:

I < ~ (EP IE)l/TpC p


P '" 4n 0

and

Taking into account that


y - 1 = 11Tpsp '" 11Tese '" 1.4 --:- 1.5 and E~ ~ Em • x ,
we get
IelIp;S (E~/Eg)Y-1.
It is naturally suggested that Eg = M pc 2 and E~ = 34 MeV. The latter case is
obtained by the assumption that the electron component is secondary and is
generated by the processes n -> fl-> e. Therefore, the average energy of elec-
trons generated by fl-> e + v + ii decay is just E~ = 34 MeV. Thus,

I II < (Ee IFF - 1 = {0.96% for y = 2.4


e p '" 0 0 0.69% for y = 2.5
which coincides with the average measured ratio being about 1% obtained by
experimentalists (see Meyer et ai., 1974).
Finally, we discuss some questions concerning the value of the effective time
Ter '" t max of acceleration during which the cosmic-ray particle's (mainly pro-
tons) energy reaches 10 19_10 20 eV. It is usually suggested that one could
expect TeJ ;S T u, where Tu is the Hubble time. In particular, this assumption is
based on the fact that in early cosmological epochs, particles would also
apparently lose energy adiabatically, due to the expansion of the Universe, so
that energy gained at very early times would not be retained. The effective time
TeJ of acceleration will therefore be appreciably less than the Hubble time. We
notice that this problem may be solved within the framework of our accele-
ration mechanism proposed above. Indeed, according to formula (11.63) the
maximum energy of the particle is E = (mc 2 /2)exp(stmax 'c), where t max '" Ter
and c'sTer = (9/8~)c[2rc-3 T er . Above we have assumed rc = ro '" 0.8 x
1O- 13 cm which gives E", 10 2 °eV for Ter = Ter'" Tu' However, generally
396 Chapter 11

speaking the value of the radius rc is a semi-empirical one in our approach


and may be changed. From this it is easy to see that the very slightest change
in the value of rc gives rise to the enormous range of the energy of the
cosmic-ray particles. Thus, by some choice of rC' say rc ~ 10- 14 -;- 1O-15 cm
we can obtain E ~ 10 19 - 10 20 eV for Ter ~ 7;,. Here it is assumed that [ =
[PI = 1.62 x 10 - 33 cm.

11.8. Space-Time Structure near Particles and its Influence on


Particle Behavior

11.8.1. Introduction
In this section we consider equation (11.59) and discuss the possibility of an
interrelationship between the properties of space-time structure and the dy-
namics of particles. This problem is not new to scientific literature. For
example, Einstein's (1924) idea to use the random fluctuations of the metric
field gllv(x) as the origin of the real quantum forces which justify the stochastic
interpretation of quantum mechanics were considered by Frederick (1976);
Vigier (1982) and others (see, the review due to Vigier, 1982). Moreover, an
analogous idea on the question has been discussed by Efinger (1981), who
wrote: 'freely moving particles are represented by solitary waves which, by
definition, preserve their shape, then one could hypothesize that associated
with these waves is a Riemannian metric gllv(x) which is nearly singular in a
region corresponding to the width of the solitary wave-amplitude (an idea
reminiscent of an old concept by Einstein, 1967)'. According to this point it is
assumed that the particles in question are geometrical objects on a
Riemannian space-time.
It is well known that in Einstein's theory of gravitation, space-time structure
near matter is changed and differs from flat space-time (for illustration see
Figure 16). The validity of this assumption in a macroscale was demonstrated
by the experiment measuring the deviation of light near the Sun, i.e., a ray of
light going past the Sun is accordingly undergoing deflection at the rate of 1.7
seconds of arc. However, the question of what effect would be expected in the
microworld is still open to discussion within this assumption, and in particular
the influence of space-time structure near particles on their dynamic behavior.
In accordance with the previous section, the next assumption is that a form
of gIlV([2jx 2) in (11.57) determines the space-time structure near particles and
gives a general effect on the particle behavior. In other words, the character
and type of the particle motion is essentially dependent on the space-time
structure near moving particles. For example, we propose that the flat space-
time structure gives rectilinear motion, and soliton-like, stochastic and other
types of particle motion may be caused by different forms of space-time
structures around the particle.
Physical Consequences of Hypothesis of Stochastic Space- Time 397

Fig. 16. The illustration of changing space-time structure near the particle.

Our aim is now to discuss equation (11.59) in the nonrelativistic limit by the
appropriate choice of the form of the function </J(x, I). It is important to notice
that the interrelationship between properties of space-time structure at small
distance and the dynamic behavior of the particles appears at a deeper level
and requires careful investigations. In this respect our approach is modest and
belongs to the semi-empirical level.

11.8.2. Stochastic Behavior of Particles and its Connection with


Stochastic Mechanical Dynamics
It is convenient to study the particle dynamic behavior arising from equation
(11.59) and depending on the concrete form of the function ¢(x, I) in the
nonrelativistic limit. Before approaching this limit for equation (11.59) we make
some remarks. Generally speaking, according to the hypothesis of the curved
space-time structure around particles, the concept of their trajectory and
velocity should be changed and generalized appropriately. In other words, it is
quite possible that owing to the metric fluctuations in the presence of the
particles, the description of the particles has a universal character in Nature
and requires probabilistic methods. In particular, the particle velocity depends
not only on the time variable t but also on the spatial variables x: v(t) --> v(x, t).
We suggest that for a complete description of the particle motion within our
approach, and at the same time with particle velocity v(x,t), two more
quantities p(x,t) and u(x,t) should be introduced, where p(x,t) is the
probability density of finding the particle at point x and at time t, u(x, t) is the
stochastic velocity given by the formula u(x, t) = DV In p(x, t) (see Chapter 8).
Thus, our basic idea is the following: fluctuations in metric take place
everywhere (increasing in the presence of the particles), and may play the role
398 Chapter 11

of the origin of the real 'quantum' forces and lead to the random behavior of
the particles; their dynamics are described by nonlinear partial differential
equations of a type of (11.59) admitting random solutions and by the equation
of continuity for p(x, t):
op/ot + div(pv) = O. (11.89)
In this case the particle velocity is given by the formula

v(t) = fd x p(x, t)v(x, t).


3

In order to obtain the nonrelativistic equation of motion in our ap-


proach we go over to the formal limit c -+ 00 in equation (11.59). The time
component of this equation in the case of c -+ 00 acquires the form
v'V In ¢ = O. (11.90)
The passage to the limit c -+ 00 in the spatial component of (11.59) gives the
following relation:

(11.91)

Here taking into account relation (11.90), from this equation we see that if the
function ¢ does not depend on the time variable, the particle velocity is
constant and the particle moves along the rectilinear trajectory. This situation
also takes place at the limit, when the value of the fundamental length I is
neglected, thus by our assumption ¢(x, I) = 1 at 1-+ 0 (x = x, t).
After simple integration of (11.91) we have

(11.92)
Here the constant may be obtained by the initial condition, say
V 2¢21t = 0 = v~. Thus, the particle velocity v(x, t) and the form of the function

¢(x, l) are related by the formula


(11.93)
Physically this relationship means that by knowing the space-time structure
near the particle we can calculate its velocity (generalized) and, on the
contrary, by the value of the particle velocity one can try to build the space-
time structure near the moving particle. Thus, it seems, there exists a profound
connection between these two concepts and they enter as a single inseparable
entity into our scheme. Therefore, generally speaking, the function ¢(x, l)
should be dependent on the properties of the particle: on its mass, an effective
size A (in the quantum mechanical case A = II/me) and the initial velocity Vo
and so on. In this book we consider three cases for the function
Physical Consequences of Hypothesis of Stochastic Space-Time 399

¢(x, I) => ¢(x, t; I, A, vo) (here the space-time dimension is two; x, t):

I. ¢l (x, t; I, A, vo ) -+ 00 for t -+ 00, and (11.94)


¢l(X, t; I, A, vo) -+ 0 at Ixl -+ 00,
II. ¢z(x, t; I, A, vo) -+ 1, for t, Ixl -+ 00, (11.95)
III. ¢3(X, t; I, A, vo) -+ 00 for t, Ixl -+ 00. (11.96)
We assume here the equality ¢i(X, t; I, A, vo) = 1 at 1-+ 0, is fulfilled for all three
cases. Now we consider the first case and simple form
(11.97)
gives
v = (vo + btx)(1 + btZ)-l, (11.98)
here b = 4D(l/A 3), D is a universal constant dimension of [cm ZIsec], obviously
that is D ~ him. Solution of the equation of continuity (11.89) with the velocity
(11.98) and the initial condition
p(x, t = 0) = a-ln- l/Z exp( _xZla Z)
has the form
(11.99)
where a = (I/A 3 )-1 / Z. According to the above deduction in our case the
stochastic velocity of the particle is given by
(11.100)
Mean value and dispersion of the quantities v(x, t) and u(x, t) are determined
by the formulas:

<v(x, t» = f dx v(x, t) p(x, t) = Vo, (11.101)

(u(x, t» = f dx u(x, t) p(x, t) = 0,

Dis. VZ(x, t) = ta Zb Ztz(1 + btZ)-l.


Moreover, by definition

(X(t» = f dx x p(x, t) = vot,

Dis. XZ(t) = ta Z(1 + bt Z ), (11.102)


(VZ(x, t» = v6 + ta Zb Ztz(1 + btZ)-l.
From formulas (11.98)--(11.102) we conclude immediately that the simple
space-time structure near moving particles which is determined by expression
(11.97) gives completely stochastic mechanical results obtained in Section 9.3 if
400 Chapter 11

the constant D is taken equal to the diffusion coefficient v = h/2m. In this case
the space-time metric near stochastic particles takes the form
(11.103)
where ¢l(X, t; I, Ilv o ) is given by formula (11.97).
It should be noted that this metrical form gives rise to the random behavior
of particles (like Brownian motion), i.e., the particle is forced to move
stochastically during its propagation in space-time. But, on the other hand,
particles always take care of the space-time structure around themselves in
order to move further. It is simple the essence of the dialectical unity of motion
and space-time.

11.8.3. Soliton-Like Behavior of Particles


Now we consider the cases (11.95) and (11.96), and choose the following simple
forms:
(11.104)

¢3(X, t,. I, Il, Vo) -_( 1 + J:I)Shb


ch 3 b ch 3(X-v
o
I 11 2 t )-1
+ b sh x

x (IX ~ zvot + b) (11.105)

where b is a positive constant. These expressions satisfy conditions (11.95),


(11.96) and ¢i = 1 (i = 2, 3) at I = 0, for any x and t. The following generalized
velocities of the particle:
Vz = vol¢z = vo [ll2 + (x - vot)Z][llz + [2 + (x - vot)Z] -1 and V3 = VOI¢3'
correspond to formulas (11.104) and (11.105), respectively. From these velo-
cities V z and V3 we have generalized 'trajectories':
IZ
[vot - Xz(x, t)] = (IlZ + IZ)l/Z

I ( vot 1 ) (11.106)
arc tg (Il Z + [2)1/2 1 + [x(x _ vot)]/(Il 2 + [2)

(11.107)
Physical Consequences of Hypothesis of Stochastic Space-Time 401

Fig. 17. Solitary type wave corresponding to formula (11.106). We see here two precise maximum-
waves, the right part of which moves at a constant velocity without change of shape, while the left
part remains near the point x = O. From this plot and formula (11.106) one can suggest that in
nature there may exist a type of particle motion which consists of two parts: rectilinear and
soliton-like; however, because of the small value of the fundamental length I the latter type motion
is not observed in practice.

Fig. 18. Behavior of the 'trajectory' (11.107) at the value of A = l. This plot is a hump-shaped wave
exactly as Scott Russell observed (see Dodd et al., 1983), which moves along the classical
trajectory x = Va t at a constant velocity without change of shape. For larger values of A the wave
is broader and higher, but becomes thinner and shorter the smaller A becomes. The left part of this
figure, it appears, has no physical meaning which is caused by the initial condition of the problem
and corresponds to some minimum located near the point x = O. This motionless minimum gives
a 'canal' over the time variable, a plot of which is not seen in the figure.
402 Chapter 11

It is easy to verify that


X z(X, t) -+ Xz(t) = vot,
in the limit 1-+0. Behaviors of the functions [vot - X z(x, t )]/1 and X 3(X, t)/l
are shown in Figures 17 and 18, respectively. From these figures we observe
that the 'trajectories' obtained by the functions ¢z and ¢3 correspond to the
soliton-like motion of the particle. However, in the case of (11.106) because of
the small value of the fundamental length I we do not, in fact, observe the
deviation from the rectilinear particle trajectory for any quantity A, even up to
A - l. But the situation is different in the case of (11.107). When
A;;S 10- 15 ..;- 10- 16 cm, the amplitude of the process (11.107) is an order of
unit, therefore we can, in principle, observe experimentally soliton-like particle
behavior which is caused by the space-time structure near the moving particle.

Fig. 19. A plot of the space-time structure, i.e., q,2(X, t) corresponding to the soliton-like motion
sketched in Figure 18. From this diagram we see that the metric ds 2 = q, 2(X, t) dS5 in the
conformally flat space-time is singular in the region corresponding to the classical trajectory
x = Vo t. According to the diagram one can conclude that in order to hold a particle in a localized
region during its propagation in space-time, or, for energy to be propagated in localized stable
'packets' without being dispersed, it needs enormous efforts for space-time, i.e., space-time sets up
an infinitely higher barrier-wall around a particle.
Physical Consequences of Hypothesis of Stochastic Space-Time 403

It should be noted that case (11.94) essentially differs from the other two
cases (11.95) and (11.96). Since in the first case the generalized trajectory
X 1 (x, t) corresponding to the velocity v1 (x, t) [v 1 (x, t) = X 1 (x, t)] becomes to
infinity at x -> ± 00, i.e., the concept of trajectory in this case loses its
significance and has no physical meaning, while the functions [vot - X 2(X, t)]
and X 3(X, t) are located along the rectilinear classical trajectory x(t) = vot.
Finally, for illustration, the space-time structure ds 2 = ¢2(X, t)l'/llv dx v dx ll ,
i.e., the function ¢2(X, t) corresponding to the particle behavior determined by
formula (11.107) is shown in Figure 19. In our semi-empirical approach the
question of a possible unique choice of the functions ¢(x, t; I, A, vo) is not
solved and seems to require another fundamental physical principle. At
present, the formulation of this principle is not known and needs deeper study.
However, we assume that this important problem may be solved alternatively,
namely by the character of particle motion. It means that one can try to
construct the space-time structure near the moving particle by means of the
value of its velocity. For example, if we observe the rectilinear trajectory of the
particle, we can conclude that the space-time structure near such a particle is
flat. The role of that particle may be played by the light quantum-photon since
according to the relation (11.92) if the constant in the right-hand side of (11.92)
is equal to the velocity of light c then the current velocity v(x, t) should be
equal to c, therefore ¢ = 1 everywhere at any t. Of course, we propose here
that the light velocity in space-time is truly constant. If the last assumption is
valid at small distances (or at very-high energies) then light does not disturb
space-time around itself, which always remains flat. In contradistinction to the
rectilinear motion other types of particle motion correspond to a curved space-
time structure near particle under consideration.
In conclusion we notice that fluctuations in metric which give rise to
stochastic and other types of particle motion, if they exist, would be detected
by ultra-high energy particles if their wavelength A were comparable to, or
shorter than, the value of the fundamental length I ~ 10 - 33 em, in other words
the properties of the particle with very small mass m ~ Mpi = hlclpi = 10 -5 gm
(A = hlmc -> 00) would be essentially insensitive to the fluctuational structure
of space-time.
Bibliography

Aharony, A., Imry, Y., and Ma, S.-K. (1976) 'Lowering of Dimensionality in Phase Transitions
with Random Fields', Phys. Rev. Lett. 37, 1364-1367.
Akama, K. (1981) 'Pregeometry Including Fundamental Gauge Bosons', Phys. Rev. D24, 3073-
3081.
Akhiezer, A. 1. and Berestetskii, V. B. (1965) Quantum Electrodynamics, Interscience, New York.
Akhiezer, A. 1. and Berestetskii, V. B. (1981) Quantum Electrodynamics (in Russian), Nauka,
Moscow.
Albeverio, S. and H¢egh-Krohn, R. (1974) 'A Remark on the Connection Between Stochastic
Mechanics and the Heat Equation', J. Math. Phys. 15, 1745-1747.
Albeverio, S. and H¢egh-Krohn, R. (1976) Mathematical Theory of Feynman Path Integrals,
Springer, Berlin.
Alebastrov, V. A. and Efimov, G. V. (1973). 'A Proof of the Unitarity of S-Matrix in a Nonlocal
Quantum Field Theory', Comm. Math. Phys. 311, 1-24.
Alebastrov, V. A. and Efimov, G. V. (1974) 'Causality in Quantum Field Theory with Nonlocal
Interactions', Comm. Math. Phys. 38, 11-28.
Alebastrov, V. A., Efimov, G. V. and Seltzer, Sh. Z. (1973) 'Nonlocal Theory of the
Electromagnetic and Weak Interactions with W-Boson', Ann. Phys. (N.Y.) 76, 251-280.
Alexandrov, P. and Hopf, H. (1935) Topologie, Vol. 1, Springer, Berlin.
Alfaro, 1. and Sakita, B. (1983) 'Derivation of Quenched Momentum Prescription by Means.of
Stochastic Quantization', Phys. Lett. 121B, 339-344.
de Alfaro, V., Fubini, S. and Furlan, G. (1980) 'A New Approach to the Theory of Gravitation',
Nuovo Cimento B57, 227-252.
de Alfaro, V., Fubini, S. and Furlan, G. (1976) 'A New Classical Solution of the Yang-Mills Field
Equations', Phys. Lett. 65B, 163-166.
Andryushin, V. 1., Bilenky, S. M. and Gershtein, S. S. (1971) 'On a Possible Method of a
Measurement of Neutrino Electromagnetic Formfactor' Pisma JE7P 13, 573-576.
Arbuzov, B. V. (1975) 'High-Energy Weak Interactions', in Procedings of the 1975 JINR-CERN
School of Physics, Alushta, USSR, Preprint of JINR, E2-9086, Dubna.
Aron, 1. C. (1981) 'Stochastic Foundation of Microphysics. A Critical Analysis', Found. Phys. 11,
699-720.
Aron, 1. C. (1966) 'Stochastic Processes in Microphysics in Connection with Relativity', Progr.
Theoret. Phys. 35, 147-171.
Aspect, A., Dalibard, 1. and Roger, G. (1982a) 'Experimental Test of Bell's Inequalities Using
Time-Varying Analysis', Phys. Rev. Lett., 49,1804-1807.
Aspect, A., Grangier, P. and Roger, G. (1982b) 'Experimental Realization of Einstein-Podolsy-
Rosen-Bohm Gedankenexperiment: A New Violation of Bell's Inequalities', Phys. Rev. Lett. 49,
91-94.
Bailey, 1. et al. (CERN-Mainz-Daresbury collaboration) (1979) 'Final Report on the CERN Muon

404
Bibliography 405

Storage Ring Including the Anomalous Magnetic Moment and the Electric Dipole Moment of
the Muon, and a Direct Test of Relativistic Time Dilation', Nuclear Phys. 8150, 1-75.
Ballentine, 1. E. (1970) 'The Statistical Interpretation of Quantum Mechanics', Rev. Modern Phys.,
42,358--380.
Baltay, C. (1979) 'Recent Results from Neutrino Experiments in Heavy Preon Bubble Chambers',
in Proceedings of the 1979 JINR-CERN School of Physics, Dobogoko, Hungary.
Banner, M. et al. (UA2 Collaboration) (1983) 'Observation of Single Isolated Electrons of High
Transverse Momentum in Events with Missing Transverse Energy at the CERN pp Collider',
Phys. Lett. 1228, 476-485.
Barber, D. P. et al. (1979a) 'Test of Quantum Electrodynamics at SI/2 = 13 and 17 GeV', Phys. Rev.
Lett. 42,1110-1113.
Barber D. P. et al. (1979b) 'Test of Universality of Charged Leptons', Phys. Rev. Lett. 43, 1915-
1918.
Barber, D. P. et al. (1980) 'Experimental Study of Heavy Charged Leptons and Search for Scalar
Partners of Muons at Petra (12 GeV « Ee.m « 36 GeV)', Phys. Rev. Lett. 45, 1904-1907.
Barbieri, R., Nanopoulos, D. V., Morchio, G. and Strocchi, F. (1980a) 'Neutrino Masses in Grand
Unified Theories', Phys. Lett. 90B, 91-97.
Barbieri, R., Ellis, J. and Gaillard, M. K. (1980b) 'Neutrino Masses and Oscillations in SU(5)',
Phys. Lett., 90B, 249-252.
Bardin, D. Yu. and Mogilevsky, O. A. (1974). 'On the Neutrino Charge Radius', Lett. Nuovo
Cimento 9, 549-553.
Barger, V., Whisnant, K., Cline, D. and Phillips, R. J. N. (1980a) 'Possible Indications of Neutrino
Oscillations", Phys. Lett. 938, 194-198.
Barger, V., Whisnant, K. and Phillips, R. J. N. (1980b) 'Three-Neutrino Oscillations and Present
Experimental Data', Phys. Rev. D22, 1636-1646.
Bargmann, V. and Wigner, E. P. (1948). 'Group Theoretical Discussion of Relativistic Equations',
Proc. Nat. Acad. Sci. U.S.A. 34, 211-.
Bartel, W. et al. (JADE collaboration) (1980) 'Test of Quantum Electrodynamics at Petra', Phys.
Lett. 928, 206-210.
Baseyan, G. Z., Matinyan, S. G. and Savvidi, G. K. (1979) 'Nonlinear Plane Waves in Yang-Mills
Massless Theory', Pisma JETP 29, 641-644.
Baseyan, G. Z. and Matinyan, S. G. (1980) 'Solutions of Classical Yang-Mills Equations
Containing Instantons and Merons, Pisma JETP 31,76-77.
Baulieu, 1. and Zwanziger, D. (1981) 'Equivalance of Stochastic Quantization and the Faddeev-·
Popov Ansatz', Nuclear Phys. 8193, 163-172.
Belavin, A. A., Polyakov, A. M., Schwartz, A. S. and Tyupkin, Yu. S. (1975) 'Pseudo-Particle
Solutions of the Yang-Mills Equations', Phys. Lett. 598, 85-87.
Bender, C. M. and Orszag, S. A. (1978) Advanced Mathematical Methods for Scientists and
Engineers McGraw-Hill, New York.
Berezin, F. A. (1966) Method of Second Quantization, Academic Press, New York and London.
Berger, Ch. et al. (PLUTO Collaboration) (1980) 'Test of QED in the Reactions e+ e- --> e+ e- and
e+ e- -->)1+)1- at CMS Energy from 9.4 to 31.6 GeV', Z. Physik C4, 269-276.
Beron, B. 1. et al. (1978) 'Experimental Tests of Quantum Electrodynamics Through the
Measurement of the Reactions e+ e- --> e+ e-, e+ e- --> yy and e+ e- --> )1 +)1- at a Center-of-mass
Energy of 5.2 GeV', Phys. Rev. D17, 2187-2198.
Berrondo, M. (1973) 'The Path Integral Method as Derived from a Stochastic Variational
- Principle', Nuovo Cimento 18B, 95-109.
Bervillier, c., Drouffe, J. M. and Zinn-Justin, J. (1978) 'Comparison Between Large-Order
Estimates and Perturbation Series in a Scalar Field Theory with Gaussian Propagator', Phys.
Rev. D17, 2144-2159.
Bilenky, S. M., Hosek, J. and Petcov, S. T. (1980) 'On the Oscillations of Neutrinos with Dirac and
Majorana Masses', Phys. Lett. 94B, 495-498.
406 Bibliography

Bilenky, S. M. and Pontecorvo, B. (1980a) 'Majorana and Dirac Masses, Neutrino Oscillations
and the Number of Charged Leptons', Phys. Lett. 958, 233-236.
Bilenky, S. M. and Pontecorvo, B. (1980b) 'On "Fast" and "Slow" Neutrinos', Lett. Nuovo Cimento
28, 601-602.
Bilenky, S. M. and Pontecorvo, B. (1978) 'Lepton Mixing and Neutrino Oscillations', Phys.
Reports 41C, 225-276.
Bloch, C. (1952) 'On Field Theories with Non-localized Interaction', Danske Vid. Selsk. Mat-Fys.
Medd. 27(8), 3-55.
Blokhintsev, D. I. (1973a) 'Stochastic Space and Nonlocal Fields', Teoret. Math. Fiz. 17, 153-159.
Blokhintsev, D. I. (1973b) Space and Time in the Microworld, D. Reidel, Dordrecht.
Blokhintsev, D. I. (1975) 'The Stochastic Spaces', Particles and Nuclei 5, 242- (ed. by APS, New
York).
Blokhintsev, D.1. (1947) 'The Wave Field with Mass Spectrum', JETP 17, 115-120 (English Trans.
J. Phys. 11, 72-).
BogolubQv, N. N. and Shirko.v, D. V. (1980) Introduction to the Theory of Quantized Fields, 3rd
ed., Wiley-Interscience, I., New York.
Bogolubov, N. N., Medvedev, B. V. and Polivanov, M. K. (1958) Problems in the Theory of
Disperision Relations, Institute for Advanced Studies, Princeton.
Bogolubov, N. N., Logunov, A. A. and Todorov, I. T. (1975) Introduction to Axiomatic Quantum
Field Theory, Benjamin, New York.
Bogoliubov, N. N. and Parasiuk, O. A. (1957) 'Uber die Multiplikation der Kausalfunktionen in
der Quantentheorie der Felder', Acta Math. 97, 227-.
Bohm, D. (1952) 'Suggested Interpretation of Quantum Theory in terms of "Hidden Variables"',
(Parts I and II) Phys. Rev., 85, 166-193.
Bohm, D. and Vigier, J.-P. (1954) 'Model of the Causal Interpretation of Quantum Theory in
Terms of a Fluid with Irregular Fluctuations', Phys. Rev. 96, 208-216.
Born, M. (1949) 'Reciprocity Theory of Elementary Particles' Rev. Modern Phys. 21, 463-473.
Born, M. (1938) 'A Suggestion for Unifying Quantum Theory and Relativity', Proc. Roy. Soc.
London A165, 291-303.
Bourrett, R. C. (1964) 'Quantized Fields as Random Classical Fields', Phys. Lett. 12, 323-325.
Boyer, T. H. (1975) 'Random Electrodynamics: The Theory of Classical Electrodynamics with
Classical Electromagnetic Zero-Point Radiation', Phys. Rev. D11, 790-808.
Boy<,:r, T. H. (1975) 'General Connection Between Random Electrodynamics and Quantum
Electrodynamics for Free Electromagnetic Fields and for Dipole Oscillator Systems', Phys. Rev.
Dll, 809-830.
Bracci, L., Fiorentini, G., Mezzorani, G. and Quarati, P. (1983) 'Bounds on a Hypothetical
Fundamental Length', Phys. Lett. 1338, 231-233.
Braffort, P. and Tzara, C. (1954) 'Energie de L'oscillateur Harmonique Dans Ie Vide', C. R. Acad.
Sci. Paris 239, 1779-1780.
Braffort, P. et al. (1965). L'energie Moyenne d'un Oscillateur Harmonique Non-Relativiste en
Electromagnetique Aleatoire. C. R. Acad. Sci. Paris 261, 4339-4341.
Bricman, C. et al. (Particle Data Group) (1980) 'Review of Particle Properties', Rev. Modern Phys.,
52, No.2 sl-s286.
Brodsky, S. J. and Drell, S. D. (1970) 'The Present Status of Quantum Electrodynamics', Annual
Review of Nuclear Science, Vol. 20, E. Segre (ed.), Stanford Univ. Press, Palo Alto, California,
pp. 147-194.
de Broglie, L. (1964) 'La Thermodynamique "Cachee" des Particules', Ann. Inst. H. Poincare 1,
1-19.
de Broglie, L. (1967) 'Notes des Membres et Correspondants et Notes Presentees ou Transmises
par Leurs Soins', C. R. Acad. Sci. Paris 8264, 1041-1044.
Brooke, J. A., Guz, W. and Prugovecki, E. (1982) 'The Reciprocity Principle in Stochastic
Quantum Mechanics', Hadronic J. 5, 1717-1733.
Brout, R., Englert, F., Frere, J.-M., Gunzig, E., Nardone, P., Truffin, C. and Spindei, Ph. (1980)
Bibliography 407

'Cosmogenesis and the Origin of the Fundamental Length Scale', Nuclear Phys. 8170, 228-264.
Bunge, M. (1956). 'Survey of the Interpretations of Quantum Mechanics', Amer. J. Phys. 24,272-
286.
Buras, A. I., Ellis, I., Gaillard, M. K. and Nanopoulos, D. V. (1978) 'Aspects of the Grand
Unification of Strong, Weak and Electromagnetic Interactions', Nuclear Phys. 8135, 66-92.
Busser, F. W. (1980 'High-Energy Neutrino Experiments', in Field Theory and Strong Interactions,
Paul Urban Graz (ea)., Springer-Verlag, Vienna and New York.
Callan, C. G., Dashen, R. and Gross, D. I. (1978) 'Towards a Theory of the Strong Interactions',
Phys. Rev. D17, 2717-2763.
Calmet, I., Narison, S., Perrottet, M. and Rafael, E. (1977) The Anomalous Magnetic Moment of
the Muon: A Review of the Theoretical Contributions' Rev. Modern Phys., 49, 21-29.
Cameron, R. H. and Martin, W. T. (1949) 'Transformations of Wiener Integrals by Non-Linear
Transformations', Trans. Amer. Math. Soc. 66, 252-283.
Caubet, I.-P. (1976) Le Mouvement Brownien Relativiste, Lecture Notes in Mathematics, N. 559,
Springer-Verlag, Berlin.
Cavalieri, G. (1981) The Propagator of Stochastic Electrodynamics', Phys. Rev. D23, 363-372.
Cerofolini, G. F. (1980) 'Quantum and Subquantum Mechanics', Nuovo Cimento 858, 286-300.
Chaichian, M. and Nelipa, N. F. (1984) Introduction to Gauge Field Theories, Springer-Verlag,
Berlin, Heidelberg and New York.
Chandrasekhar, S. (1943) 'Stochastic Problems in Physics and Astronomy', Rev. Modern Phys., 15,
N. 1, 1-89.
Chang, L.-N. and Chang, N.-P. (1980) 'B-L Nonconservation and Neutron Oscillation', Phys. Lett.
928,103-106.
Cheng, T. P. and Ling-Fong Li. (1980) 'Neutrino Masses, Mixings, and Oscillations in
SU(2) x U(l) Models of Electroweak Interactions', Phys. Rev. D22, 2860-2868.
Cheng, T. P. and Ling-Fong Li. (1977) 'Nonconservation of Separate 11- and e-Lepton Numbers in
Gauge Theories with V + A Currents', Phys. Rev. Lett. 38, 381-384.
Cheon, IL-T. (1978) 'Hypothesis of the Fundamental Length and Quantum Electrodynamics',
Internat. J. Theoret. Phys. 17, 611-629.
Cheon IL-T. (1979) 'Special Relativity on the Discrete Space-Time and a Fundamental Length',
Lett. Nuovo Cimento 26, 604-608.
Cheon, IL-T. (1982) 'On the Motion of a Particle with the Shadow of Vacuum Fluctuation', Lett.
Nuovo Cimento 34, 513-519.
Chirikov, B. V. and Shepelyansky, D. L. (1981) 'Stochastic Oscillation of Yang-Mills Classical
Fields', Pisma JETP 34,171-175.
Chretien, M. and Peierls, R. E. (1953) 'Properties of Form Factors in Non-local Theories', Nuovo
Cimento, 10, 668-676.
Claverie, P. and Diner, S. (1976) '''Statistical and Stochastic Aspects of the Delocalization
Problem in Quantum Mechanics" and "The Classical Limit in the Framework of Stochastic
Mechanics''', in Localization and Delocalization in Quantum Chemistry. Ionized and Excited
States, Vol. 2, O. Chalvet, and R. Daudel, et al. (eds.), D. Reidel, Dordrecht, pp. 395-448 and 449-
460.
Claverie, P. and Diner, S. (1978) 'Stochastic Electrodynamics and Quantum Theory', Internat. J
Quantum Chem., 12, Suppl. 1,41-82.
Claverie, P. and Diner. S. (1973, 1975) 'Nates des Membres et Correspond ants et Notes Presentees
ou Transmises par Leurs Soins', c.R. Acad. Sci. Paris 8277, 579-582, 8280, 1-4.
Cole, E.A.B. (1972) 'The Classification of Displacements and Rotations in a Cellular Space-Time',
Internat. J. Theoret. Phys. 5, 437-446.
Coleman, S. (1977) 'Non-Abelian Plane Waves', Phys. Lett. 708, 59-60.
Cufaro Petroni, N. and Vigier, I.-P. (1979) 'Causal Superluminal Interpretation of the Einstein-
Podolsky-Rosen Paradox', Lett. Nuovo Cimenta 26, 149-154.
Cufaro Petroni, N. and Vigier, I.-P. (1983a) 'Causal Action-At-Distance Interpretation of the
Aspect-Rapisarda Experiments', Phys. Lett. 93A, 383-387.
408 Bibliography

Cufaro Petroni, N. and Vigier, 1.-P. (1983b) 'Dirac's Aether in Relativistic Quantum Mechanics',
Found. Phys., 13, 253-286.
Dankel, T.-l. (1970) 'Mechanics on Manifold and the Incorporation of Spins into Nelson's
Stochastic Mechanics', Arch. Rational Mech. Anal. 37, 192-22l.
Dankel, T.-l. (1977) 'Higher Spin States in the Stochastic Mechanics of the Bopp--Haag Spin
Model', J. Math. Phys. 18, 253-255.
Daum, M., Eaton, G. H., Frosch, R., Hirschmann, H., McCulloch, 1., Minehart, R. C, and Steiner,
E. (1978) 'Precision Measurement of the Muon Momentum in Pion Decay at Rest', Phys. Lett.,
74B, 126-129.
Davidson, M. (1978). 'On the Equivalence of Quantum Mechanics and a Certain Class of Markov
Processes', J. Math. Phys., 19, 1975-1978.
Davidson, M. (1979) 'A Generalization of the Hnyes-Nelson Stochastic Model of Quantum
Mechanics', Lett. Math. Phys. 3, 271-277, 367-376.
Davidson, M. (1980a) The Generalized Hnyes-Nelson Model for Free Scalar Field Theory', Lett.
Math. Phys. 4, 101-106.
Davidson, M. (1980b) 'Generalized Stochastic Mechanics on Riemannian Maniforld', Lett. Math.
Phys. 4, 475-483.
Davidson, M. (1981) 'Stochastic Quantization of the Electromagnetic Field' J. Math. Phys. 22,
2588-2593.
Davidson, M. (1982) 'Stochastic Quantization of the Linearized Gravitation Field', J. Math. Phys.
23, 132-137.
DeWitt-Morette, C (1972) 'Feynman's Path Integrals. Definition Without Limiting Procedure',
Comm. Math. Phys. 28, 47-67.
DeWitt-Morette, C and Elworthy, K. D. (eds.), (1981) 'New Stochastic Methods in Physics', Phys.
Reports e77, 122-375.
DeWitt-Morette, C, Maheshwai, A. and Nelson, B. (1979) 'Path Integration in Non-Relativistic
Quantum Mechanics', Phys. Reports 50, 255-372.
Dineykhan, M. and Namsrai, Kh. (1977) 'Formulation of Gauge Invariant Quantum
Electrodynamics in the Stochastic Space', Teoret. Mat. Fiz. 33, 32-41 [Theoret. Math. Phys.
(U.S.A.) 33, 862-].
Dineykhan, M., Namsrai, Kh. and Omboo, Z. (1977) 'Construction of Gradient-Invariant
Quantum Electrodynamics for Particles with Spin 0 and 1 in Stochastic Space', Communication
of the lINR, P2-10963, Dubna.
Dodd, R. K., Eilbeck, 1. C, Gibbon, 1. D. and Morris, H. C (1983) Solitons and Nonlinear Wave
Equations, Academic Press, New York and London.
Dohrn, D. and Guerra, F. (1978) 'Nelson's Stochastic Mechanics on Riemannian Manifold', Lett.
Nuovo Cimento 22, 121-127.
Doob,l. (1953) Stochastic Processes, Wiley, New York.
Duffin, R. (1938) 'On the Characteristic Matrices of Covariant Systems', Phys. Rev. 54, 1114--.
Dunford, N. and Schwartz, 1. (1958) Linear Operators, Vol. 1, Interscience, New York.
Dynkin, E. B. (1965) Markov Processes, Springer-Verlag, Berlin and New York.
Ebel, G. et al. (1970) 'Compilation of Coupling Constants and Low-Energy Parameters', Springer
Tracts in Modern Physics 55, 239-290.
Eden, R. 1. and LandsholT, P. V. (1965) The Problem of Causality in S-matrix Theory'. Ann. Phys.
(N.¥.) 31, 370-390.
Efimov, G. V. (1963) 'On the Development of a Nonlocal Quantum Field Theory Without
Ultraviolet Divergences', JETP 44, 2107-2117.
Efimov, G. V. (1968) 'On a Class of Relativistic Invariant Distributions', Comm. Math. Phys. 7,
138-15l.
Efimov, G. V. (1970) 'On Representation of Entire Functions of an Arbitrary Order of Growth',
Dokl. Akad. Nauk SSSR 195, 536-539.
Efimov, G. V. (1977a) NonlocalInteractions of Quantized Fields, Nauka, Moscow.
Bib liogr aph y 409

Efimov, G. V. (1977b) 'Strong Coupling in the Quantum Field Theory with Nonlocal
Nonpolynomial Interaction', Comm. Math. Phys. 57, 235-258.
Efimov, G. V. (1972) 'On the Construction of Nonlocal Quantum Electrodynamics', Ann. Phys.
(N.Y.) 71, 466-485.
Efimov, G. V. (1979) 'Vacuum Energy in gq:>j-Theory for g --> 00', Comm. Math. Phys. 65, 15-
44.
Efimov, G. V. (1974) 'Quantization of Nonlocal Field Theory', Internat. J. Theoret. Phys., 10, 19-
37.
Efimov, G. V. and Ivanov, M. A. (1981). 'Vacuum Energy Yd in the Yukawa Model in the Strong
Coupling Limit', Preprint of the JINR, P2-81-707, Dubna.
Efimov, G. V., Malyshkin, V. G., Mogilevsky, O. A., Namsrai, Kh. and Yumatov, A. Yu. (1973)
'Anomalous Magnetic Moments of Leptons and the Lamb Shift in Nonlocal Theory of
Current-Current Weak Interactions', Nuclear Phys., 859, 1-22.
Efimov, G. V. and Namsrai, Kh. (1982). The Schrodinger Equation in the Quantum Field Theory
with Nonlocal Interactions, Teoret. Mat. Fiz. 50, 221-229.
Efimov, G. V. and Namsrai, Kh. (1975). 'To the Construction of Nonlocal Theory of Quantum
Electrodynamics of Particles with spin 0 and 1', Teoret. Mat. Fiz. 22, 186-202 (Theoret. Math.
Phys. U.S.A. 22, 129-141).
Efimov G. V. and Seltzer, Sh. Z. (1971) 'Gauge Invariant Nonlocal Theory of the Weak Interactions',
Ann. Phys. (N.Y.) 67,124-144.
Efinger, H. J. (1981). 'Solitary Waves on a Curved Space-Time', Found. Phys., 11, 791-795.
Ehrlich, R. (1978) 'Are There an Elementary Length 10 = 0.66 fm and an Elementary Time
'0 = 0.66 fmlc Associated with the Strong Interactions?' Phys. Rev. D18, 320-325.
Ehrlich, R. (1976) 'Possible Evidence for the Quantization of Particle Lifetimes', Phys. Rev. D13,
50-55.
Einstein, A. (1924) Schweiz. Natur. Gesell. Verhand. 85, 85-.
Einstein, A. (1950, 1979) The Meaning of Relativity, 3rd and 4th edn, Princeton Univ. Press,
Princeton, New York.
Einstein, A. (1956) Investigations on the Theory of Brownian Movement, Dover, New York.
Einstein, A. (1967) The Meaning of Relativity, 6th edn. rev., Appendix ii (1956), Science Paperbacks
and Methuen and Co.
Einstein, A., Podolsky, B. and Rosen; N. (1935); 'Can Quantum-Mechanical Description of
Physical Reality Be Considered Complete?' Phys. Rev., 47, 777-780.
Ellis, J., Gaillard, M. K., Nanopoulos, D. V. and Rudaz, S. (1980) 'Uncertainties in the Proton
Lifetime', Nuclear Phys. 8176, 61-99.
Exner, P. (1984) Open Quantum Systems and Feynman Integrals, D. Reidel, Dordrecht.
Faber, S.M. and Gallagher, J. S. (1979) 'Masses and Mass-to-Light Ratios of Galaxies', Ann. Rev.
Astronom. Astrophys. 17, 135-187.
Faddeev, L. D. and Popov, V. N. (1967) 'Feynman Diagrams for the Yang-Mills Field', Phys. Lett.
258,29-30.
Faddeev, L. D. and Slavanov, A. A. (1980) Introduction to the Theory of Gauge Theories,
Benjamin, Reading, Mass.
Fainberg, V. Va. and Soloviev, M. A. (1978) 'How Can Local Properties Be Described in Field
Theories Without Strict Locality?', Ann. Phys. (N.¥.) 113, 421-447.
Fainberg, V. Va, and Soloviev, M. A. (1977) 'Causality, Localizability and Holomorphically
Convex Hills', Comm. Math. Phys. 57, 149-159.
de Falso, D., de Martino, S. and de Siena, S. (1982) 'Position-Momentum Uncertainty Relation in
Stochastic Mechanics', Phys. Rev. Lett. 49, 181-183.
de Falso, D., de Martino, S. and de Siena, S. (1983) 'Momentum from Sample Paths in Stochastic
Mechanics', Lett. Nuovo Cimento 36, 457-460.
Faustov, R. N. (1972) 'Energy Levels and Electromagnetic Properties of the Hydrogen-Like
Atoms', Fizika Element. Chast. Atomnogo Yadra 3, 238-268.
410 Bibliography

Feynman, R. P. (1948) 'Space-Time Approach to Non-Relativistic Quantum Mechanics', Rev.


Modern Phys. 20, 367-387.
Feynman, R. P. (1950) 'Mathematical Formulation of the Quantum Theory of Electromagnetic
Interactions', Phys. Rev. 80, 440-457.
Feynman, R. P. (1951) 'An Operator Calculus Having Applications in Quantum Electrodynamics',
Phys. Rev. 84, 108-128.
Feynman, R. P. and Hibbs, A. R. (1965) Quantum Mechanics and Path Integrals, McGraw-Hill,
New York.
Fenyes, I. (1952) 'Eine Wahrscheinlichkeitstheoretische Begriindund und Interpretation der
Quantenmechanik', Z. Physik 132, 81-106.
Finkelstein, D. (1969) 'Space-Time Code', Phys. Rev. 184, 1261-1271.
Finkelstein, D. (1972) 'Space-Time Code, II and III', Phys. Rev. OS, 320-328, 2922-2931.
Finkelstein, D. (1974) 'Space-Time Code, IV', Phys. Rev. 09, 2219-2230.
Finkelstein, D., Frye, G. and Susskind, L. (1974) 'Space-Time Code, V', Phys. Rev. 09, 2231-2236.
Fiorini, E. (1979) 'Low Energy Weak Interactions', in Proceedings Neutrino 1979 Conference,
Bergen, Norway, Vol. 1, A. Haatuft and C. Jarlskog (eds.), p. 236.
Floratos, E. and Iiiopoulos, J. (1983) 'Equivalence of Stochastic and Canonical Quantization in
Perturbation Theory', Nuclear Phys. B214, 392-404.
Fliigge, G. (1980) 'Recent e+e- Physics', in Proceedings of the 8th Internatinal Winter Meeting in
Fundamental Physis, Ronda, Spain, p. 169.
Fox, R. F. (1978) 'Gaussian Stochastic Processes in Physics', Phys. Reports 48, 179-283.
Fradkin, E. S. (1954) The Green Function for Interacting Nucleons with Mesons', Doklad. Akad.
Nauk SSSR 98, 47-50.
Fradkin, E. S. (1954) 'On the Renormalization of Quantum Electrodynamics', JETP 26,751-754.
Frederick, C. (1976) 'Stochastic Space-Time and Quantum Theory', Phys. Rev. 013, 3183-3191.
Fridman, A. A. (1965) The World as Space- Time, Nauka, Moscow.
Friedberg, R. and Lee, T. D. (1983) 'Discrete Quantum Mechanics', Nuclear Phys. B225 [FS9], 1-
52.
Frohlich, 1. (1974) 'Verification of Axioms for Euclidean and Relativistic Fields and Haag's
Theorem in a Class of P(,p)2 Models', Ann. Inst. H. Poincare A21, 271-317.
Fubini, S. (1974) 'Summary Talk: Present Trends in Particle Physics', in Proceedings of the XV II
International Conference on High Energy Physics, London, Chilton, London.
Fujiwara, K. (1980) 'Is the Light Velocity in Vacuum Really a Constant: Possible Break-Down of
the Linear wok Relation at Extremely High Frequencies', Found. Phys. 10, 309-332.
Fujikawa, K. and Shrock, R. E. (1980) 'Magnetic Moment of a Massive Neutrino and Neutrino-
Spin Rotation', Phys. Rev. Lett. 45, 963-966.
Fiirth, R. (1933) 'Uber Einige Bezeihungen Zwischen Klassischer Statistik and Quantenmecha-
nik', Z. Physik 81,143-162.
Gel'fand, I. M. and Shilov, G. E. (1964) Generalized Functions, Vol. 1, Academic Press, New York.
Gel'fand, I. M. and Shilov, G. E. (1968) Generalized Functions, Vol. 2, Academic Press, New York.
Gel'fand, I. M. and Yaglom, A. M. (1956) 'Integration in Functional Spaces and Its Application in
Quantum Physics', S. Uspehi Mat. Nauk 67, 77-114 (English Translation (1960), J. Math. Phys.
1,48-69).
Gell-Mann, M., Ramond, P. and Siansky, R. (1979) 'Complex Spinors and Unified Theories', in
Supergravity, D. Z. Freedmann and P. van Nieuwenhuizen (eds), North-Holland, Amsterdam,
pp. 315-321.
Georgi, H. and Glashow, S. L. (1974) 'Unity of all Elementary Particle Forces', Phys. Rev. Lett. 32,
438-441.
Georgi, H. and Nanopoulos, D. V. (1979) 'Ordinary Predictions from Grand Principles: t-quark
Mass in 0(10)', Nuclear Phys. B155, 52-74.
Ghirardi, G. c., Omero, c., Rimini, A. and Weber, T. (1978) 'Stochastic Interpretation of
Quantum Mechanics: A Critical Review', Riv. Nuovo Cimento I, 1-34.
Bibliography 411

Gihman, I. I. and Skorokhod, A. V. (1972) Stochastic Differential Equations, Springer-Verlag,


Berlin.
Gihman, I. I. and Skorokhod, A. V. (1975) The Theory of Stochastic Processes, Springer-Verlag,
New York.
Ginzburg, V. L. (1975), 'Evaporation of Black Holes and the Fundamental Length', Pisma
JETP 22, 514-515.
Ginzburg, V. L. and Ptuskin, V. S. (1976) 'On the Origin of Cosmic Rays: Some Problems in
High-Energy. Astrophysics', Rev. Modern Phys., 48, 161-189.
Ginzburg, V. L. and Syrovatskii (1964) The Origin of Cosmic Rays, Pergamon Press, London.
Glaser, V. (1974) 'On the Equivalence of the Euclidean and Wightman Formulation of Field
Theory', Comm. Math. Phys. 37, 257-272.
Glashow, S. L. (1961) 'Partial Symmetries of Weak Interactions', Nuclear Phys. 22, 579-588.
Glashow, S. L. (1980) Towards a Unified Theory: Threads in a Tapestry', Rev. Modern Phys. 52,
539-543.
Glashow, S. L., Illiopoulos, J. J. and Maiani, L. (1970). 'Weak Interactions with Lepton-Hadron
Symmetry', Phys. Rev., D2, 1285-1292.
Glimm, J. and Jaffe, A. (1981) Quantum Physics (A Functional Integral Point of View), Springer-
Verlag, New York, Heidelberg and Berlin.
Goldhaber, M. and Sulak, L. R. (1981) 'An Overview of Current Experiments in Search of Proton
Decay', Comment in Nuclear and Particle Physics 10, 215-225.
Goldman, T. J. and Ross, D. A. (1979) 'A New Estimate of the Proton Lifetime', Phys. Lett. 84B,
208-210.
Grabert, H., Hanggi, P. and Talkner, P. (1979) 'Is Quantum Mechanics Equivalent to a Classical
Stochastic Process?', Phys. Rev. A19, 2440-2445.
Gribov, V. N. (1977) 'Un stability of the Non-Abelian Gauge Theories and Impossibility of Choice
of the Coulomb Gauge', in Materials for the XII Winter School of the Leningrad Nuclear
Research Institute, pp. 147-162.
Gribov, V. N. (1978) 'Quantization of Non-Abelian Gauge Theories', Nuclear Phys. B139, 1-19.
Gueret, Ph. and Vigier, I.-P. (1982) 'De Broglie's Wave Particle Duality in the Stochastic
Interpretation of Quantum Mechanics: A Testable Physical Assumption', Found. Phys. 12,
1057-1083.
Guerra, F. (1981) 'Structural Aspects of Stochastic Mechanics and Stochastic Field Theory', Phys.
Reports 77, 263-312.
Guerra, F. and Loffredo, M. I. (1980) 'Stochastic Equations for the Maxwell Field', Lett. Nuovo
Cimento 27, 41-45.
Guerra, F. and Ruggiero, P. (1978) 'A Note on Relativistic Markov Processes', Lett. Nuovo
Cimento 23, 529-534.
Guerra, F. and Ruggiero, P. (1973) 'New Interpretation of the Euclidean-Markov Field in the
Framework of Physical Minkowski Space-Time', Phys. Rev. Lett. 31, 1022-1025.
Guha, A. and Lee, S.-C. (1982) 'Stochastic Quantization of Matrix and Lattice Gauge Models',
Preprint of State Univ. of New York, ITP-SB-82-71, Stony Brook, New York, 11794.
Guth, A. H. (1981) 'Inflationary Universe: A Possible Solution to the Horizon and Flatness
Problems', Phys. Rev. D23, 347-356.
Guth, A. H. (1982) '10- 35 Seconds After the Big Bang', in The Birth of Universe, I. Audouze and J.
Tran Thank Van (eds.). (Proceedings of the Seventeeth Rencontre de Moriond Astrophysics
Meeting Les Arcs-Savoie-France, March, 14-26, 1982), Gif Sur Yvette Editions Frontieres, pp.
25-43.
Harish-Chandra (1946) The Correspondence Between the Particle and the Wave Aspects of the
Meson and the Photon', Proc. Roy. Soc. Al86, 502-.
Heisenberg, W. (1936), 'Zur Theorie der "Schauer" in der Hohenstrahlung', Z. Phys. 101,533-540.
Hillas, A. M. (1974) 'Survey of Data on Primary Cosmic-Ray Nuclei above 1014 eV', Phi/os. Trans.
Roy. Soc. Land. A277, 413-428.
412 Bibliography

Hillas, A. M. (1975) 'Some Recent Developments in Cosmic Rays', Phys. Reports 20C, 59-136.
't Hooft, G. (1976) 'Symrrletry Breaking Through Bell-lackiw Anomalies', Phys. Rev. Lett. 37,
8-11.
't Hooft, G. (1979) 'Quantum Gravity: A Fundamental Problem and Some Radical Ideas', in
Recent Development of Gravitation, M. Levy and S. Deser (eds.), (NATO Advanced Study
Institutes Series B: Physics, Vol. 44), Plenum Press, New York and London, pp. 323-345.
't Hooft, G. (1974) 'Magnetic Monopoles in Unified Gauge Theories', Nuclear Phys. B79, 276-284.
Hori, T. (1983) 'BRST Invariance and Stochastic Quantization of Supergravity' Phys. Lett. 123B,
391-395.
Hsu, J. P. and Mac, E. (1979) 'Fundamental Length, Bubble Electrons and Nonlocal Quantum
Electrodynamics', Nuovo Cimenta, B49, 55-67.
Ibragimov, I. I. (1962) External Properties of Entire Functions of the Finite Order of Growth,
Academic Press of Az. SSR, Baky.
Ingraham, R. I. (1967) Renorrnalization Theory of Quantum Field with a Cutoff, Gordon and
Breach, New York.
loffe, B. L. (1973) 'Weak Interactions at Small Distances', Uspehi Fiz. N auk 110, 357-404.
Ito, K. (1961) Lectures on Stochastic Processes, Tata, Bombay.
Jackiw, R. (1977) 'Quantum Meaning of Classical Field Theory', Rev. Modern Phy. 49, 681-706.
Jaffe, A. M. (1966) 'Form Factors at Large Momentum Transfer', Phys. Rev. Lett. 17, 661-663.
Jaffe, A. M. (1967) 'High-Energy Behavior in Quantum Field Theory. I Strictly Localizable Fields',
Phys. Rev. 158, 1454-1461.
Jammer, M. (1974) The Philosophy of Quantum Mechanics 1. Wiley, New York.
Jannussis, A. D. and Papatheou, V. (1982) 'Remarks on the Complex Time of Physical
Phenomena', Lett. Nuovo Cimento 35, 485-487.
Japan group (Fukai, T., Nakazato, N., Namiki, M., Ohba, I., Okano, K., Yamanaka, Y.) (1982)
'Stochastic Quantization Method .. .', Preprints of the Univ. of Waseda, WU-HEP-82-4, 82-5,
82-6,82-7, Tokyo, 160, Japan; Progr. Theoret. Phys. (1983), 69,1580-1616.
Jost, R. (1965) The General Theory of Quantized Fields, American Mathematical Society,
Providence, Rhode Island.
Kac, M. (1959) Probability and Related Topics in Physical Sciences, Interscience, New York.
Kadyshevsky, V. G. (1980) 'A New Approach to the Theory of Electromagnetic Interactions',
Fiz. Element. Chast. Atomnogo Yadra 11, 5-39.
Kang, K., Kim, J., Kim, J. E., Soh, K. S. and Song, H. S. (1980) 'Natural Mass Scale and Neutrino
Oscillation', Phys. Rev. D22, 2869-2875.
Kantorovich, L. V. and Akilov, G. P. (1964) Functional Analysis in Normed Spaces, MacMillan,
New York.
Kazarnovsky, M. V., Kuzmin, V. A., Chetyrkin, K. G. and Shaposhnikov, M. E. (1980) 'On the
Neutron-Antineutron Oscillation', Pisma JE7P 32, 88-91
Kemmer, N. (1939) 'The Particle Aspect of Meson Theory', Proc. Roy. Soc. A173, 91-116.
Kershaw, D. (1964) 'Theory of Hidden Variables', Phys. Rev. B136, 1850-1856.
Kim, R., Mathur, V. and Okubo, S. (1974) 'Electromagnetic Properties of the Neutrino from
Neutral-Current Experiments', Phys. Rev. D9, 3050-3053.
Kinoshita, T. (1979) 'Anomalous Magnetic Moment of an Electron and High Precision Test of
Quantum Electrodynamics', in Luminy CNRS Collogium.
Kinoshita, T. (1950) 'On the Interaction of Mesons with the Electromagnetic Field. 1', Progr.
Theoret. Phys. 5, 473-488.
Kinoshita, T. and Nambu, Y. (1950) 'On the Interaction of Mesons with Electromagnetic Field. II',
Progr. Theoret. Phys. 5, 749-768.
Kirzhnits, D. A. (1966) 'A Non-local Quantum Field Theory', Uspehi Fiz. Nauk 90, 129-142.
Kirzhnits, D. A. and Linde, A. D. (1972) 'Macroscopic Consequences of the Weinberg Model',
Phys. Lett., 42B, 471-474.
Bibliography 413

Kogut, 1. and Susskind, L. (1975) 'Hamiltonian Formulation of Wilson's Lattice Gauge Theories',
Phys. Rev. Dll, 395-408.
Kolmogorov, A. N. (1933) Grundbegriffe der Wahrscheinlichkeitsrechnung, Berlin; (1956)
Foundations oj the Theory oj Probability, Chelsea Publishing Co., New York.
Komar, A. A. and Markov, M. A. (1959) 'An Example of a Field Theory with Indefinite Metric in
Hilbert Space II', Nuclear Phys. 12, 190-203.
Kozik, V. S.,Lubimov, V.A.,Novikov,E.G.,Nozik, V.Z. and Tretiakov,I. F. (1980),AnEstimateofthe
ve-Massfrom the Ii-Spectrum of Tritium in the Valine Molecule', Jadernaya Fiz. 32, 301-303;Phys.
Lett. 94B, 266-268.
Kracklauer, A. F. (1974) 'Comment on Derivation of the Schrodinger Equation from Newtonian
Mechanics', Phys. Rev., DI0, 1358-1360.
Kraus, 1. (1975) 'Delbriick Scattering and Tests of the Electron Propagator', Nuclear Phys. BS9,
133-154.
Kroll, N. M. (1966) 'Ad Hoc Modifications of Quantum Electrodynamics', Nuovo Cimento A45,
65-92.
Lacroix, R. (1979) 'Sur L'existence d'une Longueur Elementaire et d'un Intervalle de Temps
'Elementaire. II', Canad. J. Phys. 57, 1681-1685.
Lai, C. H. (ed.) (1983) Gauge Theory oj Weak and Electromagnetic Interactions (Selected Papers),
World Scientific Publishing Co. Pte Ltd, Singapore.
Landau, L. D. and Lifschitz, E. M. (1971) The Classical Theory oj Fields (3rd edn) Pergamon Press,
Oxford.
Langacker, P. (1981) 'Grand Unified Theories and Proton Decay', Phys. Rep. cn, 185-385.
Langouche, F., Roekaerts, D. and Tirapegui, E. (1979) 'Functional Integral Methods for Stochastic
Fields', Physica 95A, 252-274.
Lautrup, B. E., Peterman, A. and de Rafael, E. (1972) 'Recent Developments in the Comparison
Between Theory and Experiments in Quantum Electrodynamics', Phys. Rep. C3, 193-259.
Lavenda, B. H. (1980) 'On the Equivalence Between Classical Markov Processes and Quantum
Mechanics', Lett. Nuovo Cimento 27, 433-436.
Lee, B. W., Pakvasa, S., Shrock, R. E. and Sugawara, H. (1977) 'Muon and Electron Number
Nonconservation in a V-A Gauge Model', Phys. Rev. Lett. 38, 937-939.
Lee. T. D. (1981) 'Is the Vacuum a Physical Medium?', in Stotistical Mechanics oj Quarks and
Hadrons, H. Satz, (ed.), (Proceedings of the International Symposium Held at the Univ. of
Bielfeld, FRG), Nord-Holland, Amsterdam, New York and Oxford, pp. 3-15.
Lee, V.l. (1980) 'Physical Foundation of Quantum Theory: Stochastic Formulation and Proposed
Experimental Test', Found. Phys. 10, 77-107.
Lehr, W. 1. and Park, 1. L. (1977) 'A Stochastic Derivation of Klein-Gordon Equation', J. Math.
Phys. IS, 1235-1240.
Leschke, H. and Schmutz, M. (1977) 'Operator Orderings and Functional Formulations of
Quantum and Stochastic Dynamics', Z. Physik B27, 85-94.
Levine, M. J., Park, H. Y. and Roskies, R. Z. (1082) 'High-Precision Evalution of Contributions
to g-2 of the Electron in Sixth-Order', Phys. Rev. D25, 2205-2207.
Li, F. (1982) The Effective Schwarzschild Radius and the Internal Interaction of Leptons', Lett.
Nuovo Cimento 34, 507-508.
Linde, A. D. (1982) 'A New Inflationary Universe Scenario; A Possible Solution of Horizon,
Flatness, Homogeneity, Isotropy and Primordial Monopole Problems', Phys. Lett. 10SB, 389-
393.
Magg, M., Ringhofer, K. and Salecker, H. (1972) '/i-Mesic Hydrogen and g-2 of the Muon as
Possible Tests of the Electron Propagator', Nuclear Phys. B40, 367-374.
March, A. Z. (1934) 'Die Geometrie Kleinster Riiume. I, II', Z. Physik 104, 93-99, 161-168.
March, A. Z. (1937) 'Zur Grundlegung und Anwendung Einer Statistischen Metrik', Z. Physik 105,
620.
414 Bibliography

Marciano, W. J. (1979) 'Weak Mixing Angle and Grand Unified Gauge Theories', Phys. Rev. D20,
274-288.
Markov, M. A. (1958) Gyperons and K-Mesons, Fizmatgiz, Moscow.
Markov, M. A. (1959) 'An Example of a Field Theory with Indefinite Metric in Hilbert Space. r,
Nuclear Phys., 10, 140--150.
Markov, M. A. (1982) 'Maximum Density of Matter as an Universal Law of Nature', Pisma JETP
36, 214-216.
Markov, M. A. (1984) 'Some Problems of the Recent Theory of Gravitation', Priroda, No.4 (824),
3-10.
Marshall, T. W. (1963) 'Random Electrodynamics', Proc. Roy. Soc. A276, 475-491.
Marshall, T. W. (1965) 'Stochastic Electrodynamics', Proc. Cambridge Phi/os. Soc. 61, 537-546.
Marzke, J. and Wheeler, J. A. (1964) Gravitation and Relativity, H.-Y. Chiu and W. F. Hoffman
(eds.), W. A. Benjamin, New York.
Matinyan, S. G., Savvidi, G. K. and Ter-Arutinyan-Savvidi, N. G. (1981a) 'Classical Yang-Mills
Mechanics. Nonlinear Colour Oscillations', JE1P 80,830--838.
v1atinyan, S. G., Savvidi, G. K. and Ter-Arutinyan-Savvidi, N. G. (1981b) 'Stochasticity of
Classical Yang-Mills Mechanics and its Elimination by Higgs Mechanism', Pisma JETP 34,
613-616.
Mc'Clain, B., Niemi, A. and Taylor, C. (1982) 'Stochastic Quantization of Gauge Theories', Ann.
Phys. (N.Y.) 140, 232-246.
Mc'Clain, B., Niemi, A., Taylor, C. and Wijewardhana, L. C. R. (1983) 'Superspace, Dimensional
Reduction, and Stochastic Quantization', Nuclear Phys. B217, 430--460.
Mc'Kean (Jr.), H. P. (1969) Stochastic Integrals, Academic Press, New York and London.
Meiman, N. N. (1964) 'The Causality Principle and Asymptotic Behavior of the Scattering
Amplitude', JETP 47, 1966-1983.
Menger, K. (1942) 'Statistical Metrics', Proc. Nat. Acad. Sci. U.S.A. 28, 535.
Menger, K. (1949) 'Probability Geometry', Proc. Nat. Acad. Sci. U.S.A. 37, 226.
Meyer, P., Ramaty, R. and Webber, W. R. (1974) Phys. Today, 27,23.
Mezincescy, L. and Ogievetsky, V. I. (1975) 'A Symmetry Between Bosons and Fermions and
Superfields', Uspehi Fiz. Nauk, 117, 637-683.
Michelson, D. M. and Sivashinsky, G. I. (1977) 'Nonlinear Analysis of Hydrodynamic Instability
in Laminar Flames. II. Numerical Experiments', Acta Astronautica 4, 1207-1221.
Mielnik, B. and Tengstrand, G. (1980) 'Nelson-Brown Motion: Some Question Marks', Internat. J.
Theoret. Phys. 19, 239-250.
Miettinen, H. I. and Thomas, G. H. (1980) 'Evidence that Hadronic Interiors have a Denser
Matter than Charge Distribution', Nuclear Phys. Bl66, 365-377.
Migdal, A. B. (1981) 'Instability of Quantum Yang-Mills Fields and Vacuum Fluctuations', in
Statistical Mechanics of Quarks and Hadrons, H. Satz (ed.), (Proceedings of the International
Symposium Held at the Univ. of Bielfeld FRG), North-Holland, Amsterdam, New York and
Oxford, pp. 349-353.
Misner, Ch. W., Thorne, K. S. and Wheeler, J. A. (1973) Gravitation, W. H. Freeman and Co., San
Francisco.
Miura, T. (1979) 'Relativistic Path Integrals', Progr. Theoret. Phys. 61, 1521-1535.
Mohapatra, R. N. and Marshak, R. E. (1980) 'Phenomenology of Neutron Oscillations', Phys.
Lett. 94B, 183-186.
Moore, S. M. (1979) 'Can Stochastic Physics a Complete Theory of Nature?', Found. Phys. 9, 237-
259.
Moore, S. M. (1980) 'Stochastic Fields from Stochastic Mechanics', J. Math. Phys. 21, 2102-2110.
Morrison, D. R. O. (1980) 'Review of Neutrino Oscillations and Hunt for the Tay Neutrino',
Preprint of CERN, EP-80 -190.
Morse, Ph. M. and Feschbach, H. (1953) Methods of Theoretical Physics, McGraw-Hili, New
York.
Bibliography 415

Nady, K. L. (1966) State Vector Spaces with Indefinite Metric in Quantum Field Theory, Akademia
Kiado, Budapest.
Nage1s, M. M. et al. (1976) 'Compilation of Coupling Constants and Low-Energy Parameters',
Nuclear Phys. 8109, 1-90.
Nakano, T. (1959) 'Quantum Field Theory in Terms of Euclidean Parameters', Prog. Theoret.
Phys. 21, 241-259.
Nakano, Y. (1983) 'One-Time Characteristic Functional in the Stochastic Quantization', Progr.
Theoret. Phys. 69, 316--365.
Namsrai, Kh. (1980a) 'Relativistic Dynamics of Stochastic Particles. Found. Phys. 10, 353-361.
Namsrai, Kh. (1980b) 'A Stochastic Derivation of the Sivashinsky Equation for the Self-Turbulent
Motion of a Free Particle', Found. Phys. 10, 732-742.
Namsrai, Kh. (1980c) 'Relativistic Feynman-Type Integrals', Internat. J. Theoret. Phys. 19, 397-
404.
Namsrai, Kh. (1981a) The Hypothesis of Stochastic Space and General Connection of Stochastic
Theory', Phys. Lett. 82A, 103-106.
Namsrai, Kh. (1981b) 'A Stochastic Model for the Motion of Two Relativistic Particles', J. Phys.
A: Mathematical General, 14, 1307-1311.
Namsrai, Kh. (1981c) 'A Nonlocal Stochastic Model for the Free Scalar Field Theory', Internat. J.
Theoret. Phys., 20, 365-375.
Namsrai, Kh. (1981d) 'Stochastic Mechanics', Fiz. Element. Chast. Atomnogo Yadra, 12, 1116--1156
(Theoret. Math. Phys. (U.S.A.) 12,449-464).
Namsrai, Kh. (1981e) 'Hierarchical Scales and Family of Black Holes', Internat. J. Theoret. Phys.,
20, 749-754.
Namsrai, Kh. (1982) 'Proton Decay in the Theory of Gravitation and the Hypothesis of
Fundamental Length', Phys. Lett. 88A, 269-271.
Namsrai, Kh. and Dineykhan, M. (1983) 'Electromagnetic and Weak Interactions in Stochastic
Space-Time: A Review', Internat. J. Theoret. Phys. 22, 131-192.
Nash, C. (1978) Relativistic Quantum Fields, Academic Press, New York and San Francisco.
Nelson, E. (1964) 'Feynman Integrals and the Schrodinger Equation', J. Math. Phys. 5, 332-343.
Nelson, E. (1966) 'Derivation of the Schrodinger Equation from Newtonian Mechanics', Phys.
Rev., 150, 1079-1085.
Nelson, E. (1967) Dynamical Theories of Brownian Motion, Princeton Univ. Press, Princeton, New
Jersey.
Nelson, E. (1971) 'Quantum Fields and Markoff Fields', in American Mathematical Society,
Summer Institute on Partial Differential Equations held at Berkeley.
Nelson, E. (1973) 'Construction of Quantum Fields from Markoff Fields', J. Funct. Anal. 12, 97-
112.
Nelson, E. (1973) 'The Free Markoff Field', J. Funct. Anal. 12, 211-227.
Nelson, E. (1979) 'Connection Between Brownian Motion and Quantum Mechanics', in Einstein
Symposium, H. Nelkowski (ed.), Berlin.
Neutrino 82, Proceedings of the International Conference, 14-19 June, 1982), A. Frenkel and L.
Jenik (eds.), Vols I and II, Supplement, Balatonfiired, Budapest.
Nielsen, H. B. and Olesen, P. (1973) 'Vortex-Line Models for Dual Strings', Nuclear Phys. 861,
45-61.
Niemi, A. and Wijewardhana, L.c.R. (1982) 'Stochastic Regularization of Quantum Field Theory',
Ann. Phys. (N.¥.) 140, 247-265.
Olver, F. W. J. (ed.) (1960) Royal Society's Mathematical Tables, VoL 7, Bessel Functions; Part III,
Zeros. and Associated Values, Univ. Press, Cambridge.
Onofri, E. (1979) The Stochastic Interpretation of Quantum Mechanics: A Reply to Ghirardi et
al.', Lett. Nuovo Cimento 24, 253-254.
Osterwalder, K. and Schrader, R. (1973) 'Axioms for Euclidean Green's Functions', Comm. Math.
Phys. 31, 83-112.
416 Bibliography

Osterwalder, K. and Schrader, R. (1975) 'Axioms for Euclidean Green's Functions. II', Commun.
Math. Phys. 42, 281-305.
Pais, A. and Uhlenbeck, G. E. (1950) 'On Field Theories with Non-Localized Action', Phys. Rev.,
79, 145-165.
Paley, R. E. A. C. and Wiener, N. (1934) Fourier Transforms in the Complex Domain, American
Mathematical Society, New York.
Paley, R. E. A. c., Wiener, N. and Zygmund, A. (1933) 'Note on Random Functions', Math. Z. 37,
647-668.
Parisi, G. (1977) The Perturbative Expansion and the Infinite Coupling Limit', Phys. Lett., 698,
329-331.
Parisi, G. and Sourlas, N. (1979) 'Random Magnetic Fields, Supersymmetry, and Negative
Dimensions', Phys. Rev. Lett. 43, 744-745.
Parisi, G. and Sour las, N. (1982) 'Supersymmetric Field Theory and Stochastic Differential
Equations', Nuclear Phys. 8206, 321-332.
Parisi, G. and Wu, Y.-S. (1981) Sci. Sinica 24, 483-
Particle Data Group (1982) 'Review of Particle Properties', Phys. Lett. 1118, i-xxii, 1-294.
Pati, J. C. and Salam, A. (1973a) 'Is Baryon Number Conserved?', Phys. Rev. Lett. 31, 661-
664.
Pati, 1. C. and Salam, A. (1973b) 'Unified Lepton-Hadron Symmetry and a Gauge Theory of the
Basic Interactions', Phys. Rev. D8, 1240-1251.
Pati, 1. C., Salam, A. and Strathdee, 1. (1981) 'Probings Through Proton Decay and n-n
Oscillations', Nuclear Phys. 8185, 445-472.
de la Pena-Auerbach, L. (1969) 'New Formulation of Stochastic Theory and Quantum
Mechanics', J. Math. Phys. 10, 1620-1630.
de la Pena-Auerbach, L. (1971) 'Stochastic Theory of Quantum Mechanics for Particles with Spin',
J. Math. Phys., 12, 453-461.
de la Pen a-Auerbach, L. and Cetto, A. M. (1972) 'Stronger Form for the Position-Momentum
Uncertainty Relation', Phys. Lett. 39A, 65-66.
de la Pena-Auerbach, L. and Cetto, A. M. (1974) 'Stochastic Treatment of the Quantum
Mechanical Harmonic Oscillator: Mass Renormalization and Lamb Shift', Phys. Lett. A47,
183-184.
de la Pen a-Auerbach, L. and Cetto, A. M. (1977) 'Derivation of Quantum Mechanics from
Stochastic Electrodynamics', J. Math. Phys. 18, 1612-1622.
de la Pena-Auerbach, L. and Cetto, A. M. (1978) 'Why SchrOdinger's Equation?', Internat. J.
Quant. Chem. 12, Supp!. 1, 23-37, Discussion, 39-40.
de la Pen a-Auerbach, L. and Cetto, A. M. (1982) 'Does Quantum Mechanics Accept a Stochastic
Support?', Found. Phys. 12 1017-1037.
de la Pen a-Auerbach, L. and Cetto, A. M. (1975) 'Stochastic Theory for Classical and Quantum
Mechanical Systems', Found. Phys. 5, 355-370.
Petrina, D. Ya., Ivanov, S. S. and Rebenko, A. L. (1979) Equationsfor Coefficient Functions of the
Scattering Matrix, Nauka, Moscow.
Poincare, H. (1898) La Mesure du Temps Rev. de Metaphysique et de Morale V, VI.
Polyakov, A. M. (1975a) 'Compact Gauge Fields and the Infrared Catastrophe', Phys. Lett. 598,
82-84.
Polyakov, A. M. (1975b) 'Isomeric States of Quantum Fields', JETP 68, 1975-1990.
Pontecorvo, B. (1967) 'Neutrino Experiments and the Question of Lepton Charge Conservation',
JETP 53,1717-1725.
Prugovecki, E. (1978) 'Consistent Formulation of Relativistic Dynamics for Massive Spin-Zero
Particles in External Fields', Phys. Rev. D18, 3655-3675.
Prugovecki, E. (1978) 'Relativistic Quantum Kinematics on Stochastic Phase Spaces for Massive
Particles', J. Math. Phys. 19, 2260-2270.
Prugovecki, E. (1981) 'Quantum Space-Time Operationally Based on Propagators for Extended
Test Particles', Hadronic J. 4, 1018-1104.
Bibliography 417

Prugovecki, E. (1984) Stochastic Quantum Mechanics and Quantum Spacetime, D. Reidel,


Dordrecht.
Quigg, C. and Jackson, J. D. (1968) 'Decays of Neutral Pseudoscalar Mesons into Lepton Pairs',
Preprint of California Univ., Berkeley, Lawrence Radiation Laboratory, UCRL, Report No.
18487.
Recami, E. (1981) 'Elementary Particles as Micro-Universes', Preprint of Institute Nazional di
Fisica Nucleare INFN/AE-81/3. (Gennaio, Sezione de Catania, Italia).
Reed, M. and Simon, B. (1972) Methods of Modern Mathematical Physics, Vol. I, Functional
Analysis, Academic Press, New York.
Reed, M. and Simon, B. (1975) Methods of Modern Mathematical Physics, Vol. 2, Fourier Analysis.
Self-Adjointness, Academic Press, New York.
Reines, F. and Crouch, M. F. (1974) 'Baryon-Conservation Limit', Phys. Rev. Lett. 32, 493-494.
Reines, F., Learned, J. and Soni, A. (1979) 'Limits on Nonconservation of Baryon Number', Phys.
Rev. Lett. 43, 907-910.
Reines, F., Sobel, H. W. and Pasierb, E. (1980) 'Evidence for Neutrino Instability', Phys. Rev. Lett.
45, 1307-1311.
Rice, S. O. (1944) 'Mathematical Analysis of Random Noise, I, II', Bell System Tech. J., 23, 282-
332.
Rice, S. O. (1945) 'Mathematical Analysis of Random Noise, III: Statistical Properties of Random
Noise Currents', Bell System Tech. J. 24, 46-108.
Ringhofer, K. and Salecker, H. (1980) 'What Can Be Tested in Quantum Electrodynamics?' Found.
Phys. 10, 185-196.
Robiscoe, R. T. (1968) 'Reconciliation of Experimental Lamb Shifts', Phys. Rev. 168,4-11.
Robiscoe, R. T. and Shyn, T. W. (1970) 'Kinematics Corrections to Atomic Beam Experiments',
Phys. Rev. Lett. 24, 559-562.
Rosen, N. (1974) 'A Classical Picture of Quantum Mechanics', Nuovo Cimento 19B, 90-98.
Ross-Bonney, A. A. (1975) 'Does God Play Dice? A Discussion of Some Interpretation of
Quantum Mechanics', Nuovo Cimento 30B, 55-79.
Roy Choudhury, R. K. and Roy, S. (1980) 'Stochastic Space-Time and the Concept of Potential in
Classical Mechanics', Phys. Rev. D22, 2384-2386.
Roy, S. (1979) 'Relativistic Brownian Motion and the Foundations of Quantum Mechanics',
Nuovo Cimento B51, 29-44.
Rueda, A. (1977) 'Particle Acceleration Mechanism in Stochastic Electrodynamics', (Preprint of
the Univ. de los Andes, Departamento de Fisica, AR-I, Apartado Aero 4976), Bogota, D. E.
Colombia, S. A.
Rueda, A. (1978) 'Model of Einstein and Hopf for Protons in Zero-Point Field and Cosmic-Ray
Spectrum', Nuovo Cimento 48A, 155-182.
Rueda, A. (1981) 'Behavior of Classical Particles Immersed in the Classical Electromagnetic Zero-
Point Field', Phys. Rev. A23, 2020-2040.
Rueda, A. and Lecompte, A. (1979) 'On Approximations in the Model of Einstein and Hopf',
Nuovo Cimento 52A, 264-275.
De Rujula, A. and Glashow, S. L. (1980) 'Neutrino Weight Watching', Nature, 286, No. 5775, 755-
756.
De Rujula, A., Lusignoli, M., Maiani, L., Petcov, S. T. and Petronzio, R. (1980) 'A Fresh Look at
Neutrino Oscillations', Nuclear Phys. B168, 54-68.
Salam, A. (1968) in Elementary Particle Physics (Nobel Symposium, No.8), N. Svartholm (ed.),
Almquist and Wiksell, Stockholm, p. 367.
Salam. A. (1980) 'Gauge Unification of Fundamental Forces', Rev. Modern Phys. 52, 525-538.
Salam, A. and Strathdee, J. (1976) 'Black Holes as Solitons', Phvs. Lett., 6IB, 375-376.
Santamato, E. and Lavenda, B. H. (1981) The Underlying Brownian Motion of Nonrelativistic
Quantum Mechanics', Found. Phys. 11, 653-678.
Santos, E. (1972) 'Brownian Motion and Stochastic Theory', in Irreversibility in the Many-Body
Problem, J. Biel, and J. Rae (eds.) Plenum Press, New York, pp. 459-470.
418 Bibliography

Santos, E. (1974) The Harmonic Oscillator in Stochastic Electrodynamics', Nuovo Cimento B19,
57~89.
Santos, E. (1974) 'Foundations of Stochastic Electrodynamics, 1: General Formalism and the Free
Radiation Field', Nuovo Cimento 22B, 201~214.
Sawada, O. and Fukugita, M. (1980) 'Neutron Oscillation as a Course of Cosmic Ray
Antinucleons', (Preprint of National Laboratory of High Energy Physics), KEK. TH-19,
Tsukuba.
Scadron, M. D. (1980) Advanced Quantum Theory and Its Applications. Through Feynmhn
Diagrams, Springer-Yerlag, New York, Heidelberg and Berlin.
Schramm, D. N. (1979) 'Current Outlook for Neutrino Astronomy', in Proceedings of Neutrino
1979 Conference, Vol. 1, A. Haatuft and C. Jarlskog, (eds.), Bergen, Norway, p. 503.
Schramm, R. and Steigman, G. (1980) 'Gravitiy Foundation Prize Essay' (unpublished).
Schrodinger, E. (1931) Uber die Umkehrung der Naturgesetze, Berliner Sitzungsberichte, pp. 144-
153.
Schwartz, L. (1957, 1959) Theorie des Distributions, Vols. I~II. Hermann, Paris.
Schweizer, B. (1967) 'Probability Metric Space', in The First 25 Years of the University of
Massachusetts.
Schwinger, J. (1959) 'Euclidean Quantum EleGtrodynamics', Phys. Rev. 115, 721~731.
Sehgal, L. M. (1969) 'Electromagnetic Contribution to the Decays Ks -> IT and KL --> If",
Phys. Rev. 183, 1511~1513.
Sherstnev, A. N. (1963) 'On a Notion of a Random Normalized Space', Dokl. Akad. Nauk SSSR,
149, 280-283.
Shucker, D. S. (1978) 'Stochastic Mechanics', Princeton Thesis.
Shucker, D. S. (1980a) 'Davidson's Generalization of the Fenyes~Nelson Stochastic Model of
Quantum Mechanics', Lett. Math. Phys. 4, 61~65.
Shucker, D. S. (1980b) 'Stochastic Mechanics of Systems with Zero Potential', J. Funct. Anal. 38,
146~155.
Shuryak, E. V. (1980) 'Quantum Chromodynamics and the Theory of Superdense Matter', Phys.
Reports C61, 71~158.
Simon, B. (1974) The P(CP)2 Euclidean (Quantum) Field Theory, Princeton Univ. Press, Princeton,
New Jersey.
Simon, B. (1979) Functional Integration and Quantum Physics, Academic Press, New York, San
Francisco and London.
Singer, I. M. (1978) 'Some Remarks on the Gribov Ambiguity', Comm. Math. Phys. 60, 7~12.
Sivashinsky, G. I. (1977) 'Nonlinear Analysis of Hydrodynamic Instability in Laminar Flames. I.
Derivation of Basic Equations', Acta Astronautica 4, 1177~1206.
Sivashinsky, G. I. (1978) Self-Turbulence in the Motion of a Free Particle', Found. Phys. 8, 735~
744.
Sivashinsky, G. I. (1980) Turbulence in the Motion of a Free Particle and de Broglie Waves', Lett.
Nuovo Cimento 27, 504-508.
Skagerstam, Bo-S. K. (1975) 'Stochastic Processes on a Classical Configuration Space', Pre print of
Institute for Theoretical Physics, Report 75-21, Gothenburg, Sweden.
Slavnov, D. A. (1974) 'The Generalized Pauli~Villars Regularization', (Ph.D. Thesis), Moscow
State Univ.
Smirnov, A. Ya. (1980) 'Superheavy Fermions and Proton Lifetime', Pisma JETP 31, 781~783.
Smoluchowsky, M. V. (1923) Abhandlung Uber dir Brownsche Bewegung und Verwandte
Erscheinugen, Akademii-Verlag, Leipzig.
Snyder, H. S. (1947) 'Quantized Space-Time', Phys. Rev. 71, 38- .
Snyder, H. S. (1947) The Electromagnetic Field in Quantized Space-Time', Phys. Rev. 72, 68~ 71.
Soloviev, M. A. (1971) 'On the Choice of the Class of Generalized Functions Compatible with
Microcausality Condition and Spectrality', Tearet. Mat. Fiz. 7, 183~194.
Soloviev, M. A. (1980) 'Relativistic Invariant Formulation of Causality in Nonlocal Field Theory
of Exponential Growth', Teare!. Mat. Fiz. 43, 202~209.
Bibliography 419

Stapp, H. P. (1965) 'Space and Time in S-Matrix Theory', Phys. Rev. B139, 257-270.
Streater, R. and Wightman, A. S. (1964) CPT, Spin and Statistics and All That, Benjamin, New
York.
Surdin, M. (1971) 'Le Champ Electromagnetique Fluctuant de L'univers (Essai d'une
Electrodynamique Stochastique),. Anns. Inst. H. Poincare 15, 203-241.
Surdin, M. (1978) 'The Steady State Universe Revisited, with Stochastic Electrodynamics as a
Guide', Found. Phys. 8, 341-357.
Sutton, Ch. (1980) 'Neutrinos: Do They Rule the Universe', New Scientist, 86, No. 1206, 308- .
Symanzik, K. (1966) 'Euclidean Quantum Field Theory. I. Equations for a Scalar Model', J. Math.
Phys. 7, 510-525.
Symanzik, K. (1969) 'Euclidean Quantum Field Theory', in Proceedings oJ the International School
oj Physics "Enrico Fermi"', Varenna Course XLV, R. Jost (ed.), Academic Press, New York.
Takano, Y. (1961) 'The Singularity of Propagators in Field Theory and the Structure of Space-
time', Prog. Theoret. Phys. 26, 304-314; 27, 212-213.
Takano, Y. (1967) 'Fluctuation of Space-Time and Elementary Particles, I, II', Progr. Theoret.
Phys. 38, 1185-1187.
Tamm, I. E. (1965) 'On the Curved Momentum Space', in Proceedings oj the International
ConJerence on Elementary Particles, Kyoto, Kyoto, Japan, pp. 314-326.
Taylor, J. G. (1966) 'On the Existence of Field Theory, I. The Analytic Approach', J. Math. Phys.
7,1720-1729.
Tennakone, K. (1974) 'Electron, Muon, Proton, and Strong Gravity', Phys. Rev. D10, 1722-1725.
Terazawa, H. (1981) 'Pregeometry', Preprint of the Institute for Nuclear Study, INS-Report No.
429, Univ. of Tokyo.
Thomsen, D. E. (1980) 'Ups and Downs of Neutrino Oscillation', Sci. News 117, No. 24, 377.
Ting, S. C. C. (1982) 'Test of Quantum Electrodynamics and the Study of Heavy Leptons', Riv.
Nuovo Cimento 5, 1-36.
Titchmarsh, E. (1939) The Theory oj Functions, Oxford Univ. Press.
Turner, M. S. and Wilczek, F. (1982) 'Is Our Vacuum Metastable?' Nature 298, 633-634.
Uhlenbeck, G. E. and Ornstein, L. S. (1930) 'On the Theory of the Brownian Motion', Phys. Rev.
36, 823-841.
Van Dyck, R. S. et al. (1979) 'Progress of the Electron Spin Anomaly Experiment', Bull. Amer.
Phys. Soc. 24, 758.
Van Dyck, R. S., Schwinberg, P. B. and Dehmelt, H. G. (1977) 'Precise Measurements of Axial
Magnetron, Cyclotron, and Spin-Cyclotron-beat Frequencies on a Isolated I-MeV Electron',
Phys. Rev. Lett. 38, 310-314.
Van Nieuwenhuizen, P. and Freedman, D. Z. (eds.) (1979) Supergravity North-Holland,
Amsterdam.
Varadham, S. R. S. (1980) Diffusion Problems and Partial Differential Equations, Published for Tata
Institute of Fundamental Research by Springer-Verlag, p. 251.
Vialtsev, A. I. (1965) Discrete Space- Time, Nauka, Moscow.
Vigier, J.-P. (1979) 'Model of Quantum Statistics in Terms of a Fluid with Irregular Stochastic
Fluctuations Propagating at the Velocity of Light: A Derivation of Nelson's Equations', Lett.
Nuovo Cimento 24, 265-272.
Vigier, J.-P. (1982a) 'Non locality, Causality and Aether in Quantum Mechanics', Astronom. Nach.
303, I, 55-80.
Vigier, J.-P. (1982b) Louis de Broglie - Physicist and Thinker', Found. Phys. 12, 923-930.
Vladimirov, V. S. (1979) Generalized Functions in Mathematical Physics, Nauka, Moscow.
Vuilleumier, J. L. (1979) 'Search for Neutrino Oscillations - A Progress Report', in Proceedings oj
Neutrino 1979 Coriference, Vol. 1, A. Haatuft and C. Jarlskog (eds.), Bergen, Norway, p. 185.
Wachsmuch, H. (1979) 'Recent Data on Prompt Single Lepton Production in Hadron-Nucleus
Collisions', in Proceedings oj the 1979 International Symposium on Lepton and Photon
Interactions at High Energy, Fermi National Acceleration Laboratory, Batavia, Illinois, USA,
pp. 541-552.
420 Bibliography

Walstad, A. (1980) 'Time's Arrow in an Oscillating Universe', Found. Phys. 10, 743-750.
Wanders, G. (1959) 'On the Problem of Causality', Nuovo Cimento, 14, 168-184.
Wang, M. C. and Uhlenbeck, G. E. (1945) 'On the Theory of the Brownian Motion. II', Rev.
Modern Phys. 17, 323--342.
Wataghin, G. (1934) 'Bemerkung iiber die Selbstenergie der Electronen', Z. Physik 88, 92-98.
Weinberg, S. (1967) 'A Model of Leptons', Phys. Rev. Lett. 19, 1264--1266.
Weinberg, S. (1972) Gravitation and Cosmology, Wiley, New York.
Weinberg, S. (1979}'Baryon- and Lepton-Nonconserving Processes', Phys. Rev. Lett. 43, 1566--1570.
Weinberg, S. (1980) 'Conceptual Foundations of the Unified Theory of Weak and Electromagnetic
Interactions', Rev. Modern Phys. 52, 515--532.
Weinberg, S. (1981) The Decay of Proton. Sci. Amer. 244, 64--75.
Weizel, W. (1953) 'Ableitung der Quantemtheorie aus einem Klassischen, Kausal Determinierten
Modell, I, II', Z. Physik 134, 264--285; 135, 270--273.
Welton, T. A. (1948) 'Some Observable Effects of the Quantum-Mechanical Fluctuations of the
Electromagnetic Field', Phys. Rev. 74, 1157-1167.
Weisskopf, V. F. (1949) 'Recent Developments in the Theory of the Electron', Rev. Modern Phys.
21, 305--321.
Weyl, H. (1929) 'Electron and Gravitation I', Z. Physik 56,330-- .
Weyl, H. (1950) 'A Remark on the Coupling of Gravitation and Electron', Phys. Rev. 77, 699-701.
Wheeler, J. A. (1964) 'Geometrodynamics and the Issue of the Final State', in Relativity, Groups
and Topology, C. DeWitt and B. S. DeWitt (eds.), Gordon and Breach, New York.
Wick, W. D. (1981) 'Convergence to Equilibrium of the Stochastic Heisenberg Model', Comm.
Math. Phys. 81, 361-377.
Wiener, N. (1923) 'Differential Space', J. Mathematical and Physical Sci. 2,132-174.
Wightman, A. S. (1964) Introduction to Some Aspects of the Relativistic Dynamics of Quantized
Fields, Lectures at the French Summer School of Theoretical Physics, Cargese, Corsica.
Wightman, A. S. (1972) Constructive Field Theory. Introduction to the Problems Coral Gables
Lectures.
Wilczek, F. and Zee, A. (1979) 'Operator Analysis of Nucleon Decay', Phys. Rev. Lett. 43, 1571-
1573.
Wilson, K. G. (1974) 'Confinement of Quarks', Phys. Rev. DlO, 2445-2459.
Winter, K. (1979) 'Neutrino Physics', in Proceedings of the 1979 JINR-CERN School of Physics,
Dobogoko, Hungary.
Witten, E. (1980) 'Neutrino Masses in the Minimal 0(10) Theory', Phys. Lett. 91B, 81-84.
Wolf, G. (1980 'Selected Topics on e+e- Physics', Preprint DESY 80/13, Hamburg.
Wu, T. T. and Yang, C. N. (1969) 'Some Solutions of the Classical Isotopic Gauge Field
Equations', in Properties of Matter Under Unusual Conditions, H. Mark and S. Fernbach (eds.),
Interscience, New York, pp. 349-354.
Yanagida, T. and Yoshimura, M. (1980) 'Neutrino Mixing in a Class of Grand Unified Theories'
Phys. Lett., 97B, 99-102.
Yang, C. N. and Mills, R. L. (1954) 'Conservation of the Isotopic Spin and the Isotopic Gauge
Invariance', Phys. Rev. 96, 191-194.
Yasue, K. (1979) 'Stochastic Quantization: A Review', Internat. J. Theoret. Phys. 18, 861-913.
Yukawa, H. (1950) 'Quantum Theory of Non-local Fields, I, II', Phys. Rev. 77, 219-226, 1047-
1052.
Yukawa, H. (1966), Pre print of Research Institute for Fundamental Physics, Kyoto Univ., RIFP-
55.
Zwanziger, D. (1981) 'Covariant Quantization of Gauge Fields Without Gribov Ambiguity',
Nuclear Phys. B192, 259-269.
Index

Acceleration mean 266, 277 see also S-matrix


mechanism of cosmic rays 358, 384 Central limit theorem 241
passim, 390 passim Characteristic function of a random
Action classical 164, 179 IT. variable 215, 216, 226, 230 ff.
Anharmonic oscillator 192, 195 passim Charge radius of neutrino 146 passim
Annihilation and creation operators 93 Charged closed loop 111, 136 passim, 147
Autonomous equation 389 passim
Averaging in stochastic space-time 7, 10, Chapman-Kolmogorov equation 233, 235
12 If. passim, 298
over time 235 Chebyshev's inequality 218
Christolfel symbol 386
Bargmann-Wigner equation 122, 123 Clilford algebra 170 passim
Baulieu, L. 324, 327, 336 passim, 342, 343, Compton wave-length 258, 276, 371, 372,
347 398, 403
Berezin F. A. 171 passim Condition expectation 162 passim 220 If.
Bernoulli equation 393 probability 220, 223
Bessel function 12, 28, 85, 359, 374, 375, Constructive quantum field theory 3, 93
381 Convergence over probability 217 passim
Big-Bang cosmology 259 strong 17
Black holes small 365 passim Convolution 25 passim
Blokhintsev, D. I. 5, 9, 11, 50, 72, 254 Correlation function 227 If.
Bochner's theorem 215 Cosmic ray energy spectrum 358, 391
Bogolubov, N. N. 3, 4, 15, 21 If. passim
Bohm, D. 254, 256, 277 Cosmological term 257 passim
Boltzmann ergodic hypothesis 235 Coulomb gauge 331, 335
Born's reciprocity principle 379 potential 14, 68, 377
de Broglie, L. 254, 257, 277 Covariance 217, 218 passim, 227, 242, 269,
Brownian motion 161, 232, 241 If. 273, 334, 362
d-dimensional 247 passim see also correlation function
type motion 260 If. Cylinder function 243

Cameron-Martin formula 252 Davidson, M, 4, 11, 253, 267, 324, 327


Canonical commutation relations 4, 55, 57, passim
72, 75, 93 d-operation 110 passim
Cauchy problem 47, 53, 92, 93, 244, 268, Delta function 16, 22 If.
309 passim, 350, 351, 386 Density of matter 357, 362 passim
Causality condition 4, 31, 61, 73 passim, Derivative backward 265 If.
88 passim directional 321 passim

421
422 Index

Derivative backward (contd.) Feynman-Kac formula 244, 249 passim


forward 265 If. Feynmann-Kac-Ito formula 251
functional (variational) 158 If. Feynman diagrams 107, 113 If.
Ito's 277 general parametric formula 115
Radon-Nikodym 220, 252 path integral 161 passim, 179
with respect to a Grassmann vari- relativistic 301 passim
able 166 passim, 174 passim process 296, 302 passim
Dilferential form 321 passim Finite function 21
Dilfusion coefficient 161, 164, 243, 253 If. Floratos, E. 179, 325 passim, 336, 343
equation 161, 255, 263 If. Fluctuation in Euclidean space 276, 283
process 245, 302 passim passim
at imaginary time 303 passim in geometry 9
Dirac equation 306 in metric 381, 384 passim, 397 passim
Dispersion of random variable 216 in space-time see stochastic space-time
Doob's inequality 224 in topology 9
Drift process 252 random 260 passim
Duffin-Kemmer equation 120 Fock space 93, 96, 101, 103
Dynamics nonrelativistic 277 passim Fokker-Planck equation 266 If.
relati vistic 283 passim Form factor 7, 12, 14, 47 If.
Fourier transform 24, 28, 32, 39, 41, 44 ff.
Efimov, G. V. 5, 7, 11, 12, 13 If. of generalized function 12, 23 passim,
Efinger, H. J. 260, 376, 386, 396 41,44
Einstein, A. 8, 9, 10, 255, 257 If. Free field 30, 94, 274
Electroweak theory 3, 105 passim, 134 process 317
Entire function 12, 14, 37 If. Frederick, C. 9, 12
space 34 passim Frohlich's reconstruction theorem 272
Equation of motion 277 passim, 383 Fujiwara, K. 376
passim, 299, 301, 306 passim, Functional differentiation 158, 161
322, 354, 385 passim, 392, 393, 398 passim, 175 passim, 179 If.
of continuity 266 If. Functional integral 175 If.
with retardation 47, 92, 99. order of 20
Equidistance behavior of particle mass space 19 passim
spectrum 358, 378 passim linear 19, 30, 32
Equilibrium distribution 326, 341, 343 Fundamental length 5, 7, 8, 277 If.
passim
Equivalence of stochastic and canonical Galilean transformation 7, 299
quantization 342 passim Garding-Wightman axiom 3, 29, 272
Euclidean Green's function 4, 94, 185,254 passim
If. Gauge ambiguity 327
Markov field 3, 268 passim, 327 passim field 336 If.
space 11, 270 invariance 105 passim, 135
Euler-Lagrange equation 264, 267 passim 148 If.
Event 213 passim see also S-matrix
Evolution equation 47, 93, 97, 98, 99 length (weak interaction length) 135
Exotic decays 142 passim theory 3, 105 passim, 134, 135, 326 If.
Expectation 162 passim, 217 If. transformation 264
vacuum 13 If. global 105
Extended object 4, 13, 64, 68, 94 If. local 106
see also nonlocal field Gaussian integral 169, 177
measure 14, 97, 226 passim, 243 ff.
Faddeev-Popov procedure 336, 342, 347 random variable 162, 216 passim, 219
de Falso, D. 296, 317 passim 225 passim
Fenyes, I. 4, 253, 256, 281 covariance of 217, 218 If.
Index 423

Gaussian integral (contd.) Imaginary time 6, 11, 93 ff.


jointly 217 Improper limit 61, 80
independent of 219 see also weak limit
process 226 passim, 269 Indefinite metric 8, 11, 48, 50, 55
Generalized function 12, 15, 19 space 56
passim 27, 39 passim Indicator of events 215, 220
see also Local and nonlocal Inflationary model 259
generalized functions Integration over Grassman algebra 167
Generating functional for Green's passim, 175 passim
function 180 passim change variable 168 passim
Ghost fields 342 Interaction Lagrangian 46 ff.
states 48, 56, 63 picture 60, 97
Gihman, I. I. 213, 231, 244 Isomorphism 24
GIM mechanism 152 Ito's lemma 248 passim
Glashow, Sh. 3, 105, 134, 152, 365 derivatives 277
Glimm-Jaffe theory 3 integral 248 passim
see also constructive quantum field process 264
theory rule 246, 277 passim
Grand unified theory 259, 358, 365, 369
passim, Jaffe, A. M. 3, 31, 33, 42, 157
Grassmann algebra 165 passim
with infinite number generators 171
passim Kallen-Lehmann representation 30
with involution 173 passim Kershaw, D. 4, 253, 277 ff.
measure 178, 207. Klein-Gordon equation 46, 63, 269, 277,
Gravitation length see Planck length 286, 302 passim
vacuum 257 passim Kolmogorov, A. N. 213, 230
repulsion 259 passim Kroll's prescription 108 passim
Green's function 13, 27, 50, 58, 61, 63 ff.
see also propagator of particles Lamb shift 6, 119 passim
advanced 58 Langevin equation 325, 355
causal local 27, 50, 58, 65, 83 Linearization of equation 291
passim 94 ff. Lipschitz condition 264
nonlocal 13, 63, 81 passim, 98, 112 Jr. Local generalized function 15, 16, 20
retarded 58 passim, 27 passim, 35, 40 ff.
Gribov, V. N. 327, 336, 352 Locality 4,27 passim, 31 passim, 72, 75,
Guerra, F. 4, 179, 211, 253, 263 272
passim, 269 ff. Lorentz condition 48, 286

Hamiltonian 55 Jr. MacDonald function 94, 271


Hamilton-Jacobi equation 294, 297, 300 Macrocausality condition 5, 76 passim
Hankel function 28, 83, 271 Magnetic moment of leptons 6, 116
Harmonic oscillator 268 passim passim, 138 passim
Heisenberg, W. 5, 31, 65, 72, 180 of neutrino 146 passim
Hidden variable theory 5, 254, 256, 267, March, A. Z. 9
277 Markov, M. A. 5, 9, 364, 365, 374, 384
Hierarchical scale 365 passim Markov principle 4, 233 passim
Hierarchies of masses 365 passim process 4, 233 passim, 257, 261, 264,
of lengths 365 passim 275
Holder inequality 188 temporally homogeneous 233
Hubble constant 387 time-reversed 238 passim
Martingales 224 passim, 242
Iliopoulos, 1. 179, 325, 326, 327 passim de Martino, S. 317 passim
424 IndeJl

Matinyan, S. G. 324, 352 passim Parisi, G. 324 passim, 336 passim


Maxwell equation 331 passim Path integral see functional integral
Measurable map 215 passim Pauli-Villars regularization 6
Mellin representation 14,71,81 de la Peiia-Auerbach, L. 4, 253 If.
Metric averaged 10, 13, 260, 373 Planck constant 5, 253, 281, 365
passim, 385, 396 passim length 358, 364, 387, 388, 396
Microcausality condition 4, 30, 31 mass 367, 375, 403
passim, 61, 73 passim, 88, 93 Poincare-Lorentz transformation 8
Minkowski space-time 8, 10, 93 If. Point-like lepton 6, 14
metric 8, 260 If. Poisson's process 232
Minimal electromagnetic interaction 106 transition function 237
Misner, Ch. W. 8, 9, 321 passim Pontryagin index 349
Moments of Gaussian processes 227 Positive definite 217, 218, 227, 229, 273
passim Poynting's vector 352
of random variable 216 passim Progeometry 8
Monomials 165 passim Probability density 263 ff.
Multiplication operation 25 passim, 79, conditional 220 passim
80, 87, 95, 171 distribution 215 ff.
Multiplier function 25, 78 joint 216 passim
measure 7, 10, 13 passim, 35, 39,.98 If.
space 214 If.
Nelson, E. 3, 4, 242, 244, 253, 263 ff.
transition 232 If.
Newton constant 257, 387
Projecting operator 122
equation 299
Propagator of particles 13 If.
law 266,277
Proper time 276,283, .291,321 passim
potential 371
Proton decay 358, 369 passim
Nonlinear partial differential equation 253,
Lifetime 358, 373
267, 279, 285, 290, 294, 299 ff.
Prugovecki,1i 9
Nonlocal generalized function 12, 34, 39
passim 48 ff.
Quantization of nonlocal fields 47 passim
Euclidean-Markov field 359 passim
canonical 54, 179, 336, 342
field 7, 13, 49 passim, 77, 98, 112 ff.
Quark confinement 259, 380
interaction 5 passim, 46 If.
quantum potential 295
Radon-Nikodym theorem 222
space-time metric 10, 13, 260, 373
Random process 226 If.
passim, 385, 396 passim
variable 214 passim
stochastic electrodynamics 363, 364
independent 219 passim
zero-point field 362 passim
orthodiagonal 218
Nonlocality 5, 6, 8 passim, 46, 63, 72,
walk 241, 260 passim
100, 112 If.
Regularization procedure 47, 51 passim,
Nordstrom-Reissner solution 366
60, 64, 70, 81 passim, 113 If.
Norm of function 17 passim, 32, 34, 160
Riccati-type equation 386
of state vector 97 passim
Riemannian metric 260, 385, 386, 396
in L z 17, 218 passim, 247
S-matrix 6, 29 ff.
Order of growth for entire functions 38 causal 31, 61, 73 passim, 88
Ornstein-Uhlenbeck velocity process 229, correspondence principle 47
269 functional form 86 passim, 179 passim
Oscillation of neutrino 140 passim gauge invariant 107 passim, 113, 120,
of neutron 358, 370 passim 135 passim
time 372 stability condition 65
Oscillator process 229 translation invariant 79
Osterwalder-Schrader positivity 272 unitary 59, 61, 78, passim, 86 passim
Index 425

Salam, A. 3, 105, 134, 365, 387 Stochasticity in physics 6, 10, 255 passim
Schrodinger equation 97 passim, 164, Submartingale 224 passim
253, 267 fT. Superfield 372
Schwarzschild radius 366, 379, 387 Supermartingale 224 passim
Schwinger, 1. 3, 92, 181, 254 Support of generalized function (or
Schwinger function 269 passim, 329, 331 functional) 27, 30, 35, 77
nonlocal 360 passim of function 21, 30, 31, 76
Self-energy diagram 115, 116, 130, 131, 136 Switching ofT (or on) the interaction 60,
of Bosons 130, 131 64, 73, 78 fT.
Self-turbulance in motion 296 passim
Semi-norm 17 Takano, Y. 5, 9, 10, 374
Shucker, D. S. 267, 296, 317 passim Tempered distribution 21, 31,40
de Siena, S. 317 passim Test function 16 passim, 34 passim
Simon, B. 93, 157, 216, 218, 244, 271 particle 375 passim, 379 passim
Sivashinsky equation 296 passim Thorne, K. S. 8, 9, 321 passim
Skew symmetric 165, 172 passim Time-reversal transformation 279
Smoluchowski equation 233, 263 fT. Tomonaga-Schwinger equation 92
type equation 276 fT. T -ordering operation 13, 59 fT.
Solitary waves 260, 296, 396, 400 passim Transition density 236 passim
Soliton-like motion 400 passim function 235 passim
Space countably normalized 18, 20 see also probability transition
normalized 17, 20 conservative 236, 237, 238
Space-time cellular 9 normal 236, 237, 238
code 9 Triangle inequality 17
conformally flat 260, 358, 385 passim Truncation moment 233
discrete 8 Two-particle problem in stochastic
"foam-like" 9 theory 287 passim, 291 passim
higher dimensional 9, 10 Type of entire functions 38
lattice 8
stochastic 4, 6, 7, 9, 11 fT. Ultraviolet divergences 4, 6, 28 fT.
unification 8 Uncertainty relations for position and
Spectrum density of zero-point electro- momentum in stochastic mechanics
magnetic field 254, 362 passim 318 passim
Spread-out object see non local and Vacuum energy density 189 passim, 207,
extended objects 258 passim
Stability of matter see proton decay polarization 6, 7, 107, 113 passim, 120
Stochastic calculus 244 passim of Bosons 126, 127
derivative 245 passim, 277 tunneling phenomena 347 passim
differential equation 245, 264 fT. state 56,93
electrodynamics 254, 256, 362, 363 Velocity backward 277 fT.
electromagnetic field 331 passim current 266 fT.
integral 245 passim, 250 passim of light 5, 8
mechanics 275 ff. forward 277 fT.
metric 10 fT. stochastic 266 fT.
motion 308, 309, 397 passim Vigier, J.-P. 4, 253 passim 287 ff.
see also Brownian motion
process see random process Ward-Takahashi identity 107 passim
with independent increments 229 Weak limit 15, 63, 79, 326, 343
passim Weinberg, S. 3, 105, 134, 257, 335
quantization 264 fT. Weinberg-Salam-Glashow theory 3, 105
scalar field 268 fT. passim, 134
solutions to Yang-Mills equation 352 Weizel, W. 257
passim Weyl, H. 302, 304
426 Index

Wheeler, J. A. 8, 9, 321 passim Wu, Y.-S. 324 passim, 336 passim


Wick-ordered exponential 248
Wiener integral 163, 243 passim, 351 Yang-Mills theory 3, 107, 348, 352 passim
measure 162, 243 Yasue, K. 4, 324, 337, 347 passim
process 232, 239 passim, 264 Yukawa model 203 passim
transition function 237, 239 potential 14, 68, 371, 377
Wightman, A. S. 3, 60
axioms 29, 272 Zero-point electromagnetic field 254
function 29, 50, 269 passim passim, 362 passim
White noise 242, 325 Zwanziger, D. 324, 327, 336 passim, 342,
Wormholes 9 343, 347

You might also like